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

Relaxation dynamics of SIR-flocks with random epidemic states

Seung-Yeal Ha
Department of Mathematical Sciences and Research Institute of Mathematics,
Seoul National University, Seoul 08826, Republic of Korea
[email protected]
Hansol Park
Department of Mathematical Sciences
Seoul National University, Seoul 08826, Republic of Korea
[email protected]
 and  Seoyeon Yang
Department of Mathematical Sciences
Seoul National University, Seoul 08826, Republic of Korea
[email protected]
Abstract.

We study the collective dynamics of a multi-particle system with three epidemic states as an internal state. For the collective modeling of active particle system, we adopt modeling spirits from the swarmalator model and the SIR epidemic model for the temporal evolution of particles’ position and internal states. Under suitable assumptions on system parameters and non-collision property of initial spatial configuration, we show that the proposed model does not admit finite-time collisions so that the standard Cauchy-Lipschitz theory can be applied for the global well-posedness. For the relaxation dynamics, we provide several sufficient frameworks leading to the relaxation dynamics of the proposed model. The proposed sufficient frameworks are formulated in terms of system parameters and initial configuration. Under such sufficient frameworks, we show that the state configuration relaxes to the fixed constant configuration via the exponentially perturbed gradient system and explicit dynamics of the SIR model. We present explicit lower and upper bounds for the minimal and maximal relative distances.

Key words and phrases:
SIR model, swarmalator, epidemic
2020 Mathematics Subject Classification:
70G60, 34D06, 70F10
Acknowledgment. The work of S.-Y. Ha is supported by National Research Foundation of Korea (NRF-2017R1A5A1015626).

1. Introduction

The purpose of this paper is to continue the studies begun in [19, 20] on the collective dynamcis modeling of active particles with internal states. Collective behaviors of complex systems are ubiquitous in nature, e.g., crowd dynamics [2], aggregation of bacteria [8, 9, 10, 26], flocking of birds [5, 6, 12, 13, 15, 22, 23, 24, 29, 30, 34, 37], synchronization of fireflies [11, 38] and swarming of fish [16], etc. See survey articles [1, 3, 7, 17, 21, 33, 36]. In recent years, thanks to the emerging applications to the decentralized control of multi-particle systems, collective behaviors have received a lot of attention from diverse scientific disciplines such as applied mathematics, biology, control theory and statistical physics etc. In this work, we are interested in the first-order modeling of active particles with random epidemic states (susceptible(SS), infected(II) and recovered(RR)) as an internal state, i.e., we assume that particle ii can take aforementioned three epidemic states with certain probabilities denoted by Si,IiS_{i},I_{i} and RiR_{i} whose precise meaning will be clarified in a minute. Thus, the state of particle ii is represented by position vector xidx_{i}\in\mathbb{R}^{d} and probability vector for epidemic states Wi=(Si,Ii,Ri)[0,1]3W_{i}=(S_{i},I_{i},R_{i})\in[0,1]^{3}, respectively. In what follows, we briefly describe how to model the dynamics of position vectors and epidemic vectors via continuous dynamical systems. Let NN be the size of system, i.e., the total number of particles in a given ensemble {(Wi,xi)}i=1N\{(W_{i},x_{i})\}_{i=1}^{N}.

First, we use the position dynamics xi=xi(t)x_{i}=x_{i}(t) of the swarmalator model [31, 32] which describes the attractive and repulsive forces:

(1.1) x˙i=κ2N1jNjiΨaijxjxixjxiα+κ3N1jNjiΨrijxixjxixjβ,i[N]:={1,2,,N},\dot{x}_{i}=\displaystyle\frac{\kappa_{2}}{N}\sum_{\begin{subarray}{c}1\leq j\leq N\\ j\neq i\end{subarray}}\Psi^{ij}_{a}\frac{x_{j}-x_{i}}{\|x_{j}-x_{i}\|^{\alpha}}+\frac{\kappa_{3}}{N}\sum_{\begin{subarray}{c}1\leq j\leq N\\ j\neq i\end{subarray}}\Psi^{ij}_{r}\frac{x_{i}-x_{j}}{\|x_{i}-x_{j}\|^{\beta}},\quad i\in[N]:=\{1,2,\cdots,N\},

where κ2,κ3\kappa_{2},\kappa_{3} are nonnegative coupling strengths, and Ψaij,Ψrij\Psi^{ij}_{a},~{}\Psi^{ij}_{r} represent attractive and repulsive weights whose explicit functional forms will be discussed in Section 3, and we assume that positive system parameters α,β\alpha,\beta satisfy the relation:

(1.2) 1α<β1\leq\alpha<\beta

so that repulsive force is dominant in a small relative distance regime. Here \|\cdot\| denotes the standard 2\ell^{2}-norm in d{\mathbb{R}}^{d}.

Next, we use the modeling spirit of the SIR model for the dynamics of epidemic state WiW_{i}. We introduce convex set 𝒮{\mathcal{S}} consisting of all admissible state vectors:

(1.3) 𝒮:={W:=(S,I,R)[0,1]3:S+I+R=1},\mathcal{S}:=\Big{\{}W:=(S,I,R)\in[0,1]^{3}:~{}~{}S+I+R=1\Big{\}},

and we set

Si:the probability that the i-th particle is in susceptible state,Ii:the probability that the i-th particle is in infected state,Ri:the probability that the i-th particle is in recovered state,\displaystyle\begin{aligned} &S_{i}:~{}\mbox{the probability that the $i$-th particle is in susceptible state},\\ &I_{i}:~{}\mbox{the probability that the $i$-th particle is in infected state,}\\ &R_{i}:~{}\mbox{the probability that the $i$-th particle is in recovered state},\end{aligned}

and

Wi:=(Si,Ii,Ri):the epidemic probability vector of the i-th particle.W_{i}:=(S_{i},I_{i},R_{i})~{}:~{}\mbox{the epidemic probability vector of the $i$-th particle}.

Then, we assume that the dynamics of WiW_{i} is governed by the following coupled system:

(1.4) S˙i=κ1j=1NaijSiIj,I˙i=κ1j=1NaijSiIjbiIi,R˙i=biIi,i[N].\dot{S}_{i}=-\kappa_{1}\sum_{j=1}^{N}a^{ij}S_{i}I_{j},\quad\dot{I}_{i}=\kappa_{1}\sum_{j=1}^{N}a^{ij}S_{i}I_{j}-b^{i}I_{i},\quad\dot{R}_{i}=b^{i}I_{i},~{}~{}i\in[N].

Finally, we couple two systems (1.1) and (1.4) via nonnegative system functions Ψaij,Ψrij,aij\Psi^{ij}_{a},~{}\Psi^{ij}_{r},~{}a^{ij} and bib^{i} by imposing suitable functional dependences between position variable {xi}\{x_{i}\} and internal variable {Wi}\{W_{i}\}:

(1.5) Ψaij:=Ψa(Wi,Wj),Ψrij:=Ψr(Wi,Wj),aij:=a(xixj),i,j[N].\Psi_{a}^{ij}:=\Psi_{a}(W_{i},W_{j}),\quad\Psi^{ij}_{r}:=\Psi_{r}(W_{i},W_{j}),\quad a^{ij}:=a(\|x_{i}-x_{j}\|),\quad i,j\in[N].

Moreover, we also assume that there exist positive constants εa,εr,Ma\varepsilon_{a},\varepsilon_{r},M_{a} and MrM_{r} such that

0<εaΨaijMa,0<εrΨrijMr,i,j[N].0<\varepsilon_{a}\leq\Psi_{a}^{ij}\leq M_{a},\quad 0<\varepsilon_{r}\leq\Psi_{r}^{ij}\leq M_{r},\quad i,j\in[N].

See Section 3 for explicit functional relations in (1.5).

Finally, we combine all three ingredients (1.1), (1.4) and (1.5) together to write down the dynamical system for (Wi,xi)(W_{i},x_{i}) with suitable initial data:

(1.6) {S˙i=κ1j=1NaijSiIj,I˙i=κ1j=1NaijSiIjbiIi,R˙i=biIi,t>0,i[N],x˙i=κ2N1jNji(Ψaijxjxixjxiα)+κ3N1jNji(Ψrijxixjxixjβ),(Wi,xi)(0)=(Wi0,xi0)𝒮×d.\begin{cases}\displaystyle\dot{S}_{i}=-\kappa_{1}\sum_{j=1}^{N}a^{ij}S_{i}I_{j},\quad\dot{I}_{i}=\kappa_{1}\sum_{j=1}^{N}a^{ij}S_{i}I_{j}-b^{i}I_{i},\quad\dot{R}_{i}=b^{i}I_{i},\quad t>0,~{}~{}i\in[N],\\ \displaystyle\dot{x}_{i}=\displaystyle\frac{\kappa_{2}}{N}\sum_{\begin{subarray}{c}1\leq j\leq N\\ j\neq i\end{subarray}}\left(\Psi^{ij}_{a}\cdot\frac{x_{j}-x_{i}}{\|x_{j}-x_{i}\|^{\alpha}}\right)+\frac{\kappa_{3}}{N}\sum_{\begin{subarray}{c}1\leq j\leq N\\ j\neq i\end{subarray}}\left(\Psi^{ij}_{r}\cdot\frac{x_{i}-x_{j}}{\|x_{i}-x_{j}\|^{\beta}}\right),\\ \displaystyle(W_{i},x_{i})(0)=(W_{i}^{0},x_{i}^{0})\in{\mathcal{S}}\times\mathbb{R}^{d}.\end{cases}

Throughout the paper, we call the coupled system (1.6) as the “SIR-flock model” for simplicity.

The goal of this work is to provide sufficient frameworks leading to the emergent dynamics of the SIR-flock model (1.6). Now, we briefly discuss four main results.

First, we show that if spatial configuration is noncollisional initially, there will be no finite-time collisions so that system (1.6) is globally well-posed by the Cauchy-Lipschitz theory (see Theorem 2.2).

Second, we present a sufficient conditions for the relaxation of epidemic state WiW_{i} toward a constant epidemic state. More specifically, our sufficient framework for the relaxation of WiW_{i} is expressed in terms of network topology (aij)(a^{ij}) and a recovering vector (bi)(b^{i}):

(1.7) aij:={1(xixj+L)γifij,0otherwise,andmin1iNbi>κ1(N1)Lγ,\displaystyle a^{ij}:=\begin{cases}\displaystyle\frac{1}{(\|x_{i}-x_{j}\|+L)^{\gamma}}\quad&\text{if}\quad i\neq j,\\ 0&\text{otherwise},\end{cases}\hskip 28.45274pt\mbox{and}\quad\min_{1\leq i\leq N}b^{i}>\frac{\kappa_{1}(N-1)}{L^{\gamma}},

where LL is a positive constant and γ\gamma is a nonnegative constant. Under this framework (1.7), we show that the epidemic state WiW_{i} relaxes to a constant state (see Theorem 4.1): there exist a constant state Wi=(Si,0,Ri)W_{i}^{\infty}=(S_{i}^{\infty},0,R_{i}^{\infty}) and λ>0\lambda>0 such that

limt(Si(t),Ii(t),Ri(t))=(Si,0,Ri),i[N]and|1Ni=1NIi(t)|eλt,t0.\lim_{t\rightarrow\infty}(S_{i}(t),I_{i}(t),R_{i}(t))=(S^{\infty}_{i},0,R^{\infty}_{i}),\quad i\in[N]\quad\mbox{and}\quad\Big{|}\frac{1}{N}\sum_{i=1}^{N}{I_{i}(t)}\Big{|}\leq e^{-\lambda t},~{}~{}t\geq 0.

Third, we show that if initial configuration and system parameters satisfy suitable conditions, then minimal and maximal relative distances are positive uniformly in time, i.e., there exist positive constants δ1\delta_{1} and δ\delta_{\infty} such that

minijinf0t<xi(t)xj(t)δ1andmaxijsup0t<xi(t)xj(t)δ,\min_{i\not=j}\inf_{0\leq t<\infty}\|x_{i}(t)-x_{j}(t)\|\geq\delta_{1}\quad\mbox{and}\quad\max_{i\not=j}\sup_{0\leq t<\infty}\|x_{i}(t)-x_{j}(t)\|\leq\delta_{\infty},

see Theorem 5.1.

Fourth, we show that the spatial configuration relaxes to a constant configuration asymptotically under a suitable condition on the initial configuration and system parameters (see Theorem 6.1).

The rest of this paper is organized as follows. In Section 2, we briefly review basic properties of the SIR epidemic model and the swarmalator model on the asymptotic relaxation of state variables and present a global well-posedness by verifying nonexistence of finite-time collisions. In Section 3, we discuss the modeling spirit of system parameters and coupling functions. In Section 4, we study the relaxation of epidemic states toward a constant state and in particular, we provide a sufficient condition leading to asymptotic removal of infected particles. In Section 5, we study the existence of positive lower bound and upper bound for the minimal and maximal relative distances, respectively. In Section 6, we study a relaxation of spatial configuration toward a fixed spatial configuration using the perturbed gradient flow theory. In Section 7, we provide several numerical examples and compare them with analytical results obtained in previous sections. Finally, Section 8 is devoted to a brief summary of our main results and some remaining issues for a future work.

2. Preliminaries

In this section, we briefly review basic properties on two related models “the SIR epidemic model” and “the swarmalator model” which correspond to the subsystems of system (1.6).

2.1. The SIR epidemic model

In this subsection, we briefly discuss the SIR model [4, 35] which is a prototype model for the spread of disease or virus. Kermack and McKendrick’s work in 1927 motivated a large number of modelings in epidemics [25]. In particular, there are some of the works for diffusive SIR models on stability and numerical methods [18, 14, 28]. First, we introduce the following three observables:

S\displaystyle S =S(t):ratio of susceptible individuals in a total population,\displaystyle=S(t):\text{ratio of {\it susceptible} individuals in a total population},
I\displaystyle I =I(t):ratio of infected individuals in a total population,\displaystyle=I(t):\text{ratio of {\it infected} individuals in a total population},
R\displaystyle R =R(t):ratio of recovered individuals in a total population.\displaystyle=R(t):\text{ratio of {\it recovered} individuals in a total population}.

Then, we assume that the state (S,I,R)(S,I,R) is governed by the Cauchy problem to the coupled system of ODEs:

(2.1) {S˙=aSI,t>0,I˙=aSIbI,R˙=bI,(S,I,R)(0)=(S0,I0,R0),\displaystyle\begin{cases}\displaystyle\dot{S}=-aSI,\quad t>0,\\ \displaystyle\dot{I}=aSI-bI,\\ \displaystyle\dot{R}=bI,\\ \displaystyle(S,I,R)(0)=(S^{0},I^{0},R^{0}),\end{cases}

where aa and bb are positive constants. This model was designed to explain the spreading of epidemic diseses. We can express the SIR model via the following pictorial diagram.

SSIIRRaSIaSIbIbI

Next, we list basic properties of the SIR model in the following lemma.

Proposition 2.1.

Let (S,I,R)(S,I,R) be a global solution of (2.1) with initial data satisfying

S00,I00,R00andS0+I0+R0=1.S^{0}\geq 0,\quad I^{0}\geq 0,\quad R^{0}\geq 0\quad\mbox{and}\quad S^{0}+I^{0}+R^{0}=1.

Then, one has

S(t)+I(t)+R(t)=1,S(t)0,I(t)0,R(t)0,t0andlimt+I(t)=0.S(t)+I(t)+R(t)=1,\quad S(t)\geq 0,~{}~{}I(t)\geq 0,~{}~{}R(t)\geq 0,\quad t\geq 0\quad\mbox{and}\quad\lim_{t\to+\infty}I(t)=0.
Proof.

(i) We add three equations in (2.1) to get

S˙+I˙+R˙=0.\dot{S}+\dot{I}+\dot{R}=0.

This yields

(2.2) S(t)+I(t)+R(t)=S0+I0+R0=1for all t0.S(t)+I(t)+R(t)=S^{0}+I^{0}+R^{0}=1\quad\mbox{for all $t\geq 0$}.

(ii) We integrate the first two equations in (2.1) to obtain

(2.3) S(t)=S0exp[a0tI(τ)𝑑τ]0,I(t)=I0exp[0t(aS(τ)b)𝑑τ]0.\displaystyle\begin{aligned} S(t)&=S^{0}\exp\Big{[}-a\int_{0}^{t}I(\tau)d\tau\Big{]}\geq 0,\\ I(t)&=I^{0}\exp\Big{[}\int_{0}^{t}\Big{(}aS(\tau)-b\Big{)}d\tau\Big{]}\geq 0.\end{aligned}

The nonnegativity of RR follows from the third equation and the nonnegativity of (2.3)2\eqref{B-3}_{2}:

R˙=bI0.{\dot{R}}=bI\geq 0.

On the other hand, it follows from (2.2) and (2.3) that RR is bounded by 11:

R(t)=1S(t)I(t)1.R(t)=1-S(t)-I(t)\leq 1.

Since R˙0\dot{R}\geq 0 and is bounded above by 1, there exists an asymptotic constant state RR^{\infty} such that

(2.4) R:=limtR(t)andlimtR˙(t)=0.\exists~{}R^{\infty}:=\lim_{t\rightarrow\infty}R(t)\quad\mbox{and}\quad\lim_{t\rightarrow\infty}\dot{R}(t)=0.

(iii) By (2.4)\eqref{B-4}, one has

limtI(t)=1blimtR˙(t)=0.\lim_{t\to\infty}I(t)=\frac{1}{b}\lim_{t\to\infty}\dot{R}(t)=0.

