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

\fail

Improved Hierarchical ADMM for Nonconvex Cooperative Distributed Model Predictive Control

Xiaoxue Zhang, Jun Ma, Zilong Cheng, Sunan Huang, Clarence W. de Silva, and Tong Heng Lee X. Zhang, Z. Cheng, and T. H. Lee are with the Integrative Sciences and Engineering Programme, NUS Graduate School, National University of Singapore, 119077 (e-mail: [email protected]; [email protected]; [email protected]).J. Ma is with the Department of Mechanical Engineering, University of California, Berkeley, CA 94720 USA (e-mail: [email protected]).S. Huang is with the Temasek Laboratories, National University of Singapore, Singapore, 117411 (e-mail: [email protected]).Clarence W. de Silva is with the Department of Mechanical Engineering, University of British Columbia, Vancouver, BC, Canada V6T 1Z4 (e-mail: [email protected]).This work has been submitted to the IEEE for possible publication. Copyright may be transferred without notice, after which this version may no longer be accessible.
Abstract

Distributed optimization is often widely attempted and innovated as an attractive and preferred methodology to solve large-scale problems effectively in a localized and coordinated manner. Thus, it is noteworthy that the methodology of distributed model predictive control (DMPC) has become a promising approach to achieve effective outcomes, e.g., in decision-making tasks for multi-agent systems. However, the typical deployment of such distributed MPC frameworks would lead to the involvement of nonlinear processes with a large number of nonconvex constraints. To address this important problem, the development and innovation of a hierarchical three-block alternating direction method of multipliers (ADMM) approach is presented in this work to solve this nonconvex cooperative DMPC problem in multi-agent systems. Here firstly, an additional slack variable is introduced to transform the original large-scale nonconvex optimization problem. Then, a hierarchical ADMM approach, which contains outer loop iteration by the augmented Lagrangian method (ALM) and inner loop iteration by three-block semi-proximal ADMM, is utilized to solve the resulting transformed nonconvex optimization problem. Additionally, it is analytically shown and established that the requisite desired stationary point exists for convergence in the algorithm. Finally, an approximate optimization stage with a barrier method is then applied to further significantly improve the computational efficiency, yielding the final improved hierarchical ADMM. The effectiveness of the proposed method in terms of attained performance and computational efficiency is demonstrated on a cooperative DMPC problem of decision-making process for multiple unmanned aerial vehicles (UAVs).

Index Terms:
Alternating direction method of multipliers (ADMM), distributed optimization, model predictive control (MPC), multi-agent system, collision avoidance.

I Introduction

With the rapid development of communication and computation technologies, cooperative decision making of large-scale multi-agent systems has become a promising trend in the development of automated systems [1, 2, 3]. Nevertheless, the significant challenge involved in cooperative decision making, in handling the coupling conditions among the interconnected agents with an appropriate guarantee of efficient computation still remains. In the existing literature, a significant amount of effort has been put into distributed control and coordinated networks for large-scale interconnected multi-agent systems, with the objective of optimizing the global cost while satisfying the required safety constraints collectively. Essentially for such approaches, the control and optimization algorithms deployed are distributed in structure, and the deployment utilizes only the local observations and information [4, 5]. With the appropriate communication technology in place, the agents are developed to suitably make decisions while attempting to satisfy all of the constraints or requirements invoked [6, 7]. On the other hand, a remarkable alternate development that utilizes an approach termed as distributed model predictive control (DMPC) has also been shown to display great promise to be effectively used to handle the input and state constraints with multiple objectives; and possibly also applicable to this important problem of cooperative decision making of large-scale multi-agent systems [8]. In addition, via the DMPC framework, there is great potential that the computational burden can be significantly relieved by decomposing the centralized system into several interconnected subsystems (with also a similar situation of availability of knowledge of necessary information among the connected agents).

For the DMPC problem with interconnected agents, the promising optimization approach of the alternating direction method of multipliers (ADMM) algorithm has the potential for very effective utilization. In this context, it can be noted that a large number of research works have demonstrated that the ADMM methodology has the capability to rather effectively determine the optimal solution to many challenging optimization problems, such as distributed optimization [9], decentralized control [10, 11], and statistical learning problems [12, 13]. The key idea of the ADMM approach is to utilize a decomposition coordination principle, wherein local sub-problems with smaller sizes are solved, and then the solutions of these subproblems can be coordinated to obtain the solution to the original problem. Notably, the ADMM solves the sub-problems by optimizing blocks of variables, such that efficient distributed computation can be attained. The convergence of the ADMM with two blocks has been investigated for convex optimization problems in [14, 15], and typical applications of such two-block ADMM approaches in distributed convex optimization problems are reported extensively in [16, 17]. Several researchers have also made rather noteworthy extensions from this two-block convex ADMM to multi-block convex ADMM in [18, 19]. However, although seemingly evident as a possibility with great potential, there has not been much substantial work on innovating the ADMM framework for nonconvex DMPC. Classical ADMM is known only to admit a linear convergence rate for convex optimization problems.

For nonconvex optimization problems, certain additional conditions and assumptions need to be considered to guarantee the required convergence [20, 21, 22, 23, 24, 25, 26]. More specifically, the work in [20] studies the two-block ADMM with one of the blocks defined as the identity matrix; and it also states that the objective function and the domain of the optimization variable need to be lower semi-continuous. Yet further, a nonconvex Bregman-ADMM is proposed in [21] by introducing a Bregman divergence term to the augmented Lagrangian function to promote the descent of certain potential functions. Moreover, the work in [22] proposes two proximal-type variants of ADMM with ϵ\epsilon-stationarity conditions to solve the nonconvex optimization problem with affine coupling constraints under the assumption that the gradient of the objective function is Lipschitz continuous. Additionally, in [23], a multi-block ADMM is presented to minimize a nonconvex and nonsmooth objective function subject to certain coupling linear equality constraints; and the work in [24] presents an ADMM-type algorithm for nonconvex optimization problem in the application environment of modern machine learning problems, though requiring the assumption that the penalty parameter should be large enough. Besides, an augmented Lagrangian based alternating direction inexact newton (ALADIN) method is introduced in [27] to solve the optimization problems with a separable nonconvex objective function and coupled affine constraints.

Noting all these additional conditions and assumptions that typically can arise, it is the situation where the ADMM approach cannot yet be directly applied to solve a large-scale constrained nonconvex program without further relaxation or reformulation. Also, the ADMM with more than two blocks is typically implemented with sequential optimization; and while a direct parallelization of a multi-block ADMM formulation is also sometimes considered, the aforementioned approach has the disadvantage that it is typically less likely to converge [28, 29]. In order to address all these limitations, a hierarchical three-block ADMM approach is utilized in our paper to solve the nonconvex optimization problem that arises for decision making in multi-agent systems. Firstly, an additional slack variable is innovatively introduced to transform the original large-scale nonconvex optimization problem, such that the intractable nonconvex coupling constraints are suitably related to the distributed agents. Then, the approach with a hierarchical ADMM that contains the outer loop iteration by the augmented Lagrangian method (ALM), and inner loop iteration by three-block semi-proximal ADMM, is utilized to solve the resulting transformed nonconvex optimization problem. Next, the approximate optimization with a barrier method is then applied to improve the computational efficiency. Finally, a multi-agent system involving decision making for multiple unmanned aerial vehicles (UAVs) is utilized to demonstrate the effectiveness of the proposed method in terms of attained performance and computational efficiency.

The remainder of this paper is organized as follows. Section II describes a multi-agent system that contains multiple interconnected agents equipped with sensors and communication devices, and also formulates a large-scale nonconvex optimization problem for decision making. Section III gives a more compact form of the optimization problem, which is further transformed by introducing slack variables. In section IV, the hierarchical ADMM is used to solve the transformed optimization problem. Next, an improved version of such hierarchical ADMM is presented to accelerate the computation process based on the barrier method in Section V. Then, in Section VI, a multi-UAV system is used as an example to show the effectiveness and computational efficiency of the hierarchical ADMM and improved hierarchical ADMM. Finally, pertinent conclusions of this work are given in Section VII.

II Problem Formulation

The following notations are used in the remaining content. a×b\mathbb{R}^{a\times b} denotes the set of real matrices with aa rows and bb columns, a\mathbb{R}^{a} means the set of aa-dimensional real column vectors. The symbol X0X\succ 0 and X0X\succeq 0 mean that the matrix XX is positive definite and positive semi-definite, respectively. x>yx>y and xyx\geq y mean that vector xx is element-wisely greater and no less than the vector yy. AA^{\top} and xx^{\top} denote the transpose of the matrix AA and vector xx. IaI_{a} represents the aa-dimensional identity matrix; 1a1_{a} and 1(a,b)1_{(a,b)} denote the aa-dimensional all-one vector and the aa-by-bb all-one matrix, respectively; 0a0_{a} and 0(a,b)0_{(a,b)} represent the aa-dimensional all-zero vector and the aa-by-bb all-zero matrix, respectively. The operator X,Y\langle X,Y\rangle denotes the Frobenius inner product, i.e., X,Y=Tr(XY)\langle X,Y\rangle=\operatorname{Tr}(X^{\top}Y) for all X,Ya×bX,Y\in\mathbb{R}^{a\times b}; the operator X\|X\| is the Frobenius norm of matrix XX. \otimes denotes the Kronecker product and \odot denotes the Hadamard product (Schur product). blockdiag(X1,X2,,Xn)\operatorname{blockdiag}(X_{1},X_{2},\cdots,X_{n}) denotes a block diagonal matrix with diagonal entries X1,X2,,XnX_{1},X_{2},\cdots,X_{n}; diag(a1,a2,,an)\operatorname{diag}(a_{1},a_{2},\cdots,a_{n}) denotes a diagonal matrix with diagonal entries a1,a2,,ana_{1},a_{2},\cdots,a_{n}. a\mathbb{Z}^{a} and ab\mathbb{Z}_{a}^{b} mean the sets of positive integers {1,2,,a}\{1,2,\cdots,a\} and {a,a+1,,b}\{a,a+1,\cdots,b\}, respectively. +\mathbb{Z}_{+} denotes the set of positive integers {1,2,}\{1,2,\cdots\}. The operator {xi}iab\{x_{i}\}_{\forall i\in\mathbb{Z}_{a}^{b}} means the concatenation of the vector xix_{i} for all iabi\in\mathbb{Z}_{a}^{b}, i.e., {xi}iab=(xa,xa+1,,xb)=[xaxa+1xb]\{x_{i}\}_{\forall i\in\mathbb{Z}_{a}^{b}}=(x_{a},x_{a+1},\cdots,x_{b})=\begin{bmatrix}x_{a}^{\top}&x_{a+1}^{\top}&\cdots&x_{b}^{\top}\end{bmatrix}^{\top}.

II-A Modelling of Interconnected Agents

In order to represent the constraint or information topology of multiple interconnected agents, an undirected graph 𝒢(𝒱,)\mathcal{G}(\mathcal{V},\mathcal{E}) can be utilized. The node set 𝒱={1,2,,N}\mathcal{V}=\{1,2,\cdots,N\} denotes the agents, and the edge set \mathcal{E} denotes the coupling constraints (information flow) between two interconnected agents, which is defined as

{(i,j)(t),dsafepipjdcmu(i,j)(t),otherwise,\begin{cases}(i,j)\in\mathcal{E}(t),&\;d_{\mathrm{safe}}\leq\|p_{i}-p_{j}\|\leq d_{\mathrm{cmu}}\\ (i,j)\notin\mathcal{E}(t),&\;\text{otherwise},\end{cases} (1)

where NN and MM are the number of agents and coupling constraints in the multi-agent system, respectively, pi,pjp_{i},p_{j} are the position vectors of the iith agent and jjth agent, dcmud_{\mathrm{cmu}} and dsafed_{\mathrm{safe}} mean the maximum communication distance and minimum safe distance between two agents, respectively. Therefore, based on the communication topology, an adjacency matrix of the network (denoted by 𝒟\mathcal{D}) can be obtained, which is a square symmetric matrix. The elements of 𝒟\mathcal{D} indicate whether pairs of vertices are adjacent/connected or not in the graph. Therefore, the neighbor nodes of the iith agent are the corresponding indexes of nonzero elements of the iith row in 𝒟\mathcal{D}, which is denoted by ν(i)={j|(i,j),j𝒱}\nu(i)=\{j|(i,j)\in\mathcal{E},\forall j\in\mathcal{V}\}.

II-B Problem Description

Each connected agent in the set 𝒱\mathcal{V} has its origin and target state, which means each agent needs to achieve its task by coordinating with other connected agents in 𝒱\mathcal{V}. Moreover, the information of the connected agents (neighbors) is necessary for each agent to make decisions, due to the requirement of the communication topology (1). Since the communication between connected agents occurs during the whole process, the present state information of neighbors needs to be conveyed to the iith agent to make decisions. Thus, the cooperative task can be formulated as a cooperative DMPC problem for each agent i𝒱i\in\mathcal{V}. Then, given the initial state xi(t)x_{i}(t), the DMPC problem for the iith agent at timestamp tt can be written as

minui,xi\displaystyle\min_{u_{i},x_{i}}\quad i𝒱Ji(xi,ui)\displaystyle\sum_{i\in\mathcal{V}}J_{i}(x_{i},u_{i}) (2a)
s.t.\displaystyle\operatorname{s.t.}\quad xi(τ+1)=Aixi(τ)+Biui(τ)\displaystyle x_{i}(\tau+1)=A_{i}x_{i}(\tau)+B_{i}u_{i}(\tau)
u¯iui(τ)u¯i\displaystyle\underline{u}_{i}\leq u_{i}(\tau)\leq\overline{u}_{i}
x¯ixi(τ+1)x¯i\displaystyle\underline{x}_{i}\leq x_{i}(\tau+1)\leq\overline{x}_{i}
(xi(τ+1),xj(τ+1))\displaystyle(x_{i}(\tau+1),x_{j}(\tau+1))\in\mathbb{C}
jν(i),τtt+T1,\displaystyle\forall j\in\nu(i),\forall\tau\in\mathbb{Z}_{t}^{t+T-1},

where the subscript ()i(\cdot)_{i} means the corresponding variable of the iith agent, xinx_{i}\in\mathbb{R}^{n} and uimu_{i}\in\mathbb{R}^{m} denote the state and control input vectors of the iith agent, respectively. Ain×n,Bin×mA_{i}\in\mathbb{R}^{n\times n},B_{i}\in\mathbb{R}^{n\times m} are the state matrix and control input matrix, respectively. tt denotes the initial time stamp and TT is the prediction horizon. Notice that (2a) is the objective function for the state and control variables; (II-B) denotes the kinematics of the agents; (II-B) and (II-B) denote the bound constraints of state and control input vectors, x¯i\overline{x}_{i} (x¯i\underline{x}_{i}) and u¯i\overline{u}_{i} (u¯i\underline{u}_{i}) are the upper (lower) bound of state and control input vectors, respectively; (II-B) denotes the coupling constraints of two interconnected agents (the iith agent and the jjth agent with jν(i)j\in\nu(i)) and \mathbb{C} denotes the set of the coupling constraints. The cost function of the iith agent Ji(xi,ui)J_{i}(x_{i},u_{i}) is a summation of the stage cost term i\ell_{i} and the terminal cost term T,i\ell_{T,i}, i.e.,

