
We study dynamical behaviors of the ensemble of thermomechanical Cucker-Smale (in short TCS) particles with singular power-law communication weights in velocity and temperatures. For the particle TCS model, we present several sufficient frameworks for the global regularity of solution and a finite-time breakdown depending on the blow-up exponents in the power-law communication weights at the origin where the relative spatial distances become zero. More precisely, when the blow-up exponent in velocity communication weight is greater than unity and the blow-up exponent in temperature communication weights is more than twice of blow-up exponent in velocity communication, we show that there will be no finite time collision between particles, unless there are collisions initially. In contrast, when the blow-up exponent of velocity communication weight is smaller than unity, we show that there can be a collision in finite time. For the kinetic TCS equation, we present a local-in-time existence of a unique weak solution using the suitable regularization and compactness arguments.
Citation: Young-Pil Choi, Seung-Yeal Ha, Jeongho Kim. Propagation of regularity and finite-time collisions for the thermomechanical Cucker-Smale model with a singular communication[J]. Networks and Heterogeneous Media, 2018, 13(3): 379-407. doi: 10.3934/nhm.2018017
[1] | 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 |
[2] | 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 |
[3] | 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 |
[4] | Young-Pil Choi, Seung-Yeal Ha, Jeongho Kim . Propagation of regularity and finite-time collisions for the thermomechanical Cucker-Smale model with a singular communication. Networks and Heterogeneous Media, 2018, 13(3): 379-407. doi: 10.3934/nhm.2018017 |
[5] | Hyunjin Ahn, Woojoo Shim . Interplay of a unit-speed constraint and time-delay in the flocking model with internal variables. Networks and Heterogeneous Media, 2024, 19(3): 1182-1230. doi: 10.3934/nhm.2024052 |
[6] | 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 |
[7] | 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 |
[8] | 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 |
[9] | 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 |
[10] | 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 |
We study dynamical behaviors of the ensemble of thermomechanical Cucker-Smale (in short TCS) particles with singular power-law communication weights in velocity and temperatures. For the particle TCS model, we present several sufficient frameworks for the global regularity of solution and a finite-time breakdown depending on the blow-up exponents in the power-law communication weights at the origin where the relative spatial distances become zero. More precisely, when the blow-up exponent in velocity communication weight is greater than unity and the blow-up exponent in temperature communication weights is more than twice of blow-up exponent in velocity communication, we show that there will be no finite time collision between particles, unless there are collisions initially. In contrast, when the blow-up exponent of velocity communication weight is smaller than unity, we show that there can be a collision in finite time. For the kinetic TCS equation, we present a local-in-time existence of a unique weak solution using the suitable regularization and compactness arguments.
Collective behaviors of classical and quantum many-body systems are ubiquitous in our biological, chemical and physical complex systems in nature, e.g., flocking of birds, swarming of fish, aggregation of bacteria, or synchronization of neurons, synchronization of Josephson junction arrays, etc. [28,29,30]. In previous literature, many mathematical models were proposed to describe these collective dynamics. Among those models, our main interest in this paper lies on the thermomechanical Cucker-Smale (in short, TCS) model which is recently introduced in [19]. Consider an ensemble consisting of
$ dxidt=vi,t>0,i=1,2,⋯,N,dvidt=κ1NN∑j=1ϕ(rij)(vjθj−viθi),dθidt=κ2NN∑j=1ζ(rij)(1θi−1θj), $ | (1) |
subject to initial data:
$ (x_i(0), v_i(0), \theta_i(0)) = :(x_i^0, v_i^0, \theta_i^0), \;\;\; \sum\limits_{i = 1}^N v_i^0 = 0. $ | (2) |
Here
$ \phi(s): = \frac{1}{s^\alpha}, \;\;\; \zeta(s) = \frac{1}{s^\beta}, \;\;\; \text{with} \;\;\; \alpha, \beta > 0, \;\;\; s > 0. $ | (3) |
As its name suggests, TCS model is one of generalizations for the Cucker-Smale (in short, C-S) model introduced in [9]:
$ dxidt=vi,t>0,i=1,2,⋯,N,dvidt=κNN∑j=1ϕ(rij)(vj−vi). $ | (4) |
Note that for the same constant temperatures
$ ∂tf+∇x⋅(vf)+∇v⋅(F[f]f)+∂θ(G[f]f)=0,x,v∈Rd,θ∈R+,t>0,F[f](z,t):=−κ1∫R2d×R+ϕ(|x−x∗|)(vθ−v∗θ∗)f(z∗,t)dz∗,G[f](x,θ,t):=κ2∫R2d×R+ζ(|x−x∗|)(1θ−1θ∗)f(z∗,t)dz∗,z=(x,v,θ),dz:=dxdvdθ, $ | (5) |
subject to the initial condition:
$ \label{6} f(z, 0) = : f^0(z), \;\;\; z = (x, v, \theta)\in \mathbb{R}^{2d}\times\mathbb{R}_+. $ | (6) |
Below, we briefly discuss our main issues to be explored throughout the paper. In this paper, we are interested in the following two questions for the Cauchy problems (1) - (2) and (5) - (6):
● (Q1) : Can the strong singularity in
● (Q2) : Can we establish the well-posedness of the kinetic TCS equation at least locally in time?
Next, we comment on the above questions. The collision avoidance between particles has been studied for the Cucker-Smale model and its variants in [1,8,3,26,25] due to the possible applications of the C-S model (4) to the traffic control of unmanned aerial vehicles and robot systems. More precisely, some special class of initial configurations are taken to guarantee collision avoidance for strong communication weight with
The main novelty of this paper are three-fold. First, we show that the particle TCS model (1) - (3) with a strong communication weights
$ \|X\|_{[l]}(t) : = \sqrt{\sum\limits_{i, j\in[l]}|x_i(t)-x_j(t)|^2} \;\;\; \text{and} \;\;\; \|V\|_{[l]}(t): = \sqrt{\sum\limits_{i, j\in[l]}|v_i(t)-v_j(t)|^2}, $ |
Then, we will show that
$ \label{7} |\Phi(\|X\|_{[l]}(t))|\le \int_0^t\phi(\|X\|_{[l]}(\tau))\|V\|_{[l]}(\tau)d\tau +|\Phi(\|X\|_{[l]}(0))|. $ | (7) |
where
$ \label{8} \lim\limits_{r \to 0+} \Phi(r) = \infty, \;\;\; \text{for} \alpha \geq 1. $ | (8) |
By laborious estimates, we can show that the R.H.S. of (7) is finite for \; all
$ \lim\limits_{t \to t_0-} |\Phi(\|X\|_{[l]}(t))| = \infty $ |
which gives a contradiction. As long as there are no collisions, the R.H.S. of (1) is still locally Lipschitz. Hence, the classical Cauchy-Lipschitz theory can be applied to (1) - (3), and this yields the global smooth solutions. Second, for weakly singular communication
$ W_1(\mu_1, \mu_2): = \inf\limits_{\gamma\in \Gamma(\mu_1, \mu_2)} \int_{\mathbb{R}^{2d}}|x-y| d\gamma(x, y), $ |
for two probability measures
The rest of this paper is organized as follows. In Section 2, we provide a several preliminaries for the TCS model, which will be crucially used in the later sections. In Section 3, we briefly review some relevant results for the C-S model and present our main results on the global existence of solutions and local existence of weak solutions for the particle system (1)-(2). In Section 4, we provide the conditions on the communication weight
Notation: Throughout this paper,
$ \|f\|_{L^1\cap L^p}: = \|f\|_{L^1}+\|f\|_{L^p},\;\;\; \text{and} \;\;\; \|f\|: = \|f\|_{L^\infty(0, \tau;L^1\cap L^p)}. $ |
Moreover, we will use the notation of volume element
In this section, we briefly review theoretic minimum of the TCS model and kinetic TCS equation and Wasserstein distances.
First, we study a priori estimates for the particle TCS model (1) which will be crucially used in later sections. For position, velocity and temperature configurations
$ D(X(t)): = \max\limits_{1\le i, j\le N}|x_i(t)-x_j(t)|,\;\;\; D(V(t)): = \max\limits_{1\le i, j\le N}|v_i(t)-v_j(t)|, $ |
and
$ D({\Theta}(t)): = \max\limits_{1\le i, j\le N}|\theta_i(t)-\theta_j(t)|. $ |
The most basic property of system (1) is the monotonicity and boundedness of temperature, position and velocity diameters.
