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

Efficient space-time discretizations for tracking the boundaries of reachable sets

Janosch Rieger    Kyria Wawryk
Abstract

The reachable sets of nonlinear control systems can in general only be numerically approximated, and are often very expensive to calculate. In this paper, we propose an algorithm that tracks only the boundaries of the reachable sets and that chooses the temporal and spatial discretizations in a non-uniform way to reduce the computational complexity.

Mathematics Subject Classification: 93B03, 65L50.

Keywords: Reachable sets of control systems; Numerical approximation; Boundary tracking; Efficient discretization; Euler’s method.

1 Introduction

The reachable set of a control system represents all possible states the system can attain at a given time. Reachable sets have applications in several different fields, which are often, but not always, related to safety and security [5, 6, 13]. Since explicit formulas for reachable sets are not usually available, we instead must rely on numerical approximation. These approximations are expensive to compute as they need to track an entire set that evolves according to a multivalued flow. In particular, they require a spatial discretization of the set as well as a temporal discretization of the dynamics.

The numerical methods for the computation of reachable sets are split into those which are designed specifically for linear control systems, and those which are applicable to nonlinear systems.

The reachable sets of linear systems are convex and behave in a rather straightforward way. This enables the creation of numerical methods which are theoretically robust and perform well in practice, taking advantage of the fact that the sets can be approximated by relatively simple shapes such as zonotopes [1, 7, 19] or ellipsoids [10], or by a finite number of support points [2, 11].

The shape and behavior of reachable sets of nonlinear control systems is more complex, and so it is unclear how they can be discretized and evolved in the most efficient way. For this reason, several different families of numerical methods have been proposed.

One such family applies the linear methods discussed above to local linearizations of nonlinear systems. Due to the linearization errors, the performance of these methods can depend on the degree of their nonlinearity (see [18] for an analysis).

Another family of algorithms, which are based on optimal control [3, 14], can significantly improve practical performance at the expense of losing guaranteed convergence. These methods use solvers for optimal control problems to approximate the reachable set at a particular time, and thus eliminate the need for spatial discretization at intermediate timesteps at the risk of getting caught in local minima and losing a significant part of the set.

Finally, Runge-Kutta methods for nonlinear differential equations can be generalized to reachable set approximation. Typically, first-order methods are used, as even second-order convergence has prohibitively strict requirements on the control system [20, 21]. Spatial discretization for these methods generally involves representing the reachable set as a subset of a grid [4, 9]. Since these subsets can become very large, storage and computational cost can be reduced by an intelligent choice of discretization [17], or by considering only the boundaries of the reachable sets [15].

In this paper, we aim to combine the benefits of the algorithms proposed in [15] and [17] in a new numerical method for reachable set computation. The boundary tracking algorithm from [15] is essentially a fully discretized set-valued Euler scheme, but it reduces computational complexity by evolving only the boundary of the reachable set in time, and evaluating only the boundary of the multivalued vector field of the control system. The adaptive refinement strategy pursued in [17] aims to identify a space-time discretization with a given a priori error tolerance that minimizes the computational cost of the set-valued Euler scheme. To develop an algorithm that possesses both advantages, we proceed as follows.

After setting our notation and introducing reachable sets and their Euler approximations in Sections 2 and 3, we first have to overcome two technical obstacles in Section 4: The boundary tracking algorithm as published in [15] stores boundaries of discrete reachable sets in an ad-hoc fashion, without exploring the space of all such boundaries from a theoretical perspective. In order to take the algorithm further, we must do so now, and embed it into the framework of [16] for working with boundaries of discrete sets. Furthermore, it turns out that a naive generalization of the algorithm to nonuniform meshes fails, and we introduce appropriate modifications.

In Section 5, we develop an abstract scheme that refines a given space-time discretization until a prescribed error tolerance is met. It is a greedy algorithm in the sense that every refinement step subdivides the time interval where the quotient of the decrease in error and the increase in computational cost is maximized. In Section 6, we derive a concrete estimator for the computational cost of the boundary tracking algorithm for nonuniform discretizations, and prove that it meets the requirements of the abstract refinement scheme. The resulting algorithm is conceptually similar to the method from [17]. It alternates between computing approximate boundaries of the reachable sets for a given space-time discretization and, based on this knowledge of their geometry, refining the mesh. We demonstrate that this process converges in finite time.

Finally, we present two numerical examples in Section 7, which suggest that our algorithm has the potential to outperform the Euler scheme with uniform discretization from [4], the boundary tracking algorithm with uniform discretization from [15] and the Euler scheme with adaptive discretization from [17].

We consciously break away from the tradition to keep algorithms close to the position in the document where they are mentioned. Instead, we display them grouped together on pages 1 to 4, which allows the reader to retrace how the building blocks of the overall algorithm communicate.

2 Notation

We denote 0:={0,1,2,}\mathbb{N}_{0}:=\{0,1,2,\ldots\}, 1={1,2,3,}\mathbb{N}_{1}=\{1,2,3,\ldots\}, 2={2,3,4,}\mathbb{N}_{2}=\{2,3,4,\ldots\} and :={1,0,1,}\mathbb{Z}:=\{\ldots-1,0,1,\ldots\}. When JJ\subset\mathbb{R} is an interval, and when it is clear that kk\in\mathbb{N} or kk\in\mathbb{Z}, we will write kJk\in J instead of kJk\in J\cap\mathbb{N} or kJk\in J\cap\mathbb{Z}. In addition to the usual dd-dimensional vector space d\mathbb{R}^{d}, we introduce the space

0,d:={(x0,x1,,xd):x0,x1,,xd}\mathbb{R}^{0,d}:=\{(x_{0},x_{1},\ldots,x_{d}):x_{0},x_{1},\ldots,x_{d}\in\mathbb{R}\}

and define the cumulative sum Σ+:d0,d\Sigma_{+}:\mathbb{R}^{d}\to\mathbb{R}^{0,d} by (Σ+x)k:=j=1kxj\left(\Sigma_{+}x\right)_{k}:=\sum_{j=1}^{k}x_{j} for all k[0,d]k\in[0,d]. We also denote

+d:={xd:xi0i[1,d]},>0d:={xd:xi>0i[1,d]},\mathbb{R}_{+}^{d}:=\{x\in\mathbb{R}^{d}:x_{i}\geq 0\ \forall\,i\in[1,d]\},\quad\mathbb{R}_{>0}^{d}:=\{x\in\mathbb{R}^{d}:x_{i}>0\ \forall\,i\in[1,d]\},

and we extend this notation to 0,d\mathbb{R}^{0,d} in the obvious way. We equip d\mathbb{R}^{d} with

:d+given byx:=maxi[1,d]|xi|,\|\cdot\|_{\infty}:\mathbb{R}^{d}\to\mathbb{R}_{+}\quad\text{given by}\quad\|x\|_{\infty}:=\max_{i\in[1,d]}|x_{i}|,

and we define the ball with center xdx\in\mathbb{R}^{d} and radius r>0r>0 by

Br(x):={yd:xyr}.B_{r}(x):=\{y\in\mathbb{R}^{d}:\|x-y\|_{\infty}\leq r\}.

The cardinality of a set MdM\subset\mathbb{R}^{d} is denoted #M\#M, and its boundary M\partial M.

Let 𝒦(d)\mathcal{K}(\mathbb{R}^{d}) and 𝒦𝒞(d)\mathcal{KC}(\mathbb{R}^{d}) denote the collections of all nonempty and compact subsets of d\mathbb{R}^{d}, and all nonempty, compact and convex subsets of d\mathbb{R}^{d}, respectively. The Hausdorff distance distH:𝒦(d)×𝒦(d)+\operatorname{dist}_{H}:\mathcal{K}(\mathbb{R}^{d})\times\mathcal{K}(\mathbb{R}^{d})\to\mathbb{R}_{+} is given by

distH(M,M~)=max{dist(M,M~),dist(M~,M)},\operatorname{dist}_{H}(M,\tilde{M})=\max\{\operatorname{dist}(M,\tilde{M}),\,\operatorname{dist}(\tilde{M},M)\},

where

dist(M,M~)=supxMinfx~M~xx~.\operatorname{dist}(M,\tilde{M})=\sup_{x\in M}\inf_{\tilde{x}\in\tilde{M}}\|x-\tilde{x}\|_{\infty}.

To approximate 𝒦(d)\mathcal{K}(\mathbb{R}^{d}), for every ρ>0\rho>0, we consider the grid Δρ:=ρd\Delta_{\rho}:=\rho\mathbb{Z}^{d} and the collections of digital images

Sρ:={MΔρ:M}andSρ+:=2Δρ.S_{\rho}:=\{M\subset\Delta_{\rho}:M\neq\emptyset\}\quad\text{and}\quad S_{\rho}^{+}:=2^{\Delta_{\rho}}.

By CρC_{\rho} we denote the collection of all sets MSρM\in S_{\rho}, i.e. sets with the property that for any x,zMx,z\in M there exists a finite sequence (yk)k=0n(y_{k})_{k=0}^{n} with y0=xy_{0}=x, yn=zy_{n}=z and yk+1yk=ρ\|y_{k+1}-y_{k}\|_{\infty}=\rho for all k[0,n1]k\in[0,n-1]. These are called chain-connected sets in [15, Definition 9] and 88-connected sets (in the plane) in [8, Section 2.2.1].

Finally, a set-valued map F:d𝒦(d)F:\mathbb{R}^{d}\to\mathcal{K}(\mathbb{R}^{d}) is called LL-Lipschitz with L+L\in\mathbb{R}_{+} if

distH(F(x),F(y))Lxyx,yd,\operatorname{dist}_{H}(F(x),F(y))\leq L\|x-y\|_{\infty}\quad\forall\,x,y\in\mathbb{R}^{d},

and when MdM\subset\mathbb{R}^{d}, we denote F(M):=xMF(x)F(M):=\cup_{x\in M}F(x).

3 Reachable sets and their approximations

In the following, we briefly introduce reachable sets and a common approach to their numerical computation.

3.1 Reachable sets of differential inclusions

We consider the differential inclusion

x˙(t)F(x(t))for almost everyt(0,T),x(0)X0,\dot{x}(t)\in F(x(t))\ \text{for almost every}\ t\in(0,T),\quad x(0)\in X_{0}, (1)

where X0𝒦(d)X_{0}\in\mathcal{K}(\mathbb{R}^{d}) is path-connected, and F:d𝒦𝒞(d)F:\mathbb{R}^{d}\to\mathcal{KC}(\mathbb{R}^{d}) is LL-Lipschitz and satisfies the uniform bound

supxdsupfF(x)fP.\sup_{x\in\mathbb{R}^{d}}\sup_{f\in F(x)}\|f\|_{\infty}\leq P. (2)

We are interested in the following sets associated with inclusion (1).

Definition 3.1.

The solution set and the reachable sets of inclusion (1) are

𝒮F(T)\displaystyle\mathcal{S}^{F}(T) :={x()W1,1([0,T],d):x()solves inclusion (1)},\displaystyle:=\{x(\cdot)\in W^{1,1}([0,T],\mathbb{R}^{d}):x(\cdot)\ \text{solves inclusion \eqref{eq:ODI}}\},
F(t)\displaystyle\mathcal{R}^{F}(t) :={x(t):x()𝒮F(T)},t[0,T].\displaystyle:=\{x(t):x(\cdot)\in\mathcal{S}^{F}(T)\},\quad t\in[0,T].

3.2 The Euler scheme for uniform grids

We discretize the space 𝒦(d)\mathcal{K}(\mathbb{R}^{d}) in a way that is similar to outer Jordan digitization, see [8, Definition 2.8].

Lemma 3.2.

For any numbers α,ρ>0\alpha,\rho>0 with αρ/2\alpha\geq\rho/2, the mapping

πρα:𝒦(d)Sρgiven byπρα(M):=Bα(M)Δρ\pi^{\alpha}_{\rho}:\mathcal{K}(\mathbb{R}^{d})\to S_{\rho}\quad\text{given by}\quad\pi^{\alpha}_{\rho}(M):=B_{\alpha}(M)\cap\Delta_{\rho}

is well-defined, and for all path-connected M𝒦(d)M\in\mathcal{K}(\mathbb{R}^{d}), we have πρα(M)Cρ\pi^{\alpha}_{\rho}(M)\in C_{\rho}.

The Euler map approximates the flow of inclusion (1).

Definition 3.3.

For any given α,h,ρ>0\alpha,h,\rho>0, we define the set-valued functions ϕh:d𝒦𝒞(d)\phi_{h}:\mathbb{R}^{d}\to\mathcal{KC}(\mathbb{R}^{d}) and Φh,ρα:dSρ+\Phi_{h,\rho}^{\alpha}:\mathbb{R}^{d}\to S_{\rho}^{+} by

ϕh(x):=x+hF(x)andΦh,ρα(x):=πρα(ϕh(x)).\displaystyle\phi_{h}(x):=x+hF(x)\quad\text{and}\quad\Phi_{h,\rho}^{\alpha}(x):=\pi^{\alpha}_{\rho}(\phi_{h}(x)).

The following proposition is nearly identical to [15, Proposition 12] and is essential for the boundary tracking algorithm to be discussed in Section 4 to work.

Proposition 3.4.

Let h,ρ>0h,\rho>0, and define

α(h,ρ):=(1+Lh)ρ2.\alpha(h,\rho):=(1+Lh)\tfrac{\rho}{2}. (3)

Then for any MCρM\in C_{\rho} and αα(h,ρ)\alpha\geq\alpha(h,\rho), we have Φh,ρα(M)Cρ\Phi_{h,\rho}^{\alpha}(M)\in C_{\rho}.

The following Euler scheme for approximating the reachable sets from Definition 3.1 employs a uniform discretization: Every timestep has length h>0h>0, and every point computed is an element of Δρ\Delta_{\rho}.

Definition 3.5.

Let n1n\in\mathbb{N}_{1}, let ρ>0\rho>0, and let h:=T/nh:=T/n. Then the set of all Euler approximations to inclusion (1) is

𝒮h,ρF(T):={(yk)k=0n:y0πρρ/2(X0),ykΦh,ρα(h,ρ)(yk1)k[1,n]},\mathcal{S}_{h,\rho}^{F}(T):=\{(y_{k})_{k=0}^{n}:y_{0}\in\pi^{\rho/2}_{\rho}(X_{0}),\ y_{k}\in\Phi^{\alpha(h,\rho)}_{h,\rho}(y_{k-1})\ \forall\,k\in[1,n]\},

and the corresponding approximate reachable sets are

h,ρF(k):={yk:(yj)j=0n𝒮h,ρF(T)},k[0,n].\mathcal{R}_{h,\rho}^{F}(k):=\{y_{k}:(y_{j})_{j=0}^{n}\in\mathcal{S}_{h,\rho}^{F}(T)\},\quad k\in[0,n]. (4)

The discrete solution set Sh,ρF(T)S_{h,\rho}^{F}(T) is only used for theoretical purposes. In practice, the discrete reachable sets are computed using the recursion

h,ρF(0)\displaystyle\mathcal{R}_{h,\rho}^{F}(0) =πρρ/2(X0)\displaystyle=\pi^{\rho/2}_{\rho}(X_{0}) (5)
h,ρF(k)\displaystyle\mathcal{R}_{h,\rho}^{F}(k) =Φh,ρα(h,ρ)(h,ρF(k1))k[1,n].\displaystyle=\Phi^{\alpha(h,\rho)}_{h,\rho}(\mathcal{R}_{h,\rho}^{F}(k-1))\ \forall\,k\in[1,n].

Since X0X_{0} is path-connected, and in view of Lemma 3.2 and Proposition 3.4, we have h,ρF(k)Cρ\mathcal{R}_{h,\rho}^{F}(k)\in C_{\rho} for all k[0,n]k\in[0,n].

4 Boundary tracking for nonuniform grids

In this section we introduce a storage format for boundaries of digital images, and summarize how the algorithm from [15] tracks the boundaries of the reachable sets from (4). Finally, we show how the restriction operator from [16] helps to overcome technical difficulties that arise when nonuniform discretizations are considered.

4.1 Boundary pairs of digital images

We restate the definition of boundary layers of digital images from [15]. They are related to the so-called distance transform [8, Section 3.4.2].

Definition 4.1.

For every ρ>0\rho>0 and MΔρM\subset\Delta_{\rho}, we define

ρ0M\displaystyle\partial_{\rho}^{0}M :={xM:zMcΔρwithxz=ρ},\displaystyle:=\{x\in M:\exists\,z\in M^{c}\cap\Delta_{\rho}\ \text{with}\ \|x-z\|_{\infty}=\rho\},
ρjM\displaystyle\partial_{\rho}^{j}M :={xMcΔρ:dist(x,ρ0M)=jρ},j1,\displaystyle:=\{x\in M^{c}\cap\Delta_{\rho}:\operatorname{dist}(x,\partial_{\rho}^{0}M)=j\rho\},\quad j\in\mathbb{N}_{1},
ρjM\displaystyle\partial_{\rho}^{-j}M :={xM:dist(x,ρ0M)=jρ},j1.\displaystyle:=\{x\in M:\operatorname{dist}(x,\partial_{\rho}^{0}M)=j\rho\},\quad j\in\mathbb{N}_{1}.

The following theorem summarizes the relevant results from [16]:

  • i)

    Every digital image, i.e. every element of MSρM\in S_{\rho}, is uniquely characterized by the pair (ρ0M,ρ1M)(\partial^{0}_{\rho}M,\partial^{1}_{\rho}M).

  • ii)

    There exists a well-defined mapping Rρρ\partial R_{\rho}^{\rho^{\prime}} that maps the boundary pair of a set MSρM\in S_{\rho} to the boundary pair of a metrically (6) and topologically (7) similar set Rρρ(M)SρR_{\rho}^{\rho^{\prime}}(M)\in S_{\rho^{\prime}} that is discretized at a different resolution.

For an illustration of the theorem see the first diagram in Figure 1.

Theorem 4.2.

For every ρ>0\rho>0, the mapping

Trρ:SρSρ+×Sρ+,Trρ(M):=(ρ0M,ρ1M),\operatorname{Tr}_{\rho}:S_{\rho}\to S_{\rho}^{+}\times S_{\rho}^{+},\quad\operatorname{Tr}_{\rho}(M):=(\partial^{0}_{\rho}M,\partial^{1}_{\rho}M),

is injective, and we write bdρ:=Trρ(Sρ)\operatorname{bd}_{\rho}:=\operatorname{Tr}_{\rho}(S_{\rho}) and bdcρ:=Trρ(Cρ)\operatorname{bdc}_{\rho}:=\operatorname{Tr}_{\rho}(C_{\rho}).

For nested grids ΔρΔρ\Delta_{\rho^{\prime}}\subsetneq\Delta_{\rho} with ρρ2\rho^{\prime}\in\rho\mathbb{N}_{2}, the operator Rρρ:SρSρR_{\rho}^{\rho^{\prime}}:S_{\rho}\to S_{\rho^{\prime}} given by Rρρ(M):=Bρ/2(M)ΔρR_{\rho}^{\rho^{\prime}}(M):=B_{\rho^{\prime}/2}(M)\cap\Delta_{\rho^{\prime}} satisfies

MSρ:distH(Rρρ(M),M)ρ/2,\displaystyle\forall\,M\in S_{\rho}:\quad\operatorname{dist}_{H}(R_{\rho}^{\rho^{\prime}}(M),M)\leq\rho^{\prime}/2, (6)
MCρ:Rρρ(M)Cρ,\displaystyle\forall\,M\in C_{\rho}:\quad R_{\rho}^{\rho^{\prime}}(M)\in C_{\rho^{\prime}}, (7)

and the lifted operator

Rρρ:bdρbdρ,Rρρ:=TrρRρρTrρ1\partial R_{\rho}^{\rho^{\prime}}:\operatorname{bd}_{\rho}\to\operatorname{bd}_{\rho^{\prime}},\quad\partial R_{\rho}^{\rho^{\prime}}:=\operatorname{Tr}_{\rho^{\prime}}\circ R_{\rho}^{\rho^{\prime}}\circ\operatorname{Tr}_{\rho}^{-1} (8)

is well-defined.

The mapping Rρρ\partial R_{\rho}^{\rho^{\prime}} can be implemented efficiently, see [16, Algorithm 1].