Ji(xi,ui)\displaystyle\quad J_{i}(x_{i},u_{i})
=i(xi(t+1),,xi(t+T1),ui(t),,ui(t+T1))\displaystyle=\ell_{i}\left(x_{i}(t+1),\cdots,x_{i}(t+T-1),u_{i}(t),\cdots,u_{i}(t+T-1)\right)
+T,i(xi(t+T)),\yesnumber\displaystyle+\ell_{T,i}(x_{i}(t+T)),\yesnumber

where the initial state xi(t)x_{i}(t) is known. The stage cost function i\ell_{i} is defined as

i(xi,ui)\displaystyle\ell_{i}(x_{i},u_{i}) =\displaystyle= τ=t+1t+T1xi(τ)xi,ref(τ),Qi(xi(τ)xi,ref(τ))\displaystyle\sum_{\tau=t+1}^{t+T-1}\langle x_{i}(\tau)-x_{i,\text{ref}}(\tau),\>Q_{i}\left(x_{i}(\tau)-x_{i,\text{ref}}(\tau)\right)\rangle
+τ=tt+T1ui(τ),Riui(τ),\yesnumber\displaystyle+\sum_{\tau=t}^{t+T-1}\langle u_{i}(\tau)\>,\>R_{i}u_{i}(\tau)\rangle,\yesnumber

where xi,refx_{i,\mathrm{ref}} is the reference state the iith agent tries to follow, and Qin×nQ_{i}\in\mathbb{R}^{n\times n} and Rim×mR_{i}\in\mathbb{R}^{m\times m} are the diagonal weighting matrices for the state and control input vectors, respectively. The terminal cost function T,i\ell_{T,i} is

T,i(xi(t+T))\displaystyle\ell_{T,i}(x_{i}(t+T)) =\displaystyle= xi(t+T)xi,ref(t+T),\displaystyle\langle x_{i}(t+T)-x_{i,\text{ref}}(t+T)\>,\>
Qi,T(xi(t+T)xi,ref(t+T)),\yesnumber\displaystyle\>\>Q_{i,T}\left(x_{i}(t+T)-x_{i,\text{ref}}(t+T)\right)\rangle,\yesnumber

where Qi,TQ_{i,T} is the diagonal weighting matrix for the terminal state vector.

In a specific multi-agent system, \mathbb{C} can be specified as

={(xi(τ),xj(τ))|dsafeLp(xi(τ)xj(τ))dcmu,\displaystyle\mathbb{C}=\Big{\{}(x_{i}(\tau),x_{j}(\tau))\>\Big{|}\>d_{\mathrm{safe}}\leq\left\|L_{p}(x_{i}(\tau)-x_{j}(\tau))\right\|\leq d_{\mathrm{cmu}},
i𝒱,jν(i),τt+1t+T},\yesnumber\displaystyle\quad\forall i\in\mathcal{V},\forall j\in\nu(i),\forall\tau\in\mathbb{Z}_{t+1}^{t+T}\Big{\}},\yesnumber

where LpL_{p} is the locating matrix to extract the position vector pip_{i} from the state vector xix_{i}, i.e., pi(τ)=Lpxi(τ)p_{i}(\tau)=L_{p}x_{i}(\tau).

Assumption 1.

ν(i)\nu(i) in the prediction horizon [t+1,t+T][t+1,t+T] is time-invariant.

III Problem Transformation

III-A Problem Reformulation

First of all, the decision variable is set as

zi\displaystyle z_{i} =\displaystyle= (ui(t),xi(t+1),ui(t+1),xi(t+2),,\displaystyle\big{(}u_{i}(t),x_{i}(t+1),u_{i}(t+1),x_{i}(t+2),\cdots,
ui(t+T1),xi(t+T))T(m+n).\yesnumber\displaystyle\quad u_{i}(t+T-1),x_{i}(t+T)\big{)}\in\mathbb{R}^{T(m+n)}.\yesnumber

It is straightforward that Lpxi=pinp,i𝒱L_{p}x_{i}=p_{i}\in\mathbb{R}^{n_{p}},\forall i\in\mathcal{V}, where npn_{p} is the dimension of position in state vector pip_{i} and npnn_{p}\leq n, and pi(τ)pj(t)2=pi(τ)pj(t),pi(τ)pj(t)\|p_{i}(\tau)-p_{j}(t)\|^{2}=\left\langle p_{i}(\tau)-p_{j}(t)\>,\>p_{i}(\tau)-p_{j}(t)\right\rangle. For all jν(i)j\in\nu(i) and τt+1t+T\tau\in\mathbb{Z}_{t+1}^{t+T}, the coupling constraints dsafepi(τ)pj(τ)dcmud_{\mathrm{safe}}\leq\|p_{i}(\tau)-p_{j}(\tau)\|\leq d_{\mathrm{cmu}}, jν(i)\forall j\in\nu(i), i𝒱\forall i\in\mathcal{V}, τt+1t+T\forall\tau\in\mathbb{Z}_{t+1}^{t+T} can be equivalently expressed as

dsafe2(pi(t+1)pj(t))(pi(t+1)pj(t))dcmu2,\displaystyle d_{\mathrm{safe}}^{2}\leq\left(p_{i}(t+1)-p_{j}(t)\right)^{\top}\left(p_{i}(t+1)-p_{j}(t)\right)\leq d_{\mathrm{cmu}}^{2},
jν(i),\displaystyle\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\forall j\in\nu(i),
\displaystyle\vdots
dsafe2(pi(t+T)pj(t))(pi(t+T)pj(t))dcmu2,\displaystyle d_{\mathrm{safe}}^{2}\leq\left(p_{i}(t+T)-p_{j}(t)\right)^{\top}\left(p_{i}(t+T)-p_{j}(t)\right)\leq d_{\mathrm{cmu}}^{2},
jν(i),\yesnumber\displaystyle\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\forall j\in\nu(i),\yesnumber

where rir_{i} is the number of the neighbours of the iith agent, i.e., ri=|ν(i)|r_{i}=|\nu(i)| where |||\cdot| is the cardinality operator for the set ν(i)\nu(i). This means the iith agent has rir_{i} neighbors, and there are rir_{i} inequalities for each τ\tau. Thus, the coupling constraint dsafepi(τ)pj(τ)dcmud_{\mathrm{safe}}\leq\|p_{i}(\tau)-p_{j}(\tau)\|\leq d_{\mathrm{cmu}}, jν(i)\forall j\in\nu(i), i𝒱\forall i\in\mathcal{V}, τt+1t+T\forall\tau\in\mathbb{Z}_{t+1}^{t+T} denotes TriTr_{i} inequalities for all τ\tau and jj.

Define a known vector [pj]={pj(t)}jν(i)rinp[p_{j}]=\left\{p_{j}(t)\right\}_{\forall j\in\nu(i)}\in\mathbb{R}^{r_{i}n_{p}}. Notice that the position vector of all neighbors of the iith agent pj(t)p_{j}(t) are known due to the existence of communication. Thus, the coupling constraints can be rewritten as

(zi;[pj])𝒫(zi;[pj])\displaystyle\mathcal{M}\left(z_{i};[p_{j}]\right)\mathcal{P}\left(z_{i};[p_{j}]\right) \displaystyle\leq dcmu2\displaystyle d_{\mathrm{cmu}}^{2}
(zi;[pj])𝒫(zi;[pj])\displaystyle\mathcal{M}\left(z_{i};[p_{j}]\right)\mathcal{P}\left(z_{i};[p_{j}]\right) \displaystyle\geq dsafe2,\yesnumber\displaystyle d_{\mathrm{safe}}^{2},\yesnumber

with the mapping 𝒫:T(m+n)Trinp\mathcal{P}:\mathbb{R}^{T(m+n)}\rightarrow\mathbb{R}^{Tr_{i}n_{p}} characterized by

𝒫(zi;[pj])=(IT1riInp)Mpzi1T[pj],\yesnumber\displaystyle\mathcal{P}\left(z_{i};[p_{j}]\right)=(I_{T}\otimes 1_{r_{i}}\otimes I_{n_{p}})M_{p}z_{i}-1_{T}\otimes[p_{j}],\yesnumber

where the matrix MpTnp×T(m+n)M_{p}\in\mathbb{R}^{Tn_{p}\times T(m+n)} is a matrix to extract all position vectors (pi(t+1),pi(t+2),,pi(t+T))\left(p_{i}(t+1),p_{i}(t+2),\cdots,p_{i}(t+T)\right) from the decision variable ziz_{i}, and Mp=IT[0(np,m)Inp0(np,(nnp))]M_{p}=I_{T}\otimes\begin{bmatrix}{0}_{(n_{p},m)}&I_{n_{p}}&{0}_{(n_{p},(n-n_{p}))}\end{bmatrix}. Besides, the mapping :T(m+n)Tri×Trinp\mathcal{M}:\mathbb{R}^{T(m+n)}\rightarrow\mathbb{R}^{Tr_{i}\times Tr_{i}n_{p}} is given by

(zi;[pj])=(ITri1np)(ITrinp(𝒫(zi;[pj])1Trinp)).\displaystyle\mathcal{M}\left(z_{i};[p_{j}]\right)=\left(I_{Tr_{i}}\otimes 1_{n_{p}}^{\top}\right)\left(I_{Tr_{i}n_{p}}\odot\left({\mathcal{P}\left(z_{i};[p_{j}]\right)}{1}_{Tr_{i}n_{p}}^{\top}\right)\right).
\yesnumber\displaystyle\yesnumber

Therefore, the large-scale cooperative optimization problem (2a) can be reformulated as

minzi\displaystyle\min\limits_{z_{i}} i𝒱Ji(zi)\displaystyle\sum_{i\in\mathcal{V}}J_{i}(z_{i})
s.t.\displaystyle\operatorname{s.t.} Gizi=gi\displaystyle G_{i}z_{i}=g_{i} (7)
Fz,izifz,i\displaystyle F_{z,i}z_{i}\leq f_{z,i}
(zi;[pj])𝒫(zi;[pj])dcmu2\displaystyle{\mathcal{M}\left(z_{i};[p_{j}]\right)\mathcal{P}\left(z_{i};[p_{j}]\right)}\leq d_{\mathrm{cmu}}^{2}
(zi;[pj])𝒫(zi;[pj])dsafe2\displaystyle{\mathcal{M}\left(z_{i};[p_{j}]\right)\mathcal{P}\left(z_{i};[p_{j}]\right)}\geq d_{\mathrm{safe}}^{2}
jν(i),τt+1t+T,\displaystyle\forall j\in\nu(i),\forall\tau\in\mathbb{Z}_{t+1}^{t+T},

where Ji(zi)=zizi,ref,Hi(zizi,ref)J_{i}(z_{i})=\langle z_{i}-z_{i,\text{ref}}\>,\>H_{i}(z_{i}-z_{i,\text{ref}})\rangle; Hi=H_{i}= blockdiag(Ri,Qi,Ri,Qi,,Ri,Qi,Ri,QT,i)\operatorname{blockdiag}(R_{i},Q_{i},R_{i},Q_{i},\cdots,R_{i},Q_{i},R_{i},Q_{T,i}) is the weighting matrix for the optimization variable ziz_{i}; zi,refz_{i,\text{ref}} is the reference vector, i.e., zi,ref=(0m,xi,ref(t+1),z_{i,\text{ref}}=\big{(}0_{m},x_{i,\text{ref}}(t+1), 0m,xi,ref(t+2),,0m,xi,ref(t+T))0_{m},x_{i,\text{ref}}(t+2),\cdots,0_{m},x_{i,\text{ref}}(t+T)\big{)}. Due to the fact that the agents’ dynamic models are linear time-invariant during the optimization horizon [t,t+T][t,t+T], we can rewrite (II-B) as the equality constraint Gizi=giG_{i}z_{i}=g_{i} in (III-A), where

Gi\displaystyle G_{i} =\displaystyle= IT[BiIn]+blockdiag(0(n,m),\displaystyle I_{T}\otimes\begin{bmatrix}-B_{i}&I_{n}\end{bmatrix}\;+\operatorname{blockdiag}\Big{(}0_{(n,m)},
[Ai0(n,m)],,[Ai0(n,m)]T2,[Ai0(n,m+n)])\displaystyle\underbrace{\begin{bmatrix}-A_{i}&0_{(n,m)}\end{bmatrix},\cdots,\begin{bmatrix}-A_{i}&0_{(n,m)}\end{bmatrix}}_{T-2},\begin{bmatrix}-A_{i}&0_{(n,m+n)}\end{bmatrix}\Big{)}

and gi=[(Aixi(t))0n0n0nT1]g_{i}=\Big{[}\left(A_{i}x_{i}(t)\right)^{\top}\quad\underbrace{0_{n}^{\top}\quad 0_{n}^{\top}\quad\cdots\quad 0_{n}^{\top}}_{T-1}\Big{]}^{\top} with a known state vector xi(t)x_{i}(t) at current timestamp tt. Notice that GiTn×T(m+n)G_{i}\in\mathbb{R}^{Tn\times T(m+n)} and giTng_{i}\in\mathbb{R}^{Tn}. Additionally, (II-B) and (II-B) can be rewritten as Fz,izifz,iF_{z,i}z_{i}\leq f_{z,i} in (III-A), where

Fz,i=[IT(m+n)IT(m+n)],fz,i=[1T[u¯ix¯i]1T[u¯ix¯i]].\displaystyle F_{z,i}=\begin{bmatrix}I_{T(m+n)}\\ -I_{T(m+n)}\end{bmatrix},f_{z,i}=\begin{bmatrix}{1_{T}\otimes\begin{bmatrix}\overline{u}_{i}^{\top}&\overline{x}_{i}^{\top}\end{bmatrix}^{\top}}\\ {1_{T}\otimes\begin{bmatrix}-\underline{u}_{i}^{\top}&-\underline{x}_{i}^{\top}\end{bmatrix}^{\top}}\end{bmatrix}.
Remark 1.

It is well-known that increasing the prediction horizon TT can enlarge the domain of attraction of the MPC controller. However, this will increase the computational burden. Weighting the terminal cost can also enlarge the domain of attraction of the MPC controller without the occurrence of the terminal constraints; thus, the stabilizing weighting factor of a given initial state can be included, which has been proved in [30] that the asymptotic stability of the nonlinear MPC controller can be achieved.

