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

Collision Avoidance for Elliptical Agents with Control Barrier Function Utilizing Supporting Lines

Koju Nishimoto1, Riku Funada1, Tatsuya Ibuki2, and Mitsuji Sampei1 *This work was supported by JSPS KAKENHI Grant Number 21K20425. To be presented at the 2022 American Control Conference.1 Department of Systems and Control Engineering, Tokyo Institute of Technology nishimoto.k.ab@m.titech.ac.jp.{funada},{sampei}@sc.e.titech.ac.jp.2 Department of Electronics and Bioinformatics, Meiji University ibuki@meiji.ac.jp.
Abstract

This paper presents a collision avoidance method for elliptical agents traveling in a two-dimensional space. We first formulate a separation condition for two elliptical agents utilizing a signed distance from a supporting line of an agent to the other agent, which renders a positive value if two ellipses are separated by the line. Because this signed distance could yield a shorter length than the actual distance between two ellipses, the supporting line is rotated so that the signed distance from the line to the other ellipse is maximized. We prove that this maximization problem renders the signed distance equivalent to the actual distance between two ellipses, hence not causing the conservative evasive motion. Then, we propose the collision avoidance method utilizing novel control barrier functions incorporating a gradient-based update law of a supporting line. The validity of the proposed methods is evaluated in the simulations.

I INTRODUCTION

Guaranteeing collision avoidance in multi-agent systems is of significant importance to ensuring safety in many application fields, including environmental monitoring [1, 2], autonomous transportation [3, 4], robot navigation [5], and precision agriculture [6]. In these challenging and complex realms, multi-agent systems are demanded to embrace agents of different capabilities, shapes, and sizes to enhance performance [7]. In the presence of such heterogeneity, the collision avoidance protocol should guide agents not to collide with each other while incorporating their forms.

To achieve real-time collision avoidance for multi-agent systems, various approaches have been developed, including the methods utilizing artificial potential fields (APFs) [8] and control barrier functions (CBFs) [9]. APFs were first presented in [10], and have been employed to the multi-agent systems in the context of the formation [11] and flocking control [12, 13], where a repulsive potential function is designed for steering agents not to collide with each other. A local repulsive function activated only in the sensing regions of each agent is proposed in [14]. On the other hand, CBFs were recently proposed in [15], which confines the state of the system in the set defined by the CBFs. The work [16] developed the collision avoidance methods for multi robot systems, which can be implemented in a distributed fashion. The authors in [17] developed the hybrid CBFs to achieve the collision avoidance of the agents with limited sensing ranges. The comparative study of APFs and CBFs in obstacle avoidance scenarios is conducted in [18]. Most of the mentioned papers assume the agent as a single point, a circular disk, or a sphere. Although these methods can be employed to any agents by overestimating the original shape of agents to a sphere enclosing them, this approach could render too conservative evasive behavior if they have a nonspherical, especially elongated, body.

To mitigate this conservativeness, we propose the collision avoidance method capable of embracing the agents with heterogeneous elliptical shapes, as shown in Fig. 1. Because the CBF provides the ability to synthesize the number of safety-critical constraints [19], e.g., the constraints ensuring the battery of the robot never depletes [20], together with the collision avoidance, we opt for a CBF-based approach. While, in the context of APFs, the work [21] proposes the flocking control for ellipsoidal agents with achieving collision avoidance, the condition used for repulsive potential functions becomes complicated and is not straightforwardly extendable to CBFs. The work [22] employs the result in the computer graphics field to develop the separation condition of elliptical agents. However, the physical interpretation of the metric utilized in the collision avoidance law is not readily understandable since it does not provide the distance between agents. The extent-compatible CBF is developed in [23], where it can prevent the collision among agents having volume by solving a sum-of-squares optimization (SOS) program. Although this method can be applied to elliptical agents, the computational burden stemming from SOS programs might prevent implementations to high order systems. The work also proposes the sampling-based methods while the designer needs to assume the bounded control input. The complexities and limitations in the existing collision avoidance methods for elliptical agents partially come from the difficulties in measuring the distance between two separated ellipses. Although the computational methods deriving numerical solutions of the distance between two ellipses are proposed in the computer graphics field [24, 25], the analytical solution of the distance is difficult to derive in a simple form.

To alleviate the difficulties of evaluating a distance between two ellipses, we propose a novel CBF that incorporates a signed distance from a supporting line of an elliptical agent to the other agent, as illustrated in Fig. 2. Because a naive selection of the supporting line could yield a shorter length than the actual distance between two ellipses, we propose a gradient ascent-based update law, where the supporting line is rotated on the boundary so that the distance between the line and the other ellipse is maximized. We then prove that the maximum value derived from this optimization problem is equivalent to the actual distance between two ellipses. A novel CBF incorporating the gradient ascent input to rotate the supporting line is designed. In addition, we prove that the proposed CBF is a valid one, namely, there always exists the control input to make the collision-free set forward invariance. Numerical simulations demonstrate that the proposed method achieves the collision avoidance between elliptical agents without exhibiting conservative evasive motions.

Refer to caption
Figure 1: Proposed scenario. The agents characterized as ellipses with heterogeneous shapes avoid collisions with each other.

II Preliminary

II-A Problem Formulation

In this paper, we present a collision avoidance method among elliptical agents, labeled through the index set 𝒩={1n}\mathcal{N}=\{1\cdots n\}, in 2-D Euclidean space 2{\mathbb{R}}^{2}, as illustrated in Fig. 1. We denote the world coordinate frame as Σw\Sigma_{w}. We also define the coordinate frame of agent ii as Σi\Sigma_{i}, arranged at the center of agent ii so that its xix_{i}-axis corresponds with the major axis of the ellipse. The relative pose of Σi\Sigma_{i} with respect to Σw\Sigma_{w} is described as (𝒑i,Ri(θ)):2×SO(2)(\bm{p}_{i},R_{i}(\theta)):{\mathbb{R}}^{2}\times SO(2) with the position 𝒑i=[pix,piy]T2\bm{p}_{i}=[p_{ix},p_{iy}]^{T}\in{\mathbb{R}}^{2} and the orientation

Ri(θi)=[cos(θi)sin(θi)sin(θi)cos(θi)],θi(π,π].\displaystyle R_{i}(\theta_{i})=\begin{bmatrix}\cos{(\theta_{i})}&-\sin{(\theta_{i})}\\ \sin{(\theta_{i})}&\cos{(\theta_{i})}\end{bmatrix},~{}\theta_{i}\in(-\pi,\pi]. (1)

The state of agent ii is defined as 𝒙i=[pix,piy,θi]T\bm{x}_{i}=[p_{ix},p_{iy},\theta_{i}]^{T}. We suppose that the motion of agent ii can be represented according to a single integrator dynamics,

𝒙˙i=[uix,uiy,uiθ]T\displaystyle\dot{\bm{x}}_{i}=\left[u_{ix},u_{iy},u_{i\theta}\right]^{T} (2)

with the velocity input [uix,uiy]T[u_{ix},u_{iy}]^{T} and the angular velocity input uiθu_{i\theta}. We denote the control input for agent ii as 𝒖i=[uix,uiy,uiθ]T\bm{u}_{i}=[u_{ix},u_{iy},u_{i\theta}]^{T}. Agent ii occupies the elliptical region i\mathcal{E}_{i} described as

i={𝑿2(𝑿𝒑i)TRiQi2RiT(𝑿𝒑i)10},\displaystyle\mathcal{E}_{i}\!=\!\left\{\bm{X}\!\in\!\mathbb{R}^{2}\!\mid\!(\bm{X}\!-\!\bm{p}_{i})^{T}R_{i}Q_{i}^{-2}R_{i}^{T}(\bm{X}\!-\!\bm{p}_{i})\!-\!1\!\leq\!0\right\}, (3)

where the constant matrix Qi=diag(qix,qiy)Q_{i}={\rm diag}(q_{ix},q_{iy}) is defined with qixq_{ix} and qiyq_{iy} which specify the length of the major axis and the minor axis.