4.2 Boundary tracking for uniform grids

In view of Proposition 3.4, the operator

Ψh,ρ:bdcρbdcρ,Ψh,ρ(D):=TrρΦh,ρα(h,ρ)Trρ1(D)\Psi_{h,\rho}:\operatorname{bdc}_{\rho}\to\operatorname{bdc}_{\rho},\quad\Psi_{h,\rho}(D):=\operatorname{Tr}_{\rho}\circ\,\Phi^{\alpha(h,\rho)}_{h,\rho}\circ\operatorname{Tr}_{\rho}^{-1}(D) (9)

with α(h,ρ)\alpha(h,\rho) as in (3) is well-defined. It completes the second diagram in Figure 1. The boundary tracking algorithm from [15] uses Ψh,ρ\Psi_{h,\rho} to lift recursion (5), which evolves the full reachable sets, to a recursion

Trρ(h,ρF(0))\displaystyle\operatorname{Tr}_{\rho}(\mathcal{R}_{h,\rho}^{F}(0)) =Trρ(πρρ/2(X0)),\displaystyle=\operatorname{Tr}_{\rho}(\pi^{\rho/2}_{\rho}(X_{0})),
Trρ(h,ρF(k))\displaystyle\operatorname{Tr}_{\rho}(\mathcal{R}_{h,\rho}^{F}(k)) =Ψh,ρ(Trρ(h,ρF(k1)))k[1,n].\displaystyle=\Psi_{h,\rho}(\operatorname{Tr}_{\rho}(\mathcal{R}_{h,\rho}^{F}(k-1)))\quad\forall\,k\in[1,n].

Since X0X_{0} is path-connected, it follows from Lemma 3.2 and (9) that

Trρ(h,ρF(k))bdcρk[0,n].\operatorname{Tr}_{\rho}(\mathcal{R}_{h,\rho}^{F}(k))\in\operatorname{bdc}_{\rho}\quad\forall\,k\in[0,n]. (10)

bdρ{\operatorname{bd}_{\rho}}bdρ{\operatorname{bd}_{\rho^{\prime}}}Sρ{S_{\rho}}Sρ{S_{\rho^{\prime}}}Rρρ\scriptstyle{\partial R_{\rho}^{\rho^{\prime}}}Trρ\scriptstyle{\operatorname{Tr}_{\rho}}Rρρ\scriptstyle{R_{\rho}^{\rho^{\prime}}}Trρ\scriptstyle{\operatorname{Tr}_{\rho^{\prime}}}

bdcρ{\operatorname{bdc}_{\rho}}bdcρ{\operatorname{bdc}_{\rho}}Cρ{C_{\rho}}Cρ{C_{\rho}}Ψh,ρ\scriptstyle{\Psi_{h,\rho}}Trρ\scriptstyle{\operatorname{Tr}_{\rho}}Φh,ρα(h,ρ)\scriptstyle{\Phi^{\alpha(h,\rho)}_{h,\rho}}Trρ\scriptstyle{\operatorname{Tr}_{\rho}}

bdcρ{\operatorname{bdc}_{\rho}}bdcρ{\operatorname{bdc}_{\rho^{\prime}}}Cρ{C_{\rho}}Cρ{C_{\rho^{\prime}}}Ψh,ρ,ρ\scriptstyle{\Psi_{h,\rho,\rho^{\prime}}}Trρ\scriptstyle{\operatorname{Tr}_{\rho}}Φh,ρ,ρα(h,ρ)\scriptstyle{\Phi^{\alpha(h,\rho)}_{h,\rho,\rho^{\prime}}}Trρ\scriptstyle{\operatorname{Tr}_{\rho^{\prime}}}

Figure 1: Mathematically equivalent computations carried out either with entire sets (bottom row) or with their boundaries (top row).

The following theorem is the main result from [15], which allows us to evaluate Ψh,ρ(D)\Psi_{h,\rho}(D) directly, i.e. without knowledge of Trρ1(D)\operatorname{Tr}_{\rho}^{-1}(D). Statement (10) is essential for its proof, which relies on topological arguments.

Theorem 4.3.

Let h(0,14L]h\in(0,\frac{1}{4L}] and ρ>0\rho>0, denote α=α(h,ρ)\alpha=\alpha(h,\rho) from (3) and

κ=κ(h,ρ):=2(1+Lh)(α1Lh+ρ),\kappa=\kappa(h,\rho):=2(1+Lh)(\tfrac{\alpha}{1-Lh}+\rho), (11)

and let D=(D0,D1)bdcρD=(D_{0},D_{1})\in\operatorname{bdc}_{\rho}. Furthermore, consider the set-valued mappings ϕh:d𝒦(d)\phi_{h}^{\partial}:\mathbb{R}^{d}\to\mathcal{K}(\mathbb{R}^{d}) and Φh,ρ,α:dSρ+\Phi_{h,\rho}^{\partial,\alpha}:\mathbb{R}^{d}\to S_{\rho}^{+} given by

ϕh(x):=ϕh(x)andΦh,ρ,α(x):=πρα(ϕh(x)),\phi_{h}^{\partial}(x):=\partial\phi_{h}(x)\quad\text{and}\quad\Phi_{h,\rho}^{\partial,\alpha}(x):=\pi_{\rho}^{\alpha}(\phi_{h}^{\partial}(x)),

and the sets

D1:=ρ1(Tr1(D))andD2:=ρ2(Tr1(D)).D_{-1}:=\partial_{\rho}^{-1}(\operatorname{Tr}^{-1}(D))\quad\text{and}\quad D_{2}:=\partial_{\rho}^{2}(\operatorname{Tr}^{-1}(D)). (12)

Then we can form the auxiliary collections

W00:=Φh,ρ,α+κ(D0)Φh,ρα(D0),W1=Φh,ρ,α(D1D2),W01:=Φh,ρ,α(D1),\begin{gathered}W_{0}^{0}:=\Phi^{\partial,\alpha+\kappa}_{h,\rho}(D_{0})\cap\Phi^{\alpha}_{h,\rho}(D_{0}),\quad W_{1}=\Phi^{\partial,\alpha}_{h,\rho}(D_{1}\cup D_{2}),\\ W_{0}^{-1}:=\Phi^{\partial,\alpha}_{h,\rho}(D_{-1}),\end{gathered} (13)

and we can express Ψh,ρ(D)=D~\Psi_{h,\rho}(D)=\tilde{D} with the pair D~=(D~0,D~1)\tilde{D}=(\tilde{D}_{0},\tilde{D}_{1}) given by

D~1\displaystyle\tilde{D}_{1} :={xW1:dist(x,W00W01)=ρ},\displaystyle:=\{x\in W_{1}:\operatorname{dist}(x,W_{0}^{0}\cup W_{0}^{-1})=\rho\},
D~0\displaystyle\tilde{D}_{0} :={xW00W01:dist(x,D~1)=ρ}\displaystyle:=\{x\in W_{0}^{0}\cup W_{0}^{-1}:\operatorname{dist}(x,\tilde{D}_{1})=\rho\}

in terms of D1,D0,D1D_{-1},\ D_{0},\ D_{1}, and D2D_{2}.

Theorem 4.3 is the theoretical foundation for Algorithm 1, which is displayed on page 1. Note that it is easy and inexpensive to compute the sets in (12). Figure 2 provides a sketch of the sets used to calculate (13).

(a) ϕh(x)\phi_{h}(x) (gray), and ϕh(x)\phi_{h}^{\partial}(x) (dashed)
2α+κ2\alpha+\kappa
(b) Blowup required for xD0x\in D_{0}
2α2\alpha
(c) Blowup required for xDkx\in D_{k}, k{1,1,2}k\in\{-1,1,2\}
Figure 2: Sets used in the calculation of W00,W01W_{0}^{0},\ W_{0}^{-1} and W1W_{1} in equations (13) and (18).

4.3 An Euler scheme for nonuniform grids

We present a numerical method that generalizes the Euler scheme from Section 3.2 to nonuniform temporal and spatial discretization parameters

𝜹:=(𝒉,𝒕,𝝆)>0n×+0,n×>00,n,{\bm{\delta}}:=(\bm{h},\bm{t},\bm{\rho})\in\mathbb{R}^{n}_{>0}\times\mathbb{R}_{+}^{0,n}\times\mathbb{R}_{>0}^{0,n},

and is specifically designed to support the boundary tracking algorithm to be proposed in Section 4.4. The vector 𝒉=(h1,,hn)\bm{h}=(h_{1},\ldots,h_{n}) contains stepsizes with k=1nhk=T\sum_{k=1}^{n}h_{k}=T, the vector 𝒕=(t0,,tn)=+𝒉\bm{t}=(t_{0},\ldots,t_{n})=\sum_{+}\bm{h} stores the temporal nodes, and the vector 𝝆=(ρ0,,ρn)\bm{\rho}=(\rho_{0},\ldots,\rho_{n}) stores the mesh sizes of the spatial grids Δρ0,,Δρn\Delta_{\rho_{0}},\ldots,\Delta_{\rho_{n}} associated with the temporal nodes t0,,tnt_{0},\ldots,t_{n}.

The following mapping is to be interpreted as a single Euler step with stepsize h>0h>0 and blow-up α>0\alpha>0 from the grid Δρ\Delta_{\rho} to the grid Δρ\Delta_{\rho^{\prime}}. The reason for its particular construction will become apparent in Theorem 4.8.

Definition 4.4.

Let α,h,ρ,ρ>0\alpha,h,\rho,\rho^{\prime}>0 satisfy ρ/ρ(0,1)1\rho^{\prime}/\rho\in(0,1)\cup\mathbb{N}_{1} as well as αmin(ρ,ρ)/2\alpha\geq\min(\rho,\rho^{\prime})/2. We define the set-valued function Φh,ρ,ρα:dSρ\Phi_{h,\rho,\rho^{\prime}}^{\alpha}:\mathbb{R}^{d}\to S_{\rho^{\prime}} by

Φh,ρ,ρα(x):={Rρρ(Φh,ρα(x)),ρ<ρ,Φh,ρα(x),ρρ\Phi_{h,\rho,\rho^{\prime}}^{\alpha}(x):=\begin{cases}R_{\rho}^{\rho^{\prime}}(\Phi_{h,\rho}^{\alpha}(x)),&\rho<\rho^{\prime},\\ \Phi_{h,\rho^{\prime}}^{\alpha}(x),&\rho\geq\rho^{\prime}\end{cases}

with Φh,ρα\Phi_{h,\rho}^{\alpha} from Definition 3.3.

Our next result generalizes Proposition 3.4.

Corollary 4.5.

Let h,ρ,ρ>0h,\rho,\rho^{\prime}>0 with ρ/ρ(0,1)1\rho^{\prime}/\rho\in(0,1)\cup\mathbb{N}_{1}, and let α(h,ρ)\alpha(h,\rho) be as in Proposition 3.4. Then for any MCρM\in C_{\rho}, we have Φh,ρ,ρα(h,ρ)Cρ.\Phi_{h,\rho,\rho^{\prime}}^{\alpha(h,\rho)}\in C_{\rho^{\prime}}.

Proof.

When ρ<ρ\rho<\rho^{\prime}, then Φh,ρα(h,ρ)(x)Cρ\Phi_{h,\rho}^{\alpha(h,\rho)}(x)\in C_{\rho} by Proposition 3.4, and by statement (7), we have Rρρ(Φh,ρα(h,ρ)(x))CρR_{\rho}^{\rho^{\prime}}(\Phi_{h,\rho}^{\alpha(h,\rho)}(x))\in C_{\rho^{\prime}}. If ρρ\rho\geq\rho^{\prime}, then α(h,ρ)α(h,ρ)\alpha(h,\rho)\geq\alpha(h,\rho^{\prime}), and hence Φh,ρα(h,ρ)(x)Cρ\Phi_{h,\rho^{\prime}}^{\alpha(h,\rho)}(x)\in C_{\rho^{\prime}} directly by Proposition 3.4. ∎

We generalize Definition 3.5 to nonuniform discretizations.

Definition 4.6.

Let n1n\in\mathbb{N}_{1}, and let 𝛅=(𝐡,𝐭,𝛒)>0n×+0,n×>00,n{\bm{\delta}}=(\bm{h},\bm{t},\bm{\rho})\in\mathbb{R}^{n}_{>0}\times\mathbb{R}_{+}^{0,n}\times\mathbb{R}_{>0}^{0,n} be such that k=1nhk=T\sum_{k=1}^{n}h_{k}=T as well as 𝐭=Σ+𝐡\bm{t}=\Sigma_{+}\bm{h} and ρk+1/ρk(0,1)1\rho_{k+1}/\rho_{k}\in(0,1)\cup\mathbb{N}_{1} for all k[0,n1]k\in[0,n-1]. Then the set of all Euler approximations to inclusion (1) is

𝒮𝜹F(T):={(yk)k=0n:y0πρ0ρ0/2(X0),ykΦhk,ρk1,ρkα(hk,ρk1)(yk1)k[1,n]},\mathcal{S}_{\bm{\delta}}^{F}(T):=\{(y_{k})_{k=0}^{n}:y_{0}\in\pi_{\rho_{0}}^{\rho_{0}/2}(X_{0}),\ y_{k}\in\Phi_{h_{k},\rho_{k-1},\rho_{k}}^{\alpha(h_{k},\rho_{k-1})}(y_{k-1})\ \forall\,k\in[1,n]\},

and the discrete approximation to F(tk)\mathcal{R}^{F}(t_{k}) from Definition 3.1 is given by

𝜹F(k):={yk,y𝒮𝜹F(T)},k[0,n].\mathcal{R}_{\bm{\delta}}^{F}(k):=\{y_{k},y\in\mathcal{S}_{\bm{\delta}}^{F}(T)\},\quad k\in[0,n]. (14)

The proof of the following error bound is very similar to proofs in the literature, see e.g. [4] and [9]. The additive error terms χj(𝜹)\chi_{j}({\bm{\delta}}) occur whenever the operator Rρkρk+1R_{\rho_{k}}^{\rho_{k+1}} is invoked, see Definition 4.4.

Theorem 4.7.

For any n1n\in\mathbb{N}_{1}, any 𝛅:=(𝐡,𝐭,𝛒)>0n×+0,n×>00,n{\bm{\delta}}:=(\bm{h},\bm{t},\bm{\rho})\in\mathbb{R}_{>0}^{n}\times\mathbb{R}_{+}^{0,n}\times\mathbb{R}_{>0}^{0,n} satisfying the requirements of Definition 4.6, and any k[0,n]k\in[0,n], the exact reachable sets of inclusion (1) with property (2) and the discrete reachable sets from Definition 4.6 satisfy

distH(F(tk),𝜹F(k))\displaystyle\operatorname{dist}_{H}(\mathcal{R}^{F}(t_{k}),\mathcal{R}_{\bm{\delta}}^{F}(k)) (15)
ρ02eLtk+j=1keL(tktj)[(eLhj1)(Phj+(1+Lhj)2Lhjρj12+χj(𝜹)]\displaystyle\leq\tfrac{\rho_{0}}{2}e^{Lt_{k}}+\sum_{j=1}^{k}e^{L(t_{k}-t_{j})}[(e^{Lh_{j}}-1)(Ph_{j}+\tfrac{(1+Lh_{j})^{2}}{Lh_{j}}\tfrac{\rho_{j-1}}{2}+\chi_{j}({\bm{\delta}})]

for k[0,n]k\in[0,n], where

χj(𝜹):={ρj2,ρj1<ρj,0,otherwise.\chi_{j}({\bm{\delta}}):=\begin{cases}\tfrac{\rho_{j}}{2},&\rho_{j-1}<\rho_{j},\\ 0,&\text{otherwise}.\end{cases}

As in Section 3.2, the discrete reachable sets (14) satisfy a recursion

𝜹F(0)\displaystyle\mathcal{R}_{\bm{\delta}}^{F}(0) =πρ0ρ0/2(X0),\displaystyle=\pi^{\rho_{0}/2}_{\rho_{0}}(X_{0}), (16)
𝜹F(k)\displaystyle\mathcal{R}_{\bm{\delta}}^{F}(k) =Φhk,ρk1,ρkα(hk,ρk1)(𝜹F(k1))k[1,n].\displaystyle=\Phi^{\alpha(h_{k},\rho_{k-1})}_{h_{k},\rho_{k-1},\rho_{k}}(\mathcal{R}_{\bm{\delta}}^{F}(k-1))\quad\forall\,k\in[1,n].

By Lemma 3.2 and Corollary 4.5, we have 𝜹F(k)Cρk\mathcal{R}_{\bm{\delta}}^{F}(k)\in C_{\rho_{k}} for all k[0,n]k\in[0,n].

4.4 Boundary tracking for nonuniform grids

In view of Corollary 4.5, the mapping

Ψh,ρ,ρ(D):=TrρΦh,ρ,ρα(h,ρ)Trρ1(D)Dbdcρ\Psi_{h,\rho,\rho^{\prime}}(D):=\operatorname{Tr}_{\rho^{\prime}}\circ\Phi_{h,\rho,\rho^{\prime}}^{\alpha(h,\rho)}\circ\operatorname{Tr}_{\rho}^{-1}(D)\quad\forall\,D\in\operatorname{bdc}_{\rho}

is an operator Ψh,ρ,ρ:bdcρbdcρ\Psi_{h,\rho,\rho^{\prime}}:\operatorname{bdc}_{\rho}\to\operatorname{bdc}_{\rho^{\prime}}. Its action is illustrated in the third diagram in Figure 1. It allows us to lift iteration (16) to the recursion

Trρ0(𝜹F(0))\displaystyle\operatorname{Tr}_{\rho_{0}}(\mathcal{R}_{\bm{\delta}}^{F}(0)) =Trρ0(πρ0ρ0/2(X0)),\displaystyle=\operatorname{Tr}_{\rho_{0}}(\pi^{\rho_{0}/2}_{\rho_{0}}(X_{0})), (17)
Trρk(𝜹F(k))\displaystyle\operatorname{Tr}_{\rho_{k}}(\mathcal{R}_{{\bm{\delta}}}^{F}(k)) =Ψhk,ρk1,ρk(Trρk1(𝜹F(k1)))k[1,n],\displaystyle=\Psi_{h_{k},\rho_{k-1},\rho_{k}}(\operatorname{Tr}_{\rho_{k-1}}(\mathcal{R}_{\bm{\delta}}^{F}(k-1)))\quad\forall\,k\in[1,n],

which is displayed on page 2 as Algorithm 2. Again because of Lemma 3.2, we have Trρk(𝜹F(k))bdcρk\operatorname{Tr}_{\rho_{k}}(\mathcal{R}_{{\bm{\delta}}}^{F}(k))\in\operatorname{bdc}_{\rho_{k}} for all k[0,n]k\in[0,n]. Note that these iterates are the boundary pairs of the sets 𝜹F(k)\mathcal{R}_{{\bm{\delta}}}^{F}(k) which satisfy the error bound (15).

To carry out recursion (17) without knowledge of the set h,ρF(k1)\mathcal{R}_{h,\rho}^{F}(k-1), we would ideally like to generalize Theorem 4.3 to nonuniform grids. An inspection of its proof, given in [15], shows that the theorem remains valid when ρk+1ρk\rho_{k+1}\leq\rho_{k}. Elementary counterexamples show, however, that the theorem is false when ρk+1>ρk\rho_{k+1}>\rho_{k}. The construction in Definition 4.4 circumvents this problem, as it allows the representation (19) below.

Theorem 4.8.

