Network Flows that Solve Linear Equations

Similar documents
Introduction to Real Analysis Alternative Chapter 1

Controlling and Stabilizing a Rigid Formation using a few agents

Convergence Rate of Nonlinear Switched Systems

Functional Analysis. Franck Sueur Metric spaces Definitions Completeness Compactness Separability...

1 Lyapunov theory of stability

DS-GA 1002 Lecture notes 0 Fall Linear Algebra. These notes provide a review of basic concepts in linear algebra.

Consensus-Based Distributed Optimization with Malicious Nodes

Multi-Robotic Systems

Decentralized Stabilization of Heterogeneous Linear Multi-Agent Systems

2. Dual space is essential for the concept of gradient which, in turn, leads to the variational analysis of Lagrange multipliers.

Chapter III. Stability of Linear Systems

Constrained Consensus and Optimization in Multi-Agent Networks

Structural and Multidisciplinary Optimization. P. Duysinx and P. Tossings

E-Companion to The Evolution of Beliefs over Signed Social Networks

Reaching a Consensus in a Dynamically Changing Environment A Graphical Approach

Part III. 10 Topological Space Basics. Topological Spaces

Dynamic Coalitional TU Games: Distributed Bargaining among Players Neighbors

Lecture 2: Linear Algebra Review

SPECTRAL PROPERTIES OF THE LAPLACIAN ON BOUNDED DOMAINS

Linear Algebra Massoud Malek

1 The Observability Canonical Form

1 Directional Derivatives and Differentiability

SPECTRAL THEOREM FOR COMPACT SELF-ADJOINT OPERATORS

AN ELEMENTARY PROOF OF THE SPECTRAL RADIUS FORMULA FOR MATRICES

Introduction and Preliminaries

Lebesgue Measure on R n

Complex Laplacians and Applications in Multi-Agent Systems

BOUNDARY VALUE PROBLEMS ON A HALF SIERPINSKI GASKET

(x k ) sequence in F, lim x k = x x F. If F : R n R is a function, level sets and sublevel sets of F are any sets of the form (respectively);

On Backward Product of Stochastic Matrices

Metric Spaces and Topology

Connectedness. Proposition 2.2. The following are equivalent for a topological space (X, T ).

In English, this means that if we travel on a straight line between any two points in C, then we never leave C.

The goal of this chapter is to study linear systems of ordinary differential equations: dt,..., dx ) T

Maximum Capability of Feedback Control for Network Systems

Chapter 1. Preliminaries. The purpose of this chapter is to provide some basic background information. Linear Space. Hilbert Space.

Appendix B Convex analysis

The nonsmooth Newton method on Riemannian manifolds

LECTURE 15: COMPLETENESS AND CONVEXITY

Measure Theory on Topological Spaces. Course: Prof. Tony Dorlas 2010 Typset: Cathal Ormond

Exercise Solutions to Functional Analysis

NORMS ON SPACE OF MATRICES

Converse Lyapunov theorem and Input-to-State Stability

Boolean Inner-Product Spaces and Boolean Matrices

On Krause s Multi-Agent Consensus Model With State-Dependent Connectivity

Locally convex spaces, the hyperplane separation theorem, and the Krein-Milman theorem

Set, functions and Euclidean space. Seungjin Han

Near-Potential Games: Geometry and Dynamics

Disturbance Attenuation Properties for Discrete-Time Uncertain Switched Linear Systems

Monotone Control System. Brad C. Yu SEACS, National ICT Australia And RSISE, The Australian National University June, 2005

Part V. 17 Introduction: What are measures and why measurable sets. Lebesgue Integration Theory

Lebesgue Measure on R n

Analysis-3 lecture schemes

Lecture 4. Chapter 4: Lyapunov Stability. Eugenio Schuster. Mechanical Engineering and Mechanics Lehigh University.

Output Input Stability and Minimum-Phase Nonlinear Systems

Differential Topology Final Exam With Solutions

Stability of Feedback Solutions for Infinite Horizon Noncooperative Differential Games

U.C. Berkeley CS294: Spectral Methods and Expanders Handout 11 Luca Trevisan February 29, 2016

An introduction to some aspects of functional analysis

CHAPTER 7. Connectedness

A Geometric Framework for Nonconvex Optimization Duality using Augmented Lagrangian Functions

08a. Operators on Hilbert spaces. 1. Boundedness, continuity, operator norms

Duke University, Department of Electrical and Computer Engineering Optimization for Scientists and Engineers c Alex Bronstein, 2014

Vectors. January 13, 2013

DO NOT OPEN THIS QUESTION BOOKLET UNTIL YOU ARE TOLD TO DO SO

Exercises: Brunn, Minkowski and convex pie

Spectral Clustering on Handwritten Digits Database

Problem Set 6: Solutions Math 201A: Fall a n x n,

Elements of Convex Optimization Theory

The Liapunov Method for Determining Stability (DRAFT)

Tangent spaces, normals and extrema

PROBLEMS. (b) (Polarization Identity) Show that in any inner product space

Course Summary Math 211

An introduction to Birkhoff normal form

In particular, if A is a square matrix and λ is one of its eigenvalues, then we can find a non-zero column vector X with

Empirical Processes: General Weak Convergence Theory

DUALITY IN COUNTABLY INFINITE MONOTROPIC PROGRAMS

Eigenvalues and Eigenfunctions of the Laplacian

Topological properties

The Strong Largeur d Arborescence

Existence and Uniqueness

Equivalence of dynamical systems by bisimulation

Locally Lipschitzian Guiding Function Method for ODEs.

Generalized Input-to-State l 2 -Gains of Discrete-Time Switched Linear Control Systems

Proofs for Large Sample Properties of Generalized Method of Moments Estimators

Eigenvalue comparisons in graph theory

15-780: LinearProgramming

Lecture 1: Entropy, convexity, and matrix scaling CSE 599S: Entropy optimality, Winter 2016 Instructor: James R. Lee Last updated: January 24, 2016

A strongly polynomial algorithm for linear systems having a binary solution

Lecture Notes in Advanced Calculus 1 (80315) Raz Kupferman Institute of Mathematics The Hebrew University

On reduction of differential inclusions and Lyapunov stability

Relations Graphical View

Matrices and Vectors. Definition of Matrix. An MxN matrix A is a two-dimensional array of numbers A =

Chapter 1. Measure Spaces. 1.1 Algebras and σ algebras of sets Notation and preliminaries

SHARP BOUNDARY TRACE INEQUALITIES. 1. Introduction

Math 341: Convex Geometry. Xi Chen

Examples include: (a) the Lorenz system for climate and weather modeling (b) the Hodgkin-Huxley system for neuron modeling

Global Stability and Asymptotic Gain Imply Input-to-State Stability for State-Dependent Switched Systems

Convex Functions and Optimization

4th Preparation Sheet - Solutions

Transcription:

Network Flows that Solve Linear Equations Guodong Shi, Brian D. O. Anderson and Uwe Helmke Abstract We study distributed network flows as solvers in continuous time for the linear algebraic equation arxiv:1510.05176v3 [cs.sy] 0 Sep 016 z = Hy. Each node i has access to a row h T i of the matrix H and the corresponding entry z i in the vector z. The first consensus + projection flow under investigation consists of two terms, one from standard consensus dynamics and the other contributing to projection onto each affine subspace specified by the h i and z i. The second projection consensus flow on the other hand simply replaces the relative state feedback in consensus dynamics with projected relative state feedback. Without dwell-time assumption on switching graphs, we prove that all node states converge to a common solution of the linear algebraic equation, if there is any. The convergence is global for the consensus + projection flow while local for the projection consensus flow in the sense that the initial values must lie on the affine subspaces. If the linear equation has no exact solutions, we show that the node states can converge to a ball around the least squares solution whose radius can be made arbitrarily small through selecting a sufficiently large gain for the consensus + projection flow for a fixed bidirectional graph. Semi-global convergence to approximate least squares solutions is also demonstrated for switching balanced directed graphs under suitable conditions. It is also shown that the projection consensus flow drives the average of the node states to the least squares solution with a complete graph. Numerical examples are provided as illustrations of the established results. 1 Introduction In the past decade, distributed consensus algorithms have attracted a significant amount of research attention [8 11], due to their wide applications in distributed control and estimation [1,13], distributed signal processing [14], and distributed optimization methods [15, 16]. The basic idea is that a network of interconnected nodes with different initial values can reach a common value, namely an agreement A brief version of the current manuscript was presented at the American Control Conference in Boston, July 016 [1]. G. Shi is with the Research School of Engineering, College of Engineering and Computer Science, The Australian National University, Canberra 000, Australia. E-mail: guodong.shi@anu.edu.au. B. D. O. Anderson is with the National ICT Australia NICTA and the Research School of Engineering, College of Engineering and Computer Science, The Australian National University, Canberra 000, Australia. E-mail: brian.anderson@anu.edu.au. U. Helmke is deceased but was with Institute of Mathematics, University of Würzburg, 97074 Würzburg, Germany. 1

or a consensus, via local information exchange as long as the communication graph is well connected. The consensus value is informative in the sense that it can depend on all nodes initial states even if the value is not the exact average [17]; the consensus processes are robust against random or deterministic switching of network interactions as well as against noises [18 1]. As a generalized notion, constrained consensus seeks to reach state agreement in the intersection of a number of convex sets, where each set serves as the supporting state space for a particular node []. It was shown that with doubly stochastic arc weights, a projected consensus algorithm, where each node iteratively projects a weighted average of its neighbor s values onto its supporting set, can guarantee convergence to constrained consensus when the intersection set is bounded and contains an interior point []. This idea was then extended to the continuous-time case under which the graph connectivity and arc weight conditions can be relaxed with convergence still being guaranteed [3]. The projected consensus algorithm was shown to be convergent even under randomized projections [4]. In fact, those developments are closely related to a class of so-called alternating projection algorithms, which was first studied by von Neumann in the 1940s [5] and then extensively investigated across the past many decades [6 9]. A related but different intriguing problem lies in how to develop distributed algorithms that solve a linear algebraic equation z = Hy, H R N m, z R N with respect to variable y R m, to which tremendous efforts have been devoted with many algorithms developed under different assumptions of information exchanges among the nodes [30 39]. Naturally, we can assume that a node i has access to a row vector, h T i of H as well as the corresponding entry, z i in z. Each node will then be able to determine with straightforward calculations an affine subspace whose elements satisfy the equation h T i y = z i. This means that, as long as the original equation z = Hy has a nonempty solution set, solving the equation would be equivalent to finding a point in the intersection of all the affine subspaces and therefore can be put into the constrained consensus framework. If however the linear equation has no exact solution and its least squares solutions are of interest, we ended up with a convex optimization problem with quadratic cost and linear constraints. Many distributed optimization algorithms, such as [16,, 40 46], developed for much more complex models, can therefore be directly applied. It is worth emphasizing that the above ideas of distributed consensus and optimization were traced back to the seminal work of Bertsekas and J. Tsitsiklis [47, 48]. In this paper, we study two distributed network flows as distributed solvers for such linear algebraic equations in continuous time. The first so-called consensus + projection flow consists of two additive terms, one from standard consensus dynamics and the other contributing to projection onto each affine subspace. The second projection consensus flow on the other hand simply replaces the relative state feedback in consensus dynamics with projected relative state feedback. Essentially only relative state information is exchanged among the nodes for both of the two flows, which justifies their full distribut-

edness. To study the asymptotic behaviours of the two distributed flows with respect to the solutions of the linear equation, new challenges arise in the loss of intersection boundedness and interior points for the exact solution case as well as in the complete loss of intersection sets for the least squares solution case. As a result, the analysis cannot be simply mapped back to the studies in [, 3]. The contributions of the current paper are summarized as follows. Under mild conditions on the communication graphs without requiring a dwell-time for switching graphs and only requiring a positively lower bound on the integral of arc weights over certain time intervals, we prove that all node states asymptotically converge to a common solution of the linear algebraic equation for the two flows, if there is any. The convergence is global for the consensus + projection flow, and local for the projection consensus flow in the sense that the initial values must be put into the affine subspaces which is a very minor restriction indeed. We manage to characterize the node limits for balanced or fixed graphs. If the linear equation has no exact solutions, we show that the node states can be forced to converge to a ball of fixed but arbitrarily small radius surround the least squares solution by taking the gain of the consensus dynamics to be sufficiently large for consensus + projection flow under fixed and undirected graphs. Semi-global convergence to approximate least squares solutions is established for switching balanced directed graphs under suitable conditions. A minor, but more explicit result arises where it is also shown that the projection consensus flow drives the average of the node states to the least squares solution with complete communication graphs. These results rely on our development of new technique of independent interest for treating the interplay between consensus dynamics and the projections onto affine subspaces in the absence of intersection boundedness and interior point assumptions. All the convergence results can be sharpened to provide exponential convergence with suitable conditions on the switching graph signals. The remainder of this paper is organized as follows. Section introduces the network model, presents the distributed flows under consideration, and defines the problem of interest. Section 3 discusses the scenario where the linear equation has at least one solution. Section 4 then turns to the least squares case where the linear equation has no solution at all. Finally, Section 5 presents a few numerical examples illustrating the established results and Section 6 concludes the paper with a few remarks. Notation and Terminology A directed graph digraph is an ordered pair of two sets denoted by G = V, E []. Here V = {1,..., N} is a finite set of vertices nodes. Each element in E is an ordered pair of two distinct nodes in V, called an arc. A directed path in G with length k from v 1 to v k+1 is a sequence of distinct nodes, v 1 v... v k+1, such that v m, v m+1 E, for all m = 1,..., k. A digraph G is termed strongly connected if for any two 3