We propose a control method that prevents a collision between elliptical agents described with (3). If the minimum distance between i\mathcal{E}_{i} and j\mathcal{E}_{j} is described as wij(𝒙i,𝒙j)w_{ij}^{*}(\bm{x}_{i},\bm{x}_{j}), the safe set restricting collisions between agents ii and jj can be captured by the set

𝒮ij={𝒙i,𝒙j3wij(𝒙i,𝒙j)0},\displaystyle\mathcal{S}_{ij}=\left\{\bm{x}_{i},\bm{x}_{j}\in\mathbb{R}^{3}\mid w_{ij}^{*}(\bm{x}_{i},\bm{x}_{j})\geq 0\right\}, (4)

where this set has to be rendered forward invariant. For this goal, we leverage control barrier functions (CBFs) being introduced in the next subsection.

II-B Control Barrier Functions

Refer to caption
(a)
Refer to caption
(b)
Figure 2: The supporting line lijl_{ij} separating two elliptical agents i\mathcal{E}_{i} and j\mathcal{E}_{j}. (a) shows the distance hij(ϕij)h_{ij}(\phi_{ij}) between the ellipse j\mathcal{E}_{j} and a supporting line lijl_{ij}, the normal vector and the tangent point of which is denoted as 𝒛ij\bm{z}_{ij} and 𝒎ij\bm{m}_{ij}, respectively. Although the distance hij(ϕij)h_{ij}(\phi_{ij}) can be derived from (13), hij(ϕij)h_{ij}(\phi_{ij}) could be shorter than the actual distance between two ellipses. (b) illustrates the update law of the supporting line lijl_{ij}, where lijl_{ij} is rotated on the boundary of i\mathcal{E}_{i} so that hij(ϕij)h_{ij}(\phi_{ij}) approaches the actual distance between two ellipses, by maximizing hij(ϕij)h_{ij}(\phi_{ij}) with ϕij\phi_{ij}. Note that hij(ϕij)h_{ij}(\phi_{ij}) takes a positive value if and only if the line lijl_{ij} separates two ellipses.

CBFs have been utilized for ensuring the forward invariance property to the set 𝒮\mathcal{S}, in which the state 𝒙\bm{x} should be confined during the task execution of agents. We assume that a set 𝒮\mathcal{S} can be expressed as the superlevel set of a continuously differentiable function h:nh:{\mathbb{R}}^{n}\to{\mathbb{R}}, namely, 𝒮={𝒙nh(𝒙)0}{\mathcal{S}}=\{\bm{x}\in\mathbb{R}^{n}\mid h(\bm{x})\geq 0\}. Then, CBF is defined as follows.

Definition 1.

[15, Def. 5] Given the control affine system

𝒙˙=f(𝒙)+g(𝒙)𝒖,\displaystyle\dot{\bm{x}}=f(\bm{x})+g(\bm{x})\bm{u}, (5)

where ff and gg are locally Lipschitz, 𝐱n\bm{x}\in{\mathbb{R}}^{n} and 𝐮m\bm{u}\in{\mathbb{R}}^{m}, together with the set 𝒮\mathcal{S}. Then, the function hh is a control barrier function (CBF) defined on a set 𝒮¯\bar{\mathcal{S}} with 𝒮𝒮¯n\mathcal{S}\subseteq\bar{\mathcal{S}}\subset\mathbb{R}^{n}, if there exists an extended class 𝒦\mathcal{K} function α\alpha, such that

sup𝒖[Lfh(𝒙)+Lgh(𝒙)𝒖+α(h(𝒙))]0,𝒙𝒮¯\displaystyle\sup_{\bm{u}}\left[L_{f}h(\bm{x})+L_{g}h(\bm{x})\bm{u}+\alpha(h(\bm{x}))\right]\geq 0,~{}~{}~{}\forall\bm{x}\in\bar{\mathcal{S}} (6)

where Lfh(𝐱)L_{f}h(\bm{x}) and Lgh(𝐱)L_{g}h(\bm{x}) are the Lie derivatives of hh along f(𝐱)f(\bm{x}) and g(𝐱)g(\bm{x}), respectively.

The forward invariance of the set 𝒮\mathcal{S} is ensured through the following corollary.

Corollary 1.

[15, Cor. 2] Given a set 𝒮\mathcal{S}, if hh is a CBF on 𝒮¯\bar{\mathcal{S}}, then any Lipschitz continuous controller u(𝐱):𝒮¯Uu(\bm{x}):\bar{\mathcal{S}}\to U such that

Lfh(𝒙)+Lgh(𝒙)𝒖(𝒙)+α(h(𝒙))0,\displaystyle L_{f}h(\bm{x})+L_{g}h(\bm{x})\bm{u}(\bm{x})+\alpha(h(\bm{x}))\geq 0, (7)

will render the set 𝒮\mathcal{S} forward invariant.

The condition (7) guaranteeing forward invariance of the set 𝒮\mathcal{S} can be synthesized to the control law through the optimization-based controller leveraging Quadratic Programming (QP). Let us denote the nominal input as 𝒖nom\bm{u}_{\mathrm{nom}} and wish to modify it minimally invasive way so as to satisfy the condition (7). This goal can be achieved by employing the input 𝒖\bm{u}^{*} derived from the following QP

𝒖=\displaystyle\bm{u}^{*}= argmin𝒖𝒖𝒖nom(x)2,\displaystyle\mathop{\rm arg~{}min}\limits_{\bm{u}}~{}\|\bm{u}-\bm{u}_{\mathrm{nom}}(x)\|^{2}, (8a)
s.t.Lfh(𝒙)+Lgh(𝒙)𝒖+α(h(𝒙))0.\displaystyle~{}\mbox{s.t.}~{}L_{f}h(\bm{x})+L_{g}h(\bm{x})\bm{u}+\alpha(h(\bm{x}))\geq 0. (8b)

III Collision Avoidance for Elliptical Agents

In this section, we formulate a novel CBF that ensures the forward invariance of the set 𝒮ij\mathcal{S}_{ij} in (4), namely preventing agent ii from colliding with agent jj. As mentioned previously and from [21, 22], it is difficult to derive the analytical solution of the distance between two ellipses, namely wijw_{ij}^{*}, in a form simple enough to be employed as a CBF. Furthermore, numerical solutions of wijw_{ij}^{*} cannot be employed as a CBF. To mitigate the difficulties, we design a novel CBF that incorporates a signed distance from a supporting line of agent ii to agent jj, depicted as hijh_{ij} in Fig. 2. Because hijh_{ij} could take a shorter length than wijw_{ij}^{*} with naive choices of the supporting line, we propose the procedure that drives hijh_{ij} to wijw_{ij}^{*} based on the gradient of an optimization problem.

