24
An efficient iterative method for the formulation of flow networks Wei Hu a , Dong Wang b , Xiao-Ping Wang a,b,* a Department of Mathematics, The Hong Kong University of Science and Technology, Clear Water Bay, Kowloon, Hong Kong. b School of Science and Engineering, The Chinese University of Hong Kong, Shenzhen, Guangdong 518172, China. Abstract We propose an efficient iterative convolution thresholding method for the formulation of flow networks where the fluid is modeled by the Darcy–Stokes flow with the presence of volume sources. The method is based on the minimization of the dissipation power in the fluid region with a Darcy term. The flow network is represented by its characteristic function and the energy is approximated under this representation. The minimization problem can then be approximately solved by alternating: 1) solving a Brinkman equation to model the Darcy–Stokes flow and 2) updating the characteristic function by a simple convolution and thresholding step. The proposed method is simple and easy to implement. We prove mathematically that the iterative method has the total energy decaying property. Numerical experiments demonstrate the performance and robustness of the proposed method and interesting structures are observed. Keywords: topology optimization, convolution, thresholding, Darcy–Stokes flow 1. Introduction The flow networks are the cornerstone infrastructure of transporting substances and information. They are classified into two categories, social transport networks [1] and biological transport networks [2], based on whether they are created and designed by humans. Examples of social transport networks include but not limited to electrical power-grids, railway and highway networks, information networks such as the internet, and supply and demand networks. Biological transport networks are widely present in the organism, e.g.leaf venation, blood vessel systems, neural networks etc. The primary design purpose of a flow network is to balance the cost of constructing and maintaining the structure, the efficiency of transportation, and resilience to local failure and disturbance. Explosive growing demand for transferring matter and messages put urgent needs in studies on transport networks. Material properties are closely related to their structure. Studies on networks’ structures not * Corresponding author Email addresses: [email protected] (Wei Hu), [email protected] (Dong Wang), [email protected] (Xiao-Ping Wang ) 1

An efficient iterative method for the formulation of flow

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: An efficient iterative method for the formulation of flow

An efficient iterative method for the formulation of flow networks

Wei Hua, Dong Wangb, Xiao-Ping Wanga,b,∗

aDepartment of Mathematics, The Hong Kong University of Science and Technology, Clear Water Bay, Kowloon, HongKong.

bSchool of Science and Engineering, The Chinese University of Hong Kong, Shenzhen, Guangdong 518172, China.

Abstract

We propose an efficient iterative convolution thresholding method for the formulation of flow networks where

the fluid is modeled by the Darcy–Stokes flow with the presence of volume sources. The method is based

on the minimization of the dissipation power in the fluid region with a Darcy term. The flow network is

represented by its characteristic function and the energy is approximated under this representation. The

minimization problem can then be approximately solved by alternating: 1) solving a Brinkman equation

to model the Darcy–Stokes flow and 2) updating the characteristic function by a simple convolution and

thresholding step. The proposed method is simple and easy to implement. We prove mathematically

that the iterative method has the total energy decaying property. Numerical experiments demonstrate the

performance and robustness of the proposed method and interesting structures are observed.

Keywords: topology optimization, convolution, thresholding, Darcy–Stokes flow

1. Introduction

The flow networks are the cornerstone infrastructure of transporting substances and information. They

are classified into two categories, social transport networks [1] and biological transport networks [2], based

on whether they are created and designed by humans. Examples of social transport networks include but not

limited to electrical power-grids, railway and highway networks, information networks such as the internet,

and supply and demand networks. Biological transport networks are widely present in the organism, e.g.leaf

venation, blood vessel systems, neural networks etc. The primary design purpose of a flow network is to

balance the cost of constructing and maintaining the structure, the efficiency of transportation, and resilience

to local failure and disturbance.

Explosive growing demand for transferring matter and messages put urgent needs in studies on transport

networks. Material properties are closely related to their structure. Studies on networks’ structures not

∗Corresponding authorEmail addresses: [email protected] (Wei Hu), [email protected] (Dong Wang), [email protected]

(Xiao-Ping Wang )

1

Page 2: An efficient iterative method for the formulation of flow

only help us build efficient and robust networks but also shed insights into the understanding of biological

transport networks where natural selection plays the role of optimization processes. For example, the cubic

laws for blood vessel which was discovered in experiments can be explained by balancing the delivering

energy cost of blood flow and the metabolic cost of blood cells [3, 4].

The difficulties and challenges of designing flow networks are the non-convex optimization property and

ill-posedness (no solution, e.g.Steiner tree model [5] or multiple solutions [6–8]) of the formulated mathe-

matical models. It is necessary to adopt filtering techniques and/or design suitable objective functionals

to resolve the well-posedness. To tackle the non-convexity, stability, convergence and ability to avoid local

optima are critical for a practical numerical algorithm.

There are many algorithms developed for constructing transport networks. In [2], discrete adaption

dynamics, which reacts to local stimulus for network formation, is proposed. The continuous analogy has

also been studied and analyzed [9–12]. Tree-like structures have been observed using their models. They

also discover that loops emerge when strong fluctuations of flow sources present. In [13], a flow network

system consists of Darcy and Stokes region is analyzed, and optimal layout is obtained through sensitivity

analysis and the first-order method MMA [14]. Many other methods have also been developed in the context

of shape/topology/structure optimization: density interpolation [15]; level-set method approach [16, 17];

topological derivatives [18]; phase field method [19]; evolutionary structural optimization (ESO) approaches

[20], and several others. To ensure well-posedness and mesh-independent solutions, regularization is usually

needed in topology optimization. Many robust regularization approaches were introduced. For example,

regularizing the optimization problem by introducing fictitious interface energy [21] or penalizing the original

objective function by perimeter control functions based on phase field methods [22].

In this paper, we formulate the formation of flow networks as a constrained optimization problem. To

be specific, we assume the fluid in the network is either a Stokes flow or governed by Darcy’s law. The

interface is implicitly represented by characteristic functions of corresponding domains and the network is

formed by minimizing the total potential energy of the system or the dissipation of the fluids under volume

constraints. It is motivated by the MBO method for approximating mean curvature flow using characteristic

functions [23] and Esedoglu and Otto’s [24] novel interpretation using minimizing movement for problems

of multiphase flow with arbitrary surface tensions. The MBO method has attracted much attention due

to its simplicity and unconditional stability. It has subsequently been extended to deal with many other

applications, including image processing [25–27], problems of anisotropic interface motions [28, 29], the

wetting problem on solid surfaces [30, 31], and so on.