2.2. The swarmalator model

Let xix_{i} and θi\theta_{i} be the position and phase of the ii-th swarmalator, respectively. Then, its dynamics is governed by the Cauchy problem to the Kuramoto-type swarmalator model:

(2.5) {x˙i=wi+1N1jNji[Ψ~a(θjθi)xjxixjxiαΨ~r(θjθi)xjxixjxiβ],t>0,θ˙i=νi+κN1jNjisin(θjθi)xjxiγ,i[N],(xi(0),θi(0))=(xi0,θi0).\begin{cases}\displaystyle{\dot{x}}_{i}=w_{i}+\frac{1}{N}\sum_{\begin{subarray}{c}1\leq j\leq N\\ j\neq i\end{subarray}}\left[\tilde{\Psi}_{a}(\theta_{j}-\theta_{i})\frac{x_{j}-x_{i}}{\|x_{j}-x_{i}\|^{\alpha}}-\tilde{\Psi}_{r}(\theta_{j}-\theta_{i})\frac{x_{j}-x_{i}}{\|x_{j}-x_{i}\|^{\beta}}\right],\quad t>0,\\ \displaystyle{\dot{\theta}}_{i}=\nu_{i}+\frac{\kappa}{N}\sum_{\begin{subarray}{c}1\leq j\leq N\\ j\neq i\end{subarray}}\frac{\sin(\theta_{j}-\theta_{i})}{\|x_{j}-x_{i}\|^{\gamma}},\quad i\in[N],\\ \displaystyle(x_{i}(0),\theta_{i}(0))=(x_{i}^{0},\theta_{i}^{0}).\end{cases}

Here, constants wiw_{i} and νi\nu_{i} are the natural velocity and frequency of the ii-th particle, respectively, \|\cdot\| denotes the standard Euclidean 2\ell^{2}-norm in d\mathbb{R}^{d}. System parameters and interaction functions Ψ~a{\tilde{\Psi}}_{a} and Ψ~r{\tilde{\Psi}}_{r} satisfy

(2.6) 1α<β,Ψ~a(θ)=Ψ~a(θ),Ψ~r(θ)=Ψ~r(θ),θ,0<m~aΨ~a(θ)M~a<,0<m~rΨ~r(θ)M~r<.\displaystyle\begin{aligned} &1\leq\alpha<\beta,\quad{\tilde{\Psi}}_{a}(\theta)={\tilde{\Psi}}_{a}(-\theta),\quad\tilde{\Psi}_{r}(\theta)=\tilde{\Psi}_{r}(-\theta),\quad\theta\in\mathbb{R},\\ &0<\tilde{m}_{a}\leq\tilde{\Psi}_{a}(\theta)\leq\tilde{M}_{a}<\infty,\quad 0<\tilde{m}_{r}\leq\tilde{\Psi}_{r}(\theta)\leq\tilde{M}_{r}<\infty.\end{aligned}

If we simply set γ=0\gamma=0, system (2.5)2\eqref{B-5}_{2} becomes the Kuramoto model [27]:

θ˙i=νi+κNj=1Nsin(θjθi).{\dot{\theta}}_{i}=\nu_{i}+\frac{\kappa}{N}\sum_{j=1}^{N}{\sin(\theta_{j}-\theta_{i})}.

Due to the singular terms in the swarmalator model (2.5)1\eqref{B-5}_{1}, it is important to make sure that there is no finite-time collision which can be quantified in the following proposition.

First, we set several parameters:

(2.7) γ1:=2m~aN,γ2:=2M~rN,γ3:=D(W):=maxi,j|wiwj|,𝒬:=N(N1)2,B~:=m~r2(M~a+M~r)+𝒟(W),D1(t):=min1ijNxi(t)xj(t),D(t):=max1ijNxi(t)xj(t),Λ:=(12)𝒬max{D1(0),1}𝒬((max{(m~r2M~a)1βα,1})𝒬(B~N)Qβ1.\displaystyle\begin{aligned} &\gamma_{1}:=\frac{2\tilde{m}_{a}}{N},\quad\gamma_{2}:=\frac{2\tilde{M}_{r}}{N},\quad\gamma_{3}:=D(W):=\max_{i,j}|w_{i}-w_{j}|,\\ &{\mathcal{Q}}:=\frac{N(N-1)}{2},\quad\tilde{B}:=\frac{\tilde{m}_{r}}{2(\tilde{M}_{a}+\tilde{M}_{r})+{\mathcal{D}}(W)},\\ &D_{1}(t):=\min_{1\leq i\not=j\leq N}\|x_{i}(t)-x_{j}(t)\|,\quad D(t):=\max_{1\leq i\not=j\leq N}\|x_{i}(t)-x_{j}(t)\|,\\ &\Lambda:=\Big{(}\frac{1}{2}\Big{)}^{{\mathcal{Q}}}\max\{D_{1}(0),1\}^{\mathcal{Q}}\Big{(}\Big{(}\max\Big{\{}\Big{(}\frac{\tilde{m}_{r}}{2\tilde{M}_{a}}\Big{)}^{\frac{1}{\beta-\alpha}},1\Big{\}}\Big{)}^{\mathcal{Q}}\Big{(}\frac{\tilde{B}}{N}\Big{)}^{\frac{Q}{\beta-1}}.\end{aligned}

Next, we state two results on the positivity of minimal and maximal distance between particles.

Proposition 2.2.

(Positivity of minimal distance) [20] Suppose system parameters satisfy (LABEL:B-6), and the initial data (X0,Θ0)(X^{0},\Theta^{0}) is non-collisional:

κ>0,min1ijNxi0xj0>0.\kappa>0,\qquad\min_{1\leq i\not=j\leq N}\|x^{0}_{i}-x^{0}_{j}\|>0.

Then, there exists a global solution (X,Θ)(X,\Theta) to (2.5) - (LABEL:B-6) and δ~1>0\tilde{\delta}_{1}>0 such that

inf0t<min1ijNxi(t)xj(t)δ~1.\inf_{0\leq t<\infty}\min_{1\leq i\not=j\leq N}\|x_{i}(t)-x_{j}(t)\|\geq\tilde{\delta}_{1}.
Proposition 2.3.

[20] Suppose system parameters and initial data satisfy the following framework:

  • (1)({\mathcal{F}}1) : For α=1\alpha=1,

    (2.8) Λ>(M~rm~a)βα,γ1>γ3.\Lambda>\left(\frac{\tilde{M}_{r}}{\tilde{m}_{a}}\right)^{\beta-\alpha},\quad\gamma_{1}>\gamma_{3}.
  • (2)({\mathcal{F}}2) : For α>1\alpha>1,

    (2.9) {D(X0)<y,Λ>(M~rm~a)βα,γ1(γ1(α1)γ2(β1))(βα)(α1)+γ2(γ1(α1)γ2(β1))(βα)(β1)+γ30,\begin{cases}\displaystyle D(X^{0})<y^{*},\quad\Lambda>\left(\frac{\tilde{M}_{r}}{\tilde{m}_{a}}\right)^{\beta-\alpha},\\ \displaystyle-\gamma_{1}\left(\frac{\gamma_{1}(\alpha-1)}{\gamma_{2}(\beta-1)}\right)^{(\beta-\alpha)(\alpha-1)}+\gamma_{2}\left(\frac{\gamma_{1}(\alpha-1)}{\gamma_{2}(\beta-1)}\right)^{(\beta-\alpha)(\beta-1)}+\gamma_{3}\leq 0,\end{cases}

    where yy^{*} is the largest root of the equation:

    γ1y1α+γ2y1β+γ3=0,-\gamma_{1}y^{1-\alpha}+\gamma_{2}y^{1-\beta}+\gamma_{3}=0,

and let (X,Θ)(X,\Theta) be a solution to (2.5) - (LABEL:B-6). Then, there exists a positive constant δ~\tilde{\delta}_{\infty} such that

sup0t<D(t)δ~.\sup_{0\leq t<\infty}D(t)\leq\tilde{\delta}_{\infty}.
Proof.

For a proof, we refer to [20]. ∎

Consider the following ansatz for Ψ~a\tilde{\Psi}_{a} and Ψ~r\tilde{\Psi}_{r}:

Ψ~a(θ):=1+Jcosθ,Ψ~r:=1,|J|<1.\tilde{\Psi}_{a}(\theta):=1+J\cos\theta,\quad\tilde{\Psi}_{r}:=1,\quad|J|<1.

In this case, system (2.5) becomes

(2.10) {x˙i=wi+1N1jNji[(1+Jcos(θjθi))xjxixjxiαxjxixjxiβ],t>0,θ˙i=νi+κN1jNjisin(θjθi)xjxiγ.\begin{cases}\displaystyle{\dot{x}}_{i}=w_{i}+\frac{1}{N}\sum_{\begin{subarray}{c}1\leq j\leq N\\ j\neq i\end{subarray}}\left[(1+J\cos(\theta_{j}-\theta_{i}))\frac{x_{j}-x_{i}}{\|x_{j}-x_{i}\|^{\alpha}}-\frac{x_{j}-x_{i}}{\|x_{j}-x_{i}\|^{\beta}}\right],\quad t>0,\\ \displaystyle{\dot{\theta}}_{i}=\nu_{i}+\frac{\kappa}{N}\sum_{\begin{subarray}{c}1\leq j\leq N\\ j\neq i\end{subarray}}\frac{\sin(\theta_{j}-\theta_{i})}{\|x_{j}-x_{i}\|^{\gamma}}.\end{cases}

By Proposition 2.1 and convergence result of the perturbed gradient system, system (2.10) exhibits phase synchronization.

Theorem 2.1.

[20] Suppose system parameters, natural velocities and initial data satisfy one of the frameworks (2.8) and (2.9), and

κ>0,D(ν)=0and0<D(Θ0)<π,\kappa>0,\quad D(\nu)=0\quad\mbox{and}\quad 0<D(\Theta^{0})<\pi,

and let (X,Θ)(X,\Theta) be a solution to (2.10). Then there exists asymptotic state XX^{\infty} such that

limtX(t)=X.\lim_{t\rightarrow\infty}X(t)=X^{\infty}.
Proof.

For a proof, we refer to Section 4.2 of [20]. ∎

2.3. A global well-posedness

In this subsection, we discuss a global well-posedness of system (1.6). Since the R.H.S. of (1.6)2\eqref{A-1}_{2} contains the term xjxi\|x_{j}-x_{i}\| in the denominators, as long as we can rule out the possibility of finite-time collisions, we obtain a global well-posedness using the standard Cauchy-Lipschitz theory. In the sequel, we show that system (1.6) does not admit a finite-time collision, as long as there is no collisions initially. This will be done using a contradiction argument and Gronwall’s inequality.

Suppose that initial spatial configuration satisfy

min1ijNxi0xj0>0.\min_{1\leq i\not=j\leq N}\|x_{i}^{0}-x_{j}^{0}\|>0.

Then, by the continuity of solution, there will be no collisions between particles at least small time interval [0,ε)[0,\varepsilon) with ε1\varepsilon\ll 1. Then, using the Cauchy-Lipschitz theory, we can show that system (1.6) has a local smooth solution {(Wi,xi)}\{(W_{i},x_{i})\} in the time-interval [0,ε)[0,\varepsilon). Now, in order to show a global well-posedness, it suffices to show that there will be no finite-time collisions. Suppose there is a finite-time collision and let t0t_{0} be the first collision time. We take one of the particles that make a collision and fix it by pp. Then, we define a set containing all the particles involved in the collision at time t0t_{0} by 𝒞{\mathcal{C}}. Now, we set the following handy notation:

:={1,2,,N}𝒞,χ𝒞:=i,j𝒞xixj2,ij:=1jNji.{\mathcal{R}}:=\{1,2,\cdots,N\}\setminus{\mathcal{C}},\qquad\chi_{\mathcal{C}}:=\sqrt{\sum_{i,j\in{\mathcal{C}}}\|x_{i}-x_{j}\|^{2}},\qquad\sum_{i\not=j}:=\sum_{\begin{subarray}{c}1\leq j\leq N\\ j\neq i\end{subarray}}.

Then, it is easy to see that

(2.11) χ𝒞(t0)=0,χ𝒞(t)>0,t(0,t0).\chi_{\mathcal{C}}(t_{0})=0,\quad\chi_{\mathcal{C}}(t)>0,\quad t\in(0,t_{0}).

By direct calculations, one has

(2.12) dχ𝒞2dt=2i,j𝒞(xixj)(xi˙xj˙)=2i,j𝒞(xixj)[κ2Nki(Ψaikxkxixkxiα)κ3Nki(Ψrikxkxixkxiβ)]2i,j𝒞(xixj)[κ2Nkj(Ψajkxkxjxkxjα)κ3Nkj(Ψrjkxkxjxkxjβ)]=2i,j𝒞(xixj)[κ2NkC,ki(Ψaikxkxixkxiα)κ2NkC,kj(Ψaikxkxjxkxjα)]+2i,j𝒞(xixj)[κ2NkR(Ψaikxkxi|xkxi|α)κ2NkR(Ψaikxkxjxkxjα)]2i,j𝒞(xixj)[κ3NkC,ki(Ψrikxkxixkxiβ)κ3NkC,kj(Ψrikxkxjxkxjβ)]2i,j𝒞(xixj)[κ3NkR(Ψrikxkxixkxiβ)κ3NkR(Ψrikxkxjxkxjβ)]=:11+12+13+14.\displaystyle\begin{aligned} \frac{d\chi_{\mathcal{C}}^{2}}{dt}&=2\sum_{i,j\in{\mathcal{C}}}{(x_{i}-x_{j})\cdot(\dot{x_{i}}-\dot{x_{j}})}\\ &=2\sum_{i,j\in{\mathcal{C}}}{(x_{i}-x_{j})\cdot\left[\frac{\kappa_{2}}{N}\sum_{k\neq i}{\left(\Psi_{a}^{ik}\cdot\frac{x_{k}-x_{i}}{\|x_{k}-x_{i}\|^{\alpha}}\right)-\frac{\kappa_{3}}{N}\sum_{k\neq i}\left(\Psi_{r}^{ik}\cdot\frac{x_{k}-x_{i}}{\|x_{k}-x_{i}\|^{\beta}}\right)}\right]}\\ &-2\sum_{i,j\in{\mathcal{C}}}{(x_{i}-x_{j})\cdot\left[\frac{\kappa_{2}}{N}\sum_{k\neq j}{\left(\Psi_{a}^{jk}\cdot\frac{x_{k}-x_{j}}{\|x_{k}-x_{j}\|^{\alpha}}\right)-\frac{\kappa_{3}}{N}\sum_{k\neq j}\left(\Psi_{r}^{jk}\cdot\frac{x_{k}-x_{j}}{\|x_{k}-x_{j}\|^{\beta}}\right)}\right]}\\ &=2\sum_{i,j\in{\mathcal{C}}}{(x_{i}-x_{j})\cdot\left[\frac{\kappa_{2}}{N}\sum_{k\in C,k\neq i}{\left(\Psi_{a}^{ik}\cdot\frac{x_{k}-x_{i}}{\|x_{k}-x_{i}\|^{\alpha}}\right)}-\frac{\kappa_{2}}{N}\sum_{k\in C,k\neq j}{\left(\Psi_{a}^{ik}\cdot\frac{x_{k}-x_{j}}{\|x_{k}-x_{j}\|^{\alpha}}\right)}\right]}\\ &+2\sum_{i,j\in{\mathcal{C}}}{(x_{i}-x_{j})\cdot\left[\frac{\kappa_{2}}{N}\sum_{k\in R}{\left(\Psi_{a}^{ik}\cdot\frac{x_{k}-x_{i}}{|x_{k}-x_{i}|^{\alpha}}\right)}-\frac{\kappa_{2}}{N}\sum_{k\in R}{\left(\Psi_{a}^{ik}\cdot\frac{x_{k}-x_{j}}{\|x_{k}-x_{j}\|^{\alpha}}\right)}\right]}\\ &-2\sum_{i,j\in{\mathcal{C}}}{(x_{i}-x_{j})\cdot\left[\frac{\kappa_{3}}{N}\sum_{k\in C,k\neq i}{\left(\Psi_{r}^{ik}\cdot\frac{x_{k}-x_{i}}{\|x_{k}-x_{i}\|^{\beta}}\right)}-\frac{\kappa_{3}}{N}\sum_{k\in C,k\neq j}{\left(\Psi_{r}^{ik}\cdot\frac{x_{k}-x_{j}}{\|x_{k}-x_{j}\|^{\beta}}\right)}\right]}\\ &-2\sum_{i,j\in{\mathcal{C}}}{(x_{i}-x_{j})\cdot\left[\frac{\kappa_{3}}{N}\sum_{k\in R}{\left(\Psi_{r}^{ik}\cdot\frac{x_{k}-x_{i}}{\|x_{k}-x_{i}\|^{\beta}}\right)}-\frac{\kappa_{3}}{N}\sum_{k\in R}{\left(\Psi_{r}^{ik}\cdot\frac{x_{k}-x_{j}}{\|x_{k}-x_{j}\|^{\beta}}\right)}\right]}\\ &=:\mathcal{I}_{11}+\mathcal{I}_{12}+\mathcal{I}_{13}+\mathcal{I}_{14}.\end{aligned}

In the following lemma, we provide estimates for 1i{\mathcal{I}}_{1i}.

Lemma 2.1.

The term 1i{\mathcal{I}}_{1i} with 1i41\leq i\leq 4 satisfies

(i)11+132Ma|𝒞|Ni,k𝒞kixkxi2α+2mr|𝒞|Ni,k𝒞kixkxi2β,(ii)12+144|||𝒞|N(κ2Maδα1+κ3Mrδβ1)χ𝒞,\displaystyle\begin{aligned} &(i)~{}\mathcal{I}_{11}+\mathcal{I}_{13}\geq-\frac{2M_{a}|{\mathcal{C}}|}{N}\sum_{\begin{subarray}{c}i,k\in{\mathcal{C}}\\ k\neq i\end{subarray}}{\|x_{k}-x_{i}\|^{2-\alpha}}+\frac{2m_{r}|{\mathcal{C}}|}{N}\sum_{\begin{subarray}{c}i,k\in{\mathcal{C}}\\ k\neq i\end{subarray}}{\|x_{k}-x_{i}\|^{2-\beta}},\\ &(ii)~{}{\mathcal{I}}_{12}+{\mathcal{I}}_{14}\geq-\frac{4|{\mathcal{R}}|\cdot|{\mathcal{C}}|}{N}\left(\frac{\kappa_{2}M_{a}}{\delta^{\alpha-1}}+\frac{\kappa_{3}M_{r}}{\delta^{\beta-1}}\right)\chi_{\mathcal{C}},\end{aligned}

where |A||A| is the cardinality of the set AA and δ:=infi𝒞,jxixj\displaystyle\delta:=\inf_{\begin{subarray}{c}i\in{\mathcal{C}},\\ j\in{\mathcal{R}}\end{subarray}}\|x_{i}-x_{j}\|.