Lemma 2.1. For a positive constant
1. The temperature diameter
$D(\Theta(t)) \le D(\Theta(s)), \;\;\; 0\le s\le t.$ |
2. The velocity and position diameters are bounded:
$ \sup\limits_{0 \leq t \leq T} D(V(t)) \le C_T, \;\;\; \sup\limits_{0\le t\le T} D(X(t))\le \tilde{C}_T$ |
where
Proof. (1) For a given
$\theta_M(t): = \max\limits_{1\le i\le N}\theta_i(t), \;\;\; \theta_m(t): = \min\limits_{1\le i\le N}\theta_i(t).$ |
Then, it follows from
$ \frac{d\theta_M}{dt} = \frac{\kappa_2}{N}\sum\limits_{j = 1}^N \zeta(r_{Mj})(\frac{1}{\theta_M}-\frac{1}{\theta_j}) \leq 0, $ |
and
$ \frac{d\theta_m}{dt} = \frac{\kappa_2}{N}\sum\limits_{j = 1}^N \zeta(r_{mj})(\frac{1}{\theta_m}-\frac{1}{\theta_j}) \geq 0. $ |
These yield
$ \theta_m(s) \le \theta_m(t)\le \theta_i(t)\le \theta_M(t)\le \theta_M(s) , \;\;\; t \geq s \geq 0.$ |
Then, we have
$ D(\Theta(t)) \leq D(\Theta(s)), \;\;\; t \geq s \geq 0. $ |
(2) It follows from the differential inequalities in [12,Lemma 3.2] that we have
$ \frac{d}{dt}D(V) \le -\frac{\kappa_1\phi(D_X)}{\theta_M^0}D_V+2\kappa_1\frac{D(\Theta)}{(\theta_m^0)^2}D(V) \le 2\kappa_1\frac{D(\Theta^0)}{(\theta_m^0)^2}D(V). $ |
Then, Gronwall's lemma yields
$ D(V(t)) \leq D(V^0) \exp(2\kappa_1\frac{D(\Theta^0)}{(\theta_m^0)^2} T ) = : C_T.$ |
Finally, note that the following differential inequality in [12,Lemma 3.2] also holds:
$\frac{d}{dt}D(X)\le D(V).$ |
Then, we use the boundedness of velocity diameter to get
$D(X(t))\le D(X(0))+C_TT, $ |
which implies the boundedness of diameter of position.
Next, for position and temperature vectors
$ \label{B-3-1} {\mathcal P}(t) : = {\mathcal P}(X(t), \Theta(t)) = \sum\limits_{i, j = 1}^N\zeta(r_{ij}(t))\frac{|\theta_i(t)-\theta_j(t)|^2}{\theta_i(t)\theta_j(t)}. $ |
Lemma 2.2. Suppose that the coupling strength
$ \int_0^\infty {\mathcal P}(t)dt \leq \frac{N}{\kappa_2} \sum\limits_{i = 1}^N|\theta_i^0|^2. $ |
Proof. We multiply (1)
$ddtN∑i=1|θi|2=2κ2NN∑i,j=1ζ(rij)(1θi−1θj)θi=2κ2NN∑i,j=1ζ(rij)(θj−θiθiθj)θi=2κ2NN∑i,j=1ζ(rij)(θi−θjθjθi)θj=−κ2NN∑i,j=1ζ(rij)|θi−θj|2θiθj=−κ2NP.$ |
We integrate the above relation in time to get
$\sum\limits_{i = 1}^N|\theta_i(t)|^2+\frac{\kappa_2}{N}\int_0^t {\mathcal P}(s)ds = \sum\limits_{i = 1}^N|\theta_i^0|^2, $ |
which yields the desired estimate.
Next, we study the propagation of velocity and temperature moments along the kinetic TCS equation. For this, we set
$⟨1⟩:=∫R2d×R+fdz,⟨v⟩:=∫R2d×R+vfdz,⟨v2⟩:=∫R2d×R+|v|2fdz,⟨θ⟩:=∫R2d×R+θfdz,⟨θ2⟩:=∫R2d×R+θ2fdz,and⟨logθ⟩:=∫R2d×R+(logθ)fdz.$ |
In next lemma, we study propagation of above moments along (5).
Lemma 2.3. Let
$ \frac{d\langle 1 \rangle}{dt} = 0, \;\;\; \frac{d\langle v \rangle}{dt} = 0, \;\;\; \frac{d\langle \theta \rangle}{dt} = 0, \;\;\; \frac{d\langle \theta^2 \rangle}{dt}\le0, \;\;\; and \;\;\; \frac{d\langle \log \theta \rangle}{dt}\ge 0. $ |
Proof. (ⅰ) The conservation of mass follows from the divergence form of (5).
(ⅱ) We multiply (5) by
$d⟨v⟩dt=∫R2d×R+v∂tfdz=−∫R2d×R+v(v⋅∇xf)+v∇v⋅(F[f]f)−v∂θ(G[f]f)dz=d∫R2d×R+F[f]fdz=dκ1∫R4d×(R+)2ϕ(|x−x∗|)(v∗θ∗−vθ)f(z,t)f(z∗,t)dzdz∗=0.$ |
The case for
(ⅲ) We multiply (5) by
$d⟨θ2⟩dt=−∫R2d×R+θ2v⋅∇xf+θ2∇v⋅(F[f]ff)+θ2∂θ(G[f]f)dz=2∫R2d×R+θG[f]fdz=2∫R4d×(R+)2ζ(|x−x∗|)(1θ−1θ∗)θf(z,t)f(z∗,t)dzdz∗=−∫R4d×(R+)2ζ(|x−x∗|)(θ−θ∗)2θθ∗f(z,t)f(z∗,t)dzdz∗≤0.$ |
(ⅳ) Finally, we can estimate
$d⟨logθ⟩dt=−∫R2d×R+(logθ)v⋅∇xf+(logθ)∇v⋅(F[f]f)+(logθ)∂θ(G[f]f)dz=∫R2d×R+1θG[f]fdz=∫R4d×(R+)2ζ(|x−x∗|)(1θ−1θ∗)1θf(z,t)f(z∗,t)dzdz∗=12∫R4d×(R+)2ζ(|x−x∗|)(1θ−1θ∗)2f(z,t)f(z∗,t)dzdz∗≥0.$ |
In this subsection, we introduce forward characteristic curves associated with the kinetic TCS equation (5). First, we define forward characteristics
$ \label{9} {dx(s)ds=v(s),s>0,dv(s)ds=F[f](x(s),v(s),θ(s)),dθ(s)ds=G[f](x(s),θ(s)),(x(0),v(0),θ(0))=(x,v,θ). $ | (9) |
Moreover, we also define the section of support of
$Ωx(t):=¯{x∈Rd|f(x,v,θ,t)≠0},Ωv(t):=¯{v∈Rd|f(x,v,θ,t)≠0},Ωθ(t):=¯{θ∈R+|f(x,v,θ,t)≠0}.$ |
Lemma 2.4. Let
$\Omega_\theta(s)\subseteq \Omega_\theta(t), \;\;\; 0 \leq t \leq s.$ |
Proof. In fact, the maximum value of
$12dθ(s)2ds=θ(s)dθ(s)ds=∫R2d×R+ζ(|x(s)−x∗|)(1θ(s)−1θ∗)θ(s)f(z∗,t)dz∗=∫R2d×R+ζ(|x(s)−x∗|)(θ∗−θ(s)θ(s)θ∗)θ(s)f(z∗,t)dz∗≤0,$ |
where we use
In this section, we provide some basic properties of Wasserstein metric between two measures for later use. First, we begin with definition of push-forward measure in the following definition.
Definition 2.5. Let
$\mu_2(B) = \mu_1(f^{-1}(B)), \;\;\; B\subset \mathbb{R}^d.$ |
Next, we list several results about push-forward measure and Wasserstein metric without proofs.
Proposition 1. [31,32] The following assertions hold.
(1) Let
$\mu_2 = f\#\mu_1 \Longleftrightarrow \int_{\mathbb{R}^d}\phi\, d\mu_2 = \int_{\mathbb{R}^d}\phi \circ f d\mu_1, \;\;\; \forall \phi\in C_b(\mathbb{R}^d).$ |
(2) Suppose that
$W_p^p(f_1\#\mu_0, f_2\#\mu_0)\le \int_{\mathbb{R}^d\times\mathbb{R}^d}|x-y|^pd\gamma(x, y) = \int_{\mathbb{R}^d}|f_1(x)-f_2(x)|^pd\mu_0(x), $ |
where
(3) Let
$\int_{\mathbb{R}^d}|x|d\mu_n(x)\to\int_{\mathbb{R}^d}|x|d\mu(x), \;\;\; as \;\;\; n\to \infty.$ |
Note that the space
$W_1(\mu_1, \mu_2) = \sup\Big\{\int_{\mathbb{R}^d}\phi(x)d(\mu_1-\mu_2)(x) \Big| \phi\in \text{Lip}(\mathbb{R}^d, \mathbb{R}), \; \|\phi\|_{Lip}\le1\Big\}, $ |
where
In this section, we briefly present our main results on the collision avoidance and local well-posedness of the particle TCS model and kinetic TCS equation, respectively.
First, we briefly review the previous results [3,17,24,25,26] on the particle and kinetic C-S models with power-law communication weights from three perspectives: " flocking dynamics, non-existence and it local-in-time well-posedness ". As far as the authors know, the particle C-S model with a singular communication weight was first treated in [17] to study the flocking dynamics and then the global regularity and emergence of finite-time collisions has been established in a series of works by Peszek and his collaborators [3,25,26]. In a recent work [24], Mucha and Peszek studied the existence of measure-valued solutions and weak-atomic uniqueness for the kinetic C-S equation with singular communication weights. In the sequel, we explicitly state our main results for the particle and kinetic TCS models. The detailed proofs will be given in the later sections.
In this subsection, we provide two results on the collision avoidance and asymptotic mono-cluster flocking of the particle TCS model.
Our first result deals with the collision avoidance and finite-time collisions of the particle model (1) depending on the blow-up exponents at the singular point of the communication weight
Theorem 3.1. (Collision avoidance and collisions) The following assertions hold.