III-A Separation Conditions for Two Elliptical Agents

Refer to caption
Figure 3: The parameter ϕij\phi_{ij}, which specifies a point on the boundary of the ellipse. The point on the unit circle is specified by 𝒗ij\bm{v}_{ij} in (10), which angle from the xx-axis is ϕij\phi_{ij}. Then, a point on the ellipse is specified by transforming 𝒗ij\bm{v}_{ij} with the positive definite matrix Q¯i\bar{Q}_{i}.

We first introduce a supporting line lijl_{ij} of agent ii, which contacts i\mathcal{E}_{i} at the point 𝒎ij\bm{m}_{ij}, as depicted in Fig. 2. The point 𝒎ij\bm{m}_{ij} can be defined as

𝒎ij(𝒙i,ϕij)\displaystyle\bm{m}_{ij}(\bm{x}_{i},\phi_{ij}) =Q¯i𝒗ij+𝒑i\displaystyle=\bar{Q}_{i}\bm{v}_{ij}+\bm{p}_{i} (9)
𝒗ij(ϕij)\displaystyle\bm{v}_{ij}(\phi_{ij}) =[cos(ϕij),sin(ϕij)]T\displaystyle=\left[\cos(\phi_{ij}),\sin(\phi_{ij})\right]^{T} (10)

with a positive definite matrix Q¯i=RiQiRiT\bar{Q}_{i}=R_{i}Q_{i}R_{i}^{T} and a parameter ϕij(π,π]\phi_{ij}\in(-\pi,\pi]. Here, the parameter ϕij\phi_{ij} is introduced to specify a point on the boundary of the ellipse i\mathcal{E}_{i}, where the graphical interpretation is shown in Fig. 3. Then, the supporting line lijl_{ij} is described as

lij={X2|𝒗ijTQ¯i1X(1+𝒗ijTQ¯i1𝒑i)=0},\displaystyle l_{ij}\!=\!\left\{X\in\mathbb{R}^{2}~{}|~{}\bm{v}_{ij}^{T}\bar{Q}_{i}^{-1}X-\left(1+\bm{v}_{ij}^{T}\bar{Q}_{i}^{-1}\bm{p}_{i}\right)\!=\!0\right\}, (11)

which is determined by 𝒙i\bm{x}_{i} and ϕij\phi_{ij}.

Refer to caption
(a) hij(ϕij)<0h_{ij}(\phi_{ij})<0
Refer to caption
(b) hij(ϕij)<0h_{ij}(\phi_{ij})<0
Refer to caption
(c) hij(ϕij)>0h_{ij}(\phi_{ij})>0
Figure 4: The separation condition evaluated by hijh_{ij}, the minimum signed distance from the line lijl_{ij}. The signed distance hijh_{ij} takes zero on lijl_{ij} and takes the larger value as a point to be evaluated moves to the upper direction specified by the green normal vector. (a) and (b): Since the proposed signed distance provides a negative value to a point in the same half-plane with the ellipse i\mathcal{E}_{i}, 𝒏ijj\bm{n}_{ij}\in\mathcal{E}_{j} equipped with the minimum signed distance hijh_{ij} is the one furthest from lijl_{ij}. (c): When the supporting line separates two ellipses, hijh_{ij} returns the distance between lijl_{ij} and j\mathcal{E}_{j}.

Let us derive the separation condition evaluated with the signed distance from the supporting line lijl_{ij}, which provides a positive value to a point in the different half-plane with i\mathcal{E}_{i}, and a negative value otherwise as shown in Fig. 4. The point 𝒏ijj\bm{n}_{ij}\in\mathcal{E}_{j} that minimizes the signed distance from the supporting line lijl_{ij} is determined by

𝒏ij(𝒙i,𝒙j,ϕij)=1Q¯jQ¯i1𝒗ijQ¯j2Q¯i1𝒗ij+𝒑j.\displaystyle\bm{n}_{ij}(\bm{x}_{i},\bm{x}_{j},\phi_{ij})=-\frac{1}{\left\|\bar{Q}_{j}\bar{Q}_{i}^{-1}\bm{v}_{ij}\right\|}\bar{Q}_{j}^{2}\bar{Q}_{i}^{-1}\bm{v}_{ij}+\bm{p}_{j}. (12)

Then, the minimum signed distance from lijl_{ij} is calculated by

hij(𝒙i,𝒙j,ϕij)=Q¯jQ¯i1𝒗ij+(𝒑j𝒑i)TQ¯i1𝒗ij1Q¯i1𝒗ij,\displaystyle h_{ij}(\bm{x}_{i},\bm{x}_{j},\phi_{ij})\!=\!\frac{-\left\|\bar{Q}_{j}\bar{Q}_{i}^{-1}\bm{v}_{ij}\right\|\!+\!(\bm{p}_{j}-\bm{p}_{i})^{T}\bar{Q}_{i}^{-1}\bm{v}_{ij}\!-\!1}{\left\|\bar{Q}_{i}^{-1}\bm{v}_{ij}\right\|}, (13)

which satisfies hij(𝒙i,𝒙j,ϕij)>0h_{ij}(\bm{x}_{i},\bm{x}_{j},\phi_{ij})>0 if and only if i\mathcal{E}_{i} and 𝒏ij\bm{n}_{ij} are in the different half-plane separated by the line lijl_{ij} as shown in Fig. 4. In other words, if there exists ϕij\phi_{ij} that fulfills hij(𝒙i,𝒙j,ϕij)>0h_{ij}(\bm{x}_{i},\bm{x}_{j},\phi_{ij})>0, then two ellipses are separated by the line lijl_{ij}. Note that 𝒏ijj\bm{n}_{ij}\in\mathcal{E}_{j} is not the closest point to the supporting line lijl_{ij} as depicted in Fig. 4(a) and (b). In the remaining of this subsection, we analyze the property of hij(𝒙i,𝒙j,ϕij)h_{ij}(\bm{x}_{i},\bm{x}_{j},\phi_{ij}) assuming two agents ii and jj are fixed, hence denote hij(𝒙i,𝒙j,ϕij)h_{ij}(\bm{x}_{i},\bm{x}_{j},\phi_{ij}) as hij(ϕij)h_{ij}(\phi_{ij}).

The fact that hij(ϕij)>0h_{ij}(\phi_{ij})>0 encodes the separation condition between two agents motivates us to employ hij(ϕij)h_{ij}(\phi_{ij}) as a CBF for ensuring collision avoidance between elliptical agents. However, the naive selection of the parameter ϕij\phi_{ij}, namely the line lijl_{ij}, could yield a shorter distance than the actual distance wijw_{ij}^{*} as depicted in Fig. 2(a). Because of this difference from wijw_{ij}^{*}, hij(ϕij)h_{ij}(\phi_{ij}) could overestimate the risk of collisions and hence cause too conservative evasive motion. To mitigate this gap, we propose the following optimization problem that intends to rotate the supporting line lijl_{ij} on the boundary of i\mathcal{E}_{i} so that hij(ϕij)h_{ij}(\phi_{ij}) is maximized as

maxϕij\displaystyle\max_{\phi_{ij}}~{} hij(ϕij),\displaystyle h_{ij}(\phi_{ij}), (14a)
s.t. 𝒗ij=[cos(ϕij),sin(ϕij)]T,\displaystyle\bm{v}_{ij}=\left[\cos(\phi_{ij}),\sin(\phi_{ij})\right]^{T}, (14b)

where its graphical interpretation is depicted in Fig. 2(b).

