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

Planning and Control of Multi-Robot-Object Systems under Temporal Logic Tasks and Uncertain Dynamics

Christos K. Verginis, Yiannis Kantaros, and Dimos V. Dimarogonas C. K. Verginis is with the Division of Signals and Systems, Department of Electrical Engineering, Uppsala University, Uppsala, Sweden. Y. Kantaros is with the Department of Electrical and Systems Engineering, Washington University, St. Louis, MO, USA. D. V. Dimarogonas is with the Division of Decision and Control Systems, Department of Electrical Engineering and Computer Science, KTH Royal Institute of Technology, Stockholm, Sweden.
Abstract

We develop an algorithm for the motion and task planning of a system comprised of multiple robots and unactuated objects under tasks expressed as Linear Temporal Logic (LTL) constraints. The robots and objects evolve subject to uncertain dynamics in an obstacle-cluttered environment. The key part of the proposed solution is the intelligent construction of a coupled transition system that encodes the motion and tasks of the robots and the objects. We achieve such a construction by designing appropriate adaptive control protocols in the lower level, which guarantee the safe robot navigation/object transportation in the environment while compensating for the dynamic uncertainties. The transition system is efficiently interfaced with the temporal logic specification via a sampling-based algorithm to output a discrete path as a sequence of synchronized actions of the robots; such actions satisfy the robots’ as well as the objects’ specifications. The robots execute this discrete path by using the derived low level control protocol. Simulation results verify the proposed framework.

Index Terms:
Multi-agent systems, multi-agent-object systems, linear temporal logic, robot navigation, cooperative manipulation, object transportation, multi.

I Introduction

Temporal-logic-based motion planning has gained significant amount of attention over the last decade, as it provides a fully automated correct-by-design controller synthesis approach for autonomous robots. Temporal logics, such as linear temporal logic (LTL), provide formal high-level languages that can describe planning objectives more complex than the well-studied navigation algorithms, and have been used extensively both in single- as well as in multi-robot setups (see, indicatively, [1, 2, 3, 4, 5, 6, 7, 8, 9, 10]). The task is given as a temporal logic formula, which is coupled with a discrete representation of the underlying system, abstracted from the underlying continuous dynamics, in order to derive an appropriate high-level discrete path.

There exist numerous works that consider temporal-logic-based tasks, both for single- and multi-robot systems [1, 2, 3, 5, 7, 9]. Nevertheless, the related works consider temporal logic-based motion planning for fully actuated, autonomous robotic agents. Consider, however, cases where some unactuated objects must undergo a series of processes in a workspace with autonomous robots (e.g., car factories or automated warehouses). In such cases, the robots, except for satisfying their own motion specifications, are also responsible for coordinating with each other in order to transport the objects around the workspace. When the unactuated objects’ specifications are expressed using temporal logics, then the abstraction of the robots’ behavior becomes much more complex, since it has to take into account the objects’ goals.

In addition, the spatial discretization of a multi-agent system to an abstracted higher level system necessitates the design of appropriate continuous-time controllers for the transition of the agents among the states of discrete system. Many works in the related literature, however, either assume that there exist such continuous controllers or adopt non-realistic assumptions. For instance, many works either do not take into account continuous agent dynamics or consider single or double integrators [3, 9, 1, 8, 10], which can deviate from the actual dynamics of the agents, leading thus to poor performance in real-life scenarios. Discretized abstractions, including design of the discrete state space and/or continuous-time controllers, can be found in [7, 11, 12, 13, 14] for general systems and [15, 16] for multi-agent systems. Moreover, many works adopt dimensionless point-mass agents and therefore do not consider inter-agent collision avoidance [7, 9, 10], which can be a crucial safety issue in applications involving autonomous robots.

Since we aim at incorporating the unactuated objects’ specifications in our framework, the robots have to perform (cooperative) transportation of the objects around the workspace, while avoiding collisions with obstacles. Cooperative transportation/manipulation has been extensively studied in the literature (see, for instance, [17, 18, 19, 20, 21, 22, 23, 24, 25, 26, 27]), with collision avoidance specifications being incorporated in [28, 24, 25]. Cooperative object transportation under temporal logics has also been considered in our previous work [29].

This paper proposes a novel algorithm for the motion and task planning of a multi-robot-object system, i.e., a system that consists of multiple robots and unactuated objects, subject to tasks that are expressed by LTL constraints. We consider that the robots and the objects evolve subject to continuous-time and -space dynamics that suffer from uncertainties and in an environment cluttered with obstacles. Our contribution is summarized as follows. First, we abstract the continuous dynamics of the multi-robot-object system into a finite transition system that encodes the coupled behavior of the robots and the objects in the environment. We achieve such an abstraction by developing continuous feedback-control protocols that guarantee i) the navigation of the robots and ii) the cooperative transportation of the objects by the robots in the environment. These protocols combine barrier functions and point-world transformations to guarantee collision-avoidance properties as well as adaptive-control methodologies to compensate for the dynamic uncertainties. The robot navigation and cooperative object transportation constitute robot action primitives that enable the transitions in the aforementioned transition system. Second, we compose the resulting transition system with an automaton-based representation of the underlying LTL task. Finally, we use a sampling-based procedure to search the composition of the transition system and the task’s automaton for the derivation of an optimal high-level plan, as a sequence of robot actions, that satisfies the task.

This paper extends our previous work [30] in the following directions: firstly, we consider transportation of the objects by multiple robots (as opposed to [30], which consider single-robot object transportation). Secondly, we consider more complex, obstacle-cluttered environments. Thirdly, we consider dynamic uncertainty in the dynamics of the robots and objects. Finally, instead of model-checking tools based on graph search algorithms, we use a sampling-based procedure to derive an optimal task-satisfying plan, yielding significantly lower complexity in terms of runtime and memory.

The rest of the paper is organized as follows. Section II provides necessary background and formulates the problem, while Section III presents the proposed solution. Section IV provides simulation results and, finally, Section V concludes the paper.

II Preliminaries and Problem Formulation

II-A Task Specification in LTL

We focus on the task specification ϕ\phi given as a Linear Temporal Logic (LTL) formula. The basic ingredients of a LTL formula are a set of atomic propositions Ψ\Psi and several boolean and temporal operators. LTL formulas are formed according to the following grammar [31]: ϕ:true|a|ϕ1ϕ2|¬ϕ|ϕ|ϕ1𝒰ϕ2\phi:\coloneqq\mathrm{true}\ |\ a\ |\ \phi_{1}\ \land\phi_{2}\ |\ \neg\phi\ |\ \circ\phi\ |\ \phi_{1}\mathcal{U}\phi_{2}, where aΨa\in\Psi, ϕ1\phi_{1} and ϕ2\phi_{2} are LTL formulas and \circ, 𝒰\mathcal{U} are the next and until operators, respectively. Definitions of other useful operators like \square (always), \Diamond (eventually) and \Rightarrow (implication) are omitted and can be found at [31]. The semantics of LTL are defined over infinite words over 2Ψ2^{\Psi}. Intuitively, an atomic proposition ψΨ\psi\in\Psi is satisfied on a word w=w1w2w=w_{1}w_{2}\dots if it holds at its first position w1w_{1}, i.e. ψw1\psi\in w_{1}. Formula ϕ\phi holds true if ϕ\phi is satisfied on the word suffix that begins in the next position w2w_{2}, whereas ϕ1𝒰ϕ2\phi_{1}\mathcal{U}\phi_{2} states that ϕ1\phi_{1} has to be true until ϕ2\phi_{2} becomes true. Finally, ϕ\Diamond\phi and ϕ\square\phi holds on ww eventually and always, respectively. For a full definition of the LTL semantics, we refer the reader to [31].

II-B Problem Formulation

Consider N>1N>1 robotic agents operating in a compact 22D workspace 𝒲2\mathcal{W}\subset\mathbb{R}^{2} with an outer boundary with M>0M>0 objects, and 𝒩{1,,N}\mathcal{N}\coloneqq\{1,\dots,N\}, {1,,M}\mathcal{M}\coloneqq\{1,\dots,M\}; The objects are represented by rigid bodies whereas the robotic agents are fully actuated and holonomic, equipped with a transportation tool, such as a robotic arm. In addition, the workspace is populated with JJ\in\mathbb{N} connected, closed sets {𝔒j}j𝒥\{\mathfrak{O}_{j}\}_{j\in\mathcal{J}}, with 𝒥{1,,J}\mathcal{J}\coloneqq\{1,\dots,J\}, representing obstacles, and we define the free space as 𝒲f𝒲\j𝒥𝔒j\mathcal{W}_{\textup{f}}\coloneqq\mathcal{W}\backslash\bigcup_{j\in\mathcal{J}}\mathfrak{O}_{j}.

We further assume that there exist K>1K>1 points within 𝒲free\mathcal{W}_{\textup{free}}, denoted by pπkp_{\pi_{k}}, corresponding to certain properties of interest (e.g., gas station, repairing area, etc.), with 𝒦{1,,K}\mathcal{K}\coloneqq\{1,\dots,K\}. Since, in practise, these properties are naturally inherited to some neighborhood of the respective point of interest, we define for each k𝒦k\in\mathcal{K} the region of interest πk\pi_{k}, corresponding to pπkp_{\pi_{k}}, as the closed ball πk¯(pπk,rπk)𝒲free\pi_{k}\coloneqq\bar{\mathcal{B}}(p_{\pi_{k}},r_{\pi_{k}})\subset\mathcal{W}_{\textup{free}}, with rπk>0r_{\pi_{k}}>0 positive radii, k𝒦\forall k\in\mathcal{K}. These properties of interest are expressed as boolean variables via finite sets of atomic propositions. In particular, we introduce disjoint sets of atomic propositions Ψi,Ψjo\Psi_{i},\Psi^{o}_{j}, expressed as boolean variables, that represent services provided to agent i𝒩i\in\mathcal{N} and object jj\in\mathcal{M} in Π{π1,,πK}\Pi\coloneqq\{\pi_{1},\dots,\pi_{K}\}. The services provided at each region πk\pi_{k} are given by the labeling functions i:Π2Ψi,jo:Π2Ψjo\mathcal{L}_{i}:\Pi\rightarrow 2^{\Psi_{i}},\mathcal{L}^{o}_{j}:\Pi\rightarrow 2^{\Psi^{o}_{j}}, which assign to each region πk,k𝒦\pi_{k},k\in\mathcal{K}, the subset of services Ψi\Psi_{i} and Ψjo\Psi^{o}_{j}, respectively, that can be provided in that region to agent i𝒩i\in\mathcal{N} and object jj\in\mathcal{M}, respectively. In addition, we consider that the agents and the object are initially (t=0t=0) in the regions of interest πinit(i),πinito(j)\pi_{init(i)},\pi_{init_{o}(j)}, where the functions init:𝒩𝒦init:\mathcal{N}\to\mathcal{K}, inito:𝒦init_{o}:\mathcal{M}\to\mathcal{K} specify the initial region indices. In the following, we present the modeling of the coupled dynamics of the object and the robots.

We denote by xi2x_{i}\in\mathbb{R}^{2} the position of robot ii’s center of mass. In this work we explicitly consider the actions of robot navigation, as well as (cooperative) transportation of an object via xix_{i}, where the joint variables of the potential mounted robotic arm are assumed to be fixed. We consider that the robotic arm joints are used only for grasping/releasing an object (when the respective mobile base is fixed), actions that we do not explicitly model. The dynamics describing the motion of robot ii’s center of mass are given by

mix¨i+fi(xi,x˙i)=uihi,\displaystyle m_{i}\ddot{{x}}_{i}+f_{i}(x_{i},\dot{{x}}_{i})=u_{i}-h_{i}, (1)

where mim_{i} is the unknown mass of robot ii, fi:42f_{i}:\mathbb{R}^{4}\to\mathbb{R}^{2} are unknown friction-like functions, ui2u_{i}\in\mathbb{R}^{2} is the control input of robot ii, and hi2h_{i}\in\mathbb{R}^{2} is the force exerted by robot ii at the grasping point with object jj in case of contact. The aforementioned dynamics concern the cases when (i) robot ii is navigating to some pre-defined point, and (ii) robot ii is transporting, possibly collaboratively with other robots, an object jj. In both of these cases, the joint variables of the mounted robotic arm are assumed to be constant. The procedures of grasping/releasing an object, where the robotic arm would have to be activated, are not considered here.

We consider that each robot ii, for a given xi{x}_{i}, covers a spherical region of constant radius ri>0r_{i}\in\mathbb{R}_{>0} that bounds its volume, i.e., ¯(xi,ri)\bar{\mathcal{B}}({x}_{i},r_{i}), i𝒩\forall i\in\mathcal{N}; Moreover, we consider that the agents have specific power capabilities, which for simplicity, we associate to positive integers ζi>0\zeta_{i}>0, i𝒩i\in\mathcal{N}, via an analogous relation. The overall configuration is x[x1,,xN]{x}\coloneqq[{x}^{\top}_{1},\dots,{x}^{\top}_{N}]^{\top} 2N\in\mathbb{R}^{2N}.

Regarding the objects, we denote by xjo2x^{o}_{j}\in\mathbb{R}^{2} the position of the jjth object’s center of mass, j\forall j\in\mathcal{M}. We consider the following second-order Newton-Euler dynamics:

mjox¨jo+fjo(xjo,x˙jo)=hjo,\displaystyle m^{o}_{j}\ddot{{x}}^{o}_{j}+f^{o}_{j}(x^{o}_{j},\dot{x}^{o}_{j})=h^{o}_{j}, (2)

where mjom^{o}_{j} is the unknown mass of object jj, fjo:42f^{o}_{j}:\mathbb{R}^{4}\to\mathbb{R}^{2} are unknown friction terms, and hjo2h^{o}_{j}\in\mathbb{R}^{2} are the forces exerted to the jjth object’s center of mass, in case of contact with a robot. The overall object configuration is denoted by xo[x1o,,xMo]x^{o}\coloneqq[x^{o}_{1},\dots,x^{o}_{M}]^{\top}, and x¯[x,(xo)]\bar{x}\coloneqq[x^{\top},(x^{o})^{\top}]^{\top}.