1. Suppose that the blow-up exponents
$ 1 \leq \alpha \leq \frac{\beta}{2} \;\;\; and \;\;\; x_i^0 \neq x_j^0 \;\;\; for \; all \;\;\; 1 \leq i \neq j \leq N. $ |
Then, there exists a unique global solution
$ x_i(t) \neq x_j(t) \;\;\; for \; all \;\;\; 1 \leq i \neq j \leq N \;\;\; and \;\;\; t \geq 0. $ |
2. Suppose that the exponent
$ 0 < \alpha < 1 \;\;\; and \;\;\; N = 2, \;\;\; d = 1. $ |
Then, there exists initial data
$ x_1(t_c) = x_2(t_c), \;\;\; for \; some\; time \;t_c > 0. $ |
Remark 1. For the second assertion, note that if all initial temperatures are equal with
As a direct corollary of Theorem 3.1, we have the following equivalence relation.
Corollary 1. Let
(i) The exponents
(ii) For any
Next, we introduce a concept of mono-cluster flocking for the TCS model (1) in the following definition.
Definition 3.2. [19] Let
$ \sup\limits_{0\leq t < \infty}|x_i(t) -x_j(t)| < \infty, \;\;\; \lim\limits_{t\rightarrow\infty}(|v_i(t) -v_j(t)|+| \theta_i(t) -\theta_j(t)|) = 0. $ |
Remark 2. The mono-cluster flocking dynamics for the particle, kinetic, and hydrodynamic C-S models have been studied in [9,11,15,17,18,20,21,22,23,27].
Our second result is concerned about the emergent dynamics of the TCS model (1) - (2) with singular communications.
Theorem 3.3. (Emergence of mono-cluster flocking) Suppose that the exponents
$ 1≤α≤β2,x0i≠x0jforall1≤i≠j≤N,D(Θ0)≤(θ0m)2ϕ(D(X0))2Cθ0M,D(V0)≤κ12θ0M∫D∞XD(X0)ϕ(s)ds,83Cϕ(D(X0))≤ϕ(D∞X)<ϕ(D(X0)), $ | (10) |
where
$sup0≤t<∞D(X(t))≤D∞X,D(Θ(t))≤D(Θ0)e−κ2(θ0M)2ζ(D∞X)t,D(V(t))≤D(V0)exp(−κ1ϕ(D∞X)θ0Mt+2κ1(θ0M)2D(Θ0)κ2(θ0m)2ζ(D∞X)),∀t≥0. $ |
Remark 3. Note that the existence of global solutions is guaranteed by Theorem 3.1 at least under the conditions in (10) in the interactions kernels
We first recall the definition of weak solutions to (5) in the following definition.
Definition 3.4 (Weak solution) For a given finite
1.
2. For any
$ ∫R2d×R+f(z,τ)Φ(z,τ)dz−∫R2d×R+f0(z)Φ(z,0)dz=∫τ0∫R2d×R+f(∂tΦ+v⋅∇xΦ+F[f]⋅∇vΦ+G[f]∂θΦ)dzdt. $ | (11) |
Then, our third result is concerned with local-in-time existence of weak solutions.
Theorem 3.5. Suppose that
$1≤p≤∞,0<α,β<dp∗−1,f0∈(L1+∩Lp∩P1)(R2d×R+),suppvf0(x,⋅,θ)⊂Rd,suppθf0(x,v,⋅)⊂R+,foreachx,v∈Rd,θ∈R+,$ |
where
$ \sup\limits_{0 \leq t \leq \tau}W_1(f_1(t), f_2(t)) \leq G_0 W_1(f_1^0, f_2^0), $ |
where
In this section, we provide the detailed proof of Theorem 3.1. For this, we use a similar strategy to that in [3] which is based on the construction of a system of locally dissipative differential inequalities (SDDI) for quantifying collisions between particles. Note that this idea is proposed in [17] to show the flocking behavior of the original Cucker-Smale model.
For a fixed
$\lim\limits_{t\to t_0-}(x_i(t)-x_j(t)) = 0.$ |
Next, we will consider colliding particles and non-colliding particles separately. For the colliding particles, we will use the standard index changing trick
$ [l]: = \lbrace j\in \{1, \cdots, N\} | r_{jl}\to 0 \;\; \text{as} \; t\to t_0^- \rbrace, $ |
i.e.,
$ r_{jl}(t)\ge \delta > 0 \;\;\; \text{in}\; [0, t_0) \; \text{for all} \; j\notin[l]. $ |
Without loss of generality, we may assume
$ \|X\|_{[l]}(t): = \sqrt{\sum\limits_{i, j\in[l]}|x_i(t)-x_j(t)|^2} \;\;\; \text{and} \;\;\; \|V\|_{[l]}(t): = \sqrt{\sum\limits_{i, j\in[l]}|v_i(t)-v_j(t)|^2}. $ | (12) |
Note that
Lemma 4.1. Suppose that the exponents
$ 1\le \alpha\le \frac{\beta}{2}, $ |
and let
$ ddt‖X‖[l]≤‖V‖[l],0<t<t0,ddt‖V‖[l]≤−C1ϕ(‖X‖[l])‖V‖[l]+C2√P(X,Θ)+C3, $ | (13) |
where
Proof. ● Step A (Derivation of (13)
$ \frac{d}{dt}\|X\|_{[l]}\le \|V\|_{[l]}. $ |
● Step B (Derivation of (13)
$ddt‖v‖2[l]=2∑i,j∈[l]⟨vi−vj,κ1NN∑k=1ϕ(rki)(vkθk−viθi)−κ1NN∑k=1ϕ(rkij)(vkθk−vjθj)⟩=2κ1N(∑i,j,k∈[l]+∑i,j∈[l]k∉[l])[ϕ(rki)⟨vkθk−viθi,vi−vj⟩−ϕ(rkj)⟨vkθk−vjθj,vi−vj⟩]=:I11+I12.$ |
$I11=2κ1N∑i,j,k∈[l]ϕ(rki)⟨vkθk−viθi,vi−vj⟩=−2κ1N∑i,j,k∈[l]ϕ(rki)⟨vkθk−viθi,vk−vj⟩=κ1N∑i,j,k∈[l]ϕ(rki)⟨vkθk−viθi,vi−vk⟩=−κ1|[l]|N∑i,j∈[l]ϕ(rij)⟨viθi−vjθj,vi−vj⟩=−κ1|[l]|N∑i,j∈[l]ϕ(rij)θi|vi−vj|2+κ1|[l]|N∑i,j∈[l]ϕ(rij)θi−θjθiθj⟨vj,vi−vj⟩≤−κ1|[l]|N∑i,j∈[l]ϕ(rij)θi|vi−vj|2+κ1|[l]|N(∑i,j∈[l]ϕ2(rij)|θi−θj|2θ2iθ2j|vj|2)12‖V‖[l],$ |
where we used the relation:
$ \frac{v_i}{\theta_i} - \frac{v_j}{\theta_j} = \frac{v_i - v_j}{\theta_i} + \Big( \frac{1}{\theta_i} - \frac{1}{\theta_j} \Big) v_j = \frac{v_i - v_j}{\theta_i} + \frac{(\theta_j - \theta_i)}{\theta_i \theta_j} v_j. $ |
We use Lemma 2.1 and the assumption
$ |v_j| = \frac{1}{N}\vert \sum\limits_{k = 1}^N(v_j-v_k) \vert\le\frac{1}{N}\sum\limits_{k = 1}^N|v_j-v_k|\le D(V)\le C_{t_0}, \;\;\;\text{and} \;\;\; \phi^2\le \zeta $ | (14) |
for
$ \mathcal{I}_{11}\le -\frac{\kappa_1|[l]|}{N}\sum\limits_{i, j\in[l]}\frac{\phi(r_{ij})}{\theta_i}|v_i-v_j|^2+\frac{\kappa_1C_{t_0}^2|[l]|}{T_m^0}{\left( {\underbrace {\sum\limits_{i,j = 1}^N \zeta ({r_{ij}})\frac{{|{\theta _i} - {\theta _j}{|^2}}}{{{\theta _i}{\theta _j}}}}_{ ={\mathcal P} (X,\Theta )}} \right)^{1/2}} \|V\|_{[l]}. $ |
Now, we define positive constants
$ C_1: = \frac{\kappa_1|[l]|}{2N \theta_M^0} \;\;\; \text{and} \;\;\; C_2: = \frac{\kappa_1C_{t_0}^2|[l]|}{2\theta_m^0}. $ |
This gives
$ \mathcal{I}_{11}\le -2C_1\sum\limits_{i, j\in[l]}\phi(r_{ij})|v_i-v_j|^2+2C_2 {\mathcal P}^{1/2}\|V\|_{[l]}. $ |
● (Estimation for
$I12=2κ1N∑i,j∈[l]k∉[l][ϕ(rki)⟨vkθk−viθi,vi−vj⟩−ϕ(rkj)⟨vkθk−vjθj,vi−vj⟩]=2κ1N∑i,j∈[l]k∉[l][ϕ(rki)⟨vjθj−viθi,vi−vj⟩+(ϕ(rki)−ϕ(rkj))⟨vkθk−vjθj,vi−vj⟩]=:I121+I122.$ |
$ I121=−2κ1N∑i,j∈[l]k∉[l]ϕ(rki)θi|vi−vj|2+2κ1N∑i,j∈[l]k∉[l]ϕ(rki)θi−θjθiθj⟨vj,vi−vj⟩≤2κ1N∑i,j∈[l]k∉[l]ϕ(rki)θi−θjθiθj⟨vj,vi−vj⟩≤2κ1(N−|[l]|)Nδα(θ0m)2D(Θ)D(V)∑i,j∈[l]|vi−vj|≤2κ1(N−|[l]|)Nδα(θ0m)2D(Θ0)Ct0|[l]|‖V‖[l], $ | (15) |
where we used Cauchy-Schwarz inequality in the last inequality.