Proof.

We provide estimate for 1i\mathcal{I}_{1i} (i=1,2,3,4i=1,2,3,4) as follows:

(i) (Estimate of 11+13\mathcal{I}_{11}+\mathcal{I}_{13}): Note that

2i,j𝒞(xixj)[κ2Nk𝒞,ki(Ψaikxkxixkxiα)]\displaystyle 2\sum_{i,j\in{\mathcal{C}}}{(x_{i}-x_{j})\cdot\left[\frac{\kappa_{2}}{N}\sum_{k\in{\mathcal{C}},k\neq i}{\left(\Psi^{ik}_{a}\cdot\frac{x_{k}-x_{i}}{\|x_{k}-x_{i}\|^{\alpha}}\right)}\right]}
=2κ2Ni,j𝒞k𝒞ki(Ψaik(xixj)(xkxi)xkxiα)=2κ2Ni,j,k𝒞ki(Ψaik(xixj)(xkxi)xkxiα)\displaystyle\hskip 28.45274pt=\frac{2\kappa_{2}}{N}\sum_{i,j\in{\mathcal{C}}}\sum_{\begin{subarray}{c}k\in{\mathcal{C}}\\ k\neq i\end{subarray}}{\left(\Psi_{a}^{ik}\cdot\frac{(x_{i}-x_{j})\cdot(x_{k}-x_{i})}{\|x_{k}-x_{i}\|^{\alpha}}\right)}=\frac{2\kappa_{2}}{N}\sum_{\begin{subarray}{c}i,j,k\in{\mathcal{C}}\\ k\neq i\end{subarray}}{\left(\Psi_{a}^{ik}\cdot\frac{(x_{i}-x_{j})\cdot(x_{k}-x_{i})}{\|x_{k}-x_{i}\|^{\alpha}}\right)}
=κ2Ni,j,k𝒞ki(Ψaik(xixj)(xkxi)xkxiα+Ψaik(xkxj)(xixk)xixkα)\displaystyle\hskip 28.45274pt=\frac{\kappa_{2}}{N}\sum_{\begin{subarray}{c}i,j,k\in{\mathcal{C}}\\ k\neq i\end{subarray}}{\left(\Psi_{a}^{ik}\cdot\frac{(x_{i}-x_{j})\cdot(x_{k}-x_{i})}{\|x_{k}-x_{i}\|^{\alpha}}+\Psi_{a}^{ik}\cdot\frac{(x_{k}-x_{j})\cdot(x_{i}-x_{k})}{\|x_{i}-x_{k}\|^{\alpha}}\right)}
=κ2Ni,j,k𝒞ki(Ψaikxkxiα2)=|𝒞|κ2Ni,k𝒞ki(Ψaikxkxiα2).\displaystyle\hskip 28.45274pt=-\frac{\kappa_{2}}{N}\sum_{\begin{subarray}{c}i,j,k\in{\mathcal{C}}\\ k\neq i\end{subarray}}{\left(\frac{\Psi_{a}^{ik}}{\|x_{k}-x_{i}\|^{\alpha-2}}\right)}=-\frac{|{\mathcal{C}}|\kappa_{2}}{N}\sum_{\begin{subarray}{c}i,k\in{\mathcal{C}}\\ k\neq i\end{subarray}}{\left(\frac{\Psi_{a}^{ik}}{\|x_{k}-x_{i}\|^{\alpha-2}}\right)}.

Therefore, one has

(2.13) 11=2|𝒞|κ2Ni,k𝒞ki(Ψaikxkxiα2).\mathcal{I}_{11}=-\frac{2|{\mathcal{C}}|\kappa_{2}}{N}\sum_{\begin{subarray}{c}i,k\in{\mathcal{C}}\\ k\neq i\end{subarray}}{\left(\frac{\Psi_{a}^{ik}}{\|x_{k}-x_{i}\|^{\alpha-2}}\right)}.

Similarly, one has

(2.14) 13=2|𝒞|κ3Ni,k𝒞ki(Ψrikxkxiβ2).\mathcal{I}_{13}=\frac{2|{\mathcal{C}}|\kappa_{3}}{N}\sum_{\begin{subarray}{c}i,k\in{\mathcal{C}}\\ k\neq i\end{subarray}}{\left(\frac{\Psi_{r}^{ik}}{\|x_{k}-x_{i}\|^{\beta-2}}\right)}.

We combine (2.13) and (2.14) to obtain

11+13\displaystyle\mathcal{I}_{11}+\mathcal{I}_{13} =2|𝒞|κ2Ni,k𝒞ki(Ψaikxkxiα2)+2|𝒞|κ3Ni,k𝒞ki(Ψrikxkxiβ2)\displaystyle=-\frac{2|{\mathcal{C}}|\kappa_{2}}{N}\sum_{\begin{subarray}{c}i,k\in{\mathcal{C}}\\ k\neq i\end{subarray}}{\left(\frac{\Psi_{a}^{ik}}{\|x_{k}-x_{i}\|^{\alpha-2}}\right)}+\frac{2|{\mathcal{C}}|\kappa_{3}}{N}\sum_{\begin{subarray}{c}i,k\in{\mathcal{C}}\\ k\neq i\end{subarray}}{\left(\frac{\Psi_{r}^{ik}}{\|x_{k}-x_{i}\|^{\beta-2}}\right)}
2Ma|𝒞|Ni,k𝒞kixkxi2α+2mr|𝒞|Ni,k𝒞kixkxi2β.\displaystyle\geq-\frac{2M_{a}|{\mathcal{C}}|}{N}\sum_{\begin{subarray}{c}i,k\in{\mathcal{C}}\\ k\neq i\end{subarray}}{\|x_{k}-x_{i}\|^{2-\alpha}}+\frac{2m_{r}|{\mathcal{C}}|}{N}\sum_{\begin{subarray}{c}i,k\in{\mathcal{C}}\\ k\neq i\end{subarray}}{\|x_{k}-x_{i}\|^{2-\beta}}.

(ii) (Estimate of 12+14\mathcal{I}_{12}+\mathcal{I}_{14}): Since |xkxi|δ|x_{k}-x_{i}|\geq\delta, one has

12\displaystyle{\mathcal{I}}_{12} =2i,j𝒞(xixj)[κ2NkR(Ψaikxkxixkxiα)κ2Nk(Ψaikxkxjxkxjα)]\displaystyle=2\sum_{i,j\in{\mathcal{C}}}{(x_{i}-x_{j})\cdot\left[\frac{\kappa_{2}}{N}\sum_{k\in R}{\left(\Psi_{a}^{ik}\cdot\frac{x_{k}-x_{i}}{\|x_{k}-x_{i}\|^{\alpha}}\right)}-\frac{\kappa_{2}}{N}\sum_{k\in{\mathcal{R}}}{\left(\Psi_{a}^{ik}\cdot\frac{x_{k}-x_{j}}{\|x_{k}-x_{j}\|^{\alpha}}\right)}\right]}
4κ2Ni,j𝒞kMaxixjδα1=4κ2||MaNδα1i,j𝒞xixj4κ2|||𝒞|MaNδα1χ𝒞,\displaystyle\geq-\frac{4\kappa_{2}}{N}\sum_{\begin{subarray}{c}i,j\in{\mathcal{C}}\\ k\in{\mathcal{R}}\end{subarray}}{M_{a}\frac{\|x_{i}-x_{j}\|}{\delta^{\alpha-1}}}=-\frac{4\kappa_{2}|{\mathcal{R}}|M_{a}}{N\delta^{\alpha-1}}\sum_{i,j\in{\mathcal{C}}}\|x_{i}-x_{j}\|\geq-\frac{4\kappa_{2}|{\mathcal{R}}|\cdot|{\mathcal{C}}|M_{a}}{N\delta^{\alpha-1}}\chi_{\mathcal{C}},

where we used the Cauchy-Schwarz inequality to see

i,j𝒞xixj|𝒞|χ𝒞.\sum_{i,j\in{\mathcal{C}}}\|x_{i}-x_{j}\|\leq|{\mathcal{C}}|\chi_{\mathcal{C}}.

Similarly, one has

12+144|||𝒞|N(κ2Maδα1+κ3Mrδβ1)χ𝒞.\displaystyle{\mathcal{I}}_{12}+{\mathcal{I}}_{14}\geq-\frac{4|{\mathcal{R}}|\cdot|{\mathcal{C}}|}{N}\left(\frac{\kappa_{2}M_{a}}{\delta^{\alpha-1}}+\frac{\kappa_{3}M_{r}}{\delta^{\beta-1}}\right)\chi_{\mathcal{C}}.

It follows from (2.12) and Lemma 2.1 that

(2.15) dχC2dt2Ma|𝒞|Ni,k𝒞kixkxi2α+2mr|𝒞|Ni,k𝒞kixkxi2β4|||𝒞|N(κ2Maδα1+κ3Mrδβ1)χ𝒞=c1i,kCkixkxi2α+c2i,kCkixkxi2βc3χC,\displaystyle\begin{aligned} \frac{d\chi_{C}^{2}}{dt}&\geq-\frac{2M_{a}|{\mathcal{C}}|}{N}\sum_{\begin{subarray}{c}i,k\in{\mathcal{C}}\\ k\neq i\end{subarray}}{\|x_{k}-x_{i}\|^{2-\alpha}}+\frac{2m_{r}|{\mathcal{C}}|}{N}\sum_{\begin{subarray}{c}i,k\in{\mathcal{C}}\\ k\neq i\end{subarray}}{\|x_{k}-x_{i}\|^{2-\beta}}\\ &\hskip 14.22636pt-\frac{4|{\mathcal{R}}|\cdot|{\mathcal{C}}|}{N}\left(\frac{\kappa_{2}M_{a}}{\delta^{\alpha-1}}+\frac{\kappa_{3}M_{r}}{\delta^{\beta-1}}\right)\chi_{\mathcal{C}}\\ &=-c_{1}\sum_{\begin{subarray}{c}i,k\in C\\ k\neq i\end{subarray}}{\|x_{k}-x_{i}\|^{2-\alpha}}+c_{2}\sum_{\begin{subarray}{c}i,k\in C\\ k\neq i\end{subarray}}{\|x_{k}-x_{i}\|^{2-\beta}}-c_{3}\chi_{C},\end{aligned}

where c1,c2,and c3c_{1},c_{2},\text{and }c_{3} are positive constants defined as follows.

c1:=2Ma|𝒞|N,c2:=2mr|𝒞|N,c3:=4|||𝒞|N(κ2Maδα1+κ3Mrδβ1).c_{1}:=\frac{2M_{a}|{\mathcal{C}}|}{N},\quad c_{2}:=\frac{2m_{r}|{\mathcal{C}}|}{N},\quad c_{3}:=\frac{4|{\mathcal{R}}|\cdot|{\mathcal{C}}|}{N}\left(\frac{\kappa_{2}M_{a}}{\delta^{\alpha-1}}+\frac{\kappa_{3}M_{r}}{\delta^{\beta-1}}\right).

On the other hand, since 1α<β1\leq\alpha<\beta in (1.2), there exists a small positive constant ε1\varepsilon\ll 1 such that for t(t0ε,t0)t\in(t_{0}-\varepsilon,t_{0}),

(2.16) c1i,kCkixkxi2α<c24i,kCkixkxi2β,c3χC<c24i,kCkixkxi2β.c_{1}\sum_{\begin{subarray}{c}i,k\in C\\ k\neq i\end{subarray}}{\|x_{k}-x_{i}\|^{2-\alpha}}<\frac{c_{2}}{4}\sum_{\begin{subarray}{c}i,k\in C\\ k\neq i\end{subarray}}{\|x_{k}-x_{i}\|^{2-\beta}},\quad c_{3}\chi_{C}<\frac{c_{2}}{4}\sum_{\begin{subarray}{c}i,k\in C\\ k\neq i\end{subarray}}{\|x_{k}-x_{i}\|^{2-\beta}}.

Therefore, we combine (2.15) and (2.16) to find

dχC2dt>c22i,kCkixkxi2β,t(t0ε,t0).\frac{d\chi_{C}^{2}}{dt}>\frac{c_{2}}{2}\sum_{\begin{subarray}{c}i,k\in C\\ k\neq i\end{subarray}}{\|x_{k}-x_{i}\|^{2-\beta}},\quad t\in(t_{0}-\varepsilon,t_{0}).

Moreover, there exists a positive constant ε(0,ε)\varepsilon^{*}\in(0,\varepsilon) such that for t(t0ε,t0)t\in(t_{0}-\varepsilon^{*},t_{0}),

xixj2βxixj2,for t(t0ε,t0).\|x_{i}-x_{j}\|^{2-\beta}\geq\|x_{i}-x_{j}\|^{2},\>\text{for }t\in(t_{0}-\varepsilon^{*},t_{0}).

Thus, we have

ddtχC2>c22i,kCkixkxi2βc22χC2, for t(t0ε,t0).\frac{d}{dt}\chi_{C}^{2}>\frac{c_{2}}{2}\sum_{\begin{subarray}{c}i,k\in C\\ k\neq i\end{subarray}}{\|x_{k}-x_{i}\|^{2-\beta}}\geq\frac{c_{2}}{2}\chi_{C}^{2},\quad\text{ for }t\in(t_{0}-\varepsilon^{*},t_{0}).

i.e.,

ddtχC2>c22χC2, for t(t0ε,t0).\frac{d}{dt}\chi_{C}^{2}>\frac{c_{2}}{2}\chi_{C}^{2},\quad\text{ for }t\in(t_{0}-\varepsilon^{*},t_{0}).

We integrate the above differential inequality from t0εt_{0}-\varepsilon^{*} to t0t_{0} to get

ec22t0χC2(t0)ec22(t0ε)χC2(t0ε).e^{-\frac{c_{2}}{2}t_{0}}\chi_{C}^{2}(t_{0})\geq e^{-\frac{c_{2}}{2}(t_{0}-\varepsilon^{*})}\chi_{C}^{2}(t_{0}-\varepsilon^{*}).

On the other hand, it follows from (2.11) that the left-hand side is zero, but the right-hand side is strictly positive, which is contradictory. Therefore, there are no finite-time collisions between particles from the well-prepared initial configuration. Finally, we can summarize the previous argument as follows.

Theorem 2.2.

(A global well-posedness) Suppose that initial spatial configuration is noncollisional in the sense that

min1ijNxi0xj0>0.\min_{1\leq i\not=j\leq N}\|x_{i}^{0}-x_{j}^{0}\|>0.

Then, there exists a unique global solution {(Wi,xi)}\{(W_{i},x_{i})\} to system (1.6) in any finite-time interval.

3. Modeling of system functions

In this section, we discuss meaning of the system parameters and system functions appearing in the SIR-flock (1.6):

(aij):interaction topology,(bi):recovering vector,Ψa,Ψr:coupling weight functions.(a^{ij}):~{}\mbox{interaction topology},\quad(b^{i}):~{}\mbox{recovering vector},\quad\Psi_{a},~{}~{}\Psi_{r}:~{}\mbox{coupling weight functions}.

3.1. Interaction topology and recovering vector

For the modeling purpose, we assume that the nonnegative value aij=aij(xi(t)xj(t))a^{ij}=a^{ij}(\|x_{i}(t)-x_{j}(t)\|) depends on the relative distance xixj\|x_{i}-x_{j}\| between the ii-th particle and the jj-th particle. To be definiteness, we set

(3.1) aij={1(xixj+L)γifij,0otherwise,\displaystyle a^{ij}=\begin{cases}\displaystyle\frac{1}{(\|x_{i}-x_{j}\|+L)^{\gamma}}\quad&\text{if}\quad i\neq j,\\ 0&\text{otherwise},\end{cases}

where LL is a positive constant and γ\gamma is a nonnegative constant. Since the ii-th particle can not be affected by itself, we assume the diagonal entries of (aij)(a^{ij}) are zero:

aii=0,i[N].a^{ii}=0,\quad\forall~{}i\in[N].

Now we discuss the natural recovering vector b=(b1,,bN)b=(b^{1},\cdots,b^{N}). Suppose that every particles have own immune system, the disease will disappear automatically. Thus, we assume

bi>0,i[N].b^{i}>0,\quad\forall~{}i\in[N].

If the immune system of the ii-th particle is well-functioning, then bib^{i} has a large value. In contrast, if the immune system of the ii-th particle is not well-functioning, then bimpimwwwib^{i}mpimwwwi will have a small value. In this work, we assume that the natural recovering vector bib_{i} is a positive constant (see the following figure in the sequel).

SiS_{i}IiI_{i}RiR_{i}SjS_{j}IjI_{j}RjR_{j}aijSiIja^{ij}S_{i}I_{j}biIib^{i}I_{i}ajiSjIia^{ji}S_{j}I_{i}bjIjb^{j}I_{j}

3.2. Coupling weight functions

Recall that the dynamics of xix_{i} is governed by the following system:

(3.2) x˙i=κ2Nji(Ψaijxjxixjxiα)κ3Nji(Ψrijxjxixjxiβ).\dot{x}_{i}=\frac{\kappa_{2}}{N}\sum_{j\neq i}\left(\Psi_{a}^{ij}\frac{x_{j}-x_{i}}{\|x_{j}-x_{i}\|^{\alpha}}\right)-\frac{\kappa_{3}}{N}\sum_{j\neq i}\left(\Psi_{r}^{ij}\frac{x_{j}-x_{i}}{\|x_{j}-x_{i}\|^{\beta}}\right).

Now, the matter of question is how to model Ψa\Psi_{a} and Ψr\Psi_{r} in terms of WiW_{i}^{\prime}s.