The functions fif_{i} and fjof^{o}_{j} are assumed to satisfy the following assumption:

Assumption 1

The functions fi,fjo:42f_{i},f^{o}_{j}:\mathbb{R}^{4}\to\mathbb{R}^{2} are analytic and satisfy

fi(𝗑,𝗒)αi𝗒,\displaystyle\|f_{i}(\mathsf{x},\mathsf{y})\|\leq\alpha_{i}\|\mathsf{y}\|, (3)
fjo(𝗑,𝗒)αjo𝗒,\displaystyle\|f^{o}_{j}(\mathsf{x},\mathsf{y})\|\leq\alpha^{o}_{j}\|\mathsf{y}\|, (4)

𝗑,𝗒2\forall\mathsf{x},\mathsf{y}\in\mathbb{R}^{2}, where αi\alpha_{i}, αjo\alpha^{o}_{j} are unknown positive constants, i𝒩,j\forall i\in\mathcal{N},j\in\mathcal{M}.

The aforementioned assumption is inspired by standard friction-like terms, which can be approximated by continuously differentiable velocity functions [32].

Similarly to the robots, each object’s volume is represented by the spherical set of constant radius rjo>0r^{o}_{j}\in\mathbb{R}_{>0}, i.e., ¯(xjo,rjo)\bar{\mathcal{B}}({x}^{o}_{j},r^{o}_{j}), j\forall j\in\mathcal{M}.

Next, we provide the coupled dynamics between an object jj\in\mathcal{M} and a subset 𝒜𝒩\mathcal{A}\subseteq\mathcal{N} of robots that grasp it rigidly. For these robots, it holds that hjo=i𝒜hih^{o}_{j}=\sum_{i\in\mathcal{A}}h_{i}, and since the joint variables of the robotic arms are fixed, xjo=xi+dijox^{o}_{j}=x_{i}+d^{o}_{ij}, x˙jo=x˙i\dot{x}^{o}_{j}=\dot{x}_{i} and x¨jo=x¨i\ddot{x}^{o}_{j}=\ddot{x}_{i}, where dijo2d^{o}_{ij}\in\mathbb{R}^{2} is the constant offset between xjox^{o}_{j} and xix_{i}, i𝒜\forall i\in\mathcal{A}. Therefore, one obtains that

m𝒜,jx¨jo+f𝒜,j(xjo,x˙jo)=i𝒜ui,{m}_{\mathcal{A},j}\ddot{x}^{o}_{j}+{f}_{\mathcal{A},j}(x^{o}_{j},\dot{x}^{o}_{j})=\sum_{i\in\mathcal{A}}u_{i}, (5)

where m𝒜,jmjo+i𝒜,jmi{m}_{\mathcal{A},j}\coloneqq m^{o}_{j}+\sum_{i\in\mathcal{A},j}m_{i}, f𝒜,jfjo(xjo,x˙jo)+i𝒜,jfi(xjodijo,x˙jo){f}_{\mathcal{A},j}\coloneqq f^{o}_{j}(x^{o}_{j},\dot{x}^{o}_{j})+\sum_{i\in\mathcal{A},j}f_{i}(x^{o}_{j}-d^{o}_{ij},\dot{x}^{o}_{j}). Note that Assumption 1 implies that

f𝒜,j(xjo,x˙jo)α𝒜,jx˙jo,\|{f}_{\mathcal{A},j}(x^{o}_{j},\dot{x}^{o}_{j})\|\leq\alpha_{\mathcal{A},j}\|\dot{x}^{o}_{j}\|, (6)

for an unknown positive constant α𝒜,j\alpha_{\mathcal{A},j}.

Regarding the volume of the coupled robots-object system, we consider that it is bounded by a sphere of radius centered at xjo{x}^{o}_{j} with constant radius r𝒜,j>0r_{\scriptscriptstyle\mathcal{A},j}\in\mathbb{R}_{>0}, i.e., (xjo,r𝒜,j)\mathcal{B}({x}^{o}_{j},r_{\scriptscriptstyle\mathcal{A},j}), which is large enough to cover the volume of the coupled system.

Moreover, in order to take into account the introduced robots’ power capabilities ζi\zeta_{i}, i𝒩i\in\mathcal{N}, we consider a function Λ{𝖳𝗋𝗎𝖾,𝖥𝖺𝗅𝗌𝖾}\Lambda\in\{\mathsf{True},\mathsf{False}\} that outputs whether the robots that grasp an object are able to transport the object, based on their power capabilities. For instance, Λ(mjo,ζ𝒜)=𝖳𝗋𝗎𝖾\Lambda(m^{o}_{j},\zeta_{\mathcal{A}})=\mathsf{True}, where mjo>0m^{o}_{j}\in\mathbb{R}_{>0} is the mass of object jj and ζ𝒜[ζi]i𝒜\zeta_{\mathcal{A}}\coloneqq[\zeta_{i}]^{\top}_{i\in\mathcal{A}}, implies that the robots 𝒜\mathcal{A} have sufficient power capabilities to cooperatively transport object jj.

Next, we define the boolean functions 𝒜𝒢j:02𝒩0\mathcal{AG}_{j}:\mathbb{R}_{\geq 0}\to 2^{\mathcal{N}_{0}}, with 𝒩0𝒩{0}\mathcal{N}_{0}\coloneqq\mathcal{N}\cup\{0\}, to denote which robots grasp an object jj\in\mathcal{M}; 𝒜𝒢j=0\mathcal{AG}_{j}=0 means that no robots grasp object jj. Note also that i𝒜𝒢ji𝒜𝒢j=𝖥𝖺𝗅𝗌𝖾,j\{j}i\in\mathcal{AG}_{j}\Leftrightarrow i\notin\mathcal{AG}_{j^{\prime}}=\mathsf{False},\forall j^{\prime}\in\mathcal{M}\backslash\{j\}, i.e., robot ii can grasp at most one object at a time. We further denote 𝒜𝒢[𝒜𝒢1,,𝒜𝒢M](2𝒩0)M\mathcal{AG}\coloneqq[\mathcal{AG}_{1},\dots,\mathcal{AG}_{M}]^{\top}\in(2^{\mathcal{N}_{0}})^{M}.

In the following, we use the term “entity” to refer to single robots, objects as well as systems comprised by robots that grasp an object (robots-object systems). The number of these systems depends on the variables 𝒜𝒢\mathcal{AG}. Given a grasping configuration 𝒜𝒢(2𝒩0)M\mathcal{AG}\in(2^{\mathcal{N}_{0}})^{M}, consider T¯(𝒜𝒢)\bar{T}(\mathcal{AG}) number of entities, indexed by the set 𝒯(𝒜𝒢){1,,T¯(𝒜𝒢)}\mathcal{T}(\mathcal{AG})\coloneqq\{1,\dots,\bar{T}(\mathcal{AG})\}. Each entity (robot, object, or robots-object system) is characterized by the respective configuration (xix_{i}, xjox_{j}^{o}, xjox_{j^{\prime}}^{o}) and radius (rir_{i}, rjor^{o}_{j}, r𝒜,jr_{\mathcal{A},j^{\prime}}), respectively, which we denote for simplicity by the generic variables xiex_{i}^{e}, rier_{i}^{e}, for all i𝒯(𝒜𝒢)i\in\mathcal{T}(\mathcal{AG}).

We further define the free space for each entity i(𝒜𝒢)(𝒲f\j𝒯(𝒜𝒢)\{i}¯(xje,rje))rie\mathcal{F}_{i}(\mathcal{AG})\coloneqq\big{(}\mathcal{W}_{\textup{f}}\backslash\bigcup_{j\in\mathcal{T}(\mathcal{AG})\backslash\{i\}}\bar{\mathcal{B}}(x_{j}^{e},r^{e}_{j})\big{)}\ominus r^{e}_{i}, where the incorporation of rie\ominus r^{e}_{i} enlarges the obstacles and the other entities with the radius rier_{i}^{e}.

We give now the definitions for the transitions of the robots and the objects between the regions of interest.

Definition 1

(Navigation) Let 𝒜𝒢(t0)(2𝒩0)M\mathcal{AG}(t_{0})\in(2^{\mathcal{N}_{0}})^{M} such that

xie(t0)i(𝒜𝒢(t0))¯i(xie(t0),rie)πik}i𝒯(𝒜𝒢(t0))\displaystyle\left.\begin{matrix}[l]&x^{e}_{i}(t_{0})\in\mathcal{F}_{i}(\mathcal{\mathcal{AG}}(t_{0}))\\ &\bar{\mathcal{B}}_{i}(x_{i}^{e}(t_{0}),r_{i}^{e})\subset\pi_{i_{k}}\end{matrix}\ \ \ \right\}i\in\mathcal{T}(\mathcal{AG}(t_{0}))

where ik𝒦i_{k}\in\mathcal{K}, for all i𝒯(𝒜𝒢)i\in\mathcal{T}(\mathcal{AG}) and some t00t_{0}\in\mathbb{R}_{\geq 0}. Then, entity j𝒯(𝒜𝒢)j\in\mathcal{T}(\mathcal{AG}) executes a transition from πjk\pi_{j_{k}} to πjk\pi_{j_{k^{\prime}}}, with jk𝒦\{jk}j_{k^{\prime}}\in\mathcal{K}\backslash\{j_{k}\}, if there exists a finite tft0t_{f}\geq t_{0} such that

¯j(xje(tf),rje)πjk\displaystyle\bar{\mathcal{B}}_{j}(x_{j}^{e}(t_{f}),r_{j}^{e})\subset\pi_{j_{k^{\prime}}}
xj(t)j(𝒜𝒢)\displaystyle x_{j}(t)\in\mathcal{F}_{j}(\mathcal{AG})

for all t[t0,tf]t\in[t_{0},t_{f}].

Definition 2

(Grasping) Let 𝒜𝒢(t0)(2𝒩0)M\mathcal{AG}(t_{0})\in(2^{\mathcal{N}_{0}})^{M} such that

xie(t0)i(𝒜𝒢(t0))¯i(xie(t0),rie)πik}i𝒯(𝒜𝒢(t0))\displaystyle\left.\begin{matrix}[l]&x^{e}_{i}(t_{0})\in\mathcal{F}_{i}(\mathcal{AG}(t_{0}))\\ &\bar{\mathcal{B}}_{i}(x_{i}^{e}(t_{0}),r_{i}^{e})\subset\pi_{i_{k}}\end{matrix}\ \ \ \right\}i\in\mathcal{T}(\mathcal{AG}(t_{0}))
j𝒜𝒢(t0)\displaystyle\hskip 14.22636ptj\notin\mathcal{AG}_{\ell}(t_{0})
xj(t0),xo(t0)π\displaystyle\hskip 14.22636ptx_{j}(t_{0}),x_{\ell}^{o}(t_{0})\in\pi_{\ell}

for some j𝒩j\in\mathcal{N}, \ell\in\mathcal{M}, 𝒦\ell\in\mathcal{K}, t00t_{0}\in\mathbb{R}_{\geq 0} where ik𝒦i_{k}\in\mathcal{K}, for all i𝒯(𝒜𝒢)i\in\mathcal{T}(\mathcal{AG}). Then, agent jj grasps object \ell, denoted by j𝑔j\xrightarrow{g}\ell, if there exists a finite tft0t_{f}\geq t_{0} such that j𝒜𝒢(tf)j\in\mathcal{AG}_{\ell}(t_{f}).

Similarly, we can define the releasing action j𝑟j\xrightarrow[]{r}\ell for an agent j𝒩j\in\mathcal{N} and object \ell\in\mathcal{M}. Loosely speaking, the aforementioned definitions correspond to specific action primitives of the robots, namely robot navigation or object transportation, which define the navigation transition, and grasping and releasing actions. When the navigation transition from πjk\pi_{j_{k}} to πjk\pi_{j^{\prime}_{k}} corresponds to a robot navigation for robot j𝒩j\in\mathcal{N}, we denote it by πjkjπjk\pi_{j_{k}}\rightarrow_{j}\pi_{j^{\prime}_{k}}; when it corresponds to a cooperative transportation of object jj\in\mathcal{M} by a subset of robots 𝒜\mathcal{A}, we denote it by πjk𝑇𝒜,jπjk\pi_{j_{k}}\xrightarrow{T}_{\mathcal{A},j}\pi_{j^{\prime}_{k}}.

We also assume the existence of a procedure 𝒫s{\mathcal{P}}_{s} that outputs whether or not a set of non-intersecting spheres fits in a larger sphere as well as possible positions of the spheres in the case they fit. More specifically, given a region of interest πk\pi_{k} and a number N~\widetilde{N}\in\mathbb{N} of sphere radii (of robots, objects, or object-robots systems) the procedure can be seen as a function 𝒫s[𝒫s,0,𝒫s,1]{\mathcal{P}}_{s}\coloneqq[\mathcal{P}_{s,0},{\mathcal{P}}^{\top}_{s,1}]^{\top}, where 𝒫s,0:0N~+1{𝖳𝗋𝗎𝖾,𝖥𝖺𝗅𝗌𝖾}\mathcal{P}_{s,0}:\mathbb{R}^{\widetilde{N}+1}_{\geq 0}\to\{\mathsf{True},\mathsf{False}\} outputs whether the spheres fit in the region πk\pi_{k} whereas 𝒫s,1{\mathcal{P}}_{s,1} provides possible configurations of the robots and the objects or 0{0} in case the spheres do not fit. For instance, Ps,0(rπ2,r1,r3,r1o,r5o)P_{s,0}(r_{\pi_{2}},r_{1},r_{3},r^{o}_{1},r^{o}_{5}) determines whether the robots 1,31,3 and the objects 1,51,5 fit in region π2\pi_{2}, without colliding with each other; (x1e,x2e,x3e,x4e)=(x1,x3,x1o,x5o)=Ps,1(rπ2,r1e,r2e,r3e,r4e)=Ps,1(rπ2,r1,r3,r1o,r5o)(x^{e}_{1},x^{e}_{2},x^{e}_{3},x^{e}_{4})=(x_{1},x_{3},{x}^{o}_{1},{x}^{o}_{5})=P_{s,1}(r_{\pi_{2}},r^{e}_{1},r^{e}_{2},r^{e}_{3},r^{e}_{4})=P_{s,1}(r_{\pi_{2}},r_{1},r_{3},r^{o}_{1},r^{o}_{5}) provides a set of configurations such that ¯i(xie,rie)\bar{\mathcal{B}}_{i}(x^{e}_{i},r^{e}_{i}) i{1,,4}i\in\{1,\dots,4\}, and the pairwise intersections of these sets are empty. The problem of finding an algorithm 𝒫s\mathcal{P}_{s} is a special case of the sphere packing problem [33]. Note, however, that we are not interested in finding the maximum number of spheres that can be packed in a larger sphere but, rather, in the simpler problem of determining whether a set of spheres can be packed in a larger sphere.