$ I122≤2κ1N∑i,j∈[l]k∉[l]|ϕ(rki)−ϕ(rkj)||⟨vkθk−vjθj,vi−vj⟩|≤2κ1N∑i,j∈[l]k∉[l]|ϕ(rki)−ϕ(rkj)||vk−vjθk+vj(1θk−1θj)||vi−vj|≤2κ1N∑i,j∈[l]k∉[l]Lδ|xi−xj|(D(V)θ0m(1+D(Θ)θ0m))|vi−vj|≤2κ1Lδ(N−|[l]|)NCt0θ0m(1+D(Θ0)θ0m)∑i,j∈[l]|xi−xj||vi−vj|≤2κ1Lδ(N−|[l]|)NCt0θ0m(1+D(Θ0)θ0m)˜Ct0|[l]|‖V‖[l], $ | (16) |
where
$ L_\delta : = ||\phi||_{\text{Lip}(|r| \geq \delta)}. $ |
Now we combine (15) and (16) to obtain
$I12≤2κ1(N−|[l]|)Nδα(θ0m)2D(Θ0)Ct0|[l]|‖V‖[l]+2κ1Lδ(N−|[l]|)NCt0θ0m(1+D(Θ0)θ0m)˜Ct0|[l]|‖V‖[l]=:2C3‖V‖[l].$ |
Finally, we combine all the above estimates to get
$ \frac{d}{dt}\|V\|_{[l]}^2\le \mathcal{I}_{11}+\mathcal{I}_{12}\le -2C_1\phi(\|x\|_{[l]})\|V\|_{[l]}^2+2C_2 {\mathcal P}^{1/2}\|V\|_{[l]}+2C_3\|V\|_{[l]}, $ |
or equivalently,
$ \frac{d}{dt}\|V\|_{[l]}\le -C_1\phi(\|X\|_{[l]})\|V\|_{[l]}+C_2 \sqrt{ {\mathcal P}}+C_3. $ |
We are now ready to provide the proof of the first part of Theorem 3.1. We apply Grönwall's inequality for
$ ‖V‖[l](t)≤‖V‖[l](s)e−C1∫tsϕ(‖X‖[l](τ))dτ+∫ts(C2√P(τ)+C3)e−C1∫tτϕ(‖X‖[l](σ)dσ)dτ. $ | (17) |
We set
$ \label{D-6} \Phi(x): = \int^x \phi(y)\, dy. $ | (18) |
On the other hand, we use
$ |\Phi(\|X\|_{[l]}(t))|\le \underbrace{\int_s^t\phi(\|X\|_{[l]}(\tau))\|V\|_{[l]}(\tau)d\tau}_{ = :{\mathcal J}(s, t)}+|\Phi(\|X\|_{[l]}(s))|. $ | (19) |
Next, we claim:
$ |{\mathcal J}(t, s)|\le C_{\mathcal J} \;\;\; \text{for} \; 0\le s, t \le t_0. $ |
Proof of claim: First, we set
$ \mathcal{B}(s, t): = e^{-C_1\int_s^t\phi(\|X\|_{[l]}(\sigma))d\sigma}. $ |
Then, it is easy to see that
$ \label{20} \partial_t \mathcal{B} = -C_1\phi(\|X\|_{[l]}(t))\mathcal{B}(s, t) \;\;\;\text{and} \;\;\; \mathcal{B}(\tau, t)\mathcal{B}(s, \tau) = \mathcal{B}(s, t), \;\;\; \text{for} \; s\le \tau\le t . $ | (20) |
Note that
$‖V‖[l](t)≤‖V‖[l](s)B(s,t)+∫ts(C2√P(τ)+C3)B(τ,t)dτ≤C4B(s,t)+∫ts(C2√P(τ)+C3)B(τ,t)dτ.$ |
Thus, we have
$ J(s,t)≤∫tsϕ(‖X‖[l](τ))[C4B(s,τ)+∫τs(C2√P(σ)+C3)B(σ,τ)dσ]dτ=:I31+I32. $ | (21) |
Next, we estimate the terms
● (Estimation for
$ \label{121} \mathcal{I}_{31} = C_4 \int_s^t\phi(\|X\|_{[l]}(\tau))\mathcal{B}(\tau, s)\, d\tau = -\frac{C_4}{C_1}\int_s^t\partial_\tau(\mathcal{B}(\tau, s))\, d\tau\le\frac{C_4}{C_1}. $ | (22) |
● (Estimation for
$ I32=∫tsϕ(‖X‖[l](τ))[∫τs(C2√P(σ)+C3)B(τ,s)B(σ,s)dσ]dτ=∫tsϕ(‖x‖[l](τ))B(τ,s)[∫τs(C2√P(σ)+C3)1B(σ,s)dσ]dτ=−1C1B(t,s)∫ts(C2√P(σ)+C3)1B(σ,s)dσ+1C1∫ts[C2√P(τ)+C3]B(τ,s)B(τ,s)dτ≤1C1∫ts(C2√P(τ)+C3)dτ≤C5, $ | (23) |
where we used integration by parts and
$ {\mathcal J}(s, t)\le \frac{C_4}{C_1}+C_5. $ |
However, if
$ \Phi(s) = \left\{ s1−α1−αifα>1,logsifα=1. \right. $ |
Hence we have
$ \label{123} |\Phi(\|X\|_{[l]}(t))|\to \infty \;\;\; \text{as} \;t\to t_0^-. $ | (24) |
On the other hand in (19), we have
$ |\Phi(\|X\|_{[l]}(t))| < \infty, $ |
which is contradictory to (24).
In this section, we provide some initial configurations leading to collision between TCS particles in finite time. Note that if all initial temperatures are the same, then the TCS system (1)-(2) can be reduced to the particle C-S model with singular communication weight which has been extensively studied in [1,2,3,26,25]. In particular, the authors in [26] showed that the finite time collision between the Cucker-Smale particles with singular weights under certain assumptions on the initial configurations. For the same initial temperatures, as a direct application of [26,Proposition 3.1], we have a counterexample leading to the finite time collision.
More precisely, consider a two-body system on the real line, and its initial configuration
$ x_1^0 > x_2^0, \;\;\; v_2^0-v_1^0 = \frac{\kappa_1}{\theta_0(1-\alpha)}(x_1^0-x_2^0)^{1-\alpha}, \;\;\; \theta_1^0 = \theta_2^0 = :\theta^0 > 0, \;\;\; 0 < \alpha < 1. $ |
Then there exists a finite time
$ ˙x1=v1,˙x2=v2,t>0,˙v1=κ12ϕ(|x1−x2|)(v2θ2−v1θ1),˙v2=κ12ϕ(|x1−x2|)(v1θ1−v2θ2),˙θ1=κ22ζ(|x1−x2|)(1θ1−1θ2),˙θ2=κ22ζ(|x1−x2|)(1θ2−1θ1),ϕ(|x1−x2|)=1|x1−x2|α,ζ(|x1−x2|)=1|x1−x2|β,0<α<1,β>0. $ | (25) |
In the sequel, we will show that there exists an initial configuration leading to the finite time collision for the system (25).