By introducing an auxiliary variable zf,i2T(m+n)z_{f,i}\in\mathbb{R}^{2T(m+n)}, we have

minzi\displaystyle\min\limits_{z_{i}} i𝒱Ji(zi)+δ+2T(m+n)(zf,i)\displaystyle\sum_{i\in\mathcal{V}}J_{i}(z_{i})+\delta_{\mathbb{R}_{+}^{2T(m+n)}}(z_{f,i})
s.t.\displaystyle\operatorname{s.t.} 𝒜izi+izf,ihi=0\displaystyle\mathcal{A}_{i}z_{i}+\mathcal{B}_{i}z_{f,i}-h_{i}=0
(zi;[pj])𝒫(zi;[pj])dcmu2\displaystyle{\mathcal{M}\left(z_{i};[p_{j}]\right)\mathcal{P}\left(z_{i};[p_{j}]\right)}\leq d_{\mathrm{cmu}}^{2}
(zi;[pj])𝒫(zi;[pj])dsafe2\displaystyle{\mathcal{M}\left(z_{i};[p_{j}]\right)\mathcal{P}\left(z_{i};[p_{j}]\right)}\geq d_{\mathrm{safe}}^{2}
jν(i),\yesnumber\displaystyle\forall j\in\nu(i),\yesnumber

where 𝒜i=[GiFz,i]\mathcal{A}_{i}=\begin{bmatrix}G_{i}\\ F_{z,i}\end{bmatrix}, i=[0(Tn,2T(m+n))I2T(m+n)]\mathcal{B}_{i}=\begin{bmatrix}0_{(Tn,2T(m+n))}\\ I_{2T(m+n)}\end{bmatrix}, and hi=[gifz,i]h_{i}=\begin{bmatrix}g_{i}\\ f_{z,i}\end{bmatrix}, δ+2T(m+n)(zf,i)\delta_{\mathbb{R}_{+}^{2T(m+n)}}(z_{f,i}) is an indicator function defined as

δ+2T(m+n)(zf,i)={0,if zf,i+2T(m+n),otherwise.\displaystyle\delta_{\mathbb{R}_{+}^{2T(m+n)}}(z_{f,i})=\begin{cases}0,&\text{if }z_{f,i}\in\mathbb{R}_{+}^{2T(m+n)}\\ \infty,&\text{otherwise}.\end{cases} (8)

Since the second inequality in (III-A) is nonlinear and nonconvex, we define a set

𝒵i={ziT(m+n)|(zi)𝒫(zi)dcmu2(zi)𝒫(zi)dsafe2}.\displaystyle\mathcal{Z}_{i}=\left\{z_{i}\in\mathbb{R}^{T(m+n)}\middle|\begin{array}[]{l}\mathcal{M}(z_{i})\mathcal{P}(z_{i})\leq d_{\mathrm{cmu}}^{2}\\ \mathcal{M}(z_{i})\mathcal{P}(z_{i})\geq d_{\mathrm{safe}}^{2}\end{array}\right\}.

Then, the problem (III-A) can be simplified as

minzi\displaystyle\min\limits_{z_{i}} Ji(zi)+δ𝒵i(zi)+δ+2T(m+n)(zf,i)\displaystyle J_{i}(z_{i})+\delta_{\mathcal{Z}_{i}}(z_{i})+\delta_{\mathbb{R}_{+}^{2T(m+n)}}(z_{f,i})
s.t.\displaystyle\operatorname{s.t.} 𝒜izi+izf,ihi=0,\yesnumber\displaystyle\mathcal{A}_{i}z_{i}+\mathcal{B}_{i}z_{f,i}-h_{i}=0,\yesnumber

where δ𝒵i(zi)\delta_{\mathcal{Z}_{i}}(z_{i}) is an indicator function of ziz_{i} onto the set 𝒵i\mathcal{Z}_{i}.

III-B Optimization Problem Transformation

Due to the fact that the direct use of the conventional ADMM to the nonconvex optimization problem (III-A) directly cannot ensure the convergence, problem transformation is significant in the constrained nonconvex optimization problem [22]. Thus, a slack variable si2Tm+3Tns_{i}\in\mathbb{R}^{2Tm+3Tn} is introduced to the problem (III-A), and then the transformed optimization problem is given by

minzi\displaystyle\min\limits_{z_{i}} Ji(zi)+δ𝒵i(zi)+δ+2T(m+n)(zf,i)\displaystyle J_{i}(z_{i})+\delta_{\mathcal{Z}_{i}}(z_{i})+\delta_{\mathbb{R}_{+}^{2T(m+n)}}(z_{f,i})
s.t.\displaystyle\operatorname{s.t.} 𝒜izi+izf,ihi+si=0\displaystyle\mathcal{A}_{i}z_{i}+\mathcal{B}_{i}z_{f,i}-h_{i}+s_{i}=0
si=0.\yesnumber\displaystyle s_{i}=0.\yesnumber

After introducing sis_{i}, the linear coupling constraints in (III-A) change from 2 blocks (ziz_{i} and zf,iz_{f,i}) to 3 blocks (ziz_{i}, zf,iz_{f,i} and sis_{i}). Since the block about sis_{i} is an identity matrix I2Tm+3TnI_{2Tm+3Tn} whose image is the whole space, there always exists an sis_{i} such that 𝒜izi+izf,ihi+si=0\mathcal{A}_{i}z_{i}+\mathcal{B}_{i}z_{f,i}-h_{i}+s_{i}=0 is satisfied, given any ziz_{i} and zf,iz_{f,i}. In addition, the constraint si=0s_{i}=0 can be treated separately from the equality constraint 𝒜izi+izf,ihi+si=0\mathcal{A}_{i}z_{i}+\mathcal{B}_{i}z_{f,i}-h_{i}+s_{i}=0. Therefore, we can consider the constraints in the problem (III-B) by separating them into two groups of constraints. The first group considers the equality constraint and the nonconvex constraint involved by the indicator function δ𝒵i(zi)\delta_{\mathcal{Z}_{i}}(z_{i}) and δ+2T(m+n)(zf,i)\delta_{\mathbb{R}_{+}^{2T(m+n)}}(z_{f,i}), and the other group deals with the constraint si=0s_{i}=0. As for the first sub-problem which does not include the constraint si=0s_{i}=0, some existing techniques of the ADMM can be employed. On the other hand, the constraint si=0s_{i}=0 can be transformed by dualizing the constraint si=0s_{i}=0 with λi\lambda_{i} and adding a quadratic penalty βi2si2\frac{\beta_{i}}{2}\|s_{i}\|^{2}. Then the problem (III-B) can be rewritten as

minzi,zf,i,si\displaystyle\min\limits_{z_{i},z_{f,i},s_{i}} Ji(zi)+δ𝒵i(zi)+δ+2T(m+n)(zf,i)\displaystyle J_{i}(z_{i})+\delta_{\mathcal{Z}_{i}}(z_{i})+\delta_{\mathbb{R}_{+}^{2T(m+n)}}(z_{f,i})
+λi,si+βi2si2\displaystyle+\langle\lambda_{i},s_{i}\rangle+\frac{\beta_{i}}{2}\|s_{i}\|^{2}
s.t.\displaystyle\operatorname{s.t.} 𝒜izi+izf,ihi+si=0,\yesnumber\displaystyle\mathcal{A}_{i}z_{i}+\mathcal{B}_{i}z_{f,i}-h_{i}+s_{i}=0,\yesnumber

where λi2Tm+3Tn\lambda_{i}\in\mathbb{R}^{2Tm+3Tn} is the dual parameter and βi+\beta_{i}\in\mathbb{R}_{+} is the penalty parameter. Then, the ALM is used to solve the second sub-problem.

The augmented Lagrangian function of problem (III-B) is

ρi\displaystyle\mathcal{L}_{\rho_{i}} =\displaystyle= Ji(zi)+δ𝒵i(zi)+δ+2T(m+n)(zf,i)+λi,si\displaystyle J_{i}(z_{i})+\delta_{\mathcal{Z}_{i}}(z_{i})+\delta_{\mathbb{R}_{+}^{2T(m+n)}}(z_{f,i})+\langle\lambda_{i},s_{i}\rangle
+βi2si2+yi,𝒜izi+izf,ihi+si\displaystyle+\frac{\beta_{i}}{2}\|s_{i}\|^{2}+\langle y_{i},\mathcal{A}_{i}z_{i}+\mathcal{B}_{i}z_{f,i}-h_{i}+s_{i}\rangle
+ρi2𝒜izi+izf,ihi+si2,\yesnumber\displaystyle+\frac{\rho_{i}}{2}\|\mathcal{A}_{i}z_{i}+\mathcal{B}_{i}z_{f,i}-h_{i}+s_{i}\|^{2},\yesnumber

where yi2Tm+3Tny_{i}\in\mathbb{R}^{2Tm+3Tn} is the dual parameter of constraint 𝒜izi+izf,ihi+si=0\mathcal{A}_{i}z_{i}+\mathcal{B}_{i}z_{f,i}-h_{i}+s_{i}=0 and ρi\rho_{i}\in\mathbb{R} is the penalty parameter. Then, the first-order optimality conditions at a stationary solution (zik,zf,ik,sik,yik)(z_{i}^{k},z_{f,i}^{k},s_{i}^{k},y_{i}^{k}) are given by

0Ji(zik)+𝒵i(zik)+𝒜iyik\displaystyle 0\in\nabla J_{i}(z_{i}^{k})+\mathbb{N}_{\mathcal{Z}_{i}}(z_{i}^{k})+\mathcal{A}_{i}^{\top}y_{i}^{k} (10a)
0iyik++2T(m+n)(zf,ik)\displaystyle 0\in\mathcal{B}_{i}^{\top}y_{i}^{k}+\mathbb{N}_{\mathbb{R}_{+}^{2T(m+n)}}\left(z_{f,i}^{k}\right)
λik+βksik+yik=0\displaystyle\lambda_{i}^{k}+\beta^{k}s_{i}^{k}+y_{i}^{k}=0
𝒜izik+izf,ikhi+sik=0,\displaystyle\mathcal{A}_{i}z_{i}^{k}+\mathcal{B}_{i}z_{f,i}^{k}-h_{i}+s_{i}^{k}=0,

where 𝒵i(zik)\mathbb{N}_{\mathcal{Z}_{i}}(z_{i}^{k}) and +2T(m+n)(zf,ik)\mathbb{N}_{\mathbb{R}_{+}^{2T(m+n)}}\left(z_{f,i}^{k}\right) denote the normal cone of the set 𝒵i\mathcal{Z}_{i} and +2T(m+n)\mathbb{R}_{+}^{2T(m+n)} with respect to zikz_{i}^{k} and zf,ikz_{f,i}^{k}, respectively, and the superscript ()k(\cdot)^{k} means the variable in the kkth iteration.

Remark 2.

A solution that satisfies the optimality conditions (10a) may not necessarily satisfy the primal feasibility of the constraint si=0s_{i}=0. However, ALM can prompt the slack variable sis_{i} to zero through updating its dual parameter λi\lambda_{i}.

Based on Remark 2, the update of λi\lambda_{i} helps to prompt sis_{i} to zero with λik+1=λik+βiksik\lambda_{i}^{k+1}=\lambda_{i}^{k}+\beta_{i}^{k}s_{i}^{k}.

IV Hierarchical ADMM

For the problem (III-B), we treat it into two groups. The first one is to solve the problem (III-B) and update zi,zf,i,si,yiz_{i},z_{f,i},s_{i},y_{i} in sequence with keeping λi\lambda_{i} and βi\beta_{i} constant, and the other group is to update the parameter λi\lambda_{i} and βi\beta_{i}. In order to clarify the two groups, the iteration number of the first group (inner iteration) is indexed by rr, and the one of the second group (outer iteration) is kk.

1. Update ziz_{i}:

zik,r+1\displaystyle z_{i}^{k,r+1} =\displaystyle= argminziρik(zi,zf,ik,r,sik,r;yik,r,λik,ρik,βik)\displaystyle\underset{z_{i}}{\operatorname{argmin}}\;\mathcal{L}_{\rho_{i}^{k}}\left(z_{i},z_{f,i}^{k,r},s_{i}^{k,r};y_{i}^{k,r},\lambda_{i}^{k},\rho_{i}^{k},\beta_{i}^{k}\right)
=\displaystyle= argminzi𝒵iJi(zi)+ρik2𝒜izi+izf,ik,r\displaystyle\underset{z_{i}\in\mathcal{Z}_{i}}{\operatorname{argmin}}\;J_{i}(z_{i})+\frac{\rho_{i}^{k}}{2}\left\|\mathcal{A}_{i}z_{i}+\mathcal{B}_{i}z_{f,i}^{k,r}\right.
+sik,rhi+yik,rρik2.\yesnumber\displaystyle\left.\quad\quad+s_{i}^{k,r}-h_{i}+\frac{y_{i}^{k,r}}{\rho_{i}^{k}}\right\|^{2}.\yesnumber

2. Update zf,iz_{f,i}:

zf,ik,r+1\displaystyle z_{f,i}^{k,r+1} =\displaystyle= argminzf,iρik(zik,r+1,zf,i,sik,r;yik,r,λik,ρik,βik)\displaystyle\underset{z_{f,i}}{\operatorname{argmin}}\;\mathcal{L}_{\rho_{i}^{k}}\left(z_{i}^{k,r+1},z_{f,i},s_{i}^{k,r};y_{i}^{k,r},\lambda_{i}^{k},\rho_{i}^{k},\beta_{i}^{k}\right)
=\displaystyle= argminzf,iδ+2T(m+n)(zf,i)+ρik2𝒜izik,r+1\displaystyle\underset{z_{f,i}}{\operatorname{argmin}}\;\delta_{\mathbb{R}_{+}^{2T(m+n)}}(z_{f,i})+\frac{\rho_{i}^{k}}{2}\left\|\mathcal{A}_{i}z_{i}^{k,r+1}\right.
+izf,i+sik,rhi+yik,rρik2.\yesnumber\displaystyle\left.\quad+\mathcal{B}_{i}z_{f,i}+s_{i}^{k,r}-h_{i}+\frac{y_{i}^{k,r}}{\rho_{i}^{k}}\right\|^{2}.\yesnumber