distinct nodes i, j V, there is a path from i to j. A digraph is called bidirectional when i, j E if and only if j, i E for all i and j. A strongly connected bidirectional digraph is simply called connected. All vectors are column vectors and denoted by bold, lower case letters, i.e., a, b, c, etc.; matrices are denoted with bold, upper case letters, i.e., A, B, C, etc.; sets are denoted with A, B, C, etc. Depending on the argument, stands for the absolute value of a real number or the cardinality of a set. The Euclidean inner product between two vectors a and b in R m is denoted as a, b, and sometimes simply a T b. The Euclidean norm of a vector is denoted as. Problem Definition.1 Linear Equations Consider the following linear algebraic equation: z = Hy 1 with respect to variable y R m, where H R N m and z R N. We know from the basics of linear algebra that overall there are three cases. I There exists a unique solution satisfying Eq. 1: rankh = m and z spanh. II There is an infinite set of solutions satisfying Eq. 1: rankh < m and z spanh. III There exists no solution satisfying Eq. 1: z / spanh. We denote h T 1 h T N z 1 z N h H = T z, z =.. with h T i being the i-th row vector of H. For the ease of presentation and with inessential loss of generality we assume throughout the rest of the paper that hi = 1, i = 1,..., N. Introduce } A i := {y : h T i y = z i for each i = 1,..., N, which is an affine subspace. It is then clear that Case I is equivalent to the condition that A := N A i is a singleton, and that Case II is equivalent to the condition that 4

A := N A i is an affine space with a nontrivial dimension. For Case III, a least squaress solution of 1 can be defined via the following optimization problem: which yields a unique solution y = H T H 1 Hz if rankh = m. min y R m z Hy, Consider a network with nodes indexed in the set V = { 1,..., N }. Each node i has access to the value of h i and z i without the knowledge of h j or z j from other nodes. Each node i holds a state x i t R m and exchanges this state information with other neighbor nodes, these being determined by the edges of the graph of the network. We are interested in distributed flows for the x i t that asymptotically solve the equation 1, i.e., x i t approaches some solution of 1 as t grows.. Network Communication Structures Let Θ denote the set of all directed graphs associated with node set V. Node interactions are described by a signal σ : R 0 Θ. The digraph that σ defines at time t is denoted as G σt = V, E σt, where E σt is the set of arcs. The neighbor set of node i at time t, denoted N i t, is given by } N i t := {j : j, i E σt. This is to say, at any given time t, node i can only receive information from the nodes in the set N i t. Let R 0 and R + be the sets of nonnegative and positive real numbers, respectively. Associated with each ordered pair j, i there is a function a ij : R 0 R + representing the weight of the possible connection j, i. We impose the following assumption, which will be adopted throughout the paper without specific further mention. Weights Assumption. The function a ij is continues except for at most a set with measure zero over R 0 for all i, j V; there exists a > 0 such that a ij t a for all t R 0 and all i, j V. Denote I j,i Eσt as an indicator function for all i j V, where I j,i Eσt = 1 if j, i E σt and I j,i Eσt communication structures. = 0 otherwise. We impose the following definition on the connectivity of the network Definition 1 i An arc j, i is said to be a δ-arc of G σt for the time interval [t 1, t if t t 1 a ij ti j,i Eσt dt δ. ii G σt is δ-uniformly jointly strongly connected δ-ujsc if there exists T > 0 such that the δ-arcs of G σt on time interval [s, s + T form a strongly connected digraph for all s 0; iii G σt is δ-bidirectionally infinitely jointly connected δ-bijc if G σt is bidirectional for all t 0 and the δ-arcs of G σt on time interval [s, form a connected graph for all s 0. 5

Figure 1: An illustration of the consensus + projection flow left and the projection consensus flow right. The blue arrows mark the vector of ẋ i for the two flows, respectively..3 Distributed Flows Define the mapping P Ai : R m R m as the projection onto the affine subspace A i. Let K > 0 be a given constant. We consider the following continuous-time network flows: [ Consensus + Projection Flow]: ẋ i = K [ Projection Consensus Flow]: ẋ i = The first term of 3 is a standard consensus flow [8], while along a ij t x j x i + P Ai x i x i, i V; 3 a ij t P Ai x j P Ai x i, i V. 4 ẋ i = P Ai x i x i, 5 x i t will be asymptotically projected onto A i since 5 is a gradient flow with P Ai v v = v A i /, where v A i is a C convex function. The flow 4 simply replaces the relative state in standard consensus flow with relative projective state. Notice that a particular equilibrium point of these equations is given by x 1 = x = = x N = y, where y is a solution of Eq. 1. The aim of the two flows is to ensure that the x i t asymptotically tend to a solution of Eq. 1. Note that, the projection consensus flow 4 can be rewritten as 1 ẋ i = P Ai 1 A rigorous treatment will be given later in Lemma 4. a ij tx j x i a ij t P Ai 0. 6

Therefore, in both of the consensus + projection and the projection consensus flows, a node i essentially only receives the information a ij t x j t x i t from its neighbors, and the flows can then be utilized in addition with the h i and z i it holds. In this way, the flows 3 and 4 are distributed. Without loss of generality we assume the initial time is t 0 = 0. We denote by xt = x T 1 t xt N tt a trajectory along any of the two flows..4 Discussions.4.1 Geometries The two network flows are intrinsically different in their geometries. In fact, the consensus + projection flow is a special case of the optimal consensus flow proposed in [3] consisting of two parts, a consensus part and another projection part. The projection consensus flow, first proposed and studied in [39] for fixed bidirectional graphs, is the continuous-time analogue of the projected consensus algorithm proposed in []. Because it is a gradient descent on a Riemannian manifold if the communication graph is undirected and fixed, there is guaranteed convergence to an equilibrium point. The two flows are closely related to the alternating projection algorithms, first studied by von Neumann in the 1940s [5]. We refer to [9] for a thorough survey on the developments of alternating projection algorithms. We illustrate the intuitive difference between the two flows in Figure 1..4. Relation with Previous Work The dynamics described in systems 3 and 4 are linear, in contrast to the nonlinear dynamics associated with the general convex sets studied in [, 3]. However, we would like to emphasize that new challenges arise with the systems 3 and 4 compared to the work of [,3]. First of all, for both of the Cases I and II, A := N A i contains no interior point. This interior point condition however is essential to the results in []. Next, for Case II where A := N A i is an affine space with a nontrivial dimension, the boundedness condition for the intersection of the convex sets no longer holds, which plays a key role in the analysis of [, 3]. Finally, for Case III, A := N A i becomes an empty set. The least squares solution case then completely falls out from the discussions of constrained consensus in [, 3]..4.3 Grouping of Rows We have assumed hitherto that each node i only has access to the value of h i and z i from the equation 1 and therefore there are a total of N nodes. Alternatively and as a generalization, there can be 7

n N nodes with node i holding n i 1 rows, denoted h i1 z i1,..., h ini z ini of H z with ik {1,..., Z}. In this case, we can revise the definition of A i to } A i := {y : h T ik y = z ik, k = 1,..., n i, which is nonempty if 1 has at least one solution. Then the two flows 3 and 4 can be defined in the same manner for the n nodes. Of course, if n i is large, finding an initial condition consistent with A i becomes computationally more burdensome. Let n { i1,..., ini } = { 1,..., Z }. Note that, the A i are still affine subspaces, while solving 1 exactly continues to be equivalent to finding a point in n A i. Consequently, all our results for Cases I and II apply also to this new setting with row grouping. 3 Exact Solutions In this section, we show how the two distributed flows asymptotically solve the equation 1 under quite general conditions for Cases I and II. 3.1 Singleton Solution Set We first focus on the case when Eq. 1 admits a unique solution y, or equivalently, A := N A i = {y } is a singleton. For the consensus + projection flow 3, the following theorem holds. Theorem 1 Let I hold. Then along the consensus + projection flow 3, there holds lim x it = y, i V for all initial values if G σt is either δ-ujsc or δ-bijc. The projection consensus flow 4, however, can only guarantee local convergence for a particular set of initial values. The following theorem holds. Theorem Let I hold. Suppose x i 0 A i for all i. Then along the projection consensus flow 4, there holds if G σt is either δ-ujsc or δ-bijc. lim x it = y, i V Note that, it is not necessary to require the { i1,..., in i } to be disjoint. 8

