
We studied the dynamics of thermodynamic Cucker–Smale (TCS) particles moving with a constant speed constraint. The TCS model describes the collective dynamics of the population of birds with a time varying internal variable, and it was first introduced as the generalization of the Cucker–Smale (CS) model. In this paper, we considered a modification of the TCS model in which each agent moves at a constant speed, such as the Vicsek model, and we additionally considered the effect of time-delays due to the finiteness of the information propagation speed between agents. Then, we presented several sufficient conditions in terms of initial data and system parameters to exhibit asymptotic flocking. We presented two kinds of results for this purpose. One was an estimate of the diameter of the velocity and temperature configuration, and the other was an estimate of the diameter of the configuration within the time-delay bound τ.
Citation: Hyunjin Ahn, Woojoo Shim. Interplay of a unit-speed constraint and time-delay in the flocking model with internal variables[J]. Networks and Heterogeneous Media, 2024, 19(3): 1182-1230. doi: 10.3934/nhm.2024052
[1] | Shenglun Yan, Wanqian Zhang, Weiyuan Zou . Multi-cluster flocking of the thermodynamic Cucker-Smale model with a unit-speed constraint under a singular kernel. Networks and Heterogeneous Media, 2024, 19(2): 547-568. doi: 10.3934/nhm.2024024 |
[2] | Hyunjin Ahn . Non-emergence of mono-cluster flocking and multi-cluster flocking of the thermodynamic Cucker–Smale model with a unit-speed constraint. Networks and Heterogeneous Media, 2023, 18(4): 1493-1527. doi: 10.3934/nhm.2023066 |
[3] | Hyunjin Ahn . Asymptotic flocking of the relativistic Cucker–Smale model with time delay. Networks and Heterogeneous Media, 2023, 18(1): 29-47. doi: 10.3934/nhm.2023002 |
[4] | Hyunjin Ahn, Se Eun Noh . Finite-in-time flocking of the thermodynamic Cucker–Smale model. Networks and Heterogeneous Media, 2024, 19(2): 526-546. doi: 10.3934/nhm.2024023 |
[5] | Hyeong-Ohk Bae, Seung Yeon Cho, Jane Yoo, Seok-Bae Yun . Effect of time delay on flocking dynamics. Networks and Heterogeneous Media, 2022, 17(5): 803-825. doi: 10.3934/nhm.2022027 |
[6] | Hyunjin Ahn . Uniform stability of the Cucker–Smale and thermodynamic Cucker–Smale ensembles with singular kernels. Networks and Heterogeneous Media, 2022, 17(5): 753-782. doi: 10.3934/nhm.2022025 |
[7] | Young-Pil Choi, Cristina Pignotti . Emergent behavior of Cucker-Smale model with normalized weights and distributed time delays. Networks and Heterogeneous Media, 2019, 14(4): 789-804. doi: 10.3934/nhm.2019032 |
[8] | J. J. P. Veerman, B. D. Stošić, A. Olvera . Spatial instabilities and size limitations of flocks. Networks and Heterogeneous Media, 2007, 2(4): 647-660. doi: 10.3934/nhm.2007.2.647 |
[9] | Yipeng Chen, Yicheng Liu, Xiao Wang . The critical delay of the consensus for a class of multi-agent system involving task strategies. Networks and Heterogeneous Media, 2023, 18(2): 513-531. doi: 10.3934/nhm.2023021 |
[10] | Don A. Jones, Hal L. Smith, Horst R. Thieme . Spread of viral infection of immobilized bacteria. Networks and Heterogeneous Media, 2013, 8(1): 327-342. doi: 10.3934/nhm.2013.8.327 |
We studied the dynamics of thermodynamic Cucker–Smale (TCS) particles moving with a constant speed constraint. The TCS model describes the collective dynamics of the population of birds with a time varying internal variable, and it was first introduced as the generalization of the Cucker–Smale (CS) model. In this paper, we considered a modification of the TCS model in which each agent moves at a constant speed, such as the Vicsek model, and we additionally considered the effect of time-delays due to the finiteness of the information propagation speed between agents. Then, we presented several sufficient conditions in terms of initial data and system parameters to exhibit asymptotic flocking. We presented two kinds of results for this purpose. One was an estimate of the diameter of the velocity and temperature configuration, and the other was an estimate of the diameter of the configuration within the time-delay bound τ.
Emergent behaviors of interacting multi-agent systems are often observed in our daily life. For example, aggregation of bacteria [34], flocking of migrating birds [17], schooling of moving fish [18,33], synchronization of fireflies and pacemaker cells [6,19,38], etc. For a brief introduction to this subject, we also refer to the articles and books [1,4,20,29,30,32,36,37]. In this work, we are interested in self-propelled flocking dynamics in which all agents move at a common velocity with limited surroundings and simple rules. After the seminal work done by Vicsek [35], many mathematical models describing flocking behavior have been widely studied in the mathematics community. Among them, the Cucker–Smale (CS) model [17] is one of the most successful models that represents flocking which has been studied from various perspectives, to name a few, local interactions [27], kinetic descriptions [7,24], hydrodynamic descriptions [21,26], stochastic descriptions [8], time-delay [10,13,14], etc.
In [23], Ha and Ruggeri proposed a CS-type flocking model with internal variable, which was called a temperature variable in their context. They considered the standard balance laws (mass, momentum, energy conservation) for the finitely many mixtures of spatially homogeneous ideal gases, together with specific constitutive equations suggested in [31] consistent with the entropy principle in thermodynamics. To fix the idea, let xi,vi,Ti be the position, velocity, and temperature of the i-th flocking agent, respectively. Then, the thermodynamic Cucker-Smale model (TCS in short) is given by the following second-order ordinary differential equations (ODEs) for position–velocity–temperature variables {(xi,vi,Ti)}Ni=1:
{dxidt=vi,t>0,i∈[N]:={1,…,N},(1.1a)dvidt=κ1NN∑j=1ϕ(‖xi−xj‖)(vjTj−viTi),(1.1b)ddt(Ti+12‖vi‖2)=κ2NN∑j=1ζ(‖xi−xj‖)(1Ti−1Tj),(1.1c)(xi(0),vi(0),Ti(0))=(x0i,v0i,T0i)∈Rd×Rd×(0,∞),(1.1d) |
where κ1 and κ2 are nonnegative scale parameters of the velocity and temperature communication weights, respectively. As in the CS model, the authors in [23] assumed that ϕ,ζ:[0,∞)→[0,∞) are nonnegative, locally Lipschitz, and monotonically decreasing, so that (1.1) is locally well-posed and the interaction becomes weaker as the distance between the two agents increases. However, it might be questionable whether the fact that
each bird′s motion depends on the temperature of the other birds |
is natural in terms of a flocking model. Therefore, when we think of the TCS-type models as the flocking models, we would like to suggest another interpretation for the temperature variable Ti. For instance, we may interpret 1/Ti as a time-varying measure of how attractive each bird is relative to other birds in the population. Then, each bird decides which direction to accelerate in proportion to the velocities and attractiveness of others, even for birds with the same velocity, and it becomes a matter of finding the initial conditions of all birds' velocities and attractiveness to converge to the same value. However, out of respect for the first authors' expressions, we decided to continue to refer to it as the TCS model throughout the paper.
Meanwhile, recent experiments on starling flocks [5,9] indicate that the speed fluctuations of birds are very small during their flights, demonstrating the need for us to study constant speed flocking models as [35]. Since the velocities of CS and TCS models converge to the same value under well-prepared initial conditions, the CS model and the TCS model can be said to have similar asymptotic behavior to their constant speed counterparts. Recently, [2] studied a new TCS-type flocking model in which the speed of each particle is constant. The main idea of [2] to create the new model was the same idea that led to the unit speed CS model [12] from the CS model: replace the right-hand side of (1.1b) by its orthogonal projection onto vi, i.e.,
replaceϕ(‖xi−xj‖)(vjTj−viTi)toϕ(‖xi−xj‖)Tj(vj−⟨vj,vi⟩‖vi‖2vi). |
Note that because it is an orthogonal projection to vi, the vi term disappears from the original TCS model, and the only interaction that appears to be asymmetric remains. As a result, the TCS model with a unit-speed constraint was given as
{dxidt=vi,t>0,i∈[N],dvidt=κ1NN∑j=1ϕ(‖xi−xj‖)Tj(vj−⟨vj,vi⟩‖vi‖2vi),dTidt=κ2NN∑j=1ζ(‖xi−xj‖)(1Ti−1Tj),(xi(0),vi(0),Ti(0))=(x0i,v0i,T0i)∈Rd×Sd−1×(0,∞), | (1.2) |
where the assumptions for κ1, κ2, ϕ, and ζ are the same as for the system (1.1), and Sd−1 is the (d−1)-dimensional unit sphere embedded in Rd, i.e.,
Sd−1:={x=(x1,…,xd)|d∑i=1|xi|2=1}. |
In addition to [2] and [12], there have been several mathematical studies on CS-type models with constant speed constraint. For example, [15] provided a sufficient condition to exhibit a bi-cluster flocking of unit speed CS ensemble, and [22] found a critical coupling strength to exhibit an asymptotic flocking for the same model. In particular, [11] conducted a study on a modified unit speed CS model which also considers the constant time-delay effects. Such time-delay effects are in fact prevalent in both physical and biological systems due to the fundamental constraint that information transmission cannot occur instantaneously during mutual communication processes [16,25]. This limitation on the speed of communication propagation has found significant application within the mathematical biology literature.
In this paper, we study the unit speed TCS model with nonconstant time-delays. More precisely, we define τji(t) as the time-delay for the i-th agent to detect a signal from the j-th agent at time t≥0, where we assume τii(t)=0 for all i∈[N] to avoid a self-processing time-delay. Furthermore, we assume τji(t):[0,∞)→[0,∞) nonnegative, continuous, and uniformly bounded by a constant τ for each pair (i,j)∈[N]2, i.e.,
τji(⋅)∈C([0,∞);[0,∞)),supt≥0maxi,j∈[N]τji(t)≤τ,τii(⋅)≡0,i,j∈[N]. |
Taking account these time-delay effects to Eq (1.2), we propose the following unit speed TCS model with time-delay (in short TCSUT):
{dxidt=vi,t>0,i∈[N],dvidt=κ1NN∑j=1ϕ(‖xi−xj(t−τji(t))‖)Tj(t−τji(t))(vj(t−τji(t))−⟨vj(t−τji(t)),vi⟩vi‖vi‖2),dTidt=κ2NN∑j=1ζ(‖xi−xj(t−τji(t))‖)(1Ti−1Tj(t−τji(t))),(xi(t),vi(t),Ti(t))=(x0i(t),v0i(t),T0i(t))∈Rd×Sd−1×(0,∞),t∈[−τ,0]. | (1.3) |
For this flocking model, our main concern in this paper is to find sufficient conditions on the communication weights and initial data that guarantee the asymptotic flocking to occur. To do this, we first define the asymptotic flocking phenomenon rigorously.
Definition 1.1. (Asymptotic flocking) Let Z={(xi,vi,Ti)}Ni=1 be a solution to the system (1.3). The configuration Z exhibits asymptotic flocking if
supt≥0maxi,j∈[N]‖xi(t)−xj(t)‖<∞,limt→∞maxi,j∈[N]‖vj(t)−vi(t)‖=0,limt→∞maxi,j∈[N]|Tj(t)−Ti(t)|=0. |
Note that the nonconstant time-delay in (1.3) means that each i-th bird receives information about the other birds at different points in time to determine its velocity and temperature. Therefore, if we want to analyze the diameters
DX(t):=maxi,j∈[N]‖xi(t)−xj(t)‖,DV(t):=maxi,j∈[N]‖vi(t)−vj(t)‖,DT(t):=maxi,j∈[N]|Ti(t)−Tj(t)| | (1.4) |
to show the asymptotic flocking in Definition 1.1, there are some additional technical difficulties compared to the constant time-delay models. Since it may not be enough to simply use the diameters in (1.4) to fully control the right-hand side of (1.3), we need to measure the upper bound of the error that each velocity and temperature can have due to the time-delay effect. For simplicity, we set
X:=(x1,…,xN),V:=(v1,…,vN),T:=(T1,…,TN), |
and for every Z∈{X,V,T}, we define the delayed diameter DτZ and perturbation ΔτZ as
DτZ(t):=maxs∈[t−τ,t]maxi,j∈[N]‖zi(s)−zj(t)‖,ΔτZ(t):=maxs∈[t−τ,t]maxi∈[N]‖zi(s)−zi(t)‖, | (1.5) |
which satisfy
DZ(t)≤DτZ(t),t>0,Z∈{X,V,T}. |
Then, the right-hand side of (1.3) can be controlled by using Eq (1.5), since
‖vi(t)−vj(t−τji(t))‖≤DτV(t)≤DV(t)+ΔτV(t),t>0,i,j∈[N],|Ti(t)−Tj(t−τji(t))|≤DτT(t)≤DT(t)+ΔτT(t),t>0,i,j∈[N], |
and the main result of this paper is to find sufficient conditions for DX to be uniformly bounded and DV,DT to converge to zero.
To this end, we will consider two different approaches: constructing a differential inequality for diameters (DX,DV,DT) and constructing a differential inequality for delayed diameters (DτX,DτV,DτT). The analysis of diameters is similar to the method used in [3] and other related works, but the analysis of delayed diameter is, to the best of the authors' knowledge, the first attempt on studies of flocking models with time-delays.