In order to obtain a close-form solution to the problem (IV), we define 𝒯ik𝕊2T(m+n)\mathcal{T}_{i}^{k}\in\mathbb{S}^{2T(m+n)} as a given positive semidefinite linear operator. Then, it follows that

zf,ik,r+1\displaystyle z_{f,i}^{k,r+1} =\displaystyle= argminzf,iδ+2T(m+n)(zf,i)+12zf,izf,ik,r𝒯ik2\displaystyle\arg\min\limits_{z_{f,i}}\delta_{\mathbb{R}_{+}^{2T(m+n)}}(z_{f,i})+\frac{1}{2}\left\|z_{f,i}-z_{f,i}^{k,r}\right\|^{2}_{\mathcal{T}_{i}^{k}}
+ρik2𝒜izik,r+1+izf,i+sik,rhi+yik,rρik2.\yesnumber\displaystyle+\frac{\rho_{i}^{k}}{2}\left\|\mathcal{A}_{i}z_{i}^{k,r+1}+\mathcal{B}_{i}z_{f,i}+s_{i}^{k,r}-h_{i}+\frac{y_{i}^{k,r}}{\rho_{i}^{k}}\right\|^{2}.\yesnumber

Then, the optimality condition to the sub-problem of zf,iz_{f,i} updating is given by

0\displaystyle 0 \displaystyle\in δ+2T(m+n)(zf,i)+ρkiTizf,i+𝒯ik(zf,izf,ik,r)\displaystyle\partial\delta_{\mathbb{R}_{+}^{2T(m+n)}}(z_{f,i})+\rho^{k}\mathcal{B}_{i}^{T}\mathcal{B}_{i}z_{f,i}+\mathcal{T}_{i}^{k}\left(z_{f,i}-z_{f,i}^{k,r}\right)
+ρkiT(𝒜izik,r+1+sik,rhi+yik,rρik).\yesnumber\displaystyle+\rho^{k}\mathcal{B}_{i}^{T}\left(\mathcal{A}_{i}z_{i}^{k,r+1}+s_{i}^{k,r}-h_{i}+\frac{y_{i}^{k,r}}{\rho_{i}^{k}}\right).\yesnumber

Setting 𝒯ik=αikI2T(m+n)ρikiTi\mathcal{T}_{i}^{k}=\alpha_{i}^{k}I_{2T(m+n)}-\rho_{i}^{k}\mathcal{B}_{i}^{T}\mathcal{B}_{i} where αik\alpha_{i}^{k} is the largest eigenvalue of the matrix ρikiTi\rho_{i}^{k}\mathcal{B}_{i}^{T}\mathcal{B}_{i}, we have

0\displaystyle 0 \displaystyle\in (δ+2T(m+n)+αikI2T(m+n))zf,i(αikIρikiTi)zf,ik,r\displaystyle\left(\partial\delta_{\mathbb{R}_{+}^{2T(m+n)}}+\alpha_{i}^{k}I_{2T(m+n)}\right)z_{f,i}-\left(\alpha_{i}^{k}I-\rho_{i}^{k}\mathcal{B}_{i}^{T}\mathcal{B}_{i}\right)z_{f,i}^{k,r}
+ρkiT(𝒜izik,r+1+sik,rhi+yik,rρik).\yesnumber\displaystyle+\rho^{k}\mathcal{B}_{i}^{T}\left(\mathcal{A}_{i}z_{i}^{k,r+1}+s_{i}^{k,r}-h_{i}+\frac{y_{i}^{k,r}}{\rho_{i}^{k}}\right).\yesnumber

Then, it follows that

zf,ik,r+1\displaystyle z_{f,i}^{k,r+1} =\displaystyle= (αik1δ+2T(m+n)+I2T(m+n))1\displaystyle\left({\alpha_{i}^{k}}^{-1}\partial\delta_{\mathbb{R}_{+}^{2T(m+n)}}+I_{2T(m+n)}\right)^{-1}
((I2T(m+n)αik1ρikiTi)zf,ik,r\displaystyle\Bigg{(}\left(I_{2T(m+n)}-{\alpha_{i}^{k}}^{-1}\rho_{i}^{k}\mathcal{B}_{i}^{T}\mathcal{B}_{i}\right)z_{f,i}^{k,r}
αik1ρkiT(𝒜izik,r+1+sik,rhi+yik,rρik)).\displaystyle\;-{\alpha_{i}^{k}}^{-1}\rho^{k}\mathcal{B}_{i}^{T}\left(\mathcal{A}_{i}z_{i}^{k,r+1}+s_{i}^{k,r}-h_{i}+\frac{y_{i}^{k,r}}{\rho_{i}^{k}}\right)\Bigg{)}.
\yesnumber\displaystyle\yesnumber
Lemma 1.

The projection operator Proj𝒞()\operatorname{Proj}_{\mathcal{C}}(\cdot) with respect to the convex set 𝒞\mathcal{C} can be expressed as

Proj𝒞()=(I+αδ𝒞)1,\displaystyle\operatorname{Proj}_{\mathcal{C}}(\cdot)=(I+\alpha\partial\delta_{\mathcal{C}})^{-1}, (11)

where \partial is the sub-differential operator, and α\alpha\in\mathbb{R} is any arbitrary number.

Proof.

Define a finite dimensional Euclidean space 𝒳\mathcal{X} equipped with an inner product and its induced norm such that 𝒞𝒳\mathcal{C}\subset\mathcal{X}. For any x𝒳x\in\mathcal{X}, there exists z𝒳z\in\mathcal{X} such that z(I+αδ𝒞)1(x)z\in(I+\alpha\partial\delta_{\mathcal{C}})^{-1}(x). Then, it follows that

x\displaystyle x \displaystyle\in (I+αδ𝒞)(z)=z+αδ𝒞.\yesnumber\displaystyle(I+\alpha\partial\delta_{\mathcal{C}})(z)=z+\alpha\partial\delta_{\mathcal{C}}.\yesnumber

The projection operator Proj𝒞(z)\operatorname{Proj}_{\mathcal{C}}(z) is

Proj𝒞(z)\displaystyle\operatorname{Proj}_{\mathcal{C}}(z) =\displaystyle= argmin𝑧{δ𝒞(x)+12αzx2}.\displaystyle\underset{z}{\operatorname{argmin}}\left\{\delta_{\mathcal{C}}(x)+\frac{1}{2\alpha}\|z-x\|^{2}\right\}. (12)

Due to the fact that the optimization problem (12) is strictly convex, the sufficient and necessary condition of (12) is

0αδ𝒞(x)+zx.\displaystyle 0\in\alpha\partial\delta_{\mathcal{C}}(x)+z-x. (13)

Since (13) is equivalent to (12) and the projection onto a convex set is unique, the operator (I+αδ𝒞)1(I+\alpha\partial\delta_{\mathcal{C}})^{-1} is a point-to-point mapping. This completes the proof of Lemma 1. ∎

Based on Lemma 1, zf,iz_{f,i} can be determined as

zf,ik,r+1\displaystyle z_{f,i}^{k,r+1} =Proj+2T(m+n){(I2T(m+n)αik1ρikiTi)zf,ik,r\displaystyle=\operatorname{Proj}_{\mathbb{R}_{+}^{2T(m+n)}}\Bigg{\{}\left(I_{2T(m+n)}-{\alpha_{i}^{k}}^{-1}\rho_{i}^{k}\mathcal{B}_{i}^{T}\mathcal{B}_{i}\right)z_{f,i}^{k,r}
αik1ρkiT(𝒜izik,r+1+sik,rhi+yik,rρik)},\yesnumber\displaystyle-{\alpha_{i}^{k}}^{-1}\rho^{k}\mathcal{B}_{i}^{T}\left(\mathcal{A}_{i}z_{i}^{k,r+1}+s_{i}^{k,r}-h_{i}+\frac{y_{i}^{k,r}}{\rho_{i}^{k}}\right)\Bigg{\}},\yesnumber

where the projection operator Proj+2T(m+n)(z)\operatorname{Proj}_{\mathbb{R}_{+}^{2T(m+n)}}(z) means the projection of zz onto a closed convex set +2T(m+n)\mathbb{R}_{+}^{2T(m+n)}.

Lemma 2.

With 𝒞=+n\mathcal{C}=\mathbb{R}_{+}^{n}, then for any znz\in\mathbb{R}^{n}, the projection of zz onto 𝒞\mathcal{C} is given by

Proj𝒞(z)=max{z,0}.\displaystyle\operatorname{Proj}_{\mathcal{C}}(z)=\max\{z,0\}. (14)

According to Lemma 2, zf,ik,r+1z_{f,i}^{k,r+1} can be easily obtained.

3. Update sis_{i}:

sik,r+1\displaystyle s_{i}^{k,r+1} =\displaystyle= argminsiρik(zik,r+1,zf,ik,r+1,si;yik,r,λik,ρik,βik)\displaystyle\underset{s_{i}}{\operatorname{argmin}}\;\mathcal{L}_{\rho_{i}^{k}}\left(z_{i}^{k,r+1},z_{f,i}^{k,r+1},s_{i};y_{i}^{k,r},\lambda_{i}^{k},\rho_{i}^{k},\beta_{i}^{k}\right)
=\displaystyle= ρikρik+βik(𝒜izik,r+1+izf,ik,r+1hi+yik,rρik)\displaystyle-\frac{\rho_{i}^{k}}{\rho_{i}^{k}+\beta_{i}^{k}}\left(\mathcal{A}_{i}z_{i}^{k,r+1}+\mathcal{B}_{i}z_{f,i}^{k,r+1}-h_{i}+\frac{y_{i}^{k,r}}{\rho_{i}^{k}}\right)
λikρik+βik.\yesnumber\displaystyle\;-\frac{\lambda_{i}^{k}}{\rho_{i}^{k}+\beta_{i}^{k}}.\yesnumber

4. Update yiy_{i}:

yik,r+1\displaystyle y_{i}^{k,r+1} =\displaystyle= argminyiρik(zik,r+1,zf,ik,r+1,sik,r+1;yi,λik,ρik,βik)\displaystyle\underset{y_{i}}{\operatorname{argmin}}\;\mathcal{L}_{\rho_{i}^{k}}\left(z_{i}^{k,r+1},z_{f,i}^{k,r+1},s_{i}^{k,r+1};y_{i},\lambda_{i}^{k},\rho_{i}^{k},\beta_{i}^{k}\right)
=\displaystyle= yik,r+ρik(𝒜izik,r+1+izf,ik,r+1+sik,r+1hi).\displaystyle y_{i}^{k,r}+\rho_{i}^{k}\left(\mathcal{A}_{i}z_{i}^{k,r+1}+\mathcal{B}_{i}z_{f,i}^{k,r+1}+s_{i}^{k,r+1}-h_{i}\right).
\yesnumber\displaystyle\yesnumber
Assumption 2.

Given λik\lambda_{i}^{k}, βik\beta_{i}^{k} and ρik\rho_{i}^{k}, the ziz_{i}-update can find a stationary solution zirz_{i}^{r} such that

0\displaystyle 0 \displaystyle\in ziρik(zir,zf,ir1,sir1,yir1)\displaystyle\partial_{z_{i}}\mathcal{L}_{\rho_{i}^{k}}(z_{i}^{r},z_{f,i}^{r-1},s_{i}^{r-1},y_{i}^{r-1})
ρik(zir,zf,ir1,sir1,yir1)\displaystyle\mathcal{L}_{\rho_{i}^{k}}(z_{i}^{r},z_{f,i}^{r-1},s_{i}^{r-1},y_{i}^{r-1}) \displaystyle\leq ρik(zir1,zf,ir1,sir1,yir1),\displaystyle\mathcal{L}_{\rho_{i}^{k}}(z_{i}^{r-1},z_{f,i}^{r-1},s_{i}^{r-1},y_{i}^{r-1}),
\yesnumber\displaystyle\yesnumber

for all a+a\in\mathbb{Z}_{+}.

Lemma 3.

For all a+a\in\mathbb{Z}_{+}, the following conditions holds:

iyir1+ρii(𝒜izir+izf,ir+sir1hi),\displaystyle\left\langle\mathcal{B}_{i}^{\top}y_{i}^{r-1}+\rho_{i}\mathcal{B}_{i}^{\top}\left(\mathcal{A}_{i}z_{i}^{r}+\mathcal{B}_{i}z_{f,i}^{r}+s_{i}^{r-1}-h_{i}\right),\right.
(zf,izf,ir)0,zf,i+2T(m+n)\displaystyle\left.(z_{f,i}-z_{f,i}^{r})\right\rangle\geq 0,\;\forall z_{f,i}\in\mathbb{R}_{+}^{2T(m+n)}
λi+βisir+yir=0.\yesnumber\displaystyle\lambda_{i}+\beta_{i}s_{i}^{r}+y_{i}^{r}=0.\yesnumber
Proof.

This lemma can be straightforwardly derived based on the optimality conditions of ziz_{i}- and zf,iz_{f,i}-updates, i.e., (III-B) and (III-B), and thus the proof is omitted. ∎

Theorem 1.

With ρik2βik\rho_{i}^{k}\geq\sqrt{2}\beta_{i}^{k}, the inner iteration of problem (III-B) converges to the stationary point (zik,zf,ik,sik,yik)(z_{i}^{k},z_{f,i}^{k},s_{i}^{k},y_{i}^{k}) of the transformed problem (III-B).

Proof.