The main observation is that the approximate energy functional with perimeter regularization (using

Esedoglu and Otto’s formula in [24]) is concave with respect to the characteristic function of the domain

and convex with respect to the fluid velocity field. This allows us to design an iterative method to minimize

the energy functional effectively and robustly.

2

Page 3: An efficient iterative method for the formulation of flow

In our method, we minimize the total potential power with a perimeter regularization, subject to some

constraints (i.e.; 1. Stokes region volume is fixed, 2. Brinkman equation is used to model the velocity field in

the whole domain, and 3. The presence of constant or delta volume sources, which implies the divergence of

the velocity field is a delta function or constant). Borrowing the idea from threshold dynamics [24, 26], the

perimeter term is based on the diffusion of the characteristic functions of the fluid regions. We introduce the

Brinkman equation to “interpolate” between the Stokes equation in the flow region and Darcy flow through

a porous medium (a weakened nonfluid region). In each iteration of the proposed algorithm, we solve the

Brinkman equation for the whole domain by the finite-element method and update the fluid-solid regions

by a convolution step followed by a thresholding step (See Algorithm 1). The proposed scheme is simple,

easy to implement, and stable.

From both theoretical well-posedness and numerical stability concerns, we apply diffusion as a filter to

the fluid-solid region. It couples well with the diffusive approximation of perimeter. The diffusion step size

is an adjustable parameter in our method, based on which we manipulate this parameter and improve our

algorithm to a self-warm-starting version (See Algorithm 2). Numerical results show that the method is

efficient, robust, and insensitive to the initial guess and parameters. In particular, we further prove the

unconditional energy-decaying property of the proposed algorithm.

The paper is organized as follows. In Section 2, we introduce the model and derive the approximate energy

for the total energy. Based on this approximation, we derive an efficient iterative convolution thresholding

method and prove its unconditional stability (i.e.; the energy decaying property). We discuss the numerical

implementation and demonstrate the efficiency and the performance of the proposed method in Section 3.

Finally, we draw some conclusions and discussions in Section 4.

2. Derivation of the method

In this section, we discuss the derivation of the proposed method. We first introduce the mathematical

model for modeling the fluid flow with fluid generation and consumption. Then, we discuss the relaxed and

the approximate model. In Section 2.3 we derive the iterative convolution thresholding method to solve the

approximate problem iteratively. The unconditional stability is proved in Section 2.4.

2.1. The mathematical model

To introduce the optimization problem, we first introduce the following forward problem. In a fixed

domain Ω consisting of fluid region Ωf and solid region Ωs (i.e.; Ω = Ωf ∪ Ωs), we consider a generalized

Stokes equation, which is similar to the Brinkman system, to “interpolate” between the Stokes equation in

3

Page 4: An efficient iterative method for the formulation of flow

Ωf and the Darcy’s law in Ωs:

αu− 2∇ · (µε(u)) +∇p = g,

∇ · u = s,

u|ΓD= uD,

−pn + 2µε(u)n|Γt = t

(1)

where α denotes inverse permeability, µ is dynamic viscosity, p is pressure, ε(u) = ∇u+(∇u)T

2 is strain, g is

body force, and s is volume source in the system. The boundary is splitted into two parts ∂Ω = ΓD ∪ Γn,

where Dirichlet and Neumann boundary conditions are specified respectively. In the fluid region, α = 0, the

partial differential equation reduces to Stokes equation. In solid region, µ = 0, the equation is the Darcy’s

law. In the extreme case where α = ∞ with no body force, the differential equation leads to u = 0. It is

easy to check that, when Ωf and Ωs are fixed, the solution of this system of partial differential equations is

the minimizer of the following total potential energy functional,

E(u,Ωf ,Ωs) =1

2

∫Ω

αu · u + 2µε(u) : ε(u) dx−∫

Ω

g · u dx−∫

Γt

t · u dS (2)

Note here E depends on Ωf ,Ωs through α, µ as well as u which varies with α, µ.

Now, we consider the topology optimization problem to find the optimal interface between the fluid

domain Ωf and the solid domain Ωs in the sense of minimizing E(u,Ωf ,Ωs) subject to the constraints in (1)

and an additional volume constraint. Certain regularization techniques are required from both theoretical

and numerical points of view.

Theoretically, it is well-known that such topology optimization problems lack the existence of solutions

property without suitable restrictions on admissible configurations or appropriate regularization. Numer-

ically, discontinuity brings difficulty in convergence and stability of the numerical scheme. We adopt two

techniques to address this issue. First, we will apply a Gaussian filter on the solutions (interpolation, see

Section 2.2.1). Also, we may add perimeter penalty to energy functional (2) to smooth the problem. To be

specific, the goal of this paper is to develop an efficient method to solve approximately,min

u,Ωf ,Ωs

E(u,Ωf ,Ωs) + γ|∂Ωf |,

s.t.,u solves (1), |Ωf | = ω|Ω|(3)

where |∂Ωf | denotes the perimeter or surface area of the fluid-solid interface, |Ωf | denotes the volume of the

fluid domain, and ω is the fraction for the volume constraint.

2.2. The relaxed and approximate model

Motivated by the idea from the threshold dynamics methods developed in [24, 26], we use the character-

istic functions for the fluid region and the solid region to implicitly represent the interface.

4

Page 5: An efficient iterative method for the formulation of flow

Define an admissible set B as follows:

B :=(v1, v2) ∈ BV (Ω) | vi(x) = 0, 1, v1(x) + v2(x) = 1 a.e. in Ω, and

∫Ω

v1 dx = ω|Ω|, (4)

where BV (Ω) is the space of functions with bounded variation in Ω, and ω is the fixed volume fraction of

the fluid region. We introduce χ1(x) to denote the characteristic function of the fluid region Ωf , i.e.;

χ1(x) = χΩf:=

1, if x ∈ Ωf ,

0, otherwise,

and χ2(x) as the characteristic function of Ωs, i.e.χ2(x) = 1 − χ1(x). The interface Γ is then implicitly

represented by χ1 and χ2. Let χ = (χ1, χ2) and we have χ ∈ B. The problem (3) can be reformulated as,minu,χ∈B

E(u, χ) + γTV (χ),

s.t.,u solves (1),

(5)

where the total variation TV (χ) is defined as,

TV (χ) = supφ∈C1

c (Ω),|φ|≤1

∫Ω

χ1∇ · φdx.