The rest of this paper is organized as follows. In Section 2, we provide some basic properties and previous flocking estimate for the unit speed TCS model (1.2) without time-delay. In Section 3, we present several preparatory lemmas which are crucially used to guarantee the global well-posedness of the system (1.3) and also used to construct differential inequalities for the delayed diameters in Section 4. In Section 4, we derive a system of differential inequalities for the delayed-diameters (DτX,DτV,DτT) and present the first main result. In Section 5, we present a system of differential inequalities for the diameters (DX,DV,DT) and the second main result. We also present several numerical experiments in Section 6 to demonstrate our theoretical results. Finally, Section 7 is devoted to summarizing the main results of this paper and some discussion on the remaining issues to be investigated in a future work. In Appendix A and B, we provide the detailed proof of Lemma 3.1 and Lemma 4.1, respectively.
Notation: Throughout this paper, we employ the following notation for simplicity.
‖⋅‖:=standard l_2 -norm,⟨⋅,⋅⟩=standard inner product,[N]:={1,…,N},zτjij(t):=zj(t−τji(t))forZ=(z1,…,zN)∈{X,V,T},t≥0,ψτjiji(t):=ψ(‖xτjij(t)−xi(t)‖)forψ∈{ϕ,ζ},t≥0,Tm=mini∈[N]Ti,TM=maxi∈[N]Ti,Tτm(t)=mins∈[t−τ,t]Tm(s),TτM(t)=maxs∈[t−τ,t]TM(s). |
In this section, we briefly review several basic properties and previous results of the TCS model with a unit speed constraint obtained in [2]. Readers with sufficient background knowledge in this subject may skip this section.
In this subsection, we introduce the unit speed TCS model in the absence of time-delay. Recall that the unit speed TCS model was given as
{dxidt=vi,t>0,i∈[N],dvidt=κ1NN∑j=1ϕ(‖xi−xj‖)Tj(vj−⟨vj,vi⟩‖vi‖2vi),dTidt=κ2NN∑j=1ζ(‖xi−xj‖)(1Ti−1Tj),(xi(0),vi(0),Ti(0))=(x0i,v0i,T0i)∈Rd×Sd−1×(0,∞). | (2.1) |
Even if we consider Ti as an internal variable other than temperature, the 1st and 2nd laws of thermodynamics still hold for the original TCS model, since it is derived from the standard balance laws for spatially homogeneous gas mixtures. Similarly, in the model with the constant speed condition added, the law of energy conservation and entropy principle hold for Ti's, which are as follows.
Proposition 2.1. Let (X,V,T) be a solution to the system (2.1). Then, the following assertions hold:
1. (Energy conservation) The total internal energy ∑Ni=1Ti is conserved.
2. (Entropy principle) The total entropy S:=∑Ni=1lnTi is monotonically increasing.
However, these laws can be considered only as long as each Ti exists as a positive real number, since the solution of (2.1) blows up if any Ti approaches zero. Therefore, in order to see how the solution of the equation behaves for t→∞, it must be ensured that all Ti's cannot be smaller than some positive real number, which can be seen in the following proposition.
Proposition 2.2. Let {(x0i,v0i,T0i)}Ni=1⊂Rd×Sd−1×(0,∞) be an arbitrary initial configuration. Then, the equation (2.1) admits a unique global solution on t>0 and
0<Tm(0)≤Tm(t)≤Ti(t)≤TM(t)≤TM(0),∀i∈[N],t≥0. |
What Proposition 2.2 means is that we can always be guaranteed that the extreme values of the temperatures will approach each other, even though they may not actually converge to the same value. Therefore, a natural question to ask is whether we can expect a similar behavior for velocity vectors, and the answer to this question can be found in the following proposition.
Proposition 2.3. Let (X,V,T) be a solution to (2.1) subject to initial data {(x0i,v0i,T0i)}Ni=1 satisfying
mini,j∈[N]⟨v0i,v0j⟩>0. |
Then, the function
A(t):=mini,j∈[N]⟨vi(t),vj(t)⟩=1−12DV(t)2 |
is monotonically increasing in t.
In this subsection, we review previous flocking estimates for (2.1) obtained from [2] and discuss whether they can be improved. We begin with a system of differential inequalities in terms of DX,DV and DT.
Proposition 2.4. (Differential inequalities for diameters) Let (X,V,T) be a solution to the system (2.1) subject to initial data {(x0i,v0i,T0i)}Ni=1 satisfying A(0)>0. Then, we have
|dDXdt|≤DV,dDVdt≤−κ1A(0)ϕ(DX)TM(0)DV,dDTdt≤−κ2ζ(DX)TM(0)2DT. |
In particular, the function
L±(DX,DV):=DV±κ1A(0)TM(0)Φ(DX) |
is monotonically decreasing with respect to t, where
Φ(a):=∫a0ϕ(s)ds,∀a≥0. |
Since L(DX,DV) is monotonically decreasing, Φ(DX) has a uniform in time upper bound determined by the initial data. Therefore, if this upper bound is smaller than Φ(∞)∈[0,∞], we know that the diameter DX is uniformly bounded in time. Then, Proposition 2.4 determines the exponential decay rate of the diameters DV and DT.
Proposition 2.5. (Asymptotic flocking) Let (X,V,T) be a solution to the system (2.1) subject to initial data {(x0i,v0i,T0i)}Ni=1 satisfying
0≤TM(0)κ1∫∞DX(0)ϕ(s)ds<A(0)DV(0)=1−12DV(0)2DV(0). | (2.2) |
Then, we have
DX(t)≤D∞X,DV(t)≤DV(0)exp(−κ1A(0)ϕ(D∞X)TM(0)t),DT(t)≤DT(0)exp(−κ2ζ(D∞X)TM(0)2t), |
where D∞X is the unique positive number satisfying
TM(0)κ1∫D∞XDX(0)ϕ(s)ds=1−12DV(0)2DV(0). |
Remark 2.1. We make a few remarks about the optimality of Proposition 2.5.
(1) In [2], the author provided two sufficient frameworks for the asymptotic flocking of (2.1). One is the result stated in Proposition 2.5, which uses the Lyapunov functional L+, and there was another result using 'the bootstrapping argument' according to their context. However, with more precise calculations we can show that the result is weaker under stronger conditions than Proposition 2.5.
(2) In fact, according to the proof of Proposition 2.4, one can obtain
dDV(t)dt≤−κ1A(t)ϕ(DX(t))TM(0)DV(t)=−κ1ϕ(DX(t))TM(0)(1−12D2V(t))DV(t). |
Then, the function
L(DX,DV):=1√2ln(√2+DV√2−DV)+κ1Φ(DX)TM(0) |
is monotonically decreasing with respect to t, and we have
DV(t)22−DV(t)2≤DV(0)22−DV(0)2exp(−2κ1ϕ(D∞X)TM(0)t), |
whenever the initial data satisfies
0≤∫∞DX(0)ϕ(s)ds>TM(0)√2κ1ln(√2+DV(0)√2−DV(0)). |
In Section 3, we will show that the generalizations of Proposition 2.2 and Proposition 2.3 hold in the time-delayed model (1.3). These will be crucially used to prove the well-posedness of (1.3) and to find some flocking estimates corresponding to Proposition 2.5.
In this section, we present the global well-posedness and some basic properties of the system (1.3). To fix the idea, consider a following time-delayed ODE:
dx(t)dt=f(x(t),x(t−τ(t))),τ(⋅)∈C(R;[0,τ]). |
Then, x(t−τ(t)) does not play a crucial role to the well-posedness of the solution. Instead, it is more appropriate to view it as an independently given function that does not depend on the value of x(t). Therefore, the local well-posedness of the ODE can be obtained when f(y,z) is Lipschitz in y, continuous in z, and t↦x(t−τ(t)) is continuous. Because of the above reasons, it is clear that the global well-posedness of the system (1.3) can be obtained by the invariance of ‖vi‖ and the boundedness of 1/Ti for each i∈[N], and the invariance of each ‖vi‖ is given by the relation
12d‖vi‖2dt=κ1NN∑j=1ϕτjijiTτjij⟨vτjij−⟨vτjij,vi⟩vi‖vi‖2,vi⟩=0,∀i∈[N]. |
Then, due to the uniqueness of the solution and the unit-speed constraint for initial data, we can rewrite the system (1.3) as the following simplified form:
{dxidt=vi,t>0,i∈[N],dvidt=κ1NN∑j=1ϕτjijiTτjij(vτjij−⟨vτjij,vi⟩vi),dTidt=κ2NN∑j=1ζτjiji(1Ti−1Tτjij),(xi(t),vi(t),Ti(t))=(x0i(t),v0i(t),T0i(t))∈Rd×Sd−1×(0,∞),∀t∈[−τ,0]. | (3.1) |
In [10], the authors proved the monotonicity of the maximum and minimum temperatures for the TCS model with time-delay, which does not have the invariance of speeds as Eq (3.1). Although not rigorously proved in previous works, their basic idea was to use the following lemma to the minimum and maximum temperatures. We here present a rigorous proof for the completeness of this paper and for the use of this lemma in future work.
Lemma 3.1. Let f∈C(R;R), and define F=T[f]:R→R as
T[f](t):=mins∈[t−τ,t]f(s),∀t∈R. |
Then, the following assertions hold:
1. If f(t0)>F(t0) for some t0∈R, there exists δ>0 such that
F(t0)≤F(t),∀t∈[t0,t0+δ]. |
2. F is monotonically increasing if D+f(t∗)≥0 for all t∗∈R satisfying F(t∗)=f(t∗).
3. If f∈LipL(R,R) for some L>0, then F∈LipL(R,R).
Proof. Since the proof is lengthy and technical, we leave it to Appendix A.
Remark 3.1. To preempt some possible misconceptions about the relation between f and T[f], we present the following examples.
1. Convergence of F=T[f] does not imply the convergence of f. For example, if f(t)=cos(2πtτ), the function F=T[f] is a constant function F≡−1. In fact, if F is monotonically increasing, we have
limt→∞F(t)=limt→∞mint≤sF(s)=limt→∞inft−τ≤sf(s)=lim inft→∞f(t). |
2. There may not be a 'first point' that does not satisfy the condition in Lemma 3.1 (2). For example, if f(t)=−max{0,t}2, the function f is monotonically decreasing. Then, T[f]≡f and f′(t)=−2max{0,t}≥0 for only t≤0, and therefore
F(t)=f(t)for allt∈R,butD+f(t)≥0for onlyt≤0. |
Using Lemma 3.1, one can prove monotonic increase in minimum temperature and monotonic decrease in maximum temperature, similar to the original models without delay [2,23]. More precisely, we can prove that Tτm is monotonically increasing and TτM is monotonically decreasing when the initial values of both functions are positive real numbers.
Lemma 3.2. (Monotonicity of Tτm and TτM) Let (X,V,T) be a solution to Eq (3.1), where the initial data
(x0i,v0i,T0i):[−τ,0]→Rd×Sd−1×(0,∞) |
is continuous for every i∈[N]. Then, Tτm and −TτM are monotonically increasing, i.e.,
0<Tτm(0)≤Tτm(t)≤Ti(t)≤TτM(t)≤TτM(0),∀t≥0,i∈[N]. |
Proof. Let us denote by
f(t):=minj∈[N]Tj(t),F:=T[f]. |
Then, the constraint for the initial data can be written as F(0)>0, and we want to prove F(t)≥F(0) for every t≥0. If ε>0 and t0 is the first time satisfying F(t)=F(0)−ε>0, then whenever F(t∗)=f(t∗) in t∗∈[0,t0], we can consider the maximal index set I⊂[N] such that
F(t∗)=f(t∗)=Tj(t∗),∀j∈I. |
Then, since each Tk(t) is differentiable in t, we have
Tk(t)=Tk(t∗)+dTkdt(t∗)(t−t∗)+o(|t−t∗|),∀k∈[N], |
and by taking the minimum in k, one can obtain
f(t)=f(t∗)+minj∈IdTjdt(t∗)(t−t∗)+o(|t−t∗|),∀0<t−t∗≪1. |
Therefore, the Dini derivative D+f at t=t∗ satisfies
D+f(t∗)=minj∈IdTjdt(t∗)=minj∈Iκ2NN∑k=1ζτkjkj(1Tj(t∗)−1Tτkjk(t∗))=minj∈Iκ2NN∑k=1ζτkjkj(1F(t∗)−1Tτkjk(t∗))≥0, |
where we used Tτkjk(t∗)≥F(t∗)>F(0)−ε>0 in the last inequality. By using Lemma 3.1, the function F is therefore monotonically increasing in [0,t0], which contradicts to F(t0)=F(0)−ε<F(0). Therefore, we have
0<Tτm(0)=F(0)≤F(t)=Tτm(t)≤Ti(t),∀t≥0,i∈[N]. |
In addition, by letting each s≥0 be a new starting time, this also shows that F is monotonically increasing in [0,∞):
0<Tτm(s)=F(s)≤F(t)=Tτm(t)≤Ti(t),∀t≥s≥0,i∈[N]. |
We can also obtain the monotonic decreasing property of TτM by choosing f=−TM and F=T[f]=−TτM.