With Assumption 2, we have

ρik(zir1,zf,ir1,sir1,yir1)ρik(zir,zf,ir1,sir1,yir1).\displaystyle\mathcal{L}_{\rho_{i}^{k}}\left(z_{i}^{r-1},z_{f,i}^{r-1},s_{i}^{r-1},y_{i}^{r-1}\right)\geq\mathcal{L}_{\rho_{i}^{k}}\left(z_{i}^{r},z_{f,i}^{r-1},s_{i}^{r-1},y_{i}^{r-1}\right).
\yesnumber\displaystyle\yesnumber

Besides, we also have

ρik(zir,zf,ir1,sir1,yir1)ρik(zir,zf,ir,sir1,yir1)\displaystyle\mathcal{L}_{\rho_{i}^{k}}\left(z_{i}^{r},z_{f,i}^{r-1},s_{i}^{r-1},y_{i}^{r-1}\right)-\mathcal{L}_{\rho_{i}^{k}}\left(z_{i}^{r},z_{f,i}^{r},s_{i}^{r-1},y_{i}^{r-1}\right)
=\displaystyle= iyir1+ρiki(𝒜izir+izf,ir+sir1hi),\displaystyle\Big{\langle}\mathcal{B}_{i}^{\top}y_{i}^{r-1}+\rho_{i}^{k}\mathcal{B}_{i}^{\top}\left(\mathcal{A}_{i}z_{i}^{r}+\mathcal{B}_{i}z_{f,i}^{r}+s_{i}^{r-1}-h_{i}\right),
zf,ir1zf,ir+ρik2i(zf,ir1zf,ir)2\displaystyle\quad z_{f,i}^{r-1}-z_{f,i}^{r}\Big{\rangle}+\frac{\rho_{i}^{k}}{2}\left\|\mathcal{B}_{i}(z_{f,i}^{r-1}-z_{f,i}^{r})\right\|^{2}
\displaystyle\geq ρik2i(zf,ir1zf,ir)2.\yesnumber\displaystyle\frac{\rho_{i}^{k}}{2}\left\|\mathcal{B}_{i}(z_{f,i}^{r-1}-z_{f,i}^{r})\right\|^{2}.\yesnumber

Note that the equality in (IV) holds because of the following claim:

𝒜izir+sir1hi+izf,ir12𝒜izir+sir1hi+izf,ir2\displaystyle\left\|\mathcal{A}_{i}z_{i}^{r}+s_{i}^{r-1}-h_{i}+\mathcal{B}_{i}z_{f,i}^{r-1}\right\|^{2}-\left\|\mathcal{A}_{i}z_{i}^{r}+s_{i}^{r-1}-h_{i}+\mathcal{B}_{i}z_{f,i}^{r}\right\|^{2}
=2𝒜izir+sir1hi+izf,ir,izf,ir1izf,ir\displaystyle=2\left\langle\mathcal{A}_{i}z_{i}^{r}+s_{i}^{r-1}-h_{i}+\mathcal{B}_{i}z_{f,i}^{r},\mathcal{B}_{i}z_{f,i}^{r-1}-\mathcal{B}_{i}z_{f,i}^{r}\right\rangle
+izf,ir1izf,ir2.\yesnumber\displaystyle\quad+\left\|\mathcal{B}_{i}z_{f,i}^{r-1}-\mathcal{B}_{i}z_{f,i}^{r}\right\|^{2}.\yesnumber

In addition, the inequality in (IV) can be obtained from Lemma 3. Now, the descent of ziz_{i} and zf,iz_{f,i} updates has been proved. Next, we will show the descent of sis_{i} and yiy_{i} updates.

ρik(zir,zf,ir,sir1,yir1)ρik(zir,zf,ir,sir,yir)\displaystyle\mathcal{L}_{\rho_{i}^{k}}\left(z_{i}^{r},z_{f,i}^{r},s_{i}^{r-1},y_{i}^{r-1}\right)-\mathcal{L}_{\rho_{i}^{k}}\left(z_{i}^{r},z_{f,i}^{r},s_{i}^{r},y_{i}^{r}\right)
=\displaystyle= λik,(sir1sir)+βik2(sir12sir2)\displaystyle\left\langle\lambda_{i}^{k},(s_{i}^{r-1}-s_{i}^{r})\right\rangle+\frac{\beta_{i}^{k}}{2}\left(\left\|s_{i}^{r-1}\|^{2}-\|s_{i}^{r}\right\|^{2}\right)
+yir,sir1sir+ρik2sir1sir2\displaystyle\>+\left\langle y_{i}^{r},s_{i}^{r-1}-s_{i}^{r}\right\rangle+\frac{\rho_{i}^{k}}{2}\left\|s_{i}^{r-1}-s_{i}^{r}\right\|^{2}
ρik𝒜izir+izf,ir+sirhi2\displaystyle\>-\rho_{i}^{k}\left\|\mathcal{A}_{i}z_{i}^{r}+\mathcal{B}_{i}z_{f,i}^{r}+s_{i}^{r}-h_{i}\right\|^{2}
\displaystyle\geq (ρik2βik2ρik)sir1sir2.\yesnumber\displaystyle\left(\frac{\rho_{i}^{k}}{2}-\frac{{\beta_{i}^{k}}^{2}}{\rho_{i}^{k}}\right)\left\|s_{i}^{r-1}-s_{i}^{r}\right\|^{2}.\yesnumber

Remarkably, the first equality holds due to (IV) and the following derivation:

ρik𝒜izir+izf,irhi+sir,𝒜izir+izf,irhi+sir1\displaystyle-\rho_{i}^{k}\langle\mathcal{A}_{i}z_{i}^{r}+\mathcal{B}_{i}z_{f,i}^{r}-h_{i}+s_{i}^{r},\mathcal{A}_{i}z_{i}^{r}+\mathcal{B}_{i}z_{f,i}^{r}-h_{i}+s_{i}^{r-1}\rangle
+ρik2(𝒜izir+izf,irhi+sir12\displaystyle+\frac{\rho_{i}^{k}}{2}\Big{(}\left\|\mathcal{A}_{i}z_{i}^{r}+\mathcal{B}_{i}z_{f,i}^{r}-h_{i}+s_{i}^{r-1}\right\|^{2}
𝒜izir+izf,irhi+sir2)\displaystyle\quad-\left\|\mathcal{A}_{i}z_{i}^{r}+\mathcal{B}_{i}z_{f,i}^{r}-h_{i}+s_{i}^{r}\right\|^{2}\Big{)}
=\displaystyle= ρik2sir1sir2ρik𝒜izir+izf,irhi+sir2.\yesnumber\displaystyle\frac{\rho_{i}^{k}}{2}\left\|s_{i}^{r-1}-s_{i}^{r}\right\|^{2}-\rho_{i}^{k}\left\|\mathcal{A}_{i}z_{i}^{r}+\mathcal{B}_{i}z_{f,i}^{r}-h_{i}+s_{i}^{r}\right\|^{2}.\yesnumber

To prove the inequality in (IV), we define a function f(si)=λir1,si+βi2si2f(s_{i})=\langle\lambda_{i}^{r-1},s_{i}\rangle+\frac{\beta_{i}}{2}\|s_{i}\|^{2}. The gradient of f(si)f(s_{i}) is f(sir)=λir1+βisir=yir\nabla f(s_{i}^{r})=\lambda_{i}^{r-1}+\beta_{i}s_{i}^{r}=-y_{i}^{r} due to (III-B). It follows that f(sir1)f(sir)+(yir)(sir1sir)0f(s_{i}^{r-1})-f(s_{i}^{r})+(y_{i}^{r})^{\top}(s_{i}^{r-1}-s_{i}^{r})\geq 0, as f(si)f(s_{i}) is convex. Therefore, the inequality in (IV) holds.

Under the condition ρi2βi\rho_{i}\geq\sqrt{2}\beta_{i}, the descent of ρik\mathcal{L}_{\rho_{i}^{k}} with updating (zik,zf,ik,sik,yik)(z_{i}^{k},z_{f,i}^{k},s_{i}^{k},y_{i}^{k}) can be proved, which means the augmented Lagrangian function is sufficient descent.

Since the function f(si)f(s_{i}), a part of the augmented Lagrangian function, is Lipschitz differentiable with βi\beta_{i}, and λi\lambda_{i} is bounded by [λi¯,λi¯][\underline{\lambda_{i}},\overline{\lambda_{i}}], it is obvious that ρik(zir,zf,ir,sir,yir)\mathcal{L}_{\rho_{i}^{k}}\left(z_{i}^{r},z_{f,i}^{r},s_{i}^{r},y_{i}^{r}\right) is lower bounded [28]. Thus, the solution of the inner loop of this transformed problem (III-B) will converge to the stationary point (zik,zf,ik,sik,yik)(z_{i}^{k},z_{f,i}^{k},s_{i}^{k},y_{i}^{k}). This completes the proof of Theorem 1. ∎

5. Update λi\lambda_{i} and βi\beta_{i}: Then, in the outer iterations, the dual variable λik\lambda_{i}^{k} is updated. In order to drive the slack variable sis_{i} to zero, the penalty parameter βik\beta_{i}^{k} is multiplied by a ratio γ>1\gamma>1 if the currently computed siks_{i}^{k} from the inner iterations does not decrease enough from the previous outer iteration sik1s_{i}^{k-1}, i.e., sik>ωsik1,ω(0,1)\|s_{i}^{k}\|>\omega\|s_{i}^{k-1}\|,\omega\in(0,1).

λik+1\displaystyle\lambda_{i}^{k+1} =\displaystyle= Proj[λ¯i,λ¯i](λik+βiksik)\displaystyle\operatorname{Proj}_{[\underline{\lambda}_{i},\overline{\lambda}_{i}]}(\lambda_{i}^{k}+\beta_{i}^{k}s_{i}^{k})
βk+1\displaystyle\beta^{k+1} =\displaystyle= {γβik,sik>ωsik1βik,sikωsik1,\yesnumber\displaystyle\begin{cases}\gamma\beta_{i}^{k},&\|s_{i}^{k}\|>\omega\|s_{i}^{k-1}\|\\ \beta_{i}^{k},&\|s_{i}^{k}\|\leq\omega\|s_{i}^{k-1}\|,\end{cases}\yesnumber

where the projection operator Proj[λ¯i,λ¯i](λik+βiksik)\operatorname{Proj}_{[\underline{\lambda}_{i},\overline{\lambda}_{i}]}(\lambda_{i}^{k}+\beta_{i}^{k}s_{i}^{k}) means the projection of λik+βiksik\lambda_{i}^{k}+\beta_{i}^{k}s_{i}^{k} onto a closed convex set [λ¯i,λ¯i][\underline{\lambda}_{i},\overline{\lambda}_{i}].

Lemma 4.

Given a hypercube set [a,b][a,b] with a,bna,b\in\mathbb{R}^{n} and a<ba<b, for any znz\in\mathbb{R}^{n}, the projection of zz onto the hypercube [a,b][a,b] is given by

Proj[a,b](z)\displaystyle~{}\operatorname{Proj}_{[a,b]}(z) =\displaystyle= {z(i),a(i)z(i)b(i)a(i),z(i)<a(i)b(i),z(i)>b(i),\displaystyle\begin{cases}z_{(i)},&a_{(i)}\leq z_{(i)}\leq b_{(i)}\\ a_{(i)},&z_{(i)}<a_{(i)}\\ b_{(i)},&z_{(i)}>b_{(i)},\\ \end{cases} (15)

where the subscript (i)\cdot_{(i)} means the iith component in vector zz, aa and bb.

Theorem 2.

Suppose that the point (zik,zf,ik,sik,yik)(z_{i}^{k},z_{f,i}^{k},s_{i}^{k},y_{i}^{k}) satisfies the optimality condition (10a). For any ϵ1,ϵ2,ϵ3>0\epsilon_{1},\epsilon_{2},\epsilon_{3}>0, after finite outer iterations kk, appropriate stationary point (zik,zf,ik,sik,yik)(z_{i}^{k},z_{f,i}^{k},s_{i}^{k},y_{i}^{k}) can be found such that the optimality condition of the original problem (III-A) before introducing the slack variable sis_{i} is also satisfied [28], i.e.,

0Ji(zik)+𝒵i(zik)+𝒜iyik\displaystyle 0\in\nabla J_{i}(z_{i}^{k})+\mathbb{N}_{\mathcal{Z}_{i}}(z_{i}^{k})+\mathcal{A}_{i}^{\top}y_{i}^{k} (16a)
0iyik++2T(m+n)(zf,ik)\displaystyle 0\in\mathcal{B}_{i}^{\top}y_{i}^{k}+\mathbb{N}_{\mathbb{R}_{+}^{2T(m+n)}}\left(z_{f,i}^{k}\right)
𝒜izik+izf,ikhi=0.\displaystyle\mathcal{A}_{i}z_{i}^{k}+\mathcal{B}_{i}z_{f,i}^{k}-h_{i}=0.
Proof.

Assume that (zik,zf,ik,sik,yik)(z_{i}^{k},z_{f,i}^{k},s_{i}^{k},y_{i}^{k}) converges to the stationary point (zi,zf,i,si,yi)(z_{i}^{\star},z_{f,i}^{\star},s_{i}^{\star},y_{i}^{\star}). We have zi𝒵i,zf,i+2T(m+n)z_{i}^{\star}\in\mathcal{Z}_{i},z_{f,i}\in\mathbb{R}_{+}^{2T(m+n)} and 𝒜izi+izf,ihi+si=0\mathcal{A}_{i}z_{i}^{\star}+\mathcal{B}_{i}z_{f,i}^{\star}-h_{i}+s_{i}^{\star}=0.

For the case where βik\beta_{i}^{k} is bounded, we have zik0z_{i}^{k}\rightarrow 0 and zi=0z_{i}^{\star}=0. For the case where βik\beta_{i}^{k} is not bounded, based on the optimality condition (10a), it follows that

λikβik+sik,r+yik,rβik=0.\displaystyle\frac{{\lambda_{i}}^{k}}{\beta_{i}^{k}}+s_{i}^{k,r}+\frac{{y_{i}}^{k,r}}{\beta_{i}^{k}}=0. (18)

Then, we have si=0s_{i}^{\star}=0 by taking the limitation of (18) to zero, as yikyiy_{i}^{k}\rightarrow y_{i}^{\star} and λik[λ¯i,λ¯i]\lambda_{i}^{k}\in[\underline{\lambda}_{i},\overline{\lambda}_{i}]. Since we have si=0s_{i}^{\star}=0 in both of the two cases, the optimality condition with respect to yiy_{i}, i.e., (2), can be satisfied. Besides, the optimality condition with respect to ziz_{i} and zf,iz_{f,i}, i.e., (16a) and (2), can also be satisfied by taking k0k\rightarrow 0 on (10a) and (III-B), respectively. This completes the proof of Theorem 2. ∎

Remark 3.

Our algorithm explicitly demonstrates how to apply the hierarchical ADMM in the multi-agent system with the presence of nonconvex collision-avoidance constraints. The closed-form solution of every variable is provided. In this approach, a proximal term is introduced to the two-level algorithm presented in [28] and a more compact convergence condition is proposed, i.e., ρi2βi\rho_{i}\geq\sqrt{2}\beta_{i}. Therefore, the Theorem 1 in our paper proves the convergence of the inner iteration under the condition ρi2βi\rho_{i}\geq\sqrt{2}\beta_{i}. Theorem 2 is proved under different assumptions, compared with the algorithm proposed in [28]. Thus, we believe our algorithm and theorems provide more details about the hierarchical ADMM algorithms for the multi-gent systems.

The stopping criterion of this ADMM algorithm is given by

ρik𝒜iT(izf,ir+sirizf,ir+1sir+1)ϵ1k\displaystyle\left\|\rho_{i}^{k}\mathcal{A}_{i}^{T}\left(\mathcal{B}_{i}z_{f,i}^{r}+s_{i}^{r}-\mathcal{B}_{i}z_{f,i}^{r+1}-s_{i}^{r+1}\right)\right\|\leq\epsilon_{1}^{k} (19a)
ρikiT(sirsir+1)ϵ2k\displaystyle\left\|\rho_{i}^{k}\mathcal{B}_{i}^{T}\left(s_{i}^{r}-s_{i}^{r+1}\right)\right\|\leq\epsilon_{2}^{k}
𝒜izir+1+izf,ir+1hi+sir+1ϵ3k,\displaystyle\left\|\mathcal{A}_{i}z_{i}^{r+1}+\mathcal{B}_{i}z_{f,i}^{r+1}-h_{i}+s_{i}^{r+1}\right\|\leq\epsilon_{3}^{k},\textbf{}

where ϵek0,e{1,2,3}\epsilon_{e}^{k}\rightarrow 0,\forall e\in\{1,2,3\}.