While, in the next subsection, we introduce the input to ϕij\phi_{ij} for maximizing hij(ϕij)h_{ij}(\phi_{ij}), we first establish the connection between the optimization problem (14) and the actual distance between two agents, namely wijw_{ij}^{*}. For this goal, we introduce the following optimization problem, which optimal solution 𝒘\|\bm{w}^{*}\| is equal to wijw_{ij}^{*}.

min𝝃,𝜼,𝒘\displaystyle\min_{\bm{\xi},\bm{\eta},\bm{w}}~{} 𝒘,\displaystyle\|\bm{w}\|, (15a)
s.t. fi(𝝃)0,fj(𝜼)0,\displaystyle f_{i}(\bm{\xi})\leq 0,~{}f_{j}(\bm{\eta})\leq 0, (15b)
𝜼𝝃=𝒘,\displaystyle\bm{\eta}-\bm{\xi}=\bm{w}, (15c)

where fk(𝝃):=(𝝃𝒑k)TQ¯k2(𝝃𝒑k)10f_{k}(\bm{\xi}):=(\bm{\xi}-\bm{p}_{k})^{T}\bar{Q}_{k}^{-2}(\bm{\xi}-\bm{p}_{k})-1\leq 0 signifies the condition 𝝃k\bm{\xi}\in\mathcal{E}_{k}. Then, the following theorem formalizes the relationship between wijw_{ij}^{*} and hij(ϕij)h_{ij}(\phi_{ij}).

Theorem 1.

Suppose that two ellipses i\mathcal{E}_{i} and j\mathcal{E}_{j} have no overlap, namely ij=\mathcal{E}_{i}\cap\mathcal{E}_{j}=\emptyset holds. Then, the optimization problem (14) is the dual of the problem (15). Furthermore, the strong duality holds between the optimization problems (15) and (14), namely the following condition holds

wij=hijhij(ϕij).\displaystyle w_{ij}^{*}=h_{ij}^{*}\geq h_{ij}(\phi_{ij}). (16)
Proof.

The dual function of the problem (15) is

g(λi,λj,𝒛)=inf𝝃,𝜼,𝒘(𝒘+λifi(𝝃)+λjfj(𝜼)+𝒛T(𝜼𝝃𝒘))\displaystyle\begin{aligned} g(\lambda_{i},\lambda_{j},\bm{z})=\inf_{\bm{\xi},\bm{\eta},\bm{w}}\left(\|\bm{w}\|\right.&\left.+\lambda_{i}f_{i}(\bm{\xi})+\lambda_{j}f_{j}(\bm{\eta})\right.\\ &\left.+\bm{z}^{T}(\bm{\eta}-\bm{\xi}-\bm{w})\right)\end{aligned} (17)
={inf𝝃(λifi(𝝃)𝒛T𝝃)+inf𝜼(λjfj(𝜼)+𝒛T𝜼)𝒛1,λi,λj0otherwise,\displaystyle=\begin{cases}\begin{aligned} \inf_{\bm{\xi}}&\left(\lambda_{i}f_{i}(\bm{\xi})-\bm{z}^{T}\bm{\xi}\right)\\ &+\inf_{\bm{\eta}}\left(\lambda_{j}f_{j}(\bm{\eta})+\bm{z}^{T}\bm{\eta}\right)\end{aligned}&\begin{aligned} \|\bm{z}\|&\leq 1,\\ \lambda_{i},\lambda_{j}&\geq 0\end{aligned}\\ -\infty&\mbox{otherwise}\end{cases}, (18)

where λi\lambda_{i}, λj\lambda_{j} and 𝒛\bm{z} are Lagrange multipliers. To simplify the first term inf𝝃(λifi(𝝃)𝒛T𝝃)\inf_{\bm{\xi}}\left(\lambda_{i}f_{i}(\bm{\xi})-\bm{z}^{T}\bm{\xi}\right) in (18), we introduce 𝝃¯=Q¯i1𝝃\bar{\bm{\xi}}=\bar{Q}_{i}^{-1}\bm{\xi}, 𝒑¯i=Q¯i1𝒑i\bar{\bm{p}}_{i}=\bar{Q}_{i}^{-1}\bm{p}_{i}, and 𝒛¯=Q¯i𝒛\bar{\bm{z}}=\bar{Q}_{i}\bm{z}. Then, the first term is transformed as

inf𝝃(λifi(𝝃)𝒛T𝝃)\displaystyle\inf_{\bm{\xi}}\left(\lambda_{i}f_{i}({\bm{\xi}})-\bm{z}^{T}{\bm{\xi}}\right)
=\displaystyle= inf𝝃(λi(𝝃𝒑i)TQ¯i2(𝝃𝒑i)λi𝒛T𝝃)\displaystyle\inf_{\bm{\xi}}\left(\lambda_{i}({\bm{\xi}}-\bm{p}_{i})^{T}\bar{Q}_{i}^{-2}({\bm{\xi}}-\bm{p}_{i})-\lambda_{i}-\bm{z}^{T}{\bm{\xi}}\right)
=\displaystyle= 𝒛¯T𝒑¯i𝒛¯2+4λi24λi.\displaystyle-\bar{\bm{z}}^{T}\bar{\bm{p}}_{i}-\frac{\|\bar{\bm{z}}\|^{2}+4\lambda_{i}^{2}}{4\lambda_{i}}. (19)

Following the similar path to (19), the second term in (18) is expressed as

inf𝜼(λjfj(𝜼)+𝒛T𝜼)=𝒛^T𝒑^j𝒛^2+4λj24λj,\displaystyle\inf_{\bm{\eta}}\left(\lambda_{j}f_{j}(\bm{\eta})+\bm{z}^{T}\bm{\eta}\right)=\hat{\bm{z}}^{T}\hat{\bm{p}}_{j}-\frac{\|\hat{\bm{z}}\|^{2}+4\lambda_{j}^{2}}{4\lambda_{j}}, (20)

where 𝒑^j=Q¯j1𝒑j\hat{\bm{p}}_{j}=\bar{Q}_{j}^{-1}\bm{p}_{j} and 𝒛^=Q¯j𝒛\hat{\bm{z}}=\bar{Q}_{j}\bm{z}. Therefore, the dual problem can be expressed as follows:

maxz,λi,λj\displaystyle\max_{z,\lambda_{i},\lambda_{j}}~{} 𝒛¯T𝒑¯i𝒛¯2+4λi24λi+𝒛^T𝒑^j𝒛^2+4λj24λj,\displaystyle-\bar{\bm{z}}^{T}\bar{\bm{p}}_{i}-\frac{\|\bar{\bm{z}}\|^{2}+4\lambda_{i}^{2}}{4\lambda_{i}}+\hat{\bm{z}}^{T}\hat{\bm{p}}_{j}-\frac{\|\hat{\bm{z}}\|^{2}+4\lambda_{j}^{2}}{4\lambda_{j}}, (21a)
s.t. 𝒛¯=Qi𝒛,𝒛^=Qj𝒛,𝒛1,λi,λj0.\displaystyle\bar{\bm{z}}=Q_{i}\bm{z},~{}\hat{\bm{z}}=Q_{j}\bm{z},~{}\|\bm{z}\|\leq 1,~{}\lambda_{i},\lambda_{j}\geq 0. (21b)

Focusing on the second and fourth terms in (21a), we define a function M(a,x)M(a,x) as

M(a,x)=a+4x24x(a,x0),\displaystyle M(a,x)=-\frac{a+4x^{2}}{4x}~{}~{}~{}(a,x\geq 0), (22)

and consider maxa,xM(a,x)\max_{a,x}M(a,x).