Remark 1 Convergence along the projection consensus flow relies on specific initial values due to the existence of equilibriums other than the desired consensus states within the set A: if x i 0 are all equal, then obviously they will stay there for ever along the flow 4. It was suggested in [39] that one can add another term in the projection consensus flow and arrive at ẋ i = a ij t P Ai x j P Ai x i + P Ai x i x i 6 for i V, then convergence will be global under 6. We would like to point out that 6 has a similar structure as the consensus + projection flow with the consensus dynamics being replaced by projection consensus. In fact, our analysis of the consensus + projection flow developed in this paper can be refined and generalized to the flow 6 and then leads to results under the same graph conditions for both exact and least squares solutions. 3. Infinite Solution Set We now turn to the scenario when Eq. 1 has an infinite set of solutions, i.e., A := N A i is an affine space with a nontrivial dimension. We note that in this case A is no longer a bounded set; nor does it contain interior points. This is in contrast to the situation studied in [, 3]. For the consensus + projection flow 3, we present the following result. Theorem 3 Let II hold. Then along the consensus + projection flow 3 and for any initial value x0, there exists y x0, which is a solution of 1, such that if G σt is either δ-ujsc or δ-bijc. lim x it = y x0, i V For the projection consensus flow, convergence relies on restricted initial nodes states. Theorem 4 Let II hold. Then along the projection consensus flow 4 and for any initial value x0 with x i 0 A i for all i, there exists y x0, which is a solution of 1, such that if G σt is either δ-ujsc or δ-bijc. lim x it = y x0, i V 3.3 Discussion: Convergence Speed/The Limits For any given graph signal G σt, the value of y x0 in Theorems 3 and 4 depends only on the initial value x0. We manage to provide a characterization to y x0 for balanced switching graphs or fixed graphs. Denote P A as the projection operator over A. 9

Theorem 5 The following statements hold for both the consensus + projection and the projection consensus flows. i Suppose G σt is balanced, i.e., a ijt = i N j t a jit for all t 0 and for all i, j V. Suppose in addition that G σt is either δ-ujsc or δ-bijc. Then lim x it = N P A xi 0 /N, i V. ii Suppose G σt G for some fixed, strongly connected, digraph G and for any i, j V, a ij t a ij for some constant a ij. Let w := w 1... w N T with N w i = 1 be the left eigenvector corresponding to the simple eigenvalue zero of the Laplacian 3 L of the digraph G. Then we have lim x it = N w i P A xi 0, i V. Due to the linear nature of the systems 3 and 4, it is straightforward to see that the convergence stated in Theorems 1,, 3 and 4 is exponential if G σt is periodic and δ-ujsc. It is however difficult to provide tight approximations of the exact convergence rates because of the general structure with switching interactions adopted by our network model. For time-invariant networks with constant edge weights, the two flows become linear time-invariant, and then the convergence rate is determined by the spectrum of the state transition matrix: the rate of convergence for the consensus + projection flow is determined together by both the graph Laplacian and the structure of the linear manifolds A i and one in fact cannot dominate another; the rate of convergence for the projection consensus flow is upper bounded by the smallest nonzero eigenvalue of the graph Laplacian [39]. 3.4 Preliminaries and Auxiliary Lemmas Before presenting the detailed proofs for the stated results, we recall some preliminary theories from affine subspaces, convex analysis, invariant sets, and Dini derivatives, as well as a few auxiliary lemmas which will be useful for the analysis. 3.4.1 Affine Spaces and Projection Operators An affine space is a set X that admits a free transitive action of a vector space V. A set K R d is said to be convex if 1 λx + λy K whenever x K, y K and 0 λ 1 [4]. For any set S R d, the intersection of all convex sets containing S is called the convex hull of S, denoted by cos. Let K be a 3 The Laplacian matrix L associated with the graph G under the given arc weights is defined as L = D A where A = [I j,i E a ij ] and D = diag N j=1 I j,1 E a 1j,..., N j=1 I j,n E a Nj. In fact, we have wi > 0 for all i V if G is strongly connected [6]. 10

closed convex subset in R d and denote x K. = miny K x y as the distance between x R d and K. There is a unique element P K x K satisfying x P K x = x K associated to any x R d [3]. The map P K is called the projector onto K 4. The following lemma holds [3]. Lemma 1 i P K x x, P K x y 0, x R d, y K. ii P K x P K y x y, x, y R d. iii The function x K is continuously differentiable at every x Rd with x K = x P K x. 3.4. Invariant Sets and Dini Derivatives Consider the following autonomous system ẋ = fx, 7 where f : R d R d is a continuous function. Let xt be a solution of 7 with initial condition xt 0 = x 0. Then Ω 0 R d is called a positively invariant set of 7 if, for any t 0 R and any x 0 Ω 0, we have xt Ω 0, t t 0, along every solution xt of 7. as The upper Dini derivative of a continuous function h : a, b R a < b at t is defined D + ht = lim sup s 0 + ht + s ht. s When h is continuous on a, b, h is non-increasing on a, b if and only if D + ht 0 for any t a, b. Hence if D + ht 0 for all t 0 and ht is lower bounded then the limit of ht exists as a finite number when t tends to infinity. The next result is convenient for the calculation of the Dini derivative [7]. Lemma Let V i t, x : R R d R i = 1,..., n be a continuously differentiable function with respect to both t and x and V t, x = max,...,n V i t, x. Let xt be an absolutely continuous function. If It = {i {1,,..., n} : V t, xt = V i t, xt} is the set of indices where the maximum is reached at t, then D + V t, xt = max i It Vi t, xt. 3.4.3 Key Lemmas The following lemmas, Lemmas 3, 4, 5 can be easily proved using the properties of affine spaces, which turn out to be useful throughout our analysis. We therefore collect them below and the details of their proofs are omitted. Lemma 3 Let K := {y R m : a T y = b} be an affine subspace, where a R m with a = 1, and b R. Denote P K : R m K as the projection onto K. Then P K y = I aa T y + ba for all y R m. Consequently there holds P K y u = P K y P K u + P K 0 for all y, u R m. 4 The projections P Ai and P A introduced earlier are consistent with this definition. 11