If two state vectors WiW_{i} and WjW_{j} are similar, then particles xix_{i} and xjx_{j} will attract each others, whereas if two state vectors WiW_{i} and WjW_{j} are dissimilar, then xix_{i} and xjx_{j} will repel each other. For definiteness, we set

(3.3) {Ψaij=Ψa(Wi,Wj)=(Si+Ri)(Sj+Rj)+IiIj+εa,i,j[N],Ψrij=Ψr(Wi,Wj)=(Si+Ri)Ij+(Sj+Rj)Ii+εr,i,j[N],\displaystyle\begin{cases}\Psi_{a}^{ij}=\Psi_{a}(W_{i},W_{j})=(S_{i}+R_{i})(S_{j}+R_{j})+I_{i}I_{j}+\varepsilon_{a},\quad i,j\in[N],\\ \Psi_{r}^{ij}=\Psi_{r}(W_{i},W_{j})=(S_{i}+R_{i})I_{j}+(S_{j}+R_{j})I_{i}+\varepsilon_{r},\quad i,j\in[N],\end{cases}

where α\alpha and β\beta are positive constants. Here εa\varepsilon_{a} and εr\varepsilon_{r} are positive constants presenting social distancing. These social distancing constants play an important role in preventing collisions between particles. Note that the ansatz (3.3) satisfies symmetry:

(3.4) Ψaij=Ψaji,Ψrij=Ψrji,i,j[N].\Psi_{a}^{ij}=\Psi_{a}^{ji},\quad\Psi_{r}^{ij}=\Psi_{r}^{ji},\quad i,j\in[N].

Next, we will impose conditions on α\alpha and β\beta later in (3.2). To illustrate the functional relations in the right-hand side of (3.3), we consider an ensemble which is partitioned into two sub-ensembles:

{Infected particles}or{susceptible or recovered particles}.\{\mbox{Infected particles}\}\quad\mbox{or}\quad\{\mbox{susceptible or recovered particles}\}.

Recall that IiI_{i} and Si+RiS_{i}+R_{i} are probabilities that ii-th particle are in infected state and are not in infected state, respectively.

Consider the inner product-like function GG as follows:

(3.5) G(Wi,Wj)=(Si+Ri,Ii)(Sj+Rj,Ij)=(Si+Ri)(Sj+Rj)+IiIj=(1Ii)(1Ij)+IiIj.\displaystyle\begin{aligned} G(W_{i},W_{j})&=(S_{i}+R_{i},I_{i})\cdot(S_{j}+R_{j},I_{j})=(S_{i}+R_{i})(S_{j}+R_{j})+I_{i}I_{j}\\ &=(1-I_{i})(1-I_{j})+I_{i}I_{j}.\end{aligned}

Note that G(Wi,Wj)G(W_{i},W_{j}) becomes larger when IiIjI_{i}\simeq I_{j} and Si+RiSj+RjS_{i}+R_{i}\simeq S_{j}+R_{j}. If WiW_{i} and WjW_{j} are similar, then G(Wi,Wj)G(W_{i},W_{j}) becomes larger, and one has

(3.6) 0G(Wi,Wj)=(1Ii)(1Ij)+IiIj[(1Ii)+Ii][(1Ij)+Ii]=1.0\leq G(W_{i},W_{j})=(1-I_{i})(1-I_{j})+I_{i}I_{j}\leq\big{[}(1-I_{i})+I_{i}\big{]}\cdot\big{[}(1-I_{j})+I_{i}\big{]}=1.

Now, we set

(3.7) Ψaij=εa+G(Wi,Wj)0,Ψrij=1+εrG(Wi,Wj)0.\Psi_{a}^{ij}=\varepsilon_{a}+G(W_{i},W_{j})\geq 0,\quad\Psi_{r}^{ij}=1+\varepsilon_{r}-G(W_{i},W_{j})\geq 0.

Then, Ψa\Psi_{a} and Ψr\Psi_{r} satisfy the following monotonicity properties:

  1. (1)

    If WiW_{i} and WjW_{j} become similar, Ψa\Psi_{a} increases and Ψr\Psi_{r} decreases.

  2. (2)

    If WiW_{i} and WjW_{j} become dissimilar, Ψa\Psi_{a} decreases and Ψr\Psi_{r} increases.

The relations (3.5) and (3.7) yield (3.3). On the other hand, it follows from (3.3) and (3.6) that the coupling weight functions Ψa\Psi_{a} and Ψr\Psi_{r} admit positive lower bound and upper bounds:

(3.8) 0<εaΨa1+εa=:Ma,0<εrΨr1+εr=:Mr.0<\varepsilon_{a}\leq\Psi_{a}\leq 1+\varepsilon_{a}=:M_{a},\qquad 0<\varepsilon_{r}\leq\Psi_{r}\leq 1+\varepsilon_{r}=:M_{r}.
Proposition 3.1.

Suppose that system parameters satisfy

(3.9) Ψ=0,κ1=aN1,bi=b,i[N],\Psi=0,\quad\kappa_{1}=\frac{a}{N-1},\quad b^{i}=b,\quad i\in[N],

and let (S,I,R)(S,I,R) and (Si,Ii,Ri)(S_{i},I_{i},R_{i}) be the solutions of system (2.1) and system (1.6) with the initial data:

S(0)=S0,I(0)=I0,R(0)=R0,Si0=S0,Ii0=I0,Ri0=R0,i[N].\displaystyle\begin{aligned} &S(0)=S^{0},\quad I(0)=I^{0},\quad R(0)=R^{0},\\ &S_{i}^{0}=S^{0},\quad I_{i}^{0}=I^{0},\quad R_{i}^{0}=R^{0},\quad i\in[N].\end{aligned}

Then we have

(3.10) Si(t)=S(t),Ii(t)=I(t),Ri(t)=R(t),t0,i[N].S_{i}(t)=S(t),\quad I_{i}(t)=I(t),\quad R_{i}(t)=R(t),\quad t\geq 0,~{}~{}i\in[N].
Proof.

Suppose the conditions (3.9) hold. Then, the relation (3.1) and (2.1) become

aij={1ifij,0otherwise,and{S˙i=aN11jNjiSiIj,t>0,I˙i=aN11jNjiSiIjbIi,R˙i=bIi,i[N].a^{ij}=\begin{cases}1\quad&\text{if}\quad i\neq j,\\ 0&\text{otherwise},\end{cases}\quad\mbox{and}\quad\begin{cases}\displaystyle\dot{S}_{i}=-\frac{a}{N-1}\sum_{\begin{subarray}{c}1\leq j\leq N\\ j\neq i\end{subarray}}S_{i}I_{j},\quad t>0,\\ \displaystyle\dot{I}_{i}=\frac{a}{N-1}\sum_{\begin{subarray}{c}1\leq j\leq N\\ j\neq i\end{subarray}}S_{i}I_{j}-bI_{i},\\ \displaystyle\dot{R}_{i}=bI_{i},\quad i\in[N].\end{cases}

Since (Si,Ii,Ri)(S_{i},I_{i},R_{i}) and (S,I,R)(S,I,R) satisfy the same system (2.1) with the same initial data, by the uniqueness of ODE solution, we have the desired uniqueness (3.10). ∎

3.3. Symptom expression vector

Next, we introduce the symptom expression vector s=(s1,,sN)s=(s_{1},\cdots,s_{N}) motivated by the ongoing pandemic COVID-19. In April 2020, the daily COVID-19 confirmed number per day of Korea was less than 20. Most of confirmed people were isolated, however patients with no symptoms of COVID-19 were still spreading the virus. That is why the daily confirmed number per day does not goes to 0 directly. So we need to consider the ratio of symptom expressions. Some people show symptoms well when they got the virus, in contrast some people does not show any symptoms although they got the virus. So even for the same inputs, different degree of output can emerge. Thus, we define the symptom expression vector that are related to previous phenomena. We define define each component si[0,1]s_{i}\in[0,1] to following the following properties:

  • If the ii-th particle shows the symptom well, then we put large value to sis_{i}.

  • If the ii-th particle does not show the symptom well, then we put small value to sis_{i}.

We may use the symptom expression vector ss to express the condition of being suspected as a patient. In this paper, we assume that the symptom expression vector ss is a time-independent constant vector. We assume that the condition of being suspected as a patient only depends on the product siIis_{i}I_{i}, and we also define some fixed threshold constant cc to make decision. If siIi(t)cs_{i}I_{i}(t)\geq c, then the ithi^{th} person will be confirmed at time tt.

4. Relaxation of epidemic states

In this section, we study the relaxation dynamics of the SIR-flock model (1.6). First, we show that 𝒮\mathcal{S} is invariant along (1.6).

Lemma 4.1.

Let {(Wi,xi)}\{(W_{i},x_{i})\} be a solution of system (1.6). Then, one has

(4.1) Wi(t)𝒮,ddt(i=1Nxi)=0,t>0,i[N],W_{i}(t)\in\mathcal{S},\qquad\frac{d}{dt}\Big{(}\sum_{i=1}^{N}{x_{i}}\Big{)}=0,\quad t>0,\quad i\in[N],

where 𝒮{\mathcal{S}} is the set defined in (1.4).

Proof.

We basically use the same arguments as in Proposition 2.1.

(i) (Verification of the second relation in (4.1)): It follows from (1.6)1\eqref{A-1}_{1} that

ddt(Si+Ii+Ri)=0,t>0,i[N].\frac{d}{dt}(S_{i}+I_{i}+R_{i})=0,\quad t>0,\quad i\in[N].

This yields

(4.2) Si(t)+Ii(t)+Ri(t)=1,t0.S_{i}(t)+I_{i}(t)+R_{i}(t)=1,\quad t\geq 0.

Now, we need to show the positivity of Si,IiS_{i},I_{i} and RiR_{i}.

\bullet (Positivity of SiS_{i}): Due to (4.2), it suffices to check the positivity of Si,IiS_{i},I_{i} and RiR_{i}. Now, we integrate (1.6)1\eqref{A-1}_{1} to find

Si(t)=Si0exp(κ10tj=1Naij(τ)Ij(τ)dτ)0,fort0.S_{i}(t)=S_{i}^{0}\exp\left(-\kappa_{1}\int_{0}^{t}\sum_{j=1}^{N}a^{ij}(\tau)I_{j}(\tau)d\tau\right)\geq 0,\quad\mbox{for}~{}t\geq 0.

Thus, one has

Si(t)0,fort0,i[N].S_{i}(t)\geq 0,\qquad\mbox{for}~{}t\geq 0,\quad i\in[N].

\bullet (Positivity of IiI_{i}): It follows from (2.1)2 that

I˙i+biIi=κ1j=1NaijSiIj.\dot{I}_{i}+b^{i}I_{i}=\kappa_{1}\sum_{j=1}^{N}a^{ij}S_{i}I_{j}.

This yields

ddt(Iiebit)=κ1j=1NaijSiIjebit.\frac{d}{dt}(I_{i}e^{b^{i}t})=\kappa_{1}\sum_{j=1}^{N}a^{ij}S_{i}I_{j}e^{b^{i}t}.

Now, we introduce a time-varying minimal index mt[N]m_{t}\in[N] such that

Imt(t):=min1iNIi(t).I_{m_{t}}(t):=\min_{1\leq i\leq N}I_{i}(t).

Since each Ii(t)I_{i}(t) is analytic, there exists the refinement of time-interval

0=t0<t1<t2<0=t_{0}<t_{1}<t_{2}<\cdots

such that the index of ImtI_{m_{t}} is not changed on each interval [tj,tj+1)[t_{j},t_{j+1}), and

ddt(Imtebmtt)=κ1j=1NamtjSmtIjebmtt,t>0.\frac{d}{dt}(I_{m_{t}}e^{b^{m_{t}}t})=\kappa_{1}\sum_{j=1}^{N}a^{m_{t}j}S_{m_{t}}I_{j}e^{b^{m_{t}}t},\quad t>0.

Therefore, we have

ddt(Imtebmtt)=κ1j=1NamtjSmtIjebmttκ1j=1NamtjSmtImtebmtt=Imtebmtt(κ1j=1NamtjSmt).\displaystyle\frac{d}{dt}(I_{m_{t}}e^{b^{m_{t}}t})=\kappa_{1}\sum_{j=1}^{N}a^{m_{t}j}S_{m_{t}}I_{j}e^{b^{m_{t}}t}\geq\kappa_{1}\sum_{j=1}^{N}a^{m_{t}j}S_{m_{t}}I_{m_{t}}e^{b^{m_{t}}t}=I_{m_{t}}e^{b^{m_{t}}t}\Big{(}\kappa_{1}\sum_{j=1}^{N}a^{m_{t}j}S_{m_{t}}\Big{)}.

This yields

ImtebmttIm0exp(κ1j=1N0tamτj(τ)Smτ𝑑τ)0.\displaystyle I_{m_{t}}e^{b^{m_{t}}t}\geq I_{m_{0}}\exp\left(\kappa_{1}\sum_{j=1}^{N}\int_{0}^{t}\ a^{m_{\tau}j}(\tau)S_{m_{\tau}}d\tau\right)\geq 0.

Therefore, one has

Ii(t)Imt(t)0,i[N].I_{i}(t)\geq I_{m_{t}}(t)\geq 0,\quad\forall~{}i\in[N].

\bullet (Positivity of RiR_{i}): Since Si+Ii+Ri=1,Si0S_{i}+I_{i}+R_{i}=1,~{}S_{i}\geq 0 and Ii0I_{i}\geq 0, one has

Ri0.R_{i}\geq 0.

Therefore, we combine all the estimates for Si,IiS_{i},I_{i} and RiR_{i} to get

Wi𝒮.W_{i}\in{\mathcal{S}}.

(ii) (Verification of the second relation in (4.1)): Recall the equation for xix_{i}:

x˙i=κ2Nji(Ψaijxjxixjxiα)κ3Nji(Ψrijxjxixjxiβ).\dot{x}_{i}=\displaystyle\frac{\kappa_{2}}{N}\sum_{j\neq i}\left(\Psi_{a}^{ij}\frac{x_{j}-x_{i}}{\|x_{j}-x_{i}\|^{\alpha}}\right)-\frac{\kappa_{3}}{N}\sum_{j\neq i}\left(\Psi_{r}^{ij}\frac{x_{j}-x_{i}}{\|x_{j}-x_{i}\|^{\beta}}\right).

Since the right-hand side of the above equation is skew-symmetry with respect to interchange map iji~{}\longleftrightarrow~{}j using (3.4), the total sum ixi\sum_{i}x_{i} is time-invariant. ∎

Now, we discuss the asymptotic convergence of the probability vector WiW_{i} toward a constant probability vector W𝒮W^{\infty}\in{\mathcal{S}} in (1.4).

Theorem 4.1.

Let {(Wi,xi)}\{(W_{i},x_{i})\} be a solution of system (1.6). Then, the following assertions hold.

  1. (1)

    There exists a constant state {(Si,0,Ri)}\{(S_{i}^{\infty},0,R_{i}^{\infty})\} with 0Si,Ri10\leq S^{\infty}_{i},~{}R^{\infty}_{i}\leq 1 such that

    limt(Si(t),Ii(t),Ri(t))=(Si,0,Ri),i[N].\lim_{t\rightarrow\infty}(S_{i}(t),I_{i}(t),R_{i}(t))=(S^{\infty}_{i},0,R^{\infty}_{i}),\quad i\in[N].
  2. (2)

    If the recovering value bib_{i} satisfies

    (4.3) min1iNbi>κ1(N1)Lγ,\min_{1\leq i\leq N}b^{i}>\frac{\kappa_{1}(N-1)}{L^{\gamma}},

    then there exists a positive constant λ:=min1iNbiκ1(N1)Lγ>0\displaystyle\lambda:=\min_{1\leq i\leq N}b^{i}-\frac{\kappa_{1}(N-1)}{L^{\gamma}}>0 such that

    |1Ni=1NIi(t)|eλt.\Big{|}\frac{1}{N}\sum_{i=1}^{N}{I_{i}(t)}\Big{|}\leq e^{-\lambda t}.
Proof.

(i) Since IjI_{j} and SiS_{i} are non-negative,

S˙i=κ1(j=1NaijIj)Si0,\dot{S}_{i}=-\kappa_{1}\left(\sum_{j=1}^{N}a^{ij}I_{j}\right)S_{i}\leq 0,

which means SiS_{i} is non-increasing. Since SiS_{i} is bounded below by zero, there are some constants SiS^{\infty}_{i} that Si(t)S_{i}(t) converges to SiS_{i}^{\infty}, as tt goes \infty by monotone convergence theorem. Similarly,

R˙i=biIi0,\dot{R}_{i}=b^{i}I_{i}\geq 0,

and RiR_{i} is non-decreasing and bounded above, there are some constants RiR^{\infty}_{i} that Ri(t)R_{i}(t) converges to RiR_{i}^{\infty}, as tt goes \infty. Moreover, as RiR_{i} converges,

0=limtR˙i(t)=limtbiIi(t),thuslimtIi(t)=0.\displaystyle 0=\lim_{t\rightarrow\infty}\dot{R}_{i}(t)=\lim_{t\rightarrow\infty}b^{i}I_{i}(t),\quad\text{thus}\quad\lim_{t\rightarrow\infty}I_{i}(t)=0.

(ii) Recall the equation for IiI_{i}:

I˙i+biIi=κ1j=1NaijSiIj.\dot{I}_{i}+b^{i}I_{i}=\kappa_{1}\sum_{j=1}^{N}a^{ij}S_{i}I_{j}.

Now, we use the method of integrating factor to derive Gronwall’s inequality for i=1NebitIi\sum_{i=1}^{N}{e^{b^{i}t}I_{i}}:

ddt(i=1NebitIi)=κ1ebiti,j𝒩aijSiIjebiti,j𝒩jiκ1LγIj=κ1(N1)Lγ(i=1NebitIi).\frac{d}{dt}\left(\sum_{i=1}^{N}{e^{b^{i}t}I_{i}}\right)=\kappa_{1}e^{b^{i}t}\sum_{i,j\in\mathcal{N}}{a^{ij}S_{i}I_{j}}\leq e^{b^{i}t}\sum_{\begin{subarray}{c}i,j\in\mathcal{N}\\ j\neq i\end{subarray}}{\frac{\kappa_{1}}{L^{\gamma}}I_{j}}=\frac{\kappa_{1}(N-1)}{L^{\gamma}}\left(\sum_{i=1}^{N}{e^{b^{i}t}I_{i}}\right).

Thus, we have

(i=1NebitIi)(i=1NIi(0))eκ1(N1)LγtNeκ1(N1)Lγt.\displaystyle\left(\sum_{i=1}^{N}{e^{b^{i}t}I_{i}}\right)\leq\left(\sum_{i=1}^{N}{I_{i}(0)}\right)e^{\frac{\kappa_{1}(N-1)}{L^{\gamma}}t}\leq Ne^{\frac{\kappa_{1}(N-1)}{L^{\gamma}}t}.