What we want to emphasize in this proof is that we assumed the first point t0 where the value of F is less than or equal to F(0)−ε and then applied Lemma 3.1 on the interval [0,t0] to prove its contradiction. This is because, as we pointed out in Remark 3.1. (2), there may not be the first point at which the condition of Lemma 3.1(2) does not hold.
Then, by applying the classical Cauchy-Lipschitz theory, the global existence of the solution to (3.1) can be guaranteed from the invariance of the speed and the existence of positive lower bound for temperatures. However, the use of Lemma 3.1 goes beyond simply showing the monotonicity of extreme temperatures. To be specific, every open hemisphere U⊂Sd−1 contains all v1(t),…,vN(t) uniformly in t≥0, whenever U contains all initial velocities.
Lemma 3.3. Let (X,V,T) be a solution to Eq (3.1), where the initial data
(x0i,v0i,T0i):[−τ,0]→Rd×Sd−1×(0,∞) |
is continuous for every i∈[N]. If initial data satisfies
mins∈[−τ,0]minj∈[N]⟨u,vj(s)⟩>0 | (3.2) |
for some u∈Sd−1, the function
t↦mins∈[t−τ,t]minj∈[N]⟨u,vj(s)⟩ |
is monotonically increasing.
Proof. We can employ a similar argument to Lemma 3.2. In this case, we set
f(t):=minj∈[N]⟨u,vj(t)⟩,F:=T[f]. |
If ε>0 and t0 is the first time satisfying F(t)=F(0)−ε>0, then whenever F(t∗)=f(t∗) in t∗∈[0,t0], we can consider the maximal index set I⊂[N] such that
F(t∗)=f(t∗)=⟨u,vj(t∗)⟩,∀j∈I. |
Then, the Dini derivative D+f at t=t∗ satisfies
D+f(t∗)=minj∈I⟨dvjdt(t∗),u⟩=minj∈Iκ1NN∑k=1ϕτkjkjTτkjk⟨vτkjk(t∗)−⟨vτkjk(t∗),vj(t∗)⟩vj(t∗),u⟩≥minj∈Iκ1NN∑k=1ϕτkjkjTτkjkF(t∗)(1−⟨vτkjk(t∗),vj(t∗)⟩)≥0, |
where we used ⟨u,vτkjk(t∗)⟩≥F(t∗)≥0 in the second inequality. By using Lemma 3.1, the function F is monotonically increasing in [0,t0], which contradicts to F(t0)=F(0)−ε<F(0). Therefore, we have F(t)≥F(0) for all t≥0, and by letting each s≥0 be a new starting time, this also shows that F is monotonically increasing in [0,∞).
What should be noted here is that Lemma 3.3 alone cannot prove the desired flocking phenomenon, since the unit vector u∈Sd−1 satisfying (3.2) is not unique if it exists. Instead, it is better to use an indicator that can show that individual velocities are getting closer to each other. The most intuitive way to achieve this is to assume that all initial velocities can be used as the vector 'u' in Eq (3.2), so that all velocities approach each other. Under such initial conditions, the lemma below implies that the inner product ⟨vi(t1),vj(t2)⟩ is strictly positive for all i,j∈[N] and t1,t2≥−τ.
Lemma 3.4. Let (X,V,T) be a solution to (3.1), where the initial data
(x0i,v0i,T0i):[−τ,0]→Rd×Sd−1×(0,∞) |
is continuous for every i∈[N]. If the initial data satisfies
mins1∈[−τ,0]s2∈[−τ,0]mini,j∈[N]⟨vi(s1),vj(s2)⟩>0, | (3.3) |
then the function
Aτ,τ(t1,t2):=mins1∈[t1−τ,t1]s2∈[t2−τ,t2]mini,j∈[N]⟨vi(s1),vj(s2)⟩ |
is monotonically increasing in both t1 and t2.
Proof. Since Aτ,τ(t1,t2)=Aτ,τ(t2,t1) for all t1,t2≥0, we only need to show the monotonic increasing property in one variable, which is sufficient to verify
Aτ,τ(0,t2)≥Aτ,τ(0,0),∀t2≥0, |
by taking each positive number as a new starting time as in Lemma 3.2 and Lemma 3.3. For every s1∈[−τ,0] and i∈[N], we apply Lemma 3.3 to u=vi(s1) and obtain
t↦mins2∈[t−τ,t]minj∈[N]⟨vi(s1),vj(s2)⟩ |
is monotonically increasing. Therefore, we have
Aτ,τ(0,t2)=mins1∈[−τ,0]mini∈[N]mins2∈[t2−τ,t2]minj∈[N]⟨vi(s1),vj(s2)⟩≥mins1∈[−τ,0]mini∈[N]mins2∈[−τ,0]minj∈[N]⟨vi(s1),vj(s2)⟩=Aτ,τ(0,0), |
which is the desired result.
Before we close this section, we provide several estimates which will play crucial roles in Section 4. Recall that we defined the diameter and perturbation functions affected by the time-delay as follows: for Z=(z1,…,zN)∈{X,V,T},
DτZ(t):=maxs∈[t−τ,t]maxi,j∈[N]‖zi(s)−zj(t)‖,ΔτZ(t):=maxs∈[t−τ,t]maxi∈[N]‖zi(s)−zi(t)‖. |
Consequently, one can easily obtain ΔτZ≤DτZ by definition, and Lemma 3.4 yields
DτV(t)2=2−2mins∈[t−τ,t]mini,j∈[N]⟨vi(s),vj(t)⟩≤2−2Aτ,τ(t,t)≤2−2Aτ,τ(0,0), |
provided that Eq (3.3) holds, i.e., Aτ,τ(0,0)>0. In addition, we can estimate perturbation functions ΔτV(t) and ΔτT(t) by using the integration of ‖˙zi‖ in [t−τ,t].
Lemma 3.5. Let (X,V,T) be a solution to (3.1), where the initial data
(x0i,v0i,T0i):[−τ,0]→Rd×Sd−1×(0,∞) |
is Lipschitz continuous for every i∈[N]. Then, for every t≥0, we have
ΔτV(t)≤max{τ−t,0}maxi∈[N]‖v0i‖Lip+(N−1)κ1ϕ(0)NTτm(0)∫tmax{t−τ,0}DτV(s)ds,ΔτT(t)≤max{τ−t,0}maxi∈[N]‖T0i‖Lip+(N−1)κ2ζ(0)NTτm(0)2∫tmax{t−τ,0}DτT(s)ds. |
Proof. For given Z∈{V,T} and t≥0, suppose we have
ΔτZ(t)=‖zi(s)−zi(t)‖,s∈[t−τ,t]. |
Then, by using the triangle inequality, one has
‖zi(s)−zi(t)‖≤∫max{s,0}s‖˙zi(u)‖du+∫tmax{s,0}‖˙zi(u)‖du≤max{τ−t,0}‖z0i‖Lip+∫tmax{t−τ,0}‖˙zi(u)‖du. |
Below, we estimate the upper bound of ‖˙zi(u)‖ in u≥0 one by one. For the velocity derivative, we have
‖dvi(t)dt‖=‖κ1NN∑j=1ϕτjijiTτjij(vτjij−⟨vτjij,vi⟩vi)‖≤κ1ϕ(0)NTτm(0)∑j∈[N]j≠i‖vτjij−⟨vτjij,vi⟩vi‖=κ1ϕ(0)NTτm(0)∑j∈[N]j≠i√1−⟨vτjij,vi⟩2≤κ1ϕ(0)NTτm(0)∑j∈[N]j≠i√2−2⟨vτjij,vi⟩≤(N−1)κ1ϕ(0)NTτm(0)DτV(t),∀t≥0, |
where we used
ϕ(‖xi(t)−xj(t−τji(t))‖)≤ϕ(0),Tj(t−τji(t))≥Tτm(0),2−2⟨vj(t−τji(t)),vi(t)⟩=‖vj(t−τji(t))−vi(t)‖2≤DτV(t)2, |
in the first and the last inequality, respectively. Similarly, for the temperature derivative, we apply Lemma 3.2 to Eq (3.1) to obtain
|dTi(t)dt|=κ2N∑j∈[N]j≠iζ(‖xi(t)−xj(t−τji(t))‖)|Ti(t)−Tj(t−τji(t))Ti(t)Tj(t−τji(t))|≤(N−1)κ2ζ(0)NTτm(0)2DτT(t),∀t≥0, |
where we used
ζ(‖xi(t)−xj(t−τji(t))‖)≤ζ(0),Ti(t)Tj(t−τji(t))≥Tτm(0)2,|Ti(t)−Tj(t−τji(t))|≤DτT(t),∀t≥0, |
in the last inequality.
Remark 3.2. Since ˙z0i(s) is not given in the differential equation (3.1), we need to use the Lipschitz constant of the initial data {z0i}Ni=1 to evaluate ΔτZ(t) in t∈[0,τ). In addition, the integration of ‖˙vi(u)‖ in u∈[s,t] is greater than or equal to the 'geodesic distance' between vi(s) and vi(t), since velocities are moving on the unit sphere Sd−1. Therefore, it is possible to modify Lemma 3.5 to
cos−1(1−12ΔτV(t)2)≤(N−1)κ1ϕ(0)NTτm(0)∫tt−τDτV(s)ds,t≥τ. |
Finally, the last thing we need to check is that the diameter functions DτZ are absolutely continuous, so it satisfies the fundamental theorem of calculus. If we verify this, we can use the system of differential inequalities for DτX,DτV, and DτT, which we will prove in the next section, to find sufficient framework that guarantees the desired flocking phenomenon.
Lemma 3.6. Let (X,V,T) be a solution to (3.1), where the initial data
(x0i,v0i,T0i):[−τ,0]→Rd×Sd−1×(0,∞) |
is Lipschitz continuous for every i∈[N]. Then, DτX,DτV, and DτT are globally Lipschitz in t≥0.
Proof. By using the uniform boundedness of dzidt obtained from Lemma 3.2, one can see that
s1↦−‖zi(s1)−w‖ |
is globally Lipschitz. Since the minimum of finitely many Lipschitz continuous functions is Lipschitz by the relation
min{f1,f2}=f1+f2−|f1−f2|2, |
we have the global Lipschitz continuity of the function
s↦−maxi∈[N]‖zi(s)−w‖, |
for every fixed w. In addition, by using Lemma 3.1(3), one can also obtain the Lipschitz continuity of
t↦maxs∈[t−τ,t]maxi∈[N]‖zi(s)−w‖. |
Now, let us denote a nonempty, compact set CZ(t) by
CZ(t):={zi(s):s∈[t−τ,t],i∈[N]}. |
Then, the function
t↦maxz∈CZ(t)‖z−w‖ |
is globally Lipschitz, and the function
s2↦maxz∈CZ(t1)‖z−zj(s2)‖ |
is also globally Lipschitz since zj(s2) is Lipschitz in s2 and the Lipschitz constant of the function w↦maxz∈CZ(t)‖z−w‖ is not greater than 1. Therefore, all delayed diameter functions are globally Lipschitz, i.e.,
DτZ(t)=maxj∈[N]maxz∈CZ(t)‖z−zj(t)‖ |
is globally Lipschitz in t.
In this section, we present a system of differential inequalities on delayed diameters DτX,DτV, and DτT to deduce suitable sufficient frameworks for the asymptotic flocking of the system (3.1). Unlike the commonly used diameter DZ(t)=maxi,j∈[N]‖zi(t)−zj(t)‖, each delayed diameter does not specify at which point in time the distance between two vectors are evaluated. Therefore, in order to estimate the derivative of the delayed diameter, we need to know the behavior of the function given by a maximum value of a differentiable function over a certain range.
Lemma 4.1. Let f:R×R→R be a continuous function, and define
S(t):=[t−τ,t]×{t}∪{t}×[t−τ,t],m[f](t):=max(s1,s2)∈S(t)f(s1,s2),t∈R. |
If there is a continuous function λ:R×R→R and t0∈R such that
lim suph→0+f(t1+h,t2+h)−f(t1,t2)h≤λ(t1,t2),∀(t1,t2)∈R2,λ(t1,t2)≤cwheneverm[f](t0)=f(t1,t2)and(t1,t2)∈S(t0), |
we have
D+m[f](t0)≤c. |
Proof. Basically, the main idea to prove this lemma is finding a uniform upper bound of the Dini derivative of h↦f(t1+h,t2+h) for all points (t1,t2) which maximizes f on S(t). However, since we are considering the values of f on an infinite set S(t) and analyzing the temporal evolution of their maximum, we want the Dini derivative to be uniformly bounded by a reasonable value in the neighborhood of every maximum point in S(t). This is why we required a continuous upper bound λ, and the technical difficulties of the proof can be resolved with some simple preparation and with the help of the Berge maximum theorem (see [28] for details). We leave the detailed proof to Appendix B.
Now, we present the differential inequalities of the delayed diameters. For every Z∈{X,V}, DτZ(t)2 can be represented as the maximum value of
S(t)∋(t1,t2)↦fZ(t1,t2):=maxi,j∈[N]‖zi(t1)−zj(t2)‖2. |
Once we find a continuous upper bound λ of the Dini derivative of the function fZ in the (1,1) direction, we can use the maximum of λ over the maximizing set of fZ as the upper bound of the Dini derivative of (DτZ)2. Therefore, by expressing the value of the upper bound λ in terms of the value of fZ, we can obtain a differential inequality for the delayed diameters.