Our goal is to control the multi-robot-object system defined above such that the robots and the objects obey a given specification over their atomic propositions Ψi,Ψjo,i𝒩,j\Psi_{i},\Psi^{o}_{j},\forall i\in\mathcal{N},j\in\mathcal{M}. Given the trajectories xi(t),xjo(t),t0{x}_{i}(t),{x}^{o}_{j}(t),t\in\mathbb{R}_{\geq 0}, of robot ii and object jj, respectively, their corresponding behaviors are given by the infinite sequences

bi(xi(t),σi)(xi(ti,1),σi,1)(xi(ti,2),σi,2),\displaystyle b_{i}\coloneqq({x}_{i}(t),\sigma_{i})\coloneqq({x}_{i}(t_{i,1}),\sigma_{i,1})({x}_{i}(t_{i,2}),\sigma_{i,2})\dots,
bjo(xjo(t),σjo)(xjo(tj,1o),σj,1o)(xjo(tj,2o),σj,2o),\displaystyle b^{o}_{j}\coloneqq({x}^{o}_{j}(t),\sigma^{o}_{j})\coloneqq({x}^{o}_{j}(t^{o}_{j,1}),\sigma^{o}_{j,1})({x}^{o}_{j}(t^{o}_{j,2}),\sigma^{o}_{j,2})\dots,

with ti,+1>ti,0,tj,+1o>tj,o0,t_{i,\ell+1}>t_{i,\ell}\geq 0,t^{o}_{j,\ell+1}>t^{o}_{j,\ell}\geq 0,\forall\ell\in\mathbb{N}, representing specific time stamps. The sequences σi,σjo\sigma_{i},\sigma^{o}_{j} are the services provided to the agent and the object, respectively, over their trajectories, i.e., σi,2Ψi,σj,lo2Ψjo\sigma_{i,\ell}\in 2^{\Psi_{i}},\sigma^{o}_{j,l}\in 2^{\Psi^{o}_{j}} with 𝒜i(qi(ti,))πki,,σi,i(πki,)\mathcal{A}_{i}({q}_{i}(t_{i,\ell}))\subset\pi_{k_{i,\ell}},\sigma_{i,\ell}\in\mathcal{L}_{i}(\pi_{k_{i,\ell}}) and 𝒪j(xjo(tj,lo))πkj,lo,σj,lojo(πkj,lo),ki,,kj,lo𝒦,,l,i𝒩,j\mathcal{O}_{j}({x}^{o}_{j}(t^{o}_{j,l}))\subset\pi_{k^{o}_{j,l}},\sigma^{o}_{j,l}\in\mathcal{L}^{o}_{j}(\pi_{k^{o}_{j,l}}),k_{i,\ell},k^{o}_{j,l}\in\mathcal{K},\forall\ell,l\in\mathbb{N},i\in\mathcal{N},j\in\mathcal{M}, where i\mathcal{L}_{i} and jo\mathcal{L}^{o}_{j} are the previously defined labeling functions. The following Lemma then follows:

Lemma 1

The behaviors bi,bjob_{i},b^{o}_{j} satisfy formulas ϕi,ϕoj\phi_{i},\phi_{o_{j}} if σiϕi\sigma_{i}\models\phi_{i} and σjoϕjo\sigma^{o}_{j}\models\phi^{o}_{j}, respectively.

The control objectives are given as LTL formulas ϕi,ϕjo\phi_{i},\phi^{o}_{j} over Ψi,Ψjo\Psi_{i},\Psi^{o}_{j}, respectively, i𝒩,j\forall i\in\mathcal{N},j\in\mathcal{M}. The LTL formulas ϕi,ϕjo\phi_{i},\phi^{o}_{j} are satisfied if there exist behaviors bi,bjob_{i},b^{o}_{j} of agent ii and object jj that satisfy ϕi,ϕjo\phi_{i},\phi^{o}_{j}. We are now ready to give a formal problem statement:

Problem 1

Consider NN robotic agents and MM objects in 𝒲\mathcal{W} satisfying

¯i(xi(0),ri)πinit(i),¯j(xjo(0),rjo)πinito(j),x˙i(0)=0,\displaystyle\bar{\mathcal{B}}_{i}(x_{i}(0),r_{i})\subset\pi_{\text{init}(i)},\bar{\mathcal{B}}_{j}({x}^{o}_{j}(0),r^{o}_{j})\subset\pi_{\text{init}_{o}(j)},\dot{{x}}_{i}(0)={0},
x˙jo=0,i𝒩,j,\displaystyle\dot{x}^{o}_{j}={0},\forall i\in\mathcal{N},j\in\mathcal{M},

in collision-free initial configurations. Given the disjoint sets Ψi,Ψjo\Psi_{i},\Psi^{o}_{j}, NN LTL formulas ϕi\phi_{i} over Ψi\Psi_{i} and MM LTL formulas ϕjo\phi^{o}_{j} over Ψjo\Psi^{o}_{j}, develop a control strategy that achieves behaviors bi,bjob_{i},b^{o}_{j} which yield the satisfaction of ϕi,ϕjo,i𝒩,j\phi_{i},\phi^{o}_{j},\forall i\in\mathcal{N},j\in\mathcal{M}.

Note that it is implicit in the problem statement the fact that the agents/objects starting in the same region can actually fit without colliding with each other. Technically, it holds that 𝒫s,0(rπk,[ri]i{i𝒩:init(i)=k},[rjo]j{j:inito(j)=k})=𝖳𝗋𝗎𝖾\mathcal{P}_{s,0}(r_{\pi_{k}},[r_{i}]_{i\in\{i\in\mathcal{N}:\text{init}(i)=k\}},[r^{o}_{j}]_{j\in\{j\in\mathcal{M}:\text{init}_{o}(j)=k\}})=\mathsf{True}, k𝒦\forall k\in\mathcal{K}.

III Main Results

III-A Continuous Control Design

The first ingredient of our solution is the development of feedback control laws that establish the navigation transition of Def. 1, i.e., robot navigation and cooperative object transportations. We do not focus on the grasping/releasing actions (see Def. 2) and we refer to some existing methodologies that can derive the corresponding control laws (e.g., [34, 35]).

The control design is based on the integration of the adaptive control scheme and point world transformation proposed in [36] and [37], respectively. The former accommodates the uncertain robot and object dynamics and the latter deals with the complex environment obstacles. Since [37] consider single-robot navigation in complex environments, we consider here a sequential execution of the navigation and object-transportation transitions111The correctness of the multi-robot adaptive control scheme of [36] has only been proven for simple, spherical environments.. That is, only one entity is allowed to navigate from one RoI to another, viewing the rest of the entities as fixed obstacles.

1) Robot Navigation

Assume that the conditions of Problem 1 hold for some t00t_{0}\in\mathbb{R}_{\geq 0}, i.e., all robots and objects are located in some RoI with zero velocity. We first consider the control problem of safe navigation for a robot i𝒩i\in\mathcal{N} satisfying i𝒜𝒢j(t0)i\notin\mathcal{AG}_{j}(t_{0}) for all jj\in\mathcal{M}, i.e., not grasping any objects, from region iki_{k} to iki^{\prime}_{k} (transition πikiπik\pi_{i_{k}}\rightarrow_{i}\pi_{i^{\prime}_{k}}), with ¯i(xi(t0),ri)πik\bar{\mathcal{B}}_{i}(x_{i}(t_{0}),r_{i})\subset\pi_{i_{k}} and iki_{k}, iki^{\prime}_{k} 𝒦\in\mathcal{K}. Let xie=xix^{e}_{i}=x_{i}. As stated before, the other entities are fixed and viewed as static obstacles by robot ii. Moreover, to account for safety specifications, we wish the robot to avoid entering the RoI πk\pi_{k}, k𝒦\{ik,ik}k\in\mathcal{K}\backslash\{i_{k},i^{\prime}_{k}\}. Therefore, the free space of robot ii becomes i=(𝒲f\𝒪i)ri\mathcal{F}_{i}=(\mathcal{W}_{\textup{f}}\backslash\mathcal{O}_{i})\ominus r_{i}, where

𝒪i(𝒯(𝒜𝒢)\{i}¯(xe,re))(k𝒦\{ik,ik}πk)\displaystyle\mathcal{O}_{i}\coloneqq\left(\bigcup_{\ell\in\mathcal{T}(\mathcal{AG})\backslash\{i\}}\bar{\mathcal{B}}(x^{e}_{\ell},r^{e}_{\ell})\right)\bigcup\left(\bigcup_{k\in\mathcal{K}\backslash\{i_{k},i_{k}^{\prime}\}}\pi_{k}\right)

Let the desired navigation goal of the robot be xd,iπikx_{\textup{d,i}}\in\pi_{i^{\prime}_{k}}, which will be provided by the procedure 𝒫s\mathcal{P}_{s}, as explained in the previous section. Next, we use the workspace transformation χ=(x)\chi=\mathcal{H}(x) from [37] that converts the environment to the open unit disk 𝒟([0,0],1)\mathcal{D}\coloneqq\mathcal{B}([0,0]^{\top},1), and the obstacles (environment obstacles, entities other than robot ii, and RoI other than πik\pi_{i_{k}}, πik\pi_{i^{\prime}_{k}}) to J¯\bar{J} points b𝒟b_{\ell}\in\mathcal{D}, 𝒥¯{1,,J¯}\ell\in\bar{\mathcal{J}}\coloneqq\{1,\dots,\bar{J}\}, robot ii to the point χi(xi)\chi_{i}\coloneqq\mathcal{H}(x_{i}), and the goal of robot ii to χd(xd,i)\chi_{\textup{d}}\coloneqq\mathcal{H}(x_{\textup{d,i}}). In the subsequent analysis we omit the robot index ii for ease of presentation.

Let now a constant r¯>0\bar{r}>0 satisfying

bibj>2r¯,i,j𝒥¯,ij\displaystyle\|b_{i}-b_{j}\|>2\bar{r},\forall i,j\in\bar{\mathcal{J}},i\neq j
1bi>2r¯,i𝒥¯.\displaystyle 1-\|b_{i}\|>2\bar{r},\forall i\in\bar{\mathcal{J}}.

The transformed free space of the robot can be defined as 𝒟\{b1,,bJ¯}\mathcal{F}_{\mathcal{H}}\coloneqq\mathcal{D}\backslash\{b_{1},\dots,b_{\bar{J}}\}. We define next the set 𝒥¯0{0}𝒥¯\bar{\mathcal{J}}_{0}\coloneqq\{0\}\cup\bar{\mathcal{J}} as well as the distances d:0d_{\ell}:\mathcal{F}_{\mathcal{H}}\to\mathbb{R}_{\geq 0}, 𝒥¯0\ell\in\bar{\mathcal{J}}_{0}, with d(χ)χb2d_{\ell}(\chi)\coloneqq\|\chi-b_{\ell}\|^{2}, 𝒥¯\forall\ell\in\bar{\mathcal{J}}, and d0(x)1χ2d_{0}(x)\coloneqq 1-\|\chi\|^{2}. Note that, by keeping d(χ)>0d_{\ell}(\chi)>0, d0(χ)>0d_{0}(\chi)>0, we guarantee that χ\chi\in\mathcal{F}_{\mathcal{H}} and hence the safety of the robot. We also define the constant

r¯dmin{1χd2,min𝒥¯{d(χd)}}\bar{r}_{\textup{d}}\coloneqq\min\left\{1-\|\chi_{\textup{d}}\|^{2},\min_{\ell\in\bar{\mathcal{J}}}\left\{d_{\ell}(\chi_{\textup{d}})\right\}\right\} (7)

as the minimum distance of the goal to the transformed obstacles/workspace boundary. We revisit now the notion of the 22nd-order navigation function from [36]:

Definition 3

A 22nd-order navigation function is a function ϕ:0\phi:\mathcal{F}_{\mathcal{H}}\to\mathbb{R}_{\geq 0} of the form

φ(χ)k1χχd2+k2𝒥¯0β(d(χ)),\varphi(\chi)\coloneqq k_{1}\|\chi-\chi_{\textup{d}}\|^{2}+k_{2}\sum_{\ell\in\bar{\mathcal{J}}_{0}}\beta(d_{\ell}(\chi)),

where β:>00\beta:\mathbb{R}_{>0}\to\mathbb{R}_{\geq 0} is a (at least) twice contin. differentiable function and k1,k2k_{1},k_{2} are positive constants, with the followings properties:

  1. 1.

    β((0,τ])\beta((0,\tau]) is strictly decreasing, lim𝗑0β(𝗑)=\lim_{{\mathsf{x}}\to 0}{\beta}({\mathsf{x}})=\infty, and β(𝗑)=β(τ)\beta({\mathsf{x}})=\beta(\tau), 𝗑τ\forall{\mathsf{x}}\geq\tau, for some τ>0\tau>0,

  2. 2.

    φ(χ)\varphi(\chi) has a global minimum at χ=χdint()\chi=\chi_{\textup{d}}\in\textup{int}(\mathcal{F}_{\mathcal{H}}) where φ(χd)=0\varphi(\chi_{\textup{d}})=0,

  3. 3.

    if β(dk(χ))0\beta^{\prime}(d_{k}(\chi))\neq 0 and β′′(dk(χ))0\beta^{\prime\prime}(d_{k}(\chi))\neq 0 for some k𝒥¯0k\in\bar{\mathcal{J}}_{0}, then β(d(χ))=β′′(d(χ))=0\beta^{\prime}(d_{\ell}(\chi))=\beta^{\prime\prime}(d_{\ell}(\chi))=0, for all 𝒥¯0\{k}\ell\in\bar{\mathcal{J}}_{0}\backslash\{k\}.

  4. 4.

    The function β~:(0,τ)0\widetilde{\beta}:(0,\tau)\to\mathbb{R}_{\geq 0}, with β~(𝗑)β′′(𝗑)𝗑𝗑\widetilde{\beta}(\mathsf{x})\coloneqq\beta^{\prime\prime}(\mathsf{x})\mathsf{x}\sqrt{\mathsf{x}} is strictly decreasing.