Lemma 4 Let K be an affine subspace in R m. Take an arbitrary point k K and define Y k := {y k : y K}, which is a subspace in R m. Denote P K : R m K and P Yk : R m Y k as the projectors onto K and Y k, respectively. Then P Yk y u = P K y P K u for all y, u R m. Lemma 5 Let K 1 and K be two affine subspaces in R m with K 1 K. Denote P K1 : R m K 1 and, y P K : R m K as the projections onto K 1 and K, respectively. Then P K1 y = P K1 PK y R m. Lemma 6 Suppose either I or II holds. Take y as an arbitrary solution of 1 and let r > 0 be { arbitrary. Define M r := w R m : w y } r. Then i M r N = M r M r is a positively invariant set for the consensus + projection flow 3; ii M r N A1 A N is a positively invariant set for the projection consensus flow 4. Proof. Let xt = x T 1 t... xt N tt be a trajectory along the flows 3 or 4. Define f 1 t := max i V x i t y. Denote It := { i : f t = x i t y }. i From Lemma we obtain that along the flow 3 D + f t = max x i t y, K i It [ a = max K a ij t i It [ b max K a ij t i It [ c max h Ti x i t y ] i It a ij t x j t x i t + P Ai x i t x i t x i t y, x j t y x i t y h T i x i t y ] x j t y x i t y h T i x i t y ] 0, 8 where a follows from the fact that x i t y, P Ai x i t x i t = h T i x it y in view of Lemma 3 and the fact that y is a solution of 1; b makes use of the elementary inequality α, β α + β /, α, β R m ; 9 c follows from the definition of f t and It. We therefore know from 8 that f t is always a non-increasing functions. This implies that, if xt 0 M r N, i.e., xi t 0 y r for all i, then xi t y r for all i and all t t0. Equivalently, M r N is a positively invariant set for the flow 3. 1

ii Let xt 0 A 1 A N. The structure of the flow 4 immediately tells that xt A 1 A N for all t t 0. Furthermore, again by Lemma we obtain D + f t = max x i t y, i It a = max i It b = max i It [ c max i It d [ max i It a ij t x i t y, I h i h T i a ij t P Ai x j P Ai x i xj t x i t a ij t x i t y T xj I h i h T i t y x i t y a ij t I h i h T i x j t y I h i h T i x i t y ] a ij t x j t y x i t y ] 0, 10 where a follows from Lemma 3, b holds due to the fact that I h i h T i is a projection matrix, c is again based on the inequality 9, and d holds because x i t A i so that I h i h T i xi t y = x i t y and that I h i h T i is a projection matrix so that I h i h T i xj t y x j t y. We can therefore readily conclude that M r N A1 A N is a positively invariant set for the projection consensus flow 4. 3.5 Proofs of Statements We now have the tools in place to present the proofs of the stated results. 3.5.1 Proof of Theorems 1 and When I holds, A := N A i = {y } is a singleton, which is obviously a bounded set. The consensus + projection flow 3 is a special case of the optimal consensus flow proposed in [3]. Theorem 1 readily follows from Theorems 3.1 and Theorems 3. in [3] by adapting the treatments to the Weights Assumption adopted in the current paper using the techniques established in [1]. We therefore omit the details. Being a special case of Theorem 4, Theorem holds true as a direct consequence of Theorem 4, whose proof will be presented below. 3.5. Proof of Theorem 3 Recall that y is an arbitrary solution of 1. With Lemma 6, for any initial value x0, the set M max i V x i0 N, 13

which is obviously bounded, is a positively invariant set for the consensus + projection flow 3. Define A i := A i M max i V x i0, i = 1,..., N and A := A M max i V x i 0. As a result, recalling Theorem 3.1 and Theorem 3. from [3] 5, if G σt is either δ-ujsc or δ-bijc, there hold i lim x i t A = 0 for all i V; ii lim x i t x j t = 0 for all i and j. We still need to show that the limits of the x i t indeed exist. Introduce { } S i := y : h T i y = 0, i V and S := N S i. With Lemma 5, we see that P S PSi x i y P S xi y = P S xi y P S xi y = 0. 11 As a result, taking P S from both the left and right sides of 3 6, we obtain d dt P Sx i t y = K = K a ij t P S x j t y P S x i t y + P S PSi x i y P S xi y a ij t P S x j t y P S x i t y. 1 This is to say, if G σt is either δ-ujsc or δ-bijc, we can invoke Theorem 4.1 for δ-ujsc graphs and Theorem 5. for δ-bijc graphs in [1] to conclude: for any initial value x0, there exists p 0 x0 S such that lim P Sx i t y = p 0, i = 1,..., N. 13 While on the other hand lim x i t A lim x i t A = 0 leads to xi 0 = lim t P A x i t xi = lim t y PA x i t y = lim xi t y P S x i t y, 14 where the last equality follows from Lemma 4. We conclude from 13 and 14 that We have now completed the proof of Theorem 3. lim x it = p 0 + y := y x0, i V. 15 5 Again, the arguments in [3] were based on switching graph signals with dwell time and absolutely bounded weight functions. We can however borrow the treatments in [1] to the generalized graph and arc weight model considered here under which we can rebuild the results in [3]. More details will be provided in the proof of Theorem 4. 6 Note that, P S is a projector onto a subspace. 14

3.5.3 Proof of Theorem 4 We continue to use the notation A i := A i M max i V x i0, i = 1,..., N and A := A M max i V x i 0 introduced in the proof of Theorem 3. In view of Lemma 6, for any initial value x0, A 1 A N is a positively invariant set for the projection consensus flow 4. Define h t := max i V Let I 0 t = { i : h t = xi t A }. We obtain from Lemma 1.iii that D + h t = max x i t P A x i t, i I 0 t 1 x i t. 16 A a ij t P Ai x j P Ai x i. 17 In 17, the argument used to establish 10 can be carried through when y is replaced by P A x i t. Accordingly, we obtain 7, D + h t 0. This immediately implies that there is a constant h 0 such that lim h t = h /. As a result, for any ɛ > 0, there exists T ɛ > 0 such that x i t A h + ɛ, t T ɛ. 18 Now that h is a nonnegative constant satisfying lim h t = h /, we can in fact show that h = 0 with suitable graph conditions, as summarized in the following two technical lemmas. Details of their proofs can be found in the Appendix. Lemma 7 Let G σt be δ-ujsc. Then h = 0. In fact, h = 0 due to the following two contradictive conclusions: i If h > 0, then there holds lim x i t A = h for all i V along the projection consensus flow 4. ii If lim x i t A = h for all i V along the projection consensus flow 4, then h = 0. Lemma 8 Suppose G σt is δ-bijc. Then h = 0 along the projection consensus flow 4. Recall y A. Applying P S to both the left and right sides of 4, we have d dt P Sx i y = a ij t P S x j y P S x i y 19 7 Note that the inequalities in 10 hold without relying on the fact that y is a constant. 15

where we have used Lemma 4 to derive P S PAi x j y = P A PAi x j y = P A x j y = P S x j y. 0 We can again make use of the argument in the proof of Theorem 3 and conclude that the limits of the x i t exist and they are obviously the same. We have now completed the proof of Theorem 4. 3.5.4 Proof of Theorem 5 We provide detailed proof for the projection+consensus flow. The analysis for the projection consensus flow can be similarly established in view of 19. i With 1, we have d dt P Sx i t y = K a ij t P S x j t y P S x i t y 1 along the projection+consensus flow. Note that 1 is a standard consensus flow with arguments being P S x i t y, i = 1,..., N. As a result, there holds lim P Sx i t y = = N P S x i 0 y /N N PA x i 0 y /N if G σt is balanced for all t [10]. As a result, similar to 15, we have lim x it = p 0 + y := = N PA x i 0 y /N + y N P A x i 0/N, i V. ii If G σt G for some fixed digraph G and for any i, j V, a ij t a ij for some constant a ij, then along 1 we have [10] lim P Sx i t y = N w i P S x i 0 y = N w i P A x i 0 y, where w := w 1... w N T is the left eigenvector corresponding to eigenvalue zero for the Laplacian matrix L. This immediately gives us which completes the proof. lim x it = N w i P A x i 0, i V, 3 16

