A local sensitivity analysis for the kinetic Kuramoto equation with random inputs

  • Received: 01 May 2018 Revised: 01 January 2019
  • Primary: 35Q82, 35Q92, 37H99

  • We present a local sensivity analysis for the kinetic Kuramoto equation with random inputs in a large coupling regime. In our proposed random kinetic Kuramoto equation (in short, RKKE), the random inputs are encoded in the coupling strength. For the deterministic case, it is well known that the kinetic Kuramoto equation exhibits asymptotic phase concentration for well-prepared initial data in the large coupling regime. To see a response of the system to the random inputs, we provide propagation of regularity, local-in-time stability estimates for the variations of the random kinetic density function in random parameter space. For identical oscillators with the same natural frequencies, we introduce a Lyapunov functional measuring the phase concentration, and provide a local sensitivity analysis for the functional.

    Citation: Seung-Yeal Ha, Shi Jin, Jinwook Jung. A local sensitivity analysis for the kinetic Kuramoto equation with random inputs[J]. Networks and Heterogeneous Media, 2019, 14(2): 317-340. doi: 10.3934/nhm.2019013

    Related Papers:

    [1] Seung-Yeal Ha, Shi Jin, Jinwook Jung . A local sensitivity analysis for the kinetic Kuramoto equation with random inputs. Networks and Heterogeneous Media, 2019, 14(2): 317-340. doi: 10.3934/nhm.2019013
    [2] Seung-Yeal Ha, Jeongho Kim, Jinyeong Park, Xiongtao Zhang . Uniform stability and mean-field limit for the augmented Kuramoto model. Networks and Heterogeneous Media, 2018, 13(2): 297-322. doi: 10.3934/nhm.2018013
    [3] Young-Pil Choi, Seung-Yeal Ha, Seok-Bae Yun . Global existence and asymptotic behavior of measure valued solutions to the kinetic Kuramoto--Daido model with inertia. Networks and Heterogeneous Media, 2013, 8(4): 943-968. doi: 10.3934/nhm.2013.8.943
    [4] Seung-Yeal Ha, Jaeseung Lee, Zhuchun Li . Emergence of local synchronization in an ensemble of heterogeneous Kuramoto oscillators. Networks and Heterogeneous Media, 2017, 12(1): 1-24. doi: 10.3934/nhm.2017001
    [5] Seung-Yeal Ha, Yongduck Kim, Zhuchun Li . Asymptotic synchronous behavior of Kuramoto type models with frustrations. Networks and Heterogeneous Media, 2014, 9(1): 33-64. doi: 10.3934/nhm.2014.9.33
    [6] Seung-Yeal Ha, Se Eun Noh, Jinyeong Park . Practical synchronization of generalized Kuramoto systems with an intrinsic dynamics. Networks and Heterogeneous Media, 2015, 10(4): 787-807. doi: 10.3934/nhm.2015.10.787
    [7] Tingting Zhu . Synchronization of the generalized Kuramoto model with time delay and frustration. Networks and Heterogeneous Media, 2023, 18(4): 1772-1798. doi: 10.3934/nhm.2023077
    [8] Hirotada Honda . Global-in-time solution and stability of Kuramoto-Sakaguchi equation under non-local Coupling. Networks and Heterogeneous Media, 2017, 12(1): 25-57. doi: 10.3934/nhm.2017002
    [9] Xiaoxue Zhao, Zhuchun Li . Synchronization of a Kuramoto-like model for power grids with frustration. Networks and Heterogeneous Media, 2020, 15(3): 543-553. doi: 10.3934/nhm.2020030
    [10] Hirotada Honda . On Kuramoto-Sakaguchi-type Fokker-Planck equation with delay. Networks and Heterogeneous Media, 2024, 19(1): 1-23. doi: 10.3934/nhm.2024001
  • We present a local sensivity analysis for the kinetic Kuramoto equation with random inputs in a large coupling regime. In our proposed random kinetic Kuramoto equation (in short, RKKE), the random inputs are encoded in the coupling strength. For the deterministic case, it is well known that the kinetic Kuramoto equation exhibits asymptotic phase concentration for well-prepared initial data in the large coupling regime. To see a response of the system to the random inputs, we provide propagation of regularity, local-in-time stability estimates for the variations of the random kinetic density function in random parameter space. For identical oscillators with the same natural frequencies, we introduce a Lyapunov functional measuring the phase concentration, and provide a local sensitivity analysis for the functional.