An example for β\beta that satisfies properties 1) and 4), is

β(𝗑)\displaystyle\beta(\mathsf{x})\coloneqq {(6𝗑515𝗑4+10𝗑3)1,𝗑11,𝗑1,\displaystyle\begin{cases}\displaystyle(6\mathsf{x}^{5}-15\mathsf{x}^{4}+10\mathsf{x}^{3})^{-1},&\mathsf{x}\leq 1\\ 1,&\mathsf{x}\geq 1,\end{cases} (8)

By appropriately choosing τ\tau, only one β(d(χ))\beta(d_{\ell}(\chi)), 𝒥¯0\ell\in\bar{\mathcal{J}}_{0} affects the robotic agent for each χ\chi\in\mathcal{F}_{\mathcal{H}}, with β(d(χd))=β′′(d(χd))=0{\beta^{\prime}(d_{\ell}(\chi_{\textup{d}}))}={\beta^{\prime\prime}(d_{\ell}(\chi_{\textup{d}}))}=0. Hence, properties 2) and 3) of Def. 3 are satisfied.

Proposition 1 ([36])

By choosing τ\tau as τ(0,min{r¯2,r¯d})\tau\in(0,\min\{\bar{r}^{2},\bar{r}_{\textup{d}}\}), we guarantee that at each χ\chi\in\mathcal{F}_{\mathcal{H}} there is at most one 𝒥¯0\ell\in\bar{\mathcal{J}}_{0} such that d(χ)τd_{\ell}(\chi)\leq\tau, implying that β(d(χ))\beta^{\prime}(d_{\ell}(\chi)) and β′′(d(χ))\beta^{\prime\prime}(d_{\ell}(\chi)) are non-zero.

Intuitively, the obstacles and the workspace boundary have a local region of influence defined by the constant τ\tau.

To compensate for the unknown mass and friction terms of the dynamics (1), we define the estimates m^\hat{m}, α^\hat{\alpha} of mm and α\alpha (see Assumption 1), respectively, to be used in the control design.

Given the aforementioned definition, we design a reference signal vd:2v_{\textup{d}}:\mathcal{F}_{\mathcal{H}}\to\mathbb{R}^{2} for the robot velocity x˙\dot{x} as

vd(χ)=J(x)1(x)φ((x)),v_{\textup{d}}(\chi)=-J_{\mathcal{H}}(x)^{-1}\nabla_{\mathcal{H}(x)}\varphi(\mathcal{H}(x)), (9)

where J(x)(x)xJ_{\mathcal{H}}(x)\coloneqq\frac{\partial\mathcal{H}(x)}{\partial x} is the nonsingular Jacobian matrix of \mathcal{H}. Next, we define the respective velocity error evx˙vd(x)e_{v}\coloneqq\dot{x}-v_{\textup{d}}(x), and design the control law as u:×42u:\mathcal{F}\times\mathbb{R}^{4}\to\mathbb{R}^{2}, with

uu(x,v,m^,α^)kφJ(x)(x)φ((x))+m^v˙d(kv+32α^)ev,u\coloneqq u(x,v,\hat{m},\hat{\alpha})\coloneqq-k_{\varphi}J_{\mathcal{H}}(x)^{\top}\nabla_{\mathcal{H}(x)}\varphi(\mathcal{H}(x))+\hat{m}\dot{v}_{\text{d}}\\ -\left(k_{v}+\frac{3}{2}\hat{\alpha}\right)e_{v}, (10)

where kφk_{\varphi} is a positive gain, and m^\hat{m}, α^\hat{\alpha} evolve according to

m^˙\displaystyle\dot{\hat{m}}\coloneqq kmevv˙d\displaystyle-k_{m}e_{v}^{\top}\dot{v}_{\textup{d}} (11a)
α^˙\displaystyle\dot{\hat{\alpha}}\coloneqq kαev2,\displaystyle k_{\alpha}\|e_{v}\|^{2}, (11b)

with kmk_{m}, kαk_{\alpha} positive constants. The aforementioned control protocol guarantees safe asymptotic convergence of the robot to its goal from almost all collision-free initial conditions, i.e., except for a set of measure zero, provided that τ\tau is sufficiently small and kφ>α2k_{\varphi}>\frac{\alpha}{2} (see Theorem 2 in [36]). Therefore, since the convergence is asymptotic, we conclude that there exists a finite time instant ti>t0t_{i}>t_{0} such that ¯(xi(ti),ri)πik\bar{\mathcal{B}}(x_{i}(t_{i}),r_{i})\subset\pi_{i^{\prime}_{k}}, achieving thus the transition πikiπik\pi_{i_{k}}\to_{i}\pi_{i^{\prime}_{k}}.

2) Cooperative Object Transportation

We deal now with the control design for the cooperative object transportation problem. Consider an object jj\in\mathcal{M} grasped by a team of robots 𝒜\mathcal{A} at t00t_{0}\in\mathbb{R}_{\geq 0}, i.e., 𝒜𝒢j(t0)=𝒜\mathcal{AG}_{j}(t_{0})=\mathcal{A}, evolving subject to the dynamics (5) and satisfying ¯(xjo(t0),r𝒜,j)πjk\bar{\mathcal{B}}(x^{o}_{j}(t_{0}),r_{\mathcal{A},j})\subset\pi_{j_{k}}, for some jk𝒦j_{k}\in\mathcal{K}. The goal is the transportation of the object to some πjk\pi_{j_{k}^{\prime}}, jk𝒦j_{k}^{\prime}\in\mathcal{K}. As in the robot navigation case, we let xje=xjx^{e}_{j}=x_{j}, rje=r𝒜,jor^{e}_{j}=r^{o}_{\mathcal{A},j}, denoting the entity consisting of object jj and the robots 𝒜\mathcal{A}, viewing the rest of the entities as static obstacles. By also aiming to avoid entering the RoI πk\pi_{k}, for k𝒦\{jk,jk}k\in\mathcal{K}\backslash\{j_{k},j_{k}^{\prime}\}, the free space of the entity j=(𝒲f\𝒪j)r𝒜,jo\mathcal{F}_{j}=(\mathcal{W}_{\textup{f}}\backslash\mathcal{O}_{j})\ominus r^{o}_{\mathcal{A},j}, where

𝒪i(𝒯(𝒜𝒢)\{j}¯(xe,re))(k𝒦\{jk,jk}πk)\displaystyle\mathcal{O}_{i}\coloneqq\left(\bigcup_{\ell\in\mathcal{T}(\mathcal{AG})\backslash\{j\}}\bar{\mathcal{B}}(x^{e}_{\ell},r^{e}_{\ell})\right)\bigcup\left(\bigcup_{k\in\mathcal{K}\backslash\{j_{k},j_{k}^{\prime}\}}\pi_{k}\right)

Let the desired navigation goal of the object be xd,jx_{\textup{d},j}, provided by the procedure 𝒫s\mathcal{P}_{s}. Similarly to the robot navigation case, we use the transformation χ=(x)\chi=\mathcal{H}(x) to transform the environment to the open unit disk 𝒟\mathcal{D}, the obstacles to points bb_{\ell}, the object-robots entity to the point χ\chi, and the object goal to χd=(xd,j)\chi_{\textup{d}}=\mathcal{H}(x_{\textup{d},j}). Next, by employing the function β()\beta(), we design a reference signal vd:2v_{\textup{d}}:\mathcal{F}_{\mathcal{H}}\to\mathbb{R}^{2} for the object velocity that is identical to (9). In addition, we define the adaptation variables m^𝒜,j\hat{m}_{\mathcal{A},j} and α^𝒜,j\hat{\alpha}_{\mathcal{A},j} as the estimates of the unknown coupled mass and friction coefficients m𝒜,jm_{\mathcal{A},j} and α𝒜,j{\alpha}_{\mathcal{A},j} (see Assumption 1), respectively, and design the control law for the robots 𝒜\mathcal{A} as u:×42u:\mathcal{F}\times\mathbb{R}^{4}\to\mathbb{R}^{2}, with

u\displaystyle u_{\ell}\coloneqq u(x,v,m^𝒜,j,α^𝒜,j)\displaystyle u_{\ell}(x,v,\hat{m}_{\mathcal{A},j},\hat{\alpha}_{\mathcal{A},j})
\displaystyle\coloneqq 𝖼𝖿{kφJ(x)(x)φ((x))+m^𝒜,jv˙d\displaystyle-\mathsf{cf}_{\ell}\bigg{\{}k_{\varphi}J_{\mathcal{H}}(x)^{\top}\nabla_{\mathcal{H}(x)}\varphi(\mathcal{H}(x))+\hat{m}_{\mathcal{A},j}\dot{v}_{\text{d}}
(kv+32α^𝒜,j)ev},\displaystyle-\left(k_{v}+\frac{3}{2}\hat{\alpha}_{\mathcal{A},j}\right)e_{v}\bigg{\}},

where 𝖼𝖿\mathsf{cf}_{\ell} are load sharing coefficients satisfying 𝖼𝖿0\mathsf{cf}_{\ell}\geq 0, for all 𝒜\ell\in\mathcal{A}, and 𝒜𝖼𝖿=1\sum_{\ell\in\mathcal{A}}\mathsf{cf}_{\ell}=1; kφk_{\varphi} is a positive gain, and m^𝒜,j\hat{m}_{\mathcal{A},j}, α^𝒜,j\hat{\alpha}_{\mathcal{A},j} evolve according to (11).

By invoking the property 𝒜𝖼𝖿=1\sum_{\ell\in\mathcal{A}}\mathsf{cf}_{\ell}=1 and the fact that the object-robots system is converted to a point by the transformation \mathcal{H}, we guarantee, similar to the robot navigation case, the safe, asymptotic convergence of the object-robots entity to πjk\pi_{j^{\prime}_{k}} from almost all collision-free initial conditions, provided that kφ>α2k_{\varphi}>\frac{\alpha}{2}. Hence, there exists a finite time instant tj>t0t_{j}>t_{0} such that ¯(xj(tj),r𝒯,jo)πjk\bar{\mathcal{B}}(x_{j}(t_{j}),r^{o}_{\mathcal{T},j})\subset\pi_{j^{\prime}_{k}}, achieving thus the transition πik𝑇𝒜,jπjk\pi_{i_{k}}\xrightarrow{T}_{\mathcal{A},j}\pi_{j^{\prime}_{k}}.

III-B High-Level Plan Generation

The second part of the solution is the derivation of a high-level plan that satisfies the given LTL formulas ϕi\phi_{i} and ϕjo\phi^{o}_{j}. Thanks to (i) the proposed control laws that allow robot transitions and object transportations πkiπk\pi_{k}\rightarrow_{i}\pi_{k^{\prime}} and πk𝑇𝒜,jπk\pi_{k}\xrightarrow{T}_{\mathcal{A},j}\pi_{k^{\prime}}, respectively, and (ii) the off-the-self control laws that guarantee grasp and release actions i𝑔ji\xrightarrow{g}j and i𝑟ji\xrightarrow{r}j, we can abstract the behavior of the multi-robot-object system using a finite transition system as presented in the sequel.

Definition 4

The coupled behavior of the overall system of all the NN agents and MM objects is modeled by the transition system

𝒯𝒮=(Πs,Πsinit,s,𝒜𝒢,Ψ,,Λ,Ps,χ)\mathcal{TS}=(\Pi_{s},\Pi^{\text{init}}_{s},\rightarrow_{s},\mathcal{AG},\Psi,\mathcal{L},\Lambda,{P}_{s},\chi)

where

  1. (i)

    ΠsΠ¯×Π¯o×(2𝒩0)M\Pi_{s}\subset\bar{\Pi}\times\bar{\Pi}^{o}\times(2^{\mathcal{N}_{0}})^{M} is the set of states; Π¯Π1××ΠN\bar{\Pi}\coloneqq\Pi_{1}\times\cdots\times\Pi_{N} and Π¯oΠ1o××ΠMo\bar{\Pi}^{o}\coloneqq\Pi^{o}_{1}\times\cdots\times\Pi^{o}_{M} are the set of states-regions that the agents and the objects can be at, with Πi=Πjo=Π,i𝒩,j\Pi_{i}=\Pi^{o}_{j}=\Pi,\forall i\in\mathcal{N},j\in\mathcal{M};

    By defining π¯(πk1,,πkN),π¯o(πk1o,,πkMo)\bar{\pi}\coloneqq\left(\pi_{k_{1}},\cdots,\pi_{k_{N}}\right),\bar{\pi}_{o}\coloneqq(\pi_{\scriptscriptstyle k^{o}_{1}},\cdots,\pi_{\scriptscriptstyle k^{o}_{M}}), then the coupled state πs(π¯,π¯o,𝒜𝒢)\pi_{s}\coloneqq(\bar{\pi},\bar{\pi}_{o},\mathcal{AG}) belongs to Πs\Pi_{s}, i.e., (π¯,π¯o,𝒜𝒢)Πs(\bar{\pi},\bar{\pi}_{o},\mathcal{AG})\in\Pi_{s} if

    1. (a)

      𝒫s,0(rπk,[ri]i{i𝒩:ki=k},[rjo]j{j:kjo=k})=\mathcal{P}_{s,0}\Big{(}r_{\pi_{k}},[r_{i}]_{i\in\{i\in\mathcal{N}:k_{i}=k\}},[r^{o}_{j}]_{j\in\{j\in\mathcal{M}:k^{o}_{j}=k\}}\Big{)}=\top, i.e., the respective robots and objects fit in the region, k𝒦\forall k\in\mathcal{K},

    2. (b)

      ki=kjok_{i}=k^{o}_{j} for all i𝒩,ji\in\mathcal{N},j\in\mathcal{M} such that i𝒜𝒢ji\in\mathcal{AG}_{j}, i.e., a robot must be in the same region with the object it grasps,

  2. (ii)

    ΠsinitΠs\Pi^{\text{init}}_{s}\subset\Pi_{s} is the initial set of states at t=0t=0, which, owing to (i), satisfies the conditions of Problem 1,

  3. (iii)

    sΠs×Πs\rightarrow_{s}\subset\Pi_{s}\times\Pi_{s} is a transition relation defined as follows: given the states πs,π~sΠ\pi_{s},\widetilde{\pi}_{s}\in\Pi, with

