Citation: Yenisetti SC, Muralidhara. Beneficial Role of Coffee and Caffeine in Neurodegenerative Diseases: A Minireview[J]. AIMS Public Health, 2016, 3(2): 407-422. doi: 10.3934/publichealth.2016.2.407
[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, 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 |
[3] | 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 |
[4] | 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 |
[5] | 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 |
[6] | 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 |
[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] | Linglong Du, Anqi Du, Zhengyan Luo . Global well-posedness of strong solution to the kinetic Cucker-Smale model with external potential field. Networks and Heterogeneous Media, 2025, 20(2): 460-481. doi: 10.3934/nhm.2025021 |
[10] | Matteo Novaga, Emanuele Paolini, Eugene Stepanov, Vincenzo Maria Tortorelli . Isoperimetric planar clusters with infinitely many regions. Networks and Heterogeneous Media, 2023, 18(3): 1226-1235. doi: 10.3934/nhm.2023053 |
Emergent dynamics in interacting multi-agent systems are frequently observed in nature. Examples include the aggregation of bacteria [39], flocking of birds and vehicular flocking [7,14,19,34], schooling of fish [20,38] and the synchronization of fireflies and pacemaker cells [1,8,21,37,43]. To more introduce related literature, we refer to [22,35,41,42]. Herein, we are primarily concerned with "flocking" in which agents exhibit ordered movements and form appropriate groups. After the work of Vicsek et al. in [40], many studies on models representing flocking have been actively conducted for decades. Among them, the Cucker–Smale model [19] has received significant attention in math and physics communities due to its dissipative and simple velocity structure. Essentially, the Cucker–Smale model is a flocking dynamic system for position and velocity based on the Newtonian sense, which is governed by
$ {dxidt=vi,t>0,i∈{1,⋯,N},dvidt=κNN∑j=1ψ(‖xi−xj‖)(vj−vi),(xi(0),vi(0))=(x0i,v0i)∈Rd×Rd, $ | (1.1) |
where $ N $ denotes the number of particles, $ \kappa $ is a nonnegative coupling strength and $ \psi $ is a communication weight. To date, there have been many works examining this system and its variants due to its dissipative structure for velocity, such as the mean-field limit [5,6,25,28,30], kinetic models [9,32], hydrodynamic descriptions [23,24,33], particle analysis [9,10,13,14,15,16,17,18], temperature field [26,31] and relativistic setting [4,5,6,27].
Since Eq (1.1), the authors of [12] noted that several Vicsek-type models with unit-speed constraints have been actively studied concerning heading angles in math community. To give a unit-speed constraint to Eq (1.1), the authors modified the velocity coupling term Eq $ (1.1)_2 $ so that the velocity of each agent has a unit-speed constraint as follows:
$ \psi(\|x_i-x_j\|)\left(v_j-v_i\right) \longrightarrow \psi(\|x_i-x_j\|)\left(v_j-\frac{\langle v_j, v_i\rangle v_i}{\|v_i\|^2}\right), $ |
where the modified term is perpendicular to $ v_i $. Thus, they proposed the following Cucker–Smale type model with constant speed and studied its flocking dynamics:
$ {dxidt=vi,t>0,i∈{1,⋯,N},dvidt=κNN∑j=1ψ(‖xi−xj‖)(vj−⟨vj,vi⟩vi‖vi‖2),(xi(0),vi(0))=(x0i,v0i)∈Rd×Rd. $ | (1.2) |
Equation (1.2) has also been studied from several perspectives; for example, particle analysis [12], the emergence of the bi-cluster flocking in [17], multi-cluster flocking and critical coupling strength in [29], time-delay effect [11] and general digraph setting [36].
However, because the above literature [11,12,17,29,36] were only motivated by the original Cucker–Smale model (1.1) without considering internal energy, the author of [2] noted the extension of the above model to a temperature field to describe more realistic flocking dynamics. For this, as a backbone model, the author first adopted a thermodynamic Cucker–Smale model proposed by [26,31] based on the theory of multi-temperature mixture of fluids under the space of homogeneity, which is given by the following second-order ODEs for position-velocity-temperature $ (x_i, v_i, T_i) $:
$ dxidt=vi,t>0,i∈[N]:={1,⋯,N}, $ | (1.3a) |
$ dvidt=κ1NN∑j=1ϕ(‖xi−xj‖)(vjTj−viTi), $ | (1.3b) |
$ ddt(Ti+12‖vi‖2)=κ2NN∑j=1ζ(‖xi−xj‖)(1Ti−1Tj), $ | (1.3c) |
$ (xi(0),vi(0),Ti(0))=(x0i,v0i,T0i)∈Rd×Rd×R+−{0}, $ | (1.3d) |
where $ \sum_{i = 1}^{N} T_i^0 = :NT^\infty, $ $ N $ denotes the number of particles, $ \kappa_1, \kappa_2 $ are nonnegative coupling strengths and $ \psi, \zeta $ are communication weights. Then, motivated from the derivation idea of Eq (1.2), by modifying the velocity coupling term Eq (1.3a) as
$ \phi(\|x_i-x_j\|)\left(\frac{v_j}{T_j}-\frac{v_i}{T_i}\right) \longrightarrow \phi(\|x_i-x_j\|)\left(\frac{v_j}{T_j}-\frac{\langle v_j, v_i\rangle v_i}{T_j\|v_i\|^2}\right), $ |
the author suggested the following TCSUS model in terms of position-velocity-temperature $ (x_i, v_i, T_i) $:
$ {dxidt=vi,t>0,i∈{1,⋯,N},dvidt=κ1NN∑j=1ϕ(‖xi−xj‖)(vjTj−⟨vj,vi⟩viTj‖vi‖2),ddt(Ti+12‖vi‖2)=κ2NN∑j=1ζ(‖xi−xj‖)(1Ti−1Tj),(xi(0),vi(0),Ti(0))=(x0i,v0i,T0i)∈Rd×Sd−1×(R+−{0}), $ | (1.4) |
where $ \sum_{i = 1}^{N} T_i^0 = :NT^\infty. $ Afterward, the author immediately verified that each agent in the system (1.4) has a unit-speed. Then, from the relations,
$ \|v_i\| = 1, \quad \frac{\langle v_j, v_i\rangle v_i}{T_j\|v_i\|^2} = \frac{\langle v_j, v_i\rangle v_i}{T_j}\quad \text{and} \quad \frac{d}{dt}\left(T_i+\frac{1}{2}\|v_i\|^2\right) = \frac{dT_i}{dt}, $ |
the author simply represented the system (1.4) as follows:
$ dxidt=vi,t>0,i∈{1,⋯,N}, $ | (1.5a) |
$ dvidt=κ1NN∑j=1ϕ(‖xi−xj‖)(vj−⟨vj,vi⟩viTj), $ | (1.5b) |
$ dTidt=κ2NN∑j=1ζ(‖xi−xj‖)(1Ti−1Tj), $ | (1.5c) |
$ (xi(0),vi(0),Ti(0))=(x0i,v0i,T0i)∈Rd×Sd−1×(R+−{0}), $ | (1.5d) |
where $ \sum_{i = 1}^{N} T_i^0 = :NT^\infty. $ Here, we set $ \mathbb{R}_+: = [0, \infty) $ throughout the paper and we assume that two communication weights $ \phi $, $ \zeta:\mathbb{R}_+\to \mathbb{R}_+ $ are nonnegative, locally Lipschitz continuous and monotonically decreasing and that $ \mathbb{S}^{d-1} $ is the unit $ (d-1) $-sphere isometrically embedded in $ \mathbb{R}^d $; hence,
$ 0≤ϕ(r)≤ϕ(0)=1,(ϕ(r1)−ϕ(r2))(r1−r2)≤0,∀r,r1,r2≥0,ϕ(⋅)∈C0,1loc(R+;R+),0≤ζ(r)≤ζ(0)=1,(ζ(r1)−ζ(r2))(r1−r2)≤0,∀r,r1,r2≥0,ζ(⋅)∈C0,1loc(R+;R+),Sd−1:={x:=(x1,⋯,xd)|d∑i=1|xi|2=1,}where xi is the i −th component of x∈Rd. $ |
The system (1.5) was studied in terms of mono-cluster flocking and bi-cluster flocking in [2] and collision avoidance [3], but the multi-cluster flocking of system (1.5) has not been studied yet. Indeed, the multi-cluster flocking phenomenon is ubiquitous in daily life. Examples include opinion disagreement, schools of fish invaded by predators and flight multi-formation. In addition, a phenomenon in which individuals with the same characteristics gather together can be an example of the multi-cluster flocking.
Therefore, this paper is mainly interested in the non-emergence of mono-cluster flocking in the system (1.5) under a sufficiently small coupling strength and extending the bi-cluster flocking of [2] to general multi-cluster flocking. For this, we first introduce several basic notions concerning mono- and multi-cluster flocking as follows:
Definition 1.1. Let $ Z = \{(x_i, v_i, T_i)\}_{i = 1}^N $ be a solution to the system (1.5).
(1) The configuration $ Z $ exhibits mono-cluster flocking if the following statements hold:
$ (i)(Group formation)⟺supt∈R+max1≤i,j≤N‖xi(t)−xj(t)‖<∞,(ii)(Velocity alignment)⟺limt→∞max1≤i,j≤N‖vj(t)−vi(t)‖=0,(iii)(Temperature equilibrium)⟺limt→∞max1≤i,j≤N|Tj(t)−Ti(t)|=0. $ |
(2) The configuration $ Z $ exhibits multi-cluster flocking if there exist $ n $ cluster groups $ Z_\alpha = \{(x_{\alpha i}, v_{\alpha i}, T_{\alpha i})\}_{i = 1}^{N_\alpha} $ such that the following assertions hold for $ 1\leq n\leq N $:
$ (i)|Zα|=Nα≥1,n∑α=1|Zα|=n∑α=1Nα=N,(ii)supt∈R+max1≤k,l≤Nα‖xαk(t)−xαl(t)‖<∞,limt→∞max1≤k,l≤Nα‖vαk(t)−vαl(t)‖=0,limt→∞max1≤k,l≤Nα|Tαk(t)−Tαl(t)|=0,n≥3,1≤α≤n,(iii)inft∈R+mink,l‖xαk−xβl‖=∞,1≤k≤Nα,1≤l≤Nβ,1≤α≠β≤n. $ |
Then, we are primarily concerned with the following issue:
● (Main issue): How can we find sufficient conditions for the non-emergence of mono-cluster flocking in the system (1.5)? Additionally, under what sufficient conditions with respect to the initial data and system parameters can mono-cluster flocking emerge in system (1.5)?
The paper is organized as follows. Section 2 introduces several basic estimates for temperatures in system (1.5) and previous results studied in [2]. Section 3 gives a mono-cluster flocking estimate different from the previous paper [3] and proves the non-emergence of mono-cluster flocking under suitable sufficient conditions when $ \phi $ is integrable in system (1.5). Next, we describe several sufficient frameworks for the mono-cluster flocking of system (1.5) when the communication weight $ \phi $ is non-integrable. Section 4 reorganizes system (1.5) to the multi-cluster setting and derives some dissipative structures on each cluster group to demonstrate the multi-cluster flocking of system (1.5) under admissible data. Finally, Section 5 briefly summarizes the main results and discusses the remaining issues left for future work.
Notation. Throughout the paper, we denote the following notation for brevity:
$ ‖⋅‖=standardl2-norm,⟨⋅,⋅⟩=standard inner product,yi=i-th component ofy∈Rd,X:=(x1,⋯,xN),V:=(v1,⋯,vN),T:=(T1,⋯,TN),R+:=[0,∞),DZ(t):=max1≤i,j≤N‖zi(t)−zj(t)‖forZ=(z1,⋯,zN)∈{X,V,T}. $ |
This section reviews several basic results for the subsystem (1.5c) to guarantee its global well-posedness; these estimates will be crucial throughout this paper. Afterward, we introduce the previous bi-cluster flocking results of system (1.5) studied in [2].
This subsection deals with the entropy principle, the propagation of conserved quantity, and the uniform boundedness of temperature to the subsystem (1.5c). For this, we begin with defining the entropy of system (1.5).
Definition 2.1. [26,31] Let $ \{(x_i, v_i, T_i)\}_{i = 1}^N $ be a solution to the system (1.5). Then, the entropy is defined as
$ \mathcal{S}(t): = \sum\limits_{i = 1}^{N}\ln(T_i(t)) = \ln\left(\prod\limits_{i = 1}^{N}T_i(t)\right). $ |
Then, we present the entropy principle and conserved temperature sum as below:
Proposition 2.1. [26,31] Assume that $ \{(x_i, v_i, T_i)\}_{i = 1}^N $ is a solution to the system (1.5). Then, one has the following two assertions:
1. (Conserved temperature sum) The total sum $ \sum_{i = 1}^{N}T_i $ is conserved for $ t\geq 0 $.
$ \sum\limits_{i = 1}^{N}T_i(t) = \sum\limits_{i = 1}^N T_i^{0} = NT^\infty. $ |
2. (Entropy principle) Entropy $ \mathcal{S} $ monotonically increases for $ t\geq0 $:
$ \frac{d\mathcal{S}}{dt} = \frac{\kappa_2}{2N}\sum\limits_{i, j = 1}^{N}\zeta(\|x_j-x_i\|)\left|\frac{1}{T_i}-\frac{1}{T_j}\right|^2\geq 0. $ |
Subsequently, we offer the following uniform boundedness consisting of strictly positive lower and upper bounds for temperatures to the system (1.5):
Proposition 2.2. [26](Uniform boundedness for temperatures) Let $ Z = \{(x_i, v_i, T_i)\}_{i = 1}^N $ be a solution to system (1.5). Then, $ \min_{1\leq i\leq N}T_i(t) $ monotonically increases and $ \max_{1\leq i\leq N}T_i(t) $ monotonically decreases in time. In other words, for $ t\geq 0 $,
$ 0 < \min\limits_{1\leq i\leq N}T_i^{0} = :T_m^\infty\leq T_i(t) \leq \max\limits_{1\leq i\leq N}T_i^{0} = :T_M^\infty, \quad i = 1, \cdots, N. $ |
Since Proposition 2.2 holds, $ \phi, \zeta $ are uniformly bounded, and the speed of each agent is unit. We directly obtain the well-posedness of system (1.5) from the standard Cauchy–Lipschitz theory.
This subsection introduces the previous mono-cluster flocking and bi-cluster flocking estimated in [2]. First, we revisit the following mono-cluster flocking of the system (1.5) verified in [3]:
Proposition 2.3. [2] (Mono-cluster flocking) Suppose that $ \{(x_i, v_i, T_i)\}_{i = 1}^N $ is a global-in-time solution to the system (1.5) with the initial data $ \{(x_i^{0}, v_i^{0}, T_i^{0})\}_{i = 1}^N $ and assume that there exists a positive constant $ D_{X}^\infty > 0 $ that satisfies
$ D2V(0)<T∞mϕ(D∞X)2T∞MandDX(0)+2T∞MDV(0)κ1ϕ(D∞X)<D∞X. $ | (2.1) |
Then, we get that for $ t\in \mathbb{R}_+ $,
$ D2V(t)<2D2V(0)andDX(t)<D∞X, $ |
which yields the following mono-cluster flocking estimate of system (1.5) for $ t\in \mathbb{R}_+ $:
$ D_V(t)\leq D_V(0)\exp\left(-\frac{\kappa_1\phi(D_X^\infty)}{2T_M^\infty}\right), \quad D_T(t)\leq D_T(0)\exp\left(-\frac{\kappa_2\zeta(D_X^\infty)}{(T_M^\infty)^2}t\right). $ |
However, in Theorem 3.1, we can attain another mono-cluster flocking dynamics of system (1.5) by reducing the higher-order dissipative differential inequality in terms of velocity in Proposition 3.1 to a suitable lower-order inequality.