Let h(0,14L]h\in(0,\frac{1}{4L}], and let ρ,ρ>0\rho,\rho^{\prime}>0 satisfy ρ/ρ(0,1)1\rho^{\prime}/\rho\in(0,1)\cup\mathbb{N}_{1}. Let α=α(h,ρ)\alpha=\alpha(h,\rho) and κ=κ(h,ρ)\kappa=\kappa(h,\rho) as in (3) and (11), let ρ~:=min(ρ,ρ)\tilde{\rho}:=\min(\rho,\rho^{\prime}), and consider a pair D=(D0,D1)bdcρD=(D_{0},D_{1})\in\operatorname{bdc}_{\rho}. Denoting

D1:=ρ1(Tr1(D))andD2:=ρ2(Tr1(D)),D_{-1}:=\partial_{\rho}^{-1}(\operatorname{Tr}^{-1}(D))\quad\text{and}\quad D_{2}:=\partial_{\rho}^{2}(\operatorname{Tr}^{-1}(D)),

we can form the auxiliary collections

W00:=Φh,ρ~,α+κ(D0)Φh,ρ~α(D0),W1:=Φh,ρ~,α(D1D2),W01:=Φh,ρ~,α(D1)\begin{gathered}W_{0}^{0}:=\Phi^{\partial,\alpha+\kappa}_{h,\tilde{\rho}}(D_{0})\cap\Phi^{\alpha}_{h,\tilde{\rho}}(D_{0}),\quad W_{1}:=\Phi^{\partial,\alpha}_{h,\tilde{\rho}}(D_{1}\cup D_{2}),\\ W_{0}^{-1}:=\Phi^{\partial,\alpha}_{h,\tilde{\rho}}(D_{-1})\end{gathered} (18)

and the pair D~=(D~0,D~1)\tilde{D}=(\tilde{D}_{0},\tilde{D}_{1}) given by D~1:={xW1:dist(x,W00W01)=ρ~}\tilde{D}_{1}:=\{x\in W_{1}:\operatorname{dist}(x,W_{0}^{0}\cup W_{0}^{-1})=\tilde{\rho}\} and D~0:={xW00W01:dist(x,D~1)=ρ~}\tilde{D}_{0}:=\{x\in W_{0}^{0}\cup W_{0}^{-1}:\operatorname{dist}(x,\tilde{D}_{1})=\tilde{\rho}\}. Then we can express

Ψh,ρ,ρ(D)={D~,ρρ,Rρ~ρ(D~),else.\Psi_{h,\rho,\rho^{\prime}}(D)=\begin{cases}\tilde{D},&\rho^{\prime}\leq\rho,\\ \partial R_{\tilde{\rho}}^{\rho^{\prime}}(\tilde{D}),&\text{else}.\end{cases} (19)

in terms of D1D_{-1}, D0D_{0}, D1D_{1} and D2D_{2}, using Rρ~ρ\partial R_{\tilde{\rho}}^{\rho^{\prime}} from (8).

Proof.

A proof identical to that of Theorem 4.3 shows D~bdcρ~\tilde{D}\in\operatorname{bdc}_{\tilde{\rho}} and

Ψh,ρ,ρ~(D)=D~.\displaystyle\Psi_{h,\rho,\tilde{\rho}}(D)=\tilde{D}. (20)

If ρρ\rho^{\prime}\leq\rho, we have ρ~=ρ\tilde{\rho}=\rho^{\prime} and (20) implies (19). If ρ>ρ\rho^{\prime}>\rho, then ρ~=ρ\tilde{\rho}=\rho and using equation (8), Definition 4.4, and equation (20), we verify statement (19) by the computation

Ψh,ρ,ρ(D)=TrρΦh,ρ,ραTrρ1(D)\displaystyle\Psi_{h,\rho,\rho^{\prime}}(D)=\operatorname{Tr}_{\rho^{\prime}}\circ\Phi_{h,\rho,\rho^{\prime}}^{\alpha}\circ\operatorname{Tr}_{\rho}^{-1}(D)
=TrρRρρΦh,ραTrρ1(D)=TrρRρρΦh,ρ,ρ~αTrρ1(D)\displaystyle=\operatorname{Tr}_{\rho^{\prime}}\circ R_{\rho}^{\rho^{\prime}}\circ\Phi_{h,\rho}^{\alpha}\circ\operatorname{Tr}_{\rho}^{-1}(D)=\operatorname{Tr}_{\rho^{\prime}}\circ R_{\rho}^{\rho^{\prime}}\circ\Phi_{h,\rho,\tilde{\rho}}^{\alpha}\circ\operatorname{Tr}_{\rho}^{-1}(D)
=RρρTrρΦh,ρ,ρ~Tr1(D)=RρρΨh,ρ,ρ~(D)=Rρρ(D~).\displaystyle=\partial R_{\rho}^{\rho^{\prime}}\circ\operatorname{Tr}_{\rho}\circ\Phi_{h,\rho,\tilde{\rho}}\circ\operatorname{Tr}^{-1}(D)=\partial R_{\rho}^{\rho^{\prime}}\circ\Psi_{h,\rho,\tilde{\rho}}(D)=\partial R_{\rho}^{\rho^{\prime}}(\tilde{D}).

Algorithm 1 implements the operator Ψh,ρ,ρ\Psi_{h,\rho,\rho^{\prime}} as proposed in Theorem 4.8. In addition to an image Ψh,ρ,ρ(D0,D1)\Psi_{h,\rho,\rho^{\prime}}(D_{0},D_{1}), it returns the numbers c^R1\hat{c}_{R}\in\mathbb{N}_{1} and c^F1\hat{c}_{F}\in\mathbb{N}_{1} of grid points, which can be counted incidentally and are the key ingredients for the surrogate volumes (42) and (43). An implementation of the operator Rρρ\partial R_{\rho}^{\rho^{\prime}}, which occurs in line 1, is available in [16, Algorithm 1].

5 An abstract subdivision scheme

The primary contribution of this paper is Algorithm 4, displayed on page 4, which alternates between executing recursion (17) and refining the discretization based on information gathered in the recursion. This section develops an abstract subdivision scheme for the latter aspect, displayed as Algorithm 3 on page 3, which will be applied to concrete cost and error estimators CC and EE in Section 6.

The following discretizations are admissible for the subdivision process.

Definition 5.1.

By n>0n×+n+1×>0n+1\aleph_{n}\subset\mathbb{R}_{>0}^{n}\times\mathbb{R}_{+}^{n+1}\times\mathbb{R}_{>0}^{n+1} we denote the set of all tuples 𝛅=(𝐡,𝐭,𝛒)>0n×+n+1×>0n+1{\bm{\delta}}=(\bm{h},\bm{t},\bm{\rho})\in\mathbb{R}_{>0}^{n}\times\mathbb{R}_{+}^{n+1}\times\mathbb{R}_{>0}^{n+1} such that

𝒕=Σ+𝒉,T=tn\displaystyle\bm{t}=\Sigma_{+}\bm{h},\quad T=t_{n} (21)
𝒉14L,𝝆P8L\displaystyle\|\bm{h}\|_{\infty}\leq\tfrac{1}{4L},\quad\|\bm{\rho}\|_{\infty}\leq\tfrac{P}{8L} (22)
ρjρk{4:}j,k[0,n],\displaystyle\tfrac{\rho_{j}}{\rho_{k}}\in\{4^{\ell}:\ell\in\mathbb{Z}\}\quad\forall\,j,k\in[0,n], (23)
ρj=2LP(hj)2j[1,n],\displaystyle\rho_{j}=2LP\left(h_{j}\right)^{2}\quad\forall\,j\in[1,n], (24)

Additionally, we denote

:=n1n.\aleph:=\cup_{n\in\mathbb{N}_{1}}\aleph_{n}.

We will refine a given discretization by repeated subdivision according to the following rule.

Definition 5.2.

Given any n1n\in\mathbb{N}_{1} and 𝛅=(𝐡,𝐭,𝛒)n{\bm{\delta}}=(\bm{h},\bm{t},\bm{\rho})\in\aleph_{n}, we define the function

ψ[𝜹;j]=(ψh[𝒉;j],ψt[𝒕;j],ψρ[𝝆;j])\psi[{\bm{\delta}};j]=\left(\psi_{h}[\bm{h};j],\psi_{t}[\bm{t};j],\psi_{\rho}[\bm{\rho};j]\right) (25)

given by

ψh[𝒉;j]\displaystyle\psi_{h}[\bm{h};j] :={𝒉,j=0,(h1,,hj1,hj/2,hj/2,hj+1,,hn),j[1,n],\displaystyle:=\begin{cases}\bm{h},&j=0,\\ (h_{1},...,h_{j-1},h_{j}/2,h_{j}/2,h_{j+1},...,h_{n}),&j\in[1,n],\end{cases} (26)
ψt[𝒕;j]\displaystyle\psi_{t}[\bm{t};j] :={𝒕,j=0,(t0,,tj1,tjhj/2,tj,tj+1,,tn),j[1,n],\displaystyle:=\begin{cases}\bm{t},&j=0,\\ (t_{0},...,t_{j-1},t_{j}-h_{j}/2,t_{j},t_{j+1},...,t_{n}),&j\in[1,n],\end{cases}
ψρ[𝝆;j]\displaystyle\psi_{\rho}[\bm{\rho};j] :={(ρ0/4,ρ1,,ρn),j=0,(ρ0,,ρj1,ρj/4,ρj/4,ρj+1,,ρn),j[1,n].\displaystyle:=\begin{cases}(\rho_{0}/4,\rho_{1},...,\rho_{n}),&j=0,\\ (\rho_{0},...,\rho_{j-1},\rho_{j}/4,\rho_{j}/4,\rho_{j+1},...,\rho_{n}),&j\in[1,n].\end{cases}

Throughout the rest of this section, we consider two functions

C,E:>0C,E:\aleph\to\mathbb{R}_{>0}

that assign a number to every admissible discretization, and the change

ΔC(𝜹;j):=C(ψ[𝜹;j])C(𝜹)n1,𝜹n,j[0,n],\displaystyle\Delta C({\bm{\delta}};j):=C(\psi[{\bm{\delta}};j])-C({\bm{\delta}})\quad\forall\,n\in\mathbb{N}_{1},\ \forall\,{\bm{\delta}}\in\aleph_{n},\ \forall\,j\in[0,n],
ΔE(𝜹;j):=E(ψ[𝜹;j])E(𝜹)n1,𝜹n,j[0,n]\displaystyle\Delta E({\bm{\delta}};j):=E(\psi[{\bm{\delta}};j])-E({\bm{\delta}})\quad\forall\,n\in\mathbb{N}_{1},\ \forall\,{\bm{\delta}}\in\aleph_{n},\ \forall\,j\in[0,n]

in these functions caused by the subdivision rule ψ\psi. We assume that these functions have the following properties:

  1. P1

    There exists a function ω1C0(>0,>0)\omega_{1}\in C^{0}(\mathbb{R}_{>0},\mathbb{R}_{>0}) such that lims0ω1(s)=0\lim_{s\to 0}\omega_{1}(s)=0 and E(𝜹)ω1(𝝆)E({\bm{\delta}})\leq\omega_{1}(\|\bm{\rho}\|_{\infty}) holds for all n1n\in\mathbb{N}_{1} and 𝜹n{\bm{\delta}}\in\aleph_{n}.

  2. P2

    There exists a strictly increasing function ω2C0(+,+)\omega_{2}\in C^{0}(\mathbb{R}_{+},\mathbb{R}_{+}) such that ω2(0)=0\omega_{2}(0)=0 and ΔE(𝜹;j)ω2(ρj)\Delta E({\bm{\delta}};j)\leq-\omega_{2}(\rho_{j}) holds for all n1n\in\mathbb{N}_{1}, all 𝜹n{\bm{\delta}}\in\aleph_{n} and all j[0,n]j\in[0,n].

  3. P3

    We have ΔC(𝜹;argminiρi)>0\Delta C({\bm{\delta}};\arg\min_{i}\rho_{i})>0 for all n1n\in\mathbb{N}_{1} and 𝜹n{\bm{\delta}}\in\aleph_{n}.

  4. P4

    There exists ω4C0(>0,>0)\omega_{4}\in C^{0}(\mathbb{R}_{>0},\mathbb{R}_{>0}) such that

    ΔC(𝜹;j)ω4(ρj)ΔC(𝜹;argminiρi)\Delta C({\bm{\delta}};j)\leq\omega_{4}(\rho_{j})\Delta C({\bm{\delta}};\arg\min_{i}\rho_{i})

    holds for all n1n\in\mathbb{N}_{1}, all 𝜹n{\bm{\delta}}\in\aleph_{n} and all j[0,n]j\in[0,n].

Remark 5.3.

Though it is somewhat counterintuitive, examples show that the differences ΔC(𝛅;j)\Delta C({\bm{\delta}};j) can be negative. For this reason, property P4 does not imply P3.

The following lemma is derived from (25) by a straightforward induction.

Lemma 5.4.

Let N0N\in\mathbb{N}_{0}, and let the sequences (𝛅(m))m=0N({\bm{\delta}}^{(m)})_{m=0}^{N}, (km)m=0N(k_{m})_{m=0}^{N}, and (nm)m=0N(n_{m})_{m=0}^{N} be generated by Algorithm 3 from 𝛅(0)n0{\bm{\delta}}^{(0)}\in\aleph_{n_{0}}. Then 𝛅(m)nm{\bm{\delta}}^{(m)}\in\aleph_{n_{m}} for all m[0,N]m\in[0,N].

If the spatial discretizations generated by Algorithm 3 do not become uniformly arbitrarily fine, then there exists a particular node at which the discretization eventually remains constant.

Lemma 5.5.

Assume that Algorithm 3 does not terminate, and that the sequences (𝛅(m))m=0({\bm{\delta}}^{(m)})_{m=0}^{\infty}, (km)m=0(k_{m})_{m=0}^{\infty}, and (nm)m=0(n_{m})_{m=0}^{\infty} with 𝛅(m)=(𝐡(m),𝐭(m),𝛒(m)){\bm{\delta}}^{(m)}=(\bm{h}^{(m)},\bm{t}^{(m)},\bm{\rho}^{(m)}) it generates satisfy

limm𝝆(m)0.\lim_{m\to\infty}\|\bm{\rho}^{(m)}\|_{\infty}\neq 0. (27)

Then there exist m0m_{0}\in\mathbb{N} and jm0[0,nm0]j_{m_{0}}\in[0,n_{m_{0}}] such that for all m[m0,)m\in[m_{0},\infty), there exists jm[0,nm]j_{m}\in[0,n_{m}] with tjm(m)=tjm0(m0)t_{j_{m}}^{(m)}=t_{j_{m_{0}}}^{(m_{0})} and ρjm(m)=ρjm0(m0)\rho_{j_{m}}^{(m)}=\rho_{j_{m_{0}}}^{(m_{0})}.

Proof.

By statement (26), the sequence (𝝆(m))m=0(\|\bm{\rho}^{(m)}\|_{\infty})_{m=0}^{\infty} is a monotone decreasing sequence of positive numbers. Hence there exists s+s\in\mathbb{R}_{+} with limm𝝆(m)=s\lim_{m\to\infty}\|\bm{\rho}^{(m)}\|_{\infty}=s. By statement (27), we have s>0s>0, and it follows from statement (23) that there exists m0m_{0}\in\mathbb{N} with

𝝆(m)=sm[m0,).\|\bm{\rho}^{(m)}\|_{\infty}=s\quad\forall\,m\in[m_{0},\infty). (28)

Hence the sets

Sm:={tj(m):j[0,nm],ρj(m)=s},m[m0,),S_{m}:=\{t_{j}^{(m)}:j\in[0,n_{m}],\ \rho_{j}^{(m)}=s\},\quad m\in[m_{0},\infty),

are nonempty and closed subsets of the compact set [0,T][0,T]. By statements (26) and (28), we also have Sm+1SmS_{m+1}\subset S_{m} for all m[m0,)m\in[m_{0},\infty), so by [12, Theorem 26.9], we have m[m0,)Sm\cap_{m\in[m_{0},\infty)}S_{m}\neq\emptyset, which yields the desired result. ∎

If Algorithm 3 does not terminate, then the event that the currently finest spatial discretization is refined in line 3 occurs infinitely often.

Lemma 5.6.

If Algorithm 3 does not terminate in finite time and generates sequences (𝛅(m))m=0({\bm{\delta}}^{(m)})_{m=0}^{\infty}, (km)m=0(k_{m})_{m=0}^{\infty}, and (nm)m=0(n_{m})_{m=0}^{\infty} with 𝛅(m)=(𝐡(m),𝐭(m),𝛒(m)){\bm{\delta}}^{(m)}=(\bm{h}^{(m)},\bm{t}^{(m)},\bm{\rho}^{(m)}), then the set :={m1:ρkm(m)=minj[0,nm]ρj(m)}\mathbb{N}^{\prime}:=\{m\in\mathbb{N}_{1}:\rho_{k_{m}}^{(m)}=\min_{j\in[0,n_{m}]}\rho_{j}^{(m)}\} is infinite.

Proof.

Assume that \mathbb{N}^{\prime} is finite, and take m~:=max()+1\tilde{m}:=\max(\mathbb{N}^{\prime})+1, so that

ρkm(m)>minj[0,nm]ρj(m)m[m~,).\rho_{k_{m}}^{(m)}>\min_{j\in[0,n_{m}]}\rho_{j}^{(m)}\quad\forall\,m\in[\tilde{m},\infty). (29)

We assert by induction that this implies

ρ~:=minj[0,nm~]ρj(m~)=minj[0,nm]ρj(m)m[m~,).\tilde{\rho}:=\min_{j\in[0,n_{\tilde{m}}]}\rho_{j}^{(\tilde{m})}=\min_{j\in[0,n_{m}]}\rho_{j}^{(m)}\quad\forall\,m\in[\tilde{m},\infty). (30)

This is clearly true for m=m~m=\tilde{m}. If (30) holds for some m[m~,)m\in[\tilde{m},\infty), then by (23) and (29), we have ρkm(m)4minj[0,nm]ρj(m)\rho_{k_{m}}^{(m)}\geq 4\min_{j\in[0,n_{m}]}\rho_{j}^{(m)}, so by statement (26) we find that (30) holds with m+1m+1 in lieu of mm.

To derive a contradiction, we introduce the auxiliary function

β:[m~,),β(m):=log4(ρ0(m)ρ~)+j=1nm(2log4(ρj(m)/ρ~)1).\beta:[\tilde{m},\infty)\to\mathbb{R},\quad\beta(m):=\log_{4}(\tfrac{\rho_{0}^{(m)}}{\tilde{\rho}})+\sum_{j=1}^{n_{m}}(2^{\log_{4}(\rho_{j}^{(m)}/\tilde{\rho})}-1).

By (30) we have ρj(m)ρ~\rho_{j}^{(m)}\geq\tilde{\rho} and hence log4(ρj(m)/ρ~)0\log_{4}(\rho_{j}^{(m)}/\tilde{\rho})\geq 0 for all m[m~,)m\in[\tilde{m},\infty) and j[0,nm]j\in[0,n_{m}]. Together with (23), we obtain that

log4(ρj(m)/ρ~)0m[m~,),j[0,nm],\log_{4}(\rho_{j}^{(m)}/\tilde{\rho})\in\mathbb{N}_{0}\quad\forall\,m\in[\tilde{m},\infty),\ \forall\,j\in[0,n_{m}], (31)

and hence that β(m)0\beta(m)\in\mathbb{N}_{0} for all m[m~,)m\in[\tilde{m},\infty). Using (26), it is easy to check that

β(m+1)=β(m)1,m[m~,),\beta(m+1)=\beta(m)-1,\quad\forall\,m\in[\tilde{m},\infty),

which implies that for m^:=m~+β(m~)\hat{m}:=\tilde{m}+\beta(\tilde{m}), we have β(m^)=0\beta(\hat{m})=0. By (31), this forces ρ(m^)=ρ~𝟙\rho^{(\hat{m})}=\tilde{\rho}\mathbbm{1}. In particular, we have

ρkm^(m^)=ρ~=minj[0,nm^]ρj(m^),\rho^{(\hat{m})}_{k_{\hat{m}}}=\tilde{\rho}=\min_{j\in[0,n_{\hat{m}}]}\rho_{j}^{(\hat{m})},

which implies m^\hat{m}\in\mathbb{N}^{\prime}. Since m~m^\tilde{m}\leq\hat{m}, this contradicts the definition of m~\tilde{m}. ∎

Now we prove that Algorithm 3 terminates in finite time.

Theorem 5.7.

If the functions CC and EE satisfy properties P1P4, then, for any 𝛅(0)n0{\bm{\delta}}^{(0)}\in\aleph_{n_{0}} and ε>0\varepsilon>0, Algorithm 3 terminates after a finite number mm\in\mathbb{N} of iterations and returns a 𝛅(m)nm{\bm{\delta}}^{(m)}\in\aleph_{n_{m}} with E(𝛅(m))εE({\bm{\delta}}^{(m)})\leq\varepsilon.