Lemma 4.2. Let (X,V,T) be a solution to (3.1), where the initial data
(x0i,v0i,T0i):[−τ,0]→Rd×Sd−1×(0,∞) |
are Lipschitz continuous functions satisfying dx0idt=v0i for all i∈[N] and
Aτ,τ(0,0)=cosδ,δ∈(0,π2). |
In addition, we define a continuous function SτX as
SτX(t)=DτX(0)+∫t0DτV(s)ds,t≥0. |
Then, the following differential inequalities hold:
1. dDτX(t)dt≤DτV(t),a.e.t>0.
2. dDτV(t)dt≤−κ1ϕ(SτX(t))cosδTτM(0)DτV(t)+3κ1ϕ(0)Tτm(0)maxs∈[t−τ,t]ΔτV(s),a.e.t>τ.
3. dDτT(t)dt≤−κ2ζ(SτX(t))TτM(0)2DτT(t)+3κ2ζ(0)Tτm(0)2maxs∈[t−τ,t]ΔτT(s),a.e.t>τ.
Proof. (1) For every i,j∈N and t1,t2≥0, define
fij(t1,t2):=‖xi(t1)−xj(t2)‖2. |
Then, for the continuous function fX:=maxi,j∈[N]fij,
lim suph→0+fX(t1+h,t2+h)−fX(t1,t2)h≤maxi,j∈[N]limh→0+fij(t1+h,t2+h)−fij(t1,t2)h=maxi,j∈[N]2⟨xi(t1)−xj(t2),vi(t1)−vj(t2)⟩=:λ(t1,t2). |
In particular, for every (t1,t2)∈S(t), we have
λ(t1,t2)≤2DτX(t)DτV(t). |
Therefore, by using Lemma 4.1, one can obtain
D+[(DτX)2](t)=D+m[fX](t)≤2DτX(t)DτV(t), |
which implies
dDτX(t)dt≤DτV(t) |
whenever DτX is differentiable at t. Finally, from the Lipschitz continuity of DτX in Lemma 3.6, we have the desired result.
(2) For the velocity diameter DτV, we set
fij(t1,t2)=‖vi(t1)−vj(t2)‖2,fV=maxi,j∈[N]fij. |
Then, we have
limh→0+fij(t1+h,t2+h)−fij(t1,t2)h=2⟨vi(t1)−vj(t2),˙vi(t1)−˙vj(t2)⟩=−2⟨vi(t1),˙vj(t2)⟩−2⟨vj(t2),˙vi(t1)⟩=−2κ1N∑k∈[N]ϕτkjkj(t2)Tτkjk(t2)⟨vi(t1),vτkjk(t2)−⟨vτkjk(t2),vj(t2)⟩vj(t2)⟩−2κ1N∑k∈[N]ϕτkiki(t1)Tτkik(t1)⟨vj(t2),vτkik(t1)−⟨vτkik(t1),vi(t1)⟩vi(t1)⟩=:−2κ1N∑k∈[N]ϕτkjkj(t2)Tτkjk(t2)Iijk(t1,t2)−2κ1N∑k∈[N]ϕτkiki(t1)Tτkik(t1)Ijik(t2,t1), |
where each Iijk(t1,t2) satisfies
Iijk(t1,t2)=⟨vi(t1),vτkjk(t2)−⟨vτkjk(t2),vj(t2)⟩vj(t2)⟩=⟨vi(t1)−vj(t2),vτkjk(t2)−vk(t2)−⟨vτkjk(t2)−vk(t2),vj(t2)⟩vj(t2)⟩+⟨vi(t1),vk(t2)−⟨vk(t2),vj(t2)⟩vj(t2)⟩≥−‖vi(t1)−vj(t2)‖ΔτV(t2)+⟨vi(t1),vk(t2)−⟨vk(t2),vj(t2)⟩vj(t2)⟩. |
In particular, if fij(t1,t2)=fV(t1,t2) for some i,j∈[N] and t1,t2≥0,
limh→0+fij(t1+h,t2+h)−fij(t1,t2)h=−2κ1N∑k∈[N]ϕτkjkj(t2)Tτkjk(t2)Iijk(t1,t2)−2κ1N∑k∈[N]ϕτkiki(t1)Tτkik(t1)Ijik(t2,t1)≤(2κ1ϕ(0)Tτm(t2)ΔτV(t2)+2κ1ϕ(0)Tτm(t1)ΔτV(t1))‖vi(t1)−vj(t2)‖−2κ1Nϕ(DτX(t2))TτM(t2)∑k∈[N]⟨vi(t1),vk(t2)−⟨vk(t2),vj(t2)⟩vj(t2)⟩−2κ1Nϕ(DτX(t1))TτM(t1)∑k∈[N]⟨vj(t2),vk(t1)−⟨vk(t1),vi(t1)⟩vi(t1)⟩, |
where we used ⟨vi(t1),vj(t2)⟩≥Aτ,τ(t1,t2)>0 and
‖xi(t1)−xτkik(t1)‖≤DτX(t1),‖xj(t2)−xτkjk(t2)‖≤DτX(t2),⟨vi(t1),vk(t2)−⟨vk(t2),vj(t2)⟩vj(t2)⟩≥⟨vi(t1),vj(t2)⟩(1−⟨vk(t2),vj(t2)⟩)≥0,⟨vj(t2),vk(t1)−⟨vk(t1),vi(t1)⟩vi(t1)⟩≥⟨vi(t1),vj(t2)⟩(1−⟨vk(t1),vi(t1)⟩)≥0, |
in the last inequality. In addition, by using the inequality
⟨vi(t1),vk(t2)−⟨vk(t2),vj(t2)⟩vj(t2)⟩+⟨vj(t2),vk(t1)−⟨vk(t1),vi(t1)⟩vi(t1)⟩=(⟨vi(t1),vk(t2)⟩+⟨vk(t1),vj(t2)⟩)(1−⟨vi(t1),vj(t2)⟩)−⟨vk(t1)−vk(t2),vi(t1)−vj(t2)⟩⟨vi(t1),vj(t2)⟩≥⟨vi(t1),vj(t2)⟩‖vi(t1)−vj(t2)‖(‖vi(t1)−vj(t2)‖−‖vk(t2)−vk(t1)‖), |
we have
limh→0+fV(t1+h,t2+h)−fV(t1,t2)h≤(2κ1ϕ(0)Tτm(t2)ΔτV(t2)+2κ1ϕ(0)Tτm(t1)ΔτV(t1))√fV(t1,t2)−2κ1min{ϕ(DτX(t1))TτM(t1),ϕ(DτX(t2))TτM(t2)}×(1−12fV(t1,t2))√fV(t1,t2)(√fV(t1,t2)−maxk∈[N]‖vk(t2)−vk(t1)‖)=:λ(t1,t2). |
Therefore, by using Lemma 4.1, one can obtain
D+[(DτV)2](t)=D+m[fV](t)≤(2κ1ϕ(0)Tτm(0)ΔτV(t)+2κ1ϕ(0)Tτm(0)maxs∈[t−τ,t]ΔτV(s))DτV(t)−2κ1ϕ(SτX(t))TτM(0)(1−12DτV(t)2)DτV(t)(DτV(t)−ΔτV(t)), |
which implies the following inequalities whenever DτV is differentiable at t:
dDτV(t)dt≤(κ1ϕ(0)Tτm(0)ΔτV(t)+κ1ϕ(0)Tτm(0)maxs∈[t−τ,t]ΔτV(s))−κ1ϕ(SτX(t))TτM(0)(1−12DτV(t)2)(DτV(t)−ΔτV(t))≤−κ1ϕ(SτX(t))TτM(0)(1−12DτV(t)2)DτV(t)+3κ1ϕ(0)Tτm(0)maxs∈[t−τ,t]ΔτV(s),≤−κ1ϕ(SτX(t))cosδTτM(0)DτV(t)+3κ1ϕ(0)Tτm(0)maxs∈[t−τ,t]ΔτV(s). |
Similar to the case of DτX, we reach the desired result due to the Lipschitz continuity of Lemma 3.6.
(3) In this case, we define
fij(t1,t2)=Ti(t1)−Tj(t2),fT=maxi,j∈[N]fij. |
Then, whenever fij(t1,t2)=fT(t1,t2) and t1,t2≤t, we get
limh→0+fij(t1+h,t2+h)−fij(t1,t2)h=˙Ti(t1)−˙Tj(t2)=κ2N∑k∈[N]ζτkiki(t1)(1Ti(t1)−1Tτkik(t1))−κ2N∑k∈[N]ζτkjkj(t2)(1Tj(t2)−1Tτkjk(t2))≤κ2N∑k∈[N]ζτkiki(t1)(1Ti(t1)−1Tk(t1))−κ2N∑k∈[N]ζτkjkj(t2)(1Tj(t2)−1Tk(t2))+κ2ζ(0)(ΔτT(t1)Tτm(t1)2+ΔτT(t2)Tτm(t2)2). |
By using the relations
1Ti(t1)−1Tk(t1)=1TM(t1)−1Tk(t1)≤0≤1Tm(t2)−1Tk(t2)=1Tj(t2)−1Tk(t2), |
the above inequality reduces to
limh→0+fij(t1+h,t2+h)−fij(t1,t2)h≤κ2N∑k∈[N]ζ(SτX(t1))(1TM(t1)−1Tk(t1))−κ2N∑k∈[N]ζ(SτX(t2))(1Tm(t2)−1Tk(t2))+κ2ζ(0)(ΔτT(t1)Tτm(t1)2+ΔτT(t2)Tτm(t2)2)≤κ2N∑k∈[N]ζ(SτX(t))(1TM(t1)−1Tk(t1)−1Tm(t2)+1Tk(t2))+κ2ζ(0)(ΔτT(t1)Tτm(t1)2+ΔτT(t2)Tτm(t2)2)≤−κ2ζ(SτX(t))(1Tm(t2)−1TM(t1))+κ2ζ(0)Tτm(t1)Tτm(t2)maxk∈[N]|Tk(t1)−Tk(t2)|+κ2ζ(0)(ΔτT(t1)Tτm(t1)2+ΔτT(t2)Tτm(t2)2)=:λ(t1,t2). |
Therefore, by using Lemma 4.1, one can show that
D+[DτT](t)≤−κ2ζ(SτX(t))TτM(0)2DτT(t)+κ2ζ(0)Tτm(0)2(2ΔτT(t)+maxs∈[t−τ,t]ΔτT(s)), |
which implies
dDτT(t)dt≤−κ2ζ(SτX(t))TτM(0)2DτT(t)+3κ2ζ(0)Tτm(0)2maxs∈[t−τ,t]ΔτT(s), |
whenever DτT is differentiable at t.
Note that the differential inequality in Lemma 4.2 holds for (almost every) t>τ. This is because we need to estimate the derivative of fZ at all points on S(t). Since the ODE (3.1) only gives the derivative of ˙zi(s) for s>0, we require t−τ>0 in Lemma 4.2.
The next lemma allows us to present the upper bound of the objective function when the differential inequalities as Lemma 3.5 and Lemma 4.2 are given.
Lemma 4.3. Let y:[0,∞)→[0,∞) be a Lipschitz continuous function and f:[0,∞)→R be a continuous function. Assume that y and f satisfy
˙y(t)≤−ay(t)+maxs∈[t−τ,t]f(s),t>τ, | (4.1a) |
f(t)≤d∫tt−τy(s)ds,t>τ, | (4.1b) |
for some constants a,d>0 and τ≥0. If τ is sufficiently small to satisfy
τ<ad, |
and the following assertions hold.
(1) There are two constants b,c>0 such that
L(c)≤1d,L:(0,a)→R,L(x):=exτ(exτ−1)x(a−x),y(τ)≤becτa−c,maxt∈[0,2τ]f(t)ec(t−τ)<b. | (4.2) |
(2) Whenever b,c>0 satisfy (4.2), we have
y(t)<y(τ)e−a(t−τ)+becτa−c(e−c(t−τ)−e−a(t−τ)),t>τ,f(t)<be−c(t−τ),t≥0. | (4.3) |
Proof. Since L is continuous and
limx→0+L(x)=τa<1d,limx→a−L(x)=+∞, |
one can find a positive number c∈(0,a) satisfying L(c)≤1d, and one can also choose a sufficiently large b to satisfy (4.2). Now, assume there exists the first time t∗∈(2τ,∞) such that
f(t∗)ec(t∗−τ)=b. | (4.4) |
Then, from the definition of t∗, we have
f(t)<be−c(t−τ),t∈[0,t∗), | (4.5) |
and we substitute (4.5) to (4.1) to obtain
˙y(t)<−ay(t)+be−c(t−2τ),t∈(τ,t∗). | (4.6) |
Consequently, the Grönwall inequality (4.6) yields
y(t)<y(τ)e−a(t−τ)+becτa−c(e−c(t−τ)−e−a(t−τ))=becτa−ce−c(t−τ)+(y(τ)−becτa−c)e−a(t−τ)≤becτa−ce−c(t−τ),∀t∈(τ,t∗], | (4.7) |
where we used Eq (4.2) in the last inequality. On the other hand, by using Eq (4.7) to Eq (4.1b) at t=t∗, one can also obtain
f(t∗)≤d∫t∗t∗−τy(t)dt<d∫t∗t∗−τbecτa−ce−c(t−τ)dt=bdecτc(a−c)(e−c(t∗−2τ)−e−c(t∗−τ))=bde−c(t∗−τ)L(c)≤be−c(t∗−τ), |
which leads to a contradiction in Eq (4.4). Therefore, we have
f(t)<be−c(t−τ) |
for all t≥0, which implies the desired result for Eq (4.3).