This implies

Iii=1NIiNe(κ1(N1)Lγbi)tNe(κ1(N1)Lγmin1iNbi)t=:Neλt.\displaystyle I_{i}\leq\sum_{i=1}^{N}{I_{i}}\leq Ne^{\left(\frac{\kappa_{1}(N-1)}{L^{\gamma}}-b^{i}\right)t}\leq Ne^{\left(\frac{\kappa_{1}(N-1)}{L^{\gamma}}-\min_{1\leq i\leq N}b^{i}\right)t}=:Ne^{-\lambda t}.

Next, under a more relaxed condition compared to (4.3), we improve the second result of Theorem 4.1 as follows.

Corollary 4.1.

Let (Wi,xi)(W_{i},x_{i}) be the solution of the system (1.6). If {bi}\{b^{i}\} satisfies following condition:

(4.4) min1iNbi>κ1Lγ(max1iNjiSj0),\min_{1\leq i\leq N}b^{i}>\frac{\kappa_{1}}{L^{\gamma}}\Big{(}\max_{1\leq i\leq N}\sum_{j\neq i}S_{j}^{0}\Big{)},

then i=1NIi\sum_{i=1}^{N}{I_{i}} decays to zero exponentially fast.

Proof.

First, we write the dynamics (1.6)1\eqref{A-1}_{1} of IiI_{i}s in matrix form:

𝐈˙\displaystyle\dot{\mathbf{I}}^{\top} =κ1diag(S1,S2,,SN)𝐀𝐈b𝐈,\displaystyle=\kappa_{1}\text{diag}(S_{1},S_{2},\cdots,S_{N})\mathbf{A}\mathbf{I}^{\top}-b\mathbf{I}^{\top},

where 𝐈\mathbf{I} and 𝐀\mathbf{A} are given as follows.

𝐈(t):=(I1(t),,IN(t))and𝐀(t):=(aij(t))1i,jN.\mathbf{I}(t):=(I_{1}(t),\cdots,I_{N}(t))\quad\mbox{and}\quad\mathbf{A}(t):=(a^{ij}(t))_{1\leq i,j\leq N}.

Since each of SiS_{i} is decreasing,

Si(t)Si(0)andaijLγfor ij.S_{i}(t)\leq S_{i}(0)\quad\mbox{and}\quad a^{ij}\leq L^{-\gamma}\quad\mbox{for $i\neq j$}.

This implies that for every tt,

diag(S1(t),S2(t),,SN(t))𝐀(t)𝐀¯,\text{diag}(S_{1}(t),S_{2}(t),\cdots,S_{N}(t))\mathbf{A}(t)\leq\mathbf{\bar{A}},

where \leq is a partial order compoentwise, and

𝐀¯ij={Si0Lγ,ij,0,i=j.\displaystyle\mathbf{\bar{A}}_{ij}=\begin{cases}S_{i}^{0}L^{-\gamma},&i\neq j,\\ 0,&i=j.\end{cases}

Therfore, we obtain

(4.5) 𝐈˙=κ1diag(S1,S2,,SN)𝐀𝐈b𝐈(κ1𝐀¯bI)𝐈.\dot{\mathbf{I}}^{\top}=\kappa_{1}\text{diag}(S_{1},S_{2},\cdots,S_{N})\mathbf{A}\mathbf{I}^{\top}-b\mathbf{I}^{\top}\leq(\kappa_{1}\mathbf{\bar{A}}-bI)\mathbf{I}^{\top}.

We multiply the both sides of (4.5) by 𝟙\mathbbm{1} to yield

ddt(i=1NIi(t))i=1NλiIi(t),\displaystyle\frac{d}{dt}\left(\sum_{i=1}^{N}{I_{i}(t)}\right)\leq\sum_{i=1}^{N}\lambda_{i}I_{i}(t),

where λi\lambda_{i} is given as follows.

λi:=κ1LγjiSj0bi.\lambda_{i}:=\kappa_{1}L^{-\gamma}\sum_{j\neq i}S_{j}^{0}-b^{i}.

If

min1iNbi>κ1LγjiSj0,\displaystyle\min_{1\leq i\leq N}b^{i}>\kappa_{1}L^{-\gamma}\sum_{j\neq i}S_{j}^{0},

then λi\lambda_{i} is negative, so i=1NIi(t)\sum_{i=1}^{N}{I_{i}(t)} decays to zero exponentially fast. ∎

Remark 4.1.

Since

N1max1iNjiSj0,N-1\geq\max_{1\leq i\leq N}\sum_{j\neq i}S_{j}^{0},

the condition (4.4) gives a more relaxed condition on {bi}\{b^{i}\} compared to (4.3).

5. Quantitative estimates for relative distances

In this section, we provide explicit quantitative estimates on relative distances between particles. For a given spatial configuration {xi}\{x_{i}\}, we consider the set of all relative distances {xixj}\{\|x_{i}-x_{j}\|\} for iji\not=j, and rearrange them in an increasing order:

D1(t)D2(t)D𝒬(t)=:D(t),where𝒬:=N(N1)2.D_{1}(t)\leq D_{2}(t)\leq\cdots\leq D_{\mathcal{Q}}(t)=:D(t),\quad\mbox{where}~{}~{}{\mathcal{Q}}:=\frac{N(N-1)}{2}.

In what follows, we derive a uniform lower-bound for D1D_{1} and a uniform upper-bound for D𝒬D_{\mathcal{Q}} so that we can control the singular terms in (1.6) uniformly in time. First, we recall notation:

(5.1) Λ:=(12)𝒬max{D1(0),1}𝒬((max{(mr2Ma)1βα,1})𝒬(mr2N(Ma+Mr))𝒬β1.\Lambda:=\Big{(}\frac{1}{2}\Big{)}^{{\mathcal{Q}}}\max\{D_{1}(0),1\}^{\mathcal{Q}}\Big{(}\Big{(}\max\Big{\{}\Big{(}\frac{m_{r}}{2M_{a}}\Big{)}^{\frac{1}{\beta-\alpha}},1\Big{\}}\Big{)}^{\mathcal{Q}}\Big{(}\frac{m_{r}}{2N(M_{a}+M_{r})}\Big{)}^{\frac{{\mathcal{Q}}}{\beta-1}}.
Theorem 5.1.

Let {(Wi,xi)}\{(W_{i},x_{i})\} be a solution of system (1.6) with the initial data {(Wi0,xi0)}\{(W^{0}_{i},x^{0}_{i})\}. Then, the following assertions hold.

  1. (1)

    (Existence of a positive lower bound to D1D_{1}): If initial data satisfy

    min1ijNxi0xj0>0,\min_{1\leq i\not=j\leq N}\|x_{i}^{0}-x_{j}^{0}\|>0,

    there is a positive constant δ1>0\delta_{1}>0 such that

    inf0t<D1(t)δ1.\inf_{0\leq t<\infty}D_{1}(t)\geq\delta_{1}.
  2. (2)

    (Existence of a positive upper bound to D𝒬D_{\mathcal{Q}}): If initial data and system functions satisfy

    Λ>(κ3Mrκ2ma)βα,\Lambda>\left(\frac{\kappa_{3}M_{r}}{\kappa_{2}m_{a}}\right)^{\beta-\alpha},

    there exists a positive constant δ\delta_{\infty} such that

    sup0t<D𝒬(t)δ.\sup_{0\leq t<\infty}D_{\mathcal{Q}}(t)\leq\delta_{\infty}.
Proof.

We leave its proof in the next two subsections. ∎

5.1. A positive lower bound for minimal relative distance

In this subsection, we show that D1(t)D_{1}(t) has a positive lower bound δ1\delta_{1} which is independent of tt. Since the particles do not collide in finite-time interval, there exists an analytic solution in finite time (see Theorem 2.2). Therefore, each distance DijD_{ij} and ordered distance DiD_{i} are Lipschitz continuous. In addition, according to the analyticity, for given DijD_{ij}, there exist a sequence of times (tn)(t_{n}):

0=:t0<t1<t2<<tn<,0=:t_{0}<t_{1}<t_{2}<\cdots<t_{n}<\cdots,

such that we can decompose the whole time interval [0,)[0,\infty) into the union of subintervals

[0,)=l=1Tl,Tl=[tl1,tl),l1.[0,\infty)=\bigcup_{l=1}^{\infty}{T_{l}},\quad T_{l}=[t_{l-1},t_{l}),\quad l\geq 1.

to make sure that DijD_{ij} is not changed in each subinterval. To demonstrate the existence of positive uniform lower bound, we first provide three lemmas. First, we show a positive lower bound of maximal distance D𝒬(t)D_{\mathcal{Q}}(t):

(5.2) D(t):=D𝒬(t)=max1ijNxi(t)xj(t).\displaystyle D(t):=D_{\mathcal{Q}}(t)=\max_{1\leq i\not=j\leq N}\|x_{i}(t)-x_{j}(t)\|.

Next, we provide a series of lemmas to estimate maximal and minimal relative distances. First, we derive a differential inequality for DD.

Lemma 5.1.

Let {(Wi,xi)}\{(W_{i},x_{i})\} be a solution to system (1.6). Then, the functional DD in (5.2) satisfies

ddtD(t)2\displaystyle\frac{d}{dt}{D(t)}^{2} 4κ2MaNDα2+4κ3mrNDβ2+2Nki,j[(xixj)(xkxi)xkxiα(κ2Maκ3mrxkxiβα)]\displaystyle\geq-\frac{4\kappa_{2}M_{a}}{N\cdot D^{\alpha-2}}+\frac{4\kappa_{3}m_{r}}{N\cdot D^{\beta-2}}+\frac{2}{N}\sum_{k\neq i,j}\left[\frac{(x_{i}-x_{j})\cdot(x_{k}-x_{i})}{\|x_{k}-x_{i}\|^{\alpha}}\left(\kappa_{2}M_{a}-\frac{\kappa_{3}m_{r}}{\|x_{k}-x_{i}\|^{\beta-\alpha}}\right)\right]
+2Nki,j[(xixj)(xjxk)xkxjα(κ2Maκ3mrxkxjβα)].\displaystyle+\frac{2}{N}\sum_{k\neq i,j}\left[\frac{(x_{i}-x_{j})\cdot(x_{j}-x_{k})}{\|x_{k}-x_{j}\|^{\alpha}}\left(\kappa_{2}M_{a}-\frac{\kappa_{3}m_{r}}{\|x_{k}-x_{j}\|^{\beta-\alpha}}\right)\right].
Proof.

For given l{0}l\in\mathbb{N}\cup\{0\}, we choose ii and jj such that

D(t)=xi(t)xj(t)tTl.D(t)=\|x_{i}(t)-x_{j}(t)\|\quad\forall~{}t\in T_{l}.

Then, by direct calculation, we obtain

(5.3) ddtD(t)2=ddtxixj2=2(xixj)(xi˙xj˙)=2(xixj)[κ2Nki(Ψaikxkxixkxiα)κ3Nki(Ψrikxkxixkxiβ)]2(xixj)[κ2Nkj(Ψajkxkxjxkxjα)κ3Nkj(Ψrjkxkxjxkxjβ)]=4κ2ΨajkNxixjα2+4κ3ΨrjkNxixjβ2+2Nki,j[κ2Ψaik(xixj)(xkxi)xkxiακ3Ψrik(xixj)(xkxi)xkxiβ]2Nki,j[κ2Ψajk(xixj)(xkxj)xkxjακ3Ψrjk(xixj)(xkxj)xkxjβ]=:21+22+23,\displaystyle\begin{aligned} \frac{d}{dt}{D(t)}^{2}&=\frac{d}{dt}\|x_{i}-x_{j}\|^{2}=2(x_{i}-x_{j})\cdot(\dot{x_{i}}-\dot{x_{j}})\\ &=2(x_{i}-x_{j})\cdot\left[\frac{\kappa_{2}}{N}\sum_{k\neq i}{\left(\Psi_{a}^{ik}\frac{x_{k}-x_{i}}{\|x_{k}-x_{i}\|^{\alpha}}\right)-\frac{\kappa_{3}}{N}\sum_{k\neq i}\left(\Psi_{r}^{ik}\frac{x_{k}-x_{i}}{\|x_{k}-x_{i}\|^{\beta}}\right)}\right]\\ &\hskip 5.69046pt-2(x_{i}-x_{j})\cdot\left[\frac{\kappa_{2}}{N}\sum_{k\neq j}{\left(\Psi_{a}^{jk}\frac{x_{k}-x_{j}}{\|x_{k}-x_{j}\|^{\alpha}}\right)-\frac{\kappa_{3}}{N}\sum_{k\neq j}\left(\Psi_{r}^{jk}\frac{x_{k}-x_{j}}{\|x_{k}-x_{j}\|^{\beta}}\right)}\right]\\ &=-\frac{4\kappa_{2}\Psi^{jk}_{a}}{N\|x_{i}-x_{j}\|^{\alpha-2}}+\frac{4\kappa_{3}\Psi^{jk}_{r}}{N\|x_{i}-x_{j}\|^{\beta-2}}\\ &\hskip 5.69046pt+\frac{2}{N}\sum_{k\neq i,j}\left[\kappa_{2}\Psi_{a}^{ik}\frac{(x_{i}-x_{j})\cdot(x_{k}-x_{i})}{\|x_{k}-x_{i}\|^{\alpha}}-\kappa_{3}\Psi_{r}^{ik}\frac{(x_{i}-x_{j})\cdot(x_{k}-x_{i})}{\|x_{k}-x_{i}\|^{\beta}}\right]\\ &\hskip 5.69046pt-\frac{2}{N}\sum_{k\neq i,j}\left[\kappa_{2}\Psi_{a}^{jk}\frac{(x_{i}-x_{j})\cdot(x_{k}-x_{j})}{\|x_{k}-x_{j}\|^{\alpha}}-\kappa_{3}\Psi_{r}^{jk}\frac{(x_{i}-x_{j})\cdot(x_{k}-x_{j})}{\|x_{k}-x_{j}\|^{\beta}}\right]\\ &=:\mathcal{I}_{21}+\mathcal{I}_{22}+\mathcal{I}_{23},\end{aligned}

for tTlt\in T_{l}. We now estimate 2i\mathcal{I}_{2i} (i=1,2,3i=1,2,3) one by one.

\bullet (Estimate of 21\mathcal{I}_{21}): We use

ΨaMa,Ψrmrandxixj=D\Psi_{a}\leq M_{a},\quad\Psi_{r}\geq m_{r}\quad\mbox{and}\quad\|x_{i}-x_{j}\|=D

to get

(5.4) 214κ2MaNDα2+4κ3mrNDβ2.\mathcal{I}_{21}\geq-\frac{4\kappa_{2}M_{a}}{N\cdot D^{\alpha-2}}+\frac{4\kappa_{3}m_{r}}{N\cdot D^{\beta-2}}.

\bullet (Estimate of 22\mathcal{I}_{22}) : Since xixj\|x_{i}-x_{j}\| is the maximal distance,

(xixj)(xkxi)=(xixj)(xkxj+xjxi)=(xixj)(xkxj)|xixj|20.\displaystyle\begin{aligned} (x_{i}-x_{j})\cdot(x_{k}-x_{i})&=(x_{i}-x_{j})\cdot(x_{k}-x_{j}+x_{j}-x_{i})\\ &=(x_{i}-x_{j})\cdot(x_{k}-x_{j})-|x_{i}-x_{j}|^{2}\leq 0.\end{aligned}

Similarly, we have

(xixj)(xkxj)0.(x_{i}-x_{j})\cdot(x_{k}-x_{j})\geq 0.

Therefore, one has

(5.5) 222Nki,j[κ2Ma(xixj)(xkxi)xkxiακ3mr(xixj)(xkxi)xkxiβ]=2Nki,j[(xixj)(xkxi)xkxiα(κ2Maκ3mrxkxiβα)].\displaystyle\begin{aligned} \mathcal{I}_{22}&\geq\frac{2}{N}\sum_{k\neq i,j}\left[\kappa_{2}M_{a}\frac{(x_{i}-x_{j})\cdot(x_{k}-x_{i})}{\|x_{k}-x_{i}\|^{\alpha}}-\kappa_{3}m_{r}\frac{(x_{i}-x_{j})\cdot(x_{k}-x_{i})}{\|x_{k}-x_{i}\|^{\beta}}\right]\\ &=\frac{2}{N}\sum_{k\neq i,j}\left[\frac{(x_{i}-x_{j})\cdot(x_{k}-x_{i})}{\|x_{k}-x_{i}\|^{\alpha}}\left(\kappa_{2}M_{a}-\frac{\kappa_{3}m_{r}}{\|x_{k}-x_{i}\|^{\beta-\alpha}}\right)\right].\end{aligned}

\bullet (Estimate of 23\mathcal{I}_{23}): Similar to the estimate of 22{\mathcal{I}}_{22}, we obtain

(5.6) 232Nki,j[κ2Ma(xixj)(xjxk)xkxjακ3mr(xixj)(xjxk)xkxjβ]=2Nki,j[(xixj)(xjxk)xkxjα(κ2Maκ3mrxkxjβα)].\displaystyle\begin{aligned} \mathcal{I}_{23}&\geq\frac{2}{N}\sum_{k\neq i,j}\left[\kappa_{2}M_{a}\frac{(x_{i}-x_{j})\cdot(x_{j}-x_{k})}{\|x_{k}-x_{j}\|^{\alpha}}-\kappa_{3}m_{r}\frac{(x_{i}-x_{j})\cdot(x_{j}-x_{k})}{\|x_{k}-x_{j}\|^{\beta}}\right]\\ &=\frac{2}{N}\sum_{k\neq i,j}\left[\frac{(x_{i}-x_{j})\cdot(x_{j}-x_{k})}{\|x_{k}-x_{j}\|^{\alpha}}\left(\kappa_{2}M_{a}-\frac{\kappa_{3}m_{r}}{\|x_{k}-x_{j}\|^{\beta-\alpha}}\right)\right].\end{aligned}

In (5.3), we combine all the estimates (5.4), (5.5) and (5.6) to find the desired estimate. ∎

To derive a positive lower bound for D1D_{1}, we first verify that the maximal distance DD is bounded away from zero uniformly in time, and then we show that this positive lower bound propagates to the lower graded relative distance when we reach to the minimal relative distance D1D_{1} (the first assertion in Theorem 5.1) via the following two lemmas.
Now we derive a positive lower bound for the maximal distance D(t)D(t).