4 Least Squares Solutions In this section, we turn to Case III and consider that 1 admits a unique least squares solution y. Evidently, neither of the two continuous-time distributed flows 3 and 4 in general can yield exact convergence to the least squares solution of 1 since, even for a fixed interaction graph, y is not an equilibrium of the two network flows. It is indeed possible to find the least squares solution using double-integrator node dynamics [40, 45]. However, the use of double integrator dynamics was restricted to networks with fixed and undirected or balanced communication graphs [40, 45]. In another direction, one can also borrow the idea of the use of square-summable step-size sequences with infinite sums in discrete-time algorithms [] and build the following flow ẋ i = K a ij t x j x i + 1 t PAi x i x i, 4 for i V. The least squares case can then be solved under graph conditions of connectedness and balance [], but the convergence rate is at best O1/t. This means 4 will be fragile against noises. For the projection+consensus flow, we can show that under fixed and connected bidirectional interaction graphs, with a sufficiently large K, the node states will converge to a ball around the least squares solution whose radius can be made arbitrarily small. This approximation is global in the sense that the required K only depends on the accuracy between the node state limits and the y. Theorem 6 Let III hold with rankh = m and denote the unique least squares solution of 1 as y. Suppose G σt = G = V, E for some bidirectional, connected graph G and for any i, j V, a ij t = a ji t a ij for some constant a ij. Then along the flow 3, for any ɛ > 0, there exists K ɛ > 0 such that x := lim xt exists and xi y ɛ, i V for any initial value x0 if K K ɛ. The intuition behind Theorem 6 can be described briefly as follows see its proof presented later for a full exposure of this intuition. With fixed and undirected network topology, the consensus + projection flow 3 is a gradient flow in the form of ẋ i = xi D K x where D K x := 1 N xi + K A i {i,j} E a ij xj x i. 5 17

There holds N x i A i = h T i x i z i if each hi is normalized with a unit length. Therefore, for large K, the trajectory of 3 will asymptotically tend to be close to the solution of the following optimization problem: min y 1,...,y N R m N h T i y i z i s.t. y 1 = = y N. Any solution of the above problem consists of N copies of the solution to min Hy z y R m which is the unique least squares solution y = H T H 1 Hz if rankh = m. Therefore, a large K can drive the node states to somewhere near y along the flow 3 as stated in Theorem 6. Remark Least Squares Solutions without Normalizing h i From the above argument, for the case when h i is not normalized, we can replace the consensus + projection flow 3 with ẋ i = K a ij t x j x i xi h T i x i z i / = K a ij t x j x i h i h T i x i z i for all i V. The statement of Theorem 6 will continue to hold for the flow 6. Remark 3 Tradeoff between Convergence Rate and Accuracy Under the assumptions of Theorem 6, the flow defines a linear time-invariant system ẋ = KL I m + J x + h z 7 where L is the Laplacian of the graph G, J = diagh 1 h T 1,, h Nh T N is a block-diagonal matrix, and h z = z 1 h T 1 z Nh T N T. Therefore, the rate of convergence is given by λ min KL I m + J, which is influenced by both L the graph and J the linear equation. We also know that λ min KL I m + J in general cannot grow to infinity by increasing K due to the presence of the zero eigenvalue in L. One can however speed up the convergence by adding a weight γ to the projection term in the flow 3 and obtain ẋ i = K 6 a ij t x j x i + γ PAi x i x i, i V. 8 Certainly the convergence rate to a ball centered at y with radius ɛ of 8 can be arbitrarily large by selecting sufficiently large K and γ. However, it is clear from the proof of Theorem 6 that the required K for a given accuracy ɛ will in turn depend on γ and require a large K for a large γ. 18

We also manage to establish the following semi-global result for switching but balanced graphs under two further assumptions. [A1] The set Wy := { P AiJ P Ai1 y : i 1,..., i J V, J 1 } is a bounded set. [A] N P Ai 0 = 0. Theorem 7 Let III hold with rankh = m and denote the unique least squares solution of 1 as y. Assume [A1] and [A]. Suppose G σt is balanced for all t R + and δ-ujsc with respect to T > 0. Then for any ɛ > 0 and any x0 A 1 A N, there exist two constants K ɛ, x0 > 0 and T ɛ, x0 such that when K K and T T, there holds along the flow 3 with the initial value x0. lim sup xi t y ɛ, i V Remark 4 The two assumptions, [A1] and [A] are indeed rather strong assumptions. In general, [A1] holds if N A i is a nonempty bounded set [9], which is exactly opposite to the least squares solution case under consideration. We conjecture that at least for m = case, [A1] should hold when h 1,..., h N are distinct vectors. The assumption [A] requires a strong symmetry in the affine spaces A i, which turns out to be essential for the result to stand. For the projection consensus flow, we present the following result. Theorem 8 Let III hold with rankh = m and denote the unique least squares solution of 1 as y. Suppose G σt = G is fixed, complete, and a ij t a > 0 for all i, j V. Then for any initial value x0 A 1 A N, there holds along the flow 4. N lim x it = y 9 N Based on simulation experiments, the requirement that G σt = G does not have to be complete for the Theorem 8 to be valid for certain choices of the A i. Moreover, because the flow 4 is a gradient flow over the manifold A 1 A N, the x i t will converge but perhaps to different limits at different nodes. 4.1 Proof of Theorem 6 Suppose G σt = G for some bidirectional, connected graph G and for any i, j V, a ij t = a ji t a ij for some constant a ij. Fix ɛ > 0. Note that, we have K j N i a ij xj x i + PAi x i x i = xi D K x. 30 19