Remark 4.1. For the case when τ=0, the function L becomes a constant function 0. Then, we can choose any number c from (0,a) to satisfy L(c)≤1d, and every positive b satisfying b≥(a−c)y(0) also satisfies the condition (4.2). Therefore, we have
y(t)<y(0)e−at+(y(0)+1n)(e−(a−1n)t−e−at),∀t>0,n∈N, |
which implies
y(t)≤y(0)e−at,∀t>0. |
Since inequality (4.1) for τ=0 is
˙y(t)≤−ay(t)+f(t),f(t)≤0,∀t>0, |
we can say that Lemma 4.3 gives the optimal upper bound for τ=0.
Now, we introduce the first sufficient framework for the emergence of flocking.
● (F1): Aτ,τ(0,0)=cosδ>0,δ∈(0,π2).
● (F2): There exists a constant Dτ,∞X>0 satisfying
Φ(Dτ,∞X)>Φ(DτX(0))+TτM(0)κ1cosδ⋅2sinδ2,Φ(x):=∫x0ϕ(u)du. |
● (F3): The time-delay bound τ≥0 is sufficiently small to satisfy
Φ(Dτ,∞X)>Φ(DτX(0)+2τsinδ2)+TτM(0)κ1cosδ⋅(2sinδ2+d1τ⋅βec1τc1),a1=c1+d1τec1τ⋅ec1τ−1c1τ,a2=c2+d2τec2τ⋅ec2τ−1c2τ, | (4.8) |
for some c1∈(0,a1],c2∈(0,a2], where the constants a1,a2,d1,d2, and β are given by
a1=κ1ϕ(Dτ,∞X)cosδTτM(0),a2=κ2ζ(Dτ,∞X)TτM(0)2,d1=3(N−1)N(κ1ϕ(0)Tτm(0))2,d2=3(N−1)N(κ2ζ(0)Tτm(0)2)2,β:=max{2sinδ2,maxi∈[N]‖v0i‖Lip(N−1)κ1ϕ(0)NTτm(0)}. |
Now, we are ready to provide the first main result on the asymptotic flocking of time-delayed unit speed TCS model (3.1), by showing that DτX is uniformly bounded and DτV,DτT converge to zero}.
Theorem 4.1. Let (X,V,T) be a solution to (3.1), where the initial data
(x0i,v0i,T0i):[−τ,0]→Rd×Sd−1×(0,∞),i∈[N] |
are Lipschitz continuous functions satisfying dx0idt=v0i and (F1)–(F3). Then, we estimate
DτX(t)≤SτX(t)<Dτ,∞X∀t>0,DτV(t)≤βc1τ1−e−c1τ⋅e−c1(t−τ)∀t>τ,ΔτV(t)≤(N−1)κ1ϕ(0)NTτm(0)βτe−c1(t−2τ)∀t≥0,DτT(t)≤γc2τ1−e−c2τ⋅e−c2(t−τ)∀t>τ,ΔτT(t)≤(N−1)κ2ζ(0)NTτm(0)2γτe−c2(t−2τ)∀t≥0, |
where the constant γ is given by
γ:=max{TτM(0)−Tτm(0),maxi∈[N]‖T0i‖Lip(N−1)κ2ζ(0)NTτm(0)2}. |
Therefore, the solution (X,V,T) exhibits the asymptotic flocking.
Proof. (Step 1) First, we find the range of t such that DτV and DτT satisfy the inequalities playing the role of y(t) in Lemma 4.3. Assume there exists a minimum t∗<∞ among all t satisfying SτX(t)≥Dτ,∞X. Then, by using Lemma 3.4, we have
Φ(Dτ,∞X)=Φ(SτX(t∗))=Φ(DτX(0)+∫t∗0DτV(s)ds)≤Φ(DτX(0)+√2−2Aτ,τ(0,0)t∗)=Φ(DτX(0)+2t∗sinδ2), |
which implies t∗>τ from the condition (4.8). We then use the monotone decreasing property of ϕ,ζ and Lemma 3.5 and Lemma 4.2 to obtain
dDτV(t)dt≤−a1DτV(t)+maxs∈[t−τ,t]3κ1ϕ(0)Tτm(0)ΔτV(s),a.e.t∈(τ,t∗),3κ1ϕ(0)Tτm(0)ΔτV(t)≤d1∫tt−τDτV(s)ds,t∈(τ,t∗), | (4.9) |
for the delayed diameter DτV, and similarly,
dDτT(t)dt≤−a2DτT(t)+maxs∈[t−τ,t]3κ2ζ(0)Tτm(0)2ΔτT(s),a.e.t∈(τ,t∗),3κ2ζ(0)Tτm(0)2ΔτT(t)≤d2∫tt−τDτT(s)ds,t∈(τ,t∗), | (4.10) |
for the delayed diameter DτT.
(Step 2) Now, we will show that t∗ cannot be a finite number so that Eqs (4.9) and (4.10) hold for all t∈(τ,∞). For every positive number b1,b2 satisfying
b1>βec1τ⋅d1τ,b2>γec2τ⋅d2τ, | (4.11) |
one can verify that (4.2) holds for (b1,c1) and (b2,c2). More precisely, by using Eqs (4.8) and (4.11) and Lemma 3.4, we show that
b1ec1τa1−c1=b1d1τ⋅c1τec1τ−1>β⋅c1τ1−e−c1τ≥β≥2sinδ2≥DτV(τ). |
In addition, we use the monotonic increasing property of Aτ,τ(t,t) (see Lemma 3.4) to verify
3κ1ϕ(0)Tτm(0)ΔτV(t)≤3κ1ϕ(0)Tτm(0)[max{τ−t,0}maxi∈[N]‖v0i‖Lip+(N−1)κ1ϕ(0)NTτm(0)∫tmax{0,t−τ}DτV(u)du]≤3κ1ϕ(0)Tτm(0)[max{τ−t,0}maxi∈[N]‖v0i‖Lip+(N−1)κ1ϕ(0)NTτm(0)∫tmax{0,t−τ}√2−2Aτ,τ(0,0)du]=3κ1ϕ(0)Tτm(0)[max{τ−t,0}maxi∈[N]‖v0i‖Lip+(N−1)κ1ϕ(0)NTτm(0)(t−max{0,t−τ})(2sinδ2)]=3κ1ϕ(0)Tτm(0)[max{τ−t,0}maxi∈[N]‖v0i‖Lip+(N−1)κ1ϕ(0)NTτm(0)(min{t,τ})(2sinδ2)]≤3κ1ϕ(0)Tτm(0)[max{τ−t,0}(N−1)κ1ϕ(0)NTτm(0)β+(N−1)κ1ϕ(0)NTτm(0)(min{t,τ})β]=3κ1ϕ(0)Tτm(0)⋅(N−1)κ1ϕ(0)NTτm(0)⋅β[max{τ−t,0}+min{t,τ}]=3κ1ϕ(0)Tτm(0)⋅(N−1)κ1ϕ(0)NTτm(0)⋅β[max{τ,t}+min{t,τ}−t]=3κ1ϕ(0)Tτm(0)⋅(N−1)κ1ϕ(0)NTτm(0)⋅β[τ+t−t]=β⋅d1τ,∀t∈[0,2τ], |
which implies
maxt∈[0,2τ]3κ1ϕ(0)Tτm(0)ΔτV(t)ec1(t−τ)≤βec1τ⋅d1τ<b1. |
Similarly, by using Eqs (4.8) and (4.11) and Lemma 3.2, we have
b2ec2τa2−c2=b2d2τ⋅c2τec2τ−1>γ⋅c2τ1−e−c2τ≥γ≥TτM(0)−Tτm(0), |
and we use the monotonic decreasing property of DτT(t)=TτM(t)−Tτm(t) (see Lemma 3.2) to derive
3κ2ζ(0)Tτm(0)2ΔτT(t)≤3κ2ζ(0)Tτm(0)2[max{τ−t,0}maxi∈[N]‖T0i‖Lip+(N−1)κ2ζ(0)NTτm(0)2∫tmax{0,t−τ}DτT(u)du]≤3κ2ζ(0)Tτm(0)2[max{τ−t,0}maxi∈[N]‖T0i‖Lip+(N−1)κ2ζ(0)NTτm(0)2∫tmax{0,t−τ}DτT(0)du]=3κ2ζ(0)Tτm(0)2[max{τ−t,0}maxi∈[N]‖T0i‖Lip+(N−1)κ2ζ(0)NTτm(0)2DτT(0)(min{t,τ})]≤3κ2ζ(0)Tτm(0)2[max{τ−t,0}⋅(N−1)κ2ζ(0)NTτm(0)2⋅γ+(N−1)κ2ζ(0)NTτm(0)2⋅γ⋅min{t,τ}]=3κ2ζ(0)Tτm(0)2⋅(N−1)κ2ζ(0)NTτm(0)2⋅γ[max{τ−t,0}+min{t,τ}]=γ⋅d2τ, |
which implies
maxt∈[0,2τ]3κ2ζ(0)Tτm(0)2ΔτT(t)ec2(t−τ)≤γec2τ⋅d2τ<b2. |
Therefore, we apply Lemma 4.3 to (4.9) and (4.10) for all b1,b2 satisfying (4.11) to obtain
3κ1ϕ(0)Tτm(0)ΔτV(t)≤d1τβe−c1(t−2τ),3κ2ζ(0)Tτm(0)2ΔτT(t)≤d2τγe−c2(t−2τ),∀t∈(0,t∗). | (4.12) |
In particular, we employ the result of Lemma 4.2 to (4.12) to get
ddt[DτV(t)+κ1Φ(SτX(t))cosδTτM(0)]≤d1τβe−c1(t−2τ),∀t∈(τ,t∗). | (4.13) |
However, from direct calculation, (4.8) and (4.13) yield
DτV(t∗)+κ1Φ(SτX(t∗))cosδTτM(0)≤DτV(τ)+κ1Φ(SτX(τ))cosδTτM(0)+d1τ⋅βec1τc1(1−e−c1(t∗−τ))≤2sinδ2+κ1Φ(DτX(0)+2τsinδ2)cosδTτM(0)+d1τ⋅βec1τc1<κ1Φ(Dτ,∞X)cosδTτM(0), |
which contradicts to SτX(t∗)=Dτ,∞X obtained from the existence of t∗<∞. Therefore, we have SτX(t)<Dτ,∞X for all t>τ, and we apply Lemma 4.3 to (4.9) and (4.10) for all b1,b2 satisfying (4.11) to reach the desired results.
Remark 4.2. Theorem 4.1 is also applicable when time-delay τ is set to 0. If τ=0, the framework (F1)–(F3) becomes
A(0)>0,∫∞DX(0)ϕ(s)ds>TM(0)DV(0)κ1A(0), |
which coincides with the condition (2.2). In addition, the condition (4.8) reduces to
c1=a1,c2=a2,β=DV(0),γ=DT(0). |
Therefore, Theorem 4.1 exactly coincides with Proposition 2.5 for τ=0.
In this section, we provide a system of differential inequalities on diameters DX,DV, and DT. In this case, the Lipschitz continuity of the diameters can be easily obtained from the Lipschitz continuity of X,V,T. The goal of this section is to find a sufficient framework without the condition (F1), so that the initial data might allow ⟨vi(t1),vj(t2)⟩<0 for some i,j∈[N] and t1,t2∈[−τ,0]. More precisely, we will replace (F1) to the weaker condition that
maxi,j∈[N]⟨vi(0),vj(0)⟩>0. |
In this case, since we cannot apply the result of Lemma 3.4, the proof of Theorem 4.1 cannot be applied as is, and we must consider the possibility that Aτ,τ may oscillate at the beginning rather than increasing monotonically. The overall flow of this section is that we can still control the diameter DV under the condition that τ is sufficiently small.
Lemma 5.1. Let (X,V,T) be a solution to Eq (3.1) subject to the continuous initial data {(x0i,v0i,T0i)}Ni=1 satisfying dx0idt=v0i for all i∈[N]. Then, the following differential inequalities hold:
1. Whenever DX is differentiable at time t>−τ, we have
dDX(t)dt≤DV(t). |
2. Whenever DV is differentiable at time t>0 and DV(t)≤√2, we have
dDV(t)dt≤−κ1ϕ(DX(t)+τ)TτM(t)[1−N−12NDV(t)2]DV(t)+2(N−1)κ1ϕ(0)NTτm(t)ΔτV(t). |
3. Whenever DT is differentiable at time t>0, we have
dDT(t)dt≤−κ2ζ(DX(t)+τ)(1Tm(t)−1TM(t))+2(N−1)κ2ζ(0)N(Tτm(t))2ΔτT(t). |
Proof. We can prove this lemma by using a similar argument to Lemma 4.2.
(1) For every i,j∈[N] and t>−τ, we have
ddt‖xi(t)−xj(t)‖2=2⟨xi(t)−xj(t),vi(t)−vj(t)⟩≤2DX(t)DV(t). |
Therefore, one can obtain
D+[(DX)2](t)≤2maxi,j∈[N]⟨xi(t)−xj(t),vi(t)−vj(t)⟩≤2DX(t)DV(t), |
which implies
2DX(t)dDX(t)dt≤2DX(t)DV(t), |
whenever DX is differentiable at time t.