Subsequently, to describe the results of extending the mono-cluster flocking of Proposition 2.3 to bi-cluster flocking, we describe the admissible set $ (\mathcal {H}) $ proposed in [2]; for two cluster groups $ Z_1 = \{(x_{1i}, v_{1i}, T_{1i})\}_{i = 1}^{N_1} $ and $ Z_2 = \{(x_{2j}, v_{2j}, T_{2j})\}_{j = 1}^{N_2} $, we set the following three configuration vectors:
$ A_{\alpha}: = (a_{\alpha1}, \cdots a_{\alpha N_{\alpha}})\quad \alpha = 1, 2, \; \text{where}\; A\in\{X, V, T\}, \; \; a\in \{x, v, T\}\; \; \text{and}\; \; A: = (A_1, A_2). $ |
Next, for $ \alpha\in \{1, 2\} $, we denote $ L^\infty $ diameters regarding position-velocity-temperature for each cluster group
$ D_{X_\alpha}: = \max\limits_{1\leq i, j\leq N_\alpha}\|x_{\alpha i}-x_{\alpha j}\|, \quad D_{V_\alpha}: = \max\limits_{1\leq i, j\leq N_\alpha}\|v_{\alpha i}-v_{\alpha j}\|, \quad D_{T_\alpha}: = \max\limits_{1\leq i, j\leq N_\alpha}|T_{\alpha i}-T_{\alpha j}| $ |
and we let
$ \mathcal{D}_{X}{: = } D_{X_1}+D_{X_2}, \quad \mathcal{D}_{V}{: = } D_{V_1}+D_{V_2}, \quad \mathcal{D}_{T}{: = }D_{T_1}+D_{T_2}. $ |
Then, the admissible set $ (\mathcal {H}) $ in terms of a system parameter and initial data is given by
$ (\mathcal {H}) = :\{(X(0), V(0), T(0))\in \mathbb{R}^{2dN}\times (\mathbb{R}_{+}-\{0\})^N\; |\; (\mathcal{H}_0), (\mathcal{H}_1), (\mathcal{H}_2)\; \text{and}\; (\mathcal{H}_3)\; \text{hold.}\} $ |
● $ (\mathcal{H}_0)\; (\text{Basic notation}) $: For simplicity, we set
$ Λ0:=2NT∞MDV(0)κ1min(N1,N2)ϕ(D∞X)+16N2(T∞M)2ϕ(r02)κ1(min(N1,N2))2(ϕ(D∞X))2T∞m+8NT∞M∫∞0ϕ(s+r02)dsmin(N1,N2)ϕ(D∞X)T∞m,r0:=min1≤i≤N1,1≤j≤N2(xk1i(0)−xk2j(0)),Λ1:=κ1min(N1,N2)ϕ(D∞X)2NT∞M,Λ2:=κ1N1NT∞mΛ1+κ1N2NT∞m∫∞0ϕ(s+r02)ds,Λ3:=κ1N2NT∞mΛ1+κ1N1NT∞m∫∞0ϕ(s+r02)ds,Λ4:=min(N1,N2)κ2ζ(D∞X)N(T∞M)2,Λ5:=2κ2(1T∞m−1T∞M). $ |
● $ (\mathcal{H}_1)\; (\text{Well prepared conditions}) $: There exists a strictly positive number $ \mathcal{D}_X^\infty > 0 $ such that
$ {\mathcal{D}_X^\infty > \mathcal{D}_X(0)+\Lambda_0\quad \mbox{and}\quad \phi\; \mbox{is integrable}\; \left( \iff \int_{0}^\infty\phi(s) ds < \infty\right).} $ |
● $ (\mathcal{H}_2)\; (\text{Separated initial data}) $: For $ k\in [d] $ fixed in $ \mathcal{H}_0 $, the initial data and system parameters are chosen to be properly partitioned as follows:
$ r_0 > 0, \quad v^k_{1i}(0)-\Lambda_2 > \frac{1}{2}, \quad v^k_{2j}(0)+\Lambda_3 < -\frac{1}{2}. $ |
● $ (\mathcal{H}_3)\; (\text{Small fluctuations and coupling strength}) $: The perturbation of local velocity in each cluster group and the coupling strength are sufficiently small:
$ 2κ1T∞m∫∞r02ϕ(s)ds<DV(0)≤√T∞mmin(N1,N2)ϕ(D∞X)2max(N1,N2)T∞M. $ |
When the admissible set $ (\mathcal {H}) $ is assumed, the author of [2] verified the following bi-cluster flocking of system (1.5):
Proposition 2.4. [2] (Bi-cluster flocking) Suppose that $ Z_1 = \{(x_{1i}, v_{1i}, T_{1i})\}_{i = 1}^{N_1} $ and $ Z_2 = \{(x_{2j}, v_{2j}, T_{2j})\}_{j = 1}^{N_2} $ are a global-in-time solution to the bi-cluster dynamical system (1.5). Further, assume that the admissible set $ (\mathcal{H}) $ is valid. Then, we can get the following bi-cluster flocking result in time.
1. $ \min\limits_{1\leq i\leq N_1, 1\leq j\leq N_2}\|x_{1i}-x_{2j}\|\geq t+\frac{r_0}{2}, \quad D_X(t) < D_X^\infty. $
2. $ D_V(t)\leq D_V(0)\exp\left(-\Lambda_1t\right)+\frac{2\kappa_1}{T_m^\infty\Lambda_1}\exp\left(-\frac{\Lambda_1}{2}t\right)\phi\left(\frac{r_0}{2}\right)+\frac{2\kappa_1}{T_m^\infty\Lambda_1}\phi\left(\frac{t+r_0}{2}\right). $
3. $ D_T(t)\leq D_T(0)\exp(-\Lambda_4t)+\Lambda_5\exp\left(-\frac{\Lambda_4}{2}t\right)\zeta\left(\frac{r_0}{2}\right)+\Lambda_5\zeta\left(\frac{t+r_0}{2}\right) $.
In Section 4, we extend the sufficient frameworks for the bi-cluster flocking of Proposition 2.4 to the multi-cluster flocking result.
This section provides suitable sufficient frameworks for the mono-cluster flocking and gives sufficient conditions to guarantee the non-emergence of mono-cluster flocking to system (1.5) when $ \phi $ is integrable. Finally, in the case of system (1.5) under non-integrable $ \phi $, we present a sufficient condition independent of coupling strength for mono-cluster flocking to arise.
This subsection recalls a dissipative structure for position-velocity-temperature $ L^\infty $-diameters derived in [2] and gives a mono-cluster flocking result different from Proposition 2.3 which is the mono-cluster flocking of system (1.5) proven in [2]. For this, we begin with the following dissipative inequalities for system (1.5):
Proposition 3.1. [2] Suppose that $ \{(x_i, v_i, T_i)\}_{i = 1}^N $ is a solution to the system (1.5). Then, we have that for a.e. $ t\in \mathbb{R}_+-\{0\} $,
$ \left|\frac{dD_X}{dt}\right|\leq D_V, \quad \frac{dD_V}{dt}\leq -\kappa_1\left(\frac{\phi(D_X)}{T_M^\infty}-\frac{D_V^2}{2T_m^\infty}\right)D_V, \quad \frac{dD_T}{dt}\leq -\frac{\kappa_2\zeta(D_X)}{(T_M^\infty)^2}D_T. $ |
Now, we are ready to study the new mono-cluster flocking result of system (1.5).
Theorem 3.1. (Mono-cluster flocking) Assume that $ \{(x_i, v_i, T_i)\}_{i = 1}^N $ is a solution to the system (1.5). Suppose that there exists a nonnegative number $ D_X^\infty\in \mathbb{R}_+ $ such that the following conditions hold:
$ D2V(0)<2ϕ(D∞X)T∞mT∞M,DX(0)−√T∞MT∞mκ1√2ϕ(D∞X)log(2ϕ(D∞X)T∞m−T∞MD2V(0)(√2ϕ(D∞X)T∞m+√T∞MDV(0))2)≤D∞X. $ | (3.1) |
Then, we attain the following assertions for $ t\in \mathbb{R}_+ $:
1. $ D_X(t)\leq D_X^\infty, $
2. $ D_V(t)\leq \left(\frac{T_M^\infty}{2\phi(D_X^\infty)T_m^\infty}+\left(\frac{1}{D^2_V(0)}-\frac{T_M^\infty}{2\phi(D_X^\infty)T_m^\infty}\right)\exp\left(\frac{2\kappa_1\phi(D_X^\infty)t}{T_M^\infty}\right)\right)^{-\frac{1}{2}}, $
3. $ D_T(t)\leq D_T(0)\exp\left(-\frac{\kappa_2\zeta(D^\infty_X)}{(T_M^\infty)^2}t\right). $
Proof. $ (i) $ (The case of $ D_V(t) > 0 $ for $ t\in\mathbb{R}_+ $) First, we set $ g(t) $ as
$ g(t) = \frac{1}{D^2_V(t)}. $ |
It follows from the second assertion of Proposition 3.1 that
$ dg(t)dt≥2κ1T∞Mϕ(DX(t))g(t)−κ1T∞m,a.e.t∈R+−{0}. $ | (3.2) |
Due to inequality (3.1) and the continuity of $ D_X $, the following set:
$ S: = \{s > 0\; |\; {\rm{ (1) \ holds\ for}}\; t\in(0, s)\} $ |
is nonempty and we denote $ t^*: = \sup S > 0 $. Next, we claim that
$ t^* = +\infty. $ |
For the proof by contradiction, suppose that $ t^* < \infty $. Then, we can obtain from inequality (3.2) and the definition of $ S $ that
$ \frac{dg(t)}{dt}\geq \frac{2\kappa_1}{T_M^\infty}\phi(D_X^\infty)g(t)-\frac{\kappa_1}{T_m^\infty}, \quad a.e.\; t\in (0, t^*). $ |
Moreover, using Grönwall's lemma with the above inequality yields that
$ g(t)≥T∞M2ϕ(D∞X)T∞m+(g(0)−T∞M2ϕ(D∞X)T∞m)exp(2κ1ϕ(D∞X)tT∞M),t∈[0,t∗]. $ |
This induces that for $ t\in [0, t^*] $,
$ DV(t)≤(T∞M2ϕ(D∞X)T∞m+(1D2V(0)−T∞M2ϕ(D∞X)T∞m)exp(2κ1ϕ(D∞X)tT∞M))−12. $ | (3.3) |
Accordingly, we combine inequality (3.3) with the first assertion of Proposition 3.1 to estimate that for $ t\in [0, t^*] $,
$ DX(t)≤DX(0)+∫t0DV(s)ds≤DX(0)+∫t0(T∞M2ϕ(D∞X)T∞m+(1D2V(0)−T∞M2ϕ(D∞X)T∞m)exp(2κ1ϕ(D∞X)sT∞M))−12ds<DX(0)+∫∞0(T∞M2ϕ(D∞X)T∞m+(1D2V(0)−T∞M2ϕ(D∞X)T∞m)exp(2κ1ϕ(D∞X)sT∞M))−12ds=DX(0)−√T∞MT∞mκ1√2ϕ(D∞X)log(2ϕ(D∞X)T∞m−T∞MD2V(0)(√2ϕ(D∞X)T∞m+√T∞MDV(0))2)≤D∞X, $ |
which contradicts to $ t^* < \infty $. Therefore, $ t^* = \infty $ and for $ t\in \mathbb{R}_+ $,
$ DX(t)≤D∞X. $ | (3.4) |
Hence, one has for $ t\in \mathbb{R}_+ $,
$ DV(t)≤(T∞M2ϕ(D∞X)T∞m+(1D2V(0)−T∞M2ϕ(D∞X)T∞m)exp(2κ1ϕ(D∞X)tT∞M))−12. $ |
In addition, because the third assertion of Proposition 3.1 and inequality (3.4) hold, we derive that for a.e. $ t\in \mathbb{R}_+-\{0\} $,
$ \frac{dD_T}{dt}\leq -\frac{\kappa_2\zeta(D_X)}{(T_M^\infty)^2}D_T\leq -\frac{\kappa_2\zeta(D_X^\infty)}{(T_M^\infty)^2}D_T, $ |
which implies that for $ t\in \mathbb{R}_+ $,
$ D_T(t)\leq D_T(0)\exp\left(-\frac{\kappa_2\zeta(D^\infty_X)}{(T_M^\infty)^2}\right). $ |
$ (ii) $ (The case of $ D_V(t) = 0 $ for some $ t\in \mathbb{R}_+ $) We define $ s_* $ by
$ s_*: = \inf\{t\in \mathbb{R}_+\; |\; D_V(t) = 0\}. $ |
Then, $ s_*\in \mathbb{R}_+ $ and applying the Cauchy–Lipschitz theory implies that
$ D_V(t) = 0, \quad t\geq s_*. $ |
Finally, if we follow the arguments employed in the first case, we immediately reach the desired mono-cluster flocking estimate.
Before we end this subsection, we provide the following remark:
Remark 3.1. Although $ \frac{T_m^\infty\phi(D_{X}^\infty)}{2T_M^\infty} $ of Eq (2.1) and $ \frac{2\phi(D_X^\infty)T_m^\infty}{T_M^\infty} $ of Eq (3.1) satisfy the following inequality for $ D_X^\infty\geq 0 $:
$ \frac{T_m^\infty\phi(D_{X}^\infty)}{2T_M^\infty}\leq\frac{2\phi(D_X^\infty)T_m^\infty}{T_M^\infty}, $ |
but the following term diverges to $ -\infty $ when $ 2\phi(D_X^\infty)T_m^\infty $ and $ T_M^\infty D^2_V(0) $ are close to each other in Eq (3.1):
$ \log\left(\frac{{2\phi(D_X^\infty)T_m^\infty}-{T_M^\infty D^2_V(0)}}{\left(\sqrt{{2\phi(D_X^\infty)T_m^\infty}}+\sqrt{{T_M^\infty}}D_V(0)\right)^2}\right). $ |
Thus, it is unknown which of Proposition 2.3 and Theorem 3.1 yields better mono-cluster flocking result.
This subsection guarantees the non-emergence of mono-cluster flocking of the system (1.5) with integrable $ \phi $ and sufficient small $ \kappa_1 $. For this, we employ the main strategies implemented in [29] for the targeted system (1.5).