In the case of a>0a>0, the gradient of MM is

Mx=(a2x)(a+2x)4x2,Ma=14x.\displaystyle\frac{\partial M}{\partial x}=\frac{(\sqrt{a}-2x)(\sqrt{a}+2x)}{4x^{2}},~{}\frac{\partial M}{\partial a}=-\frac{1}{4x}. (23)

Thus, the function M(a,x)M(a,x) has no extremum for all a>0a>0, and for x0x\geq 0 it has the maximum value at x(a)=a/2x^{*}(a)={\sqrt{a}}/{2}. As a result, the following equation holds.

maxa,xM(a,x)=maxaa(a>0).\displaystyle\max_{a,x}M(a,x)=\max_{a}-\sqrt{a}~{}~{}~{}(a>0). (24)

In the case of a=0a=0, the maximum value of M(a,x)M(a,x) is as follows:

maxa,xM(a,x)=maxxx=0.\displaystyle\max_{a,x}M(a,x)=\max_{x}-x=0. (25)

Since (24) is equivalent to (25) if we substitute a=0a=0 into (24), we can summarize them as

maxa,xM(a,x)=maxaa(a0).\displaystyle\max_{a,x}M(a,x)=\max_{a}-\sqrt{a}~{}~{}~{}(a\geq 0). (26)

Considering that the second and fourth terms in (21a) are equal to M(𝒛¯2,λi)M(\|\bar{\bm{z}}\|^{2},\lambda_{i}) and M(𝒛^2,λj)M(\|\hat{\bm{z}}\|^{2},\lambda_{j}), respectively, we can simplify the problem (21) by utilizing (26) as follows:

max𝒛\displaystyle\max_{\bm{z}}~{} 𝒛¯T𝒑¯i𝒛¯+𝒛^T𝒑^j𝒛^,\displaystyle-\bar{\bm{z}}^{T}\bar{\bm{p}}_{i}-\left\|\bar{\bm{z}}\right\|+\hat{\bm{z}}^{T}\hat{\bm{p}}_{j}-\left\|\hat{\bm{z}}\right\|, (27a)
s.t. 𝒛1,\displaystyle\|\bm{z}\|\leq 1, (27b)
𝒛¯=Q¯i𝒛,𝒛^=Q¯j𝒛.\displaystyle\bar{\bm{z}}=\bar{Q}_{i}\bm{z},~{}\hat{\bm{z}}=\bar{Q}_{j}\bm{z}. (27c)

Let us parameterize 𝒛\bm{z} as

𝒛\displaystyle\bm{z} =μQ¯i1𝒗ijQ¯i1𝒗ij,\displaystyle=\frac{\mu}{\left\|\bar{Q}_{i}^{-1}\bm{v}_{ij}\right\|}\bar{Q}_{i}^{-1}\bm{v}_{ij}, (28)

with 0μ10\leq\mu\leq 1 so that the constraint (27b) is satisfied. Then, by substituting (28), 𝒛¯=Q¯i𝒛\bar{\bm{z}}=\bar{Q}_{i}\bm{z} and 𝒛^=Q¯j𝒛\hat{\bm{z}}=\bar{Q}_{j}\bm{z}, the dual problem (27) can be transformed as follows

maxμ,ϕij\displaystyle\max_{\mu,\phi_{ij}}~{} μQ¯jQ¯i1𝒗ij+(𝒑j𝒑i)TQ¯i1𝒗ij1Q¯i1𝒗ijhij(ϕij),\displaystyle\mu\underbrace{\frac{-\left\|\bar{Q}_{j}\bar{Q}_{i}^{-1}\bm{v}_{ij}\right\|+(\bm{p}_{j}-\bm{p}_{i})^{T}\bar{Q}_{i}^{-1}\bm{v}_{ij}-1}{\left\|\bar{Q}_{i}^{-1}\bm{v}_{ij}\right\|}}_{h_{ij}(\phi_{ij})}, (29a)
s.t. 𝒗ij=[cos(ϕij),sin(ϕij)]T,\displaystyle\bm{v}_{ij}=[\cos(\phi_{ij}),\sin(\phi_{ij})]^{T}, (29b)
0μ1,\displaystyle 0\leq\mu\leq 1, (29c)

where the objective function (29a) can be described as μhij(ϕij)\mu h_{ij}(\phi_{ij}). Notice that if ellipses i\mathcal{E}_{i} and j\mathcal{E}_{j} are separated, there always exists ϕij\phi_{ij} that satisfies hij(ϕij)>0h_{ij}(\phi_{ij})>0. Considering the constraint (29c) and the fact μ\mu is a variable independent of ϕij\phi_{ij}, we should set μ=1\mu=1 to maximize (29a). Therefore, by substituting μ=1\mu=1 to the problem (29), we obtain the optimization problem (14). This relation leads to the conclusion that the problem (14) is the dual of the optimization problem (15) if ij=\mathcal{E}_{i}\cap\mathcal{E}_{j}=\emptyset. Since the optimization problem (15) satisfies the Slater’s Condition [26, Sec. 5.2.3], the solution of (15) is equivalent to the solution of (14), which completes the proof. ∎

Theorem 1 implies that the proposed update law of the supporting line, described as the optimization problem (14) and Fig. 2, renders hij(ϕij)h_{ij}(\phi_{ij}) the actual distance wijw_{ij}^{*} between two ellipses. Furthermore, because of (16), the condition hij(ϕij)>0h_{ij}(\phi_{ij})>0 provides a sufficient condition for preventing collisions even if ϕij\phi_{ij} does not converge to the maximizer of (14). In the following subsection, we develop hijh_{ij} as a CBF together with presenting the update procedure of ϕij\phi_{ij}.

III-B CBFs Incorporating Rotating Supporting Lines

In this subsection, we first design the collision avoidance methods between two elliptical agents ii and jj, then extend the results to the case of nn agents. Hereafter, we regard hijh_{ij} as a function of 𝒙i,𝒙j\bm{x}_{i},\bm{x}_{j}, and ϕij\phi_{ij} as first defined in (13), though we have dropped the dependency of hijh_{ij} for notational convenience.

As mentioned in the previous subsection, we require a supporting line between agents ii and jj to evaluate the separation condition hijh_{ij}. Without loss of generality, we introduce a supporting line for the agent with a lower ID number. Then, as the supporting line is to be updated for maximizing hijh_{ij}, the model of the agent (2) now has to describe the dynamics of ϕij\phi_{ij} as well. Therefore, we introduce the augmented state 𝒙ij=[𝒙iT,𝒙jT,ϕij]T\bm{x}_{ij}=\left[\bm{x}_{i}^{T},~{}\bm{x}_{j}^{T},~{}\phi_{ij}\right]^{T} to achieve collision avoidance between two agents ii and jj (i<ji<j), where the dynamics of the augmented state is

𝒙˙ij=𝒖ij,\displaystyle\dot{\bm{x}}_{ij}=\bm{u}_{ij}, (30)

where 𝒖ij=[𝒖iT,𝒖jT,uϕij]T\bm{u}_{ij}=\left[\bm{u}_{i}^{T},\bm{u}_{j}^{T},u_{\phi_{ij}}\right]^{T}.