Proof.

Assume that Algorithm 3 does not terminate. Then it generates sequences (𝜹(m))m=0({\bm{\delta}}^{(m)})_{m=0}^{\infty}, (km)m=0(k_{m})_{m=0}^{\infty} and (nm)m=0(n_{m})_{m=0}^{\infty} with 𝜹(m)=(𝒉(m),𝒕(m),𝝆(m)){\bm{\delta}}^{(m)}=(\bm{h}^{(m)},\bm{t}^{(m)},\bm{\rho}^{(m)}) for all mm\in\mathbb{N} and

E(𝜹(m))>εm.E({\bm{\delta}}^{(m)})>\varepsilon\quad\forall\,m\in\mathbb{N}. (32)

In the following, we show that

limm𝝆(m)=0,\lim_{m\to\infty}\|\bm{\rho}^{(m)}\|_{\infty}=0, (33)

which, in conjunction with property P1, contradicts (32). This, combined with Lemma 5.4, completes the proof.

We obtain from property P2 that

ΔE(𝜹(m);j)ω2(ρj(m))<0,m1,j[0,nm],\Delta E({\bm{\delta}}^{(m)};j)\leq-\omega_{2}(\rho_{j}^{(m)})<0,\quad\forall\,m\in\mathbb{N}_{1},\ \forall\,j\in[0,n_{m}], (34)

and by construction, we have

j=0m1ΔE(𝜹(j);kj)=E(𝜹(0))E(𝜹(m))E(𝜹(0)),m1.-\sum_{j=0}^{m-1}\Delta E({\bm{\delta}}^{(j)};k_{j})=E({\bm{\delta}}^{(0)})-E({\bm{\delta}}^{(m)})\leq E({\bm{\delta}}^{(0)}),\quad\forall\,m\in\mathbb{N}_{1}. (35)

Combining (34) and (35) yields

limmΔE(𝜹(m);km)=0.\lim_{m\to\infty}\Delta E({\bm{\delta}}^{(m)};k_{m})=0. (36)

If statement (33) is false, then by Lemma 5.5, there exist m01m_{0}\in\mathbb{N}_{1} and im0[0,nm0]i_{m_{0}}\in[0,n_{m_{0}}] such that, for every m[m0,)m\in[m_{0},\infty), there exists im[0,nm]i_{m}\in[0,n_{m}] with

ρim(m)=ρim0(m0).\rho_{i_{m}}^{(m)}=\rho_{i_{m_{0}}}^{(m_{0})}. (37)

By Lemma 5.6, the set

={m1:ρkm(m)=minj[0,nm]ρj(m)}\mathbb{N}^{\prime}=\{m\in\mathbb{N}_{1}:\rho_{k_{m}}^{(m)}=\min_{j\in[0,n_{m}]}\rho_{j}^{(m)}\}

is infinite, and hence ′′:=[m0,)\mathbb{N}^{\prime\prime}:=[m_{0},\infty)\cap\mathbb{N}^{\prime} is infinite as well. By property P3, we have

ΔC(𝜹(m);km)>0m′′.\Delta C({\bm{\delta}}^{(m)};k_{m})>0\quad\forall\,m\in\mathbb{N}^{\prime\prime}. (38)

If ΔC(𝜹(m);j)0\Delta C({\bm{\delta}}^{(m)};j)\leq 0 for some m′′m\in\mathbb{N}^{\prime\prime} and j[0,nm]j\in[0,n_{m}], then the condition in line 3 of Algorithm 3 is false, and so line 3 yields the contradiction ΔC(𝜹(m);km)0\Delta C({\bm{\delta}}^{(m)};k_{m})\leq 0. Hence

ΔC(𝜹(m);j)>0m′′,j[0,nm],\Delta C({\bm{\delta}}^{(m)};j)>0\quad\forall\,m\in\mathbb{N}^{\prime\prime},\ \forall\,j\in[0,n_{m}],

so line 3 is executed in iteration mm for all m′′m\in\mathbb{N}^{\prime\prime}, and thus we have

ΔE(𝜹(m);km)ΔC(𝜹(m);km)ΔE(𝜹(m);im)ΔC(𝜹(m);im)m′′.\tfrac{\Delta E({\bm{\delta}}^{(m)};k_{m})}{\Delta C({\bm{\delta}}^{(m)};k_{m})}\leq\tfrac{\Delta E({\bm{\delta}}^{(m)};i_{m})}{\Delta C({\bm{\delta}}^{(m)};i_{m})}\quad\forall\,m\in\mathbb{N}^{\prime\prime}.

By (34) and (38), by (34) and property P2, by (37), and by (36), we have

lim′′mΔC(𝜹(m);km)ΔC(𝜹(m);im)lim′′mΔE(𝜹(m);km)ΔE(𝜹(m);im)\displaystyle\lim_{\mathbb{N}^{\prime\prime}\ni m\to\infty}\tfrac{\Delta C({\bm{\delta}}^{(m)};k_{m})}{\Delta C({\bm{\delta}}^{(m)};i_{m})}\leq\lim_{\mathbb{N}^{\prime\prime}\ni m\to\infty}\tfrac{\Delta E({\bm{\delta}}^{(m)};k_{m})}{\Delta E({\bm{\delta}}^{(m)};i_{m})}
lim′′mΔE(𝜹(m);km)ω2(ρim(m))=lim′′mΔE(𝜹(m);km)ω2(ρim0(m0))=0.\displaystyle\leq\lim_{\mathbb{N}^{\prime\prime}\ni m\to\infty}\tfrac{\Delta E({\bm{\delta}}^{(m)};k_{m})}{-\omega_{2}(\rho_{i_{m}}^{(m)})}=\lim_{\mathbb{N}^{\prime\prime}\ni m\to\infty}\tfrac{\Delta E({\bm{\delta}}^{(m)};k_{m})}{-\omega_{2}(\rho_{i_{m_{0}}}^{(m_{0})})}=0.

However, in view of the definition of ′′\mathbb{N}^{\prime\prime}, property P4 and (37) provide

ΔC(𝜹(m);km)ΔC(𝜹(m);im)ω4(ρim(m))1=ω4(ρim0(m0))1>0m′′,\tfrac{\Delta C({\bm{\delta}}^{(m)};k_{m})}{\Delta C({\bm{\delta}}^{(m)};i_{m})}\geq\omega_{4}(\rho^{(m)}_{i_{m}})^{-1}=\omega_{4}(\rho^{(m_{0})}_{i_{m_{0}}})^{-1}>0\quad\forall\,m\in\mathbb{N}^{\prime\prime},

contradicting the above statement. ∎

6 Efficient discretizations for boundary tracking

We wish to measure the quality of a discretization

𝜹=(𝒉,𝒕,𝝆)>0n×+0,n×>00,n{\bm{\delta}}=(\bm{h},\bm{t},\bm{\rho})\in\mathbb{R}^{n}_{>0}\times\mathbb{R}_{+}^{0,n}\times\mathbb{R}_{>0}^{0,n}

in terms of the a priori error of the corresponding discrete reachable sets and the cost of their computation. This will allow us to select a discretization with a high benefit-to-cost ratio. To this end, we split the a priori error bound

E(𝜹)\displaystyle E({\bm{\delta}}) =k=0nk(𝜹),\displaystyle=\sum_{k=0}^{n}\mathcal{E}_{k}({\bm{\delta}}), (39)

from Theorem 4.7 for the reachable sets computed in recursion (17) into the contribution 0(𝜹):=ρ02eLT\mathcal{E}_{0}({\bm{\delta}}):=\frac{\rho_{0}}{2}e^{LT} from the discretization of the initial set X0X_{0} and the contributions

k(𝜹):=eL(Ttk)[(eLhk1)(Phk+(1+Lhk)2Lhkρk12)+χk(𝜹)],k(0,n],\mathcal{E}_{k}({\bm{\delta}}):=e^{L(T-t_{k})}\Big{[}(e^{Lh_{k}}-1)\big{(}Ph_{k}+\tfrac{(1+Lh_{k})^{2}}{Lh_{k}}\tfrac{\rho_{k-1}}{2}\big{)}+\chi_{k}({\bm{\delta}})\Big{]},\ k\in(0,n],

from the discretization of the dynamics on each subinterval (tk1,tk](t_{k-1},t_{k}] for k(0,n]k\in(0,n], where we use the notation

χk(𝜹):={ρk2,ρk1<ρk,0,otherwise.\chi_{k}({\bm{\delta}}):=\begin{cases}\tfrac{\rho_{k}}{2},&\rho_{k-1}<\rho_{k},\\ 0,&\text{otherwise}.\end{cases}

Since there is no straightforward a priori formula for the computational cost of recursion (17) available, we derive a heuristic estimator based on a posteriori knowledge from previously performed computations in Section 6.1. We verify in Section 6.2 that this cost estimator and the a priori error estimate (39) satisfy properties P1 to P4, which means that it is safe to carry out the subdivision scheme from Section 5 with these particular inputs.

6.1 Estimating the computational cost

We first collect some preliminaries:

  • i)

    Recall that dd is the dimension of the state space, and let

    dx:=maxt[0,T]dimF(t)anddu:=maxt[0,T]maxxF(t)dimF(x).d_{x}:=\max_{t\in[0,T]}\dim\mathcal{R}^{F}(t)\quad\text{and}\quad d_{u}:=\max_{t\in[0,T]}\max_{x\in\mathcal{R}^{F}(t)}\dim F(x).

    Then the numbers

    dR:=min(d1,dx)anddF:=min(d1,du)d_{R}:=\min(d-1,d_{x})\quad\text{and}\quad d_{F}:=\min(d-1,d_{u})

    are a good guess for the dimension of the boundary of the reachable sets and the boundary of the right-hand side FF, see Figure 3 for an illustration. From here forward, we assume that dR1d_{R}\geq 1 to rule out pathological situations.

  • ii)

    We express the average thickness of the generalized annuli computed in the evaluation of the operator Ψh,ρ,ρ\Psi_{h,\rho,\rho^{\prime}} in line 1 of Algorithm 1 and (18), respectively, in terms of the function σ:(0,14L]×(0,P8L]>0\sigma:(0,\tfrac{1}{4L}]\times(0,\tfrac{P}{8L}]\to\mathbb{R}_{>0} given by

    σ(h,ρ):=2α(h,ρ)+14κ(h,ρ)=ρ4(1+Lh)(1+Lh1Lh+6).\sigma(h,\rho):=2\alpha(h,\rho)+\tfrac{1}{4}\kappa(h,\rho)=\tfrac{\rho}{4}(1+Lh)\left(\tfrac{1+Lh}{1-Lh}+6\right). (40)

    Figure 2 on page 2 provides a visualization of these annuli.

xxyy
(a) d=2d=2, dx=1d_{x}=1, dR=1d_{R}=1
xxyy
(b) d=2d=2, dx=2d_{x}=2, dR=1d_{R}=1
Figure 3: The dimension dRd_{R} of the boundary of a set is determined by the dimension dd of the state space and the dimension dxd_{x} of the set.

Now we derive the cost estimator. The exact cost, in terms of grid points computed, of carrying out recursion (17) with a given discretization 𝜹=(𝒉,𝒕,𝝆)>0n×+0,n×>00,n{\bm{\delta}}=(\bm{h},\bm{t},\bm{\rho})\in\mathbb{R}^{n}_{>0}\times\mathbb{R}_{+}^{0,n}\times\mathbb{R}_{>0}^{0,n} is

C^(𝜹)\displaystyle\hat{C}({\bm{\delta}}) :=k=0n1𝒞^k(𝜹),\displaystyle:=\sum_{k=0}^{n-1}\hat{\mathcal{C}}_{k}({\bm{\delta}}), (41)
𝒞^k(𝜹)\displaystyle\hat{\mathcal{C}}_{k}({\bm{\delta}}) :=j{1,1,2}xρkj𝜹F(k)#Φhk,ρ~k+1,αk+1(x)\displaystyle:=\sum_{j\in\{-1,1,2\}}\sum_{x\in\partial_{\rho_{k}}^{j}\!\mathcal{R}_{\bm{\delta}}^{F}(k)}\#\Phi_{h_{k},\tilde{\rho}_{k+1}}^{\partial,\alpha_{k+1}}(x)
+xρk0𝜹F(k)#(Φh,ρ~k+1,αk+1+κk+1(x)Φh,ρ~k+1αk+1(x)),\displaystyle\qquad\qquad+\sum_{x\in\partial_{\rho_{k}}^{0}\!\mathcal{R}_{\bm{\delta}}^{F}(k)}\#\big{(}\Phi_{h,\tilde{\rho}_{k+1}}^{\partial,\alpha_{k+1}+\kappa_{k+1}}(x)\cap\Phi_{h,\tilde{\rho}_{k+1}}^{\alpha_{k+1}}(x)\big{)},

where we used the notation

αk=α(hk,ρk1),κk=κ(hk,ρk1)andρ~k+1=min(ρk,ρk+1)\alpha_{k}=\alpha(h_{k},\rho_{k-1}),\quad\kappa_{k}=\kappa(h_{k},\rho_{k-1})\quad\text{and}\quad\tilde{\rho}_{k+1}=\min(\rho_{k},\rho_{k+1})

from (3) and (11). It can, however, only be determined after the reachable sets have been computed.

For this reason, we pursue an iterative approach, see Algorithm 4, in which we alternate between executing recursion (17) and refining the discretization based on information gathered in the recursion. Assuming that at some point of the iteration, we have determined

  • i)

    a discretization 𝜹()n{\bm{\delta}}^{(\ell)}\in\aleph_{n_{\ell}},

  • ii)

    the corresponding reachable sets 𝜹()F(k)\mathcal{R}_{{\bm{\delta}}^{(\ell)}}^{F}(k), k[0,n]k\in[0,n_{\ell}], from (17), and

  • iii)

    the corresponding exact complexities 𝒞^k(𝜹())\hat{\mathcal{C}}_{k}({\bm{\delta}}^{(\ell)}) for k[0,n1]k\in[0,n_{\ell}-1],

we proceed as follows:

  • i)

    First we compute the surrogate surface areas

    v^R,k()\displaystyle\hat{v}_{R,k}^{(\ell)} =14(ρk())dRj=12#ρk()j𝜹()F(k),k[0,n],\displaystyle=\tfrac{1}{4}(\rho_{k}^{(\ell)})^{d_{R}}\sum_{j=-1}^{2}\#\partial_{\rho_{k}^{(\ell)}}^{j}\mathcal{R}_{{\bm{\delta}}^{(\ell)}}^{F}(k),\quad k\in[0,n_{\ell}], (42)
    v^F,k()\displaystyle\hat{v}_{F,k}^{(\ell)} =min(ρk(),ρk+1())dF+1(hk+1())dFσ(hk+1(),ρk())𝒞^k(𝜹())j=12#ρk()j𝜹()F(k),k[0,n1]\displaystyle=\tfrac{\min(\rho_{k}^{(\ell)},\rho_{k+1}^{(\ell)})^{d_{F}+1}}{(h_{k+1}^{(\ell)})^{d_{F}}\sigma(h_{k+1}^{(\ell)},\rho_{k}^{(\ell)})}\tfrac{\hat{\mathcal{C}}_{k}({\bm{\delta}}^{(\ell)})}{\sum_{j=-1}^{2}\#\partial_{\rho_{k}^{(\ell)}}^{j}\mathcal{R}_{{\bm{\delta}}^{(\ell)}}^{F}(k)},\quad k\in[0,n_{\ell}-1] (43)

    based on the following heuristic: Equation (42) estimates the surface area of the reachable sets 𝜹()F(k)\mathcal{R}_{{\bm{\delta}}^{(\ell)}}^{F}(k) by counting the average number of grid points in the sets ρk()j𝜹()F(k)\partial_{\rho_{k}^{(\ell)}}^{j}\mathcal{R}_{{\bm{\delta}}^{(\ell)}}^{F}(k) for j[1,2]j\in[-1,2] and multiplying this number by the surface area of one face of a cube with diameter ρk()\rho_{k}^{(\ell)}. Equation (43) estimates the average area of F(x)\partial F(x) for x𝜹()F(k)x\in\partial\mathcal{R}_{{\bm{\delta}}^{(\ell)}}^{F}(k) in a similar way, taking into account the scaling of FF with hk+1()h^{(\ell)}_{k+1} and the blowup size σ(hk+1(),ρk())\sigma(h_{k+1}^{(\ell)},\rho_{k}^{(\ell)}) in evaluating Ψ\Psi with parameters hk+1()h_{k+1}^{(\ell)}, ρk()\rho_{k}^{(\ell)} and min(ρk(),ρk+1())\min(\rho_{k}^{(\ell)},\rho_{k+1}^{(\ell)}), see (18) and Figure 2.

  • ii)

    Next, setting v^F,n()=v^F,n1\hat{v}_{F,n_{\ell}}^{(\ell)}=\hat{v}_{F,n_{\ell}-1}, we interpolate the estimates (tj(),v^R,j())j=0n(t^{(\ell)}_{j},\hat{v}_{R,j}^{(\ell)})_{j=0}^{n_{\ell}} and (tj(),v^F,j())j=0n(t^{(\ell)}_{j},\hat{v}_{F,j}^{(\ell)})_{j=0}^{n_{\ell}} with piecewise linear splines

    vR(),vF():[0,T]>0v_{R}^{(\ell)},v_{F}^{(\ell)}:[0,T]\to\mathbb{R}_{>0}

    and form their product v():=vR()vF()v^{(\ell)}:=v^{(\ell)}_{R}v^{(\ell)}_{F}.

  • iii)

    Finally, for any finer discretization 𝜹=(𝒉,𝒕,𝝆)>0n×+0,n×>00,n{\bm{\delta}}=(\bm{h},\bm{t},\bm{\rho})\in\mathbb{R}^{n}_{>0}\times\mathbb{R}_{+}^{0,n}\times\mathbb{R}_{>0}^{0,n}, we estimate the number of grid points to be computed in recursion (14) by C(δ;v())C(\delta;v^{(\ell)}), where

    C(𝜹;v)\displaystyle C({\bm{\delta}};v) =k=0n1𝒞k(𝜹;v),\displaystyle=\sum_{k=0}^{n-1}{\mathcal{C}}_{k}({\bm{\delta}};v), (44)
    𝒞k(𝜹;v)\displaystyle\mathcal{C}_{k}({\bm{\delta}};v) =4v(tk)hk+1dFρkdRmin(ρk,ρk+1)dF+1σ(hk+1,ρk),\displaystyle=\tfrac{4v(t_{k})h_{k+1}^{d_{F}}}{\rho_{k}^{d_{R}}\min(\rho_{k},\rho_{k+1})^{d_{F}+1}}\sigma(h_{k+1},\rho_{k}),

    which inverts the heuristic leading to the estimates (42) and (43).