Next, we define the difference of positions, velocities and temperatures of two particles as follows:
$ x(t): = x_1(t)-x_2(t), \;\;\; v(t): = v_1(t)-v_2(t), \;\;\; \theta(t): = \theta_1(t)-\theta_2(t).$ |
Then the TCS dynamics can be rewritten in terms of
$ \label{26} \frac{dx}{dt} = v, \;\;\; \frac{dv}{dt} = \kappa_1\phi(x)(\frac{v_2\theta}{\theta_1\theta_2}-\frac{v}{\theta_1}), \;\;\; \frac{d\theta}{dt} = -\kappa_2\zeta(x)\frac{\theta}{\theta_1\theta_2}, \;\;\; t > 0. $ | (26) |
To show the finite-time collisions in one-dimensional setting for the two-particle system, we consider the following initial configuration (see Figure 1):
$ x_2^0 < x_1^0, \;\;\; v_2^0 > 0 > v_1^0, \;\;\; \theta_2^0 > \theta_1^0. $ |
In the sequel, we will show that as long as there is no collisions between two particles, the ordering of velocity and temperatures will remain as it is, i.e., before the collision, we will have
$ v_2(t) > 0, \;\;\; \theta_2(t) > \theta_1(t), \;\;\; \text{i.e., } \;\;\; \frac{v_2 \theta}{\theta_1\theta_2} > 0. $ |
● Step A: First of all, we will prove
$ \frac{d\theta}{dt} = -\kappa_2\zeta(x)\frac{\theta}{\theta_1\theta_2}\le -\kappa_2\zeta(\delta)\frac{\theta}{(\theta_1^0)^2}, \;\;\; 0 < t < t_*. $ |
Now, we use Grönwall's lemma to get
$ \theta(t)\le \theta^0 \exp(-\frac{\kappa_2\zeta(\delta)t}{(\theta_1^0)^2}), \;\;\; 0\le t < t_*. $ |
This implies
● Step B: Next we will show that the velocity of each particle maintains its sign until the collision occurs. More precisely, we will show
$v(t_*) = 0, \;\;\; \text{and} \;\;\; v(t) < 0, \; |x(t)| > \delta, \;\;\; \text{for} \;\;\; 0\le t < t_*.$ |
Then it follows from (26) that
$ \frac{dv}{dt} = \kappa_1\phi(x)(\frac{v_2 \theta}{\theta_1\theta_2}-\frac{v}{\theta_1})\le -\frac{\kappa_1\phi(x)v}{\theta_1}\le -\frac{\kappa_1\phi(\delta)v}{\theta_1^0}, \;\;\; 0 < t < t_*. $ |
Here, we use the fact that
$ v(t)\le v^0\exp(-\frac{\kappa_1\phi(\delta)t}{\theta_1^0}), \;\;\; 0\le t < t_*, $ |
which is a contradictory to
● Step C: So far, we have obtained
$ \frac{v_2\theta}{\theta_1\theta_2} < 0, $ |
This together with (26)
$ \frac{dv}{dt}\le -\frac{\kappa_1\phi(x)v}{\theta_1}\le -\frac{\kappa_1\phi(x)v}{\theta_1^0}. $ |
Now, we set
$ \displaystyle \Phi(r): = \int^r \phi(s)\, ds = \frac{1}{1-\alpha}r^{1-\alpha}, $ |
i.e.,
$ \frac{dv}{dt}\le-\frac{\kappa_1}{\theta_1^0}\frac{d}{dt}(\Phi(x)). $ |
We integrate the above equation from 0 to
$ v(t)-v^0\le-\frac{\kappa_1}{\theta_1^0}(\Phi(x(t))-\Phi(x^0)). $ | (27) |
On the other hand, under our main assumptions in Theorem 3.1, we find
$ v^0 = -\frac{\kappa_1}{\theta_1^0(1-\alpha)}(x_1^0-x_2^0)^{1-\alpha} = -\frac{\kappa_1}{\theta_1^0}\Phi(x^0). $ |
Thus, (27) is again reduced to the following sub-linear differential inequality:
$ \frac{dx}{dt} = v(t)\le-\frac{\kappa_1}{\theta_1^0}\Phi(x(t))\le-\frac{\kappa_1}{\theta_1^0(1-\alpha)}(x(t))^{1-\alpha}, $ |
which is equivalent to
$ \frac{d}{dt}(x(t)^{\alpha})\le-\frac{\kappa_1\alpha}{\theta_1^0(1-\alpha)}. $ |
This yields
$ x(t)^\alpha\le x_0^\alpha-\frac{\kappa_1\alpha}{\theta_1^0(1-\alpha)}t. $ |
Hence, we have that the collision will occur at some time
In this subsection, we study a mono-clustering of the thermomechanical Cucker-Smale model (1)-(2). Note that the global existence of solutions is guaranteed by Theorem 3.1. The proof of Theorem 3.3 is exactly the same as in [12,Theorem 3.1]. More precisely, in [12], system (1)-(2) with regular weights are taken into account and the asymptotic emergent behavior is also obtained. However, the strategy used in [12] does not depend on the singularity of
● Step A (Derivation of differential inequalities): We first derive a system of differential inequalities for the extreme values for positions, velocities, and temperatures as follows:
$ |dD(X)dt|≤D(V),t>0,dD(V)dt≤−κ1ϕ(D(X))θ0MD(V)+2κ1(θ0m)2D(Θ)D(V),dD(Θ)dt≤−κ2ζ(D(X))(θ0M)2D(Θ). $ | (28) |
● Step B (Exponential flocking from the SDDI) : The next step is showing the exponential flocking from the SDDI (28). To do this, we first assume that the following conditions for initial configuration hold: Suppose that there exist
$bD(\Theta(0))\le \frac{a\phi(D(X(0)))}{4}, \;\;\; D(V(0))\le \frac{a}{2}\int_{D(X(0))}^{X^\infty} \phi(s)\, ds, $ |
and
$\frac{2}{3}\phi(D(X(0)))\le \phi(X^\infty) < \phi(DX(0)).$ |
Then, we use the bootstrapping argument to conclude the following flocking estimation:
$\sup\limits_{0\le t < \infty}D(X(t))\le X^\infty, \;\;\; D(V(t))\le D(V(0))e^{-a\phi(X^\infty)+\frac{bD(\Theta(0))}{c\zeta(X^\infty)}}, $ |
and
$ D(\Theta(t))\le D(\Theta(0))e^{-c\zeta(X^\infty)t}, \;\;\; t\ge0.$ |
In this section, we provide a local-in-time well-posedness of weak solutions (see Definition3.4) to the kinetic TCS equation:
$ ∂tf+∇x⋅(vf)+∇v⋅(F[f]f)+∂θ(G[f]f)=0,x,v∈Rd,θ∈R+,t>0,F[f](z,t):=−κ1∫R2d×R+ϕ(x−x∗)(vθ−v∗θ∗)f(z∗,t)dz∗,G[f](x,θ,t):=κ2∫R2d×R+ζ(x−x∗)(1θ−1θ∗)f(z∗,t)dz∗, $ | (29) |
where the interaction kernels are given as follows:
$ \phi(s): = \frac{1}{s^\alpha}, \;\;\; \zeta(s) = \frac{1}{s^\beta}, \;\;\; \text{with} \;\;\; \alpha, \beta > 0. $ |
The existence of weak solutions to (29) will be obtained via a suitable weak limit of the regularized system for (29). To do so, we introduce a radially symmetric standard mollifiers
$\eta(x) = {\tilde \eta}(|x|) \ge0, \;\;\; \text{supp} \eta \subset B_1(0), \;\;\; \int_{\mathbb{R}^d}\eta(x)\, dx = 1, \;\;\; \eta_\varepsilon(x): = \frac{1}{\varepsilon^d}\eta\Big(\frac{x}{\varepsilon}\Big).$ |
Now, we use this family of mollifier to mollify the communication kernels:
$ \phi_\varepsilon: = \phi*\eta_\varepsilon \;\;\; \text{and} \;\;\; \zeta_\varepsilon = \zeta*\eta_\varepsilon \;\;\; \text{for each} \;\;\;\varepsilon > 0. $ |
With the regularized weights
$ ∂tfε+v⋅∇xfε+∇v⋅[Fε(fε)fε]+∂T[Gε(fε)fε]=0,x,v∈Rd,θ∈R+,t>0,Fε(fε)(z,t):=∫R2d×R+ϕε(|x−x∗|)(v∗θ∗−vθ)fε(z∗,t)dz∗,Gε(fε)(x,θ,t):=∫R2d×R+ζε(|x−x∗|)(1θ−1θ∗)fε(z∗,t)dz∗,fε(z,0)=:f0(z). $ | (30) |
Note that the global-in-time existence of solution to (30) can be proved by using standard method of characteristics since all of the kernels are regular enough.
We next provide uniform estimates for
Lemma 5.1. For
$\frac{dX(t)}{dt}\le CX^2(t)(Y^\frac{d}{p'}(t)+1), \;\;\; \frac{dY(t)}{dt}\le CX(t)Y(t)(Y^\frac{d}{p'}(t)+1), \;\;\; t > 0, $ |
where
$ \sup\limits_{t\in[0, \tau]} (X(t)+Y(t))\le C. $ |
Proof. We set
$\tilde{X}: = X+1, \;\;\; \tilde{Y}: = Y+1.$ |
Then,
$d˜Xdt=dXdt≤CX2(Ydp′+1)≤C˜X2˜Ydp′,t>0,d˜Ydt=dYdt≤CXY(Ydp′+1)≤C˜X˜Y(1+dp′).$ |
This yields
$ \frac{d}{dt}(\tilde{X}+\tilde{Y})\le C\tilde{X}\tilde{Y}^\frac{d}{p'}(\tilde{X}+\tilde{Y}) \;\;\; \text{for} \;\;\; t \geq 0. $ | (31) |
On the other hands, we use Young's inequality to get
$ \tilde{X}\tilde{Y}^\frac{d}{p'}\le C( \tilde{X}^{\frac{d}{p'+d}}+\tilde{Y}^{\frac{d}{p'+d}})\le C(\tilde{X}+\tilde{Y})^{\frac{d}{p'+d}}. $ |
This together with (32) gives
$\frac{d}{dt}(\tilde{X}+\tilde{Y})\le C(\tilde{X}+\tilde{Y})^{1+\frac{d}{p'+d}}.$ |
Therefore, although the value
$ \sup\limits_{0\le t\le \tau} (X(t)+Y(t))\le\sup\limits_{0\le t\le \tau} (\tilde{X}(t)+\tilde{Y}(t))\le C, $ |
which yields the desired estimate.
In next lemma, we show that
$ R_v^\varepsilon : = \max \lbrace |v_0| : v_0 \in \overline{\{ v \in \mathbb{R}^d : \exists z \in \mathbb{R}^{2d} \times \mathbb{R}_+ \text{ such that } f_\varepsilon (z, t) \neq 0\}} \rbrace. $ |
Lemma 5.2. Let
$ ddt‖fε‖L1∩Lp≤C[(Rεv)dp′+1]‖fε‖2L1∩Lp,t>0,dRεvds≤CRεv[(Rεv)dp′+1]‖fε‖L1∩Lp. $ | (32) |
Proof. Below, we will derive the differential inequalities one by one.