2.2.1. Interpolation

As we have mentioned, smooth interpolation between solid and fluid is needed. Suggested by the thresh-

old dynamics algorithm, we shall consider the following interpolation:

Xτ = Gτ ∗ χ1 : χ ∈ B, (6)

where Gτ is the heat kernel,

Gτ (x) =1

(4πτ)d/2exp(−|x|

2

4τ),

here d = 2, 3 is the dimension of space.

We next interpolate inverse permeability and viscosity as,

α(χ) = αs + (αf − αs)Gτ ∗ χ1, (7)

µ(χ) = µs + (µf − µs)Gτ ∗ χ1, (8)

where subscripts s, f correspond to parameters in solid and fluid respectively.

5

Page 6: An efficient iterative method for the formulation of flow

2.2.2. Weak and variational formulation

The weak form of (1) is: find u ∈ [H1(Ω)]d,u|ΓD= uD such that for all v ∈ [H1(Ω)]d,v|ΓD

= 0 on ΓD,

a(u,v)− b(p,v) = L(v),

b(p,u) = (s, p)L2(Ω),

(9)

where a, b are the bilinear form,

a(u,v) =

∫Ω

αu · v + 2µε(u) : ε(v) dx, (10)

b(p,v) =

∫Ω

p∇ · v dx, (11)

L is a linear functional,

L(v) =

∫Ω

g · v dx+

∫Γt

t · v dS. (12)

We note here that the energy functional (2) can be written as,

E(u,Ωf ,Ωs) =1

2a(u,u)− L(u).

The variational formulation of (1) is: find u such that it solves the minimization problem,

minv∈[H1(Ω)]d,v|ΓD

=uD

E(u,Ωf ,Ωs). (13)

2.2.3. The approximate energy

Motivated by [24, 26, 32], we approximate the TV (χ) in (6) by

|TV (χ)| ≈√π

τ

∫Ω

χ1Gτ ∗ χ2 dx, (14)

which measures the heat content “escaped” from the fluid domain χ1 to the solid domain χ2 in a short time

τ .

Now, we arrive at the following approximate optimization problem to (5):minu,χ∈B

Eτ (u, χ) : = 12a(u,u)− L(u) + γ

√πτ

∫Ωχ1Gτ ∗ χ2 dx,

s.t.,u solves (9),

(15)

Remark 2.1. For simplicity, we use the same τ in interpolation of α, µ and approximation of |TV (χ)|.

Indeed, one can use different values of τ in different terms and the algorithm is similar.

6

Page 7: An efficient iterative method for the formulation of flow

2.3. The iterative convolution thresholding method

In this section, we solve the approximate problem (15) iteratively. We decouple the dependence on u

and χ to design an efficient and unconditional stable scheme to minimize Eτ (u, χ). We use a coordinate

descent algorithm to minimize the approximate energy Eτ (u, χ) with constraints (9) on the admissible set

B. That is, given an initial guess χ0 = (χ01, χ

02), we compute a series of minimizers

u0, χ1,u1, χ2, · · · ,uk, χk+1, · · ·

such that

uk = arg minu∈SEτ (u, χk), (16)

χk+1 = arg minχ∈BEτ (uk, χ), (17)

for k = 0, 1, 2, · · · . Here, B is defined in (4) and the admissible set S is defined as

S :=

u ∈ [H1(Ω)]d | u|ΓD

= uD,

∫Ω

p(∇ · u− s) dx = 0

where s is a fixed function.

Given the k-th iteration χk, we first solve (16) to get the uk. As we discussed in Section 2.2.2, the

minimum is attained at uk which solves (9) when χk is fixed.

That is, uk can be obtained by finding u ∈ [H1(Ω)]d,u|ΓD= uD such that for all v ∈ [H1(Ω)]d,v|ΓD

= 0

on ΓD, ak(u,v)− b(p,v) = L(v),

b(p,u) = (s, p)L2(Ω),

(18)

where ak(u,v) is the corresponding form of a(u,v) at χk,

ak(u,v) =

∫Ω

α(χk)u · v + 2µ(χk)ε(u) : ε(v) dx. (19)

Given uk, we now rewrite the objective functional Eτ (χ,u) into Eτ,k(χ) as follows:

Eτ,k(χ) : =1

2

∫Ω

α(χ)uk · uk + 2µε(uk) : ε(uk)− 2g · uk dx−∫

Γt

t · uk dS + γ

√π

τ

∫Ω

χ1Gτ ∗ χ2 dx

=1

2

∫Ω

(αGτ ∗ χ1)uk · uk + (2µGτ ∗ χ1)ε(uk) : ε(uk) dx+ γ

√π

τ

∫Ω

χ1Gτ ∗ χ2 dx+ C

where C =∫

Ωαsu

k · uk + 2µsε(uk) : ε(uk)− g · uk dx−

∫Γt

t · uk dS, α = αf − αs, and µ = µf − µs.

The next step is to find χk+1 such that

χk+1 = arg minχ∈BEτ,k(χ). (20)

7

Page 8: An efficient iterative method for the formulation of flow

We note that, because χ2 = 1 − χ1, Eτ,k(χ) is strictly concave in χ1. Thus, (20) is a minimization of

a concave functional on a nonconvex admissible set B. However, we can relax it to a problem defined on a

convex admissible set by finding rk+1 such that

rk+1 = arg minr∈HEτ,k(r), (21)

where H is the convex hull of B defined as follows:

H :=(v1, v2) ∈ BV (Ω) | vi(x) ∈ [0, 1], and v1(x) + v2(x) = 1 a.e. in Ω,

∫Ω

v1 dx = ω|Ω|. (22)

The following lemma shows that the relaxed problem (21) is equivalent to the original problem (20).

Therefore, we can solve the relaxed problem (21) instead.