Input: h,ρ,ρ>0h,\rho,\rho^{\prime}>0 with ρ/ρ(0,1)1\rho^{\prime}/\rho\in(0,1)\cup\mathbb{N}_{1}, (D0,D1)bdcρ(D_{0},D_{1})\in\operatorname{bdc}_{\rho}.
Output: Ψh,ρ,ρ(D0,D1)bdcρ\Psi_{h,\rho,\rho^{\prime}}(D_{0},D_{1})\in\operatorname{bdc}_{\rho^{\prime}}, c^R1\hat{c}_{R}\in\mathbb{N}_{1}, c^F1\hat{c}_{F}\in\mathbb{N}_{1}
1
2Compute D1D_{-1} and D2D_{2} from D0D_{0} and D1D_{1} by neighbor search;
3 c^Rj=1j=2#Dj\hat{c}_{R}\leftarrow\sum_{j=-1}^{j=2}\#D_{j};
4 (α,κ,ρ~)(α(h,ρ),2(1+Lh)(α1Lh+ρ),min(ρ,ρ))(\alpha,\kappa,\tilde{\rho})\leftarrow(\alpha(h,\rho),2(1+Lh)(\frac{\alpha}{1-Lh}+\rho),\min(\rho,\rho^{\prime}));
5 Compute W00,W01W_{0}^{0},W_{0}^{-1}, and W1W_{1} as in Theorem 4.8;
6 Compute D~1\tilde{D}_{1} and D~0\tilde{D}_{0} as in Theorem 4.8;
7 c^Fj{1,1,2}xDj#Φh,ρ~,α(x)+xD0#(Φh,ρ~,α+κ(x)Φh,ρ~α(x))\hat{c}_{F}\leftarrow\sum_{j\in\{-1,1,2\}}\sum_{x\in D_{j}}\#\Phi^{\partial,\alpha}_{h,\tilde{\rho}}(x)+\sum_{x\in D_{0}}\#(\Phi^{\partial,\alpha+\kappa}_{h,\tilde{\rho}}(x)\cap\Phi^{\alpha}_{h,\tilde{\rho}}(x));
8 if ρ>ρ\rho^{\prime}>\rho then
9      Ψh,ρ,ρ(D0,D1)Rρ~ρ(D~0,D~1)\Psi_{h,\rho,\rho^{\prime}}(D_{0},D_{1})\leftarrow\partial R_{\tilde{\rho}}^{\rho^{\prime}}(\tilde{D}_{0},\tilde{D}_{1});
10else
11      Ψh,ρ,ρ(D0,D1)(D~0,D~1)\Psi_{h,\rho,\rho^{\prime}}(D_{0},D_{1})\leftarrow(\tilde{D}_{0},\tilde{D}_{1});
Evaluates Ψh,ρ,ρ\Psi_{h,\rho,\rho^{\prime}} and gathers data for cost estimator (44).
Algorithm 1 A single boundary tracking step.
Evaluates Ψh,ρ,ρ\Psi_{h,\rho,\rho^{\prime}} and gathers data for cost estimator (44).
Input: discretization 𝜹=(𝒉,𝒕,𝝆){\bm{\delta}}=(\bm{h},\bm{t},\bm{\rho})\in\aleph,
boundary of initial set (D00,D10)bdcρ0(D_{0}^{0},D_{1}^{0})\in\operatorname{bdc}_{\rho_{0}}
Output: splines vR,vF:[0,T]>0v_{R},v_{F}:[0,T]\to\mathbb{R}_{>0},
boundaries (D0k,D1k)bdcρk(D_{0}^{k},D_{1}^{k})\in\operatorname{bdc}_{\rho_{k}}, k=0,,nk=0,\ldots,n, of reachable sets
1
2for k0k\leftarrow 0 to n1n-1 do
3       (D0k+1,D1k+1,c^R,k,c^F,k)[Algorithm 1](hk+1,ρk,ρk+1,D0k,D1k)(D_{0}^{k+1},D_{1}^{k+1},\hat{c}_{R,{k}},\hat{c}_{F,k})\leftarrow[\text{Algorithm\ \ref{Alg:BdryEulrStep}}](h_{k+1},\rho_{k},\rho_{k+1},D_{0}^{k},D_{1}^{k});
4       (v^R,k,v^F,k)(14c^R,kρkdR,min(ρk,ρk+1)dF+1(hk+1)dFσ(hk+1,ρk)c^F,kc^R,k)(\hat{v}_{R,k},\hat{v}_{F,k})\leftarrow(\frac{1}{4}\hat{c}_{R,k}\rho_{k}^{d_{R}},\frac{\min(\rho_{k},\rho_{k+1})^{d_{F}+1}}{(h_{k+1})^{d_{F}}\sigma(h_{k+1},\rho_{k})}\frac{\hat{c}_{F,k}}{\hat{c}_{R,k}});
5
6c^R,nj=1j=2#(ρnj(Trρn1(D0n,D1n)))\hat{c}_{R,n}\leftarrow\sum_{j=-1}^{j=2}\#(\partial_{\rho_{n}}^{j}(\operatorname{Tr}_{\rho_{n}}^{-1}(D_{0}^{n},D_{1}^{n})));
7 (v^R,n,v^F,n)(14c^R,nρndR,v^F,n1)(\hat{v}_{R,n},\hat{v}_{F,n})\leftarrow(\frac{1}{4}\hat{c}_{R,n}\rho_{n}^{d_{R}},\hat{v}_{F,n-1});
8 (vR,vF)([linear spline](t,v^R),[linear spline](t,v^F))(v_{R},v_{F})\leftarrow(\text{[linear spline]}(t,\hat{v}_{R}),\text{[linear spline]}(t,\hat{v}_{F}));
9
Implements recursion (17) and gathers data for estimator (44).
Algorithm 2 Boundary tracking with nonuniform discretization.
Implements recursion (17) and gathers data for estimator (44).
Input: ε>0\varepsilon>0, discretization 𝜹(0)n0{\bm{\delta}}^{(0)}\in\aleph_{n_{0}}, functions CC and EE
Output: nm1n_{m}\in\mathbb{N}_{1}, discretization 𝜹(m)nm{\bm{\delta}}^{(m)}\in\aleph_{n_{m}}
1 m0m\leftarrow 0;
2 while E(𝛅(m))>εE({\bm{\delta}}^{(m)})>\varepsilon do
3       if minjΔC(𝛅(m);j)>0\min_{j}\Delta C({\bm{\delta}}^{(m)};j)>0 then
4             kmargminjΔE(𝜹(m);j)ΔC(𝜹(m);j)k_{m}\leftarrow\operatorname{argmin}_{j}\frac{\Delta E({\bm{\delta}}^{(m)};j)}{\Delta C({\bm{\delta}}^{(m)};j)};
5            
6      else
7             kmargminjΔC(𝜹(m);j)k_{m}\leftarrow\operatorname{argmin}_{j}\Delta C({\bm{\delta}}^{(m)};j);
8            
9      𝜹(m+1)ψ[𝜹(m);km]{\bm{\delta}}^{(m+1)}\leftarrow\psi[{\bm{\delta}}^{(m)};k_{m}];
10       if km>0k_{m}>0 then nm+1nm+1n_{m+1}\leftarrow n_{m}+1 else nm+1nmn_{m+1}\leftarrow n_{m};
11       mm+1m\leftarrow m+1;
12      
Finds cost-efficient discretization with guaranteed error bound.
Algorithm 3 An abstract subdivision scheme.
Finds cost-efficient discretization with guaranteed error bound.
Input: thresholds ε1>ε2>>εmax>0\varepsilon_{1}>\varepsilon_{2}>\ldots>\varepsilon_{\ell_{\max}}>0
Output: n1n\in\mathbb{N}_{1}, discretization 𝜹n{\bm{\delta}}\in\aleph_{n},
boundaries (D0k,D1k)bdcρk(D_{0}^{k},D_{1}^{k})\in\operatorname{bdc}_{\rho_{k}}, k=0,,nk=0,\ldots,n, of reachable sets
1
2(n0,)(4LT,0)(n_{0},\ell)\leftarrow(\lceil 4LT\rceil,0);
3 𝜹(0)(h(0),t(0),ρ(0))(Tn0𝟙n0,Tn0Σ+𝟙n0,2LPT2n02𝟙n0+1){\bm{\delta}}^{(0)}\leftarrow(h^{(0)},t^{(0)},\rho^{(0)})\leftarrow(\frac{T}{n_{0}}\mathbbm{1}_{n_{0}},\frac{T}{n_{0}}\Sigma_{+}\mathbbm{1}_{n_{0}},2LP\tfrac{T^{2}}{n_{0}^{2}}\mathbbm{1}_{n_{0}+1});
4
5while max\ell\leq\ell_{\max} do
6       (D00,D10)Trρ0()(πρ0()ρ0()/2(X0))(D_{0}^{0},D_{1}^{0})\leftarrow\operatorname{Tr}_{\rho_{0}^{(\ell)}}(\pi^{\rho_{0}^{(\ell)}/2}_{\rho_{0}^{(\ell)}}(X_{0}));
7       ((D0k,D1k)k=0n,vR(),vF())[Algorithm 2](𝜹(),(D00,D10))((D_{0}^{k},D_{1}^{k})_{k=0}^{n_{\ell}},v^{(\ell)}_{R},v^{(\ell)}_{F})\leftarrow[\text{Algorithm\ \ref{Alg:boundary:tracking}}]({\bm{\delta}}^{(\ell)},(D_{0}^{0},D_{1}^{0}));
8       if <max\ell<\ell_{\max} then
9            (n+1,𝜹(+1))[Algorithm 3](ε+1,𝜹(),C(;vR()vF()),E())(n_{\ell+1},{\bm{\delta}}^{(\ell+1)})\leftarrow[\text{Algorithm\ \ref{Alg:RefineDisc}}](\varepsilon_{\ell+1},{\bm{\delta}}^{(\ell)},C(\,\cdot\,;v^{(\ell)}_{R}v^{(\ell)}_{F}),E(\,\cdot\,));
10      +1\ell\leftarrow\ell+1;
11
12(n,𝜹)(nmax,𝜹(max))(n,{\bm{\delta}})\leftarrow(n_{\ell_{\max}},{\bm{\delta}}^{(\ell_{\max})});
13
Approximates reachable sets with near-optimal discretization.
Algorithm 4 Boundary tracking with iterative refinement.
Approximates reachable sets with near-optimal discretization.

6.2 Verifying hypotheses P1 to P4

In the following, we show that for a given, fixed vC0([0,T],>0),v\in C^{0}([0,T],\mathbb{R}_{>0}), the functions E()E(\,\cdot\,) and C(;v)C(\,\cdot\,;v) from (39) and (44) satisfy properties P1 to P4.

Lemma 6.1.

We have

1+sess,\displaystyle 1+s\leq e^{s}\quad\forall\,s\in\mathbb{R}, (45)
es1+s+s2s[0,3/2],\displaystyle e^{s}\leq 1+s+s^{2}\quad\forall\,s\in[0,3/2], (46)
es2(es21)(1+s/2)2s(es1)(1+s)22s3s8s(0,1/4].\displaystyle e^{\frac{s}{2}}(e^{\frac{s}{2}}-1)\tfrac{(1+s/2)^{2}}{s}-(e^{s}-1)\tfrac{(1+s)^{2}}{2s}\leq-\tfrac{3s}{8}\quad\forall\,s\in(0,1/4]. (47)
Proof.

Both (45) and (46) are well-known. Using es1=(es21)(es2+1)e^{s}-1=(e^{\frac{s}{2}}-1)(e^{\frac{s}{2}}+1) and es21s2+s24e^{\frac{s}{2}}-1\leq\tfrac{s}{2}+\tfrac{s^{2}}{4}, which is (46) with s/2s/2 in lieu of ss, we obtain

es2(es21)(1+s/2)2s(es1)(1+s)22s\displaystyle e^{\frac{s}{2}}(e^{\frac{s}{2}}-1)\tfrac{(1+s/2)^{2}}{s}-(e^{s}-1)\tfrac{(1+s)^{2}}{2s}
=12s(es21)[2es2(1+s2)2(es2+1)(1+s)2]\displaystyle=\tfrac{1}{2s}(e^{\frac{s}{2}}-1)[2e^{\frac{s}{2}}(1+\tfrac{s}{2})^{2}-(e^{\frac{s}{2}}+1)(1+s)^{2}]
=12s(es21)[(es21)s22es22ss2]\displaystyle=\tfrac{1}{2s}(e^{\frac{s}{2}}-1)[(e^{\frac{s}{2}}-1)-\tfrac{s^{2}}{2}e^{\frac{s}{2}}-2s-s^{2}]
12s(es21)[s2+s242ss2]12s(es21)3s243s8.\displaystyle\leq\tfrac{1}{2s}(e^{\frac{s}{2}}-1)[\tfrac{s}{2}+\tfrac{s^{2}}{4}-2s-s^{2}]\leq-\tfrac{1}{2s}(e^{\frac{s}{2}}-1)\tfrac{3s^{2}}{4}\leq-\tfrac{3s}{8}.

We collect some statements that follow directly from the definition (40).

Lemma 6.2.

The function σ\sigma is monotone increasing in both arguments. In particular, the constant σ¯:=σ(14L,P8L)\bar{\sigma}:=\sigma(\tfrac{1}{4L},\tfrac{P}{8L}) satisfies

σ(hk+1,ρk)σ¯k[0,n1],(𝒉,𝒕,𝝆)n,n1.\sigma(h_{k+1},\rho_{k})\leq\bar{\sigma}\quad\forall\,k\in[0,n-1],\ \forall\,(\bm{h},\bm{t},\bm{\rho})\in\aleph_{n},\ \forall\,n\in\mathbb{N}_{1}. (48)

Furthermore, for all (h,ρ)[0,14L]×(0,P8L](h,\rho)\in[0,\tfrac{1}{4L}]\times(0,\tfrac{P}{8L}], we have

74ρσ(h,ρ)=4σ(h,ρ/4),\displaystyle\tfrac{7}{4}\rho\leq\sigma(h,\rho)=4\sigma(h,\rho/4), (49)
12σ(h,ρ)σ(h/2,ρ).\displaystyle\tfrac{1}{2}\sigma(h,\rho)\leq\sigma(h/2,\rho). (50)
Proof.

The function σ\sigma is clearly monotone increasing in ρ\rho. A simple computation shows that ddsσ(s,ρ)>0\frac{d}{ds}\sigma(s,\rho)>0 for all s(0,14L]s\in(0,\tfrac{1}{4L}] and ρ(0,P8L]\rho\in(0,\frac{P}{8L}]. With this (48) then follows from (22).

The inequality in statement (49) holds since

74ρ=ρ4×1×(1+6)ρ4(1+Lh)(1+Lh1Lh+6)=σ(h,ρ),\displaystyle\tfrac{7}{4}\rho=\tfrac{\rho}{4}\times 1\times(1+6)\leq\tfrac{\rho}{4}(1+Lh)(\tfrac{1+Lh}{1-Lh}+6)=\sigma(h,\rho),

and the identity in statement (49) follows directly from the definition (40). To show (50) we use monotonicity, (24), and (49) with h/2h/2 in lieu of hh in the computation

12σ(h,ρ)12σ(14L,ρ)=11596ρ<74ρσ(h2,ρ).\tfrac{1}{2}\sigma(h,\rho)\leq\tfrac{1}{2}\sigma(\tfrac{1}{4L},\rho)=\tfrac{115}{96}\rho<\tfrac{7}{4}\rho\leq\sigma(\tfrac{h}{2},\rho).

Now we check properties P1 to P4.

Lemma 6.3.

The function EE from (39) satisfies property P1 with n\aleph_{n} as in Definition 5.1.

Proof.

Let n1n\in\mathbb{N}_{1} and 𝜹=(𝒉,𝒕,𝝆)n{\bm{\delta}}=(\bm{h},\bm{t},\bm{\rho})\in\aleph_{n}. We clearly have

ρ02eLT12eLTρ,\frac{\rho_{0}}{2}e^{LT}\leq\frac{1}{2}e^{LT}\|\rho\|_{\infty}, (51)

and using (24), we find that

k=1neL(Ttk)χk(𝜹)eLTk=1nρk2=LPeLTk=1nhk2\displaystyle\sum_{k=1}^{n}e^{L(T-t_{k})}\chi_{k}({\bm{\delta}})\leq e^{LT}\sum_{k=1}^{n}\frac{\rho_{k}}{2}=LPe^{LT}\sum_{k=1}^{n}h_{k}^{2} (52)
LPeLThk=1nhkLPTeLTh=12TeLTLPρ.\displaystyle\leq LPe^{LT}\|h\|_{\infty}\sum_{k=1}^{n}h_{k}\leq LPTe^{LT}\|h\|_{\infty}=\tfrac{1}{\sqrt{2}}Te^{LT}\sqrt{LP\|\rho\|_{\infty}}.

It follows from (22) that 0Lhk140\leq Lh_{k}\leq\frac{1}{4}. Using (46), (45) and (24), we find

k=1neL(Ttk)(eLhk1)(Phk+(1+Lhk)2Lhkρk12)\displaystyle\sum_{k=1}^{n}e^{L(T-t_{k})}(e^{Lh_{k}}-1)\Big{(}Ph_{k}+\tfrac{(1+Lh_{k})^{2}}{Lh_{k}}\tfrac{\rho_{k-1}}{2}\Big{)} (53)
eLTk=1n(1+Lhk)Lhk(Phk+(1+Lhk)2Lhkρk12)\displaystyle\leq e^{LT}\sum_{k=1}^{n}(1+Lh_{k})Lh_{k}\Big{(}Ph_{k}+\tfrac{(1+Lh_{k})^{2}}{Lh_{k}}\tfrac{\rho_{k-1}}{2}\Big{)}
=eLTk=1n(1+Lhk)(LPhk2+(1+Lhk)2ρk12)\displaystyle=e^{LT}\sum_{k=1}^{n}(1+Lh_{k})\Big{(}LPh_{k}^{2}+(1+Lh_{k})^{2}\tfrac{\rho_{k-1}}{2}\Big{)}
e4LTk=1n(LPhk2+ρk12)=e4LT(k=1nLPhk2+ρ02+k=1n1LPhk2)\displaystyle\leq e^{4LT}\sum_{k=1}^{n}\Big{(}LPh_{k}^{2}+\tfrac{\rho_{k-1}}{2}\Big{)}=e^{4LT}\Big{(}\sum_{k=1}^{n}LPh_{k}^{2}+\tfrac{\rho_{0}}{2}+\sum_{k=1}^{n-1}LPh_{k}^{2}\Big{)}
e4LT(ρ+2LPk=1nhk2)e4LT(ρ+2LPTh)\displaystyle\leq e^{4LT}\Big{(}\|\rho\|_{\infty}+2LP\sum_{k=1}^{n}h_{k}^{2}\Big{)}\leq e^{4LT}\Big{(}\|\rho\|_{\infty}+2LPT\|h\|_{\infty}\Big{)}
=e4LT(ρ+T2LPρ).\displaystyle=e^{4LT}\Big{(}\|\rho\|_{\infty}+T\sqrt{2LP\|\rho\|_{\infty}}\Big{)}.

Combining (51), (52) and (53), we see that E(𝜹)ω1(ρ)E({\bm{\delta}})\leq\omega_{1}(\|\rho\|_{\infty}) with a function

ω1(s):=12eLTs+12TeLTLPs+e4LT(s+T2LPs),\omega_{1}(s):=\frac{1}{2}e^{LT}s+\tfrac{1}{\sqrt{2}}Te^{LT}\sqrt{LPs}+e^{4LT}\Big{(}s+T\sqrt{2LPs}\Big{)},

which is continuous on >0\mathbb{R}_{>0} and satisfies lims0ω1(s)=0\lim_{s\to 0}\omega_{1}(s)=0. ∎

Lemma 6.4.

The functions ψ\psi and EE given by (25) and (39) satisfy property P2 with n\aleph_{n} as in Definition 5.1.

Proof.

In view of inequalities (54) and (60) proved below, we have

ΔE(𝜹;k)min(14ρk,110L/(2P)ρk3/2)k[0,n],\Delta E({\bm{\delta}};k)\leq-\min(\tfrac{1}{4}\rho_{k},\tfrac{1}{10}\sqrt{L/(2P)}\rho_{k}^{3/2})\quad\forall\,k\in[0,n],

and the function ω2(s):=min(14ρk,110L/(2P)ρk3/2)\omega_{2}(s):=\min(\tfrac{1}{4}\rho_{k},\tfrac{1}{10}\sqrt{L/(2P)}\rho_{k}^{3/2}) is obviously continuous and strictly increasing with ω2(0)=0\omega_{2}(0)=0. We consider three different cases.