As the nominal input for ϕij\phi_{ij} intending to maximize hijh_{ij}, we employ the following gradient-based input

unom,ϕij=γhijϕij,γ>0.\displaystyle u_{\mathrm{nom},\phi_{ij}}=\gamma\frac{\partial h_{ij}}{\partial\phi_{ij}},~{}~{}~{}\gamma>0. (31)

The input (31) drives hijh_{ij} to a local maximum point, and hence preventing too conservative evasive motions caused by the difference between wijw_{ij}^{*} and hijh_{ij}. Note that the maximizer ϕij\phi_{ij}^{*} of hijh_{ij} changes as the agent ii and jj traverse. However, as ϕij\phi_{ij} is a virtual variable not dependent on the physical dynamics of the agents, we can make ϕij\phi_{ij} converge fast enough with large γ\gamma to follow the agent movement, as we will demonstrate in the simulations in Section IV-A.

Let us denote the nominal control input for the agent ii as 𝒖nom,i\bm{u}_{{\rm nom},i}, which is designed to achieve its own objective, e.g., reaching the goal position. Combining this nominal control input 𝒖nom,i\bm{u}_{{\rm nom},i} with (31), the nominal input for the augmented state 𝒙ij\bm{x}_{ij} is expressed as 𝒖nom,ij=[𝒖nom,iT,𝒖nom,jT,unom,ϕij]T\bm{u}_{\mathrm{nom},ij}=\left[\bm{u}_{\mathrm{nom},i}^{T},~{}\bm{u}_{\mathrm{nom},j}^{T},~{}u_{\mathrm{nom},\phi_{ij}}\right]^{T}.

The goal of the collision avoidance strategy is to allow agents to execute their tasks encoded by 𝒖nom,i\bm{u}_{{\rm nom},i} while ensuring collision never occurs between agents. To achieve this objective, we propose the collision avoidance methods that render the following set 𝒮^ij\hat{\mathcal{S}}_{ij} forward invariant.

𝒮^ij={𝒙ij6×(π,π]hij(𝒙ij)0}\displaystyle\hat{\mathcal{S}}_{ij}=\left\{\bm{x}_{ij}\in\mathbb{R}^{6}\times(-\pi,\pi]\mid h_{ij}(\bm{x}_{ij})\geq 0\right\} (32)

Since wijhijw_{ij}^{*}\geq h_{ij} holds from Theorem 1, the set 𝒮ij\mathcal{S}_{ij} in (4) is also ensured to be forward invariant if we guarantee the forward invariance of 𝒮^ij\hat{\mathcal{S}}_{ij}. Similar to (8), this strategy can be achieved by 𝒖ij\bm{u}_{ij}^{*} derived from the following QP integrating hijh_{ij} as a CBF:

𝒖ij=argmin𝒖ij𝒖ij𝒖nom,ij2,\displaystyle\bm{u}_{ij}^{*}=\mathop{\rm arg~{}min}\limits_{\bm{u}_{ij}}~{}\left\|\bm{u}_{ij}-\bm{u}_{\mathrm{nom},ij}\right\|^{2}, (33a)
s.t.hij𝒙iT𝒖i+hij𝒙jT𝒖j+hijϕijuϕij+α(h)0.\displaystyle\mbox{s.t.}~{}{\frac{\partial h_{ij}}{\partial\bm{x}_{i}}}^{T}\bm{u}_{i}+{\frac{\partial h_{ij}}{\partial\bm{x}_{j}}}^{T}\bm{u}_{j}+\frac{\partial h_{ij}}{\partial\phi_{ij}}u_{\phi_{ij}}+\alpha(h)\geq 0. (33b)

Note that the constraint (33b) is derived by calculating the Lie derivative of hijh_{ij} along fij=𝟎7×1f_{ij}=\bm{0}_{7\times 1} and gij=I7g_{ij}=I_{7} in (30).

As the proposed CBF hijh_{ij} prevents collision between elliptical agents ii and jj, we need to confirm whether there always exists the control input that satisfies (6) in the Definition 1 and it provides us the forward invariance property.

Theorem 2.

The function hijh_{ij} in (13) is a valid CBF when 𝐮i,𝐮j3\bm{u}_{i},\bm{u}_{j}\in{\mathbb{R}}^{3}. Namely, for any 𝐱i,𝐱j3\bm{x}_{i},\bm{x}_{j}\in\mathbb{R}^{3}, the following conditions hold.

hij𝒙i\displaystyle\frac{\partial h_{ij}}{\partial\bm{x}_{i}} =[hij𝒑iT,hijθi]T𝟎3×1\displaystyle=\left[{\frac{\partial h_{ij}}{\partial\bm{p}_{i}}}^{T},\frac{\partial h_{ij}}{\partial\theta_{i}}\right]^{T}\neq\bm{0}_{3\times 1} (34)
hij𝒙j\displaystyle\frac{\partial h_{ij}}{\partial\bm{x}_{j}} =[hij𝒑jT,hijθj]T𝟎3×1\displaystyle=\left[{\frac{\partial h_{ij}}{\partial\bm{p}_{j}}}^{T},\frac{\partial h_{ij}}{\partial\theta_{j}}\right]^{T}\neq\bm{0}_{3\times 1} (35)
Proof.

hij/𝒑i{\partial h_{ij}}/{\partial\bm{p}_{i}} and hij/𝒑j{\partial h_{ij}}/{\partial\bm{p}_{j}} are given by

hij𝒑i\displaystyle\frac{\partial h_{ij}}{\partial\bm{p}_{i}} =1Q¯i1𝒗ijQ¯i1𝒗ij,\displaystyle=-\frac{1}{\left\|\bar{Q}_{i}^{-1}\bm{v}_{ij}\right\|}\bar{Q}_{i}^{-1}\bm{v}_{ij}, (36)
hij𝒑j\displaystyle\frac{\partial h_{ij}}{\partial\bm{p}_{j}} =1Q¯i1𝒗ijQ¯i1𝒗ij.\displaystyle=\frac{1}{\left\|\bar{Q}_{i}^{-1}\bm{v}_{ij}\right\|}\bar{Q}_{i}^{-1}\bm{v}_{ij}. (37)

Considering the fact 𝒗ij=1\left\|\bm{v}_{ij}\right\|=1 and Q¯i0\bar{Q}_{i}\succ 0,

hij𝒑i=hij𝒑j=1\displaystyle\left\|\frac{\partial h_{ij}}{\partial\bm{p}_{i}}\right\|=\left\|\frac{\partial h_{ij}}{\partial\bm{p}_{j}}\right\|=1

holds. Therefore, the conditions hij/𝒙i𝟎3×1{\partial h_{ij}}/{\partial\bm{x}_{i}}\neq\bm{0}_{3\times 1} and hij/𝒙j𝟎3×1{\partial h_{ij}}/{\partial\bm{x}_{j}}\neq\bm{0}_{3\times 1} are always satisfied. ∎

Theorem 2 signifies there always exists the control input 𝒖i,𝒖j3\bm{u}_{i},\bm{u}_{j}\in{\mathbb{R}}^{3} that renders the set 𝒮^ij\hat{\mathcal{S}}_{ij} forward invariant. Therefore, the collision is prevented if the state 𝒙ij\bm{x}_{ij} is in the safe set 𝒮^ij\hat{\mathcal{S}}_{ij} at the initial time. Note that if two ellipses do not collide with each other at the initial time, we can easily specify the initial ϕij\phi_{ij} that satisfies hij>0h_{ij}>0 by any heuristic approach.