This subsubsection offers basic notations and preliminary estimates to show the non-emergence of the mono-cluster flocking of system (1.5) when $ \phi $ is integrable. First, we consider the following subdivided $ n\geq 2 $ configurations $ \{Z_{\alpha}^0\}^n_{\alpha = 1} $ of $ Z^0 = \{(x_i^0, v_i^0, T_i^0)\}_{i = 1}^N $ satisfying
$ (x0αi,v0αi,T0αi),(x0αj,v0αj,T0αj)∈Z0α⟺v0αi=v0αj, $ |
where
$ |Z^0_\alpha| = :N_{\alpha}\geq 1, \quad Z^0 = \dot\cup_{\alpha = 1}^n Z_{\alpha}^0. $ |
In other words, we primarily deal with the initial configuration $ Z^0 $ that is not in a mono-cluster flocking state. Subsequently, we reorganize the system (1.5) to distinguish the $ n $-dynamics initiated from $ n $-subdivided initial configurations $ Z^0_\alpha $ as follows:
$ {dxαidt=vαi,t>0,i=1,⋯,Nα,α=1,⋯,n,n≥2,dvαidt=κ1NNα∑j=1ϕ(‖xαi−xαj‖)(vαj−⟨vαj,vαi⟩vαiTαj)+κ1N∑β≠αNβ∑j=1ϕ(‖xαi−xβj‖)(vβj−⟨vβj,vαi⟩vαiTβj),dTαidt=κ2NNα∑j=1ζ(‖xαi−xαj‖)(1Tαi−1Tαj)+κ2N∑β≠αNβ∑j=1ζ(‖xαi−xβj‖)(1Tαi−1Tβj),(xαi(0),vαi(0),Tαi(0))=(x0αi,v0αi,T0αi)∈Rd×Sd−1×(R+−{0}). $ | (3.5) |
In the following, we denote local averages and local deviations for $ \alpha = 1, \cdots, n $
$ x^{cen}_{\alpha} = \frac{1}{N_\alpha}\sum\limits_{i = 1}^{N_\alpha}x_{\alpha i}, \quad v^{cen}_{\alpha} = \frac{1}{N_\alpha}\sum\limits_{i = 1}^{N_\alpha}v_{\alpha i}, \quad \hat{x}_{\alpha i}: = x_{\alpha i}-x^{cen}_{\alpha}, \quad \hat{v}_{\alpha i}: = v_{\alpha i}-v^{cen}_{\alpha}, $ |
and we set the following notation to estimate the degree of separation between $ n $-subdivided initial configuration sets $ \{Z_{\alpha}^0\}^n_{\alpha = 1} $.
$ D(x0):=maxα≠β,i,j‖x0αi−x0βj‖,θ0:=minα≠βarccos⟨vcenα(0),vcenβ(0)⟩,λ0:=min(cos((δ+ϵ)θ0)−cos((1−4δ−ϵ)θ0),cos(δθ0)−cos((1−δ)θ0)−(D(x0)+2T0)⋅(N−1)κ1NT∞m), $ |
where two auxiliary parameters $ \epsilon, \delta $ $ \in (0, 1) $ will be specified later such that $ \lambda_0 > 0 $ in Section 3.2.2 and we define $ T_0 $ as
$ T_0: = \max\limits_{\alpha\neq\beta, i, j}\left\{0, -\frac{\langle x^0_{\alpha i}-x^0_{\beta j}, v^{cen}_{\alpha}(0)\rangle}{\lambda_0}\right\}. $ |
We observe that $ D(x^0) $, $ \theta_0 $ and $ \lambda_0 $ are dependent on given initial data non-mono-cluster flocking state. As we will see later, $ T_0 $ is indeed the time when two agents belonging to different cluster groups begin to move away from each other linearly and $ \lambda_0 $ is needed to estimate $ T_0 $. For the detailed descriptions, see Section 3.2.2.
Next, we set the coupling strength $ {\tilde{\kappa}_0} $ dependent on given initial data $ Z^0 = \{(x_i^0, v_i^0)\}_{i = 1}^N $ of the system (1.5) as follows:
$ (i) $ (The case of $ \min_{\alpha\neq\beta, i, j}\langle (x^0_{\alpha i}-x^0_{\beta j}), v_\alpha^{cen}\rangle < 0 $): We define $ {\tilde{\kappa}_0} $ as
$ ˜κ0=min(NT∞m(1−cos(δθ0))2(N−1)T0,NT∞m(cos(δθ0)−cos((1−δ)θ0)−λ0)(N−1)(D(x0)+2T0),λ0(cos(δθ0)−cos((δ+ϵ)θ0))(1−γN)∫∞0ϕ(s)ds),whereγN:=minαNαN. $ |
$ (ii) $ (The case of $ \min_{\alpha\neq\beta, i, j}\langle (x^0_{\alpha i}-x^0_{\beta j}), v_\alpha^{cen}\rangle\geq 0 $): We define $ {\tilde{\kappa}_0} $ as
$ ˜κ0=˜λ0(1−cos(˜δθ0))(1−γN)∫∞0ϕ(s)ds,where˜λ0:=cos(˜δθ0)−cos((1−˜δ)θ0). $ |
Herein, an auxiliary parameter $ \tilde{\delta}\in (0, 1) $ will be determined such that $ \tilde{\lambda}_0 > 0 $ later in Section 3.2.2.
Finally, we present the definitions of $ \triangle_{\alpha i, \beta j}(t) $ and $ v_\alpha^{min} $, which will be crucially used to verify the non-emergence of mono-cluster flocking in the system (1.5). We let
$ \triangle_{\alpha i, \beta j}(t): = \langle x_{\alpha i}(t)-x_{\beta j}(t), v^{cen}_{\alpha}(t)\rangle, \quad v_\alpha^{min}: = \min\limits_{1\leq i\leq N_\alpha}\langle v_{\alpha i}(t), e_{\alpha}(T_0)\rangle, $ |
where $ e_{\alpha}(t): = \frac{v^{cen}_{\alpha}(t)}{\|v^{cen}_{\alpha}(t)\|} $. Note that $ \triangle_{\alpha i, \beta j}(t) $ shows how well $ Z_\alpha(t) $ and $ Z_\beta(t) $ are separated from each other at time $ t $. Therefore, rigorous estimates concerning $ \triangle_{\alpha i, \beta j}(t) $ are important to obtain the non-emergence of mono-cluster flocking in the system (1.5).
In what follows, we demonstrate the non-emergence of the mono-cluster flocking of the TCSUS system (1.5). For this, we assume that $ T_0 > 0 $ throughout the subsubsection. If otherwise, it is a trivial case when $ T_0 = 0 $ (see Theorem 3.2). Now, we begin with the following preparatory lemmas:
Lemma 3.1. Suppose that $ Z_\alpha $ is a solution to the system (3.5) with given initial data $ Z^0_\alpha $ that is a non-mono-cluster flocking state for each $ \alpha\in \{1, \cdots, n\} $. Assume that there exists a positive number $ \delta\in \left(0, \frac{1}{3}\right) $ such that
$ 0 < \kappa_1 < \frac{NT_m^\infty(1-\cos(\delta \theta_0))}{2(N-1)T_0}. $ |
Then, one has for $ t\in [0, T_0] $ and $ \alpha\neq\beta $,
1. $ \langle v_{\alpha i}, v^{cen}_{\alpha}\rangle > \cos(\delta \theta_0) , \quad \langle v_{\beta j}, v^{cen}_{\alpha}\rangle < \cos((1-\delta) \theta_0) $,
2. $ \langle v_{\alpha i}, v_{\beta j}\rangle < \cos((1-\delta) \theta_0) , \quad \langle e_{\alpha}, e_{\beta} \rangle < \cos((1-3\delta)\theta_0) $.
Proof. To estimate the first assertion of (1), we first see that
$ dvαidt=κ1NN∑j=1ϕ(‖xαi−xj‖)(vj−⟨vj,vαi⟩vαiTj)=κ1NN∑j≠αiϕ(‖xαi−xj‖)(vj−⟨vj,vαi⟩vαiTj). $ |
Then, the triangle inequality and $ \phi\leq 1 $ yield that
$ \left\|\frac{dv_{\alpha i}}{dt}\right\|\leq \frac{(N-1)\kappa_1}{NT_m^\infty}, $ |
where we used Proposition 2.2 and $ \|v_j-\langle v_j, v_{\alpha i}\rangle v_{\alpha i}\|\leq 1. $ Thus, it follows that
$ \left|\frac{d}{dt}\langle v_{\alpha i}, v^{cen}_{\alpha}\rangle \right|\leq \frac{2(N-1)\kappa_1}{NT_m^\infty}, $ |
which implies by the condition for $ \kappa_1 $ and construction of $ Z^0_\alpha $ that for $ t\in [0, T_0] $,
$ ⟨vαi(t),vcenα(t)⟩=⟨vαi(0),vcenα(0)⟩+∫t0dds⟨vαi(s),vcenα(s)⟩ds≥⟨vαi(0),vcenα(0)⟩−2(N−1)κ1T0NT∞m=1−2(N−1)κ1T0NT∞m>cos(δθ0). $ |
To prove the second assertion of (1), we employ the same method as in the proof of the first assertion of (1) as follows:
$ \left\|\frac{dv_{\beta j}}{dt}\right\|\leq \frac{(N-1)\kappa_1}{NT_m^\infty}\quad \mbox{and then, }\quad \left|\frac{d}{dt}\langle v_{\beta j}, v^{cen}_{\alpha}\rangle \right|\leq \frac{2(N-1)\kappa_1}{NT_m^\infty}. $ |
From the definitions of $ Z^0_\alpha $ and $ \theta_0 $, we get that for $ t\in [0, T_0] $,
$ ⟨vβj,vcenα⟩≤⟨vβj(0),vcenα(0)⟩+2(N−1)κ1T0NT∞m=⟨vcenβ(0),vcenα(0)⟩+2(N−1)κ1T0NT∞m≤cos(θ0)+2(N−1)κ1T0NT∞m≤cos(θ0)+1−cos(δθ0)<cos((1−δ)θ0), $ |
where we used the assumption for $ \kappa_1 $. Next, following the proof of (1), we can also attain the first assertion of (2) for $ t\in [0, T_0] $:
$ \langle v_{\alpha i}, v_{\beta j}\rangle < \cos((1-\delta) \theta_0). $ |
Finally, to verify the second assertion of (2), we combine (1) and the first assertion of (2) to attain that for $ t\in [0, T_0] $,
$ arccos(⟨eα,eβ⟩)≥−arccos(⟨eα,vαi⟩)+arccos(⟨vαi,vβj⟩)−arccos(⟨vβj,eβ⟩)>(1−δ)θ0−2δθ0=(1−3δ)θ0. $ |
Therefore, $ \langle e_{\alpha}, e_{\beta} \rangle < \cos((1-3\delta)\theta_0) $ for $ t\in [0, T_0] $ and we conclude this lemma.
The following lemma plays a key role in deriving the desired result:
Lemma 3.2. Let $ Z_\alpha $ be a solution to the system (3.5) with given initial data $ Z^0_\alpha $ that is a non-mono-cluster flocking state for each $ \alpha = 1, \cdots, n $. Suppose that there exists a positive number $ \delta\in \left(0, \frac{1}{3}\right) $ such that
$ 0 < \kappa_1 < \min\left(\frac{NT_m^\infty(1-\cos(\delta \theta_0))}{2(N-1)T_0}, \frac{NT_m^\infty\left(\cos(\delta \theta_0)-\cos((1-\delta)\theta_0)-\lambda_0\right)}{(N-1)(D(x^0)+2T_0)} \right), \quad \lambda_0 > 0. $ |
Then, we obtain that
$ minα≠β,i,j△αi,βj(T0)>0. $ |
Proof. First, we note that
$ \|x_{\alpha i}(t)-x_{\beta j}(t)\| = \left\|x_{\alpha i}(0)-x_{\beta j}(0)+\int_{0}^t (v_{\alpha i}(s)-v_{\beta j}(s)) ds\right\|\leq D(x^0)+2T_0. $ |
Hence, we have from the arguments studied in Lemma 3.1 and the definition of $ \lambda_0 $ that
$ ddt△αi,βj=⟨vαi,vcenα⟩−⟨vβj,vcenα⟩+⟨xαi−xβj,˙vcenα⟩>cos(δθ0)−cos((1−δ)θ0)−(D(x0)+2T0)(N−1)κ1NT∞m≥λ0>0, $ |
which leads to the following result using the definition of $ T_0 $:
$ \triangle_{\alpha i, \beta j}(t) > \triangle_{\alpha i, \beta j}(0)+\lambda_0t\quad \mbox{and thus, }\quad \triangle_{\alpha i, \beta j}(T_0) > \triangle_{\alpha i, \beta j}(0)+\lambda_0T_0 > 0. $ |
From the above relation, we take $ \min_{\alpha\neq\beta, i, j} $ to derive that
$ \min\limits_{\alpha\neq\beta, i, j}\triangle_{\alpha i, \beta j}(T_0) > 0. $ |
We reach the desired lemma.
Subsequently, to prove the main result using the bootstrapping argument, we denote $ \bar{T}_0 $
$ \bar{T}_0: = \sup\left\{t\in (T_0, \infty)\; \bigg|\; \min\limits_{\alpha, i} \langle v_{\alpha i}(s), e_\alpha(T_0)\rangle > \cos((\delta+\epsilon)\theta_0), \quad s\in [T_0, t) \right\}, $ |
where an auxiliary parameter $ \epsilon\in (0, 1) $ will be determined in Lemma 3.3. Here, we observe from Lemma 3.2 that $ e_\alpha(T_0) $ is well-defined. In addition, $ \bar{T}_0 $ is well-defined due to Lemma 3.1. Indeed,
$ \langle v_{\alpha i}(T_0), e_\alpha(T_0)\rangle > \cos(\delta\theta_0) > \cos((\delta+\epsilon)\theta_0). $ |
From now on, we claim that
$ \bar{T}_0 = \infty. $ |
Lemma 3.3. Assume that $ Z_\alpha $ is a solution to the system (3.5) given initial data $ Z^0_\alpha $ that is a non-mono-cluster flocking state for each $ \alpha = 1, \cdots, n $. Suppose that there exist positive numbers $ \epsilon $ and $ \delta $ that satisfy
$ 0 < \delta < \frac{1-2\epsilon}{5}, \quad \epsilon\in\left(0, \frac{1}{2}\right), \quad 0 < \kappa_1 < \frac{NT_m^\infty(1-\cos(\delta \theta_0))}{2(N-1)T_0}, \quad \lambda_0 > 0. $ |
Then, for $ t\in [T_0, \bar{T}_0] $,
$ maxα,β,j⟨vβj,eα(T0)⟩<cos((1−4δ−ϵ)θ0),minα≠β,i,j⟨vαi−vβj,eα(T0)⟩>λ0. $ |
Proof. To get the first assertion, from the definition of $ \bar{T}_0 $ and Lemma 3.1, we estimate that
$ arccos(⟨vβj,eα(T0)⟩)≥arccos(⟨eβ(T0),eα(T0)⟩)−arccos(⟨vβj,eβ(T0)⟩)>(1−3δ)θ0−(δ+ϵ)θ0=(1−4δ−ϵ)θ0. $ |
This leads us to deduce that
$ \max\limits_{\alpha, \beta, j}\langle v_{\beta j}, e_\alpha(T_0) \rangle < \cos((1-4\delta-\epsilon)\theta_0). $ |
Additionally, the definition of $ \bar{T}_0 $ and the first assertion yield that
$ \min\limits_{\alpha\neq\beta, i, j}\langle(v_{\alpha i}-v_{\beta j}), e_\alpha(T_0)\rangle > \cos((\delta+\epsilon) \theta_0)-\cos((1-4\delta-\epsilon)\theta_0)\geq \lambda_0. $ |
We need the following lemma to verify that $ \bar{T}_0 = \infty $:
Lemma 3.4. Let $ Z_\alpha $ be a solution to the system (3.5) given initial data $ Z^0_\alpha $ that is a non-mono-cluster flocking state for each $ \alpha = 1, \cdots, n $. Assume that there exist positive numbers $ \epsilon $ and $ \delta $ that satisfy $ 0 < \delta < \frac{1-2\epsilon}{5}, \; \epsilon\in\left(0, \frac{1}{2}\right) $, and
$ 0 < \kappa_1 < \min\left(\frac{NT_m^\infty(1-\cos(\delta \theta_0))}{2(N-1)T_0}, \frac{NT_m^\infty\left(\cos(\delta \theta_0)-\cos((1-\delta)\theta_0)-\lambda_0\right)}{(N-1)(D(x^0)+2T_0)} \right), \quad \lambda_0 > 0. $ |
Then, we reach that
$ ϕM(t):=maxα≠β,i,jϕ(‖xβj−xαi‖)≤ϕ(λ0(t−T0)),t∈[T0,ˉT0). $ |
Proof. By applying Lemma 3.2 and Lemma 3.3, we induce that for $ t\in [T_0, \bar{T}_0) $,
$ ‖xαi−xβj‖≥⟨(xαi−xβj),eα(T0)⟩=⟨(xαi(T0)−xβj(T0)),eα(T0)⟩+∫tT0⟨(vαi(s)−vβj(s)),eα(T0)⟩ds>∫tT0⟨(vαi(s)−vβj(s)),eα(T0)⟩ds>λ0(t−T0). $ |
Then, this leads to the following result for $ t\in [T_0, \bar{T}_0) $ due to the monotonicity of $ \phi $:
$ ϕM(t):=maxα≠β,i,jϕ(‖xβj−xαi‖)≤ϕ(λ0(t−T0)). $ |
Hence, we conclude the desired lemma.