To summarize the above developments, the pseudocode of this algorithm is presented in Algorithm 1. After obtaining each agent’s control input and executing the control input, the agents will communicate with their neighbors to obtain the updated position information of their neighbors. Then, this process will be repeated until finishing the designed task.

Algorithm 1 Hierarchical ADMM for Multi-Agent Decision Making
  Initialize zi0T(m+n),zf,i02T(m+n),si02Tm+3Tnz_{i}^{0}\in\mathbb{R}^{T(m+n)},z_{f,i}^{0}\in\mathbb{R}^{2T(m+n)},s_{i}^{0}\in\mathbb{R}^{2Tm+3Tn}; initialize the dual variable λi0[λ¯i,λ¯i]\lambda_{i}^{0}\in[\underline{\lambda}_{i},\overline{\lambda}_{i}], where λ¯i,λ¯i2Tm+3Tn\underline{\lambda}_{i},\overline{\lambda}_{i}\in\mathbb{R}^{2Tm+3Tn}, and λ¯i>λ¯i\overline{\lambda}_{i}>\underline{\lambda}_{i}; initialize the penalty parameter βi0>0\beta_{i}^{0}>0, and the constant ω[0,1)\omega\in[0,1) and γ>1\gamma>1; initialize the sequence of tolerance {ϵ1k,ϵ2k,ϵ3k}+\{\epsilon_{1}^{k},\epsilon_{2}^{k},\epsilon_{3}^{k}\}\in\mathbb{R}_{+} with limkϵek=0,e{1,2,3}\lim_{k\rightarrow\infty}\epsilon_{e}^{k}=0,\forall e\in\{1,2,3\}. Given the terminated tolerances ϵterm,e,e={1,2,3}\epsilon_{\text{term},e},e=\{1,2,3\}, if ϵekϵterm,e\epsilon_{e}^{k}\leq\epsilon_{\text{term},e}, ϵek=ϵterm,i\epsilon_{e}^{k}=\epsilon_{\text{term},i} for e={1,2,3}e=\{1,2,3\}.
  Set the outer iteration k=0k=0.
  while Stopping criterion is violated do
     Given λik\lambda_{i}^{k} and βik\beta_{i}^{k}, initialize zik,0,zf,ik,0,sik,0z_{i}^{k,0},z_{f,i}^{k,0},s_{i}^{k,0} and yik,0y_{i}^{k,0} such that λik+βiksik,0+yik,0=0\lambda_{i}^{k}+\beta_{i}^{k}s_{i}^{k,0}+y_{i}^{k,0}=0; initialize tolerance (ϵ1k,ϵ2k,ϵ3k)(\epsilon_{1}^{k},\epsilon_{2}^{k},\epsilon_{3}^{k}); ρik2βik\rho_{i}^{k}\geq\sqrt{2}\beta_{i}^{k}.
     Set the inner iteration r=0r=0.
     while Stopping criterion is violated do
        Update zik,r+1z_{i}^{k,r+1} by (IV) via nonlinear programming solvers.
        Update zf,ik,r+1z_{f,i}^{k,r+1}, sik,r+1s_{i}^{k,r+1}, yik,r+1y_{i}^{k,r+1} by (IV), (IV), (IV), respectively.
        r=r+1r=r+1.
     end while
     Set zik+1,zf,ik+1,sik+1z_{i}^{k+1},z_{f,i}^{k+1},s_{i}^{k+1}, yik+1y_{i}^{k+1} to be zik,r+1,zf,ik,r+1,sik,r+1z_{i}^{k,r+1},z_{f,i}^{k,r+1},s_{i}^{k,r+1}, yik,r+1y_{i}^{k,r+1}, respectively.
     Update λik+1\lambda_{i}^{k+1} and βik+1\beta_{i}^{k+1} by (IV).
     k=k+1k=k+1.
  end while

V Improved Hierarchical ADMM

In the hierarchical ADMM approach, the minimization of ziz_{i} in each iteration is computationally expensive, and it requires nonlinear programming (NLP) solvers. Therefore, the improved minimization of ziz_{i} is required to improve the computational efficiency by approximating and solving this inexact minimization problem.

The original ziz_{i}-minimization problem can be expressed as

minzi\displaystyle\underset{z_{i}}{\operatorname{min}} Ji(zi)+ρik2𝒜izi+izf,ik,r+sik,rhi+yik,rρik2\displaystyle J_{i}(z_{i})+\frac{\rho_{i}^{k}}{2}\left\|\mathcal{A}_{i}z_{i}+\mathcal{B}_{i}z_{f,i}^{k,r}+s_{i}^{k,r}-h_{i}+\frac{y_{i}^{k,r}}{\rho_{i}^{k}}\right\|^{2}
s.t.\displaystyle\operatorname{s.t.} ϕ(zi;[pj])0\displaystyle\phi\left(z_{i};[p_{j}]\right)\leq 0
ψ(zi;[pj])0,\yesnumber\displaystyle\psi\left(z_{i};[p_{j}]\right)\leq 0,\yesnumber

where

ϕ(zi;[pj])\displaystyle\phi\left(z_{i};[p_{j}]\right) =\displaystyle= (zi;[pj])𝒫(zi;[pj])dcmu2\displaystyle\mathcal{M}\left(z_{i};[p_{j}]\right)\mathcal{P}\left(z_{i};[p_{j}]\right)-d_{\mathrm{cmu}}^{2}
ψ(zi;[pj])\displaystyle\psi\left(z_{i};[p_{j}]\right) =\displaystyle= dsafe2(zi;[pj])𝒫(zi;[pj]).\displaystyle d_{\mathrm{safe}}^{2}-\mathcal{M}\left(z_{i};[p_{j}]\right)\mathcal{P}\left(z_{i};[p_{j}]\right).

The KKT conditions of this ziz_{i}-minimization problem (V) are

0\displaystyle 0 =\displaystyle= Ji(zi)+ρik𝒜i(𝒜izi+izf,ik,r+sik,rhi+yik,rρik)\displaystyle\nabla J_{i}(z_{i})+{\rho_{i}^{k}}\mathcal{A}_{i}^{\top}\left(\mathcal{A}_{i}z_{i}+\mathcal{B}_{i}z_{f,i}^{k,r}+s_{i}^{k,r}-h_{i}+\frac{y_{i}^{k,r}}{\rho_{i}^{k}}\right)
+θ=1θϕμθ,iϕθ(zi;[pj])+θ=1θψκθ,iψθ(zi;[pj])\displaystyle+\sum\limits_{\theta=1}^{\theta_{\phi}}\mu_{\theta,i}\nabla\phi_{\theta}\left(z_{i};[p_{j}]\right)+\sum\limits_{\theta=1}^{\theta_{\psi}}\kappa_{\theta,i}\nabla\psi_{\theta}\left(z_{i};[p_{j}]\right)
0\displaystyle 0 =\displaystyle= μθ,iϕθ(zi;[pj]),θ=1,2,,θϕ\displaystyle\mu_{\theta,i}\phi_{\theta}\left(z_{i};[p_{j}]\right),\>\forall\theta=1,2,\cdots,\theta_{\phi}
0\displaystyle 0 =\displaystyle= κθ,iψθ(zi;[pj]),θ=1,2,,θψ.\yesnumber\displaystyle\kappa_{\theta,i}\psi_{\theta}\left(z_{i};[p_{j}]\right),\>\forall\theta=1,2,\cdots,\theta_{\psi}.\yesnumber

Notice that the hierarchical ADMM algorithm requires complete minimization of ziz_{i} in each inner iteration. However, this is neither desirable due to the computational cost nor practical, as any NLP solver finds only a point that approximately satisfies the KKT optimality conditions. In the hierarchical ADMM algorithm, we propose the use of the interior-point method to solve the ziz_{i} minimization problem (V). The interior-point method employs double-layer iterations to find the solution. In the outer iteration, a barrier technique is used to convert the inequality constraints into an additional term in the objective; the stationary points of resulting barrier problems converge to true stationary points as the barrier parameter converges to 0. In the inner iteration, a proper search method is used to obtain the optimum of the barrier problem. Since both the interior-point algorithm and the hierarchical ADMM algorithm have a double-layer structure, we consider matching these two layers.

In order to accelerate the computation process of solving (V), the barrier method can be employed to transform the inequality constraints to additional terms in the objective function. For example, in the kk-th outer loop, the function Ji(zi)J_{i}(z_{i}) can be added with a barrier term bikθϕln(ϕ(zi))-b_{i}^{k}\sum_{\theta_{\phi}}\ln(-\phi(z_{i})) based on the barrier method, where bikb_{i}^{k} is the barrier parameter with limkbik=0\lim\limits_{k\rightarrow\infty}b_{i}^{k}=0. Hence, the barrier augmented Lagrangian in the kk-th outer loop can be written as

^bik=ρikbikθ=1θϕln(ϕθ(zi)).\displaystyle{\hat{\mathcal{L}}}_{b_{i}^{k}}=\mathcal{L}_{\rho_{i}^{k}}-b_{i}^{k}\sum\limits_{\theta=1}^{\theta_{\phi}}\ln(-\phi_{\theta}(z_{i})). (22)

If the ziz_{i}-minimization step returns zik,r+1z_{i}^{k,r+1} by minimizing ^bi\hat{\mathcal{L}}_{b_{i}} with respect to ziz_{i}, then the inner iterations lead to the decrease of ^bik\hat{\mathcal{L}}_{b^{k}_{i}}, which also matches the KKT conditions of the original optimization problem. Therefore, the outer iterations can find an approximate stationary point of the original problem with the decrease of the barrier parameter bikb_{i}^{k}. Note that only ψ(zi;[pj])\psi(z_{i};[p_{j}]) is used as the inequality constraint for the purpose of demonstration.

Therefore, it remains to minimize ^bi\hat{\mathcal{L}}_{b_{i}}, which is given by

^bik\displaystyle\hat{\mathcal{L}}_{b_{i}^{k}} =\displaystyle= Ji(zi)+ρik2𝒜izi+izf,ik,r+sik,rhi+yik,rρik2\displaystyle J_{i}(z_{i})+\frac{\rho_{i}^{k}}{2}\left\|\mathcal{A}_{i}z_{i}+\mathcal{B}_{i}z_{f,i}^{k,r}+s_{i}^{k,r}-h_{i}+\frac{y_{i}^{k,r}}{\rho_{i}^{k}}\right\|^{2}
bik1Tri,ln(h(zi;[pj])dsafe2),\yesnumber\displaystyle-\left\langle b_{i}^{k}1_{Tr_{i}},\ln\left(h\left(z_{i};[p_{j}]\right)-d_{\mathrm{safe}}^{2}\right)\right\rangle,\yesnumber

where h(zi;[pj])=(zi;[pj])𝒫(zi;[pj])h\left(z_{i};[p_{j}]\right)={\mathcal{M}\left(z_{i};[p_{j}]\right)\mathcal{P}\left(z_{i};[p_{j}]\right)}.

Lemma 5.

The gradient of h(zi;[pj])h\left(z_{i};[p_{j}]\right) with respect to ziz_{i} is a linear mapping with respect to ziz_{i}.

Proof.

Define W=ITri1npTrinp×TriW=I_{Tr_{i}}\otimes 1_{n_{p}}\in\mathbb{R}^{Tr_{i}n_{p}\times Tr_{i}}, ω=(IT1riInp)Mpzi1T[pj]Trinp\omega=(I_{T}\otimes 1_{r_{i}}\otimes I_{n_{p}})M_{p}z_{i}-1_{T}\otimes[p_{j}]\in\mathbb{R}^{Tr_{i}n_{p}}, Ω=diag(ω)Trinp×Trinp\Omega=\operatorname{diag}(\omega)\in\mathbb{R}^{Tr_{i}n_{p}\times Tr_{i}n_{p}}. Then,

h(zi;[pj])=W(ITrinp(ω1Trinp))ω=WΩω.\displaystyle h\left(z_{i};[p_{j}]\right)=W^{\top}(I_{Tr_{i}n_{p}}\odot(\omega 1_{Tr_{i}n_{p}}^{\top}))\omega=W^{\top}\Omega\omega.

Since Ω=ITrinp(ω1Trinp)=ITrinp(𝟏Trinpω)\Omega=I_{Tr_{i}n_{p}}\odot(\omega 1_{Tr_{i}n_{p}}^{\top})=I_{Tr_{i}n_{p}}\odot(\mathbf{1}_{Tr_{i}n_{p}}\omega^{\top}), we have

dh(zi;[pj])\displaystyle\textup{d}h\left(z_{i};[p_{j}]\right) =\displaystyle= 2WΩ(IT1riInp)Mpdzi.\yesnumber\displaystyle 2W^{\top}\Omega(I_{T}\otimes 1_{r_{i}}\otimes I_{n_{p}})M_{p}\textup{d}z_{i}.\yesnumber

It follows

dh(zi;[pj])dzi\displaystyle\frac{\textup{d}h\left(z_{i};[p_{j}]\right)}{\textup{d}z_{i}}
=\displaystyle= 2(ITri1np)[ITrinp(((IT1riInp)Mpzi\displaystyle 2(I_{Tr_{i}}\otimes 1_{n_{p}})^{\top}\Big{[}I_{Tr_{i}n_{p}}\odot\Big{(}\big{(}(I_{T}\otimes 1_{r_{i}}\otimes I_{n_{p}})M_{p}z_{i}
1T×1[pj])1TrinpT)](IT1riInp)Mp.\yesnumber\displaystyle\;-1_{T\times 1}\otimes[p_{j}]\big{)}1_{Tr_{i}n_{p}}^{T}\Big{)}\Big{]}\cdot(I_{T}\otimes 1_{r_{i}}\otimes I_{n_{p}})M_{p}.\yesnumber

Therefore, Lemma 5 is proved. ∎

Set Υi=ψ(zi;[pj])\Upsilon_{i}=-\psi\left(z_{i};[p_{j}]\right) and define Ξi\Xi_{i} such that ΞiΥi=1Tri\Xi_{i}\odot\Upsilon_{i}=1_{Tr_{i}}. Then, the gradient of the barriered augmented Lagrangian function ^bik\hat{\mathcal{L}}_{b_{i}^{k}} is

^bik\displaystyle\nabla\hat{\mathcal{L}}_{b_{i}^{k}}
=\displaystyle= 2Hi(zizi,ref)+ρik𝒜i(𝒜izi+izf,ik,r+sik,rhi\displaystyle 2H_{i}(z_{i}-z_{i,ref})+{\rho_{i}^{k}}\mathcal{A}_{i}^{\top}\Bigg{(}\mathcal{A}_{i}z_{i}+\mathcal{B}_{i}z_{f,i}^{k,r}+s_{i}^{k,r}-h_{i}
+yik,rρik)2bikTr(1TriΞiWΩ(IT1riInp)Mp).\displaystyle+\frac{y_{i}^{k,r}}{\rho_{i}^{k}}\Bigg{)}-2b_{i}^{k}\operatorname{Tr}\left(1_{Tr_{i}}^{\top}\Xi_{i}\odot W^{\top}\Omega(I_{T}\otimes 1_{r_{i}}\otimes I_{n_{p}})M_{p}\right).
\yesnumber\displaystyle\yesnumber

A closed-form algebraic solution of problem (V) is challenging to determine. However, since the gradient ^bik\nabla\hat{\mathcal{L}}_{b_{i}^{k}} is available, some well-known gradient-based numerical methods, e.g., Barzilai-Borwein, Polak-Ribiere, and Fletcher-Reeves, can be used to calculate the optimal ziz_{i}. Therefore, the ziz_{i}-update step in Algorithm 1 can be replaced by solving an unconstrained optimization problem (V) with the known gradient (V).