    πs\displaystyle\pi_{s}\coloneqq (π¯,π¯o,𝒜𝒢)\displaystyle(\bar{\pi},\bar{\pi}_{o},\mathcal{AG})
    \displaystyle\coloneqq (πk1,,πkN,πk1o,,πkMo,𝒜𝒢1,,𝒜𝒢M),\displaystyle(\pi_{k_{1}},\dots,\pi_{k_{N}},\pi_{k^{o}_{1}},\dots,\pi_{k^{o}_{M}},\mathcal{AG}_{1},\dots,\mathcal{AG}_{M}),
    π~s\displaystyle\widetilde{\pi}_{s}\coloneqq (π¯~,π¯~o,𝒜𝒢~)\displaystyle(\widetilde{\bar{\pi}},{\widetilde{\bar{\pi}}}_{o},\widetilde{\mathcal{AG}})
    \displaystyle\coloneqq (πk~1,,πk~N,πk~1o,,πk~1o,𝒜𝒢~1,,𝒜𝒢~M),\displaystyle(\pi_{\widetilde{k}_{1}},\dots,\pi_{\widetilde{k}_{N}},\pi_{\widetilde{k}^{o}_{1}},\dots,\pi_{\widetilde{k}^{o}_{1}},\widetilde{\mathcal{AG}}_{1},\dots,\widetilde{\mathcal{AG}}_{M}), (12)

    a transition πssπs\pi_{s}\rightarrow_{s}{\pi}^{\prime}_{s} occurs if all the following hold:

    1. (a)

      i𝒩,j\nexists i\in\mathcal{N},j\in\mathcal{M} such that i𝒜𝒢ji\in\mathcal{AG}_{j}, i𝒜𝒢~ji\notin\widetilde{\mathcal{AG}}_{j} (or i𝒜𝒢ji\notin\mathcal{AG}_{j}, i𝒜𝒢~ji\in\widetilde{\mathcal{AG}}_{j}), and kik~ik_{i}\neq\widetilde{k}_{i}, i.e., there are no simultaneous grasp/release and navigation actions,

    2. (b)

      i𝒩,j\nexists i\in\mathcal{N},j\in\mathcal{M} such that i𝒜𝒢ji\in\mathcal{AG}_{j}, i𝒜𝒢~ji\notin\widetilde{\mathcal{AG}}_{j} (or i𝒜𝒢ji\notin\mathcal{AG}_{j}, i𝒜𝒢~ji\in\widetilde{\mathcal{AG}}_{j}), and ki=kjok~i=k~jok_{i}=k^{o}_{j}\neq\widetilde{k}_{i}=\widetilde{k}^{o}_{j}, i.e., there are no simultaneous grasp/release and transportation actions,

    3. (c)

      i𝒩,j,j\nexists i\in\mathcal{N},j,j^{\prime}\in\mathcal{M}, with jjj\neq j^{\prime}, such that i𝒜𝒢ji\in\mathcal{AG}_{j} and i𝒜𝒢~ji\in\widetilde{\mathcal{AG}}_{j^{\prime}}, i.e., there are no simultaneous grasp and release actions,

    4. (d)

      j\nexists j\in\mathcal{M} such that kjok~jok^{o}_{j}\neq\widetilde{k}^{o}_{j} and i𝒜𝒢j,i𝒩i\notin\mathcal{AG}_{j},\forall i\in\mathcal{N} ( or i𝒜𝒢~j,i𝒩i\notin\widetilde{\mathcal{AG}}_{j},\forall i\in\mathcal{N}), i.e., there is no transportation of a non-grasped object,

    5. (e)

      j,𝒯𝒩\nexists j\in\mathcal{M},\mathcal{T}\subseteq\mathcal{N} such that kjok~jok^{o}_{j}\neq\widetilde{k}^{o}_{j} and Λ(mjo,ζ𝒯)=\Lambda(m^{o}_{j},\zeta_{\mathcal{T}})=\bot, with (i𝒜𝒢j,i𝒜𝒢~j)i𝒯(i\in\mathcal{AG}_{j},i\in\widetilde{\mathcal{AG}}_{j})\Leftrightarrow i\in\mathcal{T}, i.e., the agents grasping an object are powerful enough to transfer it,

  4. (iv)

    ΨΨ¯Ψ¯o\Psi\coloneqq\bar{\Psi}\cup\bar{\Psi}^{o} with Ψ¯=i𝒩Ψi\bar{\Psi}=\bigcup_{i\in\mathcal{N}}\Psi_{i} and Ψ¯o=jΨjo\bar{\Psi}^{o}=\bigcup_{j\in\mathcal{M}}\Psi^{o}_{j}, are the atomic propositions of the agents and objects, respectively, as defined in Section II.

  5. (v)

    :Πs2Ψ\mathcal{L}:\Pi_{s}\rightarrow 2^{\Psi} is a labeling function defined as follows: Given a state πs\pi_{s} as in (12) and ψs(i𝒩ψi)(jψjo)\psi_{s}\coloneqq\Big{(}\bigcup_{i\in\mathcal{N}}\psi_{i}\Big{)}\bigcup\Big{(}\bigcup_{j\in\mathcal{M}}\psi^{o}_{j}\Big{)} with ψi2Ψi,ψjo2Ψjo\psi_{i}\in 2^{\Psi_{i}},\psi^{o}_{j}\in 2^{\Psi^{o}_{j}}, then ψs(πs)\psi_{s}\in\mathcal{L}(\pi_{s}) if ψii(πki)\psi_{i}\in\mathcal{L}_{i}(\pi_{k_{i}}) and ψjojo(πkjo),i𝒩,j\psi^{o}_{j}\in\mathcal{L}^{o}_{j}(\pi_{k^{o}_{j}}),\forall i\in\mathcal{N},j\in\mathcal{M}.

  6. (vi)

    Λ\Lambda and Ps{P}_{s} as defined in Section II.

  7. (vii)

    χ:(s)0\chi:(\to_{s})\to\mathbb{R}_{\geq 0} is a function that assigns a cost to each transition πssπ~s\pi_{s}\to_{s}\widetilde{\pi}_{s}. This cost might be related to the distance of the robots’ regions in πs\pi_{s} to the ones in π~s\widetilde{\pi}_{s}, combined with the cost efficiency of the robots involved in transport tasks (according to ζi,i𝒩\zeta_{i},i\in\mathcal{N}).

Next, we form the global LTL formula ϕ(i𝒩ϕi)(jϕjo)\phi\coloneqq(\land_{i\in\mathcal{N}}\phi_{i})\land(\land_{j\in\mathcal{M}}\phi^{o}_{j}) over the set Ψ\Psi. Givenϕ\phi, we define the language Words(ϕ)={σ(2Ψ)ω|σϕ}\texttt{Words}(\phi)=\left\{\sigma\in(2^{\Psi})^{\omega}|\sigma\models\phi\right\}, where (2Ψ)ω×ϕ\models\subseteq(2^{\Psi})^{\omega}\times\phi is the satisfaction relation, as the set of infinite words σ(2Ψ)ω\sigma\in(2^{\Psi})^{\omega} that satisfy the LTL formula ϕ\phi. Then, we translate ϕ\phi to a Büchi Automaton 𝒜\mathcal{BA} defined over (2Ψ)ω(2^{\Psi})^{\omega}. The Büchi Automaton is defined as follows [31]:

Definition 5 (NBA)

A Nondeterministic Bu¨\ddot{\text{u}}chi Automaton (NBA) 𝒜\mathcal{BA} is defined as a tuple B=(ΠB,ΠB0,Σ,B,ΠBF)B=\left(\Pi_{B},\Pi_{B}^{0},\Sigma,\rightarrow_{B},\Pi_{B}^{F}\right), where ΠB\Pi_{B} is the set of states, ΠB0ΠB\Pi_{B}^{0}\subseteq\Pi_{B} is a set of initial states, Σ=2Ψ\Sigma=2^{\Psi} is the alphabet, BΠB×Ψ×ΠB\rightarrow_{B}\subseteq\Pi_{B}\times\Psi\times\Pi_{B} is the transition relation, and ΠBFΠB\Pi_{B}^{F}\subseteq\Pi_{B} is a set of accepting/final states.

Given the 𝒯𝒮\mathcal{TS} and the NBA 𝒜\mathcal{BA} that corresponds to the LTL ϕ\phi, we can now define the Product Bu¨\ddot{\text{u}}chi Automaton (PBA) 𝒫𝒜𝒯𝒮𝒜\mathcal{PBA}\coloneqq\mathcal{TS}\otimes\mathcal{BA}, as follows [31]:

Definition 6 (PBA)

Given the product transition system 𝒯𝒮=(Πs,Πsinit,s,𝒜𝒢,Ψ,,Λ,Ps,χ)\mathcal{TS}=(\Pi_{s},\Pi^{\text{init}}_{s},\rightarrow_{s},\mathcal{AG},\Psi,\mathcal{L},\Lambda,{P}_{s},\chi) and the NBA B=(ΠB,ΠB0,Ψ,B,ΠBF)B=\left(\Pi_{B},\Pi_{B}^{0},\Psi,\rightarrow_{B},\Pi_{B}^{F}\right), we can define the Product Bu¨\ddot{\text{u}}chi Automaton 𝒫𝒜𝒫𝒯𝒮𝒜\mathcal{PBA}\coloneqq\mathcal{PTS}\otimes\mathcal{BA} as a tuple 𝒫𝒜=(ΠP,ΠP0,P,ΠPF)\mathcal{PBA}=(\Pi_{P},\Pi_{P}^{0},\longrightarrow_{P},\Pi_{P}^{F}) where (a) ΠP=Πs×ΠB\Pi_{P}=\Pi_{\text{s}}\times\Pi_{B} is the set of states; (b) ΠP0=Πsinit×ΠB0\Pi_{P}^{0}=\Pi_{s}^{\text{init}}\times\Pi_{B}^{0} is a set of initial states; (c) PΠP×Σ×ΠP\longrightarrow_{P}\subseteq\Pi_{P}\times\Sigma\times\Pi_{P} is the transition relation defined by the rule: (πssπs)(πB(πs)πB)πP=(πs,πB)PπP=(πs,πB)\frac{(\pi_{s}\rightarrow_{s}\pi_{s}^{\prime})\wedge(\pi_{B}\xrightarrow{\mathcal{L}\left(\pi_{s}\right)}\pi_{B}^{\prime})}{\pi_{P}=\left(\pi_{s},\pi_{B}\right)\longrightarrow_{P}\pi_{P}^{\prime}=\left(\pi_{s}^{\prime},\pi_{B}^{\prime}\right)}. Transition from state πPΠP\pi_{P}\in\Pi_{P} to πPΠP\pi_{P}^{\prime}\in\Pi_{P}, is denoted by (πP,πP)P(\pi_{P},\pi_{P}^{\prime})\in\longrightarrow_{P}, or πPPπP\pi_{P}\longrightarrow_{P}\pi_{P}^{\prime}; (d) χP(πP,πP)=χ(πs,πs)\chi_{P}(\pi_{\text{P}},\pi_{\text{P}}^{\prime})=\chi(\pi_{s},\pi_{s}^{\prime}), where πP=(πs,πB)\pi_{P}=(\pi_{s},\pi_{B}) and πP=(πs,πB)\pi_{\text{P}}^{\prime}=(\pi_{s}^{\prime},\pi_{B}^{\prime}); and (e) ΠPF=Πs×ΠBF\Pi_{P}^{F}=\Pi_{s}\times\Pi_{B}^{F} is a set of accepting/final states.

Given ϕ\phi and the PBA, an infinite path πplπs,1πs,2\pi_{\text{pl}}\coloneqq\pi_{s,1}\pi_{s,2}\dots of a 𝒯𝒮\mathcal{TS} satisfies ϕ\phi if and only if trace(πpl)Words(ϕ)\texttt{trace}(\pi_{\text{pl}})\in\texttt{Words}(\phi), which is equivalently denoted by πplϕ\pi_{\text{pl}}\models\phi, where trace(πpl)(2Ψ)ω\texttt{trace}(\pi_{\text{pl}})\in\left(2^{\Psi}\right)^{\omega} is defined as trace(πpl)=(πs,1)(πs,2)\texttt{trace}(\pi_{\text{pl}})=\mathcal{L}(\pi_{s,1})\mathcal{L}(\pi_{s,2})\dots. Specifically, if there is a path satisfying ϕ\phi, then there exists a path πplϕ\pi_{\text{pl}}\models\phi that can be written in a finite representation, called prefix-suffix structure, i.e., πpl=πplpre[πplsuf]ω\pi_{\text{pl}}=\pi_{\text{pl}}^{\text{pre}}[\pi_{\text{pl}}^{\text{suf}}]^{\omega}, where the prefix part πplpre\pi_{\text{pl}}^{\text{pre}} is executed only once followed by the indefinite execution of the suffix part πplsuf\pi_{\text{pl}}^{\text{suf}}. The prefix part πplpre\pi_{\text{pl}}^{\text{pre}} is the projection of a finite path pprep^{\text{pre}} that lives in ΠP\Pi_{P} onto Πs\Pi_{s}.

Computing a plan πpl\pi_{\text{pl}} is typically accomplished by applying graph-search methods to the PBA. Specifically, to generate a motion plan πpl\pi_{\text{pl}} that satisfies ϕ\phi, the PBA is viewed as a weighted directed graph 𝒢P={𝒱P,P,wP}\mathcal{G}_{P}=\{\mathcal{V}_{P},\mathcal{E}_{P},w_{P}\}, where the set of nodes 𝒱P\mathcal{V}_{P} is indexed by the set of states ΠP\Pi_{P}, the set of edges P\mathcal{E}_{P} is determined by the transition relation P\longrightarrow_{P}, and the weights assigned to each edge are determined by the function χP\chi_{P}. Then, to find the optimal plan τϕ\tau\models\phi, shortest paths towards final states and shortest cycles around them are computed. More details about this approach can be found in [38, 39] and the references therein. While any of the aforementioned methodologies could be used, in this work we employ STyLuS\text{STyLuS}^{*}, an algorithm that is designed to solve complex temporal planning problems in large-scale multi-robot systems and has been shown to achieve significantly lower complexity, in terms of running time and memory, than standard graph-search methods [40]. Specifically, STyLuS\text{STyLuS}^{*} is a sampling-based method that builds incrementally trees that approximate the state-space and transitions of the product automaton and does not require sophisticated graph-search techniques. Technically, STyLuS\text{STyLuS}^{*} builds a tree 𝒢T={𝒱T,T,Cost}\mathcal{G}_{T}=\{\mathcal{V}_{T},\mathcal{E}_{T},\texttt{Cost}\} first, where 𝒱TΠP\mathcal{V}_{T}\subseteq\Pi_{P} is the set of nodes, T\mathcal{E}_{T} is the set of edges, and Cost is defined as per χP\chi_{P}, determining the cost of reaching a tree node from its root. This tree is rooted at the initial state πP0\pi_{P}^{0} and is used for the synthesis of the prefix part. The tree is constructed incrementally, in a sampling-based fashion, and its construction terminates after a user-specified number of iterations nmaxn_{\text{max}}. Then, we compute paths in the constructed tree structure that connect the root to the detected, if any, final states. These paths correspond to prefix parts of candidate feasible paths. To construct the corresponding suffix paths, new trees are built, similarly, rooted at the previously detected final states aiming to compute cycles around the tree roots. Among all the detected prefix-suffix paths, STyLuS\text{STyLuS}^{*} returns the one with the minimum cost. As it was shown in [40], STyLuS\text{STyLuS}^{*} is probabilistically complete and asymptotically optimal; that is, the probability of finding a feasible and the optimal solution converges to 11 as nmaxn_{\text{max}}\to\infty.