Subsequently, we estimate the time derivative of $ v^{min}_{\alpha} $ to demonstrate the main result.
Lemma 3.5. Let $ Z_\alpha $ be a solution to the system (3.5) given initial data $ Z^0_\alpha $ that is a non-mono-cluster flocking state for each $ \alpha = 1, \cdots, n $. Then, for $ \alpha = 1, \cdots, n $, it follows that for $ t\in [T_0, \bar{T}_0) $,
$ \dot{v}^{min}_\alpha\geq -\frac{\kappa_1(1-\gamma_N)\phi_M}{T_m^\infty}. $ |
Proof. First, we fix $ \alpha\in\{1, \cdots, n\} $; then, we select index $ i_\alpha: = i_\alpha(t)\in \{1, \cdots, N_\alpha\} $ at time $ t $ such that
$ {v}^{min}_\alpha = \langle v_{\alpha i_\alpha}, e_\alpha(T_0)\rangle. $ |
Then, if we use system (3.5), Proposition 2.2, and the definitions of $ i_\alpha $ and $ \bar{T}_0 $, we obtain that
$ ˙vminα=⟨˙vαiα,eα(T0)⟩=κ1NNα∑j=1⟨ϕ(‖xαiα−xαj‖)(vαj−⟨vαj,vαiα⟩vαiαTαj),eα(T0)⟩+κ1N∑β≠αNβ∑j=1⟨ϕ(‖xαiα−xβj‖)(vβj−⟨vβj,vαiα⟩vαiαTβj),eα(T0)⟩≥κ1N∑β≠αNβ∑j=1⟨ϕ(‖xαiα−xβj‖)(vβj−⟨vβj,vαiα⟩vαiαTβj),eα(T0)⟩≥−κ1ϕMT∞m⋅(N−Nα)N≥−κ1(1−γN)ϕMT∞m, $ |
where we employed
$ \|v_{\beta j}-\langle v_{\beta j}, v_{\alpha i_{\alpha}}\rangle v_{\alpha i_{\alpha}}\|\leq 1. $ |
Thus, we get the desired lemma.
Finally, we are ready to study the non-emergence of the mono-cluster flocking of system (3.5) under the integrable communication weight $ \phi $, i.e.,
$ \|\phi\|_{L_1} = \int_{0}^\infty\phi(s)ds < \infty. $ |
Theorem 3.2. (Non-emergence of mono-cluster flocking) Assume that $ Z_\alpha $ is a solution to the system (3.5) with given initial data $ Z^0_\alpha $ that is a non-mono-cluster flocking state for each $ \alpha = 1, \cdots, n $. Suppose that $ T_0 > 0 $ and there exist positive numbers $ \epsilon $ and $ \delta $ that satisfy $ 0 < \delta < \frac{1-2\epsilon}{5} $ and $ \epsilon\in\left(0, \frac{1}{2}\right) $ such that
$ 0 < \kappa_1 < {\tilde{\kappa}_0}, \quad \lambda_0 > 0. $ |
Then, we attain that
$ minα≠β,i,jsupt∈R+‖xαi−xβj‖=∞,minα≠β,i,jlim inft→∞‖vαi−vβj‖>0. $ |
Meanwhile, when $ T_0 = 0 $, we let $ \tilde{\lambda} > 0 $ and $ \tilde{\delta}\in \left(0, \frac{1}{2}\right) $. Then, we can reach the same results as above.
Proof. To demonstrate the desired results, we divide them by the following dichotomy:
$ T_0 > 0\quad \mbox{or}\quad T_0 = 0. $ |
$ (i)\; (\mbox{The case of}\; T_0 > 0) $ For the proof by contradiction, suppose that $ \bar{T}_0 < \infty $. Then, there exist $ \alpha\in\{1, \cdots, n\} $ and $ i_\alpha\in\{1, \cdots, N_\alpha\} $ such that
$ \langle v_{\alpha i_\alpha}(\bar{T}_0), e_\alpha(T_0)\rangle = \cos((\delta+\epsilon)\theta_0). $ |
Then, we use Lemmas 3.1, 3.4 and 3.5 to obtain that for $ t\in [T_0, \bar{T}_0] $,
$ ⟨vαiα,eα(T0)⟩≥vminα≥vminα(T0)−κ1(N−Nα)NT∞m∫tT0ϕM(s)ds≥cos(δθ0)−κ1(N−Nα)NT∞mλ0‖ϕ‖L1≥cos(δθ0)−κ1(1−γN)T∞mλ0‖ϕ‖L1>cos((δ+ϵ)θ0), $ |
which gives a contradiction; therefore, $ \bar{T}_0 = \infty $. Then, the second assertion of Lemmas 3.3 and 3.4 with $ \bar{T}_0 = \infty $ yield the desired result.
$ (ii)\; (\mbox{The case of}\; T_0 = 0) $ This case is trivial, but we provide the proof rigorously to compare with the proof regarding the first assertion. Let
$ {T}^*_0: = \sup\left\{t\in \mathbb{R}_+-\{0\}\; \bigg|\; \min\limits_{\alpha, i} \langle v_{\alpha i}, e_\alpha(0)\rangle > \cos(\tilde{\delta}\theta_0), \quad t\in [0, t) \right\}, \; \mbox{where}\; \tilde{\delta}\in\left(0, \frac{1}{2}\right). $ |
It follows from the definition of $ Z_\alpha $ that $ {T}^*_0 > 0 $ exists. For the proof by contradiction, suppose that $ T_0^* < \infty $. Next, we employ the same method as utilized in proof of the first assertion of Lemma 3.1 to estimate that
$ \langle v_{\beta j}(t) , e_\alpha(0)\rangle < \cos(1-\tilde{\delta})\theta_0, \quad t\in [0, T^*_0). $ |
Hence, we have
$ \min\limits_{\alpha\neq\beta, i, j}\langle v_{\alpha i}(t)-v_{\beta j}(t), e_\alpha(0)\rangle > \cos(\tilde{\delta}\theta_0)-\cos((1-\tilde{\delta})\theta_0) = :\tilde{\lambda}_0 > 0. $ |
Then, similarly to the proof of Lemma 3.4, one can show that
$ \phi_M(t)\leq \phi(\tilde{\lambda}_0t), \quad t\in [0, T^*_0] $ |
and thus, for $ t\in [0, T_0^*) $, we can get the following estimates using the same methodologies as in the proof of Lemma 3.5:
$ ⟨vαiα,eα(0)⟩≥vminα≥vminα(0)−κ1(N−Nα)NT∞m∫t0ϕM(s)ds=1−κ1(N−Nα)NT∞m∫t0ϕM(s)ds≥1−κ1(N−Nα)NT∞mλ0‖ϕ‖L1>cos(˜δθ0), $ |
which leads to a contradiction. Therefore, $ {T}^*_0 = \infty $. Finally, if the arguments of Lemmas 3.3 and 3.4 are applied to the case of $ T_0 = 0 $, we conclude the desired result.
This subsection demonstrates a different sufficient framework than Section 3.1 for mono-cluster flocking to emerge in the system (1.5) when $ \phi $ is non-integrable by using the previous results of [3].
Proposition 3.2. [3] Let $ \{(x_i, v_i, T_i)\}_{i = 1}^N $ be a solution to the system (1.5) such that
$ \mathcal{A}(v)(0): = \min\limits_{1\leq i, j\leq N}\langle v^0_i, v^0_j \rangle > 0, \quad D_V(0) < \frac{\kappa_1\mathcal{A}(v)(0)}{T_M^\infty}\int_{D_X(0)}^\infty \phi(s)ds. $ |
Then, there exists a nonnegative number $ D_X^\infty \in \mathbb{R}_+ $ satisfying for $ t\in \mathbb{R}_+ $,
1. (Group formation) $ D_X(t)\leq D_X^\infty $,
2. (Velocity alignment) $ D_V(t)\leq D_V(0)\exp\left(-\frac{\kappa_1\mathcal{A}(v)(0)\phi(D_X^\infty)}{T_M^\infty}t\right) $,
3. (Temperature equilibrium) $ D_T(t)\leq D_T(0)\exp\left(-\frac{\kappa_2\zeta(D_X^\infty)}{(T_M^\infty)^2}t\right). $
Proof. We employ the same methodologies as the proofs of Lemma 3.1 and Theorem 3.2 in [3] to obtain the desired result. Although the previous paper [3] dealt with the singular communication weight $ \phi $ to system (1.5), the proofs of Lemma 3.1 and Theorem 3.2 in [3] can be applied, even assuming the regular communication weight case covered in this paper.
Due to Proposition 3.2, we note the following remark.
Remark 3.2. It is easy to check that we can remove the condition,
$ D_V(0) < \frac{\kappa_1\mathcal{A}(v)(0)}{T_M^\infty}\int_{D_X(0)}^\infty \phi(s)ds, $ |
when $ \phi $ is non-integrable. In other words, when $ \phi $ is non-integrable, the mono-cluster flocking of the system (1.5) emerges under the only assumption $ \mathcal{A}(v)(0) > 0 $.
Finally, we present the following mono-cluster flocking of system (1.5) under non-integrable $ \phi $:
Theorem 3.3. (Mono-cluster flocking under non-integrable $ \phi $) Assume that $ \{(x_i, v_i, T_i)\}_{i = 1}^N $ is a solution to the system (1.5) under non-integrable $ \phi $ and suppose that
$ \mathcal{A}(v)(0): = \min\limits_{1\leq i, j\leq N}\langle v^0_i, v^0_j \rangle > 0. $ |
Then, there exists a nonnegative number $ D_X^\infty \in \mathbb{R}_+ $ such that for $ t\in \mathbb{R}_+ $,
1. (Group formation) $ D_X(t)\leq D_X^\infty $,
2. (Velocity alignment) $ D_V(t)\leq D_V(0)\exp\left(-\frac{\kappa_1\mathcal{A}(v)(0)\phi(D_X^\infty)}{T_M^\infty}t\right) $,
3. (Temperature equilibrium) $ D_T(t)\leq D_T(0)\exp\left(-\frac{\kappa_2\zeta(D_X^\infty)}{(T_M^\infty)^2}t\right). $
This section provides several sufficient frameworks for the multi-cluster flocking of the system (1.5). In Section 3, we studied that mono-cluster flocking does not occur when the coupling strength $ \kappa_1 $ is less than a certain positive value in system (1.5) with integrable $ \phi $. In Section 3.2.1, we employed suitable subdivided configurations, $ \{Z_{\alpha}^0\}^n_{\alpha = 1} $, so that all initial velocities are equal to each other in each group and deduced some sufficient conditions guaranteeing the non-emergence of the mono-cluster flocking of the system. Accordingly, we may wonder what the sufficient conditions are for multi-cluster flocking to occur, so it is necessary to check how little coupling strength is required for multi-cluster flocking to occur in system (1.5). To achieve this, we reorganize the system (1.5) under integrable $ \phi $ to a multi-cluster setting and then derive suitable dissipative differential inequalities with respect position–velocity–temperature. Finally, using bootstrapping arguments for these inequalities, we deduce appropriate sufficient conditions in terms of the initial data and system parameters to guarantee the mono-cluster flocking of system (1.5). As a direct consequence, we also prove that the velocity and temperature of all agents in each cluster group converge to the same values.
This subsection converts the TCSUS model (1.5) into some multi-cluster setting. Afterward, we present basic estimates for the averages of position-velocity-temperature. For this, we begin by reorganizing the system (1.5) to the following multi-cluster setting:
$ dxαidt=vαi,t>0,i∈{1,⋯,Nα},α∈{1,⋯,n},n≥3, $ | (4.1a) |
$ ˙vαi=κ1NNα∑j=1ϕ(‖xαi−xαj‖)(vαj−⟨vαi,vαj⟩vαi)Tαj $ | (4.1b) |
$ +κ1N∑β≠αNβ∑j=1ϕ(‖xαi−xβj‖)(vβj−⟨vαi,vβj⟩vαi)Tβj, $ | (4.1c) |
$ ˙Tαi=κ2NNα∑j=1ζ(‖xαi−xαj‖)(1Tαi−1Tαj)+κ2N∑β≠αNβ∑j=1ζ(‖xαi−xβj‖)(1Tαi−1Tβj), $ | (4.1d) |
$ (xαi(0),vαi(0),Tαi(0))∈Z0α×T0αi⊂Rd×Sd−1×(R+−{0}). $ | (4.1e) |
For each cluster group $ Z_\alpha = \{(x_{\alpha i}, v_{\alpha i}, T_{\alpha i})\}_{i = 1}^{N_\alpha} $, we denote the following three configuration vectors:
$ A_{\alpha}: = (a_{\alpha1}, \cdots a_{\alpha N_{\alpha}}), \; 1\leq\alpha\leq n, \; \text{where}\; A\in\{X, V, T\}, \; \; a\in \{x, v, T\}, \; A: = (A_1, \cdots, A_\alpha). $ |
Next, we define position-velocity-temperature $ L^\infty $-diameters to each cluster group as follows:
$ (i) $ (The position-velocity-temperature diameters to the $ \alpha $-th cluster group)
$ D_{X_\alpha}: = \max\limits_{1\leq i, j\leq N_\alpha}\|x_{\alpha i}-x_{\alpha j}\|, \quad D_{V_\alpha}: = \max\limits_{1\leq i, j\leq N_\alpha}\|v_{\alpha i}-v_{\alpha j}\|, \quad D_{T_\alpha}: = \max\limits_{1\leq i, j\leq N_\alpha}|T_{\alpha i}-T_{\alpha j}|. $ |
$ (ii) $ (The local averages of velocity and temperature in each cluster group)
$ {v_{\alpha}^{cen}}: = \frac{1}{N_\alpha }\sum\limits_{i = 1}^{N_\alpha }v_{\alpha i}, \quad {T_{\alpha }^{cen}}: = \frac{1}{N_\alpha}\sum\limits_{i = 1}^{N_\alpha}T_{\alpha i}. $ |
Before we end this subsection, we offer the following lemma regarding the local averages of velocity and temperature for each cluster group. This lemma will be crucially used to prove that the velocity and temperature of all agents in each cluster group converges to some unified values.