Lemma 2.2. Let u ∈ H1uD

(Ω,Rd) be a given function and r = (r1, r2). Then

arg minr∈HEτ,k(r) = arg min

r∈BEτ,k(r). (23)

Proof. We refer a similar proof to [26].

Now, we show that (21) can be iteratively solved with a simple thresholding step. Because Eτ,k(r) is

quadratic in r and concave, we first linearize the energy Eτ,k(r) at rk by

Eτ,k(r) ≈ Eτ,k(rk) + Lτ,krk

(r − rk), (24)

where

Lτ,krk

(r) =

∫Ω

√π

τr1Gτ ∗ (1− 2rk1 ) +

1

2αr1Gτ ∗ (uk · uk) + µr1Gτ ∗ (ε(uk) : ε(uk))

)dx (25)

=

∫Ω

r1φ1 dx.

Here

φ1 = γ

√π

τGτ ∗ (1− 2rk1 ) +

α

2Gτ ∗ (uk · uk) + µGτ ∗ (ε(uk) : ε(uk)). (26)

Then (21) can be approximately reformulated into

χk+1 = arg minr∈HLτ,krk

(r) = arg minr∈H

∫Ω

r1φ1 dx. (27)

It is easy to see that (27) can be solved in a point-wise manner byχk+11 (x) = 1, if φ1(x) < δ,

χk+11 (x) = 0, otherwise,

(28)

where δ is chosen as a constant such that∫

Ωχk+1

1 dx = V0, and χk+12 = 1− χk+1

1 .

The method is summarized in Algorithm 1.

8

Page 9: An efficient iterative method for the formulation of flow

Algorithm 1: An iterative convolution thresholding method for (15)

Input: Let Ω be a fixed domain, τ > 0, ω > 0 and initial χ0.

Output: A function χ ∈ B that approximately minimizes (15) .

Set k = 0

while not converged do

1. Find u ∈ [H1(Ω)]d,u|ΓD= uD such that for all v ∈ [H1(Ω)]d,v|ΓD

= 0 on ΓD,ak(u,v)− b(p,v) = L(v),

b(p,u) = (s, p)L2(Ω).

2. Evaluate

φ1 = γ

√π

τGτ ∗ (1− 2χk1) +

α

2Gτ ∗ (uk · uk) + µGτ ∗ (ε(uk) : ε(uk)).

3. Choose a proper δ and set χk+11 (x) = 1, if φ1(x) < δ,

χk+11 (x) = 0, otherwise,

such that∫

Ωχk+1

1 dx = V0.

Set k = k + 1

To determine the value of δ, one can treat∫

Ωχk+1

1 dx−V0 as a function of δ and use an iteration method

(e.g., bisection method or Newton’s method) to find the root of∫

Ωχk+1

1 dx− V0 = 0.

For the uniform discretization of Ω, a more efficient method is the quick-sort technique proposed in [30].

Assume we have a uniform discretization of Ω with grid size h, we can approximate∫

Ωχk+1

1 dx by Mh2,

sort the values of φ1 in an ascending order and simply set χk+11 = 1 on the first M points and χk+1

2 = 1 on

the other points.

In many implementations, one may discretize the domain using nonstructural meshes. To preserve the

volume for the discretization on nonuniform grids, although the volume cannot be simply approximated by

the number of grid points times the size of each cell, a similar technique can be applied. One can still sort

the values of φ1 in ascending order, save the index into S, calculate the integrating weight at each grid point

into V, and set V = 0 and i = 0. Then, δ can be simply found by:

while V < V0; i← i+ 1; V = V + V(S(i)); end; δ = φ1(S(i+ 1)).

Besides, we note that in the original threshold dynamics method, on one hand, the algorithm can be

easily stuck when τ is very small because, in the discretized space, τ is so small that no point can switch

9

Page 10: An efficient iterative method for the formulation of flow

Algorithm 2: An iterative convolution thresholding method with warm start

Input: Given Ω as a fixed domain, an initial τinit > 0, a target τ > 0, ω > 0 and an initial shape χ0.

Output: A function χ ∈ B that approximately minimizes (5) .

Set a rough convergence criterion

Set k = 0, τcurrent = τinit

while τ < τcurrent do

1. Run Algorithm 1 with τcurrent, initial shape χk and prescribed convergence criterion to get

new shape χk+1.

2. Set k = k + 1, τcurrent = 0.5× τcurrent

from one phase to another (i.e., χ1 changes from 0 to 1 or 1 to 0) at one iteration step. On the other hand,

with a large τ , the interface can easily move but creates a large error. Hence, we apply the adaptive in time

technique [30] in numerical experiments by modifying Algorithm 1 into an adaptive algorithm by adjusting

τ during the iterations, which is summarized in Algorithm 2.

2.4. Stability analysis

In this section, we prove the unconditional stability property of the proposed algorithm. Specifically, we

have the following theorem.

Theorem 2.3. For the series of minimizers

u0, χ1,u1, χ2, · · · ,uk, χk+1, · · · ,

calculated with Algorithm 1, we have

Eτ (χk+1,uk+1) ≤ Eτ (χk,uk) (29)

for all τ > 0. Furthermore, the equality holds only when χk+1 = χk.

Proof. As we discussed in Section 2.2.2, we have that uk+1 attains the minimum value of Eτ (χk+1,u) when

χk+1 is fixed. In other words, for all u ∈ S, we have

Eτ (χk+1,uk+1) ≤ Eτ (χk+1,u). (30)

In particular, we have

Eτ (χk+1,uk+1) ≤ Eτ (χk+1,uk). (31)

10

Page 11: An efficient iterative method for the formulation of flow

For a fixed uk, from the linearization of Eτ,k(χk) in (24), we have

Eτ (χk,uk) =Lτ,kχk (χk) + γ

√π

τ

∫Ω

χk1Gτ ∗ χk1 dx,

Eτ (χk+1,uk) =Lτ,kχk (χk+1) + γ

√π

τ

∫Ω

(χk+1

1 Gτ ∗ (1− χk+11 )− χk+1

1 Gτ ∗ (1− 2χk1))dx.

Because Lτ,kχk (χk+1)− Lτ,k

χk (χk) ≤ 0, we have

Eτ (χk+1,uk)− Eτ (χk,uk) =Lτ,kχk (χk+1)− Lτ,k

χk (χk)

− γ√π

τ

∫Ω

(χk+1

1 Gτ ∗ χk+11 − 2χk+1

1 Gτ ∗ χk1 + χk1Gτ ∗ χk1)dx

≤− γ√π

τ

∫Ω

(χk+11 − χk1)Gτ ∗ (χk+1

1 − χk1) dx (32)

=− γ√π

τ

∫Ω

[Gτ/2 ∗ (χk+11 − χk1)]2 dx

≤0

where the penultimate equality is a consequence of the semi-group property of heat kernel.