Finally, note that the constructed trees explore the state space of the product automaton and, therefore, the designed prefix-suffix path is defined as an infinite sequence of product automaton states. By projecting it onto the state-space of the transition system 𝒯𝒮\mathcal{TS}, we obtain a high-level prefix-suffix plan defined as a sequence of states πplπs,1πs,2ϕ\pi_{\text{pl}}\coloneqq\pi_{s,1}\pi_{s,2}\dots\models\phi. The corresponding sequence of atomic propositions is ψpl=trace(πpl)=ψs,1ψs,2\psi_{\text{pl}}=\texttt{trace}(\pi_{\text{pl}})=\psi_{s,1}\psi_{s,2}\dots, with

πs,(π¯,π¯o,,w¯)Πs,,\displaystyle\pi_{s,\ell}\coloneqq(\bar{\pi}_{\ell},\bar{\pi}_{o,\ell},\bar{w}_{\ell})\in\Pi_{s},\forall\ell\in\mathbb{N},
ψs,(i𝒩ψi,)(jψj,o)2Ψ,(πs,),,\displaystyle\psi_{s,\ell}\coloneqq\Big{(}\bigcup\limits_{i\in\mathcal{N}}\psi_{i,\ell}\Big{)}\bigcup\Big{(}\bigcup\limits_{j\in\mathcal{M}}\psi^{o}_{j,\ell}\Big{)}\in 2^{\Psi},\mathcal{L}(\pi_{s,\ell}),\forall\ell\in\mathbb{N},

where

  • π¯πk1,,πk2,,\bar{\pi}_{\ell}\coloneqq\pi_{k_{1,\ell}},\pi_{k_{2,\ell}},\dots with ki,𝒦,i𝒩k_{i,\ell}\in\mathcal{K},\forall i\in\mathcal{N},

  • π¯o,πk1,o,πk2,o,\bar{\pi}_{o,\ell}\coloneqq\pi_{k^{o}_{1,\ell}},\pi_{k^{o}_{2,\ell}},\dots with kj,o𝒦,jk^{o}_{j,\ell}\in\mathcal{K},\forall j\in\mathcal{M},

  • w¯w1,,w2,,\bar{w}_{\ell}\coloneqq w_{1,\ell},w_{2,\ell},\dots with wi,𝒜𝒢i,i𝒩w_{i,\ell}\in\mathcal{AG}_{i},\forall i\in\mathcal{N},

  • ψi,2Ψi,i(πki,),i𝒩\psi_{i,\ell}\in 2^{\Psi_{i}},\mathcal{L}_{i}(\pi_{k_{i,\ell}}),\forall i\in\mathcal{N},

  • ψj,o2Ψjo,jo(πkj,o),j\psi^{o}_{j,\ell}\in 2^{\Psi^{o}_{j}},\mathcal{L}^{o}_{j}(\pi_{k^{o}_{j,\ell}}),\forall j\in\mathcal{M}.

The path πpl\pi_{\text{pl}} is then projected to the individual sequences of the regions πkj,1oπkj,2o\pi_{k^{o}_{j,1}}\pi_{k^{o}_{j,2}}\dots for each object jj\in\mathcal{M}, as well as to the individual sequences of the regions πki,1πki,2\pi_{k_{i,1}}\pi_{k_{i,2}}\dots and the boolean grasping variables wi,1wi,2w_{i,1}w_{i,2}\dots for each robot i𝒩i\in\mathcal{N}. The aforementioned sequences determine the behavior of robot i𝒩i\in\mathcal{N}, i.e., the sequence of actions (transition, transportation, grasp, release or stay idle) it must take.

By the definition of \mathcal{L} in Def. 4, we obtain that ψi,i(πki,),ψj,ojo(πkj,o),i𝒩,j,\psi_{i,\ell}\in\mathcal{L}_{i}(\pi_{k_{i,\ell}}),\psi^{o}_{j,\ell}\in\mathcal{L}^{o}_{j}(\pi_{k^{o}_{j,\ell}}),\forall i\in\mathcal{N},j\in\mathcal{M},\ell\in\mathbb{N}. Therefore, since ϕ=(i𝒩ϕi)(jϕoj)\phi=(\land_{i\in\mathcal{N}}\phi_{i})\land(\land_{j\in\mathcal{M}}\phi_{o_{j}}) is satisfied by ψ\psi, we conclude that ψi,1ψi,2ϕi\psi_{i,1}\psi_{i,2}\dots\models\phi_{i} and ψj,1oψj,2oϕjo,i𝒩,j\psi^{o}_{j,1}\psi^{o}_{j,2}\dots\models\phi^{o}_{j},\forall i\in\mathcal{N},j\in\mathcal{M}.

The sequences πki,1πki,2\pi_{k_{i,1}}\pi_{k_{i,2}}\dots, ψi,1ψi,2\psi_{i,1}\psi_{i,2}\dots and πkj,1oπkj,2o,ψj,1oψj,2o\pi_{k^{o}_{j,1}}\pi_{k^{o}_{j,2}}\dots,\psi^{o}_{j,1}\psi^{o}_{j,2}\dots over Π,2Ψi\Pi,2^{\Psi_{i}} and Π,2Ψjo\Pi,2^{\Psi^{o}_{j}}, respectively, produce the trajectories qi(t){q}_{i}(t) and xjo(t),i𝒩,j{x}^{o}_{j}(t),\forall i\in\mathcal{N},j\in\mathcal{M}. The corresponding behaviors are βi=(qi(t),σi)=(qi(ti,1),σi,1)(qi(ti,2),σi,2)\beta_{i}=({q}_{i}(t),\sigma_{i})=({q}_{i}(t_{i,1}),\sigma_{i,1})({q}_{i}(t_{i,2}),\sigma_{i,2})\dots and βjo\beta^{o}_{j} == (xjo(t),σjo)=(xjo(tj,1o),σj,1o)(xjo(tj,2o),σj,2o)({x}^{o}_{j}(t),\sigma^{o}_{j})=({x}^{o}_{j}(t^{o}_{j,1}),\sigma^{o}_{j,1})({x}^{o}_{j}(t^{o}_{j,2}),\sigma^{o}_{j,2})\dots, respectively, according to Section II-B, with 𝒜i(qi(ti,))πki,,σi,i(πki,)\mathcal{A}_{i}({q}_{i}(t_{i,\ell}))\subset\pi_{k_{i,\ell}},\sigma_{i,\ell}\in\mathcal{L}_{i}(\pi_{k_{i,\ell}}) and 𝒪j(xoj(toj,m))πkj,o,σj,ojo(πkj,o)\mathcal{O}_{j}({x}_{o_{j}}(t_{o_{j,m}}))\in\pi_{k^{o}_{j,\ell}},\sigma^{o}_{j,\ell}\in\mathcal{L}^{o}_{j}(\pi_{k^{o}_{j,\ell}}). Thus, it is guaranteed that σiϕi,σjoϕjo\sigma_{i}\models\phi_{i},\sigma^{o}_{j}\models\phi^{o}_{j} and consequently, the behaviors βi\beta_{i} and βjo\beta^{o}_{j} satisfy the formulas ϕi\phi_{i} and ϕjo\phi^{o}_{j}, respectively, i𝒩,j\forall i\in\mathcal{N},j\in\mathcal{M}. The aforementioned reasoning is summarized in the next theorem:

Theorem 1

The execution of the path (πpl,ψpl)(\pi_{\text{pl}},\psi_{\text{pl}}) of 𝒯𝒮\mathcal{TS} guarantees behaviors βi,βjo\beta_{i},\beta^{o}_{j} that yield the satisfaction of ϕi\phi_{i} and ϕjo\phi^{o}_{j}, respectively, i𝒩,j\forall i\in\mathcal{N},j\in\mathcal{M}, providing, therefore, a solution to Problem 1.

IV Simulations

In this section, we provide two case studies in an obstacle-cluttered office environment. We choose the atomic propositions for the robots and objects as Ψi={i-π1,,i-πK}\Psi_{i}=\{``i\text{-}\pi_{1}",\dots,``i\text{-}\pi_{K}"\} and Ψjo={Oj-π1,,Oj-πK}\Psi^{o}_{j}=\{``O_{j}\text{-}\pi_{1}",\dots,``O_{j}\text{-}\pi_{K}"\}, respectively, for i𝒩i\in\mathcal{N}, jj\in\mathcal{M}, indicating their presence in the regions of interest. For the constructed transition systems, we set the cost χ\chi as the Euclidean distance among the RoI contained in the nodes of the transitions.

For the continuous control design, we choose the robot dynamics of the form (1) with mass mi=1m_{i}=1, fi()=54sin(0.5(xi1+xi2))F(x˙i)x˙if_{i}(\cdot)=\frac{5}{4}\sin(0.5(x_{i_{1}}+x_{i_{2}}))F(\dot{x}_{i})\dot{x}_{i}, with F(x˙i)=diag{[exp(sgn(x˙i)x˙i)+1]i{1,2}}F(\dot{x}_{i})=\textup{diag}\{[\exp(-\text{sgn}(\dot{x}_{i})\dot{x}_{i})+1]_{i\in\{1,2\}}\}, where we denote (xi1,xi2)=xi(x_{i_{1}},x_{i_{2}})=x_{i}, (x˙i1,x˙i2)=x˙i(\dot{x}_{i_{1}},\dot{x}_{i_{2}})=\dot{x}_{i}, i𝒩i\in\mathcal{N}. We further choose r¯=0.1\bar{r}=0.1 and a variation of (8) for β\beta with τ=r¯2\tau=\bar{r}^{2}. Finally, we set the control gains to k1=0.01k_{1}=0.01, k2=5k_{2}=5, kϕ=1k_{\phi}=1, and km=kα=0.01k_{m}=k_{\alpha}=0.01.

IV-A Case Study I:

In this case study, we consider N=3N=3 robots, K=4K=4 regions and M=2M=2 objects. The regions of interest are circular centered at (88,280)(88,-280)m, (100,160)(100,-160)m, (200,130)(200,-130)m, and (250,285)(250,-285)m, with radius equal to 44m. The robots’ and object’ mass is taken as 11kg and 0.250.25kg, respectively, while their spherical volume’s radii as 0.750.75m and 0.20.2m, respectively. The power capabilities of the robots are 2,3,42,3,4, respectively, and the power required for each object is 5,65,6, respectively. Initially, the robots are located in regions πinit(1)=π1\pi_{init(1)}=\pi_{1}, πinit(2)=π3\pi_{init(2)}=\pi_{3}, and πinit(3)=π4\pi_{init(3)}=\pi_{4}, respectively, whereas the objects are located in regions πinito(1)=π2\pi_{init_{o}(1)}=\pi_{2}, πinito(2)=π1\pi_{init_{o}(2)}=\pi_{1}, respectively. Fig. 1 depicts the considered environment and the initial configuration of the multi-robot-object system.

Refer to caption
Figure 1: The multi-robot-object initial configuration in the first case study.
Refer to caption
Refer to caption
Figure 2: Navigation of robot 1 from π1\pi_{1} to π3\pi_{3} in the original (left) and the transformed point-world (right) environment. In the original environment, the robot’s circular volume has been transferred to the obstacles and workspace boundary.
Refer to caption
Refer to caption
Figure 3: The evolution of the control inputs u1(t)u_{1}(t) (left) and adaptation signals m^1(t)\hat{m}_{1}(t), α^1(t)\hat{\alpha}_{1}(t) (right) of robot’s 1 navigation.
Refer to caption
Refer to caption
Figure 4: Transportation of object 1 from π2\pi_{2} to π4\pi_{4} by robots 1 and 2 in the original (left) and the transformed point-world (right) environment. In the original environment, the circular volume of the coupled object-robots system has been transferred to the obstacles and workspace boundary.
Refer to caption
Refer to caption
Figure 5: The evolution of the control inputs u1(t)u_{1}(t) (left) and adaptation signals m^1(t)\hat{m}_{1}(t), α^1(t)\hat{\alpha}_{1}(t) (right) of robot 1 in the cooperative object transportation.