Usually, iterative algorithms for NLP need to be called when solving the ziz_{i} minimization problem (V), and always searching for a highly accurate solution in each ADMM iteration will result in an excessive computational cost. It is thus desirable to solve the optimization subproblem (V) in ADMM inexactly when the dual variables are yet far from the optimum, i.e., to allow zik,r+1z_{i}^{k,r+1} to be chosen such that

dxk+1ziρi.\displaystyle d_{x}^{k+1}\in\partial_{z_{i}}\mathcal{L}_{\rho_{i}}. (23)

In this paper, we use externally shrinking and summable sequence of the absolute errors, i.e.,

dzkϵ4k,k=1ϵ4k.\displaystyle\|d_{z}^{k}\|\leq\epsilon_{4}^{k},\sum_{k=1}^{\infty}\epsilon_{4}^{k}\leq\infty. (24)

such as a summable absolute error sequence can effectively reduce the total number of subroutine iterations throughout the ADMM algorithm. Such a criterion is a constructive one, rendered to guarantee the decrease of a quadratic distance between the intermediate solutions (zik,zf,ik,sik,yik)(z_{i}^{k},z_{f,i}^{k},s_{i}^{k},y_{i}^{k}) and the optimum (zi,zf,ik,sik,yik)(z_{i}^{*},z_{f,i}^{k},s_{i}^{k},y_{i}^{k}).

The above approximate NLP solution is realizable by NLP solvers where the tolerances of the KKT conditions are allowed to be specified by the user, e.g., the IPOPT solver. The pseudocode of the detailed algorithm is shown in Algorithm 2.

Algorithm 2 Improved Hierarchical ADMM for Multi-Agent Decision Making
  Initialize zi0T(m+n),zf,i02T(m+n),si02Tm+3Tnz_{i}^{0}\in\mathbb{R}^{T(m+n)},z_{f,i}^{0}\in\mathbb{R}^{2T(m+n)},s_{i}^{0}\in\mathbb{R}^{2Tm+3Tn}; initialize the dual variable λi0[λ¯i,λ¯i]\lambda_{i}^{0}\in[\underline{\lambda}_{i},\overline{\lambda}_{i}], where λ¯i,λ¯i2Tm+3Tn\underline{\lambda}_{i},\overline{\lambda}_{i}\in\mathbb{R}^{2Tm+3Tn}, and λ¯i>λ¯i\overline{\lambda}_{i}>\underline{\lambda}_{i}; initialize the penalty parameter βi0>0\beta_{i}^{0}>0, and the constant ω[0,1)\omega\in[0,1) and γ>1\gamma>1; initialize the sequence of tolerance {ϵ1k,ϵ2k,ϵ3k,epsilon4k,ϵb}+\{\epsilon_{1}^{k},\epsilon_{2}^{k},\epsilon_{3}^{k},epsilon_{4}^{k},\epsilon_{b}\}\in\mathbb{R}_{+} with limkϵek=0,e{1,2,3,4,b}\lim_{k\rightarrow\infty}\epsilon_{e}^{k}=0,\forall e\in\{1,2,3,4,b\}. Given the terminated tolerances ϵterm,e,e={1,2,3,4,b}\epsilon_{\text{term},e},e=\{1,2,3,4,b\}, if ϵekϵterm,e\epsilon_{e}^{k}\leq\epsilon_{\text{term},e}, ϵek=ϵterm,i\epsilon_{e}^{k}=\epsilon_{\text{term},i} for e={1,2,3,4,b}e=\{1,2,3,4,b\}.
  Set the outer iteration k=0k=0.
  while Stopping criterion is violated or bikϵbb_{i}^{k}\geq\epsilon_{b} do
     Given λik\lambda_{i}^{k} and βik\beta_{i}^{k}, initialize zik,0,zf,ik,0,sik,0z_{i}^{k,0},z_{f,i}^{k,0},s_{i}^{k,0} and yik,0y_{i}^{k,0} such that λik+βiksik,0+yik,0=0\lambda_{i}^{k}+\beta_{i}^{k}s_{i}^{k,0}+y_{i}^{k,0}=0; initialize tolerance (ϵ1k,ϵ2k,ϵ3k)(\epsilon_{1}^{k},\epsilon_{2}^{k},\epsilon_{3}^{k}); ρik2βik\rho_{i}^{k}\geq\sqrt{2}\beta_{i}^{k}.
     Set the inner iteration r=0r=0.
     while Stopping criterion is violated do
        Update zik,r+1z_{i}^{k,r+1} by minimize (22).
        Update zf,ik,r+1z_{f,i}^{k,r+1}, sik,r+1s_{i}^{k,r+1}, yik,r+1y_{i}^{k,r+1} by (IV), (IV), (IV), respectively.
        r=r+1r=r+1.
     end while
     Set zik+1,zf,ik+1,sik+1z_{i}^{k+1},z_{f,i}^{k+1},s_{i}^{k+1}, yik+1y_{i}^{k+1} to be zik,r+1,zf,ik,r+1,sik,r+1z_{i}^{k,r+1},z_{f,i}^{k,r+1},s_{i}^{k,r+1}, yik,r+1y_{i}^{k,r+1}, respectively.
     Update λik+1\lambda_{i}^{k+1} and βik+1\beta_{i}^{k+1} by (IV).
     k=k+1k=k+1.
  end while

In this method, a nonlinear programming problem without constraints is solved with inequality constraints handled by a fixed barrier constant throughout inner iterations. The barrier constant decreases across outer iterations. In the hierarchical ADMM, the NLP solver needs to solve an inequality-constrained NLP in each inner iteration. When the NLP solver, i.e., interior-point algorithm, is called for NLP with inequalities, the barrier constant will go through an iteration inside the solver, and thus every inner iteration will have more loops of internal iteration. By this barrier Lagrangian construction, we integrate the inner loop of the IPOPT with the inner loop of the hierarchical ADMM algorithm. In addition, in the improved hierarchical ADMM, each inner iteration only solves NLP, and an approximate solution is reached without the need of convergence. The accuracy, namely the tolerances of the errors from the KKT conditions, can be controlled when using the interior-point algorithm. Therefore, these tolerances can be set adaptively throughout the iterations (for instance, the tolerances of the 1st, 5th, and 20th iteration can be set to 0.1, 0.01, and 0.001, respectively). By this adaptive tolerance setting, we integrate the outer loop of the IPOPT with the inner loop of the hierarchical ADMM. Therefore, the improved hierarchical ADMM gives less iteration number and computation time than the hierarchical ADMM.

VI Illustrative Example

In order to show the effectiveness of the hierarchical ADMM and improved hierarchical ADMM, a multi-UAV system is presented as an application test platform for demonstration of the proposed decision-making framework. The simulations are performed in Python 3.7 environment with processor Intel Xeon(R) E5-1650 v4 @ 3.60GHz. All nonlinear problems are solved by the open source IPOPT solver.

VI-A Dynamic Model

For each quadcopter system, the system model can be expressed in the following state-space representation:

x˙=[p˙v˙ζ˙ω˙]=[R(ϕ,θ,ψ)vω×v+gR(ϕ,θ,ψ)eW(ϕ,θ,ψ)ωJ1(ω×Jω)]+B~(ueq+u),\yesnumber\displaystyle\dot{x}={\begin{bmatrix}{\dot{p}}\\ {\dot{v}}\\ {\dot{\zeta}}\\ {\dot{\omega}}\end{bmatrix}}={\begin{bmatrix}{R(\phi,\theta,\psi)}^{\top}{v}\\ {-{\omega}\times v+g{R(\phi,\theta,\psi)}e}\\ W(\phi,\theta,\psi){\omega}\\ J^{-1}(-{\omega}\times{J}{\omega})\end{bmatrix}+{\tilde{B}}{({u}_{\mathrm{eq}}+{u})}},\yesnumber

where x12x\in\mathbb{R}^{12} and u4u\in\mathbb{R}^{4} are state and control input vector for the agents, respectively. In this dynamics model, rotor thrusts of each rotor are chosen as control input vector, i.e., u=[F1F2F3F4]u=\begin{bmatrix}F_{1}&F_{2}&F_{3}&F_{4}\end{bmatrix}^{\top}. p=[pxpypz]{p}=\begin{bmatrix}p_{x}&p_{y}&p_{z}\end{bmatrix}^{\top} is a vector comprising the position in the xx-, yy-, and zz-dimension. Additionally, the velocity vector is represented by v=[vxvyvz]{v}=\begin{bmatrix}v_{x}&v_{y}&v_{z}\end{bmatrix}^{\top}. ζ=[ϕθψ]\zeta=\begin{bmatrix}\phi&\theta&\psi\end{bmatrix}^{\top} denotes a vector in terms of the roll, pitch, and yaw angles. The angular velocity vector is represented by ω=[ωxωyωz]\omega=\begin{bmatrix}\omega_{x}&\omega_{y}&\omega_{z}\end{bmatrix}^{\top}. Also, e=[001]{e}=\begin{bmatrix}0&0&1\end{bmatrix}^{\top}, ueq=[mg/4mg/4mg/4mg/4]u_{\mathrm{eq}}=\begin{bmatrix}{mg}/{4}&{mg}/{4}&{mg}/{4}&{mg}/{4}\end{bmatrix}^{\top} is the control input to balance the gravity of the agents, gg is the gravitational acceleration, mm is the mass of the quadcopter, J=diag(Jx,Jy,Jz)J=\textup{diag}(J_{x},J_{y},J_{z}) denotes the moment of inertia of the quadcopter, R(ϕ,θ,ψ)R(\phi,\theta,\psi) and W(ϕ,θ,ψ)W(\phi,\theta,\psi) denote the rotation matrices of the quadcopter. In addition, B~{\tilde{B}} is the control input matrix in the above state-space representation.

Here, (VI-A) is nonlinear and time-variant, which can be represented as a pseudo-linear form by using state-dependent coefficient factorization [31]. The state-space expression is

xt+1=(A~tΔt+I12)xt+B~Δt(ueq+ut),\displaystyle{x_{t+1}}=(\tilde{A}_{t}\Delta t+I_{12})x_{t}+\tilde{B}\Delta t(u_{\textup{eq}}+u_{t}), (25)

where Δt\Delta t is the sampling time interval. Since A~t{\tilde{A}}_{t} and B~\tilde{B} are dependent on the current state xx, this state-space representation is in a pseudo-linear form, and then we can suitably consider the system matrices to be constant during the prediction horizon. More details of the parameter settings in (VI-A) and (25) are presented in [32].

VI-B Optimization and Simulation Results

The dynamic model of each agent in the multi-UAV system is shown in (25). All the agents are set evenly distributed in a circle with the radius of 22 m and pz=0p_{z}=0 m, and each agent needs to reach its corresponding point in this same sphere. Therefore, a large number of potential collision possibilities are set up in the simulation task, and our objective is to avoid these designed potential collisions among these agents and make all agents reach their desired destinations. Here, a distributed MPC with a quadratic objective function is designed to realize the path tracking and collision avoidance tasks. Set the number of agents N=8N=8. The safety distance dsafed_{\text{safe}} is 0.20.2 m. The dimensions are n=12,m=4,np=3n=12,m=4,n_{p}=3. In addition, the agents are connected with all of the remaining agents. The adjacency matrix of the communication network 𝒟\mathcal{D} is the difference between the upper triangular matrix of the square one matrix 1(N,N)1_{(N,N)} and the identity matrix INI_{N}, i.e., 𝒟=U(1(N,N))IN\mathcal{D}=U(1_{(N,N)})-I_{N}. In this adjacency matrix 𝒟\mathcal{D}, each entry 𝒟ij\mathcal{D}_{ij} represents whether there is the connection between the iith agent and the jjth agent. If 𝒟ij=1\mathcal{D}_{ij}=1, there is a communication connection between the two agents; otherwise, there is no connection between the two agents. For example, the iith agent will get communication with the i+1,i+2,,Ni+1,i+2,\cdots,N agents. The velocity of the agents is confined into [-5, 5] m/s, and the upper and lower bounds of angles (ϕ,ψ,θ)(\phi,\psi,\theta) in three dimensions are (π,π2,π)(\pi,\frac{\pi}{2},\pi) and (π,π2,π)(-\pi,-\frac{\pi}{2},-\pi). The upper bound of control input variable is u¯=(1.96,1.96,1.96,1.96)\overline{u}=(1.96,1.96,1.96,1.96) N and its lower bound is u¯=u¯\underline{u}=-\overline{u}. Set the prediction horizon T=25T=25 with the sampling time Δt=0.05\Delta t=0.05 s. The weighting matrices Ri=0(m,m)R_{i}=0_{(m,m)} and Qi=blockdiag(Inp,0(nnp,nnp))Q_{i}=\operatorname{blockdiag}(I_{n_{p}},0_{(n-n_{p},n-n_{p})}) for all i𝒱i\in\mathcal{V}. Each element in λ¯i\overline{\lambda}_{i} and λ¯i\underline{\lambda}_{i} are set as 0.010.01 and 0.01-0.01, respectively. ω\omega and γ\gamma are set as 0.9 and 1.1. The initial tolerances ϵ10,ϵ20,ϵ30\epsilon_{1}^{0},\epsilon_{2}^{0},\epsilon_{3}^{0} are defined as 102,102,101,10^{-2},10^{-2},10^{-1}, respectively, and ϵik=ϵik1/5k1,i{1,2,3}\epsilon_{i}^{k}=\epsilon_{i}^{k-1}/5^{k-1},\forall i\in\{1,2,3\}. The terminated tolerances for outer iterations ϵterm,1\epsilon_{\text{term},1}, ϵterm,2\epsilon_{\text{term},2}, ϵterm,3\epsilon_{\text{term},3} are 106,106,10510^{-6},10^{-6},10^{-5}, respectively. The maximum iteration number is set to 10410^{4}.