Combining (31) and (32), we have

Eτ (χk+1,uk+1) ≤ Eτ (χk+1,uk) ≤ Eτ (χk,uk).

Furthermore, because

−γ√π

τ

∫Ω

[Gτ/2 ∗ (χk+11 − χk1)]2 dx

is 0 only when χk1 = χk+11 , we have that the equality in (29) holds only when χk1 = χk+1

1 .

3. Numerical experiments

In this section, we will first introduce the implementation of Algorithm 2 using the finite element method.

Then, we will display a variety of numerical examples to show the performance of the method. In all

numerical examples, we set Ω = [0, 1] × [0, 1]. All numerical results were run on a single machine with two

E5-2683v4 (16 cores each) and 128GB memory.

3.1. Finite element discretization

Based on weak formulation (9) in section 2.2.2, we will solve numerically (1) by mixed finite element

method. We will use a uniform mesh with triangular elements. Let h be parameter associated to discretiza-

tion size of the domain, Th be uniform triangulation of Ω. See Figure 1 for a typical discretization consists

of 2× 10× 10 triangles.

11

Page 12: An efficient iterative method for the formulation of flow

Figure 1: A sturctural mesh with 2 × 10 × 10 triangles. See Section 3.1.

In light of the LBB stability condition [33–36], we use standard P1−element for approximation of pressure

and P2−element for velocity, i.e. for pressure,

Ph = vh ∈ C(Ω) : vh|T ∈ P1(T ), T ∈ Th; vh|ΓD= 0,

for velocity,

Uh = vh = ((vh)1, (vh)2), (vh)1,2 ∈ C(Ω) : (vh)1,2|T ∈ P2(T ), T ∈ Th;vh|ΓD= uD,

Pk denotes polynomial of order up to k. Test space of pressure is the same as trial space Ph while test space

for velocity is,

Wh = vh = ((vh)1, (vh)2), (vh)1,2 ∈ C(Ω) : (vh)1,2|T ∈ P2(T ), T ∈ Th;vh|ΓD= 0.

To approximate the solution to the system (1) or its weak form (9) numerically, we solve the linear system:

find (uh, ph) ∈ Uh × Ph such that for all (vh, qh) ∈Wh × Ph,a(uh,vh)− b(ph,vh) = L(vh),

b(qh,uh) = (s, qh)L2(Ω).

(33)

3.2. Choices of parameters

In the following experiments, we set the external force g to be 0. α and µ are set as displayed in the

Table 1 which are consistent with those in [13]. We will first use a relatively large τ as an initial choice and

decrease it gradually with iterations. The initial τ of all experiments are set to be 0.01. All examples are

computed by Algorithm 2 using the initial shape as shown in Figure 2. The width of the initial shape is

12

Page 13: An efficient iterative method for the formulation of flow

α µ

Fluid Region 0 1.04× 10−3

Solid Region 1.04× 107 1.04× 10−11

Table 1: The parameters α and µ used in numerical experiments. See Section 3.2.

Figure 2: Initial shape for all examples. See Section 3.2.

determined by volume fraction ω. Actually, we observe that with a warm start, initial shape is not important,

even randomized initial shapes can lead same final results.

3.3. Application 1: A constant source

In this application, we consider the case where s = 1 which is a constant source. Part of the boundary

is set to be traction free so that fluids can flows out. Specifically, as shown in Figure 3, we let,

Γt = |x1 − 0.5| ≤ 0.05, x2 = 0,ΓD = ∂Ω− Γt,

uD = 0, t = 0.

This example can be understood as: the domian Ω is continuously creating fluid at each point uniformly,

this fluid then unleashes through part of the boundary, Γt. The domain is consists of fluid region and solid

(porous media) region. The fraction of fluid region is fixed. The flow network is formed to minimize total

energy needed in carrying out produced fluid. This problem has also been studied in [13].

In the implementation, we set the volume fraction ω = 0.3 and use 1024 × 1024 meshes to discretize

the domain. As the problem is symmetric about the line x = 0.5, we run our algorithm in the domain

Ω ∩ x ≥ 0.5 and extend it to the whole Ω symmetrically.x

13

Page 14: An efficient iterative method for the formulation of flow

Figure 3: A diagram for the boundary consisting Γu and Γt. See Section 3.3.

In this situation, we observe that τ plays an important role in the resulting scales of the results. Figure 4

displays the optimal networks obtained from Algorithm 2 with the choices of the smallest τ = 3.9×10−5, 9.8×

10−6, and 2.4 × 10−6, respectively. We observe relatively regular structures for all choices of τ and three

results are consistent with each other. Furthermore, as τ decreases, we observe that it generates more

fractals and the scale gets smaller.

To get some insight into it, we recall that by Algorithm 1, a new shape is obtained by thresholding after

diffusion (convolution with heat kernel) of L2 function at certain value. Namely, the boundaries of these

shapes are the level set of the diffused functions. To be more specific, we write

Wτ = Gτ ∗ f : f ∈ L2(Ω). (34)

Then the boundary of resulting shape is a level set of some function φ ∈ Wτ . The following Proposition

demonstrates the fertility of Wτ increase as τ decrease.

Proposition 3.1. If ν > τ , then Wν $Wτ .

Proof. We can write,

Gν ∗ f = Gτ ∗ (Gν−τ ∗ f),

and obviously Gν−τ ∗ f ∈ L2(Ω), thus Wν ⊂Wτ . On the other hand, if

Gτ ∗ f1 = Gν ∗ f2,

then

Gτ ∗ f1 = Gτ ∗ (Gν−τ ∗ f2),

14

Page 15: An efficient iterative method for the formulation of flow

which implies

f1 = Gν−τ ∗ f2 a.e.