The resulting transition system consists of 3,1123,112 reachable states and 154,960154,960 transitions and it was created within 1.711.71 minutes. We set the following formula for the multi-robot-object system:

ϕ=ψ¯1ψ¯2ψ¯2ψ¯4(¬ψ¯4𝒰ψ¯5)\phi=\square\Diamond\bar{\psi}_{1}\wedge\square\Diamond\bar{\psi}_{2}\wedge\square\Diamond\bar{\psi}_{2}\wedge\Diamond\bar{\psi}_{4}\wedge(\neg\bar{\psi}_{4}\mathcal{U}\bar{\psi}_{5}) (13)

with

  • ψ¯1={1-π1,2-π1,3-π1,O1-π1,O2-π4}\bar{\psi}_{1}=\{``1\text{-}\pi_{1}",``2\text{-}\pi_{1}",``3\text{-}\pi_{1}",``O_{1}\text{-}\pi_{1}",``O_{2}\text{-}\pi_{4}"\}

  • ψ¯2={1-π1,2-π3,3-π1,O1-π1,O2-π3}\bar{\psi}_{2}=\{``1\text{-}\pi_{1}",``2\text{-}\pi_{3}",``3\text{-}\pi_{1}",``O_{1}\text{-}\pi_{1}",``O_{2}\text{-}\pi_{3}"\}

  • ψ¯3={1-π3,2-π3,3-π3,O1-π4,O2-π3}\bar{\psi}_{3}=\{``1\text{-}\pi_{3}",``2\text{-}\pi_{3}",``3\text{-}\pi_{3}",``O_{1}\text{-}\pi_{4}",``O_{2}\text{-}\pi_{3}"\}

  • ψ¯4={1-π2,2-π4,3-π1,O1-π4,O2-π2}\bar{\psi}_{4}=\{``1\text{-}\pi_{2}",``2\text{-}\pi_{4}",``3\text{-}\pi_{1}",``O_{1}\text{-}\pi_{4}",``O_{2}\text{-}\pi_{2}"\}

  • ψ¯5={1-π4,2-π4,3-π2,O1-π3,O2-π4}\bar{\psi}_{5}=\{``1\text{-}\pi_{4}",``2\text{-}\pi_{4}",``3\text{-}\pi_{2}",``O_{1}\text{-}\pi_{3}",``O_{2}\text{-}\pi_{4}"\}

In words, the mission specification in (13) requires the robots and objects to satisfy ψ¯1\bar{\psi}_{1}, ψ¯2\bar{\psi}_{2}, and ψ¯3\bar{\psi}_{3} infinitely often, satisfy ψ¯4\bar{\psi}_{4} eventually, and satisfy ψ¯5\bar{\psi}_{5} before satisfying ψ¯4\bar{\psi}_{4}. The LTL formula in (13) corresponds to a NBA with 66 states - among which one is a final state - and 1818 transitions. STyLuS* found the first feasible prefix and suffix path within 1.231.23 minutes and 0.640.64 minutes on average. The action path of the robots is depicted in Table I, starting from πs,1\pi_{s,1} and satisfying {1-π1,2-π2,3-π4,O1-π2,O2-π1}\{``1\text{-}\pi_{1}",``2\text{-}\pi_{2}",``3\text{-}\pi_{4}",``O_{1}\text{-}\pi_{2}",``O_{2}\text{-}\pi_{1}"\}.

We further illustrate the continuous control design. In particular, we consider the actions of robot navigation and object transportation. Firstly, we examine the navigation of robot 1 from π1\pi_{1} to π3\pi_{3}. The results are depicted in Figs. 2, 3. The left part of Fig. 2 shows the trajectory of robot 11 in the environment, where the obstacles and boundary have absorbed the the spherical volume of the robot; the right part of Fig. 2 shows the trajectory of robot 11 in the transformed point world, where the obstacles are represented by points. Fig. 3 depicts the control input u1(t)u_{1}(t) (left) and the evolution of the adaptation signals m^1(t)\hat{m}_{1}(t), α^1(t)\hat{\alpha}_{1}(t) (right).

We further examine the transportation of object 11 by robots 11 and 22 from π2\pi_{2} to π4\pi_{4}, where we chose the load-sharing coefficients 𝖼𝖿1=𝖼𝖿2=0.5\mathsf{cf}_{1}=\mathsf{cf}_{2}=0.5. The results are depicted in Figs. 4, 5. The left part of Fig. 4 shows the trajectory of the coupled object-robots system in the environment, where the obstacles and boundary have absorbed the its spherical volume; the right part of Fig. 2 shows the trajectory of the coupled system in the transformed point world, where the obstacles are represented by points. Fig. 3 depicts the control input u1(t)u_{1}(t) (left) and the evolution of the adaptation signals m^1(t)\hat{m}_{1}(t), α^1(t)\hat{\alpha}_{1}(t) (right) of the robot 1, which are identical to the ones of robot 2.

TABLE I: The agent actions for the discrete path of the first case study
πs,\pi_{s,\ell} Actions πs,\pi_{s,\ell} Actions
πs,1\pi_{s,1} - πs,35\pi_{s,35} π21π1\pi_{2}\to_{1}\pi_{1}, 2𝑟22\xrightarrow{r}2, 3𝑟23\xrightarrow{r}2
πs,2\pi_{s,2} 1𝑔21\xrightarrow{g}2, π43π1\pi_{4}\to_{3}\pi_{1} πs,36\pi_{s,36} 1𝑔11\xrightarrow{g}1, π331\pi_{3}\to_{3}1
πs,3\pi_{s,3} π32π2\pi_{3}\to_{2}\pi_{2} 3𝑔23\xrightarrow{g}2 πs,37\pi_{s,37} π32π2\pi_{3}\to_{2}\pi_{2}, 3𝑔13\xrightarrow{g}1
πs,4\pi_{s,4} π1𝑇{1,3},2π4\pi_{1}\xrightarrow{T}_{\{1,3\},2}\pi_{4}, 2𝑔12\xrightarrow{g}1 πs,38\pi_{s,38} π1𝑇{1,3},1π4\pi_{1}\xrightarrow{T}_{\{1,3\},1}\pi_{4}
πs,5\pi_{s,5} 1𝑟21\xrightarrow{r}2, 2𝑟12\xrightarrow{r}1 πs,39\pi_{s,39} π22π3\pi_{2}\to_{2}\pi_{3}, 1𝑟11\xrightarrow{r}1, 3𝑟13\xrightarrow{r}1
πs,6\pi_{s,6} π41π2\pi_{4}\rightarrow_{1}\pi_{2}, 2𝑔12\xrightarrow{g}1, 3𝑟23\xrightarrow{r}2 πs,40\pi_{s,40} π43π3\pi_{4}\to_{3}\pi_{3}
πs,7\pi_{s,7} 1𝑔11\xrightarrow{g}1 πs,41\pi_{s,41} π41π3\pi_{4}\to_{1}\pi_{3}, 3𝑔23\xrightarrow{g}2
πs,8\pi_{s,8} π2𝑇{1,2},1π3\pi_{2}\xrightarrow{T}_{\{1,2\},1}\pi_{3} πs,42\pi^{\star}_{s,42} π32π4\pi_{3}\to_{2}\pi_{4}, 3𝑟23\xrightarrow{r}2
πs,9\pi_{s,9} 1𝑟11\xrightarrow{r}1, π43π2\pi_{4}\to_{3}\pi_{2} πs,43\pi^{\star}_{s,43} π33π4\pi_{3}\to_{3}\pi_{4}
πs,10\pi_{s,10} π31π4\pi_{3}\to_{1}\pi_{4}, 2𝑟12\xrightarrow{r}1 πs,44\pi^{\star}_{s,44} 2𝑔12\xrightarrow{g}1, 3𝑔13\xrightarrow{g}1
πs,11\pi_{s,11} π32π4\pi_{3}\to_{2}\pi_{4} πs,45\pi^{\star}_{s,45} π4𝑇{2,3},1π1\pi_{4}\xrightarrow{T}_{\{2,3\},1}\pi_{1}
πs,12\pi_{s,12} 1𝑔21\xrightarrow{g}2, π23π4\pi_{2}\to_{3}\pi_{4} πs,46\pi^{\star}_{s,46} 2𝑟12\xrightarrow{r}1, 2𝑟32\xrightarrow{r}3
πs,13\pi_{s,13} π42π3\pi_{4}\to_{2}\pi_{3}, 3𝑔23\xrightarrow{g}2 πs,47\pi^{\star}_{s,47} π13π3\pi_{1}\to_{3}\pi_{3}
πs,14\pi_{s,14} π4𝑇{1,3},2π2\pi_{4}\xrightarrow{T}_{\{1,3\},2}\pi_{2} πs,48\pi^{\star}_{s,48} 1𝑔21\xrightarrow{g}2, 2𝑔12\xrightarrow{g}1, 3𝑔23\xrightarrow{g}2
πs,15\pi_{s,15} 1𝑟21\xrightarrow{r}2, 2𝑔12\xrightarrow{g}1, 3𝑟23\xrightarrow{r}2 πs,49\pi^{\star}_{s,49} π3𝑇{1,3},2π4\pi_{3}\xrightarrow{T}_{\{1,3\},2}\pi_{4}
πs,16\pi_{s,16} 2𝑟12\xrightarrow{r}1, π23π3\pi_{2}\to_{3}\pi_{3} πs,50\pi^{\star}_{s,50} 3𝑟23\xrightarrow{r}2
πs,17\pi_{s,17} 2𝑔12\xrightarrow{g}1, 3𝑔13\xrightarrow{g}1 πs,51\pi^{\star}_{s,51} 1𝑟21\xrightarrow{r}2, 2𝑟12\xrightarrow{r}1, π43π1\pi_{4}\to_{3}\pi_{1}
πs,18\pi_{s,18} π3𝑇{2,3},1π4\pi_{3}\xrightarrow{T}_{\{2,3\},1}\pi_{4} πs,52\pi^{\star}_{s,52} π41π1\pi_{4}\to_{1}\pi_{1}, 3𝑔13\xrightarrow{g}1
πs,19\pi_{s,19} 2𝑟12\xrightarrow{r}1, 3𝑟13\xrightarrow{r}1 πs,53\pi^{\star}_{s,53} π12π4\pi_{1}\to_{2}\pi_{4}, 3𝑟13\xrightarrow{r}1
πs,20\pi_{s,20} 1𝑔21\xrightarrow{g}2, 2𝑔12\xrightarrow{g}1, π43π1\pi_{4}\to_{3}\pi_{1} πs,54\pi^{\star}_{s,54} π13π4\pi_{1}\to_{3}\pi_{4}
πs,21\pi_{s,21} π13π4\pi_{1}\to_{3}\pi_{4} πs,55\pi^{\star}_{s,55} 1𝑔11\xrightarrow{g}1, 2𝑔22\xrightarrow{g}2, 3𝑔23\xrightarrow{g}2
πs,22\pi_{s,22} 3𝑔13\xrightarrow{g}1 πs,56\pi^{\star}_{s,56} 1𝑟11\xrightarrow{r}1, π4𝑇{2,3},2π3\pi_{4}\xrightarrow{T}_{\{2,3\},2}\pi_{3}
πs,23\pi_{s,23} 1𝑟21\xrightarrow{r}2, π4𝑇{2,3},1π1\pi_{4}\xrightarrow{T}_{\{2,3\},1}\pi_{1} πs,57\pi^{\star}_{s,57} 2𝑟22\xrightarrow{r}2, 3𝑟23\xrightarrow{r}2
πs,24\pi_{s,24} 2𝑟12\xrightarrow{r}1, 3𝑟13\xrightarrow{r}1 πs,58\pi^{\star}_{s,58} 1𝑔11\xrightarrow{g}1, π33π1\pi_{3}\to_{3}\pi_{1}
πs,25\pi_{s,25} 1𝑔21\xrightarrow{g}2, π13π2\pi_{1}\to_{3}\pi_{2} πs,59\pi^{\star}_{s,59} π32π2\pi_{3}\to_{2}\pi_{2}, 3𝑔13\xrightarrow{g}1
πs,26\pi_{s,26} 1𝑔21\xrightarrow{g}2, π12π3\pi_{1}\to_{2}\pi_{3}, 3𝑔23\xrightarrow{g}2 πs,60\pi^{\star}_{s,60} π1𝑇{1,3},1π4\pi_{1}\xrightarrow{T}_{\{1,3\},1}\pi_{4}
πs,27\pi_{s,27} π2𝑇{1,3},2π4\pi_{2}\xrightarrow{T}_{\{1,3\},2}\pi_{4} πs,61\pi^{\star}_{s,61} 1𝑟11\xrightarrow{r}1, π22π3\pi_{2}\to_{2}\pi_{3}, 3𝑟13\xrightarrow{r}1
πs,28\pi_{s,28} 1𝑟21\xrightarrow{r}2, π32π1\pi_{3}\to_{2}\pi_{1}, 3𝑟23\xrightarrow{r}2 πs,62\pi^{\star}_{s,62} π43π3\pi_{4}\to_{3}\pi_{3}
πs,29\pi_{s,29} π43π1\pi_{4}\to_{3}\pi_{1} πs,63\pi^{\star}_{s,63} π41π3\pi_{4}\to_{1}\pi_{3}, 3𝑔23\xrightarrow{g}2
πs,30\pi_{s,30} π41π1\pi_{4}\to_{1}\pi_{1}, 3𝑔13\xrightarrow{g}1
πs,31\pi_{s,31} π12π4\pi_{1}\to_{2}\pi_{4}, 3𝑟13\xrightarrow{r}1
πs,32\pi_{s,32} π13π4\pi_{1}\to_{3}\pi_{4}
πs,33\pi_{s,33} π11π2\pi_{1}\to_{1}\pi_{2}, 2𝑔22\xrightarrow{g}2,3𝑔23\xrightarrow{g}2
πs,34\pi_{s,34} π4𝑇{2,3},2π3\pi_{4}\xrightarrow{T}_{\{2,3\},2}\pi_{3}

IV-B Case Study II:

In this case study, we consider N=4N=4 robots, K=3K=3 regions and M=3M=3 objects. The regions of interest are circular centered at (88,280)(88,-280)m, (100,160)(100,-160)m, and (200,130)(200,-130)m, with radius equal to 44m, as in the previous case study. The robots’ and object’ mass is taken as 11kg and 0.250.25kg, respectively, while their spherical volume’s radii as 0.750.75m and 0.20.2m, respectively. The power capabilities of the robots are 2,3,4,42,3,4,4, respectively, and the power required for each object is 5,8,10.55,8,10.5, respectively. Initially, the robots are located in regions πinit(1)=π1\pi_{init(1)}=\pi_{1}, πinit(2)=π3\pi_{init(2)}=\pi_{3}, πinit(3)=π2\pi_{init(3)}=\pi_{2}, and πinit(4)=π2\pi_{init(4)}=\pi_{2}, respectively, whereas the objects are located in regions πinito(1)=π2\pi_{init_{o}(1)}=\pi_{2}, πinito(2)=π1\pi_{init_{o}(2)}=\pi_{1}, and πinito(3)=π3\pi_{init_{o}(3)}=\pi_{3}, respectively.