(2) Let J(t) be the set of all index pairs (i,j) which maximize the difference between two velocities at time t, i.e.,
(i,j)∈J(t)⟺‖vi(t)−vj(t)‖=DV(t)≤√2. |
Then, for every (i,j)∈J(t), we can split the derivative of ‖vi−vj‖2 at time t into four parts as follows:
ddt‖vi(t)−vj(t)‖2=2⟨vi(t)−vj(t),dvidt(t)−dvjdt(t)⟩=−2κ1N∑k∈[N]−{i}ϕτkikiTτkik(t)⟨vj(t),vτkik(t)−⟨vτkik(t),vi(t)⟩vi(t)⟩−2κ1N∑k∈[N]−{j}ϕτkjkjTτkjk(t)⟨vi(t),vτkjk(t)−⟨vτkjk(t),vj(t)⟩vj(t)⟩=−2κ1N∑k∈[N]−{i}ϕτkikiTτkik(t)⟨vj(t),vk(t)−⟨vk(t),vi(t)⟩vi(t)⟩−2κ1N∑k∈[N]−{j}ϕτkjkjTτkjk(t)⟨vi(t),vk(t)−⟨vk(t),vj(t)⟩vj(t)⟩−2κ1N∑k∈[N]−{i}ϕτkikiTτkik(t)⟨vj(t),vτkik(t)−vk(t)−⟨vτkik(t)−vk(t),vi(t)⟩vi(t)⟩−2κ1N∑k∈[N]−{j}ϕτkjkjTτkjk(t)⟨vi(t),vτkjk(t)−vk(t)−⟨vτkjk(t)−vk(t),vj(t)⟩vj(t)⟩=:L11+L12+L13+L14. |
Below, we estimate these four parts separately into two parts.
⋄ (Estimate of L11+L12): In this case, one can verify the following inequalities at time t:
⟨vj,vk−⟨vk,vi⟩vi⟩≥⟨vj,vi⟩−⟨vj,vi⟩⟨vk,vi⟩=⟨vj,vi⟩(1−⟨vk,vi⟩)≥0,⟨vi,vk−⟨vk,vj⟩vj⟩≥⟨vi,vj⟩−⟨vi,vj⟩⟨vk,vj⟩=⟨vi,vj⟩(1−⟨vk,vj⟩)≥0, |
where we used ⟨vi,vj⟩=1−12D2V≥0 in the last inequality. This implies that both L11 and L12 are nonpositive numbers, and therefore
L11+L12≤−2κ1N∑k∈[N]−{i}ϕ(DX+τ)TτM⟨vj,vk−⟨vk,vi⟩vi⟩−2κ1N∑k∈[N]−{j}ϕ(DX+τ)TτM⟨vi,vk−⟨vk,vj⟩vj⟩=−2κ1ϕ(DX+τ)NTτM∑k∈[N]−{i,j}[⟨vj,vk−⟨vk,vi⟩vi⟩+⟨vi,vk−⟨vk,vj⟩vj⟩]−2κ1ϕ(DX+τ)NTτM∑k∈{i,j}[1−⟨vi,vj⟩2], | (5.1) |
where we used the monotonic decreasing property of ϕ and
‖xi−xτkik‖≤‖xi−xk‖+‖xk−xτkik‖≤DX+τ,Ti(t−τki(t))≤TτM(t),‖xj−xτkjk‖≤‖xj−xk‖+‖xk−xτkjk‖≤DX+τ,Tj(t−τkj(t))≤TτM(t), |
in the first inequality. Then, we use ⟨vi+vj,vk⟩≥2⟨vi,vj⟩ and
∑k∈{i,j}(1−⟨vi,vj⟩)={2(1−⟨vi,vj⟩)(i≠j)0(i=j)=2(1−⟨vi,vj⟩) | (5.2) |
in Eq (5.1) to obtain
L11+L12≤−2κ1ϕ(DX+τ)NTτM(1−⟨vi,vj⟩)[∑k∈[N]−{i,j}2⟨vi,vj⟩+∑k∈{i,j}(1+⟨vi,vj⟩)]=−2κ1ϕ(DX+τ)NTτM(1−⟨vi,vj⟩)[2N⟨vi,vj⟩+∑k∈{i,j}(1−⟨vi,vj⟩)]=−2κ1ϕ(DX+τ)TτM(1−⟨vi,vj⟩)[2⟨vi,vj⟩+2N(1−⟨vi,vj⟩)]. | (5.3) |
⋄ (Estimate of L13+L14): From direct calculation, we have
L13+L14=−2κ1N∑k∈[N]−{i}ϕτkikiTτkik⟨vj−vi,vτkik−vk−⟨vτkik−vk,vi⟩vi⟩−2κ1N∑k∈[N]−{j}ϕτkjkjTτkjk⟨vi−vj,vτkjk−vk−⟨vτkjk−vk,vj⟩vj⟩≤4(N−1)κ1ϕ(0)NTτm‖vi−vj‖ΔτV, | (5.4) |
where we used
ϕτkiki≤ϕ(0),Tτm≤Tτkik,‖vτkik−vk−⟨vτkik−vk,vi⟩vi‖≤‖vτkik−vk‖≤ΔτV,ϕτkjkj≤ϕ(0),Tτm≤Tτkjk,‖vτkjk−vk−⟨vτkjk−vk,vj⟩vj‖≤‖vτkjk−vk‖≤ΔτV, |
in the last inequality.
Therefore, we combine Eqs (5.3) and (5.4) to obtain
D+[(DV)2](t)≤−κ1ϕ(DX+τ)TτMD2V[2−(1−1N)D2V]+4(N−1)κ1ϕ(0)NTτmDVΔτV, |
which implies
dDVdt≤−κ1ϕ(DX+τ)TτMDV[1−N−12ND2V]+2(N−1)κ1ϕ(0)NTτmΔτV, |
whenever DV is differentiable and DV≤√2.
(3) Let M(t),m(t) be the set of all indices such that
i∈M(t),j∈m(t)⟺Ti(t)=TM(t),Tj(t)=Tm(t). |
Then, for every i∈M(t) and j∈m(t), we can split the derivative of Ti−Tj at time t into two parts as follows:
ddt(Ti−Tj)=κ2NN∑k=1ζτkiki(1Ti−1Tτkik)−κ2NN∑k=1ζτkjkj(1Tj−1Tτkjk)=κ2NN∑k=1ζτkiki(1Ti−1Tk)−κ2NN∑k=1ζτkjkj(1Tj−1Tk)+κ2NN∑k=1ζτkiki(1Tk−1Tτkik)−κ2NN∑k=1ζτkjkj(1Tk−1Tτkjk)=:L21+L22. |
Then, we estimate these two parts separately.
⋄ (Estimate of L21): From a direct calculation, it follows that
L21≤−κ2ζ(DX+τ)N[∑k∈[N]−{i}(1Tk−1Ti)+∑k∈[N]−{j}(1Tj−1Tk)]=−κ2ζ(DX+τ)N[∑k∈[N]−{i,j}(1Tk−1Ti)+∑k∈[N]−{i,j}(1Tj−1Tk)+∑k∈{i,j}(1Tj−1Ti)]=−κ2ζ(DX+τ)(1Tj−1Ti), |
where we used Tj(t)≤Tk(t)≤Ti(t) for all i∈M(t) and j∈m(t) in the first inequality.
⋄ (Estimate of L22): In this case, we use Lemma 3.2 to obtain
L22≤κ2ζ(0)N∑k∈[N]−{i}|Tk−Tτkik|TkTτkik+κ2ζ(0)N∑k∈[N]−{j}|Tk−Tτkjk|TkTτkjk≤2(N−1)κ2ζ(0)N⋅ΔτT(Tτm)2. |
Therefore, we combine these two estimates to obtain
D+[DT](t)≤−κ2ζ(DX(t)+τ)(1Tm(t)−1TM(t))+2(N−1)κ2ζ(0)N(Tτm(t))2ΔτT(t), |
which implies the desired result.
Remark 5.1. One thing to keep in mind during the proof of Lemma 5.1 is that the acceleration may not be zero even if there is a point at which the velocity diameter DV is zero, due to the time-delay effect. Therefore, even if we find i,j with DV=‖vi−vj‖, we cannot exclude the possibility that i and j are equal. Another thing to keep in mind is that the condition DV(t)≤√2 is necessary to obtain a meaningful inequality for DV because of the technical reason, otherwise the values
⟨vi,vk−⟨vk,vj⟩vj⟩=⟨vi,vk⟩−⟨vk,vj⟩⟨vi,vj⟩,⟨vj,vk−⟨vk,vi⟩vi⟩=⟨vj,vk⟩−⟨vk,vi⟩⟨vj,vi⟩, |
can be negative when ⟨vi,vk⟩=⟨vj,vk⟩=⟨vi,vj⟩<0.
Similar to what we have done in Section 4, we prepare a lemma to present the upper bound of (DX,DV,DT) and (ΔV,ΔT) by using Lemma 3.5 and Lemma 5.1.
Lemma 5.2. Let y:[0,∞)→[0,∞) be a Lipschitz continuous function and f:[0,∞)→R be a continuous function. Assume that y and f satisfy
˙y(t)≤−χy(t)+2θf(t),t>0, | (5.5a) |
f(t)≤Lmax{τ−t,0}+θ∫tmax{t−τ,0}(y(s)+f(s))ds,t>0, | (5.5b) |
for some constants χ,θ,L>0 and τ≥0. If τ is sufficiently small to satisfy
τθ(2θχ+1)<1, | (5.6) |
the following assertions hold.
(1) There are two constants ξ>0 and μ∈(0,χ) such that
[Leμτξ+θ(2θχ−μ+1)⋅eμτ−1μτ]τ≤1,y(0)≤2θξχ−μ,f(0)<ξ. | (5.7) |
(2) Whenever ξ,μ>0 satisfy Eq (5.7), we have
y(t)<y(0)e−χt+2ξθχ−μ(e−μt−e−χt),t>0,f(t)<ξe−μt,t≥0. | (5.8) |
Proof. From Eqs (5.5) and (5.6), we can choose a sufficiently large ξ satisfying
f(0)≤Lτ<Lτ1−τθ(2θχ+1)<ξ,y(0)≤2θξχ. |
Then, the condition (5.7) holds for sufficiently small positive number μ∈(0,χ). Now, assume there exists a minimum t∗∈(0,∞) among all t satisfying
f(t)=ξe−μt | (5.9) |
Then, we have
f(t)<ξe−μt,t∈[0,t∗), | (5.10) |
and we substitute (5.10) to (5.5) to obtain
˙y(t)<−χy(t)+2θξe−μt,t∈(0,t∗). |
Consequently, this Grönwall's inequality yields
y(t)<y(0)e−χt+2θξχ−μ(e−μt−e−χt)≤2θξχ−μe−μtt∈(0,t∗], | (5.11) |
where we used Eq (5.7) in the last inequality. On the other hand, by using Eqs (5.10) and (5.11) to (5.5b) at t=t∗, we have
f(t∗)≤Lmax{τ−t∗,0}+θ∫t∗max{t∗−τ,0}(y(s)+f(s))ds<Lmax{τ−t∗,0}+θ∫t∗max{t∗−τ,0}(2θξχ−μ+ξ)e−μsds=Lmax{τ−t∗,0}+θμ(2θξχ−μ+ξ)(e−μmax{t∗−τ,0}−e−μt∗)=[Lmax{τ−t∗,0}eμt∗+θμ(2θξχ−μ+ξ)(eμmin{t∗,τ}−1)]e−μt∗≤[Lτeμτ+θμ(2θξχ−μ+ξ)(eμτ−1)]e−μt∗=[Leμτ+θξ(2θχ−μ+1)(eμτ−1μτ)]τe−μt∗≤ξe−μt∗, |
where we used Eq (5.7) in the last inequality. However, this leads a contradiction to Eq (5.9), and therefore
f(t)<ξe−μt,∀t≥0, |
which implies the desired inequality (5.8).
Unlike in Section 4, we need to guarantee DV(t)≤√2 to apply Lemma 5.2 to Lemma 5.1. From Eqs (5.7) and (5.8), one can find a time-invariant upper bound for y(t). Since the function
t↦y(0)e−χt+2ξθχ−μ(e−μt−e−χt) |
has a maximum at t=t0 satisfying
−μ2θχ−μe−μt0+χ(2ξθχ−μ−y(0))e−χt0=0, |
we have
y(0)e−χt0+2ξθχ−μ(e−μt0−e−χt0)=2ξθχ−μe−μt0−(2ξθχ−μ−y(0))e−χt0=(1−μχ)2ξθχ−μe−μt0=2ξθχe−μt0≤2ξθχ. |
Therefore, we have the following time-invariant upper bound for y(t):
y(t)<2ξθχ,∀t>0. |
Now, by using Lemma 5.1 and 5.2, we can construct another sufficient framework leading to the asymptotic flocking. In this case, we assume DX(0) and DV(0) sufficiently small, instead of assuming the smallness of DτX(0) and DτV(0).
● (F1)′: DV(0)=2sinδ2,δ∈(0,π2).