Case 1: When k=0k=0, comparing terms in the sum (39) yields

ΔE(𝜹;0)=E(ψ[𝜹;0])E(𝜹)\displaystyle\Delta E({\bm{\delta}};0)=E(\psi[{\bm{\delta}};0])-E({\bm{\delta}})
=0(ψ[𝜹;0])+1(ψ[𝜹;0])0(𝜹)1(𝜹)\displaystyle=\mathcal{E}_{0}(\psi[{\bm{\delta}};0])+\mathcal{E}_{1}(\psi[{\bm{\delta}};0])-\mathcal{E}_{0}({\bm{\delta}})-\mathcal{E}_{1}({\bm{\delta}})
=ρ08eLT+eL(Tt1)((eLh11)(Ph1+(1+Lh1)2Lh1ρ08)+χ1(ψ[𝜹;0]))\displaystyle=\tfrac{\rho_{0}}{8}e^{LT}+e^{L(T-t_{1})}\Big{(}(e^{Lh_{1}}-1)\big{(}Ph_{1}+\tfrac{(1+Lh_{1})^{2}}{Lh_{1}}\tfrac{\rho_{0}}{8}\big{)}+\chi_{1}(\psi[{\bm{\delta}};0])\Big{)}
ρ02eLTeL(Tt1)((eLh11)(Ph1+(1+Lh1)2Lh1ρ02)+χ1(𝜹))\displaystyle\qquad-\tfrac{\rho_{0}}{2}e^{LT}-e^{L(T-t_{1})}\Big{(}(e^{Lh_{1}}-1)\big{(}Ph_{1}+\tfrac{(1+Lh_{1})^{2}}{Lh_{1}}\tfrac{\rho_{0}}{2}\big{)}+\chi_{1}({\bm{\delta}})\Big{)}
=38(eLT+eL(Tt1)(eLh11)(1+Lh1)2Lh1)ρ0+eL(Tt1)(χ1(ψ[𝜹;0])χ1(𝜹)).\displaystyle=-\tfrac{3}{8}\big{(}e^{LT}+e^{L(T-t_{1})}(e^{Lh_{1}}-1)\tfrac{(1+Lh_{1})^{2}}{Lh_{1}}\big{)}\rho_{0}+e^{L(T-t_{1})}\big{(}\chi_{1}(\psi[{\bm{\delta}};0])-\chi_{1}({\bm{\delta}})\big{)}.

By (23), we have

χ1(ψ[𝜹;0])χ1(𝜹)\displaystyle\chi_{1}(\psi[{\bm{\delta}};0])-\chi_{1}({\bm{\delta}})
={ρ12,ρ04<ρ1,0,otherwise{ρ12,ρ0<ρ1,0,otherwise={ρ12,ρ0=ρ1,0,otherwiseρ02,\displaystyle=\begin{cases}\tfrac{\rho_{1}}{2},&\tfrac{\rho_{0}}{4}<\rho_{1},\\ 0,&\text{otherwise}\end{cases}-\begin{cases}\tfrac{\rho_{1}}{2},&\rho_{0}<\rho_{1},\\ 0,&\text{otherwise}\end{cases}=\begin{cases}\tfrac{\rho_{1}}{2},&\rho_{0}=\rho_{1},\\ 0,&\text{otherwise}\end{cases}\leq\tfrac{\rho_{0}}{2},

and using (45) we estimate

ΔE(𝜹;0)eL(Tt1)(3838(eLh11)(1+Lh1)2Lh1+12)ρ0\displaystyle\Delta E({\bm{\delta}};0)\leq e^{L(T-t_{1})}\big{(}-\tfrac{3}{8}-\tfrac{3}{8}(e^{Lh_{1}}-1)\tfrac{(1+Lh_{1})^{2}}{Lh_{1}}+\tfrac{1}{2}\big{)}\rho_{0} (54)
eL(Tt1)(3838(1+Lh1)2+12)ρ014eL(Tt1)ρ014ρ0.\displaystyle\leq e^{L(T-t_{1})}\big{(}-\tfrac{3}{8}-\tfrac{3}{8}(1+Lh_{1})^{2}+\tfrac{1}{2}\big{)}\rho_{0}\leq-\tfrac{1}{4}e^{L(T-t_{1})}\rho_{0}\leq-\tfrac{1}{4}\rho_{0}.

Case 2: When k(0,n)k\in(0,n), comparing terms in the sum (39) yields

ΔE(𝜹;k)=E(ψ[𝜹;k])E(𝜹)\displaystyle\Delta E({\bm{\delta}};k)=E(\psi[{\bm{\delta}};k])-E({\bm{\delta}}) (55)
=k(ψ[𝜹;k])+k+1(ψ[𝜹;k])+k+2(ψ[𝜹;k])k(𝜹)k+1(𝜹).\displaystyle=\mathcal{E}_{k}(\psi[{\bm{\delta}};k])+\mathcal{E}_{k+1}(\psi[{\bm{\delta}};k])+\mathcal{E}_{k+2}(\psi[{\bm{\delta}};k])-\mathcal{E}_{k}({\bm{\delta}})-\mathcal{E}_{k+1}({\bm{\delta}}).

Substituting the definitions and canceling terms yields

k+2(ψ[𝜹;k])k+1(𝜹)\displaystyle\mathcal{E}_{k+2}(\psi[{\bm{\delta}};k])-\mathcal{E}_{k+1}({\bm{\delta}})
=eL(Ttk+1)[(eLhk+11)(Phk+1+(1+Lhk+1)2Lhk+1ρk8)+χk+2(ψ[𝜹;k])]\displaystyle=e^{L(T-t_{k+1})}\Big{[}(e^{Lh_{k+1}}\!-\!1)\big{(}Ph_{k+1}+\tfrac{(1+Lh_{k+1})^{2}}{Lh_{k+1}}\tfrac{\rho_{k}}{8}\big{)}+\chi_{k+2}(\psi[{\bm{\delta}};k])\Big{]}
eL(Ttk+1)[(eLhk+11)(Phk+1+(1+Lhk+1)2Lhk+1ρk2)+χk+1(𝜹)]\displaystyle\qquad-e^{L(T-t_{k+1})}\Big{[}(e^{Lh_{k+1}}\!-\!1)\big{(}Ph_{k+1}+\tfrac{(1+Lh_{k+1})^{2}}{Lh_{k+1}}\tfrac{\rho_{k}}{2}\big{)}+\chi_{k+1}({\bm{\delta}})\Big{]}
=eL(Ttk+1)[(eLhk+11)(1+Lhk+1)2Lhk+1(ρk8ρk2)+(χk+2(ψ[𝜹;k])χk+1(𝜹))].\displaystyle=e^{L(T-t_{k+1})}\Big{[}(e^{Lh_{k+1}}-1)\tfrac{(1+Lh_{k+1})^{2}}{Lh_{k+1}}(\tfrac{\rho_{k}}{8}-\tfrac{\rho_{k}}{2})+(\chi_{k+2}(\psi[{\bm{\delta}};k])-\chi_{k+1}({\bm{\delta}}))\Big{]}.

Using (45), we immediately obtain

(eLhk+11)(1+Lhk+1)2Lhk+1(ρk8ρk2)38(1+Lhk+1)2ρk38ρk,(e^{Lh_{k+1}}-1)\tfrac{(1+Lh_{k+1})^{2}}{Lh_{k+1}}(\tfrac{\rho_{k}}{8}-\tfrac{\rho_{k}}{2})\leq-\tfrac{3}{8}(1+Lh_{k+1})^{2}\rho_{k}\leq-\tfrac{3}{8}\rho_{k},

and using (23), we see that

χk+2(ψ[𝜹;k])χk+1(𝜹)\displaystyle\chi_{k+2}(\psi[{\bm{\delta}};k])-\chi_{k+1}({\bm{\delta}})
={ρk+12,ρk4<ρk+1,0,otherwise{ρk+12,ρk<ρk+1,0,otherwise={ρk+12,ρk=ρk+1,0,otherwiseρk2.\displaystyle=\begin{cases}\frac{\rho_{k+1}}{2},&\frac{\rho_{k}}{4}<\rho_{k+1},\\ 0,&\text{otherwise}\end{cases}-\begin{cases}\frac{\rho_{k+1}}{2},&\rho_{k}<\rho_{k+1},\\ 0,&\text{otherwise}\end{cases}=\begin{cases}\frac{\rho_{k+1}}{2},&\rho_{k}=\rho_{k+1},\\ 0,&\text{otherwise}\end{cases}\leq\frac{\rho_{k}}{2}.

Combining the above yields

k+2(ψ[𝜹;k])k+1(𝜹)18eL(Ttk+1)ρk18eL(Ttk)ρk.\mathcal{E}_{k+2}(\psi[{\bm{\delta}};k])-\mathcal{E}_{k+1}({\bm{\delta}})\leq\tfrac{1}{8}e^{L(T-t_{k+1})}\rho_{k}\leq\tfrac{1}{8}e^{L(T-t_{k})}\rho_{k}. (56)

We estimate the remaining difference

k(ψ[𝜹;k])+k+1(ψ[𝜹;k])k(𝜹)\displaystyle\mathcal{E}_{k}(\psi[{\bm{\delta}};k])+\mathcal{E}_{k+1}(\psi[{\bm{\delta}};k])-\mathcal{E}_{k}({\bm{\delta}})
=eL(Ttk+hk2)((eLhk21)(Phk2+(1+Lhk2)2ρk1Lhk)+χk(ψ[𝜹;k]))\displaystyle=e^{L(T-t_{k}+\frac{h_{k}}{2})}\Big{(}(e^{\frac{Lh_{k}}{2}}-1)\big{(}P\tfrac{h_{k}}{2}+(1+\tfrac{Lh_{k}}{2})^{2}\tfrac{\rho_{k-1}}{Lh_{k}}\big{)}+\chi_{k}(\psi[{\bm{\delta}};k])\Big{)}
+eL(Ttk)((eLhk21)(Phk2+(1+Lhk2)2ρk4Lhk)+χk+1(ψ[𝜹;k]))\displaystyle\qquad+e^{L(T-t_{k})}\Big{(}(e^{\frac{Lh_{k}}{2}}-1)\big{(}P\tfrac{h_{k}}{2}+(1+\tfrac{Lh_{k}}{2})^{2}\tfrac{\rho_{k}}{4Lh_{k}}\big{)}+\chi_{k+1}(\psi[{\bm{\delta}};k])\Big{)}
eL(Ttk)((eLhk1)(Phk+(1+Lhk)2ρk12Lhk)+χk(𝜹))\displaystyle\qquad-e^{L(T-t_{k})}\Big{(}(e^{Lh_{k}}-1)\big{(}Ph_{k}+(1+Lh_{k})^{2}\tfrac{\rho_{k-1}}{2Lh_{k}}\big{)}+\chi_{k}({\bm{\delta}})\Big{)}

by breaking it into parts. Since χk+1(ψ[𝜹;k])=0\chi_{k+1}(\psi[{\bm{\delta}};k])=0, using (22), we obtain

eL(Ttk+hk2)χk(ψ[𝜹;k])+eL(Ttk)χk+1(ψ[𝜹;k])eL(Ttk)χk(𝜹)\displaystyle e^{L(T-t_{k}+\frac{h_{k}}{2})}\chi_{k}(\psi[{\bm{\delta}};k])+e^{L(T-t_{k})}\chi_{k+1}(\psi[{\bm{\delta}};k])-e^{L(T-t_{k})}\chi_{k}({\bm{\delta}}) (57)
=eL(Ttk)eLhk2{ρk8,ρk1<ρk4,0,otherwiseeL(Ttk){ρk2,ρk1<ρk,0,otherwise,\displaystyle=e^{L(T-t_{k})}e^{\frac{Lh_{k}}{2}}\begin{cases}\frac{\rho_{k}}{8},&\rho_{k-1}<\frac{\rho_{k}}{4},\\ 0,&\text{otherwise}\end{cases}-e^{L(T-t_{k})}\begin{cases}\tfrac{\rho_{k}}{2},&\rho_{k-1}<\rho_{k},\\ 0,&otherwise,\end{cases}
{eL(Ttk)(eLhk2ρk8ρk2),ρk1<ρk,0,otherwise{0.35eL(Ttk)ρk,ρk1<ρk,0,otherwise.\displaystyle\leavevmode\resizebox{422.77661pt}{}{$\leq\begin{cases}e^{L(T-t_{k})}(e^{\frac{Lh_{k}}{2}}\tfrac{\rho_{k}}{8}-\tfrac{\rho_{k}}{2}),&\rho_{k-1}<\rho_{k},\\ 0,&otherwise\end{cases}\leq\begin{cases}-0.35e^{L(T-t_{k})}\rho_{k},&\rho_{k-1}<\rho_{k},\\ 0,&otherwise.\end{cases}$}

Using (46), we estimate

eL(Ttk)(eLhk21)(1+Lhk2)2ρk4LhkeL(Ttk)(1+Lhk2)3ρk8.\displaystyle e^{L(T-t_{k})}(e^{\frac{Lh_{k}}{2}}-1)(1+\tfrac{Lh_{k}}{2})^{2}\tfrac{\rho_{k}}{4Lh_{k}}\leq e^{L(T-t_{k})}(1+\tfrac{Lh_{k}}{2})^{3}\tfrac{\rho_{k}}{8}.

Substituting eL(Ttk+hk2)=eL(Ttk)eLhk2e^{L(T-t_{k}+\frac{h_{k}}{2})}=e^{L(T-t_{k})}e^{\frac{Lh_{k}}{2}}, it follows from (47) with LhkLh_{k} in lieu of ss that

eL(Ttk+hk2)(eLhk21)(1+Lhk2)2ρk1LhkeL(Ttk)(eLhk1)(1+Lhk)2ρk12Lhk\displaystyle\leavevmode\resizebox{422.77661pt}{}{$e^{L(T-t_{k}+\frac{h_{k}}{2})}(e^{\frac{Lh_{k}}{2}}-1)(1+\tfrac{Lh_{k}}{2})^{2}\tfrac{\rho_{k-1}}{Lh_{k}}-e^{L(T-t_{k})}(e^{Lh_{k}}-1)(1+Lh_{k})^{2}\tfrac{\rho_{k-1}}{2Lh_{k}}$}
38LhkeL(Ttk)ρk1{0,ρk1<ρk,38LhkeL(Ttk)ρk,otherwise.\displaystyle\leq-\frac{3}{8}Lh_{k}e^{L(T-t_{k})}\rho_{k-1}\leq\begin{cases}0,&\rho_{k-1}<\rho_{k},\\ -\frac{3}{8}Lh_{k}e^{L(T-t_{k})}\rho_{k},&otherwise.\end{cases}

Finally, algebraic manipulations, (24) and (45) yield

eL(Ttk+hk2)(eLhk21)Phk2\displaystyle e^{L(T-t_{k}+\frac{h_{k}}{2})}(e^{\frac{Lh_{k}}{2}}-1)P\tfrac{h_{k}}{2} (58)
+eL(Ttk)(eLhk21)Phk2eL(Ttk)(eLhk1)Phk\displaystyle\qquad+e^{L(T-t_{k})}(e^{\frac{Lh_{k}}{2}}-1)P\tfrac{h_{k}}{2}-e^{L(T-t_{k})}(e^{Lh_{k}}-1)Ph_{k}
=eL(Ttk)(eLhk1)Phk2\displaystyle=-e^{L(T-t_{k})}(e^{Lh_{k}}-1)P\tfrac{h_{k}}{2}
=eL(Ttk)(eLhk1)ρk4LhkeL(Ttk)ρk4.\displaystyle=-e^{L(T-t_{k})}(e^{Lh_{k}}-1)\tfrac{\rho_{k}}{4Lh_{k}}\leq-e^{L(T-t_{k})}\tfrac{\rho_{k}}{4}.

Because of (22) we have 38Lhk<0.35\tfrac{3}{8}Lh_{k}<0.35, so summing up inequalities (57) to (58) and expanding the cubic term yields

k(ψ[𝜹;k])+k+1(ψ[𝜹;k])k(𝜹)\displaystyle\mathcal{E}_{k}(\psi[{\bm{\delta}};k])+\mathcal{E}_{k+1}(\psi[{\bm{\delta}};k])-\mathcal{E}_{k}({\bm{\delta}}) (59)
eL(Ttk)[(1+Lhk2)3ρk8ρk4+max{0.35ρk,38Lhkρk}]\displaystyle\leq e^{L(T-t_{k})}\big{[}(1+\tfrac{Lh_{k}}{2})^{3}\tfrac{\rho_{k}}{8}-\tfrac{\rho_{k}}{4}+\max\{-0.35\rho_{k},-\tfrac{3}{8}Lh_{k}\rho_{k}\}\big{]}
=eL(Ttk)ρk[18316Lhk+332(Lhk)2+164(Lhk)3].\displaystyle=e^{L(T-t_{k})}\rho_{k}\left[-\tfrac{1}{8}-\tfrac{3}{16}Lh_{k}+\tfrac{3}{32}(Lh_{k})^{2}+\tfrac{1}{64}(Lh_{k})^{3}\right].

In view of (55), combining inequalities (56) and (59), and using (22) and (24) yields

ΔE(𝜹;k)\displaystyle\Delta E({\bm{\delta}};k) eL(Ttk)ρkLhk[316+332Lhk+164(Lhk)2]\displaystyle\leq e^{L(T-t_{k})}\rho_{k}Lh_{k}[-\tfrac{3}{16}+\tfrac{3}{32}Lh_{k}+\tfrac{1}{64}(Lh_{k})^{2}] (60)
<110eL(Ttk)ρkLhk110L/(2P)ρk3/2.\displaystyle<-\tfrac{1}{10}e^{L(T-t_{k})}\rho_{k}Lh_{k}\leq-\tfrac{1}{10}\sqrt{L/(2P)}\rho_{k}^{3/2}.

Case 3: When k=nk=n, then comparing terms in the sum (39) yields

ΔE(𝜹;n)=E(ψ[𝜹;n])E(𝜹)=n(ψ[𝜹;n])+n+1(ψ[𝜹;n])n(𝜹).\Delta E({\bm{\delta}};n)=E(\psi[{\bm{\delta}};n])-E({\bm{\delta}})=\mathcal{E}_{n}(\psi[{\bm{\delta}};n])+\mathcal{E}_{n+1}(\psi[{\bm{\delta}};n])-\mathcal{E}_{n}({\bm{\delta}}).

Since estimate (59) holds for k=nk=n for the same reasons as above, following the same process as in (60) yields the same result for k=nk=n. ∎

Lemma 6.5.

Let vC0([0,T],>0)v\in C^{0}([0,T],\mathbb{R}_{>0}). Then there exists c>0c>0 such that for any n1n\in\mathbb{N}_{1} and any 𝛅=(𝐡,𝐭,𝛒)n{\bm{\delta}}=(\bm{h},\bm{t},\bm{\rho})\in\aleph_{n}, the functions ψ\psi and CC given by (25) and (44) satisfy

c(miniρi)dRdF/2ΔC(𝜹;v;argminiρi).c(\min_{i}\rho_{i})^{-d_{R}-d_{F}/2}\leq\Delta C({\bm{\delta}};v;\operatorname{argmin}_{i}\rho_{i}). (61)

In particular, the function C(;v):>0C(\,\cdot\,;v):\aleph\to\mathbb{R}_{>0} satisfies property P3.

Proof.

Defining

VL:=inft[0,T]v(t)>0,V_{L}:=\inf_{t\in[0,T]}v(t)>0,

we will verify statement (61) with

c:=min(7(4dR+dF1)VL(2LP)dF/2,22dR+dF7VL(2LP)dF/2).c:=\min(7\tfrac{(4^{d_{R}+d_{F}}-1)V_{L}}{(2LP)^{d_{F}/2}},\tfrac{2^{2d_{R}+d_{F}}7V_{L}}{(2LP)^{d_{F}/2}}).