The reference trajectories and the resulted trajectories computed by the improved hierarchical ADMM are shown in Fig. 1 and Fig. 2. The initial positions of the 8 agents are distributed uniformly in a circle, whose radius equals 2 m, and the z-axis is 0. In this simulation task, the 8 agents need to move towards their corresponding points, which are the central symmetric points of their initial position, respectively. For example, the agents with initial position (0,2,0) m and (2,2,0\sqrt{2},\sqrt{2},0) m needs to reach (0,-2,0) m and (2,2,0-\sqrt{2},-\sqrt{2},0) m. In the two figures, each agent is represented by a different color. These lines in different colors denote the resulted trajectories of the 8 agents, respectively. From the two figures, it is obvious that all agents can effectively avoid their neighbors and maintain the safety distance. On the other hand, for the hierarchical ADMM, a similar performance of collision avoidance can be attained.

Refer to caption
Figure 1: Resulted trajectories of 8 agents in 3D view.
Refer to caption
Figure 2: Resulted trajectories of 8 agents in 2D view.

In order to show the bound of the control inputs, we only choose the control input results of the 1st agent u=[F1F2F3F4]u=\begin{bmatrix}F_{1}&F_{2}&F_{3}&F_{4}\end{bmatrix} as an example for the clarity of this figure, as shown in Fig. 3. The 4 dimensions of the control inputs are bounded in the predefined range [1.96,1.96][-1.96,1.96] N.

Refer to caption
Figure 3: Resulted control inputs of the 1st agent.

The distances between each pair of the 8 agents are shown in Fig. 4. The solid gray line in this figure represents the safety distance all of the agents need to maintain, and the other colored lines denote the distance of each pair of agents with respect to time. It shows that the distance between agents are no smaller than the safety distance.

Refer to caption
Figure 4: Distances between each pair of agents.

The comparison of computational efficiency between our proposed hierarchical ADMM and improved hierarchical ADMM can be illustrated, according to Fig. 5 and Fig. 6. Fig. 5 demonstrates the comparison of iteration numbers the inner loop and outer loop require in both hierarchical ADMM and improved hierarchical ADMM. It is straightforward to observe that the improved hierarchical ADMM can effectively reduce the iteration numbers of both the inner loop and outer loop, compared with the hierarchical ADMM. The comparison of computation time for all inner loop iterations between hierarchical ADMM and improved hierarchical ADMM is shown in Fig. 6. From this figure, it can be observed that the computation time is significantly reduced by using the improved version.

Refer to caption
Figure 5: Comparison of inner loop and outer loop iteration numbers between hierarchical ADMM and improved hierarchical ADMM.
Refer to caption
Figure 6: Comparison of computation time of all inner loop iterations between hierarchical ADMM and improved hierarchical ADMM.
TABLE I: Comparison of proposed algorithms (hierarchical ADMM and improved hierarchical ADMM) with the centralized MPC and penalty dual decomposition method [33]
Centralized MPC Penalty Dual Decomposition [33] Hierarchical ADMM Improved Hierarchical ADMM
Average outer iterations - 4.26 2.12 1.03
Average inner iterations - 17.36 6.38 1.12
𝒜izi+izf,ihi\|\mathcal{A}_{i}z_{i}+\mathcal{B}_{i}z_{f,i}-h_{i}\| 2.21E-5 3.32E-5 2.54E-5 1.47E-5
Cost value 253.12 315.35 294.23 245.36
Stopping criterions - 1E-6, 7.52E-2, 2.32E-3 1E-6, 9.38E-2, 3.5E-3 1E-6, 3.28E-2, 2.64E-3
Time 1.47s 0.22s 0.16s 0.09s

In order to show the effectiveness of the proposed improved hierarchical ADMM, the penalty dual decomposition method [33] and the centralized MPC method are used as the comparison methods. The centralized MPC method means we solve the centralized MPC problem by using a NLP solver, and the penalty dual decomposition method is also a double-loop iterative algorithm [33]. The inner loop of the penalty dual decomposition method is used to solve a nonconvex nonsmooth augmented Lagrangian problem via block-coordinate-descent-type methods, and the outer loop of this algorithm is to update the dual variables and/or the penalty parameter [33]. Table I illustrates the advantages of our proposed algorithms (hierarchical ADMM and improved hierarchical ADMM) over the penalty dual decomposition method and the centralized MPC method.

In order to demonstrate the performance in terms of computational efficiency, the comparison of computation time per step (including all outer loop iterations and inner loop iterations) under the hierarchical ADMM and the improved hierarchical ADMM, and the computation time per step under the centralized MPC (solved by the IPOPT solver) are shown in Fig. 7. From this figure, the computational efficiency of the ADMM for solving large-scale problems can be clearly observed. Moreover, the improved hierarchical ADMM is capable of further improving the efficiency, as compared with the hierarchical ADMM. With the increasing number of agents, the number of nonlinear constraints increases rapidly, which leads to difficulty in solving the problem. Therefore, the computation time is increasing with the increase in the number of agents NN, as shown in Fig. 8. However, when the centralized MPC solver is used, the growth rate of the computation time is much higher than the ones when the hierarchical ADMM and improved hierarchical ADMM are used, as shown in Fig. 8, which also demonstrates the efficiency of the two proposed methods. Besides, the efficiency of the improved hierarchical ADMM can also be proved in this figure.

Refer to caption
Figure 7: Comparison of the computation time per step among the hierarchical ADMM, improved hierarchical ADMM, and centralized MPC solver solver.
Refer to caption
Figure 8: Comparison of the computation time among the hierarchical ADMM, improved hierarchical ADMM, and centralized MPC solver with the increase of number of agents.

VII Conclusion

In this paper, a large-scale nonconvex optimization problem for multi-agent decision making is investigated and solved by the hierarchical ADMM approach. For the multi-agent system, an appropriate DMPC problem is formulated and presented to handle the constraints and achieve the collective objective. Firstly, an innovation using appropriate slack variables is deployed, yielding a resulting transformed optimization problem, which is used to deal with the nonconvex characteristics of the problem. Then, the hierarchical ADMM is applied to solve the transformed problem in parallel. Next, the improved hierarchical ADMM is proposed to increase the computational efficiency and decrease the number of inner iterations. Additionally, it is analytically shown that the appropriate desired stationary point exists for the procedures of the hierarchical stages for algorithmic convergence. Finally, comparative simulations are conducted using a multi-UAV system as the test platform, which further illustrates the effectiveness and efficiency of the proposed methodology.

References

  • [1] X. He, T. Huang, J. Yu, C. Li, and Y. Zhang, “A continuous-time algorithm for distributed optimization based on multiagent networks,” IEEE Transactions on Systems, Man, and Cybernetics: Systems, vol. 49, no. 12, pp. 2700–2709, 2017.
  • [2] H. Hong, W. Yu, G. Wen, and X. Yu, “Distributed robust fixed-time consensus for nonlinear and disturbed multiagent systems,” IEEE Transactions on Systems, Man, and Cybernetics: Systems, vol. 47, no. 7, pp. 1464–1473, 2016.
  • [3] M. Ye, G. Wen, S. Xu, and F. L. Lewis, “Global social cost minimization with possibly nonconvex objective functions: An extremum seeking-based approach,” IEEE Transactions on Systems, Man, and Cybernetics: Systems, 2020.
  • [4] J. Ma, S.-L. Chen, C. S. Teo, A. Tay, A. Al Mamun, and K. K. Tan, “Parameter space optimization towards integrated mechatronic design for uncertain systems with generalized feedback constraints,” Automatica, vol. 105, pp. 149–158, 2019.
  • [5] S. Yang, Q. Liu, and J. Wang, “Distributed optimization based on a multiagent system in the presence of communication delays,” IEEE Transactions on Systems, Man, and Cybernetics: Systems, vol. 47, no. 5, pp. 717–728, 2016.
  • [6] ——, “A multi-agent system with a proportional-integral protocol for distributed constrained optimization,” IEEE Transactions on Automatic Control, vol. 62, no. 7, pp. 3461–3467, 2016.
  • [7] D. Yuan, S. Xu, and H. Zhao, “Distributed primal–dual subgradient method for multiagent optimization via consensus algorithms,” IEEE Transactions on Systems, Man, and Cybernetics, Part B (Cybernetics), vol. 41, no. 6, pp. 1715–1724, 2011.
  • [8] E. Camponogara and L. B. De Oliveira, “Distributed optimization for model predictive control of linear-dynamic networks,” IEEE Transactions on Systems, Man, and Cybernetics-Part A: Systems and Humans, vol. 39, no. 6, pp. 1331–1338, 2009.
  • [9] M. Hong, “A distributed, asynchronous, and incremental algorithm for nonconvex optimization: an ADMM approach,” IEEE Transactions on Control of Network Systems, vol. 5, no. 3, pp. 935–945, 2017.
  • [10] J. Ma, Z. Cheng, X. Zhang, M. Tomizuka, and T. H. Lee, “On symmetric Gauss-Seidel ADMM algorithm for H{H}_{\infty} guaranteed cost control with convex parameterization,” arXiv preprint arXiv:2001.00708, 2020.
  • [11] ——, “Optimal decentralized control for uncertain systems by symmetric Gauss-Seidel semi-proximal ALM,” IEEE Transactions on Automatic Control, DOI: 10.1109/TAC.2021.3052768, 2021.
  • [12] Y. Hu, E. C. Chi, and G. I. Allen, “ADMM algorithmic regularization paths for sparse statistical machine learning,” in Splitting Methods in Communication, Imaging, Science, and Engineering.   Springer, 2016, pp. 433–459.
  • [13] J. Ding, Y. Gong, C. Zhang, M. Pan, and Z. Han, “Optimal differentially private ADMM for distributed machine learning,” arXiv preprint arXiv:1901.02094, 2019.
  • [14] B. He and X. Yuan, “On the O(1/n){O}(1/n) convergence rate of the Douglas-Rachford alternating direction method,” SIAM Journal on Numerical Analysis, vol. 50, no. 2, pp. 700–709, 2012.
  • [15] R. D. Monteiro and B. F. Svaiter, “Iteration-complexity of block-decomposition algorithms and the alternating direction method of multipliers,” SIAM Journal on Optimization, vol. 23, no. 1, pp. 475–507, 2013.
  • [16] E. Wei and A. Ozdaglar, “Distributed alternating direction method of multipliers,” in 2012 IEEE 51st IEEE Conference on Decision and Control (CDC).   IEEE, 2012, pp. 5445–5450.
  • [17] A. Makhdoumi and A. Ozdaglar, “Convergence rate of distributed ADMM over networks,” IEEE Transactions on Automatic Control, vol. 62, no. 10, pp. 5082–5095, 2017.
  • [18] B. He, M. Tao, and X. Yuan, “Alternating direction method with Gaussian back substitution for separable convex programming,” SIAM Journal on Optimization, vol. 22, no. 2, pp. 313–340, 2012.
  • [19] T. Lin, S. Ma, and S. Zhang, “On the global linear convergence of the ADMM with multiblock variables,” SIAM Journal on Optimization, vol. 25, no. 3, pp. 1478–1497, 2015.
  • [20] G. Li and T. K. Pong, “Global convergence of splitting methods for nonconvex composite optimization,” SIAM Journal on Optimization, vol. 25, no. 4, pp. 2434–2460, 2015.
  • [21] H. Wang and A. Banerjee, “Bregman alternating direction method of multipliers,” in Advances in Neural Information Processing Systems, 2014, pp. 2816–2824.
  • [22] B. Jiang, T. Lin, S. Ma, and S. Zhang, “Structured nonconvex and nonsmooth optimization: algorithms and iteration complexity analysis,” Computational Optimization and Applications, vol. 72, no. 1, pp. 115–157, 2019.
  • [23] Y. Wang, W. Yin, and J. Zeng, “Global convergence of ADMM in nonconvex nonsmooth optimization,” Journal of Scientific Computing, vol. 78, no. 1, pp. 29–63, 2019.
  • [24] X. Wang, J. Yan, B. Jin, and W. Li, “Distributed and parallel ADMM for structured nonconvex optimization problem,” IEEE Transactions on Cybernetics, 2019.
  • [25] Z. Cheng, J. Ma, X. Zhang, C. W. de Silva, and T. H. Lee, “ADMM-based parallel optimization for multi-agent collision-free model predictive control,” arXiv preprint arXiv:2101.09894, 2021.
  • [26] J. Ma, Z. Cheng, X. Zhang, M. Tomizuka, and T. H. Lee, “Alternating direction method of multipliers for constrained iterative LQR in autonomous driving,” arXiv preprint arXiv:2011.00462, 2020.
  • [27] B. Houska, J. Frasch, and M. Diehl, “An augmented lagrangian based algorithm for distributed nonconvex optimization,” SIAM Journal on Optimization, vol. 26, no. 2, pp. 1101–1127, 2016.
  • [28] K. Sun and X. A. Sun, “A two-level distributed algorithm for nonconvex constrained optimization,” arXiv preprint arXiv:1902.07654, 2019.
  • [29] W. Tang and P. Daoutidis, “Fast and stable nonconvex constrained distributed optimization: The ELLADA algorithm,” arXiv preprint arXiv:2004.01977, 2020.
  • [30] D. Limón, T. Alamo, F. Salas, and E. F. Camacho, “On the stability of constrained mpc without terminal constraint,” IEEE Transactions on Automatic Control, vol. 51, no. 5, pp. 832–836, 2006.
  • [31] X. Zhang, J. Ma, S. Huang, Z. Cheng, and T. H. Lee, “Integrated planning and control for collision-free trajectory generation in 3D environment with obstacles,” in Proceedings of International Conference on Control, Automation and Systems, 2019, pp. 974–979.
  • [32] X. Zhang, J. Ma, Z. Cheng, S. Huang, S. S. Ge, and T. H. Lee, “Trajectory generation by chance-constrained nonlinear MPC with probabilistic prediction,” IEEE Transactions on Cybernetics, DOI: 10.1109/TCYB.2020.3032711, 2020.
  • [33] Q. Shi and M. Hong, “Penalty dual decomposition method for nonsmooth nonconvex optimization-Part I: Algorithms and convergence analysis,” IEEE Transactions on Signal Processing, vol. 68, pp. 4108–4122, 2020.