We then extend the collision avoidance method (33) to the scenario with nn agents. Similar to the case of the two agents, we need to introduce a supporting line for each pair of agents, resulting in C2n{}_{n}C_{2} numbers of CBFs for the whole system. Since the agent with a smaller ID in a pair owns a supporting line, a vector augmenting ϕij\phi_{ij} of agent ii is described as ϕi=[ϕii+1,ϕii+2,,ϕin]T\bm{\phi}_{i}=[\phi_{i\hskip 1.13809pti+1},\phi_{i\hskip 1.13809pti+2},...,\phi_{i\hskip 1.13809ptn}]^{T} with ϕn=\bm{\phi}_{n}=\emptyset. In addition, we introduce a vector combining hijh_{ij} as 𝒉i=[hii+1,hii+2,,hin]T\bm{h}_{i}=[h_{i\hskip 1.13809pti+1},h_{i\hskip 1.13809pti+2},...,h_{i\hskip 1.13809ptn}]^{T}. Then, the ensemble state and CBFs of the system are expressed as 𝒙=[𝒙1T,𝒙2T,,𝒙nT,ϕ1T,ϕ2T,,ϕn1T]T\bm{x}=\left[\bm{x}_{1}^{T},\bm{x}_{2}^{T},...,\bm{x}_{n}^{T},\bm{\phi}_{1}^{T},\bm{\phi}_{2}^{T},...,\bm{\phi}_{n-1}^{T}\right]^{T} and 𝒉=[𝒉1T,𝒉2T,..,𝒉n1T]T\bm{h}=\left[\bm{h}_{1}^{T},\bm{h}_{2}^{T},..,\bm{h}_{n-1}^{T}\right]^{T}, respectively. With the introduced ensemble vectors, the collision avoidance method for nn agents can be achieved by 𝒖\bm{u}^{*} derived from

𝒖=\displaystyle\bm{u}^{*}= argmin𝒖𝒖𝒖nom2,\displaystyle\mathop{\rm arg~{}min}\limits_{\bm{u}}\|\bm{u}-\bm{u}_{\mathrm{nom}}\|^{2}, (38a)
s.t. Lf𝒉+Lg𝒉𝒖+α(𝒉)0,\displaystyle~{}\mbox{s.t. }L_{f}\bm{h}+L_{g}\bm{h}\bm{u}+\alpha(\bm{h})\geq 0, (38b)

where f=𝟎(3n+nC2)×1f=\bm{0}_{(3n+_{n}C_{2})\times 1} and gij=I(3n+nC2)g_{ij}=I_{(3n+_{n}C_{2})} since we assume a single integrator model.

IV Simulation Results

The proposed algorithm is implemented in simulations to verify that it guarantees collision avoidance between elliptical agents.

IV-A Simulation with Two Agents

We first demonstrate our proposed algorithm with the two elliptical agents, the sizes of which are characterized by Q1=diag(0.4,0.2)Q_{1}=\mathrm{diag}(0.4,0.2) for a red agent and Q2=diag(0.6,0.2)Q_{2}=\mathrm{diag}(0.6,0.2) for a blue agent, respectively. The initial condition of the simulation is depicted in Fig. 5(a), with the initial pose x1(0)=[0,1,π/4]Tx_{1}(0)=\left[0,1,{-\pi}/{4}\right]^{T}, x2(0)=[2,0.1,0]Tx_{2}(0)=\left[2,0.1,0\right]^{T}. Note that we randomly chose the initial ϕ12(0)\phi_{12}(0) from the range that yields a supporting line l12l_{12} separating two ellipses. Two agents traverse the environment so that they intersect around the center of the field. We utilize α(h12)=10h12\alpha(h_{12})=10h_{12} for an extended class 𝒦\mathcal{K} function in CBF conditions and unom,ϕ12=20(h12/ϕ12)u_{\mathrm{nom},\phi_{12}}=20\left({\partial h_{12}}/{\partial\phi_{12}}\right) for a nominal input to ϕ12\phi_{12}.

The snapshots of the simulations are presented in Fig. 5, which illustrate a supporting line incorporated in agent 1’s CBF as a black line. In addition, the distance between the supporting line and the blue ellipse is depicted as a green line. The snapshots demonstrate that the supporting line on agent ii is updated so that it separates two agents without causing any conservative transition in their evasive trajectories. The value of h12h_{12} is illustrated in Fig. 6 together with the optimal solution of (15), namely the actual distance between two agents. Although w12w_{12}^{*} and h12h_{12} differ at the initial time since we set ϕ12\phi_{12} randomly, the proposed gradient-based input (31) successfully drives h12h_{12} to the actual distance w12w_{12}^{*}. Furthermore, Fig. 6 verifies that the input (31) rotates the supporting line so that h12h_{12} follows the transition of the actual distance between two ellipses. We can also confirm that the value of h12h_{12} keeps in the positive value, hence, collision avoidance is achieved.

Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Refer to caption
(d)
Figure 5: Snapshots of the simulation, where two elliptical agent 1 and 2 are depicted in red and blue, respectively. The supporting line of agent 1 is rotated to maximize the distance, shown in the green line, between the line and agent 2.
Refer to caption
Figure 6: Evolution of h12(ϕ12)h_{12}(\phi_{12}), shown in the blue line, and the actual distance w12w_{12}^{*} between two elliptical agents, depicted as the red dashed line, calculated by the optimization problem (15). The control input (31) for ϕ12\phi_{12} makes h12(ϕ12)h_{12}(\phi_{12}) follow w12w_{12}^{*} while keeping the smaller value than w12w_{12}^{*}. Since the value of h12(ϕ12)h_{12}(\phi_{12}) remains positive during the simulation, the collision between two elliptical agents is successfully avoided.

IV-B Simulation with Four Agents

In the next simulation, we demonstrate our proposed algorithm with the four elliptical agents, the sizes of which are specified by Q1=diag(0.3,0.15)Q_{1}=\mathrm{diag}(0.3,0.15) for a red agent, Q2=diag(0.4,0.2)Q_{2}=\mathrm{diag}(0.4,0.2) for a blue agent, Q3=diag(0.4,0.2)Q_{3}=\mathrm{diag}(0.4,0.2) for a green agent and Q4=diag(0.6,0.3)Q_{4}=\mathrm{diag}(0.6,0.3) for an orange agent, respectively. The initial condition of the simulation is depicted in Fig. 7(a), with the initial pose x1(0)=[0.1,1.1,π/4]Tx_{1}(0)=\left[-0.1,1.1,-{\pi}/{4}\right]^{T}, x2(0)=[1.9,1.1,π/4]Tx_{2}(0)=\left[1.9,-1.1,-{\pi}/{4}\right]^{T}, x3(0)=[0.1,1.1,5π/4]Tx_{3}(0)=\left[-0.1,-1.1,{5\pi}/{4}\right]^{T} and x4(0)=[1.9,1.1,5π/4]Tx_{4}(0)=\left[1.9,1.1,{5\pi}/{4}\right]^{T}. All four agents traverse the environment so that they intersect around the center of the field. We utilize α(𝒉)=10𝒉\alpha(\bm{h})=10\bm{h} for an extended class 𝒦\mathcal{K} function in CBF conditions and unom,ϕij=20(hij/ϕij)u_{\mathrm{nom},\phi_{ij}}=20\left({\partial h_{ij}}/{\partial\phi_{ij}}\right) for a nominal input to each ϕij\phi_{ij}.

The snapshots of the simulation are presented in Fig. 7, where each agent is depicted with its trajectory. As illustrated in Fig. 7(b), all agents move straightly toward their goal positions until their distances become closer at the center. Then, the proposed methods modify the nominal input minimally invasive way so that the collision is avoided, as illustrated in Fig. 7(c) and (d). Note that the proposed CBF achieves the collision-free trajectories while changing the attitude of the agents in Fig. 7(c). Fig. 8 depicts the transitions of hijh_{ij}, where the proposed methods keep them in the positive value.

Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Refer to caption
(d)
Figure 7: Snapshots of the simulation, where four elliptical agents are depicted together with their trajectories.
Refer to caption
Figure 8: Evolution of hij,j𝒩\i,i𝒩h_{ij},~{}\forall j\in\mathcal{N}\backslash i,~{}\forall i\in\mathcal{N}. The proposed methods remains hijh_{ij} in the positive value during the simulation.

V Conclusion

In this paper, we proposed the collision avoidance method for elliptical agents that utilizes the novel CBF leveraging a supporting line between agents. We first introduced a supporting line of an agent to develop the separation condition of agents that is implementable as a CBF. However, we observed that a naive choice of a supporting line might render a shorter distance than the actual distance between two agents. To alleviate the conservativeness in this evaluation, we proposed the optimization problem that rotates the supporting line so that the distance between a supporting line and the other agent is maximized. We then proved that the maximum value derived from this optimization problem is equivalent to the actual distance between two agents. We presented the collision avoidance method incorporating the developed CBF together with the gradient ascent law for rotating the supporting line. Finally, numerical simulations showcased the validity of the proposed methods. Future works include extending the proposed framework to nonlinear systems through exponential CBF [27] while embracing 3D ellipsoidal agents.

VI Acknowledgement

We acknowledge Mr. Shunya Yamashita at Tokyo Institute of Technology for helpful discussions on the manuscript.

References

  • [1] O. Arslan and D. E. Koditschek, “Voronoi-based coverage control of heterogeneous disk-shaped robots,” in Proc. IEEE Int. Conf. Robotics and Automation, 2016, pp. 4259–4266.
  • [2] R. Funada, M. Santos, T. Gencho, J. Yamauchi, M. Fujita, and M. Egerstedt, “Visual coverage maintenance for quadcopters using nonsmooth barrier functions,” in Proc. IEEE Int. Conf. Robotics and Automation, 2020, pp. 3255–3261.
  • [3] T. Miyano, J. Romberg, and M. Egerstedt, “Primal-dual gradient dynamics for cooperative unknown payload manipulation without communication,” in Proc. American Control Conf., 2020, pp. 2061–2067.
  • [4] G.-B. Dai and Y.-C. Liu, “Distributed coordination and cooperation control for networked mobile manipulators,” IEEE Trans. Industrial Electronics, vol. 64, no. 6, pp. 5065–5074, 2017.
  • [5] F. S. Barbosa, L. Lindemann, D. V. Dimarogonas, and J. Tumova, “Provably safe control of Lagrangian systems in obstacle-scattered environments,” in Proc. IEEE Conf. Decision and Control, 2020, pp. 2056–2061.
  • [6] C. Zhang and J. M. Kovacs, “The application of small unmanned aerial systems for precision agriculture: A review,” Precision Agriculture, vol. 13, pp. 693–712, 2012.
  • [7] Y. Rizk, M. Awad, and E. W. Tunstel, “Cooperative heterogeneous multi-robot systems: A survey,” ACM Comput. Surv., vol. 52, no. 2, 2019.
  • [8] M. Hoy, A. S. Matveev, and A. V. Savkin, “Algorithms for collision-free navigation of mobile robots in complex cluttered environments: A survey,” Robotica, vol. 33, no. 3, pp. 463–497, 2015.
  • [9] A. D. Ames, S. Coogan, M. Egerstedt, G. Notomista, K. Sreenath, and P. Tabuada, “Control barrier functions: Theory and applications,” in Proc. European Control Conf., 2019, pp. 3420–3431.
  • [10] O. Khatib, “Real-time obstacle avoidance for manipulators and mobile robots,” in Autonomous Robot Vehicles.   ”New York, NY: Springer New York, 1986, pp. 396–404.
  • [11] D. V. Dimarogonas and K. H. Johansson, “On the stability of distance-based formation control,” in Proc. IEEE Conf. Decision and Control, 2008, pp. 1200–1205.
  • [12] R. Olfati-Saber, “Flocking for multi-agent dynamic systems: Algorithms and theory,” IEEE Trans. Automatic Control, vol. 51, no. 3, pp. 401–420, 2006.
  • [13] V. Gazi and K. M. Passino, “A class of attractions/repulsion functions for stable swarm aggregations,” Int. J. Control, vol. 77, no. 18, pp. 1567–1579, 2004.
  • [14] D. M. Stipanović, P. F. Hokayem, M. W. Spong, and D. D. Šiljak, “Cooperative avoidance control for multiagent systems,” J. Dynamic Syst., Measurement, and Control, vol. 129, no. 5, pp. 699–707, 2007.
  • [15] A. D. Ames, X. Xu, J. W. Grizzle, and P. Tabuada, “Control barrier function based quadratic programs for safety critical systems,” IEEE Trans. Automatic Control, vol. 62, no. 8, pp. 3861–3876, 2017.
  • [16] L. Wang, A. D. Ames, and M. Egerstedt, “Safety barrier certificates for collisions-free multirobot systems,” IEEE Trans. Robotics, vol. 33, no. 3, pp. 661–674, 2017.
  • [17] P. Glotfelter, I. Buckley, and M. Egerstedt, “Hybrid nonsmooth barrier functions with applications to provably safe and composable collision avoidance for robotic systems,” IEEE Robotics and Automation Letters, vol. 4, no. 2, pp. 1303–1310, 2019.
  • [18] A. Singletary, K. Klingebiel, J. Bourne, A. Browning, P. Tokumaru, and A. Ames, “Comparative analysis of control barrier functions and artificial potential fields for obstacle avoidance,” in Proc. IEEE/RSJ Int. Conf. Intelligent Robots and Syst., 2021, pp. 8129–8136.
  • [19] P. Glotfelter, J. Cortés, and M. Egerstedt, “A nonsmooth approach to controller synthesis for boolean specifications,” IEEE Trans. Automatic Control, vol. 66, no. 11, pp. 5160–5174, 2021.
  • [20] G. Notomista and M. Egerstedt, “Persistification of robotic tasks,” IEEE Trans. Control Syst. Technol., vol. 29, no. 2, pp. 756–767, 2021.
  • [21] K. D. Do, “Flocking for multiple ellipsoidal agents with limited communication ranges,” Int. Scholarly Research Notices, vol. 2013, p. 13, 2013.
  • [22] C. K. Verginis and D. V. Dimarogonas, “Closed-form barrier functions for multi-agent ellipsoidal systems with uncertain Lagrangian dynamics,” IEEE Control Syst. Letters, vol. 3, no. 3, pp. 727–732, 2019.
  • [23] M. Srinivasan, M. Abate, G. Nilsson, and S. Coogan, “Extent-compatible control barrier functions,” Syst. Control Letters, vol. 150, p. 104895, 2021.
  • [24] W. Wang, J. Wang, and M.-S. Kim, “An algebraic condition for the separation of two ellipsoids,” Computer Aided Geometric Design, vol. 18, no. 6, pp. 531–539, 2001.
  • [25] M. G. Choi, “Computing the closest approach distance of two ellipsoids,” Symmetry, vol. 12, no. 8, p. 1302, 2020.
  • [26] S. Boyd and L. Vandenberghe, Convex Optimization.   Cambridge University Press, 2004.
  • [27] Q. Nguyen and K. Sreenath, “Exponential control barrier functions for enforcing high relative-degree safety-critical constraints,” in Proc. American Control Conf., 2016, pp. 322–328.