The resulting transition system consists of 24,01224,012 reachable states and 1,805,2021,805,202 transitions and it was created within 1.52 hours. We set the following formula for the multi-robot-object system:

ϕ=\displaystyle\phi= (ψ¯1ψ¯2)(ψ¯3)(ψ¯4)(¬ψ¯1𝒰ψ¯5)\displaystyle\square\Diamond(\bar{\psi}_{1}\lor\bar{\psi}_{2})\land\square\Diamond(\bar{\psi}_{3})\land\square\Diamond(\bar{\psi}_{4})\land(\neg\bar{\psi}_{1}\mathcal{U}\bar{\psi}_{5})
(¬ψ¯6)(ψ¯7ψ¯8)ψ¯5\displaystyle\land\square(\neg\bar{\psi}_{6})\land\Diamond(\bar{\psi}_{7}\lor\bar{\psi}_{8})\land\square\Diamond\bar{\psi}_{5} (14)

with

  • ψ¯1={1-π1,2-π1,3-π1,4-π1,O1-π1,O2-π1,O3-π1}\begin{aligned} \bar{\psi}_{1}=\{\hskip 7.11317pt&``1\text{-}\pi_{1}",``2\text{-}\pi_{1}",``3\text{-}\pi_{1}",``4\text{-}\pi_{1}",\\ &``O_{1}\text{-}\pi_{1}",``O_{2}\text{-}\pi_{1}",``O_{3}\text{-}\pi_{1}"\hskip 17.07182pt\}\end{aligned}

  • ψ¯2={1-π1,2-π4,3-π3,4-π2,O1-π1,O2-π1,O3-π2}\begin{aligned} \bar{\psi}_{2}=\{\hskip 7.11317pt&``1\text{-}\pi_{1}",``2\text{-}\pi_{4}",``3\text{-}\pi_{3}",``4\text{-}\pi_{2}",\\ &``O_{1}\text{-}\pi_{1}",``O_{2}\text{-}\pi_{1}",``O_{3}\text{-}\pi_{2}"\hskip 17.07182pt\}\end{aligned}

  • ψ¯3={1-π2,2-π2,3-π2,4-π3,O1-π1,O2-π1,O3-π3}\begin{aligned} \bar{\psi}_{3}=\{\hskip 7.11317pt&``1\text{-}\pi_{2}",``2\text{-}\pi_{2}",``3\text{-}\pi_{2}",``4\text{-}\pi_{3}",\\ &``O_{1}\text{-}\pi_{1}",``O_{2}\text{-}\pi_{1}",``O_{3}\text{-}\pi_{3}"\hskip 17.07182pt\}\end{aligned}

  • ψ¯4={1-π1,2-π1,3-π2,4-π3,O1-π3,O2-π3,O3-π1}\begin{aligned} \bar{\psi}_{4}=\{\hskip 7.11317pt&``1\text{-}\pi_{1}",``2\text{-}\pi_{1}",``3\text{-}\pi_{2}",``4\text{-}\pi_{3}",\\ &``O_{1}\text{-}\pi_{3}",``O_{2}\text{-}\pi_{3}",``O_{3}\text{-}\pi_{1}"\hskip 17.07182pt\}\end{aligned}

  • ψ¯5={1-π2,2-π2,3-π3,4-π3,O1-π2,O2-π3,O3-π3}\begin{aligned} \bar{\psi}_{5}=\{\hskip 7.11317pt&``1\text{-}\pi_{2}",``2\text{-}\pi_{2}",``3\text{-}\pi_{3}",``4\text{-}\pi_{3}",\\ &``O_{1}\text{-}\pi_{2}",``O_{2}\text{-}\pi_{3}",``O_{3}\text{-}\pi_{3}"\hskip 17.07182pt\}\end{aligned}

  • ψ¯6={1-π2,2-π2,3-π2,4-π2,O1-π3,O2-π3,O3-π3}\begin{aligned} \bar{\psi}_{6}=\{\hskip 7.11317pt&``1\text{-}\pi_{2}",``2\text{-}\pi_{2}",``3\text{-}\pi_{2}",``4\text{-}\pi_{2}",\\ &``O_{1}\text{-}\pi_{3}",``O_{2}\text{-}\pi_{3}",``O_{3}\text{-}\pi_{3}"\hskip 17.07182pt\}\end{aligned}

  • ψ¯7={1-π3,2-π3,3-π2,4-π1,O1-π3,O2-π1,O3-π2}\begin{aligned} \bar{\psi}_{7}=\{\hskip 7.11317pt&``1\text{-}\pi_{3}",``2\text{-}\pi_{3}",``3\text{-}\pi_{2}",``4\text{-}\pi_{1}",\\ &``O_{1}\text{-}\pi_{3}",``O_{2}\text{-}\pi_{1}",``O_{3}\text{-}\pi_{2}"\hskip 17.07182pt\}\end{aligned}

  • ψ¯8={1-π4,2-π4,3-π3,4-π2,O1-π3,O2-π3,O3-π3}\begin{aligned} \bar{\psi}_{8}=\{\hskip 7.11317pt&``1\text{-}\pi_{4}",``2\text{-}\pi_{4}",``3\text{-}\pi_{3}",``4\text{-}\pi_{2}",\\ &``O_{1}\text{-}\pi_{3}",``O_{2}\text{-}\pi_{3}",``O_{3}\text{-}\pi_{3}"\hskip 17.07182pt\}\end{aligned}

The LTL formula in (IV-B) corresponds to a NBA with 88 states - among which one is a final state - and 2727 transitions. STyLuS* found the first feasible prefix and suffix path within 3.733.73 minutes and 1.431.43 minutes on average. We omit the action path of the robots for ease of exposition.

V Conclusion

We propose an algorithm for the control and planning of multi-robot-object systems subject LTL tasks. We develop adaptive feedback-control protocols for robot navigation and cooperative object transportation, which enable the abstraction of the underlying continuous dynamics to a finite transition system. We compose the transition system with an automaton that represents the LTL task and use a sampling-based planner to derive an optimal task-satisfying plan for the robots.

References

  • [1] G. E. Fainekos, A. Girard, H. Kress-Gazit, and G. J. Pappas, “Temporal logic motion planning for dynamic robots,” Automatica, vol. 45, no. 2, pp. 343–352, 2009.
  • [2] M. Lahijanian, M. R. Maly, D. Fried, L. E. Kavraki, H. Kress-Gazit, and M. Y. Vardi, “Iterative temporal planning in uncertain environments with partial satisfaction guarantees,” IEEE Transactions on Robotics, vol. 32, no. 3, pp. 583–599, 2016.
  • [3] S. G. Loizou and K. J. Kyriakopoulos, “Automatic synthesis of multi-agent motion tasks based on ltl specifications,” IEEE Conference on Decision and Control (CDC), vol. 1, pp. 153–158, 2004.
  • [4] Y. Diaz-Mercado, A. Jones, C. Belta, and M. Egerstedt, “Correct-by-construction control synthesis for multi-robot mixing,” IEEE Conference on Decision and Control (CDC), pp. 221–226, 2015.
  • [5] Y. Chen, X. C. Ding, A. Stefanescu, and C. Belta, “Formal approach to the deployment of distributed robotic teams,” IEEE Transactions on Robotics, vol. 28, no. 1, pp. 158–171, 2012.
  • [6] R. V. Cowlagi and Z. Zhang, “Motion-planning with linear temporal logic specifications for a nonholonomic vehicle kinematic model,” American Control Conference (ACC), pp. 6411–6416, 2016.
  • [7] C. Belta, V. Isler, and G. J. Pappas, “Discrete abstractions for robot motion planning and control in polygonal environments,” IEEE Transactions on Robotics, vol. 21, no. 5, pp. 864–874, 2005.
  • [8] A. Bhatia, M. R. Maly, L. E. Kavraki, and M. Y. Vardi, “Motion planning with complex goals,” IEEE Robotics & Automation Magazine, vol. 18, no. 3, pp. 55–64, 2011.
  • [9] I. Filippidis, D. V. Dimarogonas, and K. J. Kyriakopoulos, “Decentralized multi-agent control from local ltl specifications,” IEEE Conference on Decision and Control (CDC), pp. 6235–6240, 2012.
  • [10] M. Guo and D. V. Dimarogonas, “Multi-agent plan reconfiguration under local ltl specifications,” The International Journal of Robotics Research, vol. 34, no. 2, pp. 218–235, 2015.
  • [11] C. Belta and L. C. Habets, “Controlling a class of nonlinear systems on rectangles,” IEEE Transactions on Automatic Control, vol. 51, no. 11, pp. 1749–1759, 2006.
  • [12] G. Reißig, “Computing abstractions of nonlinear systems,” IEEE Transactions on Automatic Control, vol. 56, no. 11, pp. 2583–2598, 2011.
  • [13] A. Tiwari, “Abstractions for hybrid systems,” Formal Methods in System Design, vol. 32, no. 1, pp. 57–83, 2008.
  • [14] M. Rungger, A. Weber, and G. Reissig, “State space grids for low complexity abstractions,” IEEE Conference on Decision and Control (CDC), pp. 6139–6146, 2015.
  • [15] D. Boskos and D. V. Dimarogonas, “Decentralized abstractions for feedback interconnected multi-agent systems,” IEEE Conference on Decision and Control (CDC), pp. 282–287, 2015.
  • [16] C. Belta and V. Kumar, “Abstraction and control for groups of robots,” IEEE Transactions on robotics, vol. 20, no. 5, pp. 865–875, 2004.
  • [17] T. G. Sugar and V. Kumar, “Control of cooperating mobile manipulators,” IEEE Transactions on robotics and automation, vol. 18, no. 1, pp. 94–103, 2002.
  • [18] D. Heck, D. Kostic, A. Denasi, and H. Nijmeijer, “Internal and external force-based impedance control for cooperative manipulation,” pp. 2299–2304, 2013.
  • [19] Y. Kume, Y. Hirata, and K. Kosuge, “Coordinated motion control of multiple mobile manipulators handling a single object without using force/torque sensors,” IEEE/RSJ International Conference on Intelligent Robots and Systems (IROS), pp. 4077–4082, 2007.
  • [20] A. Tsiamis, C. K. Verginis, C. P. Bechlioulis, and K. J. Kyriakopoulos, “Cooperative manipulation exploiting only implicit communication,” IEEE/RSJ International Conference on Intelligent Robots and Systems (IROS), pp. 864–869, 2015.
  • [21] F. Ficuciello, A. Romano, L. Villani, and B. Siciliano, “Cartesian impedance control of redundant manipulators for human-robot co-manipulation,” IEEE/RSJ International Conference on Intelligent Robots and Systems (IROS), pp. 2120–2125, 2014.
  • [22] A.-N. Ponce-Hinestroza, J.-A. Castro-Castro, H.-I. Guerrero-Reyes, V. Parra-Vega, and E. Olguín-Díaz, “Cooperative redundant omnidirectional mobile manipulators: Model-free decentralized integral sliding modes and passive velocity fields,” IEEE International Conference on Robotics and Automation (ICRA), pp. 2375–2380, 2016.
  • [23] A. Marino, “Distributed adaptive control of networked cooperative mobile manipulators,” IEEE Transactions on Control Systems Technology, 2017.
  • [24] A. Nikou, C. Verginis, S. Heshmati-Alamdari, and D. V. Dimarogonas, “A nonlinear model predictive control scheme for cooperative manipulation with singularity and collision avoidance,” Mediterranean Conference on Control and Automation (MED), pp. 707–712, 2017.
  • [25] C. K. Verginis, A. Nikou, and D. V. Dimarogonas, “Communication-based decentralized cooperative object transportation using nonlinear model predictive control,” European control conference (ECC), pp. 733–738, 2018.
  • [26] C. K. Verginis, M. Mastellaro, and D. V. Dimarogonas, “Robust cooperative manipulation without force/torque measurements: Control design and experiments,” IEEE Transactions on Control Systems Technology, vol. 28, no. 3, pp. 713–729, 2019.
  • [27] S. Erhart and S. Hirche, “Model and analysis of the interaction dynamics in cooperative manipulation tasks,” IEEE Transactions on Robotics, vol. 32, no. 3, pp. 672–683, 2016.
  • [28] H. G. Tanner, S. G. Loizou, and K. J. Kyriakopoulos, “Nonholonomic navigation and control of cooperating mobile manipulators,” IEEE Transactions on robotics and automation, vol. 19, no. 1, pp. 53–64, 2003.
  • [29] C. K. Verginis and D. V. Dimarogonas, “Timed abstractions for distributed cooperative manipulation,” Autonomous Robots, pp. 1–19, 2017.
  • [30] ——, “Multi-agent motion planning and object transportation under high level goals,” IFAC World Congress, 2017.
  • [31] C. Baier, J.-P. Katoen, and K. G. Larsen, Principles of model checking. MIT press, 2008.
  • [32] C. Makkar, W. Dixon, W. Sawyer, and G. Hu, “A new continuously differentiable friction model for control systems design,” IEEE/ASME International Conference on Advanced Intelligent Mechatronics, pp. 600–605, 2005.
  • [33] D. Z. Chen, “Sphere packing problem,” Encyclopedia of Algorithms, pp. 1–99, 2008.
  • [34] M. R. Cutkosky, Robotic grasping and fine manipulation. Springer Science & Business Media, 2012, vol. 6.
  • [35] M. F. Reis, A. C. Leite, and F. Lizarralde, “Modeling and control of a multifingered robot hand for object grasping and manipulation tasks,” IEEE Conference on Decision and Control (CDC), pp. 159–164, 2015.
  • [36] C. K. Verginis and D. V. Dimarogonas, “Adaptive robot navigation with collision avoidance subject to 2nd-order uncertain dynamics,” Automatica, vol. 123, p. 109303, 2021.
  • [37] P. Vlantis, C. Vrohidis, C. P. Bechlioulis, and K. J. Kyriakopoulos, “Robot navigation in complex workspaces using harmonic maps,” IEEE International Conference on Robotics and Automation (ICRA), pp. 1726–1731, 2018.
  • [38] S. L. Smith, J. Tumova, C. Belta, and D. Rus, “Optimal path planning for surveillance with temporal-logic constraints,” The International Journal of Robotics Research, vol. 30, no. 14, pp. 1695–1708, 2011.
  • [39] A. Ulusoy, S. L. Smith, and C. Belta, “Optimal multi-robot path planning with ltl constraints: guaranteeing correctness through synchronization,” in Distributed Autonomous Robotic Systems. Springer, 2014, pp. 337–351.
  • [40] Y. Kantaros and M. M. Zavlanos, “Stylus*: A temporal logic optimal control synthesis algorithm for large-scale multi-robot systems,” The International Journal of Robotics Research, vol. 39, no. 7, pp. 812–836, 2020.