Therefore, the flow 3 is a gradient flow written as ẋ = D K x where D K x is a C convex function. Denote Z K := { x : D K x = 0 }. There must hold The following lemma holds. Lemma 9 Suppose rankh = m. Then lim xt ZK = 0. 31 i Z K is a singleton, which implies that xt converges to a fixed point asymptotically; ii K>κ 0 Z K is a bounded set for all κ 0 > 0. Proof. i Recall that L is the Laplacian of the graph G, J = diagh 1 h T 1,, h Nh T N is a block-diagonal matrix, and h z = z 1 h T 1 z Nh T N T. With Lemma 3, the equation x D K x = 0 can be written as K j N i a ij x j x i h i h T i x i = z i h i, i V, 3 or, in a compact form, Consider KL I m + J x = h z. 33 Q K x := x T KL I m + J x = N h T i x i + K {i,j} E a ij x j x i. We immediately know from the second term of Q that Qx = 0 only if x 1 = = x N. On the other hand, obviously N h T i w > 0 for any w 0 R m if rankh = m. Therefore, KL I m + J is positive-definite, which yields { 1hz } Z K = KL I m + J. ii By the Courant-Fischer Theorem see Theorem 4..11 in [5], we have λ min KL I m + J = min Q Kx x =1 [ N ] = min h T i x i + K a ij xj x i. 34 x =1 {i,j} E 0

This immediately implies λ min KL I m + J λ min κ 0 L I m + J > 0, 35 for all K κ 0 and consequently, K>κ 0 Z K is obviously a bounded set. This proves the desired lemma. Now we introduce Z := K 1 Z K. Let w = w T 1... wt N T with w i R m and define B 0 := sup max w Z i V P Ai w i w i 36 and C 0 := sup max w Z i V y w i 37 with y being the unique least squares solution of 1. We see that B 0 and C 0 are finite numbers due to the boundedness of Z. The remainder of the proof contains two steps. Step 1. Let vk = v1 TK... v NK T = KL I m + J 1 hz Z K. Then v satisfies 3, or in equivalent form, v1 P A1 v 1 T KL v P I m v = A v T. 38. vn P AN v N T Denote 8 v ave K = N v ik. Let M := {w = w T 1... w T N T : w 1 = = w N }. 39 be the consensus manifold. Denote λ L > 0 as the second smallest eigenvalue of L. We can now conclude that Kλ L v M a b = KL I m v N v i P Ai v i c NB 0 40 where a holds from the fact that the zero eigenvalue of L corresponds to a unique unit eigenvector whose entries are all the same, b is from 38, and c holds from the definition of B 0 and the fact that 8 In the rest of the proof we sometimes omit K in v ik, vk, and v avek in order to simplify the presentation. One should however keep in mind that they always depend on K. 1

v Z K. This allows us to further derive and then Kλ L v M = Kλ L N v i v ave Therefore, for any ς > 0 we can find K 1 ς > 0 that for all K K 1 ς. Step. Applying Lemma 3, we have N v i v ave NB 0, 41 NB 0 Kλ L. 4 N vi K v ave K ς, 43 Uy := z Hy = N z i h T i y = N y. 44 A i Then with 43 and noticing the continuity of U, for any ς > 0, there exists ϱ > 0 such that if Uv i Uv ave ς, i = 1,..., N 45 N vi K v ave K ϱ. 46 Consequently, we can conclude without loss of generality 9 that for any ς > 0 we can find K 1 ς > 0 so that both 43 and 45 hold when K K 1 ς. Now noticing 1 T L = 0, from 38 we have Therefore, with 43, we can find another K ς such that for all K K ς. N vi P Ai v i = 0. 47 N vave P Ai v ave ς/c0 48 9 If ϱ ς then we can replace ϱ with ς in 46 with 45 continuing to hold; if ϱ < ς then both 43 and 45 hold directly.

Take K ɛ = max{1, K 1 ɛ/, K ɛ/4}. We can finally conclude that Uy Uv i Uy Uv ave + ɛ b Uv ave y v ave + ɛ c N = vave P Ai v ave y v ave + ɛ a d ɛ C 0 y v ave + ɛ e ɛ 49 for all i V, where a is from 45, b is from the convexity of U, c is based on direct computation of U in 44, d is due to 48, and e holds because y v ave C 0 from the definition of C 0 and v ave. This completes the proof of Theorem 6. 4. Proof of Theorem 7 Consider the following dynamics for the considered network model: q i = K a ij t q j q i + wi t, i V 50 where q i R, K > 0 is a given constant, a ij t are weight functions satisfying our standing assumption, and w i t is a piecewise continuous function. The proof is based on the following lemma on the robustness of consensus subject to noises, which is a special case of Theorem 4.1 and Proposition 4.10 in [1]. Lemma 10 Let G σt be δ-ujsc with respect to T > 0. Then along 50, there holds that for any ɛ > 0, there exist a sufficiently small T ɛ > 0 and sufficiently large K ɛ such that lim sup qi t q j t ɛ wt t + for all initial value q 0 when K K ɛ and T T ɛ, where wt := max i V sup t [0, w i t. With Assumption [A1], the set x0 := [ N co ] N Wx i0 is a compact set, and is obviously positively invariant along the flow 3. Therefore, we can define { D x0 := max sup PAi } u i u i : u = u T 1... u T N T x0. i V as a finite number. Now along the trajectory xt of 3, we have P Ai x i t x i t D x0 3

for all t 0 and all i V. Invoking Lemma 10 we have for any ɛ > 0 and any initial value x0, there exist K 0 ɛ, x0 > 0 and T 0 ɛ, x0 such that if K K 0 and T T 0. Furthermore, with Lemma 4, we have which implies d PAi x i x i = K dt K a ij t x j x i lim sup x i t x j t ɛ, i, j V a ij t P Ai x j P Ai x i + 1 + K a ij t PAi 0 PAi x i x i, 51 d dt N N PAi x i x i = PAi x i x i if Assumption [A] holds and G σt is balanced. While if x0 A 1 A N, then 5 N PAi x i 0 x i 0 = 0. This certainly guarantees N PAi x i t x i t fact that lim sup x i t y ɛ, i V = 0 for all t 0 in view of 5. The proof for the can then be built using exactly the same analysis as the final part of the proof of Theorem 6. We have now completed the proof of Theorem 7. 4.3 Proof of Theorem 8 The proof is based on the following lemma. Lemma 11 There holds P Am N x i/n = N P Am x i /N for all m V. Proof. From Lemma 4 we can easily know P K y + u = P K y + P K u P K 0 for all y, u R m. As a result, we have On the other hand, N P Am x i = NP Am N x i /N N 1P K 0. 53 N P Am x i = N P Am x i N 1P K 0. 54 4

Figure : The trajectories of the node states along the consensus + projection flow left and the projection consensus flow right. The desired lemma thus holds. Suppose G σt = G is fixed, complete, and a ij t a > 0 for all i, j V. Now along the flow 4, we have d N x i dt N = = N N j=1 N j=1 a P Ai x j P Ai x i /N N a P Ai x j x i /N N [ N = a j=1 P Ai x j N j=1 x ] j N N N = a [P N j=1 x N j j=1 Ai x ] j N N 55 for any initial value x0 A 1 A N. The desired result follows straightforwardly and this concludes the proof of Theorem 8. 5 Numerical Examples In this section, we provide a few examples illustrating the established results. Example 1. Consider three nodes in the set {1,, 3} whose interactions form a fixed, three-node directed cycle. Let m =, K = 1, and a ij = 1 for all i, j V. Take h 1 = 1/ 1/ T, h = 0 1 T, h 1 = 1/ 1/ T and z 1 = 1/, z = 1, z 3 = 1/ so the linear equation 1 has a unique solution at 0, 1 corresponding to Case I. With the same initial value x 1 0 = 1 T, x 0 = 5 1 T, and x 3 0 = 4 3 T, we plot the trajectories of xt, respectively, for the consensus + projection flow 3 5