● (Derivation of (32)
$ \label{33} \frac{d}{dt}\int_{\mathbb{R}^{2d}\times\mathbb{R}_+}f_\varepsilon^p\, dz = -(p-1)\int_{\mathbb{R}^{2d}\times\mathbb{R}_+}(\nabla_v\cdot(\mathcal{F}^\varepsilon[f_\varepsilon])+\partial_\theta\mathcal{G}^\varepsilon[f_\varepsilon])f_\varepsilon^p\, dz. $ | (33) |
To estimate the R.H.S. of (33), we use the standard cutoff function
$\chi_1(x): = {1|x|≤1,0|x|>2.$ |
We use a similar strategy in [2] to estimate
$\phi*\eta_\epsilon = (\phi\chi_1)*\eta_\varepsilon+(\phi(1-\chi_1))*\eta_\varepsilon$ |
and use Young's convolution inequality to get following inequalities:
$ ‖(ϕχ1)∗ηε‖Lp′≤‖ϕχ1‖Lp′‖ηε‖L1=‖ϕχ1‖Lp′<C,‖(ϕ(1−χ1))∗ηε‖L∞≤‖ϕ(1−χ1)‖L∞≤1. $ | (34) |
Now, thanks to the boundedness of velocity support and (34), we have
$|∇v⋅(Fε[fε])|=dθ∫R2d×R+ϕε(|x−x∗|)fε(z∗,t)dz∗≤dθm(∫R2d×R+|(ϕχ1)∗ηε||fε|dz∗+∫R2d×R+|(ϕ(1−χ1))∗ηε||fε|dz∗)≤C(‖(ϕχ1)∗ηε(x)1B(0,Rεv)(v)‖Lp′‖f‖Lp+‖(ϕ(1−χ1))∗ηε‖L∞‖fε‖1)≤C(Rv)dp′‖ϕχ1‖Lp′‖fε‖Lp+‖ϕ(1−χ1)‖L∞‖fε‖1=C((Rv)dp′+1)‖fε‖L1∩Lp,$ |
for some
$|∂θGε[fε]|≤1θ2m(∫R2d×R+|(ζχ1)∗ηε||fε|dz∗+∫R2d×R+|(ζ(1−χ1))∗ηε||fε|dz∗)≤C(Rv)dp′‖ζχ1‖Lp′‖fε‖Lp+‖ζ(1−χ1)‖L∞‖fε‖1=C((Rv)dp′+1)‖fε‖L1∩Lp.$ |
Thus, we have
$ \frac{d}{dt}\|f_\varepsilon\|_{L^1\cap L^p}\le C\Big((R_v)^{\frac{d}{p'}}+1\Big)\|f_\varepsilon\|^2_{L^1\cap L^p} \;\;\; \text{for} \;\;\; t > 0, $ |
where
● (Derivation of (32)
$12dds(Rεv)2≤(Rεv)2(s)θm∫R2d×R+ϕ(|x(s)−x∗|)f(y,v∗,θ∗,s)dz∗≤C(Rεv)2((Rεv)dp′+1)‖fε‖L1∩Lp.$ |
Note that in the last inequality, we used similar estimate as in the previous step. Thus we have
$ \frac{dR_v^\varepsilon }{ds}\le CR_v^\varepsilon ((R_v^\varepsilon )^{\frac{d}{p'}}+1)\|f_\varepsilon\|_{L^1\cap L^p}, $ |
where
By direct applications of Lemma 5.1 and Lemma 5.2, we have the following uniform bound estimates and stability estimate.
Proposition 2. The following assertions hold.
1. (Uniform boundedness): Let
$ \sup\limits_{t\in [0, \tau]}\|f_\varepsilon\|_{L^1\cap L^p}\le C, \;\;\; R_v(t): = \sup\limits_{t\in[0, \tau]}|\Omega_v(t)|\le C, $ |
where
2. Let
$\frac{d}{dt}W_1(f_\varepsilon(t), f_{\varepsilon'}(t))\le C(W_1(f_\varepsilon(t), f_{\varepsilon'}(t))+\varepsilon+\varepsilon'), \;\;\; \forall 0\le t < \tau.$ |
Proof. (1) The uniform boundedness follow from Lemma 5.1 and Lemma 5.2.
(2) The stability estimate can be done as for the regular case. First, we define the family of characteristic curves
$ ddtxε(t;s,x,v,θ)=vε(t;s,x,v,θ),0≤s≤tddtvε(t;s,x,v,θ)=Fε[fε](Zε(t;s,x,v,θ),t),ddtθε(t;s,x,v,θ)=Gε[fε](xε(t;s,x,v,θ),θε(t;s,x,v,θ),t), $ | (35) |
and define
$\mathcal{T}^0(x, v, \theta) = (\mathcal{T}^0_1(x, v, \theta), \mathcal{T}^0_2(x, v, \theta), \mathcal{T}^0_3(x, v, \theta))$ |
between
$\mathcal{T}^t\#f_\varepsilon(t) = f_{\varepsilon'}(t), \;\;\; \text{where} \;\;\; \mathcal{T'}: = Z_{\varepsilon'}(t;t_0, \cdot, \cdot, \cdot)\circ \mathcal{T}^0\circ Z_\varepsilon(t_0;t, \cdot, \cdot, \cdot)$ |
Now, it follows from the Proposition 1 that we have
$ W_1(f_\varepsilon(t), f_{\varepsilon'}(t))\le \int_{\mathbb{R}^{2d}\times\mathbb{R}_+}|Z_\varepsilon(t;t_0, z)-Z_{\varepsilon'}(t;t_0, \mathcal{T}^0(z))|f_\varepsilon(z, t_0)\, dz = :Q_{\varepsilon, \varepsilon'}(t). $ |
Then, it follows from (35) that we have
$ddtQε,ε′(t)|t=t0+≤∫R2d×R+|vε(t;t0,z)−vε′(t;t0,T0(z))|fε(z,t0)dz|t=t0++∫R2d×R+|Fε[fε](Zε(t;t0,z),t)−Fε′[fε′](Zε′(t;t0,T0(z)),t)|fε(z,t0)dz|t=t0++∫R2d×R+|Gε[fε](xε(t;t0,z),θε(t;t0,z),t)−Gε′[fε′](xε′(t;t0,T0(z)),θε′(t;t0,T0(z)),t)|fε(z,t0)dz|t=t0+=:I41+I42+I43.$ |
Below, we estimate the terms
● (Estimate of
$\mathcal{I}_{41} = \int_{\mathbb{R}^{2d}\times\mathbb{R}_+}|v-\mathcal{T}_2^0(z)|f_\varepsilon(z, t_0)\, dz\le CW_1(f_\varepsilon(t_0), f_{\varepsilon'}(t_0)).$ |
● (Estimate of
$I42=∫R2d×R+|∫R2d×R+ϕε(x−y)(v∗θ∗−vθ)fε(z∗,t0)dz∗−∫R2d×R+ϕε′(T01(z)−y)(v∗θ∗−T02(z)T03(z))fε′(z∗,t0)dz∗|fε(z,t0)dz=∫R2d×R+|∫R2d×R+ϕε(x−y)(v∗θ∗−vθ)fε(z∗,t0)dz∗−∫R2d×R+ϕε′(T01(z)−T01(z∗))(T02(z∗)T03(z∗)−T02(z)T03(z))fε′(z∗,t0)dz∗|fε(z,t0)dz.$ |
Now, we define the further subterms of
$ \mathcal{I}_{42} = :\int_{\mathbb{R}^{2d}\times\mathbb{R}_+}|\mathcal{I}_{421}+\mathcal{I}_{422}|f_\varepsilon(z, t_0)\, dz \le\int_{\mathbb{R}^{2d}\times\mathbb{R}_+}|\mathcal{I}_{421}|+|\mathcal{I}_{422}|f_\varepsilon(z, t_0)\, dz, $ |
where
$I421:=∫R2d×R+(ϕε(x−y)−ϕε′(T01(z)−T01(z∗)))(v∗θ∗−vθ)fε(z∗,t0)dz∗,I422:=∫R2d×R+ϕε′(T01(z)−T01(z∗))(v∗θ∗−vθ)fε(z∗,t0)dz∗−∫R2d×R+ϕε′(T01(z)−T01(z∗))(T02(z∗)T03(z∗)−T02(z)T03(z))fε(z∗,t0)dz∗.$ |
$I421≤∫R2d×R+|(ϕε−ϕε′)(x−y)||v∗θ∗−vθ|fε(z∗,t0)dz∗+∫R2d×R+|ϕε′(x−y)−ϕε′(T01(z)−T01(z∗))||v∗θ∗−vθ|fε(z∗,t0)dz∗.$ |
Now we estimate
$ |ϕε(x)−ϕ(x)|≤∫Rd|ϕ(x−y)−ϕ(x)|θε(y)dy≤2∫Rd(1|x|1+α+1|x−y|1+α)|y|θε(y)dy≤2ε∫{y:|y|≤ε}(1|x|1+α+1|x−y|1+α)θε(y)dy≤Cε|x|1+α. $ | (36) |
Recall that the velocity support and temperature support have finite diameters at any finite time. Then, we use this fact together with the estimate (36) to obtain
$∫(R2d×R+)2|(ϕε−ϕ)(x−y)||v∗θ∗−vθ|fε(z∗,t0)fε(z,t0)dzdz∗≤Cε∫(Rd×Ωv(τ)×ΩT(τ))21|x−y|1+αfε(z∗,t0)fε(z,t0)dzdz∗≤Cε∫R2d×R+(∫{y:|x−y|<1}×Ωv(τ)×ΩT(τ)+∫{y:|x−y|≥1}×Ωv(τ)×ΩT(τ)1|x−y|1+αfε(z∗,t0)dz∗)fε(z,t0)dz≤Cε∫R2d×R+[(∫{y:|x−y|<1}1|x−y|(1+α)p′dy)1p′‖fε‖Lp+‖fε‖L1]fε(z,t0)dz≤Cε‖fε‖2≤Cε.$ |
This yields
$ \int_{(\mathbb{R}^{2d}\times\mathbb{R}_+)^2}|(\phi_\varepsilon-\phi_{\varepsilon'})(x-y)|\vert \frac{v_*}{\theta_*}-\frac{v}{\theta} \vert f_\varepsilon(z_*)\, dz_*\le C(\varepsilon+\varepsilon'). $ |