● (F2)′: There exists a constant D∞X>0 satisfying
Φ(D∞X+τ)>Φ(DX(0)+τ)+TτM(0)κ1cosδ⋅2sinδ2. |
● (F3)′: The time-delay bound τ>0 is sufficiently small to satisfy
Φ(D∞X+τ)>Φ(DX(0)+τ)+TτM(0)κ1cosδ⋅(2sinδ2+2(χ1−μ1)μ1), | (5.12a) |
[θ1L1eμ1τχ1−μ1+θ1(2θ1χ1−μ1+1)⋅eμ1τ−1μ1τ]τ=1, | (5.12b) |
[2θ2L2eμ2τ(χ2−μ2)(TτM(0)−Tτm(0))+θ2(2θ2χ2−μ2+1)⋅eμ2τ−1μ2τ]τ=1, | (5.12c) |
χ1−μ1χ1<12min{√2,√NN−1DV(0)}, | (5.12d) |
for some μ1∈(0,χ1) and μ2∈(0,χ2), where the constants χ1,χ2,θ1,θ2,L1,L2 are defined as
χ1=κ1ϕ(D∞X+τ)TτM(0)⋅(1−12DV(0)2),χ2=κ2ζ(D∞X+τ)TτM(0)2,θ1=(N−1)κ1ϕ(0)NTτm(0),θ2=(N−1)κ2ζ(0)NTτm(0)2,L1=maxi∈[N]‖v0i‖Lip,L2=maxi∈[N]‖T0i‖Lip. |
Note that from the equality conditions in (5.12), μ1 tends to χ1 during τ→+0. Therefore, (F3)′ is indeed a condition that holds for sufficiently small τ.
Finally, we are ready to provide the second main result, which also shows that DX is uniformly bounded and DV,DT converge to zero. Although the proof might not be intuitive and its implications are hard to grasp at first glance, we wrote it down in detail for the completeness.
Theorem 5.1. Let (X,V,T) be a solution to Eq (3.1), where the initial data
(x0i,v0i,T0i):[−τ,0]→Rd×Sd−1×(0,∞),i∈[N] |
are Lipschitz continuous functions satisfying dx0idt=v0i and (F1)′–(F3)′. Then, the following inequalities hold for all t>0:
DX(t)<D∞X,DV(t)<DV(0)e−χ1t+2(e−μ1t−e−χ1t),DT(t)<DT(0)e−χ2t+(TτM(0)−Tτm(0))(e−μ2t−e−χ2t),ΔV(t)<χ1−μ1θ1e−μ1t,ΔT(t)<(χ2−μ2)(TτM(0)−Tτm(0))2θ2e−μ2t. |
Therefore, the solution (X,V,T) exhibits the asymptotic flocking.
Proof. (Step 1) Assume there exists a minimum t∗<∞ among all t satisfying
DX(t)≥D∞Xor1−N−12NDV(t)2≤1−12DV(0)2orDV(t)≥√2. | (5.13) |
Then, we will show that, in fact, only the first condition among the three can be satisfied, i.e., DX(t∗)=D∞X. First, one can obtain t∗>0 by using (F1)′,(F2)′, and the continuity of DV, and we use the monotonic decreasing property of ϕ,ζ, Lemma 3.5, and Lemma 5.1 to obtain
dDV(t)dt≤−χ1DV(t)+2θ1ΔτV(t),a.e.t∈(0,t∗),ΔτV(t)≤L1max{τ−t,0}+θ1∫tmax{t−τ,0}(DV(s)+ΔτV(s))ds,t∈(0,t∗), | (5.14) |
for the diameter DV, and similarly,
dDT(t)dt≤−χ2DT(t)+2θ2ΔτT(t),a.e.t∈(0,t∗),ΔτT(t)≤L2max{τ−t,0}+θ2∫tmax{t−τ,0}(DT(s)+ΔτT(s))ds,t∈(0,t∗), | (5.15) |
for the diameter DT. Now, for the two constants ξ1,ξ2 defined as
ξ1=χ1−μ1θ1=L1eμ1τ+2θ11−θ1τ(eμ1τ−1μ1τ)τ,ξ2=(χ2−μ2)(TτM(0)−Tτm(0))2θ2=L2eμ2τ+(TτM(0)−Tτm(0))θ21−θ2τ(eμ2τ−1μ2τ)τ, |
one can verify that (5.7) holds for (L1,χ1,μ1,ξ1) and (L2,χ2,μ2,ξ2). More precisely, the first condition of (5.7) is immediately obtained from (5.12b) and (5.12c), and the third condition is obtained from
ΔτV(0)≤L1τ,ΔτT(0)≤L2τ, |
which is a consequence of Lemma 3.5 for t=0. The second condition of (5.7) can also be verified by the relation
DV(0)≤2=2ξ1θ1χ1−μ1,DT(0)≤TτM(0)−Tτm(0)=2θ2ξ2χ2−μ2. |
As a consequence, we apply Lemma 5.2 to Eqs (5.14) and (5.15) to get the following inequalities for t∈(0,t∗):
DV(t)<DV(0)e−χ1t+2(e−μ1t−e−χ1t),ΔV(t)<χ1−μ1θ1e−μ1t,DT(t)<DT(0)e−χ2t+(TτM(0)−Tτm(0))(e−μ2t−e−χ2t),ΔT(t)<(χ2−μ2)(TτM(0)−Tτm(0))2θ2e−μ2t. | (5.16) |
Then, we use Eq (5.12d) to obtain
DV(t)<2ξ1θ1χ1=2(χ1−μ1)χ1<min{√2,√NN−1DV(0)},t∈(0,t∗), |
and therefore, the only possible case to satisfy (5.13) is the first case, i.e., DX(t∗)=D∞X.
(Step 2) Therefore, we can apply Lemma 5.1 (2) to t∈(0,t∗) to obtain
dDV(t)dt≤−κ1ϕ(DX(t)+τ)TτM(t)[1−N−12NDV(t)2]DV(t)+2(N−1)κ1ϕ(0)NTτm(t)ΔτV(t)≤−κ1ϕ(DX(t)+τ)TτM(0)[1−12DV(0)2]DV(t)+2ξ1θ1e−μ1t,t∈(0,t∗), |
which implies that
ddt[DV(t)+κ1(1−12DV(0)2)TτM(0)Φ(DX(t)+τ)+2ξ1θ1μ1e−μ1t]≤0,t∈(0,t∗). | (5.17) |
However, (5.17) yields
κ1(1−12DV(0)2)TτM(0)Φ(DX(t∗)+τ)≤DV(0)+κ1(1−12DV(0)2)TτM(0)Φ(DX(0)+τ)+2ξ1θ1μ1, |
which leads a contradiction to (5.12a). Thus, there is no t∗∈(0,∞) satisfying (5.13), and the inequalities in (5.16) hold for all t>0.
Remark 5.2. Similar to the case in Remark 4.2, Theorem 5.1 is also applicable when time-delay τ is set to 0. If τ=0, the framework (F1)′−(F3)′ becomes
A(0)>0,∫∞DX(0)ϕ(s)ds>TM(0)DV(0)κ1A(0), |
which coincides with the condition (2.2). In addition, the condition (5.12) reduces to
μ1=χ1,μ2=χ2. |
Therefore, Theorem 5.1 also exactly coincides with Proposition 2.5 for τ=0.
In this section, we perform numerical simulations on (1.3) to verify the results of Theorem 4.1 and Theorem 5.1, and to explore whether there are other properties that we could not prove due to technical reasons. For the numerical implementation, we used the Euler method with time step Δt=0.05, and we fixed N,ϕ,ζ as
N=50,ϕ(r)=11+r2,ζ(r)=11+r3/2,r≥0. | (6.1) |
In addition, for the time delay τij, we set
τij(t):=τ×[2+sin(0.1(i+j)t)3],t≥0, | (6.2) |
for each i,j∈[N]. Thus, we compare how the dynamics of (1.3) vary with respect to κ1,κ2,τ, and initial data.
In particular, we prepare two types of initial data for equation (1.3), which we will refer to as 'good initial' and 'bad initial' based on the velocity range. The way we set the initial conditions for X and T is common to both types and is as follows. For each i∈[N] and n∈[−τ/Δt,0], we randomly select xi(n) from a uniform distribution on (0,1)3 and Ti(n) from a uniform distribution on (1,3.5). Then, the 'good initial' refers to the initial data for which
Lemma 3.4 and Theorem 4.1 can be applied. |
To achieve this, for each i∈[N] and n∈[−τ/Δt,0], we randomly select ˜vi(n) from a uniform distribution on (0,1)3, just like the initial positions xi(n). Then, we normalize their norms to 1 as follows:
vi(n):=˜vi(n)‖˜vi(n)‖,i∈[N],n∈[−τ/Δt,0]. | (6.3) |
If so, the discretized version of , i.e.,
is strictly positive, and we can expect that (the discrete analogue of) will be monotonically increasing, as proven in Lemma 3.4. On the other hand, by 'bad initial, ' we refer to the initial data for which
In other words, it refers to the initial conditions that do not satisfy the condition but satisfy the condition . We meet these requirements by randomly selecting from a uniform distribution on for each and , and also select randomly from a uniform distribution on . Then, we normalize their norms to as in Eq (6.3) to satisfy the unit speed constraint. In this case, we can expect that flocking can occur when is sufficiently small, according to Theorem 5.1. However, since we cannot apply Lemma 3.4, may not be increasing monotonically at the beginning. Thus, one of the goals of this section is to confirm that the initial behavior of differs between good initial and bad initial based on the applicability of Lemma 3.4. However, when we actually ran the simulations, we found that observing the behavior of is sufficient to clearly distinguish between the two cases. Therefore, to save computation time, we only display , and over (time) in all figures of this section.
In Figures 1 and 2, we fix and used the good initial data to simulate the solution of (1.3) for . The first two plots in Figure 1 are the temporal evolution of and over , and Figure 2 shows the temporal evolution of and . These plots show that the solution exhibited asymptotic flocking under the given setting. The last two plots in Figure 1 confirm that the speeds of all particles remained at 1 throughout the simulation. In the case of , it exhibits a monotonically decreasing behavior over time, while we can also observe an inflection point around . In fact, such inflection points were consistently observed in throughout the simulations, and we found that their position was always approximately at . We suspect that the value itself is not particularly important, but rather that the average of set in Eq (6.2) is , and something special seems to occur around twice that time.
In Figure 3, we fix and used the good initial data to simulate the solution of Eq (1.3) for . Once again, asymptotic flocking occurred, but the rate of convergence was significantly slower compared to the previous examples. Additionally, we observed an inflection point in around .
In Figure 4, we fix and used the good initial data to simulate the solution of (1.3) for . As increases, the time it takes for to saturate becomes much longer. Therefore, to verify whether flocking occurs, we had to observe the behavior of diameters for such a long time. For this reason, the inflection point is not clearly visible; however, the position of the dashed line set in this figure is also at . Additionally, in this figure, we can see that does not converge to 0, indicating that the solution did not exhibit flocking. This means that even if the condition is satisfied and are fixed, flocking will no longer occur if becomes sufficiently large.
In Figure 5, we fix and used the good initial data to simulate the solution of (1.3) for . In this case, although is small as in Figures 1 and 2, the small values of result in a situation where flocking does not occur. Notably, due to the small value of , did not converge to 0 in this case.
In Figures 6 and 7, we fix and used the bad initial data to simulate (1.3). In other words, we are using the same and as in Figures 1 and 2, but with the bad initial data. The first two plots in Figure 6 illustrate the behavior of and from the given bad initial data up to time , while the last two plots show the behavior up to for the same initial data. A notable feature observed in this figure is that, similar to Figure 1, the behavior of changes around the time . However, the behavior of for becomes unstable rather than just monotonically decreasing, which is a result not observed in any simulations using good initial data. Nevertheless, as demonstrated in Theorem 5.1, even with bad initial data, the asymptotic flocking can still occur if is sufficiently small and are sufficiently large.
In Figures 8 and 9, we fix and used the bad initial data to simulate (1.3). In other words, we are using the same and as in Figure 3, but with bad initial data. The first two plots in Figure 6 illustrate the behavior of and from the given bad initial data up to time , while the last two plots show the behavior up to for the same initial data. In this case, shows a very unstable oscillation before , and it even becomes larger than at . However, this effect disappears after and begins to decrease monotonically. Nevertheless, since both and do not converge to 0 in this case, asymptotic flocking does not occur. This contrasts with the occurrence of asymptotic flocking in Figure 3, demonstrating that the condition also affects the occurrence of asymptotic flocking.
Finally, in Figures 10 and 11, we fix and used the bad initial data to simulate (1.3).
That is, we are using the same and as in Figures 5, but with the bad initial data. The first two plots in Figure 10 illustrate the behavior of and from the given bad initial data up to time , while the last two plots show the behavior up to for the same initial data. In this case, shows a small oscillation before , but decrease monotonically after . However, as with the use of good initial data in Figure 5, asymptotic flocking did not occur in this case either.