Lemma 5.2.

(maximal relative distance) Suppose that initial spatial configuration is non-collisional:

min1ijNxi0xj0>0,\min_{1\leq i\not=j\leq N}\|x_{i}^{0}-x_{j}^{0}\|>0,

and let {(Wi,xi)}\{(W_{i},x_{i})\} be a solution to system (1.6). Then, one has

inf0t<D(t)δ𝒬:=min{D(0),(κ3mrκ2Ma)1βα}.\inf_{0\leq t<\infty}D(t)\geq\delta_{\mathcal{Q}}:=\min\left\{D(0),~{}~{}\left(\frac{\kappa_{3}m_{r}}{\kappa_{2}M_{a}}\right)^{\frac{1}{\beta-\alpha}}\right\}.
Proof.

We use an induction argument on the time intervals Tl=[tl1,tl)T_{l}=[t_{l-1},t_{l}).

\bullet (Initial step): we claim that

D(t)δ𝒬,for tT1=[0,t1).D(t)\geq\delta_{\mathcal{Q}},\quad\text{for }t\in T_{1}=[0,t_{1}).

Suppose not, since δ𝒬D(0)\delta_{\mathcal{Q}}\leq D(0), there exist t11t_{11} and t12t_{12} such that

0t11<t12<t1,D(t11)=δ𝒬andD(t)<δ𝒬 in (t11,t12].0\leq t_{11}<t_{12}<t_{1},\quad D(t_{11})=\delta_{\mathcal{Q}}\quad\text{and}\quad D(t)<\delta_{\mathcal{Q}}\text{ in }(t_{11},t_{12}].

Therefore, for t(t11,t12]t\in(t_{11},t_{12}],

D<δ𝒬(κ3mrκ2Ma)1βα.\displaystyle D<\delta_{\mathcal{Q}}\leq\left(\frac{\kappa_{3}m_{r}}{\kappa_{2}M_{a}}\right)^{\frac{1}{\beta-\alpha}}.

For all ki,jk\neq i,j, we get

4κ2MaNDα2+4κ3mrNDβ20,κ2Maκ3mr|xkxi|βα0,κ2Maκ3mr|xkxj|βα0,-\frac{4\kappa_{2}M_{a}}{N\cdot D^{\alpha-2}}+\frac{4\kappa_{3}m_{r}}{N\cdot D^{\beta-2}}\geq 0,\quad\kappa_{2}M_{a}-\frac{\kappa_{3}m_{r}}{|x_{k}-x_{i}|^{\beta-\alpha}}\leq 0,\quad\kappa_{2}M_{a}-\frac{\kappa_{3}m_{r}}{|x_{k}-x_{j}|^{\beta-\alpha}}\leq 0,

By Lemma 4.1, one has

ddtD20for t(t11,t12].\frac{d}{dt}D^{2}\geq 0\quad\text{for }t\in(t_{11},t_{12}].

This implies

D(t12)D(t11)=δ𝒬,D(t_{12})\geq D(t_{11})=\delta_{\mathcal{Q}},

which is contradictory to

D(t)<δ𝒬in(t11,t12].D(t)<\delta_{\mathcal{Q}}\quad\text{in}\quad(t_{11},t_{12}].

\bullet (Inductive step) : Suppose that for some l1l\geq 1,

D(t)δ𝒬fortTl=[tl1,tl).D(t)\geq\delta_{\mathcal{Q}}\quad\text{for}\quad t\in T_{l}=[t_{l-1},t_{l}).

Now we consider D(t)D(t) on Tl+1T_{l+1}. Due to the continuity of D(t)D(t) and induction hypothesis, one has

D(tl)=limttlD(t)δ𝒬.D(t_{l})=\lim_{t\rightarrow t_{l}^{-}}{D(t)}\geq\delta_{\mathcal{Q}}.

Therefore, we can use the same criteria above to get

D(t)δ𝒬fortTl+1.D(t)\geq\delta_{\mathcal{Q}}\quad\text{for}\quad t\in T_{l+1}.

Thus, one has the desired estimate. ∎

In the following lemma, we show the backward propagation of a positive lower bound for the ordered relative distance.

Lemma 5.3.

(Backward propagation of lower bounds) Suppose that initial spatial configuration is non-collisional:

min1ijNxi0xj0>0,\min_{1\leq i\not=j\leq N}\|x_{i}^{0}-x_{j}^{0}\|>0,

and let {(Wi,xi)}\{(W_{i},x_{i})\} be a solution to system (1.6), and let p[1,𝒬)p\in[1,{\mathcal{Q}}) be a fixed constant. If there exist a positive constant δq\delta_{q} such that

inf0t<Dq(t)δq for all q(p,𝒬],\inf_{0\leq t<\infty}D_{q}(t)\geq\delta_{q}\text{ for all }~{}q\in(p,{\mathcal{Q}}],

then there exists a positive constant δp\delta_{p} such that

inf0t<Dp(t)δp.\inf_{0\leq t<\infty}D_{p}(t)\geq\delta_{p}.
Proof.

We refer to the proof of Lemma 3.4 in [20]. ∎

Now, we are ready to provide a positive lower bound for the minimal relative distance D1D_{1}.

Proof of the first assertion in Theorem 5.1: Suppose initial spatial configuration is noncollisional:

min1ijNxi0xj0>0,\min_{1\leq i\not=j\leq N}\|x_{i}^{0}-x_{j}^{0}\|>0,

and let {(Wi,xi)}\{(W_{i},x_{i})\} be a solution to system (1.6). By Lemma 5.2, there exists δ𝒬>0\delta_{\mathcal{Q}}>0 such that

inf0t<D(t)δ𝒬.\inf_{0\leq t<\infty}D(t)\geq\delta_{\mathcal{Q}}.

For q=𝒬q={\mathcal{Q}}, we apply Lemma 5.3 to show that

δ𝒬1>0such thatinf0t<D𝒬1(t)δ𝒬1.\exists~{}\delta_{{\mathcal{Q}}-1}>0\quad\mbox{such that}\quad\inf_{0\leq t<\infty}D_{{\mathcal{Q}}-1}(t)\geq\delta_{{\mathcal{Q}}-1}.

Again, we apply Lemma 5.3 inductively until we reach q=1q=1 to derive the desired a positive lower bound for D1D_{1}.

5.2. A positive upper bound for maximal relative distance

In this subsection, we derive an existence of a positive upper bound for the maximal relative distance. First, we study the estimate for a differential inequality.

Lemma 5.4.

Let y:[0,)+y:[0,\infty)\rightarrow\mathbb{R}^{+} be a differentiable function satisfying the following differential inequality:

{yayp+byq,t>0,y(0)=y0,\displaystyle\begin{cases}y^{\prime}\leq-ay^{-p}+by^{-q},\quad t>0,\\ y(0)=y^{0},\end{cases}

where constants a,b,pa,b,p and qq satisfy

a>0,b>0,0p<q,a>0,\quad b>0,\quad 0\leq p<q,

Then, there exists a positive constant δ\delta such that

sup0t<y(t)δ.\sup_{0\leq t<\infty}y(t)\leq\delta.
Proof.

We use phase line analysis for z˙=azp+bzq{\dot{z}}=-az^{-p}+bz^{-q}. For this, we define

y:=(ba)1qp.y^{*}:=\left(\frac{b}{a}\right)^{\frac{1}{q-p}}.

Then, it is easy to see that

f(y)>0 for y<y,f(y)=0 and f(y)<0 for y>y.f(y)>0\quad\text{ for }y<y^{*},\quad f(y^{*})=0\quad\text{ and }\quad f(y)<0\quad\text{ for }y>y^{*}.

Therefore, yy is uniformly bounded by max{y0,y}\max\{y^{0},y^{*}\}. ∎

Now, we are ready to provide a proof for the second assertion in Theorem 5.1.

Proof of the second assertion in Theorem 5.1: Suppose initial data and system parameters satisfy

D(0)<,δ1>(κ3Mrκ2ma)βα,D(0)<\infty,\quad\delta_{1}>\left(\frac{\kappa_{3}M_{r}}{\kappa_{2}m_{a}}\right)^{\beta-\alpha},

and let {(Wi,xi)}\{(W_{i},x_{i})\} be a solution to system (1.6). Then, for t(0,)t\in(0,\infty), we choose i=iti=i_{t} and j=jtj=j_{t} such that

D(t)=xi(t)xj(t).D(t)=\|x_{i}(t)-x_{j}(t)\|.

By Lemma 5.2, one has

(5.7) ddtD(t)2=ddtxixj2=2(xixj)(xi˙xj˙)=2(xixj)[κ2Nki(Ψaikxkxixkxiα)κ3Nki(Ψrikxkxixkxiβ)]2(xixj)[κ2Nkj(Ψajkxkxjxkxjα)κ3Nkj(Ψrjkxkxjxkxjβ)]=4κ2ΨajkNxixjα2+4κ3ΨrjkNxixjβ2+2Nki,j[κ2Ψaik(xixj)(xkxi)xkxiακ3Ψrik(xixj)(xkxi)xkxiβ]2Nki,j[κ2Ψajk(xixj)(xkxj)xkxjακ3Ψrjk(xixj)(xkxj)xkxjβ]=:31+32+33,tTl.\displaystyle\begin{aligned} &\frac{d}{dt}{D(t)}^{2}=\frac{d}{dt}\|x_{i}-x_{j}\|^{2}=2(x_{i}-x_{j})\cdot(\dot{x_{i}}-\dot{x_{j}})\\ &\hskip 28.45274pt=2(x_{i}-x_{j})\cdot\left[\frac{\kappa_{2}}{N}\sum_{k\neq i}{\left(\Psi_{a}^{ik}\frac{x_{k}-x_{i}}{\|x_{k}-x_{i}\|^{\alpha}}\right)-\frac{\kappa_{3}}{N}\sum_{k\neq i}\left(\Psi_{r}^{ik}\frac{x_{k}-x_{i}}{\|x_{k}-x_{i}\|^{\beta}}\right)}\right]\\ &\hskip 34.14322pt-2(x_{i}-x_{j})\cdot\left[\frac{\kappa_{2}}{N}\sum_{k\neq j}{\left(\Psi_{a}^{jk}\frac{x_{k}-x_{j}}{\|x_{k}-x_{j}\|^{\alpha}}\right)-\frac{\kappa_{3}}{N}\sum_{k\neq j}\left(\Psi_{r}^{jk}\frac{x_{k}-x_{j}}{\|x_{k}-x_{j}\|^{\beta}}\right)}\right]\\ &\hskip 28.45274pt=-\frac{4\kappa_{2}\Psi_{a}^{jk}}{N\|x_{i}-x_{j}\|^{\alpha-2}}+\frac{4\kappa_{3}\Psi_{r}^{jk}}{N\|x_{i}-x_{j}\|^{\beta-2}}\\ &\hskip 34.14322pt+\frac{2}{N}\sum_{k\neq i,j}\left[\kappa_{2}\Psi_{a}^{ik}\frac{(x_{i}-x_{j})\cdot(x_{k}-x_{i})}{\|x_{k}-x_{i}\|^{\alpha}}-\kappa_{3}\Psi_{r}^{ik}\frac{(x_{i}-x_{j})\cdot(x_{k}-x_{i})}{\|x_{k}-x_{i}\|^{\beta}}\right]\\ &\hskip 34.14322pt-\frac{2}{N}\sum_{k\neq i,j}\left[\kappa_{2}\Psi_{a}^{jk}\frac{(x_{i}-x_{j})\cdot(x_{k}-x_{j})}{\|x_{k}-x_{j}\|^{\alpha}}-\kappa_{3}\Psi_{r}^{jk}\frac{(x_{i}-x_{j})\cdot(x_{k}-x_{j})}{\|x_{k}-x_{j}\|^{\beta}}\right]\\ &\hskip 28.45274pt=:\mathcal{I}_{31}+\mathcal{I}_{32}+\mathcal{I}_{33},\qquad t\in T_{l}.\end{aligned}

In what follows, we estimate 3i\mathcal{I}_{3i} (i=1,2,3i=1,2,3) one by one.

\bullet (Estimate of 31\mathcal{I}_{31}): We use

D=xixj,ΨamaandΨrMrD=\|x_{i}-x_{j}\|,\quad\Psi_{a}\geq m_{a}\quad\mbox{and}\quad\Psi_{r}\leq M_{r}

to get

(5.8) 314κ2maNDα2+4κ3MrNDβ2.\mathcal{I}_{31}\leq-\frac{4\kappa_{2}m_{a}}{N\cdot D^{\alpha-2}}+\frac{4\kappa_{3}M_{r}}{N\cdot D^{\beta-2}}.

\bullet (Estimate of 32\mathcal{I}_{32}): We use D(t)=xixjD(t)=\|x_{i}-x_{j}\| to see

(xixj)(xkxi)=(xixj)(xkxj+xjxi)=(xixj)(xkxj)xixj20.\displaystyle\begin{aligned} (x_{i}-x_{j})\cdot(x_{k}-x_{i})&=(x_{i}-x_{j})\cdot(x_{k}-x_{j}+x_{j}-x_{i})\\ &=(x_{i}-x_{j})\cdot(x_{k}-x_{j})-\|x_{i}-x_{j}\|^{2}\leq 0.\end{aligned}

Now, we define the angle between xixjx_{i}-x_{j} and xkxix_{k}-x_{i} by θijk\theta_{ijk}. Since

xixjxkxicos(θijk)=(xixj)(xkxi)0,\|x_{i}-x_{j}\|\cdot\|x_{k}-x_{i}\|\cos(\theta_{ijk})=(x_{i}-x_{j})\cdot(x_{k}-x_{i})\leq 0,

one has

cos(θijk)0.\cos(\theta_{ijk})\leq 0.

Thus, one has

32\displaystyle\mathcal{I}_{32} 2Nki,j[κ2ma(xixj)(xkxi)xkxiακ3Mr(xixj)(xkxi)xkxiβ]\displaystyle\leq\frac{2}{N}\sum_{k\neq i,j}\left[\kappa_{2}m_{a}\frac{(x_{i}-x_{j})\cdot(x_{k}-x_{i})}{\|x_{k}-x_{i}\|^{\alpha}}-\kappa_{3}M_{r}\frac{(x_{i}-x_{j})\cdot(x_{k}-x_{i})}{\|x_{k}-x_{i}\|^{\beta}}\right]
=2Nki,j[κ2maxixjxkxicosθijkxkxiακ3Mrxixjxkxicosθijkxkxiβ]\displaystyle=\frac{2}{N}\sum_{k\neq i,j}\left[\kappa_{2}m_{a}\frac{\|x_{i}-x_{j}\|\cdot\|x_{k}-x_{i}\|\cos{\theta_{ijk}}}{\|x_{k}-x_{i}\|^{\alpha}}-\kappa_{3}M_{r}\frac{\|x_{i}-x_{j}\|\cdot\|x_{k}-x_{i}\|\cos{\theta_{ijk}}}{\|x_{k}-x_{i}\|^{\beta}}\right]
=2Nki,j[κ2maxixjcosθijkxkxiα1κ3Mrxixjcosθijkxkxiβ1]\displaystyle=\frac{2}{N}\sum_{k\neq i,j}\left[\kappa_{2}m_{a}\frac{\|x_{i}-x_{j}\|\cos{\theta_{ijk}}}{\|x_{k}-x_{i}\|^{\alpha-1}}-\kappa_{3}M_{r}\frac{\|x_{i}-x_{j}\|\cos{\theta_{ijk}}}{\|x_{k}-x_{i}\|^{\beta-1}}\right]
=2Nki,jxixjcosθijk[κ2maxkxiα1κ3Mrxkxiβ1].\displaystyle=\frac{2}{N}\sum_{k\neq i,j}\|x_{i}-x_{j}\|\cos{\theta_{ijk}}\left[\frac{\kappa_{2}m_{a}}{\|x_{k}-x_{i}\|^{\alpha-1}}-\frac{\kappa_{3}M_{r}}{\|x_{k}-x_{i}\|^{\beta-1}}\right].

Now, we define the function

f(d):=κ2madα1κ3Mrdβ1.f(d):=\frac{\kappa_{2}m_{a}}{d^{\alpha-1}}-\frac{\kappa_{3}M_{r}}{d^{\beta-1}}.

Then, we have

f(d)>0,d>(κ3Mrκ2ma)βα.f(d)>0,\quad\forall~{}d>\left(\frac{\kappa_{3}M_{r}}{\kappa_{2}m_{a}}\right)^{\beta-\alpha}.

Since

(κ3Mrκ2ma)βα<δ1|xkxi|,\left(\frac{\kappa_{3}M_{r}}{\kappa_{2}m_{a}}\right)^{\beta-\alpha}<\delta_{1}\leq|x_{k}-x_{i}|,

we have

(5.9) 320.\mathcal{I}_{32}\leq 0.

\bullet (Estimate of 33\mathcal{I}_{33}): Similar to 32{\mathcal{I}}_{32}, we have

(5.10) 330.\mathcal{I}_{33}\leq 0.

Finally, in (LABEL:E-5), we combine all the estimates (5.8), (5.9) and (5.10) to find

ddtD24κ2maNDα2+4κ3MrNDβ2,\frac{d}{dt}D^{2}\leq-\frac{4\kappa_{2}m_{a}}{N\cdot D^{\alpha-2}}+\frac{4\kappa_{3}M_{r}}{N\cdot D^{\beta-2}},

or equivalently

ddtD2κ2maNDα1+2κ3MrNDβ1.\frac{d}{dt}D\leq-\frac{2\kappa_{2}m_{a}}{N\cdot D^{\alpha-1}}+\frac{2\kappa_{3}M_{r}}{N\cdot D^{\beta-1}}.

Now, we apply Lemma 5.4 with parameters

a=2κ2maN,b=2κ3MrN,p=α1,q=β1a=\frac{2\kappa_{2}m_{a}}{N},\quad b=\frac{2\kappa_{3}M_{r}}{N},\quad p=\alpha-1,\quad q=\beta-1

to derive the desired estimate. \quad\qed

Remark 5.1.