However, the right hand side of last equality is a function of C∞ while the left hand side is just L2. Therefore,

Wν 6= Wτ .

Figure 4: The optimal configuration obtained from Algorithm 2 with a constant source and different choices of the smallest τ .

Left: τ = 3.9 × 10−5. Middle: τ = 9.8 × 10−6. Right: τ = 2.4 × 10−6. See Section 3.3.

3.4. Effect of perimeter and convergence

As we have mentioned, adding the perimeter penalty is another way of regularizing solutions. Figure 5

lists the approximate optimal structures for γ = 1 × 102, 1 × 103, 1 × 104. It shows that adding the weight

of the perimeter term limits the growth of braches.

Figure 5: The optimal configuration obtained from Algorithm 2 with a constant source and different choices of the γ. τ =

2.4 × 10−6 for all three cases. Left: γ = 1 × 104. Middle: γ = 1 × 103. Right: γ = 1 × 102. See Section 3.4.

In Section 3.3, we observed that the Gaussian filter (interpolation) puts an effective constraint on the

admissible solutions. However, without perimeter term in the objective functional (γ = 0), different τ ’s lead

15

Page 16: An efficient iterative method for the formulation of flow

to distinct optimal shapes due to the ill-posedness of the problem. The following experiments demonstrate

that the optimization problem with perimeter regularizer will converge as τ tends to zero. We set γ = 1×102

and discretize the domain with 2× 1024× 1024. With different values of τ , all experiments give almost the

same shapes when τ is small (about < 4.9 × 10−6 in this case). Figure 6 displays the structures obtained

via using different τ . We note that the (local) optimal solution to the optimization problem with prescribed

γ can be well expressed under the filter with relatively small τ .

Figure 6: The optimal configuration obtained from Algorithm 2 with a constant source and different choices of the final τ .

γ = 1 × 102 for all three cases. Left: τ = 4.9 × 10−6. Middle: τ = 2.4 × 10−6. Right: τ = 1.2 × 10−6. See Section 3.4.

Besides, we observe that different meshes might lead to slightly different results due to the non-convexity

of the problem. Nevertheless, those optimal shapes obtained from one mesh are the stationary points of

that from another mesh. To see this, we take the final shape obtained from 2 × 512 × 512 mesh as the

initial shape to simulate over 2 × 1024× 1024 mesh. See Figure 7. We can also achieve a similar result on

2× 1024× 1024 meshes with the optimal shape from 2× 512× 512 being the initial data.

Figure 7: The optimal configuration obtained from Algorithm 2 with a constant source and varieties of meshes. γ = 1×102, τ =

4.9×10−6 for all three cases. Left: 2×1024×1024. Middle: 2×512×512. Right: 2×1024×1024, use solution from 2×512×512

as initial. See Section 3.4.

16

Page 17: An efficient iterative method for the formulation of flow

3.5. Application 2: Discrete delta sources

In this example, we consider the case where s is chosen to be a linear combination of discrete delta

sources [9]. We consider the solely zero Dirichlet condition, i.e.; ΓD = ∂Ω,uD = 0. We first write the source

as a weighted summation of several delta functions,

s =

k∑i=1

ckδxk, (35)

where δxkis an unit point measure at xk and ck is chosen such that

k∑i=1

ck = 0, (36)

which preserves the total volume of the fluid and makes Ω a closed system. In this case, the∫

Ωα|u|2dx

term in total potential energy 2 is considered as penalty for having fluid in solid region. The problem can

be viewed as: there are multiple fluid production and consumption locations in the region, one need to

design flow network with fixed volume that minimizing dissipative energy (total potential energy reduces to

dissipation in this case) in transportation.

In the implementation, we discretize the domain Ω by 2 × 512 × 512 triangles. For the numerical

consideration, we admit the Gaussian with a proper choice of variance as an approximation to the delta

function,

δxk(x) ≈ 1

4πle−|x−xk|

2

4l = Gl(x− xk). (37)

For the 2 × 512 × 512 mesh, we take l = 1 × 10−4. With this approximation, (more than 99.95% of) unit

mass is spreading to surrounding 50× 50 grid points. If not stated otherwise, in this application, we let,

ω = 0.1, τ = 4.9× 10−6.

For example in this section, we end iteration if there is no change between two consecutive iterations. As

we will later demonstrate, a much loose convergence criterion can be applied to get a satisfactory result.

To get results that are visually indistinguishable from what we present here, the number of iterations can

be much less than reported in these experiments. Besides, we note that when running Algorithm 2, by one

iteration we mean one iteration in Algorithm 1, i.e. solve (16) and (17) once.

3.5.1. 1 source with 4 sinks

In this section, we consider the case where a source is located at the center of the domain surrounded

by 4 sinks along with different directions (see the left figures in Figures 8 and 9). Specifically, the source is

located at (0.5, 0.5), and the 4 sinks lie equidistantly on the circle centered at the source with a radius of

0.25. The coefficients ck for sinks are set to −1 and thus the corresponding ck for the source is 4. In Figure 9,

17

Page 18: An efficient iterative method for the formulation of flow

the distribution of the sinks is a rotation of that in Figure 8, 30 around the source counterclockwise. Right

figures in Figures 8 and 9 are the optimal flow networks we obtained using the Algorithm 2 after 45 and

54 iterations, respectively. We observe that the configuration is invariant to the rotation, implying that the

results are reliable and self-consistent.

Also, we plot the change of energy with iterations of the first source distribution in Figure 10. It

demonstrates the energy decaying property which was proved in Theorem 2.3. Besides, we mark those steps

at which τ decrease as red. We see significant energy drops of several steps after each τ decrease, followed

by a steady change of energy. Significant shape change is observed after decreasing τ, see Figure 11.

Figure 8: Left: The distribution of the source and sinks. Right: The optimal configuration obtained from Algorithm 2 after 45

iterations. See Section 3.5.1.

Figure 9: Left: The distribution of the source and sinks. Right: The optimal configuration obtained from Algorithm 2 after 54

iterations. See Section 3.5.1.