Lemma 4.1. Assume that $ Z_\alpha = \{(x_{\alpha i}, v_{\alpha i}, T_{\alpha i})\}_{i = 1}^{N_\alpha} $ is a solution to the system (4.1). Then, each local average $ (x_{\alpha}^{cen}, v_{\alpha}^{cen}, T_{\alpha}^{cen}) $ satisfies the following relations:
$ {dxcenαdt=vcenα,t>0,α∈{1,⋯,n},n≥3,Nα˙vcenα=κ1NNα∑i=1Nα∑j=1ϕ(‖xαi−xαj‖)vαi‖vαj−vαi‖22Tαj+κ1N∑β≠αNα∑i=1Nβ∑j=1ϕ(‖xαi−xβj‖)(vβj−vαi+vαi‖vβj−vαi‖22)1Tβj,Nα˙Tcenα=κ2N∑β≠αNα∑i=1Nβ∑j=1ζ(‖xαi−xβj‖)(1Tαi−1Tβj). $ |
Proof. The first assertion is trivial. For the second assertion, we take $ \sum_{i = 1}^{N_\alpha} $ to $ \dot{v}_{\alpha i} $ and use the standard trick of interchanging $ i $ and $ j $ and dividing $ 2 $ and
$ 1-\langle v_{\alpha i}, v_{\alpha j} \rangle = \frac{\|v_{\alpha i}-v_{\alpha j}\|^2}{2}. $ |
For the third assertion, we take $ \sum_{i = 1}^{N_\alpha} $ to $ \dot{T}_{\alpha i} $ and again use the standard trick as above.
In the following, we derive several dissipative differential inequalities with respect to position–velocity–temperature to obtain suitable sufficient frameworks in terms of system parameters and initial data for the multi-cluster flocking of system (4.1). For this, we define
$ D_{X}: = \sum\limits_{\alpha = 1}^nD_{X_\alpha}, \quad D_{V}: = \sum\limits_{\alpha = 1}^nD_{V_\alpha}, \quad D_{T}{: = }\sum\limits_{\alpha = 1}^n D_{T_\alpha}. $ |
Note that the above diameter functionals $ D_{X} $, $ D_{V} $ and $ D_{T} $ measure the total deviations of position, velocity and temperature to each cluster group $ Z_\alpha $, respectively.
To reduce the TCSUS system (4.1) to its appropriate dissipative structure, we employ the following functionals: For $ \alpha = 1, \cdots, n, $
$ \Phi_{\alpha ij}(t): = \frac{\phi(\|x_{\alpha i}-x_{\alpha j}\|)}{N_\alpha}+\left(1-\frac{\sum\limits_{j = 1}^{N_\alpha}\phi(\|x_{\alpha i}-x_{\alpha j}\|)}{N_\alpha}\right)\delta_{ij}, $ |
where $ \delta_{ij} $ denotes the Kronecker delta. Next, for simplicity, we set
$ \phi_{\alpha ij}: = \phi(\|x_{\alpha i}-x_{\alpha j}\|). $ |
Then, we can easily check that $ \Phi_{\alpha ij} $ satisfies the following properties:
1. $ \Phi_{\alpha ij}\geq \frac{\phi_{\alpha ij}}{N_\alpha}, \quad \sum\limits_{j = 1}^{N_\alpha}\Phi_{\alpha ij} = 1, \quad \Phi_{\alpha ij} = \Phi_{\alpha ji}, $
2. $ \sum\limits_{j = 1}^{N_\alpha}\Phi_{\alpha ij}\frac{(v_{\alpha j}-\langle v_{\alpha j}, v_{\alpha i}\rangle v_{\alpha i})}{T_{\alpha j}} = \sum\limits_{j = 1}^{N_\alpha}\frac{\phi_{\alpha ij}}{N_\alpha}\frac{(v_{\alpha j}-\langle v_{\alpha j}, v_{\alpha i}\rangle v_{\alpha i})}{T_{\alpha j}}. $
Similarly, we can observe that the functional $ \Psi_{\alpha ij} $ defined by
$ \Psi_{\alpha ij}(t): = \frac{\zeta(\|x_{\alpha i}-x_{\alpha j}\|)}{N_\alpha}+\left(1-\frac{\sum\nolimits_{j = 1}^{N_\alpha}\zeta(\|x_{\alpha i}-x_{\alpha j}\|)}{N_\alpha}\right)\delta_{ij}, \quad \zeta_{\alpha ij}: = \zeta(\|x_{\alpha i}-x_{\alpha j}\|) $ |
satisfies the following relations:
1. $ \Psi_{\alpha ij}\geq \frac{\zeta_{\alpha ij}}{N_\alpha}, \quad \sum\limits_{j = 1}^{N_\alpha}\Psi_{\alpha ij} = 1, \quad \Psi_{\alpha ij} = \Psi_{\alpha ji} $,
2. $ \sum\limits_{j = 1}^{N_\alpha}\Psi_{\alpha ij}\left(\frac{1}{T_{\alpha i}}-\frac{1}{T_{\alpha j}}\right) = \sum\limits_{j = 1}^{N_\alpha}\frac{\zeta_{\alpha ij}}{N_\alpha}\left(\frac{1}{T_{\alpha i}}-\frac{1}{T_{\alpha j}}\right). $
We note that the above functionals of this type have already been used several times in previous literature [2,6,25,27,28]. Unlike the aforementioned previous papers, the above functionals can be applied to a multi-cluster setting.
In what follows, we induce dissipative differential inequalities in terms of $ D_X $, $ D_V $ and $ D_T $, respectively, to deduce several sufficient frameworks for the multi-cluster flocking estimate of system (4.1).
Lemma 4.2. (Dissipative structure) Suppose that $ Z_\alpha = \{(x_{\alpha i}, v_{\alpha i}, T_{\alpha i})\}_{i = 1}^{N_\alpha} $ is a solution to the system (4.1). If we set $ \phi_M $ and $ \zeta_M $ as
$ \phi_M(t): = \max\limits_{\alpha\neq \beta, i, j}\phi(\|x_{\beta j}-x_{\alpha i}\|), \quad \zeta_M(t): = \max\limits_{\alpha\neq \beta, i, j}\zeta(\|x_{\beta j}-x_{\alpha i}\|). $ |
Then, we have the following three differential inequalities for a.e. $ t\in \mathbb{R}_+-\{0\} $:
1. $ \left|\frac{dD_X}{dt}\right|\leq D_V $,
2. $ \frac{dD_V}{dt}\leq -\frac{\kappa_1\min(N_1, \cdots, N_\alpha)\phi(D_{X})}{NT_M^\infty}D_{V}+\frac{\kappa_1\max(N_1, \cdots, N_\alpha)D^3_V}{2NT_m^\infty}+\frac{2\kappa_1({n}-1)\phi_M}{T_m^\infty}, $
3. $ \frac{dD_{T}}{dt}\leq -\frac{\kappa_2\min(N_1, \cdots, N_\alpha)\zeta(D_X)}{N(T_M^\infty)^2}D_{T}+{2\kappa_2({n}-1)\zeta_M}\left(\frac{1}{T_m^\infty}-\frac{1}{T_M^\infty}\right). $
Proof. Cauchy–Schwarz's inequality immediately yields the first assertion. Next, to prove the third assertion, we choose two indices, $ M_t $ and $ m_t $, depending on $ t $, such that
$ D_{T_\alpha}(t) = T_{\alpha M_t}(t)-T_{\alpha m_t}(t), \quad 1\leq m_t, M_t\leq N_\alpha. $ |
Now, we recall the subsystem (4.1c) as follows:
$ ˙Tαi=κ2NNα∑j=1ζ(‖xαi−xαj‖)(1Tαi−1Tαj)+κ2N∑β≠αNβ∑j=1ζ(‖xαi−xβj‖)(1Tαi−1Tβj). $ |
Then, for a.e. $ t\in \mathbb{R}_+-\{0\} $, one can show that by using the definitions of $ M_t $ and $ m_t $
$ dDTαdt=κ2NNα∑j=1ζ(‖xαMt−xαj‖)(1TαMt−1Tαj)−κ2NNα∑j=1ζ(‖xαmt−xαj‖)(1Tαmt−1Tαj)+κ2N∑β≠αNβ∑j=1ζ(‖xαMt−xβj‖)(1TαMt−1Tβj)−κ2N∑β≠αNβ∑j=1ζ(‖xαmt−xβj‖)(1Tαmt−1Tβj)=:I1+I2+I3+I4. $ |
$ (i) $ (Estimate of $ \mathcal{I}_1 $+$ \mathcal{I}_2 $) Similar to the proof of Lemma 3.2 in [2], for a.e. $ t\in \mathbb{R}_+-\{0\} $,
$ I1+I2≤−κ2Nαζ(DXα)N(T∞M)2DTα. $ |
$ (ii) $ (Estimate of $ \mathcal{I}_3 $+$ \mathcal{I}_4 $) From Proposition 2.2 and the definitions of $ \phi_M $ and $ \phi_m $, for a.e. $ t\in \mathbb{R}_+-\{0\} $,
$ I3+I4≤κ2N|∑β≠αNβ∑j=1ζ(‖xαMt−xβj‖)(1TαMt−1Tβj)|+κ2N|∑β≠αNβ∑j=1ζ(‖xαmt−xβj‖)(1Tαmt−1Tβj)|≤2κ2(N−Nα)ζMN(1T∞m−1T∞M). $ |
Thus, combining $ \mathcal{I}_1 $+$ \mathcal{I}_2 $ and $ \mathcal{I}_3 $+$ \mathcal{I}_4 $ yields that for a.e. $ t\in \mathbb{R}_+-\{0\} $,
$ \frac{dD_{T_\alpha}}{dt}\leq -\frac{\kappa_2N_{\alpha}\zeta(D_{X_\alpha})}{N(T_M^\infty)^2}D_{T_\alpha}+\frac{2\kappa_2(N-N_\alpha)\zeta_M}{N}\left(\frac{1}{T_m^\infty}-\frac{1}{T_M^\infty}\right). $ |
Therefore, we take the summation from $ \alpha $ = 1 to $ n $ to the above inequality to get that for a.e. $ t\in \mathbb{R}_+-\{0\} $,
$ \frac{dD_{T}}{dt}\leq -\frac{\kappa_2\min(N_1, \cdots, N_\alpha)\zeta(D_X)}{N(T_M^\infty)^2}D_{T}+{2\kappa_2({n}-1)\zeta_M}\left(\frac{1}{T_m^\infty}-\frac{1}{T_M^\infty}\right). $ |
To verify the second assertion, we select two indices $ M_t $ and $ m_t $ depending on $ t $ satisfying
$ D_{V_\alpha}(t) = \|v_{\alpha M_t}(t)-v_{\alpha m_t}(t)\|, \quad 1\leq m_t, M_t\leq N_\alpha. $ |
We recall the following velocity coupling Eq (4.1b):
$ ˙vαi=κ1NNα∑j=1ϕ(‖xαi−xαj‖)(vαj−⟨vαi,vαj⟩vαi)Tαj+κ1N∑β≠αNβ∑j=1ϕ(‖xαi−xβj‖)(vβj−⟨vαi,vβj⟩vαi)Tβj. $ |
Hence, we attain that for a.e. $ t\in \mathbb{R}_+-\{0\} $,
$ 12dD2Vαdt=⟨vαMt−vαmt,˙vαMt−˙vαmt⟩=⟨vαMt−vαmt,κ1NNα∑j=1ϕαMtj(vαj−⟨vαMt,vαj⟩vαMtTαj)−κ1NNα∑j=1ϕαmtj(vαj−⟨vαmt,vαj⟩vαmtTαj)⟩+⟨vαMt−vαmt,κ1N∑β≠αNβ∑j=1ϕ(‖xαMt−xβj‖)(vβj−⟨vαMt,vβj⟩vαMtTβj)−κ1N∑β≠αNβ∑j=1ϕ(‖xαmt−xβj‖)(vβj−⟨vαmt,vβj⟩vαmtTβj)⟩=:J1+J2. $ |
$ (iii) $ (Estimate of $ \mathcal{J}_1 $) In the same way as the proof of Lemma 3.2 of [2], for a.e. $ t\in \mathbb{R}_+-\{0\} $,
$ \mathcal{J}_1\leq -\frac{\kappa_1N_\alpha}{N}\left(\frac{\phi(D_{X_\alpha})}{T_M^\infty}-\frac{D_{V_\alpha}^2}{2T_m^\infty}\right)D^2_{V_\alpha}. $ |
$ (iiii) $ (Estimate of $ \mathcal{J}_2 $) We employ the following identities:
$ \|{v_{\beta j}-\langle v_{\alpha M_t}, v_{\beta j}\rangle v_{\alpha M_t}}\|\leq 1, \quad \|{v_{\beta j}-\langle v_{\alpha m_t}, v_{\beta j}\rangle v_{\alpha m_t}}\|\leq 1 $ |
with Cauchy–Schwarz's inequality and Proposition 2.2 to estimate that for a.e. $ t\in \mathbb{R}_+-\{0\} $,
$ J2≤κ1DVαN‖∑β≠αNβ∑j=1ϕ(‖xαMt−xβj‖)(vβj−⟨vαMt,vβj⟩vαMtTβj)‖+κ1DVαN‖∑β≠αNβ∑j=1ϕ(‖xαmt−xβj‖)(vβj−⟨vαmt,vβj⟩vαmtTβj)‖≤2κ1(N−Nα)ϕMDVαNT∞m. $ |
Then, we combine $ \mathcal{J}_1 $ and $ \mathcal{J}_2 $ to derive that for a.e. $ t\in \mathbb{R}_+-\{0\} $,
$ dDVαdt≤−κ1NαN(ϕ(DXα)T∞M−D2Vα2T∞m)DVα+2κ1(N−Nα)ϕMNT∞m. $ |
We take the summation from $ \alpha $ = 1 to $ n $ to the above inequality to obtain that
$ dDVdt≤−κ1min(N1,⋯,Nα)ϕ(DX)NT∞MDV+κ1max(N1,⋯,Nα)D3V2NT∞m+2κ1(n−1)ϕMT∞m, $ |
because the monotonicity of $ \phi $ implies that
$ D^3_V\geq \sum\limits_{\alpha}D^3_{V_\alpha}, \quad \min(\phi(D_{X_1}), \cdots, \phi(D_{X_\alpha}))\geq \phi(D_X). $ |
Finally, we demonstrate the second assertion.
Remark 4.1. In Lemma 4.2, the two terms below
$ −κ1min(N1,⋯,Nα)ϕ(DX)NT∞MDV+κ1max(N1,⋯,Nα)D3V2NT∞m,−κ2min(N1,⋯,Nα)ζ(DX)N(T∞M)2DT $ |
are related to the velocity alignment and temperature equilibrium for each cluster group of system (4.1), respectively. Meanwhile, the following terms in Lemma 4.2
$ \frac{2({n}-1)\phi_M\kappa_1}{T_m^\infty}, \qquad {2({n}-1)\zeta_M\left(\frac{1}{T_m^\infty}-\frac{1}{T_M^\infty}\right)\kappa_2} $ |
show the tendency of the velocities and temperatures of system (4.1) to separated into $ n $ multi-cluster groups in system (4.1).