Take

k:=argminiρi.k:=\operatorname{argmin}_{i}\rho_{i}. (62)

Case 1: When k=0k=0, comparing terms in the sum (44), using (49) twice, along with (24) and (62) yields

ΔC(𝜹;v;0)=C(ψ[𝜹;0];v)C(𝜹;v)=𝒞0(ψ[𝜹;0];v)𝒞0(𝜹;v)\displaystyle\Delta C({\bm{\delta}};v;0)=C(\psi[{\bm{\delta}};0];v)-C({\bm{\delta}};v)=\mathcal{C}_{0}(\psi[{\bm{\delta}};0];v)-\mathcal{C}_{0}({\bm{\delta}};v)
=4v(0)h1dF(ρ04)dRmin(ρ04,ρ1)dF+1σ(h1,ρ04)4v(0)h1dFρ0dRmin(ρ0,ρ1)dF+1σ(h1,ρ0)\displaystyle=\tfrac{4v(0)h_{1}^{d_{F}}}{(\frac{\rho_{0}}{4})^{d_{R}}\min(\frac{\rho_{0}}{4},\rho_{1})^{d_{F}+1}}\sigma(h_{1},\tfrac{\rho_{0}}{4})-\tfrac{4v(0)h_{1}^{d_{F}}}{\rho_{0}^{d_{R}}\min(\rho_{0},\rho_{1})^{d_{F}+1}}\sigma(h_{1},\rho_{0})
=4v(0)h1dF(σ(h1,ρ04)(ρ04)dR(ρ04)dF+1σ(h1,ρ0)(ρ0)dR(ρ0)dF+1)\displaystyle=4v(0)h_{1}^{d_{F}}\left(\tfrac{\sigma(h_{1},\frac{\rho_{0}}{4})}{(\frac{\rho_{0}}{4})^{d_{R}}(\frac{\rho_{0}}{4})^{d_{F}+1}}-\tfrac{\sigma(h_{1},\rho_{0})}{(\rho_{0})^{d_{R}}(\rho_{0})^{d_{F}+1}}\right)
=4v(0)h1dFρ0dR+dF+1(4dR+dF+1σ(h1,ρ04)σ(h1,ρ0))\displaystyle=\tfrac{4v(0)h_{1}^{d_{F}}}{\rho_{0}^{d_{R}+d_{F}+1}}\left(4^{d_{R}+d_{F}+1}\sigma(h_{1},\tfrac{\rho_{0}}{4})-\sigma(h_{1},\rho_{0})\right)
=4v(0)h1dFρ0dR+dF+1(4dR+dF1)σ(h1,ρ0)7v(0)h1dF(4dR+dF1)ρ0dRdF\displaystyle=\tfrac{4v(0)h_{1}^{d_{F}}}{\rho_{0}^{d_{R}+d_{F}+1}}(4^{d_{R}+d_{F}}-1)\sigma(h_{1},\rho_{0})\geq 7v(0)h_{1}^{d_{F}}(4^{d_{R}+d_{F}}-1)\rho_{0}^{-d_{R}-d_{F}}
=7(4dR+dF1)VL(2LP)dF/2ρ1dF/2ρ0dRdF7(4dR+dF1)VL(2LP)dF/2ρ0dRdF/2.\displaystyle=7\tfrac{(4^{d_{R}+d_{F}}-1)V_{L}}{(2LP)^{d_{F}/2}}\rho_{1}^{d_{F}/2}\rho_{0}^{-d_{R}-d_{F}}\geq 7\tfrac{(4^{d_{R}+d_{F}}-1)V_{L}}{(2LP)^{d_{F}/2}}\rho_{0}^{-d_{R}-d_{F}/2}.

Case 2: When k[1,n2]k\in[1,n-2], comparing terms in the sum (44) gives

ΔC(𝜹;v;k)=i=0n1𝒞i(ψ[𝜹;k];v)i=0n1𝒞i(𝜹;v)\displaystyle\Delta C({\bm{\delta}};v;k)=\sum_{i=0}^{n-1}\mathcal{C}_{i}(\psi[{\bm{\delta}};k];v)-\sum_{i=0}^{n-1}\mathcal{C}_{i}({\bm{\delta}};v)
=𝒞k1(ψ[𝜹;k];v)+𝒞k(ψ[𝜹;k];v)+𝒞k+1(ψ[𝜹;k];v)𝒞k1(𝜹;v)𝒞k(𝜹;v).\displaystyle=\mathcal{C}_{k-1}(\psi[{\bm{\delta}};k];v)+\mathcal{C}_{k}(\psi[{\bm{\delta}};k];v)+\mathcal{C}_{k+1}(\psi[{\bm{\delta}};k];v)-\mathcal{C}_{k-1}({\bm{\delta}};v)-\mathcal{C}_{k}({\bm{\delta}};v).

Using (62), rearranging, and using (50), we obtain

Ck1(ψ[𝜹;k];v)Ck1(𝜹;v)\displaystyle C_{k-1}(\psi[{\bm{\delta}};k];v)-C_{k-1}({\bm{\delta}};v) (63)
=4v(tk1)(hk2)dFρk1dRmin(ρk1,ρk4)dF+1σ(hk2,ρk1)4v(tk1)hkdFρk1dRmin(ρk1,ρk)dF+1σ(hk,ρk1)\displaystyle=\tfrac{4v(t_{k-1})(\frac{h_{k}}{2})^{d_{F}}}{\rho_{k-1}^{d_{R}}\min(\rho_{k-1},\frac{\rho_{k}}{4})^{d_{F}+1}}\sigma(\tfrac{h_{k}}{2},\rho_{k-1})-\tfrac{4v(t_{k-1})h_{k}^{d_{F}}}{\rho_{k-1}^{d_{R}}\min(\rho_{k-1},\rho_{k})^{d_{F}+1}}\sigma(h_{k},\rho_{k-1})
=4v(tk1)(hk2)dFρk1dR(ρk4)dF+1σ(hk2,ρk1)4v(tk1)hkdFρk1dRρkdF+1σ(hk,ρk1)\displaystyle=\tfrac{4v(t_{k-1})(\frac{h_{k}}{2})^{d_{F}}}{\rho_{k-1}^{d_{R}}(\frac{\rho_{k}}{4})^{d_{F}+1}}\sigma(\tfrac{h_{k}}{2},\rho_{k-1})-\tfrac{4v(t_{k-1})h_{k}^{d_{F}}}{\rho_{k-1}^{d_{R}}\rho_{k}^{d_{F}+1}}\sigma(h_{k},\rho_{k-1})
=4v(tk1)hkdFρk1dRρkdF+1(2dF+2σ(hk2,ρk1)σ(hk,ρk1))\displaystyle=\tfrac{4v(t_{k-1})h_{k}^{d_{F}}}{\rho_{k-1}^{d_{R}}\rho_{k}^{d_{F}+1}}\left(2^{d_{F}+2}\sigma(\tfrac{h_{k}}{2},\rho_{k-1})-\sigma(h_{k},\rho_{k-1})\right)
4v(tk1)hkdFρk1dRρkdF+1(2dF+11)σ(hk,ρk1)>0.\displaystyle\geq\tfrac{4v(t_{k-1})h_{k}^{d_{F}}}{\rho_{k-1}^{d_{R}}\rho_{k}^{d_{F}+1}}(2^{d_{F}+1}-1)\sigma(h_{k},\rho_{k-1})>0.

Similarly, using (62), rearranging, and using (49) provides

Ck+1(ψ[𝜹;k];v)Ck(𝜹;v)\displaystyle C_{k+1}(\psi[{\bm{\delta}};k];v)-C_{k}({\bm{\delta}};v)
=4v(tk)hk+1dF(ρk4)dRmin(ρk4,ρk+1)dF+1σ(hk+1,ρk4)4v(tk)hk+1dFρkdRmin(ρk,ρk+1)dF+1σ(hk+1,ρk)\displaystyle=\tfrac{4v(t_{k})h_{k+1}^{d_{F}}}{(\frac{\rho_{k}}{4})^{d_{R}}\min(\frac{\rho_{k}}{4},\rho_{k+1})^{d_{F}+1}}\sigma(h_{k+1},\tfrac{\rho_{k}}{4})-\tfrac{4v(t_{k})h_{k+1}^{d_{F}}}{\rho_{k}^{d_{R}}\min(\rho_{k},\rho_{k+1})^{d_{F}+1}}\sigma(h_{k+1},\rho_{k})
=4v(tk)hk+1dF(ρk4)dR(ρk4)dF+1σ(hk+1,ρk4)4v(tk)hk+1dFρkdRρkdF+1σ(hk+1,ρk)\displaystyle=\tfrac{4v(t_{k})h_{k+1}^{d_{F}}}{(\frac{\rho_{k}}{4})^{d_{R}}(\frac{\rho_{k}}{4})^{d_{F}+1}}\sigma(h_{k+1},\tfrac{\rho_{k}}{4})-\tfrac{4v(t_{k})h_{k+1}^{d_{F}}}{\rho_{k}^{d_{R}}\rho_{k}^{d_{F}+1}}\sigma(h_{k+1},\rho_{k})
=4v(tk)hk+1dFρkdR+dF+1(4dR+dF+1σ(hk+1,ρk4)σ(hk+1,ρk))\displaystyle=\tfrac{4v(t_{k})h_{k+1}^{d_{F}}}{\rho_{k}^{d_{R}+d_{F}+1}}\left(4^{d_{R}+d_{F}+1}\sigma(h_{k+1},\tfrac{\rho_{k}}{4})-\sigma(h_{k+1},\rho_{k})\right)
=4v(tk)hk+1dFρkdR+dF+1(4dR+dF1)σ(hk+1,ρk)>0.\displaystyle=\tfrac{4v(t_{k})h_{k+1}^{d_{F}}}{\rho_{k}^{d_{R}+d_{F}+1}}(4^{d_{R}+d_{F}}-1)\sigma(h_{k+1},\rho_{k})>0.

Finally, elementary computations in combination with (49) and (24) yield

Ck(ψ[𝜹;k];v)=4v(tkhk2)(hk2)dF(ρk4)dRmin(ρk4,ρk4)dF+1σ(hk2,ρk4)\displaystyle C_{k}(\psi[{\bm{\delta}};k];v)=\tfrac{4v(t_{k}-\frac{h_{k}}{2})(\frac{h_{k}}{2})^{d_{F}}}{(\frac{\rho_{k}}{4})^{d_{R}}\min(\frac{\rho_{k}}{4},\frac{\rho_{k}}{4})^{d_{F}+1}}\sigma(\tfrac{h_{k}}{2},\tfrac{\rho_{k}}{4}) (64)
=22dR+dF+4v(tkhk2)hkdFρkdR+dF+1σ(hk2,ρk4)\displaystyle=2^{2d_{R}+d_{F}+4}\tfrac{v(t_{k}-\frac{h_{k}}{2})h_{k}^{d_{F}}}{\rho_{k}^{d_{R}+d_{F}+1}}\sigma(\tfrac{h_{k}}{2},\tfrac{\rho_{k}}{4})
22dR+dF7v(tkhk2)hkdFρkdR+dF22dR+dF7VL(2LP)dF/2ρkdRdF/2.\displaystyle\geq 2^{2d_{R}+d_{F}}\tfrac{7v(t_{k}-\frac{h_{k}}{2})h_{k}^{d_{F}}}{\rho_{k}^{d_{R}+d_{F}}}\geq\tfrac{2^{2d_{R}+d_{F}}7V_{L}}{(2LP)^{d_{F}/2}}\rho_{k}^{-d_{R}-d_{F}/2}.

Combining equations (63) - (64) implies (61).

Case 3: If k=n1k=n-1, we have

ΔC(𝜹;v;k)=𝒞k1(ψ[𝜹;k];v)+𝒞k(ψ[𝜹;k];v)𝒞k1(𝜹;v).\Delta C({\bm{\delta}};v;k)=\mathcal{C}_{k-1}(\psi[{\bm{\delta}};k];v)+\mathcal{C}_{k}(\psi[{\bm{\delta}};k];v)-\mathcal{C}_{k-1}({\bm{\delta}};v).

Equations (63) and (64) apply for the same reasons as in case 2, and so (61) holds. ∎

Lemma 6.6.

Let vC0([0,T],>0)v\in C^{0}([0,T],\mathbb{R}_{>0}). Then the function C(;v)C(\,\cdot\,;v) from (44) satisfies property P4, with ψ\psi given by (25).

Proof.

Once again, we denote

k:=argmini[0,n]ρi.k:=\operatorname{argmin}_{i\in[0,n]}\rho_{i}. (65)

Defining

VU:=supt[0,T]v(t)>0,V_{U}:=\sup_{t\in[0,T]}v(t)\in\mathbb{R}_{>0},

we aim to show that

ΔC(𝜹;v;j)ρkdRdF/2(24dR+dF+2σ¯VUTdFρjdF/2+1+4dR+1σ¯VU(2LP)dF/2ρj+4dF+2σ¯VUTdFρjdF/2+1)\Delta C({\bm{\delta}};v;j)\leq\rho_{k}^{-d_{R}-d_{F}/2}\big{(}2\tfrac{4^{d_{R}+d_{F}+2}\bar{\sigma}V_{U}T^{d_{F}}}{\rho_{j}^{d_{F}/2+1}}+\tfrac{4^{d_{R}+1}\bar{\sigma}V_{U}}{(2LP)^{d_{F}/2}\rho_{j}}+\tfrac{4^{d_{F}+2}\bar{\sigma}V_{U}T^{d_{F}}}{\rho_{j}^{d_{F}/2+1}}\big{)} (66)

holds for all j[0,n]j\in[0,n]. Combining (66) with Lemma 6.5, which states that there exists c>0c>0 such that

cρkdRdF/2ΔC(𝜹;v;k),c\rho_{k}^{-d_{R}-d_{F}/2}\leq\Delta C({\bm{\delta}};v;k),

then yields property P4 with

ω4(s):=1c(24dR+dF+2σ¯VUTdFsdF/2+1+4dR+1σ¯VU(2LP)dF/2s+4dF+2σ¯VUTdFsdF/2+1).\omega_{4}(s):=\tfrac{1}{c}\big{(}2\tfrac{4^{d_{R}+d_{F}+2}\bar{\sigma}V_{U}T^{d_{F}}}{s^{d_{F}/2+1}}+\tfrac{4^{d_{R}+1}\bar{\sigma}V_{U}}{(2LP)^{d_{F}/2}s}+\tfrac{4^{d_{F}+2}\bar{\sigma}V_{U}T^{d_{F}}}{s^{d_{F}/2+1}}\big{)}.

We proceed to verify statement (66).

Case 1: When j=0j=0, comparing terms in the sum (44), and using (48) and (65) yields

ΔC(𝜹;v;0)=𝒞0(ψ[𝜹;0];v)𝒞0(𝜹;v)𝒞0(ψ[𝜹;0];v)\displaystyle\Delta C({\bm{\delta}};v;0)=\mathcal{C}_{0}(\psi[{\bm{\delta}};0];v)-\mathcal{C}_{0}({\bm{\delta}};v)\leq\mathcal{C}_{0}(\psi[{\bm{\delta}};0];v)
=4v(0)h1dF(ρ04)dRmin(ρ04,ρ1)dF+1σ(h1,ρ04)4σ¯VUh1dF(ρ04)dRmin(ρ04,ρ1)dF+1\displaystyle=\tfrac{4v(0)h_{1}^{d_{F}}}{(\frac{\rho_{0}}{4})^{d_{R}}\min(\frac{\rho_{0}}{4},\rho_{1})^{d_{F}+1}}\sigma(h_{1},\tfrac{\rho_{0}}{4})\leq\tfrac{4\bar{\sigma}V_{U}h_{1}^{d_{F}}}{(\frac{\rho_{0}}{4})^{d_{R}}\min(\frac{\rho_{0}}{4},\rho_{1})^{d_{F}+1}}
={4dR+dF+2σ¯VUh1dFρ0dR+dF+1,ρ04ρ1,4dR+1σ¯VUh1dFρ0dRρ1dF+1,ρ04>ρ1{4dR+dF+2σ¯VUTdFρ0dF/2+1ρkdR+dF/2,ρ04ρ1,4dR+1σ¯VU(2LP)dF/2ρ0ρkdR+dF/2,ρ04>ρ1,\displaystyle=\begin{cases}\tfrac{4^{d_{R}+d_{F}+2}\bar{\sigma}V_{U}h_{1}^{d_{F}}}{\rho_{0}^{d_{R}+d_{F}+1}},&\frac{\rho_{0}}{4}\leq\rho_{1},\\ \tfrac{4^{d_{R}+1}\bar{\sigma}V_{U}h_{1}^{d_{F}}}{\rho_{0}^{d_{R}}\rho_{1}^{d_{F}+1}},&\frac{\rho_{0}}{4}>\rho_{1}\end{cases}\leq\begin{cases}\tfrac{4^{d_{R}+d_{F}+2}\bar{\sigma}V_{U}T^{d_{F}}}{\rho_{0}^{d_{F}/2+1}\rho_{k}^{d_{R}+d_{F}/2}},&\frac{\rho_{0}}{4}\leq\rho_{1},\\ \tfrac{4^{d_{R}+1}\bar{\sigma}V_{U}}{(2LP)^{d_{F}/2}\rho_{0}\rho_{k}^{d_{R}+d_{F}/2}},&\frac{\rho_{0}}{4}>\rho_{1},\end{cases}

where the last inequality in the case ρ04>ρ1\frac{\rho_{0}}{4}>\rho_{1} is derived from (24) and (65) by the computation

4dR+1σ¯VUh1dFρ0dRρ1dF+1=4dR+1σ¯VU(2LP)dF/2ρ0dRρ1dF/2+14dR+1σ¯VU(2LP)dF/2ρ0ρkdR+dF/2.\displaystyle\tfrac{4^{d_{R}+1}\bar{\sigma}V_{U}h_{1}^{d_{F}}}{\rho_{0}^{d_{R}}\rho_{1}^{d_{F}+1}}=\tfrac{4^{d_{R}+1}\bar{\sigma}V_{U}}{(2LP)^{d_{F}/2}\rho_{0}^{d_{R}}\rho_{1}^{d_{F}/2+1}}\leq\tfrac{4^{d_{R}+1}\bar{\sigma}V_{U}}{(2LP)^{d_{F}/2}\rho_{0}\rho_{k}^{d_{R}+d_{F}/2}}.

Case 2: When j[1,n1]j\in[1,n-1], comparing terms in the sum (44) yields

ΔC(𝜹;v;j)\displaystyle\Delta C({\bm{\delta}};v;j) =𝒞j1(ψ[𝜹;j];v)+𝒞j(ψ[𝜹;j];v)\displaystyle=\mathcal{C}_{j-1}(\psi[{\bm{\delta}};j];v)+\mathcal{C}_{j}(\psi[{\bm{\delta}};j];v)
+𝒞j+1(ψ[𝜹;j];v)𝒞j1(𝜹;v)𝒞j(𝜹;v).\displaystyle\quad+\mathcal{C}_{j+1}(\psi[{\bm{\delta}};j];v)-\mathcal{C}_{j-1}({\bm{\delta}};v)-\mathcal{C}_{j}({\bm{\delta}};v).

If ρj1ρj4\rho_{j-1}\leq\frac{\rho_{j}}{4}, then Lemma 6.2 yields

𝒞j1(ψ[𝜹;j];v)𝒞j1(𝜹;v)\displaystyle\mathcal{C}_{j-1}(\psi[{\bm{\delta}};j];v)-\mathcal{C}_{j-1}({\bm{\delta}};v) (67)
=4v(tj1)(hj2)dFρj1dRmin(ρj1,ρj4)dF+1σ(hj2,ρj1)4v(tj1)hjdFρj1dRmin(ρj1,ρj)dF+1σ(hj,ρj1)\displaystyle=\tfrac{4v(t_{j-1})(\frac{h_{j}}{2})^{d_{F}}}{\rho_{j-1}^{d_{R}}\min(\rho_{j-1},\frac{\rho_{j}}{4})^{d_{F}+1}}\sigma(\tfrac{h_{j}}{2},\rho_{j-1})-\tfrac{4v(t_{j-1})h_{j}^{d_{F}}}{\rho_{j-1}^{d_{R}}\min(\rho_{j-1},\rho_{j})^{d_{F}+1}}\sigma(h_{j},\rho_{j-1})
4v(tj1)(hj2)dFρj1dR+dF+1σ(hj,ρj1)4v(tj1)hjdFρj1dR+dF+1σ(hj,ρj1)0.\displaystyle\leq\tfrac{4v(t_{j-1})(\frac{h_{j}}{2})^{d_{F}}}{\rho_{j-1}^{d_{R}+d_{F}+1}}\sigma(h_{j},\rho_{j-1})-\tfrac{4v(t_{j-1})h_{j}^{d_{F}}}{\rho_{j-1}^{d_{R}+d_{F}+1}}\sigma(h_{j},\rho_{j-1})\leq 0.