For the second term of
$∫(Rd×Ωv(τ)×ΩT(τ))2|ϕε′(x−y)−ϕε′(T01(z)−T01(z∗))||v∗θ∗−vθ|fε(z,t0)fε(z∗,t0)dzdz∗≤C∫(Rd×Ωv(τ)×ΩT(τ))2|T01(z)−x||T01(z)−T01(z∗)|1+αfε(z,t0)fε(z∗,t0)dzdz∗+C∫(Rd×Ωv(τ)×ΩT(τ))2|T01(z)−x||x−y|1+αfε(z,t0)fε(z∗,t0)dzdz∗≤Cmax(‖fε‖,‖fε′‖)W1(fε(t0),fε′(t0)).$ |
Thus, we combine these estimation to get
$ \int_{\mathbb{R}^{2d}\times\mathbb{R}_+}|\mathcal{I}_{421}|f_\varepsilon(z, t_0)\, dz \le C(W_1(f_\varepsilon(t_0), f_{\varepsilon'}(t_0))+\varepsilon+\varepsilon'). $ | (37) |
$∫R2d×R+|I422|fε(z,t0)dz≤∫(R2d×R+)2|ϕε′(T01(z)−T01(z∗))||v∗θ∗−T02(z∗)T03(z∗)|fε(z,t0)fε(z∗,t0)dzdz∗+∫(R2d×R+)2|ϕε′(T01(z)−T01(z∗))||vθ−T02(z)T03(z)|fε(z,t0)fε(z∗,t0)dzdz∗=:I4221+I4222$ |
However, it is easy to see that
$I4221=∫R2d×R+(∫Rd×Ωv(τ)×ΩT(τ)|ϕε′(T01(z)−T01(z∗))|fε(z,t0)dz)×|v∗θ∗−T02(z∗)T03(z∗)|fε(z∗,t0)dz∗≤C‖fε′‖∫R2d×R+|v∗θ∗−T02(z∗)T03(z∗)|fε(z∗,t0)dz∗≤CW1(fε(t0),fε′(t0)).$ |
Similarly, we also can estimate
$ \mathcal{I}_{4222}\le CW_1(f_\varepsilon(t_0), f_{\varepsilon'}(t_0)). $ |
Hence, we have
$ \int_{\mathbb{R}^{2d}\times\mathbb{R}_+}|\mathcal{I}_{422}|f_\varepsilon(z, t_0) \, dz \le \mathcal{I}_{4221}+\mathcal{I}_{4222} \le CW_1(f_\varepsilon(t_0), f_{\varepsilon'}(t_0)). $ | (38) |
We combine (37)-(38) to get
$ \mathcal{I}_{42}\le C(W_1(f_\varepsilon(t_0), f_{\varepsilon'}(t_0))+\varepsilon+\varepsilon'). $ |
● (Estimate of
$\mathcal{I}_{43}\le C(W_1(f_\varepsilon(t_0), f_{\varepsilon'}(t_0))+\varepsilon+\varepsilon').$ |
Now, we combine the estimation of
$ \frac{d}{dt}Q_{\varepsilon, \varepsilon'}(t)\Big|_{t = t_0+}\le C(W_1(f_\varepsilon(t_0), f_{\varepsilon'}(t_0))+\varepsilon+\varepsilon'). $ |
This implies that for arbitrary
$\frac{d}{dt}W_1(f_\varepsilon(t), f_{\varepsilon'}(t))\le C(W_1(f_\varepsilon(t), f_{\varepsilon'}(t))+\varepsilon+\varepsilon'), $ |
where
Now, we are ready to prove the local-in-time existence and uniqueness of weak solution in the sense of Definition 3.4. Note that Proposition 2 implies that the family of regularized solution
● (Existence part): Fix any test function
$ ∫R2d×R+fε(z,τ)Φ(z,τ)dz−∫R2d×R+fε(z,0)Φ(z,0)dz=∫τ0∫R2d×R+fε(∂tΦ+v⋅∇xΦ+Fε[fε]⋅∇vΦ+Gε[fε]∂θΦ)dzdt. $ | (39) |
Note that we can pass the limit
$ ∫R2d×R+fε(z,τ)Φ(z,τ)dz−∫R2d×R+fε(z,0)Φ(z,0)dz⟶∫R2d×R+f(z,τ)Φ(z,τ)dz−∫R2d×R+f(z,0)Φ(z,0)dz,∫τ0∫R2d×R+fε(∂tΦ+v⋅∇xΦ)dzdt⟶∫τ0∫R2d×R+f(∂tΦ+v⋅∇xΦ)dzdt. $ | (40) |
Thus, we only need to show that as
$ ∫τ0∫R2d×R+fεFε[fε]⋅∇vΦdzdt→∫τ0∫R2d×R+fF[f]⋅∇vΦdzdt,∫τ0∫R2d×R+fεGε[fε]∂θΦdzdt→∫τ0∫R2d×R+fG[f]∂θΦdzdt. $ | (41) |
Since proofs of two limiting processes are almost same, we only focus on first limit in (41). Note that
$|∫τ0∫R2d×R+(fεFε[fε]−fF[f])⋅∇vΦdzdt|≤|∫τ0∫R2d×R+fε(Fε[fε]−F[fε])⋅∇vΦdzdt|+|∫τ0∫R2d×R+fε(F[fε]−F[f])⋅∇vΦdzdt|+|∫τ0∫R2d×R+(fε−f)F[f]⋅∇vΦdzdt|=:I51+I52+I53.$ |
Next, we estimate the terms
$ I51=|∫τ0∫(R2d×R+)2(ϕε−ϕ)(|x−x∗|)(∇vΦ)⋅(v∗θ∗−vθ)fε(z,t)fε(z∗,t)dzdz∗dt|≤Cε∫τ0∫(Rd×Ωv(τ)×ΩT(τ))21|x−x∗|1+αfε(z,t)fε(z∗,t)dzdz∗dt≤Cε‖fε‖2≤Cε→0asε→0. $ | (42) |
$ \mathcal{H}[f_\varepsilon ](z, t): = \int_{\mathbb{R}^{2d} \times \mathbb{R}_+} f_\varepsilon (z_*, t) \phi(|x_* - x|)(\frac{v}{\theta} - \frac{v_*}{\theta_*} )\cdot \nabla_{v_*}\Phi(z_*)\, dz_*. $ |
Then, we have
$ \mathcal{I}_{52} = \vert \int_0^\tau \int_{\mathbb{R}^{2d} \times \mathbb{R}_+} \mathcal{H}[f_\varepsilon ](z_*, t)(f_\varepsilon (z_*, t) - f(z_*, t))\, dz_*\, dt \vert. $ |
On the other hand, we employ similar arguments as in the proof of Lemma 5.2 and the estimate of
$ \label{43} \mathcal{I}_{52} \leq C\sup\limits_{0 \leq t \leq \tau}W_1(f^\varepsilon (t), f(t)) \to 0 \;\;\; \text{as} \;\;\; \varepsilon \to 0. $ | (43) |
$ \label{44} \mathcal{I}_{53} \leq C\sup\limits_{0 \leq t \leq \tau}W_1(f^\varepsilon (t), f(t)) \to 0 \;\;\; \text{as} \;\;\; \varepsilon \to 0. $ | (44) |
In (41), we collect all estimates (42), (43) and (44) to obtain the limiting convergence
● (Uniqueness and stability) : Let
$\frac{d}{dt}W_1(f_1(t), f_2(t))\le CW_1(f_1(t), f_2(t)), \;\;\;\; t\in [0, \tau).$ |
Then, the Grönwall lemma yields the uniqueness of the solution.
In this paper, we have studied the dynamic features of the TCS model with singular power-law kernels in their velocity and temperature dynamics. For strong singularities in communication weights, collisions cannot occur in any finite time. Thus, the classical Cauchy-Lipschitz theory can be applied to yield the global existence of smooth solutions. In contrast, when the singularity is mildly weak, finite-time collisions can still occur from some prepared initial configurations. Hence, the global smooth solutions cannot be guaranteed in general. As far as the authors know, after a finite-time collision occur, there is no existence theory after collision time. Formal BBGKY hierarchy argument yields the kinetic TCS equation with singular kernel. For this kinetic equation with singular kernel, we also provide a local existence of weak solutions. At present, we do not have a global existence theory for weak or strong solutions for the kinetic TCS equation. This will be an interesting future work to be explored.
The work of Y.-P. Choi is supported by NRF grants(2017R1C1B2012918 and 2017R1A4A1014735), and the work of S.-Y. Ha is supported by the National Research Foundation of Korea (2017R1A2B2001864).