This subsection describes suitable sufficient frameworks $ (\mathcal {H}) $ for the multi-cluster flocking estimate and then, under $ (\mathcal {H}) $, we demonstrate the multi-cluster flocking of the proposed system (4.1). For this, we first display the admissible data $ (\mathcal {H}) $ as follows:
$ (\mathcal {H}): = \{(X(0), V(0), T(0))\in \mathbb{R}^{2dN}\times (\mathbb{R}_{+}-\{0\})^N\; |\; (\mathcal{H}_0), (\mathcal{H}_1), (\mathcal{H}_2)\; \text{and}\; (\mathcal{H}_3)\; \text{hold.}\} $ |
$ (i) $ $ (\mathcal{H}_0)\; (\text{Notation}) $: For brevity, we denote the following notation:
$ Λ:=DV(0)Λ0+4(n−1)κ1T∞mΛ20ϕ(r02)+4(n−1)κ1(min1≤α≤n−1d(Iα,Iα+1))T∞mΛ0∫∞r02ϕ(s)ds,δ∈(0,1),ˉΛ0:=κ2min(N1,⋯,Nα)ζ(D∞X)N(T∞M)2,Λ0:=δκ1min(N1,⋯,Nα)ϕ(D∞X)NT∞M,Λα:=κ1NαNT∞mΛ0+κ1(N−Nα)NT∞m(min1≤α≤n−1d(Iα,Iα+1))∫∞r02ϕ(s)ds,r0:=minα<β,i,j(xkβj(0)−xkαi(0))for some fixed 1≤k≤d. $ |
$ (ii) $ $ (\mathcal{H}_1)\; (\text{Well prepared conditions}) $: There exists a strictly positive number $ D_X^\infty > 0 $ such that
$ D_X^\infty\geq D_X(0)+\Lambda, \quad \mbox{and} \quad \phi\ {\rm{is}}\ {\rm{integrable}}\; \; \text{i.e., }\; \; \|\phi\|_{L^1} < \infty . $ |
$ (iii) $ $ (\mathcal{H}_2)\; (\text{Separated initial data}) $: For fixed $ 1\leq k\leq d $ in $ (\mathcal{H}_0) $, there exist real sequences $ ({a_i})^n_{i = 1} $ and $ ({b_i})^n_{i = 1} $ such that the initial data and system parameters are selected to be split suitably as follows:
$ r0>0,a1<b1<a2<b2⋯<an<bn,Iα:=[aα,bα]⊂[−1,1],Iα∩Iβ=∅(β≠α),[vkαi(0)−Λα,vkαi(0)+Λα]⊂Iα:=[aα,bα]⊂[−1,1],α,β=1,⋯,n,i=1,⋯,Nα. $ |
$ (iii) $ $ (\mathcal{H}_3)\; (\text{Small fluctuations and coupling strength}) $: The local velocity perturbation for each cluster group and coupling strength are sufficiently small as follows:
$ 2κ1(√1+δ+1)(n−1)∫∞r02ϕ(s)dsδ(min1≤α≤n−1d(Iα,Iα+1))T∞m<DV(0)≤√2(1−δ)ϕ(D∞X)min(N1,⋯,Nα)T∞m(1+δ)max(N1,⋯,Nα)T∞M. $ |
Next, we give a brief comment regarding $ (\mathcal{H}) $. The assumption $ (\mathcal{H}_1) $ is that the sufficient condition guarantees a group formation to each cluster group. Note that $ (\mathcal{H}_2) $ implies that position initial data for each cluster group should be sufficiently separate from each other to verify the multi-cluster flocking result. Indeed, if $ v^k_{\alpha i}(0) $ is covered by $ I_\alpha: = [a_{\alpha}, b_{\alpha}] $, then we take sufficiently small $ \kappa_1 $ so that $ [v^k_{\alpha i}(0)-\Lambda_\alpha, v^k_{\alpha i}(0)+\Lambda_\alpha]\subset I_\alpha $ because $ \Lambda_\alpha $ is linearly proportional to $ \kappa_1 $. $ (\mathcal{H}_3) $ describes that the velocity perturbation between each pair of cluster groups is sufficiently small to deduce the velocity alignment for each cluster group. Here, we can find the admissible data satisfying the assumption $ (H3) $ when $ \kappa_1 $ is sufficiently small. Moreover, under sufficiently large $ r_0 $ and suitable temperature initial data and small coupling strength regime, we can check that the sufficient framework $ (\mathcal {H}) $ is admissible data.
To prove the multi-cluster flocking result, we define the following set:
$ S: = \left\{{s} > 0\; \bigg|\; \min\limits_{\alpha\neq\beta, i, j}\|x_{\alpha i}(t)-x_{\beta j}(t)\|\geq \left(\min\limits_{1\leq \alpha\leq n-1}d(I_\alpha, I_{\alpha+1})\right)t+\frac{r_0}{2}, \quad t\in[0, {s})\right\}, $ |
where $ d(I_\alpha, I_{\alpha+1}) $ is a distance between adjacent intervals $ I_\alpha $ and $ I_{\alpha+1} $. Herein, we observe that $ S $ is nonempty due to the assumption $ (\mathcal{H}_2) $ and the continuity of $ \|x_{\alpha i}(t)-x_{\beta j}(t)\| $, and we set
$ \sup{S} = :T^*. $ |
Lemma 4.3. Assume that $ Z_\alpha = \{(x_{\alpha i}, v_{\alpha i}, T_{\alpha i})\}_{i = 1}^{N_\alpha} $ is a solution to the system (4.1). Suppose that $ (\mathcal{H}_0) $, $ (\mathcal{H}_1) $, and $ (\mathcal{H}_3) $ hold. Then, it follows that
$ DV(t)<√(1+δ)DV(0),DX(t)≤D∞X,t∈[0,T∗]. $ | (4.2) |
Proof. First, we consider
$ S': = \left\{{s} > 0\; \bigg|\; \text{the desired estimates Eq (4.2) hold}, \quad t\in[0, {s}], \quad {s}\leq T^*\right\}. $ |
Let $ \sup{S'} = :T^{**} $ and suppose that $ T^{**} < T^{*} $ for the proof by contradiction. Then, one has for $ t\in [0, T^{**}] $, {
$ D2V(t)≤(1+δ)D2V(0)≤2(1−δ)ϕ(D∞X)min(N1,⋯,Nα)T∞mmax(N1,⋯,Nα)T∞M $ |
and
$ −κ1min(N1,⋯,Nα)ϕ(DX)NT∞M≤−κ1min(N1,⋯,Nα)ϕ(D∞X)NT∞M. $ |
Then, for a.e. $ t\in (0, T^{**}) $, the second assertion of Lemma 4.2 and the above estimates lead to the following inequalities:
$ dDVdt≤−κ1min(N1,⋯,Nα)ϕ(DX)NT∞MDV+κ1max(N1,⋯,Nα)D3V2NT∞m+2κ1(n−1)ϕMT∞m≤−δκ1min(N1,⋯,Nα)ϕ(D∞X)NT∞MDV+2κ1(n−1)ϕMT∞m=−Λ0DV+2κ1(n−1)ϕMT∞m. $ |
This gives from Grönwall's lemma that for $ t\in [0, T^{**}] $,
$ DV(t)≤DV(0)exp(−Λ0t)+2κ1(n−1)T∞mΛ0exp(−Λ02t)ϕ(r02)+2κ1(n−1)T∞mΛ0ϕ((min1≤α≤n−1d(Iα,Iα+1))t+r02), $ | (4.3) |
where we used the definition of $ S $ and the fact that
$ \phi_M\leq \phi\left(\left(\min\limits_{1\leq \alpha\leq n-1}d(I_\alpha, I_{\alpha+1})\right)t+\frac{r_0}{2}\right). $ |
Moreover, we again employ Grönwall's lemma to reach that for $ t\in [0, T^{**}] $,
$ DV(t)≤DV(0)exp(−Λ0t)+2κ1(n−1)(min1≤α≤n−1d(Iα,Iα+1))T∞m∫∞r02ϕ(s)ds. $ | (4.4) |
Next, using the definition of $ T^{**} $ yields that
$ D_V^2(T^{**}) = (1+\delta)D_V^2(0)\quad\text{or} \quad D_X(T^{**}) = D_X^\infty. $ |
In the former case, it is contradictory to $ (\mathcal{H}_3) $ because inequality (4.4) implies that
$ DV(T∗∗)=√1+δDV(0)≤DV(0)+2κ1(n−1)(min1≤α≤n−1d(Iα,Iα+1))T∞m∫∞r02ϕ(s)ds. $ |
In the latter case, we estimate from inequality (4.3) that for $ t\in [0, T^{**}] $,
$ DX(t)≤DX(0)+∫t0DV(s)ds≤DX(0)+∫t0[DV(0)exp(−Λ0s)+2κ1(n−1)T∞mΛ0exp(−Λ02s)ϕ(r02)+2κ1(n−1)T∞mΛ0ϕ((min1≤α≤n−1d(Iα,Iα+1))s+r02)]ds<DX(0)+Λ≤D∞X. $ | (4.5) |
Accordingly, $ D_X(T^{**}) < D_X^\infty $, which is contradictory. Finally, $ \sup S' = T^{**} = T^* $. We have reached the desired lemma.
Subsequently, we claim that $ T^* = \infty $, which is crucial to derive the multi-cluster flocking estimate of the system (4.1).
Theorem 4.1. Following Lemma 4.3, we further assume that $ (\mathcal{H}_2) $ holds. Then, we get that
$ T^* = \infty. $ |
This is equivalent to
$ \min\limits_{\alpha\neq\beta, i, j}\|x_{\alpha i}(t)-x_{\beta j}(t)\|\geq \left(\min\limits_{1\leq \alpha\leq n-1}d(I_\alpha, I_{\alpha+1})\right)t+\frac{r_0}{2}, \quad t\in \mathbb{R}_+. $ |
Proof. For the proof by contradiction, suppose that $ T^* < \infty $. From the definition of $ S $, we select four indices that satisfy
$ 1\leq \alpha^* < \beta^*\leq n, \quad i^*\in\{1, \cdots, N_{\alpha^*}\} \quad \mbox{and} \quad j^*\in\{1, \cdots, N_{\beta^*}\} $ |
such that
$ \|x_{\alpha^* i^*}(T^*)-x_{\beta^* j^*}(T^*)\| = \left(\min\limits_{1\leq \alpha\leq n-1}d(I_\alpha, I_{\alpha+1})\right)T^*+\frac{r_0}{2}. $ |
Then, we show that for the $ k\in \{1, \cdots, d\} $ chosen in $ (\mathcal{H}_0) $,
$ ‖xα∗i∗(T∗)−xβ∗j∗(T∗)‖≥xkβ∗j∗(T∗)−xkαi∗(T∗)=xkβ∗j∗(0)−xkα∗i∗(0)+∫T∗0(vkβ∗j∗(t)−vkα∗i∗(t))dt≥r0+∫T∗0(vkβ∗j∗(t)−vkα∗i∗(t))dt. $ |
Next, we integrate system (4.1b) and employ the following relation:
$ \|v_{\alpha j}-\langle v_{\alpha i}, v_{\alpha j} \rangle v_{\alpha i}\|^2 = 1-\langle v_{\alpha i}, v_{\alpha j}\rangle^2 = (1-\langle v_{\alpha i}, v_{\alpha j} \rangle)(1+\langle v_{\alpha i}, v_{\alpha j}\rangle)\leq D^2_{V_\alpha} $ |
to attain that for $ t\in [0, T^*] $,
$ |vkαi(t)−vkαi(0)|≤‖vαi(t)−vαi(0)‖≤∫t0‖˙vαi‖ds≤κ1NαNT∞m∫t0DVα(s)ds+κ1(N−Nα)NT∞m∫t0ϕM(s)ds≤κ1NαNT∞m∫∞0DVα(s)ds+κ1(N−Nα)NT∞m∫∞0ϕ(min1≤α≤n−1d(Iα,Iα+1)s+r02)ds≤κ1NαNT∞m∫∞0DV(s)ds+κ1(N−Nα)NT∞m∫∞0ϕ(min1≤α≤n−1d(Iα,Iα+1)s+r02)ds≤κ1NαNT∞mΛ+κ1(N−Nα)NT∞m∫∞0ϕ(min1≤α≤n−1d(Iα,Iα+1)s+r02)ds=κ1NαNT∞mΛ+κ1(N−Nα)NT∞m(min1≤α≤n−1d(Iα,Iα+1))∫∞r02ϕ(s)ds=:Λα, $ |
where we used $ \phi\leq 1 $, $ \|v_{\beta j}-\langle v_{\alpha i}, v_{\beta j} \rangle v_{\alpha i}\|\leq 1 $, and $ \Lambda $ was estimated in inequality (4.5). Therefore, it follows by $ (\mathcal{H}_2) $ that for $ \alpha = 1, \cdots, n $,
$ vkαi(0)+Λα≥vkαi(0)+|vkαi(t)−vkαi(0)|≥vkαi(t)=vkαi(0)+vkαi(t)−vkαi(0)≥vkαi(0)−|vkαi(t)−vkαi(0)|≥vkαi(0)−Λα⟹vkαi(t)∈Iα. $ |
Then, we derive that using the assumption $ (\mathcal{H}_2) $,
$ ‖xα∗i∗(T∗)−xβ∗j∗(T∗)‖≥r0+∫T∗0(vkβ∗j∗(t)−vkα∗i∗(t))dt>r02+min1≤α≤n−1d(Iα,Iα+1)T∗, $ |
which gives a contradiction to $ T^* < \infty $. Consequently, we conclude that $ T^* = \infty. $
Now, we are ready to prove the multi-cluster flocking dynamics under sufficient framework $ (\mathcal{H}) $ by applying Lemma 4.3 and Theorem 4.1. In addition, we verify that there exist common velocity and temperature convergence values depending on the decay rates of the integrable communication weights $ \phi $ and $ \zeta $, respectively, in each cluster group.
Theorem 4.2. Let $ Z_\alpha = \{(x_{\alpha i}, v_{\alpha i}, T_{\alpha i})\}_{i = 1}^{N_\alpha} $ be a solution to the system (4.1) and suppose that the frameworks $ (\mathcal{H}_0) $, $ (\mathcal{H}_1) $, $ (\mathcal{H}_2) $, and $ (\mathcal{H}_3) $ hold. Then, we obtain the following assertions for $ t\in \mathbb{R}_+ $:
1. (Velocity alignment for each cluster group)
$ DV(t)≤DV(0)exp(−Λ0t)+2κ1(n−1)T∞mΛ0exp(−Λ02t)ϕ(r02)+2κ1(n−1)T∞mΛ0ϕ((min1≤α≤n−1d(Iα,Iα+1))t+r02). $ |
2. (Temperature equilibrium for each cluster group)
$ DT(t)≤DT(0)exp(−ˉΛ0t)+2κ2(n−1)(1T∞m−1T∞M)exp(−ˉΛ02t)ζ(r02)+2κ2(n−1)(1T∞m−1T∞M)ζ((min1≤α≤n−1d(Iα,Iα+1))t+r02). $ |
Proof. We apply the second assertion of Lemma 4.2, the definition of the set $ S $, and Theorem 4.1 to have that for a.e. $ t\in \mathbb{R}_+-\{0\} $,
$ dDVdt≤−Λ0DV+2κ1(n−1)T∞mϕM≤−Λ0DV+2κ1(n−1)T∞mϕ(r02+min1≤α≤n−1d(Iα,Iα+1)t). $ |
From inequality (4.3), we recall that for $ t\in \mathbb{R}_+ $,
$ DV(t)≤DV(0)exp(−Λ0t)+2κ1(n−1)T∞mΛ0exp(−Λ02t)ϕ(r02)+2κ1(n−1)T∞mΛ0ϕ((min1≤α≤n−1d(Iα,Iα+1))t+r02). $ |
Hence, we reach the desired first assertion. To prove the second assertion, we employ the third assertion of Lemma 4.2, Theorem 4.1, and the second assertion of Lemma 4.3 to get that for a.e. $ t\in \mathbb{R}_+ -\{0\} $,
$ dDTdt≤−κ2min(N1,⋯,Nα)ζ(DX)N(T∞M)2DT+2κ2(n−1)ζM(1T∞m−1T∞M)≤−ˉΛ0DT+2κ2(n−1)(1T∞m−1T∞M)ζ((min1≤α≤n−1d(Iα,Iα+1))t+r02). $ |
We use Grönwall's lemma to yield that for $ t\in \mathbb{R}_+ $,
$ DT(t)≤DT(0)exp(−ˉΛ0t)+2κ2(n−1)(1T∞m−1T∞M)exp(−ˉΛ02t)ζ(r02)+2κ2(n−1)(1T∞m−1T∞M)ζ((min1≤α≤n−1d(Iα,Iα+1))t+r02). $ |
We conclude the desired second assertion.