If, on the other hand, we have ρj1>ρj4\rho_{j-1}>\frac{\rho_{j}}{4}, then (48) and (65) yield

𝒞j1(ψ[𝜹;j];v)𝒞j1(𝜹;v)𝒞j1(ψ[𝜹;j];v)\displaystyle\mathcal{C}_{j-1}(\psi[{\bm{\delta}};j];v)-\mathcal{C}_{j-1}({\bm{\delta}};v)\leq\mathcal{C}_{j-1}(\psi[{\bm{\delta}};j];v) (68)
=4v(tj1)(hj2)dFρj1dRmin(ρj1,ρj4)dF+1σ(hj2,ρj1)4dF+2σ¯VUTdFρj1dRρjdF+14dF+2σ¯VUTdFρjdF/2+1ρkdR+dF/2.\displaystyle=\tfrac{4v(t_{j-1})(\frac{h_{j}}{2})^{d_{F}}}{\rho_{j-1}^{d_{R}}\min(\rho_{j-1},\frac{\rho_{j}}{4})^{d_{F}+1}}\sigma(\tfrac{h_{j}}{2},\rho_{j-1})\leq\tfrac{4^{d_{F}+2}\bar{\sigma}V_{U}T^{d_{F}}}{\rho_{j-1}^{d_{R}}\rho_{j}^{d_{F}+1}}\leq\tfrac{4^{d_{F}+2}\bar{\sigma}V_{U}T^{d_{F}}}{\rho_{j}^{d_{F}/2+1}\rho_{k}^{d_{R}+d_{F}/2}}.

Again by (48) and (65), we obtain

𝒞j(ψ[𝜹;j];v)\displaystyle\mathcal{C}_{j}(\psi[{\bm{\delta}};j];v) =4v(tj1+hj2)(hj2)dF(ρj4)dRmin(ρj4,ρj4)dF+1σ(hj2,ρj4)\displaystyle=\tfrac{4v(t_{j-1}+\frac{h_{j}}{2})(\frac{h_{j}}{2})^{d_{F}}}{(\frac{\rho_{j}}{4})^{d_{R}}\min(\frac{\rho_{j}}{4},\frac{\rho_{j}}{4})^{d_{F}+1}}\sigma(\tfrac{h_{j}}{2},\tfrac{\rho_{j}}{4}) (69)
4σ¯VUTdF(ρj4)dR+dF+14dR+dF+2σ¯VUTdFρjdR+dF+14dR+dF+2σ¯VUTdFρjdF/2+1ρkdR+dF/2.\displaystyle\leq\tfrac{4\bar{\sigma}V_{U}T^{d_{F}}}{(\frac{\rho_{j}}{4})^{d_{R}+d_{F}+1}}\leq\tfrac{4^{d_{R}+d_{F}+2}\bar{\sigma}V_{U}T^{d_{F}}}{\rho_{j}^{d_{R}+d_{F}+1}}\leq\tfrac{4^{d_{R}+d_{F}+2}\bar{\sigma}V_{U}T^{d_{F}}}{\rho_{j}^{d_{F}/2+1}\rho_{k}^{d_{R}+d_{F}/2}}.

If ρj4ρj+1\frac{\rho_{j}}{4}\leq\rho_{j+1}, then (48) and (65) give

𝒞j+1(ψ[𝜹;j];v)𝒞j(𝜹;v)𝒞j+1(ψ[𝜹;j];v)\displaystyle\mathcal{C}_{j+1}(\psi[{\bm{\delta}};j];v)-\mathcal{C}_{j}({\bm{\delta}};v)\leq\mathcal{C}_{j+1}(\psi[{\bm{\delta}};j];v)
=4v(tj)hj+1dF(ρj4)dRmin(ρj4,ρj+1)dF+1σ(hj+1,ρj4)\displaystyle=\tfrac{4v(t_{j})h_{j+1}^{d_{F}}}{(\frac{\rho_{j}}{4})^{d_{R}}\min(\frac{\rho_{j}}{4},\rho_{j+1})^{d_{F}+1}}\sigma(h_{j+1},\tfrac{\rho_{j}}{4})
4dR+dF+2σ¯VUTdFρjdR+dF+14dR+dF+2σ¯VUTdFρjdF/2+1ρkdR+dF/2.\displaystyle\leq\tfrac{4^{d_{R}+d_{F}+2}\bar{\sigma}V_{U}T^{d_{F}}}{\rho_{j}^{d_{R}+d_{F}+1}}\leq\tfrac{4^{d_{R}+d_{F}+2}\bar{\sigma}V_{U}T^{d_{F}}}{\rho_{j}^{d_{F}/2+1}\rho_{k}^{d_{R}+d_{F}/2}}.

If, on the other hand, we have ρj4>ρj+1\frac{\rho_{j}}{4}>\rho_{j+1}, then (48) and (65) yield

𝒞j+1(ψ[𝜹;j];v)𝒞j(𝜹;v)𝒞j+1(ψ[𝜹;j];v)\displaystyle\mathcal{C}_{j+1}(\psi[{\bm{\delta}};j];v)-\mathcal{C}_{j}({\bm{\delta}};v)\leq\mathcal{C}_{j+1}(\psi[{\bm{\delta}};j];v)
=4v(tj)hj+1dF(ρj4)dRmin(ρj4,ρj+1)dF+1σ(hj+1,ρj4)4dRσ¯VUhj+1dFρjdRρj+1dF+1\displaystyle=\tfrac{4v(t_{j})h_{j+1}^{d_{F}}}{(\frac{\rho_{j}}{4})^{d_{R}}\min(\frac{\rho_{j}}{4},\rho_{j+1})^{d_{F}+1}}\sigma(h_{j+1},\tfrac{\rho_{j}}{4})\leq\tfrac{4^{d_{R}}\bar{\sigma}V_{U}h_{j+1}^{d_{F}}}{\rho_{j}^{d_{R}}\rho_{j+1}^{d_{F}+1}}
=4dRσ¯VU(2LP)dF/2ρjdRρj+1dF/2+14dRσ¯VU(2LP)dF/2ρjρkdR+dF/2.\displaystyle=\tfrac{4^{d_{R}}\bar{\sigma}V_{U}}{(2LP)^{d_{F}/2}\rho_{j}^{d_{R}}\rho_{j+1}^{d_{F}/2+1}}\leq\tfrac{4^{d_{R}}\bar{\sigma}V_{U}}{(2LP)^{d_{F}/2}\rho_{j}\rho_{k}^{d_{R}+d_{F}/2}}.

Case 3: When j=nj=n, comparing terms in the sum (44) yields

ΔC(𝜹;v;n)\displaystyle\Delta C({\bm{\delta}};v;n) =𝒞n1(ψ[𝜹;n];v)+𝒞n(ψ[𝜹;n];v)𝒞n1(𝜹;v).\displaystyle=\mathcal{C}_{n-1}(\psi[{\bm{\delta}};n];v)+\mathcal{C}_{n}(\psi[{\bm{\delta}};n];v)-\mathcal{C}_{n-1}({\bm{\delta}};v).

Statements (67), (68) and (69) hold with nn in lieu of jj for the same reasons as in case 2. ∎

We summarize the results of this section.

Theorem 6.7.

For any max1\ell_{\max}\in\mathbb{N}_{1} and ε1>ε2>>εmax>0\varepsilon_{1}>\varepsilon_{2}>\ldots>\varepsilon_{\ell_{\max}}>0, Algorithm 4 terminates in finite time and returns a 𝛅{\bm{\delta}}\in\aleph with E(𝛅)εmaxE({\bm{\delta}})\leq\varepsilon_{\ell_{\max}} and the corresponding reachable sets generated by recursion (17).

Proof.

Algorithm 4 is initialized with the values n0:=4LTn_{0}:=\lceil 4LT\rceil, 𝒉(0):=Tn0𝟙n0\bm{h}^{(0)}:=\tfrac{T}{n_{0}}\mathbbm{1}_{n_{0}}, 𝒕(0):=Σ+𝒉(0)\bm{t}^{(0)}:=\Sigma_{+}\bm{h}^{(0)} and 𝝆(0):=(2LPT2/n02)𝟙n0+1\bm{\rho}^{(0)}:=(2LPT^{2}/n_{0}^{2})\mathbbm{1}_{n_{0}+1}. It is easy to verify that we have 𝜹(0):=(𝒉(0),𝒕(0),𝝆(0))n0{\bm{\delta}}^{(0)}:=(\bm{h}^{(0)},\bm{t}^{(0)},\bm{\rho}^{(0)})\in\aleph_{n_{0}}.

Assume that Algorithm 4 has generated n1n_{\ell}\in\mathbb{N}_{1} and 𝜹()n{\bm{\delta}}^{(\ell)}\in\aleph_{n_{\ell}} for some [0,max1]\ell\in[0,\ell_{\max}-1]. By Lemma 3.2, the computation in line 4 of Algorithm 4 yields a pair (D00,D10)bdcρ0(D_{0}^{0},D_{1}^{0})\in\operatorname{bdc}_{\rho_{0}}, which is a valid input for Algorithm 2. In view of Algorithm 2, the product v():=vR()vF()v^{(\ell)}:=v_{R}^{(\ell)}v_{F}^{(\ell)} satisfies mint[0,T]v()(t)>0\min_{t\in[0,T]}v^{(\ell)}(t)>0 for all t[0,T]t\in[0,T], so by Lemmas 6.3 - 6.6 the functions C(;v()):>0C(\,\cdot\,;v^{(\ell)}):\aleph\to\mathbb{R}_{>0} and E:>0E:\aleph\to\mathbb{R}_{>0} from (44) and (39) satisfy properties P1, P2, P3, and P4. Hence Theorem 5.7 guarantees that Algorithm 3 terminates in finite time and computes 𝜹(+1)n+1{\bm{\delta}}^{(\ell+1)}\in\aleph_{n_{\ell+1}} satisfying E(𝜹(+1))ε+1E({\bm{\delta}}^{(\ell+1)})\leq\varepsilon_{\ell+1}.

Now the statement of the theorem follows by induction. ∎

7 Numerical examples

\bm{\ell} 7 8 9 10 11
ε\varepsilon_{\ell} 3.2E1 1.6E1 8.0E0 4.0E0 2.0E0
Relative error 5.9E-1 2.9E-1 1.5E-1 7.3E-2 3.7E-2
Algorithm 1
line 1
1.90E-1 1.99E0 2.41E1 3.70E2 9.53E3
Algorithm 1
line 1
1.28E-1 2.00E0 1.34E1 1.99E2 4.71E3
Algorithm 3
2.96E-3 1.01E-2 5.56E-2 2.91E-1 1.19E0
Table 1: Runtime in seconds for different components of Algorithm 4 when applied to system (70) with parameters L=4L=4 and εmax=2\varepsilon_{\ell_{\max}}=2.
Refer to caption
Figure 4: Approximate reachable sets of system (70) with L=4L=4 generated by Algorithm 4 with 3.7%3.7\% relative error. Resolution adjusted for printing purposes.
Refer to caption
Figure 5: Discretization generated by Algorithm 4 to approximate the reachable set of system 70 with L=4L=4, and with 3.7%3.7\% relative error.
Refer to caption
(a) L=1L=1
Refer to caption
(b) L=2L=2
Refer to caption
(c) L=3L=3
Refer to caption
(d) L=4L=4
Figure 6: Algorithms BA (diamonds), BU (triangles), EA (squares) and EU (circles) applied to system (70) with varying LL. Black markers represent measurements. Grey markers represent approximate inferred data.
Refer to caption
Figure 7: Approximate reachable sets of system (71) generated by Algorithm 4 with 8%8\% relative error. Resolution adjusted for printing purposes.
Refer to caption
Figure 8: Discretization produced by Algorithm 4 to approximate the reachable set of system 71 with 8%8\% relative error.
Refer to caption
Figure 9: Algorithms BA (diamonds), BU (triangles), EA (squares) and EU (circles) applied to system (71).

In this section, we compare Algorithm 4 with three other methods for reachable set computation. We introduce the following abbreviations, where B stands for the boundary method, E stands for the plain Euler scheme, A stands for adaptive, and U stands for uniform:

  • BA:

    Algorithm 4 from above,

  • BU:

    the boundary tracking algorithm with uniform discretization from [15],

  • EA:

    the Euler scheme with adaptive discretization from [17], and

  • EU:

    the Euler scheme with uniform discretization from [4] and [9].

We measure the quality of a numerical solution in terms of the a priori error bounds given in the literature for each scheme (bound (39) for Algorithm 4) and the computational cost in terms of runtime.

We first consider a very simple example in which all relevant quantities and the reachable sets are explicitly known.

Example 7.1.

Let T=1T=1 and L>0L>0, and consider the differential inclusion in 2\mathbb{R}^{2} given by

x˙i[0.9,1.0]Lxi,i=1,2,x(0)=(1,1)T.\dot{x}_{i}\in[0.9,1.0]Lx_{i},\,i=1,2,\quad x(0)=(1,1)^{T}. (70)

The exact reachable sets are

F(t)=[exp(0.9Lt),exp(Lt)]2,t[0,1],\mathcal{R}^{F}(t)=[\exp(0.9Lt),\exp(Lt)]^{2},\quad t\in[0,1],

and we have dF=dR=1d_{F}=d_{R}=1 and P=LeLP=Le^{L}.

We first consider Algorithm (4) when applied to system (70) with L=4L=4 and the error tolerance ε=2.0\varepsilon=2.0, which converts to a 3.7%3.7\% relative error. Table 1 shows how the runtime of Algorithm 4 is distributed.

  • i)

    Line 1 of Algorithm 1 is the part of the algorithm that is captured by the cost C^\hat{C} from (41) and hence by the cost estimator CC from (44).

  • ii)

    Since line 1 of Algorithm 1 performs a small fixed number of lookup operations for every unique point computed in line 1, its complexity should be a small multiple of the complexity of line 1. Our computations confirm this.

  • iii)

    The computational cost of Algorithm 3 is negligible compared to that of lines 1 and 1 of Algorithm 1.

All in all, the computations confirm that the cost estimator CC from (44) is a sensible measure of the computational complexity of Algorithm 4.

Figure 4 shows the resulting approximate reachable set boundaries generated by Algorithm 4. Sets shown are taken at times t{j16:0j16}.t\in\{\tfrac{j}{16}:0\leq j\leq 16\}.

Figure 5 shows that the adaptive refinement strategy of Algorithm 4 works as intended. Initially, the reachable set is small, and a very fine temporal and spatial discretization can be selected at a low cost. Since the local errors are propagated exponentially, choosing a small mesh size near t=0t=0 is very desirable. The opposite arguments hold for larger tt.

To test the performance of Algorithm 4 (BA) against previous algorithms BU, EA, and EU, we now consider system (70) with L=1,2,3,4L=1,2,3,4. Figure 6 shows that in terms of runtime, Algorithm BA from this paper performs best. Interestingly, the rate at which the runtime increases is very similar for both boundary methods, BA and BU, as well as for both Euler methods EA and EU.

The following example is inspired by [14, Example 4.1].

Example 7.2.

Let T=1T=1, and consider the problem

x˙1[0,1]πx2,x˙2[0,1]πx1,x(0)=(1,0)2\dot{x}_{1}\in[0,1]\pi x_{2},\quad\dot{x}_{2}\in-[0,1]\pi x_{1},\quad x(0)=(-1,0)\in\mathbb{R}^{2} (71)

with Lipschitz constant L=πL=\pi and bound P=π2.P=\pi^{2}.

The approximate reachable set boundaries generated by Algorithm 4 are shown in Figure 7 at times t{j16:0j16}t\in\{\frac{j}{16}:0\leq j\leq 16\}.

Figures 8 and 9 are the equivalent of Figures 5 and 6 for system (71), and show that Algorithm 4 still exhibits the desired behavior when applied to a less straightforward system.

8 Conclusion

Runge-Kutta methods for computing reachable sets have seen significant advancements over the past decades, as illustrated in Figures 6 and 9. The Euler scheme with uniform discretization from [4, 9] achieved meaningful error bounds only for control systems with very small control sets or over very short time intervals. The boundary tracking algorithm with uniform discretization from [15] reduced the computational cost dramatically, but computations with decent accuracy remained a challenge.

The adaptive Euler scheme from [17] outperforms [15] at low and moderate accuracy, but underperforms it asymptotically, because it tracks full-dimensional sets. Algorithm 4 presented in this paper integrates techniques from both [15] and [17] to overcome this issue. All in all, we see in Figures 6 and 9 that the computational cost for a prescribed precision has been brought down by many orders of magnitude.

While we have proved that Algorithm 4 converges, we have not been able to prove bounds on its performance that reflect its behavior in computational experiments. For now, this remains an open problem. In addition, it would be desirable to develop boundary tracking methods and adaptive refinement techniques based on semi-implicit and fully implicit schemes, which are better suited for stiff systems.

Acknowledgement

This work has partly been supported by an Australian Government Research Training Program (RTP) Scholarship.

References

  • [1] M. Althoff, O. Stursberg, and M. Buss. Computing reachable sets of hybrid systems using a combination of zonotopes and polytopes. Nonlinear Anal. Hybrid Syst., 4(2):233–249, 2010.
  • [2] R. Baier, C. Büskens, I.A. Chahma, and M. Gerdts. Approximation of reachable sets by direct solution methods for optimal control problems. Optim. Methods Softw., 22(3):433–452, 2007.
  • [3] R. Baier, M. Gerdts, and I. Xausa. Approximation of reachable sets using optimal control algorithms. Numer. Algebra Control Optim., 3(3):519–548, 2013.
  • [4] W.-J. Beyn and J. Rieger. Numerical fixed grid methods for differential inclusions. Computing, 81(1):91–106, 2007.
  • [5] R.M. Colombo, T. Lorenz, and N. Pogodaev. On the modeling of moving populations through set evolution equations. Discrete Contin. Dyn. Syst., 35(1):73–98, 2015.
  • [6] M. Gerdts and I. Xausa. Avoidance trajectories using reachable sets and parametric sensitivity analysis. In System modeling and optimization, volume 391 of IFIP Adv. Inf. Commun. Technol., pages 491–500. Springer, Heidelberg, 2013.
  • [7] A. Girard, C. Le Guernic, and O. Maler. Efficient computation of reachable sets of linear time-invariant systems with inputs. In Hybrid systems: computation and control, volume 3927 of Lecture Notes in Comput. Sci., pages 257–271. Springer, Berlin, 2006.
  • [8] R. Klette and A. Rosenfeld. Digital geometry. Morgan Kaufmann Publishers, San Francisco, CA; Elsevier Science B.V., Amsterdam, 2004. Geometric methods for digital picture analysis.
  • [9] V.A. Komarov and K.È. Pevchikh. A method for the approximation of attainability sets of differential inclusions with given accuracy. Zh. Vychisl. Mat. i Mat. Fiz., 31(1):153–157, 1991.
  • [10] A.B. Kurzhanski and P Varaiya. Ellipsoidal techniques for reachability analysis. In Hybrid Systems: Computation and Control, pages 202–214, Berlin, Heidelberg, 2000. Springer Berlin Heidelberg.
  • [11] C. Le Guernic and A. Girard. Reachability analysis of linear systems using support functions. Nonlinear Analysis: Hybrid Systems, 4(2):250–262, 2010. IFAC World Congress 2008.
  • [12] J.R. Munkres. Topology. Prentice Hall, Inc., Upper Saddle River, NJ, second edition, 2000.
  • [13] F. Parise, M.E. Valcher, and J. Lygeros. Computing the projected reachable set of stochastic biochemical reaction networks modeled by switched affine systems. IEEE Trans. Automat. Control, 63(11):3719–3734, 2018.
  • [14] W. Riedl, R. Baier, and M. Gerdts. Optimization-based subdivision algorithm for reachable sets. J. Comput. Dyn., 8(1):99–130, 2021.
  • [15] J. Rieger. Robust boundary tracking for reachable sets of nonlinear differential inclusions. Found. Comput. Math., 15(5):1129–1150, 2015.
  • [16] J. Rieger and K. Wawryk. Restriction and interpolation operators for digital images and their boundaries. arXiv, 2024, arXiv:2407.18511.
  • [17] J. Rieger and K. Wawryk. Towards optimal space-time discretization for reachable sets of nonlinear control systems. J. Comput. Dyn., 11(2):153–173, 2024.
  • [18] M. Rungger and M. Zamani. Accurate reachability analysis of uncertain nonlinear systems. In Proceedings of the 21st international conference on hybrid systems: Computation and control (part of CPS week), pages 61–70, 2018.
  • [19] M. Serry and G. Reissig. Overapproximating reachable tubes of linear time-varying systems. IEEE Trans. Automat. Control, 67(1):443–450, 2022.
  • [20] V. Veliov. Second order discrete approximations to strongly convex differential inclusions. Systems Control Lett., 13(3):263–269, 1989.
  • [21] V. Veliov. On the time-discretization of control systems. SIAM J. Control Optim., 35(5), 1997.