Note that for 1α<β1\leq\alpha<\beta, nonexistence of finite-time collisions and uniform lower bound of diameter are always warranted, however conditions on parameter is needed to guarantee uniform upper bound of diameter.

6. Relaxation of spatial configuration

In this section, we study the relaxation of spatial configuration toward a constant configuration.

Consider a perturbed gradient system with decaying forcing ff:

(6.1) x˙(t)=xV(x(t))+f(t),t>0,\dot{x}(t)=-\nabla_{x}V(x(t))+f(t),\quad t>0,

where VV is a one-body potential.

Proposition 6.1.

Suppose that external forcing ff and potential VV satisfy the following conditions:

  1. (1)

    There exist positive constants CC and λ\lambda such that

    |f(t)|Ceλt,|f(t)|\leq Ce^{-\lambda t},
  2. (2)

    There exists a compact set KK such that x(t)x(t) is contained in KK for all t0t\geq 0,

  3. (3)

    V is analytic in KK and the map t|xV(x(t))|2t\mapsto|\nabla_{x}V(x(t))|^{2} is uniformly continuous in time,

and let x=x(t)x=x(t) be a solution of (6.1). Then, there exists xKx^{\infty}\in K such that

limtx(t)=x.\lim_{t\rightarrow\infty}x(t)=x^{\infty}.
Proof.

We refer to [20] for a proof. ∎

Now we are ready to prove the convergence of spatial configuration.

Theorem 6.1.

(Relaxation of spatial configuration) Suppose initial data and system parameters satisfy

Λ>(κ3Mrκ2ma)βα,\Lambda>\left(\frac{\kappa_{3}M_{r}}{\kappa_{2}m_{a}}\right)^{\beta-\alpha},

where Λ\Lambda is a positive constant defined in (5.1), and let {(Wi,xi)}\{(W_{i},x_{i})\} be a solution to system (1.6). Then, there exists {xi}\{x_{i}^{\infty}\} such that

limtxi(t)=xi,i[N].\lim_{t\rightarrow\infty}x_{i}(t)=x_{i}^{\infty},\quad\forall~{}i\in[N].
Proof.

We consider three cases below.

\bullet Case A (α2\alpha\neq 2 and β2\beta\neq 2): We set the potential function VV and the function ff by

X˙=XV(X)+f(t),V(X):=κ2Ni=1Nji((1+εa)xjxi2α2α)κ3Ni=1Nji(εrxjxi2β2β),f(t):=κ2Ni=1Nji((Ψaij1εa)xjxixjxiα)κ3Ni=1Nji((Ψrijεr)xjxixjxiβ).\displaystyle\begin{aligned} \dot{X}&=-\nabla_{X}V(X)+f(t),\\ V(X)&:=\frac{\kappa_{2}}{N}\sum_{i=1}^{N}\sum_{j\neq i}\left((1+\varepsilon_{a})\frac{\|x_{j}-x_{i}\|^{2-\alpha}}{2-\alpha}\right)-\frac{\kappa_{3}}{N}\sum_{i=1}^{N}\sum_{j\neq i}\left(\varepsilon_{r}\frac{\|x_{j}-x_{i}\|^{2-\beta}}{2-\beta}\right),\\ f(t)&:=\frac{\kappa_{2}}{N}\sum_{i=1}^{N}\sum_{j\neq i}\left((\Psi_{a}^{ij}-1-\varepsilon_{a})\frac{x_{j}-x_{i}}{\|x_{j}-x_{i}\|^{\alpha}}\right)-\frac{\kappa_{3}}{N}\sum_{i=1}^{N}\sum_{j\neq i}\left((\Psi_{r}^{ij}-\varepsilon_{r})\frac{x_{j}-x_{i}}{\|x_{j}-x_{i}\|^{\beta}}\right).\end{aligned}

Note that XX is in Nd{\mathbb{R}^{Nd}} when each of xix_{i} is dd-dimensional. First of all, we claim that the analyticity of the potential function VV defined in Nd\mathbb{R}^{Nd}. we can restrict the solution space because of the existence of uniform lower bound of distance. We may exclude the hyperplanes Eij:={xNd:xi=xj}E_{ij}:=\{x\in\mathbb{R}^{Nd}:x_{i}=x_{j}\} for all iji\neq j. So we write the restricted space as a union of open connected domain OkO_{k}

Nd\Eij=kOk.\mathbb{R}^{Nd}\big{\backslash}\bigcup E_{ij}=\bigcup_{k}O_{k}.

By the continuity of the solution, OkO_{k} are invariant sets for XX. Without loss of generality, we may assume X(0)O1X(0)\in O_{1} and thus X(t)O1X(t)\in O_{1}. That is O1O_{1} is an open domain and VV is analytic on O1O_{1}.

Now, we will show that there exists positive constants CC and λ\lambda such that

|f(t)|Ceλt.|f(t)|\leq Ce^{-\lambda t}.

Note that we have verified that there exists a positive constant cc such that

i=1NIiNect.\sum_{i=1}^{N}I_{i}\leq Ne^{-ct}.

It’s clear that for all ii,

Ii(t)Nect=:w.I_{i}(t)\leq Ne^{-ct}=:w.

In addition, one has

Ψaij1εa=2IiIjIiIj2w2+w+w(2N2+2N)ect,Ψrijεr=Ii+Ij2IiIj2w2+w+w(2N2+2N)ect.\displaystyle\begin{aligned} &\Psi_{a}^{ij}-1-\varepsilon_{a}=2I_{i}I_{j}-I_{i}-I_{j}\leq 2w^{2}+w+w\leq(2N^{2}+2N)e^{-ct},\\ &\Psi_{r}^{ij}-\varepsilon_{r}=I_{i}+I_{j}-2I_{i}I_{j}\leq 2w^{2}+w+w\leq(2N^{2}+2N)e^{-ct}.\end{aligned}

This means Ψa(Wi,Wj)1εa\Psi_{a}(W_{i},W_{j})-1-\varepsilon_{a} and Ψr(Wi,Wj)εr\Psi_{r}(W_{i},W_{j})-\varepsilon_{r} decay in exponential order. Since we have uniform upper and lower bounds of xixj\|x_{i}-x_{j}\|, there exist positive constants CC and λ\lambda such that

|f(t)|Ceλt.|f(t)|\leq Ce^{-\lambda t}.

Lastly, upper and lower bounds guarantee the boundedness of VX(X(t))V_{X}(X(t)) and XV(X(t))\nabla_{X}V(X(t)). Similarly, ddt|XV(X(t))|2\frac{d}{dt}|\nabla_{X}V(X(t))|^{2} is bounded too. Thus we get uniform continuity of |XV(X(t))|2|\nabla_{X}V(X(t))|^{2} in time. Thus, every hypothesis in Proposition 6.1 is satisfied to get XX^{\infty} that X(t)X(t) converges.

\bullet Case B (α=2)(\alpha=2): By setting

V(X):=κ2Ni=1Nji(2(1+εa)lnxjxi)κ3Ni=1Nji(εrxjxi2β2β),\displaystyle V(X):=\frac{\kappa_{2}}{N}\sum_{i=1}^{N}\sum_{j\neq i}\left(2(1+\varepsilon_{a})\ln{\|x_{j}-x_{i}\|}\right)-\frac{\kappa_{3}}{N}\sum_{i=1}^{N}\sum_{j\neq i}\left(\varepsilon_{r}\frac{\|x_{j}-x_{i}\|^{2-\beta}}{2-\beta}\right),

we can prove the desired estimate similarly as in Case A.

\bullet Case C (β=2)(\beta=2): Similar to Case B, we can show the desired estimate with

V(X)\displaystyle V(X) :=κ2Ni=1Nji((1+εa)xjxi2α2α)κ3Ni=1Nji(2εrlnxixj).\displaystyle:=\frac{\kappa_{2}}{N}\sum_{i=1}^{N}\sum_{j\neq i}\left((1+\varepsilon_{a})\frac{\|x_{j}-x_{i}\|^{2-\alpha}}{2-\alpha}\right)-\frac{\kappa_{3}}{N}\sum_{i=1}^{N}\sum_{j\neq i}\left(2\varepsilon_{r}\ln{\|x_{i}-x_{j}\|}\right).

Next, we consider a two-particle system with the following initial SIR state:

S1(0)=s1,I1(0)=1s1,R1(0)=0,S2(0)=s2,I2(0)=1s2,R2(0)=0,\displaystyle\begin{aligned} &S_{1}(0)=s_{1},\quad I_{1}(0)=1-s_{1},\quad R_{1}(0)=0,\\ &S_{2}(0)=s_{2},\quad I_{2}(0)=1-s_{2},\quad R_{2}(0)=0,\end{aligned}

and we set

s:=max{s1,s2}.s:=\max\{s_{1},s_{2}\}.

Consider a system for spatial position:

x˙1=κ22Ψa12x2x1x1x2ακ32Ψr12x2x1x1x2β,\displaystyle\dot{x}_{1}=\frac{\kappa_{2}}{2}\Psi_{a}^{12}\frac{x_{2}-x_{1}}{\|x_{1}-x_{2}\|^{\alpha}}-\frac{\kappa_{3}}{2}\Psi_{r}^{12}\frac{x_{2}-x_{1}}{\|x_{1}-x_{2}\|^{\beta}},
x˙2=κ22Ψa12x1x2x1x2ακ32Ψr12x1x2x1x2β.\displaystyle\dot{x}_{2}=\frac{\kappa_{2}}{2}\Psi_{a}^{12}\frac{x_{1}-x_{2}}{\|x_{1}-x_{2}\|^{\alpha}}-\frac{\kappa_{3}}{2}\Psi_{r}^{12}\frac{x_{1}-x_{2}}{\|x_{1}-x_{2}\|^{\beta}}.

Note that since x˙1+x˙2=0\dot{x}_{1}+\dot{x}_{2}=0, the center of the two particles is constant:

x1(t)+x2(t)2=x10+x202,t0.\frac{x_{1}(t)+x_{2}(t)}{2}=\frac{x_{1}^{0}+x_{2}^{0}}{2},\quad t\geq 0.

Now, we set

x:=x1x2,x:=x_{1}-x_{2},

and derive the equation for xx:

x˙=x˙1x˙2=κ2Ψa12x|x|α+κ3Ψr12x|x|β=x|x|α(κ2Ψa12κ3Ψr121|x|βα).\dot{x}=\dot{x}_{1}-\dot{x}_{2}=\kappa_{2}\Psi_{a}^{12}\frac{-x}{|x|^{\alpha}}+\kappa_{3}\Psi_{r}^{12}\frac{x}{|x|^{\beta}}=-\frac{x}{|x|^{\alpha}}\left(\kappa_{2}\Psi_{a}^{12}-\kappa_{3}\Psi_{r}^{12}\frac{1}{|x|^{\beta-\alpha}}\right).

As long as each coordinate of xx is positive,

x|x|α((1+εa)εr1|x|βα)x˙x|x|α(εa(1+εr)1|x|βα).-\frac{x}{|x|^{\alpha}}\left((1+\varepsilon_{a})-\varepsilon_{r}\cdot\frac{1}{|x|^{\beta-\alpha}}\right)\leq\dot{x}\leq-\frac{x}{|x|^{\alpha}}\left(\varepsilon_{a}-(1+\varepsilon_{r})\cdot\frac{1}{|x|^{\beta-\alpha}}\right).

Thus, one has

min{x0,(εr1+εa)1βα}<x(t)<max{x0,(1+εrεa)1βα},\min\left\{x^{0},\left(\frac{\varepsilon_{r}}{1+\varepsilon_{a}}\right)^{\frac{1}{\beta-\alpha}}\right\}<x(t)<\max\left\{x^{0},\left(\frac{1+\varepsilon_{r}}{\varepsilon_{a}}\right)^{\frac{1}{\beta-\alpha}}\right\},

so the difference xx has upper and lower bounds that are uniform in time.

It follows from Lemma 5.2 that for δQ=min{D(0),(κ3mrκ2Ma)1βα}\delta_{Q}=\min\left\{D(0),{\left(\frac{\kappa_{3}m_{r}}{\kappa_{2}M_{a}}\right)}^{\frac{1}{\beta-\alpha}}\right\},

inf0t<D(t)δQ,\inf_{0\leq t<\infty}D(t)\geq\delta_{Q},

and D(t)=x(t)D(t)=x(t). Therefore, x(t)x(t) has both upper and lower bounds and by the same reason in Theorem 6.1, there exists xx^{\infty} such that x(t)x(t) converges, as tt tends to infinity.

Now, we consider an equilibrium (Si,Ii,Ri)(S_{i}^{\infty},I_{i}^{\infty},R_{i}^{\infty}):

S˙1=a12S1I2=0,I˙1=a12S1I2bI1=0.\dot{S}_{1}^{\infty}=a_{12}S_{1}^{\infty}I_{2}^{\infty}=0,\quad\dot{I}_{1}^{\infty}=a_{12}S_{1}^{\infty}I_{2}^{\infty}-bI_{1}^{\infty}=0.

Thus we have

I1=0,I2=0.I_{1}^{\infty}=0,\quad I_{2}^{\infty}=0.

This yields

Ψa12=1+εaandΨr12=εr.\Psi_{a}^{12}=1+\varepsilon_{a}\quad\mbox{and}\quad\Psi_{r}^{12}=\varepsilon_{r}.

Since x˙=0\dot{x}^{\infty}=0, we have

κ2=κ3εr(1+εa)|x|βα,i.e.,x=(κ3εrκ2(1+εa))1βα.\kappa_{2}=\frac{\kappa_{3}\varepsilon_{r}}{(1+\varepsilon_{a})|x^{\infty}|^{\beta-\alpha}},\quad\mbox{i.e.,}\quad x^{\infty}=\left(\frac{\kappa_{3}\varepsilon_{r}}{\kappa_{2}(1+\varepsilon_{a})}\right)^{\frac{1}{\beta-\alpha}}.

In addition, we can weaken the condition that guarantees the exponential decay of IiI_{i}.

Theorem 6.2.

Suppose system parameters satisfy

b>2κ1(x+L)γ.b>\frac{2\kappa_{1}}{(x^{\infty}+L)^{\gamma}}.

and let {(Wi,xi)}\{(W_{i},x_{i})\} be a solution to system (2.1). Then, there exist positive constants A,εA,\varepsilon^{*} and t0t_{0} such that

|I1(t)+I2(t)|<Aeb2t,t>t0,\left|I_{1}(t)+I_{2}(t)\right|<Ae^{-\frac{b}{2}t},\quad t>t_{0},

where AA is a positive constant defined by

A:=etε0tε(I1(t)+I2(t))𝑑t.A:=e^{t_{\varepsilon^{*}}}\int_{0}^{t_{\varepsilon^{*}}}{(I_{1}(t)+I_{2}(t))dt}.
Proof.

Note that

limtx(t)=x.\lim_{t\rightarrow\infty}x(t)=x^{\infty}.

Thus, for every ε>0\varepsilon>0, there exists positive constant tεt_{\varepsilon} with respect to ε\varepsilon such that

t>tε|x(t)x|<ε.t>t_{\varepsilon}\quad\Longrightarrow\quad|x(t)-x^{\infty}|<\varepsilon.

By the given condition, there exists a positive constant ε\varepsilon^{*} such that

κ1(xε+L)γ<b2.\frac{\kappa_{1}}{(x^{\infty}-\varepsilon^{*}+L)^{\gamma}}<\frac{b}{2}.

For t>tεt>t_{\varepsilon^{*}},

(I1(t)+I2(t))\displaystyle(I_{1}(t)+I_{2}(t))^{\prime} =κ1(|x|+L)γ(S1I2+S2I1)b(I1+I2)\displaystyle=\frac{\kappa_{1}}{(|x|+L)^{\gamma}}(S_{1}I_{2}+S_{2}I_{1})-b(I_{1}+I_{2})
<κ1(xε+L)γ(I2+I1)b(I1+I2)\displaystyle<\frac{\kappa_{1}}{(x^{\infty}-\varepsilon^{*}+L)^{\gamma}}(I_{2}+I_{1})-b(I_{1}+I_{2})
<b2(I2+I1)b(I1+I2)=b2(I2+I1).\displaystyle<\frac{b}{2}(I_{2}+I_{1})-b(I_{1}+I_{2})=-\frac{b}{2}(I_{2}+I_{1}).

This yields

I1(t)+I2(t)<Aeb2t,t>tε.I_{1}(t)+I_{2}(t)<Ae^{-\frac{b}{2}t},\quad t>t_{\varepsilon^{*}}.

Remarks.

Since

Si(t)=1Ii(t)Ri(t)=1Ii(t)b0tIi(τ)𝑑τ,S_{i}(t)=1-I_{i}(t)-R_{i}(t)=1-I_{i}(t)-b\int_{0}^{t}I_{i}(\tau)d\tau,

we get

dI1(t)dt\displaystyle\frac{dI_{1}(t)}{dt} =a12(t)(1I1(t)b0tI1(τ)𝑑τ)I2(t)bI1(t),\displaystyle=a_{12}(t)\left(1-I_{1}(t)-b\int_{0}^{t}{I_{1}(\tau)d\tau}\right)I_{2}(t)-bI_{1}(t),
dI2(t)dt\displaystyle\frac{dI_{2}(t)}{dt} =a12(t)(1I2(t)b0tI2(τ)𝑑τ)I1(t)bI2(t),\displaystyle=a_{12}(t)\left(1-I_{2}(t)-b\int_{0}^{t}{I_{2}(\tau)d\tau}\right)I_{1}(t)-bI_{2}(t),
dx(t)dt\displaystyle\frac{dx(t)}{dt} =x|x|α(κ2Ψa12κ3Ψr121|x|βα).\displaystyle=-\frac{x}{|x|^{\alpha}}\left(\kappa_{2}\Psi_{a}^{12}-\kappa_{3}\Psi_{r}^{12}\cdot\frac{1}{|x|^{\beta-\alpha}}\right).