As a direct consequence, we present the following result that the velocity and temperature of each agent in each cluster group converge to some same nonnegative value, respectively:
Corollary 4.1. Assume that $ Z_\alpha = \{(x_{\alpha i}, v_{\alpha i}, T_{\alpha i})\}_{i = 1}^{N_\alpha} $ is a solution to system (4.1). Then, under the sufficient frameworks $ (\mathcal{H}_0) $, $ (\mathcal{H}_1) $, $ (\mathcal{H}_2) $, and $ (\mathcal{H}_3) $, there exist some convergence values $ v_{\alpha}^\infty $ and $ T_{\alpha}^\infty $ for $ \alpha = 1, \cdots, n $ that satisfy that for $ t\in \mathbb{R}_+ $,
1. (Velocity convergence value for each cluster group)
$ ‖vαi(t)−v∞α‖=O(exp(−Λ02t)+ϕ((min1≤α≤n−1d(Iα,Iα+1))s+r02)+∫∞tϕ((min1≤α≤n−1d(Iα,Iα+1))s+r02)ds). $ |
2. (Temperature convergence value for each cluster group)
$ |Tαi(t)−T∞α|=O(exp(−ˉΛ02t)+ζ((min1≤α≤n−1d(Iα,Iα+1))s+r02)+∫∞tζ((min1≤α≤n−1d(Iα,Iα+1))s+r02)ds). $ |
Proof. Remember from Lemma 4.1 that
$ Nα˙vcenα=κ1NNα∑i=1Nα∑j=1ϕ(‖xαi−xαj‖)vαi‖vαj−vαi‖22Tαj+κ1NNα∑i=1Nβ∑j=1ϕ(‖xαi−xβj‖)(vβj−vαi+vαi‖vβk−vαi‖22)1Tβj. $ |
If we denote $ v_\alpha^\infty $ as
$ v∞α:=limt→∞vcenα(t)=vcenα(0)+κ1NNαNα∑i=1Nα∑j=1∫∞0ϕ(‖xαi−xαj‖)vαi‖vαj−vαi‖22Tαj+κ1NNαNα∑i=1Nβ∑j=1∫∞0ϕ(‖xαi−xβj‖)(vβj−vαi+vαi‖vβj−vαi‖22)1Tβj, $ |
then we have that
$ ‖vcenα(t)−v∞α‖≤κ1NNαNα∑i=1Nα∑j=1∫∞tϕ(‖xαi−xαj‖)vαi‖vαj−vαi‖22Tαj+κ1NNαNα∑i=1Nβ∑j=1∫∞tϕ(‖xαi−xβj‖)(vβj−vαi+vαi‖vβj−vαi‖22)1Tβj $ |
because
$ vcenα=vcenα(0)+κ1NNαNα∑i=1Nα∑j=1∫t0ϕ(‖xαi−xαj‖)vαi‖vαj−vαi‖22Tαjds+κ1NNαNα∑i=1Nβ∑j=1∫t0ϕ(‖xαi−xβj‖)(vβj−vαi+vαi‖vβj−vαi‖22)1Tβjds. $ |
Then, the multi-flocking estimate studied in Theorem 4.1 and the monotonicity and non-negativity of $ \phi $ imply that
$ \|v_{\alpha}^{cen}(t)-v^\infty_{\alpha}\| = \mathcal{O}\left(\exp\left(-\Lambda_0t\right)+\int_{t}^\infty \phi\left(\frac{\left(\min\nolimits_{1\leq \alpha\leq n-1}d(I_\alpha, I_{\alpha+1})\right)s+r_0}{2}\right) ds\right). $ |
Subsequently, we recall from Theorem 4.1 that
$ ‖vαi(t)−vcenα(t)‖=O(exp(−Λ02t)+ϕ((min1≤α≤n−1d(Iα,Iα+1))s+r02)). $ |
We combine the above estimates to derive that for $ \alpha = 1, \cdots, n $,
$ ‖vαi(t)−v∞α‖=O(exp(−Λ02t)+ϕ((min1≤α≤n−1d(Iα,Iα+1))s+r02)+∫∞tϕ((min1≤α≤n−1d(Iα,Iα+1))s+r02)ds). $ |
Similar to the above, there exists some positive value $ T_\alpha^\infty $ such that for $ \alpha = 1, \cdots, n $,
$ |Tαi(t)−T∞α|=O(exp(−ˉΛ02t)+ζ((min1≤α≤n−1d(Iα,Iα+1))s+r02)+∫∞tζ((min1≤α≤n−1d(Iα,Iα+1))s+r02)ds). $ |
We conclude the desired results.
In this paper, we have demonstrated various sufficient frameworks regarding the mono-cluster flocking, the non-emergence of mono-cluster flocking, and multi-cluster flocking of the TCSUS system. First, we presented the admissible data for the mono-cluster flocking of TCSUS to occur. From the result, we observed that the mono-cluster flocking occurs when the coupling strength is large enough, and then we were interested in how small the coupling strength must be to avoid mono-cluster flocking emerging. Second, we verified that if the coupling strength is smaller than some appropriate value in the TCSUS model with an integrable communication weight $ \phi $, then the mixed configuration gradually becomes separated after some time, and then each sub-ensemble simultaneously moves away linearly as the time increases. Hence, this showed the non-emergence of the mono-cluster flocking to the system. However, when $ \phi $ is non-integrable, we did not provide a suitable sufficient framework for the non-emergence of the mono-cluster flocking and we only gave a sufficient condition independent of the coupling strength for mono-cluster flocking to occur. Third, employing the spatial separation $ r_0 $ and velocity separations $ I_\alpha $'s, when the initial configuration is well separated given similar to multi-cluster, we proved that the multi-cluster flocking holds in the system with an integrable $ \phi $. The novelty of this paper is that we have extended the multi-cluster flocking of system (1.2) (see [29]) to a temperature field and generalize the bi-cluster flocking of system (1.5) (see [2]) to the multi-cluster flocking. We were unable to demonstrate a sufficient framework where the multi-cluster flocking emerges in a mixed initial configuration (not well separated) rather than from the multi-cluster flocking under the conditions such that the initial configuration is well separated could be an interesting research topic. This issue is left for future work.
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 Korea government (MSIT) (2022R1C12007321).
The authors declare there is no conflict of interest.
[1] | Frary CD, Johnson RK, Wang MQ. (2005) Food sources and intakes of caffeine in the diets of persons in the United States. J Am Diet Assoc 105(1):110-3. |
[2] |
Mitchell, D C, Knight, CA, Hockenberry J, Teplansky R, Hartman TJ. (2014) Beverage caffeine intakes in the U.S. Food Chem Toxicol 63: 136-42. doi: 10.1016/j.fct.2013.10.042
![]() |
[3] |
Heckman, M.A., Weil, J., Gonzalez de Mejia, E. (2010) Caffeine (1,3,7-trimethylxanthine) in foods: a comprehensive review on consumption, functionality, safety, and regulatory matters. J Food Sci 75: R77-R87. doi: 10.1111/j.1750-3841.2010.01561.x
![]() |
[4] | U.S. Food and Drug Administration. Letter from Jeanne Ireland, Assistant Commissioner for Legislation, FDA, to Senator Richard J. Durbin, August 10, 2012. Available from: http://www.durbin.senate.gov/public/index.cfm/files/serve?File_id=17eadaa1-85e7-4ceb-a827-be244fbddfa5. |
[5] | Bambha, K., Wilson, L., Unalp, A., Loomba, R.,Neuschwander-Tetri, B., Brunt, E. et al. (2014) Coffee consumption in NAFLD patients with lower insulin resistance is associated with lower risk of severe fibrosis. Liver Int 34: 1250-8. |
[6] |
Freedman ND, Park Y, Abnet CC, Hollenbeck AR, Sinha R. (2012) Association of coffee drinking with total and cause-specific mortality. N Engl J Med 366: 1891-904. doi: 10.1056/NEJMoa1112010
![]() |
[7] | Molloy JW, Calcagno CJ, Williams CD, Jones FJ, Torres DM, Harrison SA. (2012) Association of coffee and caffeine consumption with fatty liver disease, nonalcoholic steatohepatitis, and degree of hepatic fibrosis. Hepatology 55(2): 429-36. |
[8] |
Anty, R., Marjoux, S., Iannelli, A., Patouraux, S., Schneck, A., Bonnafous, S. et al. (2012) Regular coffee but not espresso drinking is protective against fibrosis in a cohort mainly composed of morbidly obese European women with NAFLD undergoing bariatric surgery. J Hepatol 57: 1090-6. doi: 10.1016/j.jhep.2012.07.014
![]() |
[9] | Birerdinc, A., Stepanova, M., Pawloski, L. and Younossi, Z. (2012) Caffeine is protective in patients with non-alcoholic fatty liver disease. Aliment Pharmacol Ther 35: 76-82. |
[10] | Catalano, D., Martines, G., Tonzuso, A., Pirri, C., Trovato, F. and Trovato, G. (2010) Protective role of coffee in non-alcoholic fatty liver disease (NAFLD). Dig Dis Sci 55: 3200-6. |
[11] | Gutierrez-Grobe, Y., Chavez-Tapia, N., Sanchez- Valle, V., Gavilanes-Espinar, J., Ponciano-Rodriguez, G., Uribe, M. et al. (2012) High coffee intake is associated with lower grade nonalcoholic fatty liver disease: the role of peripheral antioxidant activity. Ann Hepatol 11: 350-5. |
[12] | Wilson, K., Kasperzyk, J., Rider, J., Kenfield, S.,van Dam, R., Stampfer, M. et al. (2011) Coffee consumption and prostate cancer risk and progression in the Health Professionals Follow-up Study. J Natl Cancer Inst 103: 876-84. |
[13] | Liu, R., Guo, X., Park, Y., Huang, X., Sinha, R.,Freedman, N. et al. (2012) Caffeine intake, smoking, and risk of Parkinson disease in men and women. Am J Epidemiol 175: 1200-7. |
[14] | Nehlig A. (2016) Effects of coffee/caffeine on brain health and disease: What should I tell my patients? Pract Neurol 16(2): 89-95. |
[15] | Beghi, E., Pupillo, E., Messina, P., Giussani, G., Chio, A., et al. (2011) Coffee and amyotrophic lateral sclerosis: a possible preventive role. Am J Epidemiol 174: 1002-8. |
[16] |
Hu, G., Bidel, S., Jousilahti, P., Antikainen, R., Tuomilehto, J. (2007) Coffee and tea consumption and the risk of Parkinson’s disease. Mov Disord 22: 2242-8. doi: 10.1002/mds.21706
![]() |
[17] |
Vila-Luna, S., Cabrera-Isidoro, S., Vila-Luna, L., Juarez-Diaz, I., Bata- Garcia, J.L., et al. (2012) Chronic caffeine consumption prevents cognitive decline from young to middle age in rats, and is associated with increased length, branching, and spine density of basal dendrites in CA1 hippocampal neurons. Neuroscience 202: 384-95. doi: 10.1016/j.neuroscience.2011.11.053
![]() |
[18] | Popat, RA., Van Den Eeden, SK., Tanner, CM., Kamel, F., Umbach, DM., Marder, K., et al. (2011) Coffee, ADORA2A, and CYP1A2: the caffeine connection in Parkinson's disease. Eur J Neurol 18: 756-65. |
[19] |
Fisone, G., Borgkvist, A., Usiello, A. (2004) Caffeine as a psychomotor stimulant: mechanism of action. Cell Mol Life Sci 61: 857-72. doi: 10.1007/s00018-003-3269-3
![]() |
[20] |
Lee, K.W., Chen, W., Junn, E., Im, J.Y., Grosso, H., et al. (2011) Enhanced phosphatase activity attenuates α-synucleinopathy in a mouse model. J Neurosci 31: 6963-71. doi: 10.1523/JNEUROSCI.6513-10.2011
![]() |
[21] | Basurto-Islasa G, Blancharda, J, Tunga, YC. Fernandezb, JR Voronkov M, Stock M, Zhang S, Stock, JB, Iqbala K. (2014) Therapeutic benefits of a component of coffee in a rat model of Alzheimer disease. Neurobiol Aging 35(12): 2701-12. |
[22] | Teraoka, M., Nakaso, K., Kusumoto, C., Katano, S., Tajima, N., et al. (2012) Cytoprotective effect of chlorogenic acid against α-synuclein Related toxicity in catecholaminergic PC12 cells. J Clin Biochem Nutr 51(2): 122-7. |
[23] |
Ladewig, T., Kloppenburg, P., Lalley, P.M., et al. (2003) Spatial profiles of store-dependent calcium release in motoneurones of the nucleus hypoglossus from newborn mouse. J Physiol 547: 775-87. doi: 10.1113/jphysiol.2002.033605
![]() |
[24] | Cao, C., Cirrito, J.R., Lin, X., Wang, L., Verges, D.K., et al. (2009) Caffeine suppresses amyloid-beta levels in plasma and brain of Alzheimer’s disease transgenic mice. J Alzheimers Dis 17: 681-97. |
[25] | Arendash GW, Cao C. (2010) Caffeine and coffee as therapeutics against Alzheimer's disease. J Alzheimers Dis 20 Suppl 1(1): 117-26. |
[26] | Chen, X., Lan, X., Roche, I., Liu, R., Geiger, J.D. (2008) Caffeine protects against MPTP-induced blood-brain barrier dysfunction in mouse striatum. J Neurochem 107: 1147-57. |
[27] |
Goncalves, N., Simoes, A.T., Cunha, R.A., de Almeida, L.P. (2013) Caffeine and adenosine A2A receptor inactivation decrease striatal neuropathology in a lentiviral-based model of Machado-Joseph disease. Ann Neurol 73: 655-66. doi: 10.1002/ana.23866
![]() |
[28] | Zhang, X., Wu, M., Lu, F., Luo, N., He, Z.P., Yang, H. (2014) Involvement of alpha7 nAChR Signaling Cascade in Epigallocatechin Gallate Suppression of beta-Amyloid-Induced Apoptotic Cortical Neuronal Insults. Mol Neurobiol 49(1): 66-77. |
[29] | Han, K., Jia, N., Li, J., Yang, L., Min, L.Q. (2013) Chronic caffeine treatment reverses memory impairment and the expression of brain BNDF and TrkB in the PS1/APP double transgenic mouse model of Alzheimer’s disease. Mol Med Rep 8: 737-40. |
[30] | Wostyn P, Van Dam D, Audenaert K, De Deyn PP. (2011) Increased Cerebrospinal Fluid Production as a Possible Mechanism Underlying Caffeine's Protective Effect against Alzheimer's Disease. Int J Alzheimers Dis 2011:617420. |
[31] |
Leite, M.R., Wilhelm, E.A., Jesse, C.R., Brandao, R., Nogueira, C.W. (2011) Protective effect of caffeine and a selective A2A receptor antagonist on impairment of memory and oxidative stress of aged rats. Exp Gerontol 46: 309-15. doi: 10.1016/j.exger.2010.11.034