In this paper, we have demonstrated several sufficient frameworks for the asymptotic flocking dynamics of the thermodynamic CS model with a unit-speed constraint and time-delay. To do this, we first proved the monotonic property of extreme temperatures and maximal angle between velocities, and then we provided basic estimates concerning {position–velocity–temperature} diameters and perturbation functions. Then, we derived dissipative inequalities with respect to the diameters and delayed diameters, and proposed suitable ansatz for the decay rate of the perturbation function to find the sufficient framework to exhibit asymptotic flocking of (1.3). However, there are still some interesting topics that might be studied in the future. For instance, we wonder if it is possible to find differential inequalities for to show the asymptotic flocking. Since we have already shown that is monotonically increasing, we expect that if we succeed in obtaining a differential inequality for the itself, we will not need to use perturbation functions and suggest an ansatz on its exponential decay rate. We leave this issue for future work.
Hyunjin Ahn: Investigation, Funding acquisition, Writing, Validation; Woojoo Shim: Investigation, Methodology, Writing, Supervision.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
The work of H. Ahn was supported by the National Research Foundation of Korea (NRF) grant funded by the Korean government (MSIT) (2022R1C12007321).
The authors declare there is no conflict of interest.
Proof. (1) Suppose we have for some . Then, there exists such that . Since is continuous, one can find such that
Therefore, we have
(2) To show that is monotonically increasing, we claim the following:
Once we prove the above claim, we can show for all . To see this, consider a function
This function has a maximum point as a consequence of the Weierstrass extreme value theorem. In addition, since , we may assume . Therefore, we have
which implies .
Now, suppose we have for some . Then, one can obtain from the result (1), and we have
where we used in the second equality. Therefore, there exist two constants such that
(A.1) |
From (A.1), we can employ Lemma 3.1(1) to for each and obtain
and since is continuous ( Berge's maximum theorem), we have
which leads to a contradiction in .
(3) It is sufficient to prove that
If , one can find such that
due to the continuity of . Therefore, we have
where we used the Lipschitz continuity of . On the other hand, if , then
where we used
in the last equality. Therefore, we have
where we used the Lipschitz continuity of in the last inequality.
Proof. From the Berge maximum theorem, the set valued map
is upper hemi-continuous with nonempty and compact values. This means that for every , if contains , there exists a neighborhood of such that for all , is a subset of .
Now, fix arbitrary , and define . Then, we have
Since is continuous, there exists an open neighborhood
for each , such that for all . Similarly, for each , there exists an open neighborhood
such that for all . Finally, there exists an open set
such that for all . Then, is an open cover of . In addition, since is compact, we can find its finite subcover . Therefore, we can find an open subset of such that
Therefore, by using the Berge maximum theorem, there exists a positive number such that for all . This means that the Dini derivative of
is less than or equal to for all and , and therefore is monotonically decreasing in for all . By using this result, we have
which implies that for every ,
Since can be any positive number, we have
[1] |
J. A. Acebrón, L. L. Bonilla, C. J. Pérez Vicente, F. Ritort, R. Spigler, The Kuramoto model: A simple paradigm for synchronization phenomena, Rev. Mod. Phys., 77 (2005), 137–185. https://doi.org/10.1103/RevModPhys.77.137 doi: 10.1103/RevModPhys.77.137
![]() |
[2] |
H. Ahn, Emergent behaviors of thermodynamic Cucker–Smale ensemble with a unit-speed constraint, Discrete Contin. Dyn. Syst. B, 28 (2023), 4800–4825. https://doi.org/10.3934/dcdsb.2023042 doi: 10.3934/dcdsb.2023042
![]() |
[3] |
H. Ahn, Asymptotic flocking of the relativistic Cucker-Smale model with time-delay, Netw. Heterog. Media, 18 (2023), 29–47. https://doi.org/10.3934/nhm.2023002 doi: 10.3934/nhm.2023002
![]() |
[4] |
G. Albi, N. Bellomo, L. Fermo, S. Y. Ha, J. Kim, L. Pareschi, et al., 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. https://doi.org/10.1142/S0218202519500374 doi: 10.1142/S0218202519500374
![]() |
[5] |
A. Attanasi, A. Cavagna, L. Del Castello, I. Giardina, A. Jelic, S. Melillo, et al., Information transfer and behavioural inertia in starling flocks, Nat. Phys., 10 (2014), 691–696. https://doi.org/10.1038/nphys3035 doi: 10.1038/nphys3035
![]() |
[6] |
J. Buck, E. Buck, Biology of synchronous flashing of fireflies, Nature, 211 (1966), 562–564. https://doi.org/10.1038/211562a0 doi: 10.1038/211562a0
![]() |
[7] |
J. A. Carrillo, M. Fornasier, J. Rosado, G. Toscani, Asymptotic flocking dynamics for the kinetic Cucker–Smale model, SIAM J. Math. Anal., 42 (2010), 218–236. https://doi.org/10.1137/090757290 doi: 10.1137/090757290
![]() |
[8] |
P. Cattiaux, F. Delebecque, L. Pedeches, Stochastic Cucker–Smale models: old and new, Ann. Appl. Probab., 28 (2018), 3239–3286. https://doi.org/10.1214/18-AAP1400 doi: 10.1214/18-AAP1400
![]() |
[9] |
A. Cavagna, A. Cimarelli, I. Giardina, G. Parisi, R. Santagati, F. Stefanini, et al., Scale-free correlations in starling flocks, Proc. Natl. Acad. Sci., 107 (2010), 11865–11870. https://doi.org/10.1073/pnas.1005766107 doi: 10.1073/pnas.1005766107
![]() |
[10] |
H. Cho, J. G. Dong, S. Y. Ha, Emergent behaviors of a thermodynamic Cucker–Smale flock with a time-delay on a general digraph, Math. Methods Appl. Sci., 45 (2021), 164–196. https://doi.org/10.1002/mma.7771 doi: 10.1002/mma.7771
![]() |
[11] |
S. H. Choi, S. Y. Ha, Interplay of the unit-speed constraint and time-delay in Cucker–Smale flocking, J. Math. Phys., 59 (2018), 082701. https://doi.org/10.1063/1.4996788 doi: 10.1063/1.4996788
![]() |
[12] |
S. H. Choi, S. Y. Ha, Emergence of flocking for a multi-agent system moving with constant speed, Commun. Math. Sci., 14 (2016), 953–972. https://doi.org/10.4310/CMS.2016.v14.n4.a4. doi: 10.4310/CMS.2016.v14.n4.a4
![]() |
[13] |
Y. P. Choi, J. Haskovec, Cucker–Smale model with normalized communication weights and time-delay, Kinet. Relat. Models, 10 (2017), 1011–1033. https://doi.org/10.3934/krm.2017040 doi: 10.3934/krm.2017040
![]() |
[14] |
Y. P. Choi, Z. Li, Emergent behavior of Cucker–Smale flocking particles with heterogeneous time-delays, Appl. Math. Lett., 86 (2018), 49–56. https://doi.org/10.1016/j.aml.2018.06.018 doi: 10.1016/j.aml.2018.06.018
![]() |
[15] |
J. Cho, S. Y. Ha, F. Huang, C. Jin, D. Ko, Emergence of bi-cluster flocking for agent-based models with unit speed constraint, Anal. Appl., 14 (2016), 39–73. https://doi.org/10.1142/S0219530515400023 doi: 10.1142/S0219530515400023
![]() |
[16] |
K. Cooke, Z. Grossman, Discrete delay, distributed delay and stability switches, J. Math. Anal. Appl., 86 (1982), 592–627. https://doi.org/10.1016/0022-247X(82)90243-8 doi: 10.1016/0022-247X(82)90243-8
![]() |
[17] |
F. Cucker, S. Smale, Emergent behavior in flocks, IEEE Trans. Automat. Contr., 52 (2007), 852–862. https://doi.org/10.1109/TAC.2007.895842 doi: 10.1109/TAC.2007.895842
![]() |
[18] |
P. Degond, S. Motsch, Large-scale dynamics of the persistent turning walker model of fish behavior, J. Stat. Phys., 131 (2008), 989–1022. https://doi.org/10.1007/s10955-008-9529-8 doi: 10.1007/s10955-008-9529-8
![]() |
[19] |
G. B. Ermentrout, An adaptive model for synchrony in the firefly Pteroptyx malaccae, J. Math. Biol., 29 (1991), 571–585. https://doi.org/10.1007/BF00164052 doi: 10.1007/BF00164052
![]() |
[20] |
E. Ferrante, A. E. Turgut, A. Stranieri, C. Pinciroli, M. Dorigo, Self-organized flocking with a mobile robot swarm: a novel motion control method, Adapt. Behav., 20 (2012), 460–477. https://doi.org/10.1177/1059712312462248 doi: 10.1177/1059712312462248
![]() |
[21] |
A. Figalli, M. Kang, A rigorous derivation from the kinetic Cucker–Smale model to the pressureless Euler system with nonlocal alignment, Anal. PDE., 12 (2019), 843–866. https://doi.org/10.2140/apde.2019.12.843 doi: 10.2140/apde.2019.12.843
![]() |
[22] |
S. Y. Ha, D. Ko, Y. Zhang, Remarks on the critical coupling strength for the Cucker-Smale model with unit speed, Discrete Contin. Dyn. Syst., 38 (2018), 2763–2793. https://doi.org/10.3934/dcds.2018116 doi: 10.3934/dcds.2018116
![]() |
[23] |
S. Y. Ha, T. Ruggeri, Emergent dynamics of a thermodynamically consistent particle model, Arch. Ration. Mech. Anal., 223 (2017), 1397–1425. https://doi.org/10.1007/s00205-016-1062-3 doi: 10.1007/s00205-016-1062-3
![]() |
[24] |
S. Y. Ha, E. Tadmor, From particle to kinetic and hydrodynamic descriptions of flocking, Kinet. Relat. Models, 1 (2008), 415–435. https://doi.org/10.3934/krm.2008.1.415 doi: 10.3934/krm.2008.1.415
![]() |
[25] |
J. Hale, N. Sternberg, Onset of chaos in differential delay equations, J. Comput. Phys., 77 (1988), 221–239. https://doi.org/10.1016/0021-9991(88)90164-7 doi: 10.1016/0021-9991(88)90164-7
![]() |
[26] |
T. K. Karper, A. Mellet, K. Trivisa, Hydrodynamic limit of the kinetic Cucker–Smale flocking model, Math Models Methods Appl Sci, 25 (2015), 131–163. https://doi.org/10.1142/S0218202515500050 doi: 10.1142/S0218202515500050
![]() |
[27] |
S. Motsch, E. Tadmor, A new model for self-organized dynamics and its flocking behavior, J. Stat. Phys., 141 (2011), 923–947. https://doi.org/10.1007/s10955-011-0285-9 doi: 10.1007/s10955-011-0285-9
![]() |
[28] | E. A. Ok, Real Analysis with Economics Applications, Princeton University Press, Princeton, 2007,306. https://doi.org/10.1515/9781400840892 |
[29] |
R. Olfati-Saber, Flocking for multi-agent dynamic systems: algorithms and theory, IEEE Trans. Automat. Contr., 51 (2006), 401–420. https://doi.org/10.1109/TAC.2005.864190 doi: 10.1109/TAC.2005.864190
![]() |
[30] | A. Pikovsky, M. Rosenblum, J. Kurths, Synchronization: A universal concept in nonlinear sciences, Cambridge University Press, Cambridge, 2001. https://doi.org/10.1119/1.1475332 |
[31] |
T. Ruggeri, S. Simić, On the Hyperbolic System of a Mixture of Eulerian Fluids: A Comparison Between Single and Multi-Temperature Model, Math. Methods Appl. Sci., 30 (2007), 827–849. https://doi.org/10.1002/mma.813 doi: 10.1002/mma.813
![]() |
[32] |
S. H. Strogatz, From Kuramoto to Crawford: exploring the onset of synchronization in populations of coupled oscillators, Physica D, 143 (2000), 1–20. https://doi.org/10.1016/S0167-2789(00)00094-4 doi: 10.1016/S0167-2789(00)00094-4
![]() |
[33] |
J. Toner, Y. Tu, Flocks, herds, and schools: A quantitative theory of flocking, Phys. Rev. E, 58 (1998), 4828–4858. https://doi.org/10.1103/PhysRevE.58.4828 doi: 10.1103/PhysRevE.58.4828
![]() |
[34] |
C. M. Topaz, A. L. Bertozzi, Swarming patterns in a two-dimensional kinematic model for biological groups, SIAM J. Appl. Math., 65 (2004), 152–174. https://doi.org/10.1137/S0036139903437424 doi: 10.1137/S0036139903437424
![]() |
[35] |
T. Vicsek, A. Czirók, E. Ben-Jacob, I. Cohen, O. Schochet, Novel type of phase transition in a system of self-driven particles, Phys. Rev. Lett. 75 (1995), 1226–1229. https://doi.org/10.1103/PhysRevLett.75.1226 doi: 10.1103/PhysRevLett.75.1226
![]() |
[36] |
T. Vicsek, A. Zefeiris, Collective motion, Phys. Rep., 517 (2012), 71–140. https://doi.org/10.1016/j.physrep.2012.03.004 doi: 10.1016/j.physrep.2012.03.004
![]() |
[37] | A. T. Winfree, The geometry of biological time, Springer, New York, 1980. https://doi.org/10.1007/978-1-4757-3484-3 |
[38] |
A. T. Winfree, Biological rhythms and the behavior of populations of coupled oscillators, J. Theor. Biol., 16 (1967), 15–42. https://doi.org/10.1016/0022-5193(67)90051-3 doi: 10.1016/0022-5193(67)90051-3
![]() |