The Jacobian matrix of system at (I1,I2)=(0,0)(I_{1},I_{2})=(0,0) is

(ba12(1b0tI1(τ)𝑑τ)0a12(1b0tI2(τ)𝑑τ)b0).\begin{pmatrix}-b&a_{12}(1-b\int_{0}^{t}{I_{1}(\tau)d\tau})&0\\ a_{12}(1-b\int_{0}^{t}{I_{2}(\tau)d\tau})&-b&0\\ *&*&*\end{pmatrix}.

Next, we consider two cases for bb.

\bullet Case A: Consider the case

b>a12()[(1b0I1(τ)𝑑τ)(1b0I2(τ)𝑑τ)]1/2=κ1(x+L)γ[(1b0I1(τ)𝑑τ)(1b0I2(τ)𝑑τ)]1/2.\displaystyle\begin{aligned} b&>a_{12}(\infty)\Big{[}\Big{(}1-b\int_{0}^{\infty}{I_{1}(\tau)d\tau}\Big{)}\Big{(}1-b\int_{0}^{\infty}{I_{2}(\tau)d\tau}\Big{)}\Big{]}^{1/2}\\ &=\frac{\kappa_{1}}{(x^{\infty}+L)^{\gamma}}\Big{[}\Big{(}1-b\int_{0}^{\infty}{I_{1}(\tau)d\tau}\Big{)}\Big{(}1-b\int_{0}^{\infty}{I_{2}(\tau)d\tau}\Big{)}\Big{]}^{1/2}.\end{aligned}

In this case, all eigenvalues are real and eigenvalues for I!I_{!} and I2I_{2} are negative. This means that I1I_{1} and I2I_{2} decay to (0,0)(0,0) exponentially fast.

\bullet Case B: Consider the case

bκ1(x+L)γ[(1b0I1(τ)𝑑τ)(1b0I2(τ)𝑑τ)]1/2.b\leq\frac{\kappa_{1}}{(x^{\infty}+L)^{\gamma}}\Big{[}\Big{(}1-b\int_{0}^{\infty}{I_{1}(\tau)d\tau}\Big{)}\Big{(}1-b\int_{0}^{\infty}{I_{2}(\tau)d\tau}\Big{)}\Big{]}^{1/2}.

The right-hand side is less or equal to

κ12(x+L)γ[(1b0I1(τ)𝑑τ)+(1b0I2(τ)𝑑τ)].\frac{\kappa_{1}}{2(x^{\infty}+L)^{\gamma}}\Big{[}\Big{(}1-b\int_{0}^{\infty}{I_{1}(\tau)d\tau}\Big{)}+\Big{(}1-b\int_{0}^{\infty}{I_{2}(\tau)d\tau}\Big{)}\Big{]}.

This yields

0(I1+I2)(t)𝑑t2b2(x+L)γκ1.\int_{0}^{\infty}{(I_{1}+I_{2})(t)dt}\leq\frac{2}{b}-\frac{2(x^{\infty}+L)^{\gamma}}{\kappa_{1}}.

Note that the left-hand side 0(I1+I2)(t)𝑑t\int_{0}^{\infty}{(I_{1}+I_{2})(t)dt} denotes the number of total infections. Thus, we can say that Ii(t)I_{i}(t) decays to zero exponentially fast, or the total number of total infection is bounded.

7. Numerical Simulations

In this section, we provide several numerical examples to confirm analytical convergence results that we have shown in previous sections. At first, we set initial condition by (Si(0),Ii(0),Ri(0))=(1,0,0)(S_{i}(0),I_{i}(0),R_{i}(0))=(1,0,0) for uninfected 16 people and (Si(0),Ii(0),Ri(0))=(0.1,0.9,0)(S_{i}(0),I_{i}(0),R_{i}(0))=(0.1,0.9,0) for infected 4 people. Initial location is set by random seed in 3 by 3 plane. We set

κ1=1,εa=εr=0.2,\kappa_{1}=1,\quad\varepsilon_{a}=\varepsilon_{r}=0.2,

and we use the fourth order Runge-Kutta method for all simulations.

\bullet (Convergence of (Si,Ii,Ri)(S_{i},I_{i},R_{i})): In Figure 1 and Figure 2, we used

b=0.4,γ=1,L=1,b=0.4,\quad\gamma=1,\quad L=1,

here. In Theorem 4.1, we have shown that SiS_{i}, IiI_{i}, and RiR_{i} converges. To authenticate this assertion by numerical method, we observed convergence of SiS_{i}, IiI_{i}, and RiR_{i}.

Refer to caption
(a) S,I,RS,I,R for α=1\alpha=1, β=2\beta=2, κ2=1\kappa_{2}=1, κ3=5\kappa_{3}=5
Refer to caption
(b) S,I,RS,I,R for α=2\alpha=2, β=5\beta=5, κ2=10\kappa_{2}=10, κ3=1\kappa_{3}=1
Figure 1. Convergence of S,IS,I and RR

Moreover, by observing S˙i\dot{S}_{i}, I˙i\dot{I}_{i}, and R˙i\dot{R}_{i} that converges to zero, we obtain that Si,IiS_{i},I_{i} and RiR_{i} converge.

Refer to caption
(a) α=1\alpha=1, β=2\beta=2, κ2=1\kappa_{2}=1, κ3=5\kappa_{3}=5
Refer to caption
(b) α=2\alpha=2, β=5\beta=5, κ2=10\kappa_{2}=10, κ3=1\kappa_{3}=1
Figure 2. Convergence of S˙,I˙\dot{S},\dot{I} and R˙\dot{R}

\bullet (Exponential decay of iIi\sum_{i}{I_{i}}): For arbitrary NN, we showed that iIi\sum_{i}{I_{i}} decays to zero exponentially fast when b>κ1(N1)Lγb>\frac{\kappa_{1}(N-1)}{L^{\gamma}} as in Theorem 4.1. In Figure 3 and Figure 4, we set

b=1,γ=3,L=3so thatκ1(N1)Lγ0.7037<1=b,b=1,~{}\gamma=3,~{}L=3\quad\mbox{so that}\quad\frac{\kappa_{1}(N-1)}{L^{\gamma}}\approx 0.7037<1=b,

and observed two simulations based on two set of system parameters:

(α,β,κ2,κ3)=(1,2,1,5),(2,5,10,1).(\alpha,\beta,\kappa_{2},\kappa_{3})=(1,2,1,5),\quad(2,5,10,1).
Refer to caption
(a) S,I,RS,I,R for α=1\alpha=1, β=2\beta=2, κ2=1\kappa_{2}=1, κ3=5\kappa_{3}=5
Refer to caption
(b) logiIi\log\sum_{i}{I_{i}} for α=1\alpha=1, β=2\beta=2, κ2=1\kappa_{2}=1, κ3=5\kappa_{3}=5
Figure 3. Exponential decay of logiIi\log\sum_{i}{I_{i}}
Refer to caption
(a) S,I,RS,I,R for α=2\alpha=2, β=5\beta=5, κ2=10\kappa_{2}=10, κ3=1\kappa_{3}=1
Refer to caption
(b) logiIi\log\sum_{i}{I_{i}} for α=2\alpha=2, β=5\beta=5, κ2=10\kappa_{2}=10, κ3=1\kappa_{3}=1
Figure 4. Exponential decay of log(iIi)\log\Big{(}\sum_{i}{I_{i}}\Big{)}

\bullet (Effect of social distancing): Since κ2\kappa_{2} and κ3\kappa_{3} are coefficients of attracting and repulsion forces, respectively, larger ratio κ3κ2\frac{\kappa_{3}}{\kappa_{2}} represents for more intensive social distancing. We observe that the maximal value of iIi\sum_{i}{I_{i}} decreases, as κ3κ2\frac{\kappa_{3}}{\kappa_{2}} increases. We set L=1L=1 and κ2=1\kappa_{2}=1, and changed values of bb and Ψ\Psi with κ3=1,5,10,50\kappa_{3}=1,5,10,50. We plot graph for the cases of κ3κ2=1,5,10\frac{\kappa_{3}}{\kappa_{2}}=1,5,10 and 5050.

Refer to caption
(a) 1NiIi\frac{1}{N}\sum_{i}{I_{i}} for β=0.2\beta=0.2, Ψ=1\Psi=1
Refer to caption
(b) 1NiIi\frac{1}{N}\sum_{i}{I_{i}} for β=0.4\beta=0.4, Ψ=3\Psi=3
Figure 5. Average of Ii{I_{i}}, as κ3κ2\frac{\kappa_{3}}{\kappa_{2}} changes

\bullet (Behavior of the particles): We observed that infectious particles aggregated and they were isolated from non-infectious factors. It implies that this model can effectively reduce the number of infectious particles by adjusting the coefficients. We illustrated the behavior of the particles in

b=0.4,γ=3,L=1,b=0.4,~{}\gamma=3,~{}L=1,

and κ2=1\kappa_{2}=1, κ3=10\kappa_{3}=10. The purple ones represents for non-infectious particles. Moreover, we compared with the model that has no attracting and no repulsing term by setting κ2=κ3=0\kappa_{2}=\kappa_{3}=0. It is illustrated in the first graph in Figure 6 , red one is for κ2=1\kappa_{2}=1, κ3=10\kappa_{3}=10 and blue one is for κ2=κ3=0\kappa_{2}=\kappa_{3}=0. We could decrease the average of Ii(t)I_{i}(t).

Refer to caption
Figure 6. Behavior of particles in b=0.4,Ψ=3,L=1b=0.4,\Psi=3,L=1

We set

b=0.2,γ=3,L=3,α=2,β=5b=0.2,~{}\gamma=3,~{}L=3,~{}\alpha=2,~{}\beta=5

and repeated the process.

Refer to caption
Figure 7. Behavior of particles in b=0.2,Ψ=3,L=3b=0.2,\Psi=3,L=3

8. conclusion

In this paper, we have studied the emergent behaviors of a flock with SIR internal states, and have presented a new particle model with an aggregate property and epidemic internal forces by combining the SIR model and the swarmalator model. We considered that each particle has a state vector which is a kind of internal state following the SIR dynamics. From this argument, we could obtain the SIR-aggregation model by imposing the repulsive/attractive force between two particles depending on the distance between two particles and the internal states. We imposed strong repulsive force if one of them has a high probability of infection to model the social distancing, in contrast, we imposed weak repulsive force if both of them has low probabilities of infection. From this modeling, we modeled the social distancing of particles. We expect that we can do numeric experiments to find the efficiency of social distance for given strategies. We also provided the theoretical result of the SIR model, for example, nonexistence of finite-time collisions, positive lower bound for minimal relative distance, and a uniform upper bound for spatial diameter. From the numerical simulations, we could check that the isolation of the infected particle is an effective way to reduce the number of infected particles. Since we only considered that the recovered particles were never infected again, we have monotonicity on the number of susceptible/recovered particles. From those properties, we could prove the emergent dynamics. In our proposed model, we did not consider the reinfection which can happen in reality. Thus, we will leave this issue for a future work.

References

  • [1] J. A. Acebron, L. L. Bonilla, C. J. P. Pérez Vicente, F. Ritort and R. Spigler, The Kuramoto model: A simple paradigm for synchronization phenomena, Rev. Mod. Phys. 77 (2005) 137–185.
  • [2] G. Ajmone Marsan, N. Bellomo and L. Gibelli, Towards a systems approach to behavioral social dynamics, Math. Models Methods Appl. Sci. 26 (2016) 1051–1093.
  • [3] G Albi, N. Bellomo, L. Fermo, S.-Y. Ha, J. Kim, L. Pareschi, D. Poyato and J. Soler, Vehicular traffic, crowds, and swarms: From kinetic theory and multiscale methods to applications and research perspectives, Math. Models Methods Appl. Sci. 29 (2019) 1901–2005.
  • [4] K. M. Ariful Kabir, K. Kuga and J. Tanimoto, Analysis of SIR epidemic model with information spreading of awareness, Chaos, Solitons and Fractals, 119 (2019), 118-125.
  • [5] M. Ballerini, N. Cabibbo, R. Candelier, A. Cavagna, E. Cisbani, I. Giardina, V. Lecomte, A. Orlandi, G. Parisi, A. Procaccini, M. Viale, and V. Zdravkovic, Interaction ruling animal collective behavior depends on topological rather than metric distance: Evidence from a field study, Proc. Natl. Acad. Sci. USA 105 (2008) 1232–1237.
  • [6] N. Bellomo and S.-Y. Ha, A quest toward a mathematical theory of the dynamics of swarms, Math. Models Methods Appl. Sci. 27 (2017) 745–770.
  • [7] N. Bellomo, S.-Y. Ha and N. Outada, Towards a mathematical theory of behavioral swarms, To appear in ESIAM: Control, Optimization and Calculus of Variations.
  • [8] A. L. Bertozzi and J. Brandman, Finite-time blow-up of LL^{\infty}-weak solutions of an aggregation equation, Commun. Math. Sci. 8 (2010) 45–65.
  • [9] A. Bertozzi, J. A. Carrillo and T. Laurent, Blow-up in multidimensional aggregation equations with mildly singular interaction kernels, Nonlinearity 22 (2009), 683–710.
  • [10] A. L. Bertozzi, T. Laurent and J. Rosado, LpL^{p} theory for the multidimensional aggregation equation, Commun. Pure Appl. Math. 64 (2011) 45–83.
  • [11] J. Buck and E. Buck, Biology of sychronous flashing of fireflies, Nature 211 (1966) 562.
  • [12] J. A. Carrillo, Y.-P. Choi, P. B. Mucha and Jan Peszek, Sharp conditions to avoid collisions in singular Cucker–Smale interactions, Nonlinear Analysis: Real World Applications 37 (2017) 317–328.
  • [13] A. Cavagna, L. D. Castello, I. Giardina, T. Grigera, A. Jelic, S. Melillo, T. Mora, L. Parisi, E. Silvestri, M. Viale and A. M. Walczak, Flocking and turning: a new model for self-organized collective motion, J. Stat. Phys. 158 (2015) 601–627.
  • [14] S. Chinviriyasit and W. Chinviriyasit, Numerical modelling of an SIR epidemic model with diffusion, Applied Mathematics and Computation, 216 (2010), 395-409.
  • [15] F. Cucker and S. Smale, Emergent behavior in flocks, IEEE Trans. Automat. Control 52 (2007) 852–862.
  • [16] P. Degond and S. Motsch, Large-scale dynamics of the Persistent Turing Walker model of fish behavior, J. Stat. Phys. 131 (2008) 989–1022.
  • [17] R. Duan, M. Fornasier and G. Toscani, A kinetic flocking model with diffusion, Comm. Math. Phys. 300 (2010) 95–145.
  • [18] C. Gai, D. Iron and T. Kolokolnikov, Localized outbreaks in an S-I-R model with diffusion, J. of Math. Bio. 80 (2020) 1389–1411.
  • [19] S.-Y. Ha, J. Jung, J. Kim, J. Park and X. Zhang, A mean-field limit of the particle swarmalator model, To appear in Kinetic and Related Models.
  • [20] S.-Y. Ha, J. Jung, J. Kim, J. Park and X. Zhang, Emergent behaviors of the swarmalator model for position-phase aggregation, Math. Models Methods Appl. Sci. 29 (2019) 2225–2269.
  • [21] S.-Y. Ha, D. Ko, J. Park and X. Zhang, Collective synchronization of classical and quantum oscillators, EMS Surveys in Mathematical Sciences 3 (2016) 209–267.
  • [22] S.-Y. Ha and J.-G. Liu, A simple proof of Cucker-Smale flocking dynamics and mean–field limit, Commun. Math. Sci. 7 (2009) 297–325.
  • [23] S.-Y. Ha and T. Ruggeri, Emergent dynamics of a thermodynamically consistent particle model, Arch. Ration. Mech. Anal. 223 (2017) 1397–1425.
  • [24] S.-Y. Ha and E. Tadmor, From particle to kinetic and hydrodynamic description of flocking, Kinet. Relat. Models 1 (2008) 415–435.
  • [25] W. O. Kermack and A. G. McKendrick, A contribution to the mathematical theory of epidemics, Proc. Roy. Soc. Lond. A 115 (1927), 700–721.
  • [26] T. Kolokolnikov, J. A. Carrillo, A. Bertozzi, R. Fetecau, M. Lewis, Emergent behaviour in multi-particle systems with non-local interactions, Phys. D 260 (2013), 1–4.
  • [27] Y. Kuramoto, International symposium on mathematical problems in mathematical physics, Lecture Notes in Theoretical Physics 30 420 (1975).
  • [28] Z. Liu, Z. Shen, H. Wang and Z. Jin, Analysis of a local diffusive SIR model with seasonality and nonlocal incidence of infection, SIAM J. Appl. Math., 79 (2019), 2218-2241.
  • [29] S. Motsch and E. Tadmor, Heterophilious dynamics enhances consensus, SIAM Rev. 56 (2014) 577–621.
  • [30] S. Motsch and E. Tadmor, A new model for self-organized dynamics and its flocking behavior, J. Stat. Phys. 144 (2011) 923–947.
  • [31] K. P. O’Keeffe, J. H. Evers and T. Kolokolnikov, Ring states in swarmalator systems, Phys. Rev. E 98 (2018) 022203.
  • [32] K. P. O’Keeffe, H. Hong and S. H. Strogatz, Oscillators that sync and swarm, Nature Communications 8 (2017) 1504.
  • [33] A. Pikovsky, M. Rosenblum and J. Kurths, Synchronization: A universal concept in nonlinear sciences, Cambridge University Press, Cambridge, 2001.
  • [34] J. Toner and Y. Tu, Flocks, herds, and Schools: A quantitative theory of flocking, Phys. Rev. E 58 (1988) 4828–4858.
  • [35] C. Vargas-De-Le’on, On the global stability of SIS , SIR and SIRS epidemic models with standard incidence, Chaos, Solitons and Fractals, 44 (2011), 1106-1110.
  • [36] T. Vicsek and A. Zefeiris, Collective motion, Phys. Rep. 517 (2012) 71–140.
  • [37] T. Vicsek, Czirók, E. Ben-Jacob, I. Cohen and O. Schochet, Novel type of phase transition in a system of self-driven particles, Phys. Rev. Lett. 75 (1995) 1226–1229.
  • [38] A. T. Winfree, Biological rhythms and the behavior of populations of coupled oscillators, J. Theor. Biol. 16 (1967) 15–42.