![]() |
[32] |
Brothers, H.M., Marchalant, Y., Wenk, G.L. (2010) Caffeine attenuates lipopolysaccharide-induced neuroinflammation. Neurosci Lett 480: 97-100. doi: 10.1016/j.neulet.2010.06.013
![]() |
[33] |
Shin, H.J., Ryu, J.H., Kim, S.T., Zuo, Z., Do, S.H. (2013) Caffeine-induced inhibition of the activity of glutamate transporter type 3 expressed in Xenopus oocytes. Toxicol Lett 217: 143-8. doi: 10.1016/j.toxlet.2012.12.007
![]() |
[34] |
Gołembiowska, K., Dziubina, A. (2012) The effect of adenosine A2A receptor antagonists on hydroxyl radical, dopamine, and glutamate in the striatum of rats with altered function of VMAT2. Neurotox Res 22: 150-7. doi: 10.1007/s12640-012-9316-9
![]() |
[35] |
Leon, D., Albasanz, J.L., Ruiz, M.A., Iglesias, I., Martin, M. (2005) Effect of chronic gestational treatment with caffeine or theophylline on Group I metabotropic glutamate receptors in maternal and fetal brain. J Neurochem 94: 440-51. doi: 10.1111/j.1471-4159.2005.03211.x
![]() |
[36] |
Iglesias, I., Leon, D., Ruiz, M.A., Albasanz, J.L., Martin, M. (2006) Chronic intake of caffeine during gestation down regulates metabotropic glutamate receptors in maternal and fetal rat heart. Amino Acids 30: 257-66. doi: 10.1007/s00726-005-0293-7
![]() |
[37] |
Ravi, D., Muniyappa, H., Das, K.C. (2008) Caffeine inhibits UVmediated NF-kappaB activation in A2058 melanoma cells: an ATMPKC delta-p38 MAPK-dependent mechanism. Mol Cell Biochem 308: 193-200. doi: 10.1007/s11010-007-9628-x
![]() |
[38] |
Wang, L., Dai, W., Lu, L. (2005) Ultraviolet irradiation-induced K+ channel activity involving p53 activation in corneal epithelial cells. Oncogene 24: 3020-7. doi: 10.1038/sj.onc.1208547
![]() |
[39] | Miwa, S., Sugimoto, N., Yamamoto, N., Shirai, T., Nishida, H., et al. (2012) Caffeine induces apoptosis of osteosarcoma cells by inhibiting AKT/mTOR/S6K, NF-κB and MAPK pathways. Anticancer Res 32: 3643-9. |
[40] |
Zeitlin R, Patel S, Burgess S, Arendash GW, Echeverria V. (2011) Caffeine induces beneficial changes in PKA signaling and JNK and ERK activities in the striatum and cortex of Alzheimer's transgenic mice. Brain Res 1417: 127-36. doi: 10.1016/j.brainres.2011.08.036
![]() |
[41] |
Liu, W.H., Chang, L.S. (2010) Caffeine induces matrix metalloproteinase-2 (MMP-2) and MMP-9 down-regulation in human leukemia U937 cells via Ca2+/ROS-mediated suppression of ERK/c-fos pathway and activation of p38 MAPK/c-jun pathway. J Cell Physiol 224: 775-85. doi: 10.1002/jcp.22180
![]() |
[42] |
Kang, C.H., Jayasooriya, R.G., Dilshara, M.G., Choi, Y.H., Jeong, Y.K., et al. (2012) Caffeine suppresses lipopolysaccharide-stimulated BV2 microglial cells by suppressing Akt-mediated NF-κB activation and ERK phosphorylation. Food Chem Toxicol 50: 4270-6. doi: 10.1016/j.fct.2012.08.041
![]() |
[43] |
Lee, Kang-Woo., Im, Joo-Young., Woo, Jong-Mi., Grosso, H., Kim, Yoon-Seong., Cristovao, AC. (2013) Neuroprotective and Anti-inflammatory Properties of a Coffee Component in the MPTP Model of Parkinson’s Disease. Neurotherapeutics 10: 143-53. doi: 10.1007/s13311-012-0165-2
![]() |
[44] | Feve, A.P. (2012) Tyrosine Hydroxylase and Parkinsons Disease. CNS Neurol Disord Drug Targets 11: 450-5. |
[45] |
Zhu, Y., Zhang J., Zeng Y. (2012) Overview of Tyrosine Hydroxylase in Parkinson’s disease. CNS Neurol Disord Drug Targets 11: 350-8. doi: 10.2174/187152712800792901
![]() |
[46] |
Sonsalla, P.K., Wong, L.Y., Harris, S.L., Richardson, J.R., Khobahy, I., et al. (2012) Delayed caffeine treatment prevents nigral dopamine neuron loss in a progressive rat model of Parkinson’s disease. Exp Neurol 234: 482-7. doi: 10.1016/j.expneurol.2012.01.022
![]() |
[47] |
McKenzie, S., Marley, P.D. (2002) Caffeine stimulates Ca2+ entry through store-operated channels to activate tyrosine hydroxylase in bovine chromaffin cells. Eur J Neurosci 15: 1485-92. doi: 10.1046/j.1460-9568.2002.01990.x
![]() |
[48] |
Datta, U., Noailles, P.A., Rodriguez, M., Kraft, M., Zhang, Y., Angulo, J.A. (1996) Accumulation of tyrosine hydroxylase messenger RNA molecules in the rat mesencephalon by chronic caffeine treatment. Neurosci Lett 220: 77-80. doi: 10.1016/S0304-3940(96)13213-4
![]() |
[49] |
Kachroo, A., Irizarry, M.C., Schwarzschild, M.A. (2010) Caffeine protects against combined paraquat and maneb-induced dopaminergic neuron degeneration. Exp Neurol 223: 657-61. doi: 10.1016/j.expneurol.2010.02.007
![]() |
[50] | Costa, M.S., Botton, P.H., Mioranzza, S., Ardais, A.P., Moreira, J.D., et al. (2008a) Caffeine improves adult mice performance in the object recognition task and increases BDNF and TrkB independent on phospho-CREB immunocontent in the hippocampus. Neurochem Int 53: 89-94. |
[51] | Costa, M.S., Botton, P.H., Mioranzza, S., Souza, D.O., Porciuncula, L.O. (2008b) Caffeine prevents age-associated recognition memory decline and changes brain-derived neurotrophic factor and tyrosine kinase receptor (TrkB) content in mice. Neuroscience 153: 1071-8. |
[52] | Sallaberry, C., Nunes, F., Costa, M.S., Fioreze, G.T., Ardais, A.P., et al. (2013) Chronic caffeine prevents changes in inhibitory avoidance memory and hippocampal BDNF immunocontent in middle-aged rats. Neuropharmacology 64: 153-9. |
[53] |
Bairam, A., Kinkead, R., Lajeunesse, Y., Joseph, V. (2010) Neonatal caffeine treatment does not induce long-term consequences on TrkB receptors or BDNF expression in chemosensory organs of adult rats. Neurosci Lett 468: 292-6. doi: 10.1016/j.neulet.2009.11.015
![]() |
[54] |
Alzoubi KH, Srivareerat M, Aleisa Am, Alkadhi KA. (2013) chronic caffeine treatment prevents stress-induced LTP impairment: the critical role of Phosphorylated CaMKII and BDNF. J Mol Neurosci 49: 11-20. doi: 10.1007/s12031-012-9836-z
![]() |
[55] |
Moy, G.A., McNay, E.C. (2013) Caffeine prevents weight gain and cognitive impairment caused by a high-fat diet while elevating hippocampal BDNF. Physiol Behav 109: 69-74. doi: 10.1016/j.physbeh.2012.11.008
![]() |
[56] |
Krishnakumar, R., Kraus, W.L. (2010) The PARP side of the nucleus: molecular actions, physiological outcomes, and clinical targets. Mol Cell 39: 8-24. doi: 10.1016/j.molcel.2010.06.017
![]() |
[57] |
Geraets, L., Moonen, H.J., Wouters, E.F., Bast, A., Hageman, G.J. (2006) Caffeine metabolites are inhibitors of the nuclear enzyme poly(ADP-ribose)polymerase-1 at physiological concentrations.Biochem. Pharmacol 72: 902-10. doi: 10.1016/j.bcp.2006.06.023
![]() |
[58] | Cleaver, J.E., Banda, M.J., Troll, W., Borek, C. (1986) Some protease inhibitors are also inhibitors of poly(ADP-ribose) polymerase. Carcinogenesis 7: 323-5. |
[59] | Gu, J.W., Brady, A.L., Anand, V., Moore, M.C., Kelly, W.C., Adair, T.H. (1999) Adenosine upregulates VEGF expression in cultured myocardial vascular smooth muscle cells. Am J Physiol 277: H595-H602. |
[60] |
Merighi, S., Benini, A., Mirandola, P., Gessi, S., Varani, K., et al. (2006) Adenosine modulates vascular endothelial growth factor expression via hypoxia-inducible factor-1 in human glioblastoma cells. Biochem Pharmacol 72: 19-31. doi: 10.1016/j.bcp.2006.03.020
![]() |
[61] |
Grant, M.B., Tarnuzzer, R.W., Caballero, S., Ozeck, M.J., Davis, M.I., et al. (1999) Adenosine receptor activation induces vascular endothelial growth factor in human retinal endothelial cells. Circ Res 85: 699-706. doi: 10.1161/01.RES.85.8.699
![]() |
[62] |
Merighi, S., Benini, A., Mirandola, P., Gessi, S., Varani, K., et al. (2007) Caffeine inhibits adenosine-induced accumulation of hypoxiainducible factor-1alpha, vascular endothelial growth factor, and interleukin-8 expression in hypoxic human colon cancer cells. Mol Pharmacol 72: 395-406. doi: 10.1124/mol.106.032920
![]() |
[63] |
Mukhopadhyay, D., Akbarali, H. I. (1996) Depletion of (Ca2+) inhibits hypoxia-induced vascular permeability factor (vascular endothelial growth factor) gene expression. Biochem Biophys Res Commun 229: 733-8. doi: 10.1006/bbrc.1996.1873
![]() |
[64] |
Jafari, M., Rabbani, A. (2004) Studies on the mechanism of caffeine action in alveolar macrophages: caffeine elevates cyclic adenosine monophosphate level and prostaglandin synthesis. Metabolism 53: 687-92. doi: 10.1016/j.metabol.2003.08.004
![]() |
[65] |
Fiebich, B.L., Lieb, K., Hull, M., Aicher, B., van Ryn, J., et al. (2000) Effects of caffeine and paracetamol alone or in combination withacetylsalicylic acid on prostaglandin E2 synthesis in rat microglial cells. Neuropharmacology 39: 2205-13. doi: 10.1016/S0028-3908(00)00045-9
![]() |
[66] | Seevaratnam, R., Raha, S., Tarnopolsky, M.A., et al. (2009) Coffee increases antioxidant enzyme capacity in the brain of male G93A mice, an animal model of amyotrophic lateral sclerosis (ALS) (abstract). Exp Biol 23: 109.6. |
[67] | Zeidán-Chuliá F, Gelain DP, Kolling EA, Rybarczyk-Filho JL, Ambrosi P, et al. (2013) Major components of energy drinks (caffeine, taurine, and guarana) exert cytotoxic effects on human neuronal SH-SY5Y cells by decreasing reactive oxygen species production. Oxid Med Cell Longev 2013(6): 61-2. |
[68] |
Rathod, M.A., Patel, D., Das, A., Tipparaju, S.R., Shinde, S.S., Anderson, R.F. (2013) Inhibition of radical-induced DNA strand breaks by water-soluble constituents of coffee: phenolics and caffeine metabolites. Free Radic Res 47: 480-7. doi: 10.3109/10715762.2013.788167
![]() |
[69] | Silverberg, J.I., Patel, M., Brody, N., Jagdeo, J. (2012) Caffeine protects human skin fibroblasts from acute reactive oxygen species-induced necrosis. J Drugs Dermatol 11: 1342-6. |
[70] |
Mercer, J.R., Gray, K., Figg, N., Kumar, S., Bennett, M.R. (2012) The methyl xanthine caffeine inhibits DNA damage signaling and reactive species and reduces atherosclerosis in ApoE-/- mice. Arterioscler Thromb Vasc Biol 32: 2461-7. doi: 10.1161/ATVBAHA.112.251322
![]() |
[71] |
Kim, J., Lee, S., Shim, J., Kim, H.W., Kim, J., et al. (2012) Caffeinated coffee, decaffeinated coffee, and the phenolic phytochemical chlorogenic acid up-regulate NQO1 expression and prevent H2O2-induced apoptosis in primary cortical neurons. Neurochem Int 60: 466-74. doi: 10.1016/j.neuint.2012.02.004
![]() |
[72] | Shukitt-Hale B, Miller MG, Chu YF, Lyle BJ, Joseph JA. (2013) Coffee, but not caffeine, has positive effects on cognition and psychomotor behavior in aging. Age 35(6): 2183-92. |
[73] | Nawrot, P., Jordan, S., Eastwood, J., Rotstein, J., Hugenholtz, A., Feeley, M. (2003) Effects of caffeine on human health. Food Addit Contam 20: 1-30. |
[74] |
Mesas, A.E., Leon-Munoz, L.M., Rodriguez-Artalejo, F., Lopez-Garcia, E. (2011) The effect of coffee on blood pressure and cardiovascular disease in hypertensive individuals: a systematic review and meta-analysis. Am J Clin Nutr 94: 1113-26. doi: 10.3945/ajcn.111.016667
![]() |
[75] |
Brent, R.L., Christian, M.S., Diener, R.M. (2011) Evaluation of the reproductive and developmental risks of caffeine. Birth Defects Res B Dev Reprod Toxicol 92: 152-87. doi: 10.1002/bdrb.20288
![]() |
[76] |
Peck, J.D., Leviton, A., Cowan, L.D. (2010) A review of the epidemiologic evidence concerning the reproductive health effects of caffeine consumption: a 2000–2009 update. Food Chem Toxicol 48: 2549-76. doi: 10.1016/j.fct.2010.06.019
![]() |
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, 2024, 19, 1556-1801, 547, 10.3934/nhm.2024024 | |
2. | Seung-Yeal Ha, Guanghui Jin, Youwei Ma, Wook Yoon, Emergent behaviors of two thermodynamically consistent inertial spin models, 2024, 34, 1054-1500, 10.1063/5.0223494 | |
3. | Jeong Seok Han, Woojoo Shim, Hyunjin Ahn, Global-In-Time Discrete Approximation of the Cucker–Smale Model with a Unit Speed Constraint, 2025, 192, 1572-9613, 10.1007/s10955-025-03397-x |