3.5.2. 1 source with 8 sinks

In this example, as shown in Figure 12, we first change the radius in the example in Section 3.5.1 to

0.15 and then add another 4 sinks (with coefficient ck = 0.2) on the circle with radius 0.3 without rotation

18

Page 19: An efficient iterative method for the formulation of flow

Figure 10: Log total energy versus step. At red points, the τ used in Algorithm 2 is halved. See Section 3.5.1

Figure 11: Left: Shape at τ = 2.5e−3. Right: Shape at τ = 1.3e−3. See Section 3.5.1.

(i.e.; totally 2 sinks at each direction of the N, S, W, and E directions.). Then the coefficient ck at the

source is modified to 4.8 accordingly to satisfy the mass balance condition (36). Figure 12 displays the

optimal flow network we obtained through Algorithm 2 with different choices of volume fractions. We run

the experiments for volume fraction ω = 0.1 and ω = 0.05 with 52 and 70 iterations. As expected, the pipe

gets thinner after fluids passing the sinks on the first level (i.e.; sinks located at the radius 0.15). We also

observe that, as the volume of fluid decreases, the structure gets even thinner which is also consistent with

the expectation.

However, we notice that the shape around the source/sink in the fluid region is unexpected fat. We further

refine the mesh to 2× 1024× 1024 and decrease l in approximation (37) to 2e−5 to study the dependence

of the configuration on the Gaussian approximation. Figure 13 list the results using different Gaussian

approximations. In the middle one, we choose the volume fraction ω = 0.1, and the number of iteration is

19

Page 20: An efficient iterative method for the formulation of flow

Figure 12: Left: The distribution of the source and sinks. Middle: The optimal configuration obtained from Algorithm 2 after

52 iterations with ω = 0.1. Right: The optimal configuration obtained from Algorithm 2 after 70 iterations with ω = 0.05..

See Section 3.5.2.

122. In the right one, we set ω = 0.05 and the algorithm takes 217 iterations to reach the equilibrium shape.

In Figure 13, we observe that there are no unexpected fat regions when using a Gaussian with a smaller

variance, which provides a better approximation to the delta functions. One can also approximate the delta

functions by a piecewise linear hat function, a cosine approximation, a piecewise cubic function, or so on.

We refer to [37] and the references therein for more discussions on the approximation to delta functions.

Figure 13: Left: The distribution of the source and sinks. Middle: The optimal configuration obtained from Algorithm 2 after

122 iterations with ω = 0.1. Right: The optimal configuration obtained from Algorithm 2 after 217 iterations with ω = 0.05.

See Section 3.5.2.

3.5.3. 1 source with 2 sinks

In this example, we consider the source and sinks distributed as shown in Figures 14 and 15. To be

specific, two sinks with a source on their bisection. In Figure 14, the distance of two sinks is 0.2, the

distance from the source to the line of two sinks is 0.4. In Figure 15, the distance of two sinks is 0.4, the

distance from the source to the line of two sinks is 0.2. Strengths at sinks are set to 1. As we can see, when

20

Page 21: An efficient iterative method for the formulation of flow

the angle formed by these three points at the source is small, the network that the flow first merges then

bifurcates is preferred. As the angle gets larger, e.g.; 90 in Figure 14, two channels from source to two

sinks are formed.

Figure 14: Left: The distribution of the source and sinks. Middle: The optimal flow network obtained from Algorithm 2 after

158 iterations with ω = 0.1. Right: The optimal flow network obtained from Algorithm 2 after 77 iterations with ω = 0.05.

See Section 3.5.3.

Figure 15: Left: The distribution of the source and sinks. Middle: The optimal flow network obtained from Algorithm 2 after

105 iterations with ω = 0.1. Right: The optimal flow network obtained from Algorithm 2 after 114 iterations with ω = 0.05.

See Section 3.5.3.

3.5.4. Tree sturcture

In this example, we consider the source and sinks distributed as shown in Figure 16. Specifically, we put

three sinks with strength 1 (i.e.; ck = 1) surrounding a source at (0.5, 0.5) with distance 0.15, symmetrically.

For each sink, we put another two sources around it at a distance of 0.1 with a strength of 0.3. The strength

at the central source is 1.2 from a direct calculation using the balance condition (36). Figure 16 displays

the optimal shape obtained by Algorithm 2 after 78 iterations. It is consistent with the distribution of the

source and sinks. Because at such choice of the locations and setup of the strength at the source and sinks,

21

Page 22: An efficient iterative method for the formulation of flow

the flux along each edge between source and sink is the same. In Figure 16, we observe that all branches

are almost at the same scale, which is consistent with the setup and the expectation.

Figure 16: Left: The distribution of the source and sinks. Right: The optimal configuration obtained from Algorithm 2 after

78 iterations. See Section 3.5.4.

4. Conclusions and discussions

In this paper, we proposed an efficient iterative convolution thresholding method for the topology op-

timization for fluids with sources. We theoretically proved that this method enjoys the energy decaying

property in each iteration. In numerical experiments, we considered two general cases; i) the fluid is uni-

formly created all over the space and flows out through part of the boundary; ii) the sources and sinks are

described by discrete delta functions with zero boundary condition. A variety of numerical examples were

shown to verify the performance, efficiency, and energy-decaying property of the method. We observed in-

teresting fractal structures and the dependence of the configurations on parameters. It would be interesting

to have more theoretical studies on flow networks with more complicated structures.

Acknowledgement

D. Wang acknowledges support from National Natural Science Foundation of China grant 12101524 and

the University Development Fund from The Chinese University of Hong Kong, Shenzhen (UDF01001803).

X.-P. Wang was supported in part by the Hong Kong Research Grants Council (GRF grants 16302715,

16324416, 16303318 and NSFC-RGC joint research grant N-HKUST620/15). The authors would like to

thank Dan Hu for introducing the model where the source and sinks are discrete delta functions.

References

[1] M. Barthelemy, Spatial networks, Physics Reports 499 (1-3) (2011) 1–101.

22

Page 23: An efficient iterative method for the formulation of flow

[2] D. Hu, D. Cai, Adaptation and optimization of biological transport networks, Physical Review Letters 111 (13).

[3] R. Thoma, Untersuchungen uber die Histogenese und Histomechanik des Gefasssystems, Enke, 1893.

[4] C. D. Murray, The physiological principle of minimum work: I. the vascular system and the cost of blood volume,

Proceedings of the National Academy of Sciences of the United States of America 12 (3) (1926) 207.