Figure 3: The trajectories of R i t along the consensus + projection flow left and the projection consensus flow right. and the projection consensus flow 4 in Figure. The plot is consistent with the result of Theorems 1 and. Example. Consider three nodes in the set {1,, 3} whose interactions form a fixed, three-node directed cycle. Let m = 3, K = 1, and a ij = 1 for all i, j V. Take h 1 = 1/ 1/ 0 T, h = 0 1 0 T, h 1 = 1/ 1/ 0 T and z 1 = 1/, z = 1, z 3 = 1/ so corresponding to Case II, the linear equation 1 admits an infinite solution set A := { y R 3 : ht 1 h T y = For the initial value x 1 0 = 1 3 T, x 0 = 1 1 T, and x 3 0 = 1 0 1 T, we plot the trajectories of R i t := x i t zt 1 z T } 3 P A x i 0/3, i = 1,, 3 respectively, for the consensus + projection flow 3 and the projection consensus flow 4 in Figure 3. The plot is consistent with the result of Theorems 3, 4, and 5. Example 3. Consider four nodes in the set {1,, 3, 4} whose interactions form a fixed, four-node undirected cycle. Let m =, and a ij = 1 for all i, j V. Take h 1 = 1/ 1/ T, h = 1/ 1/ T, h 3 = 1/ 1/ T, h 4 = 1/ 1/ T and z 1 = 1/, z 3 = 1/, z 3 = 1/, z 3 = 1/ so corresponding to Case III the linear equation 1 has a unique least squares solution y = 0 0. For the initial value x 1 0 = 0 1 T, x 0 = 1 0 T, x 3 0 = 1 T, and x 4 0 = 1 0 T we plot the trajectories of E K t := 4 x i t y,. 6

Figure 4: The evolution of E K t for K = 1, 5, 100, respectively along the consensus + projection flow. along the consensus + projection flow 3 for K = 1, 5, 100, respectively in Figure 4. The plot is consistent with the result of Theorem 6. 6 Conclusions Two distributed network flows were studied as distributed solvers for a linear algebraic equation z = Hy, where a node i holds a row h T i of the matrix H and the entry z i in the vector z. A consensus + projection flow consists of two terms, one from standard consensus dynamics and the other as projection onto each affine subspace specified by the h i and z i. Another projection consensus flow simply replaces the relative state feedback in consensus dynamics with projected relative state feedback. Under mild graph conditions, it was shown that that all node states converge to a common solution of the linear algebraic equation, if there is any. The convergence is global for the consensus + projection flow while local for the projection consensus flow. When the linear equation has no exact solutions but has a well defined least squares approximate solution, it was proved that the node states can converge to somewhere arbitrarily near the least squares solution as long as the gain of the consensus dynamics is sufficient large for consensus + projection flow under fixed and undirected graphs. It was also shown that the projection consensus flow drives the average of the node states to the least squares solution if the communication graph is complete. Numerical examples were provided verifying the established results. Interesting future direction includes more precise comparisons of the existing continuous-time or discrete-time linear equation solvers in terms of convergence speed, computational complexity, and communication complexity. 7

Appendix A. Proof of Lemma 7. i Suppose h > 0. We show lim x i t A = h for all i V by a contradiction argument. Suppose to obtain a contradiction that there exists a node i 0 V with l := lim inf x i0 t A < h. Therefore, we can find a time instant t > T ɛ with x i0 t A h + l. 56 In other words, there is an absolute positive distance between x i0 t A and the limit h of max V x i t A. Let G σt be δ-ujsc. Consider the N 1 time intervals [t, t + T ],, [t + N T, t + N 1T ]. In view of the arguments in 10, we similarly obtain d x i0 t = x dt A i0 t P A x i0 t, a i0 jt P i0 x j P i0 x i0 j N i0 t a i0 jt x j t x A i0 t, 57 A j N i0 t which leads to d xi0 t dt A j N i0 t a i0 jt h + ɛ x i0 t A 58 for all t T ɛ noticing 18. Denoting t := t + N 1T and applying Grönwall s inequality we can further conclude that x i0 t A e t t j N i0 s a i 0 jsds x i0 t + 1 e t t j N i0 s a i 0 jsds h A + ɛ e t t j N i0 s a i 0 jsds xi0 t A + µ x i0 t A + 1 µ h + ɛ 1 e t t j N i0 s a i 0 jsds h + ɛ µ l + 1 µ h + ɛ 59 for all t [t, t ], where µ = e N 1T a. Now that G σt is δ-ujsc, there must be a node i 1 i 0 for which i 0, i 1 is a δ-arc over the time 8

interval [t, t + T. Thus, there holds d xi1 t xj a dt A i1 jt t A xi1 t A j N i1 t = I i0,i 1 E σt a i1 i 0 t x i0 t x A i1 t + A j N i1 t\{i 0 } µ I i0,i 1 E σt a i1 i 0 t l + 1 µ h + ɛ x i1 t A + a i1 jt h + ɛ x i1 t A j N i1 t\{i 0 } for t [t, t + T ]. Noticing the definition of δ-arcs and that x i1 t A Grönwall s inequality again and conclude from 60 that [e t +T t x i1 t + T µl A + µh + ɛ j N i1 s a i 1 jsds t +T [1 e t +T t µγ l + 1 µγ ] t e t f1sds f 1 tdt t t +T j N i1 s a i 1 jsds t a i1 jt x j t x A i1 t A 60 h + ɛ, we invoke the ] t e t f1sds f 1 tdt h + ɛ 61 where f 1 t := I i0,i 1 E σt a i1 i 0 t and γ = e T a 1 e δ. This further allows us to apply the estimation of x i0 t + T A over the interval [t, t ] to node i 1 for the interval [t + T, t ] and obtain x i1 t + T µ γ A l + 1 µ γ h + ɛ 6 for all t [t + T, t ]. Since G σt is δ-ujsc, the above analysis can be recursively applied to the intervals [t +T, t +T,..., [t +N T, t +N 1T, for which nodes i,..., i N 1 can be found, respectively, with V = {i 0,..., i N 1 } such that x im t A µn 1 γ for m = 0, 1,..., N 1. This implies h t µn 1 γ l + 1 µn 1 γ h + ɛ, 63 l / + 1 µn 1 γ h + ɛ / < h / 64 when ɛ is sufficiently small. However, we have known that non-increasingly there holds lim h t = h /, and therefore such i 0 does not exist, i.e., lim x i t A Lemma 7. i is proved. = h for all i V. The statement of B. Proof of Lemma 7. ii Suppose lim x i t A = h for all i V. Then for any ɛ > 0, there exists ˆT ɛ > 0 such that h ɛ x i t A h + ɛ, t ˆT ɛ. 65 9