[1] |
On collision-avoiding initial configurations to Cucker-Smale type flocking models. Comm. Math. Sci. (2012) 10: 625-643. ![]() |
[2] |
J. A. Carrillo, Y. -P. Choi and M. Hauray, Local well-posedness of the generalized CuckerSmale model with singular kernels, Mathematical Modeling of Complex Systems, 17-35,
ESAIM Proc. Surveys, 47, EDP Sci., Les Ulis, 2014. doi: 10.1051/proc/201447002
![]() |
[3] |
Sharp conditions to avoid collisions in singular Cucker-Smale interactions. Nonlinear Anal.-Real. (2017) 37: 317-328. ![]() |
[4] | J. A. Carrillo, Y. -P. Choi and S. Pérez, A review on attractive-repulsive hydrodynamics for consensus in collective behavior, in Active Particles Vol. Ⅰ - Advances in Theory, Models, Applications(tentative title), Series: Modeling and Simulation in Science and Technology, (eds. N. Bellomo, P. Degond, and E. Tadmor), Birkhäuser Basel, (2017), 259-298. |
[5] |
Asymptotic flocking dynamics for the kinetic Cucker-Smale model. SIAM J. Math. Anal. (2010) 42: 218-236. ![]() |
[6] |
Global classical solutions of the Vlasov-Fokker-Planck equation with local alignment forces. Nonlinearity (2016) 29: 1887-1916. ![]() |
[7] | Y. -P. Choi, S. -Y. Ha and Z. Li, Emergent dynamics of the Cucker-Smale flocking model and its variants, in Active Particles Vol. Ⅰ - Advances in Theory, Models, Applications(tentative title), Series: Modeling and Simulation in Science and Technology, (eds. N. Bellomo, P. Degond, and E. Tadmor), Birkhäuser Basel, (2017), 299-331. |
[8] |
Avoiding collisions in flocks. IEEE Trans. Automatic Control (2010) 55: 1238-1243. ![]() |
[9] |
Emergent behavior in flocks. IEEE Trans. Automat. Control (2007) 52: 852-862. ![]() |
[10] |
A kinetic flocking model with diffusion. Comm. Math. Phys. (2010) 300: 95-145. ![]() |
[11] |
Fluid dynamic description of flocking via Povzner-Boltzmann equation. Physica D (2011) 240: 21-31. ![]() |
[12] | S. -Y. Ha, J. Kim, C. Min, T. Ruggeri and X. Zhang, Uniform stability and mean-field limit of thermodynamic Cucker-Smale model, Submitted. |
[13] |
Emergent behaviors of Thermodynamic Cucker-Smale particles. SIAM J. Math. Anal. (2018) 50: 3092-3121. ![]() |
[14] |
Uniform stability of the Cucker-Smale model and its application to the mean-field limit. Kinet. Relat. Models (2018) 11: 1157-1181. ![]() |
[15] |
Emergent dynamics for the hydrodynamic Cucker-Smale system in a moving domain. SIAM. J. Math. Anal. (2015) 47: 3813-3831. ![]() |
[16] |
Emergence of time-asymptotic flocking in a stochastic Cucker-Smale system. Commun. Math. Sci. (2009) 7: 453-469. ![]() |
[17] |
A simple proof of Cucker-Smale flocking dynamics and mean field limit. Commun. Math. Sci. (2009) 7: 297-325. ![]() |
[18] |
From particle to kinetic and hydrodynamic description of flocking. Kinetic Relat. Models (2008) 1: 415-435. ![]() |
[19] |
Emergent dynamics of a thermodynamically consistent particle model. Arch. Ration. Mech. An. (2017) 223: 1397-1425. ![]() |
[20] |
On the Cucker-Smale flocking with alternating leaders. Quart. Appl. Math. (2015) 73: 693-709. ![]() |
[21] |
Cucker-Smale flocking under rooted leadership with fixed and switching topologies. SIAM J. Appl. Math. (2010) 70: 3156-3174. ![]() |
[22] |
Heterophilious dynamics: Enhanced Consensus. SIAM Review (2014) 56: 577-621. ![]() |
[23] |
A new model for self-organized dynamics and its flocking behavior. J. Statist. Phys. (2011) 144: 923-947. ![]() |
[24] |
The Cucker-Smale equation: Singular communication weight, measure-valued solutions and weak-atomic uniqueness. Arch. Ration. Mech. Anal. (2018) 227: 273-308. ![]() |
[25] |
Discrete Cucker-Smale flocking model with a weakly singular weight. SIAM J. Math. Anal. (2015) 47: 3671-3686. ![]() |
[26] |
Existence of piecewise weak solutions of a discrete Cucker-Smale's flocking model with a singular communication weight. J. Differ. Equat. (2014) 257: 2900-2925. ![]() |
[27] |
Cucker-Smale flocking under hierarchical leadership. SIAM J. Appl. Math. (2007) 68: 694-719. ![]() |
[28] |
Flocks, herds, and Schools: A quantitative theory of flocking. Physical Review E. (1998) 58: 4828-4858. ![]() |
[29] |
Swarming patterns in a two-dimensional kinematic model for biological groups. SIAM J. Appl. Math. (2004) 65: 152-174. ![]() |
[30] |
Novel type of phase transition in a system of self-driven particles. Phys. Rev. Lett. (1995) 75: 1226-1229. ![]() |
[31] |
C. Villani,
Topics in Optimal Transportation, American Mathematical Society, 2003. doi: 10.1007/b12016
![]() |
[32] |
C. Villani,
Optimal Transport, Old and New, Springer-Verlag, 2009. doi: 10.1007/978-3-540-71050-9
![]() |
1. | Piotr Minakowski, Piotr B. Mucha, Jan Peszek, Ewelina Zatorska, 2019, Chapter 7, 978-3-030-20296-5, 201, 10.1007/978-3-030-20297-2_7 | |
2. | Hangjun Cho, Jiu‐Gang Dong, Seung‐Yeal Ha, Emergent behaviors of a thermodynamic Cucker‐Smale flock with a time‐delay on a general digraph, 2022, 45, 0170-4214, 164, 10.1002/mma.7771 | |
3. | Junhyeok Byeon, Seung-Yeal Ha, Jeongho Kim, Asymptotic flocking dynamics of a relativistic Cucker–Smale flock under singular communications, 2022, 63, 0022-2488, 012702, 10.1063/5.0062745 | |
4. | Jiu-Gang Dong, Seung-Yeal Ha, Doheon Kim, On the Cucker--Smale Ensemble with the q-Closest Neighbors in a Self-Consistent Temperature Field, 2020, 58, 0363-0129, 368, 10.1137/18M1195462 | |
5. | Seung-Yeal Ha, Jinwook Jung, Jeongho Kim, Jinyeong Park, Xiongtao Zhang, A mean-field limit of the particle swarmalator model, 2021, 14, 1937-5077, 429, 10.3934/krm.2021011 | |
6. | Jiu-Gang Dong, Seung-Yeal Ha, Doheon Kim, From discrete Cucker-Smale model to continuous Cucker-Smale model in a temperature field, 2019, 60, 0022-2488, 072705, 10.1063/1.5084770 | |
7. | Piotr Minakowski, Piotr B. Mucha, Jan Peszek, Density-induced consensus protocol, 2020, 30, 0218-2025, 2389, 10.1142/S0218202520500451 | |
8. | Jiu-Gang Dong, Seung-Yeal Ha, Doheon Kim, Emergence of mono-cluster flocking in the thermomechanical Cucker–Smale model under switching topologies, 2021, 19, 0219-5305, 305, 10.1142/S0219530520500025 | |
9. | Jiu-Gang Dong, Seung-Yeal Ha, Doheon Kim, Jeongho Kim, Time-delay effect on the flocking in an ensemble of thermomechanical Cucker–Smale particles, 2019, 266, 00220396, 2373, 10.1016/j.jde.2018.08.034 | |
10. | Jiu-Gang Dong, Seung-Yeal Ha, Doheon Kim, Emergent behaviors of continuous and discrete thermomechanical Cucker–Smale models on general digraphs, 2019, 29, 0218-2025, 589, 10.1142/S0218202519400013 | |
11. | Hyunjin Ahn, Uniform stability of the Cucker–Smale and thermodynamic Cucker–Smale ensembles with singular kernels, 2022, 17, 1556-1801, 753, 10.3934/nhm.2022025 | |
12. | Hyunjin Ahn, Seung-Yeal Ha, Woojoo Shim, Emergent behaviors of the discrete thermodynamic Cucker–Smale model on complete Riemannian manifolds, 2021, 62, 0022-2488, 122701, 10.1063/5.0058616 | |
13. | Hyunjin Ahn, Seung-Yeal Ha, Woojoo Shim, Emergent dynamics of a thermodynamic Cucker-Smale ensemble on complete Riemannian manifolds, 2021, 14, 1937-5077, 323, 10.3934/krm.2021007 | |
14. | Hyunjin Ahn, Junhyeok Byeon, Seung-Yeal Ha, Interplay of unit-speed constraint and singular communication in the thermodynamic Cucker–Smale model, 2023, 33, 1054-1500, 10.1063/5.0165245 | |
15. | Hyunjin Ahn, Se Eun Noh, Finite-in-time flocking of the thermodynamic Cucker–Smale model, 2024, 19, 1556-1801, 526, 10.3934/nhm.2024023 |