[5] O. Sigmund, J. Petersson, Numerical instabilities in topology optimization: a survey on procedures dealing with checker-

boards, mesh-dependencies and local minima, Structural optimization 16 (1) (1998) 68–75.

[6] M. X. Goemans, Y.-s. Myung, A catalog of Steiner tree formulations, Networks 23 (1) (1993) 19–28.

[7] A. Tero, S. Takagi, T. Saigusa, K. Ito, D. P. Bebber, M. D. Fricker, K. Yumiki, R. Kobayashi, T. Nakagaki, Rules for

biologically inspired adaptive network design, Science 327 (5964) (2010) 439–442.

[8] A. Tero, T. Nakagaki, K. Toyabe, K. Yumiki, R. Kobayashi, A method inspired by physarum for solving the steiner

problem., International journal of unconventional Computing 6 (2).

[9] D. Hu, D. Cai, An optimization principle for initiation and adaptation of biological transport networks, Communications

in Mathematical Sciences 17 (5) (2019) 1427–1436.

[10] G. Albi, M. Burger, J. Haskovec, P. Markowich, M. Schlottbom, Continuum modeling of biological network forma-

tion, in: Active particles. Vol. 1. Advances in theory, models, and applications, Model. Simul. Sci. Eng. Technol.,

Birkhauser/Springer, Cham, 2017, pp. 1–48.

[11] J. Haskovec, P. Markowich, B. Perthame, M. Schlottbom, Notes on a PDE system for biological network formation,

Nonlinear Anal. 138 (2016) 127–155.

[12] J. Haskovec, P. Markowich, B. Perthame, Mathematical analysis of a PDE system for biological network formation, Comm.

Partial Differential Equations 40 (5) (2015) 918–956.

[13] N. Wiker, A. Klarbring, T. Borrvall, Topology optimization of regions of darcy and stokes flow, International Journal for

Numerical Methods in Engineering 69 (7) (2007) 1374–1404.

[14] K. Svanberg, The method of moving asymptotes—a new method for structural optimization, International Journal for

Numerical Methods in Engineering 24 (2) (1987) 359–373.

[15] M. P. Bendsøe, N. Kikuchi, Generating optimal topologies in structural design using a homogenization method, Computer

Methods in Applied Mechanics and Engineering 71 (2) (1988) 197–224.

[16] M. Y. Wang, X. Wang, D. Guo, A level set method for structural topology optimization, Computer Methods in Applied

Mechanics and Engineering 192 (1-2) (2003) 227–246.

[17] R. Picelli, S. Townsend, C. Brampton, J. Norato, H. Kim, Stress-based shape and topology optimization with the level

set method, Computer Methods in Applied Mechanics and Engineering 329 (2018) 1–23.

[18] O. Sigmund, K. Hougaard, Geometric properties of optimal photonic crystals, Physical Review Letters 100 (15).

[19] B. Bourdin, A. Chambolle, Design-dependent loads in topology optimization, ESAIM: Control, Optimisation and Calculus

of Variations 9 (2003) 19–48.

[20] Y. Xie, G. Steven, A simple evolutionary procedure for structural optimization, Computers & Structures 49 (5) (1993)

885–896.

[21] T. Yamada, K. Izui, S. Nishiwaki, A. Takezawa, A topology optimization method based on the level set method in-

corporating a fictitious interface energy, Computer Methods in Applied Mechanics and Engineering 199 (45-48) (2010)

2876–2891.

[22] A. Takezawa, S. Nishiwaki, M. Kitamura, Shape and topology optimization based on the phase field method and sensitivity

analysis, Journal of Computational Physics 229 (7) (2010) 2697–2718.

[23] B. Merriman, J. K. Bence, S. Osher, Diffusion generated motion by mean curvature, uCLA CAM Report 92-18 (1992).

[24] S. Esedoglu, F. Otto, Threshold dynamics for networks with arbitrary surface tensions, Communications on Pure and

Applied Mathematics 68 (5) (2015) 808–864.

23

Page 24: An efficient iterative method for the formulation of flow

[25] S. Esedoglu, Y.-H. R. Tsai, et al., Threshold dynamics for the piecewise constant mumford–shah functional, Journal of

Computational Physics 211 (1) (2006) 367–384.

[26] D. Wang, H. Li, X. Wei, X.-P. Wang, An efficient iterative thresholding method for image segmentation, Journal of

Computational Physics 350 (1) (2017) 657–667.

[27] D. Wang, X.-P. Wang, The iterative convolution-thresholding method (ictm) for image segmentation, arXiv preprint

arXiv:1904.10917.

[28] B. Merriman, S. J. Ruuth, Convolution-generated motion and generalized Huygens’ principles for interface motion, SIAM

Journal on Applied Mathematics 60 (3) (2000) 868–890.

[29] M. Elsey, S. Esedoglu, Threshold dynamics for anisotropic surface energies, Mathematics of Computation 87 (312) (2017)

1721–1756.

[30] X. Xu, D. Wang, X.-P. Wang, An efficient threshold dynamics method for wetting on rough surfaces, Journal of Compu-

tational Physics 330 (1) (2017) 510–528.

[31] D. Wang, X.-P. Wang, X. Xu, An improved threshold dynamics method for wetting dynamics, Journal of Computational

Physics 392 (2019) 291–310.

[32] H. Chen, H. Leng, D. Wang, X.-P. Wang, An efficient threshold dynamics method for topology optimization for fluids,

arXiv preprint arXiv:1812.09437.

[33] F. Brezzi, On the existence, uniqueness and approximation of saddle-point problems arising from lagrangian multipliers,

Publications mathematiques et informatique de Rennes 8 (R2) (1974) 129–151.

[34] I. Babuska, The finite element method with lagrangian multipliers, Numerische Mathematik 20 (3) (1973) 179–192.

[35] F. Brezzi, M. Fortin (Eds.), Mixed and Hybrid Finite Element Methods, Springer New York, 1991.

[36] K. Stewartson, O. A. Ladyzhenskaya, R. A. Silverman, The mathematical theory of viscous incompressible flow, The

Mathematical Gazette 56 (395) (1972) 83.

[37] S. Zahedi, A.-K. Tornberg, Delta function approximations in level set methods by distance function extension, Journal of

Computational Physics 229 (6) (2010) 2199–2219.

24