    Collective behaviors of oscillatory complex systems are ubiquitous in our nature, e.g., flashing of fireflies, chorusing of crickets, synchronous firing of cardiac pacemaker and metabolic synchrony in yeast cell suspension [1,7,15,32,33] etc. Aforementioned collective patterns come down to synchronization phenomena. The jargon ``synchronization" represents the adjustment of rhythms in an ensemble of weakly coupled oscillators. Compared to long human history, a rigorous treatment for synchronization started only several decades ago in the pioneering works by Kuramoto and Winfree in [25,26,37]. They introduced simple, continuous dynamical systems for weakly coupled oscillators, and explained how collective behaviors in such simple models can emerge from initial configurations. Recently, emergent dynamics of coupled oscillators on networks has become an active, emerging research field in diverse disciplines such as biology, nonlinear dynamics, statistical physics and sociology. After Kuramoto and Winfree's seminal works, many phenomenological models have been used in the study of synchronization. Among them, we are mainly interested in the prototype model, namely the Kuramoto model. In order to fix the idea, let $ z $ be a random vector which is a measurable vector-valued function on a sample space $ \Omega $ in $ \mathbb R^d $ and we denote its probability density function (pdf) by $ \pi(z) $. For the simplicity of presentation, we assume that $ z $ is one-dimensional. Due to the randomness in system parameters such as the natural frequency and coupling strength, the phases of oscillators are random processes as well. More precisely, let $ \theta_i = \theta_i(t,z) $ be the phase of the $ i $-th Kuramoto oscillator whose dynamics is governed by the random Kuramoto model (in short, RKM):

    $ tθi(t,z)=νi+1NNj=1κij(z)sin(θj(t,z)θi(t,z)),t>0,
    $
    (1)

    where $ \nu_i $ is an intrinsic natural frequency of the $ i $-th oscillator whose pdf is $ g = g(\nu) $, and nonnegative random field $ \kappa_{ij} = \kappa_{ij}(z) $ measures the random coupling strength between $ i $ and $ j $-th oscillators. Throughout the paper, we assume that $ \kappa_{ij} $ is symmetric in $ i $ and $ j $:

    $ \kappa_{ij} = \kappa_{ji}, \quad 1 \leq i, j \leq N. $

    For the deterministic case where all randomness were quenched, i.e.,

    $ \nu_i = \mbox{constant}, \quad \kappa_{ij}(z) = \kappa_{ij}, $

    emergent dynamics of (1) has been extensively studied in [2,6,10,11,12,13,14,20,23,28,29,30,34,35,36] where the complete synchronization and stability conditions were proposed. The pathwise well-posedness of (1) can be done using the standard Cauchy-Lipschitz theory. In authors' recent work [18], a local sensitivity analysis for (1) has been addressed. In this paper, we are interested in the correpsonding mean-field equation which can effectively describe the dynamics of system (1) with $ N \gg 1 $. Let $ \mathbb T = \mathbb R/(2\pi \mathbb Z) $, and $ f = f(t,\theta, \nu, z) $ be an one-oscillator probability density function on the extended phase space $ \mathbb T \times \mathbb R \times \Omega $ at time $ t $. Then, the density function $ f $ satisfies the RKKE [1,27]:

    $ tf+θ(ω[f]f)=0,(θ,ν,z)T×R×Ω,t>0,ω[f](t,θ,ν,z)=νκ(z)T×Rsin(θθ)f(t,θ,ν,z)dνdθ.
    $
    (2)

    When the extrinsic randomness in $ \kappa(z) $ and initial data $ f^0(z) $ are quenched, well-posedness and dynamic features of the kinetic Kuramoto equation have been studied in [4,5,8,27].

    In this paper, we address a local sensitivity analysis for (2) to see the effect of random parameter in $ f = f(t, \theta, \nu, z) $ which is one of topics in uncertainty quantification (UQ). More precisely, we study the propagation of regularity of $ f $ in the random space, and provide concentration and stability estimates of $ z $-variations $ \{ \partial_z^{\alpha} f \} $. The UQ for mean-field flocking models were first addressed in [3,9] where particle based gPC methods were discussed. On the other hand, systematic local sensitivity analysis for the Cucker-smale and Kuramoto models have been addressed in authors' series of recent works [16,17,18]. Thus, the current work is a continuation of this systematic research on the local sensitivity analysis for flocking and synchronization models.

    The main results of this paper are two-fold. First, we present pathwise well-posedness and stability estimate of the RKKE by establishing a priori estimates (see Theorem 3.2, Theorem 3.3 and Theorem 3.4): for $ T \in (0, \infty),\; z \in \Omega,\; \; l \leq k $,

    $ sup0t<Tlzf(t,z)Wkl,θ,νC(z,T),sup0t<Tf(t)Hlπ(Lθ,ν)C(T)f0Hlπ(Lθ,ν),sup0t<Tkl=0lz(f˜f)(t,z)Wkl,C(z,T)kl=0lz(f0˜f0)(z)Wkl,,
    $

    where $ f $ and $ {\tilde f} $ are solution processes to (2) corresponding to initial data $ f^0 $ and $ {\tilde f}^0 $, respectively.

    Second, we consider identical oscillator with $ \varphi(\nu) = \delta_0 $. In this case, we can write $ f(t,\nu, \theta, z) = \rho(t,\theta,z)\varphi(\nu) $ and $ \rho $ satisfies

    $ tρ+θ(˜ω[ρ]ρ)=0,(θ,ν,z)T×R×Ω,t>0,˜ω[ρ](t,θ,z)=κ(z)2π0sin(θθ)ρ(t,θ,z)dθ.
    $

    Now, we introduce a Lyapunov functional $ {\mathcal L} $ measuring the phase concentration:

    $ L[ρ](t,z):=T|θθρ,c(t,z)|2ρ(t,θ,z)dθ,θρ,c(t,z):=Tθρ(t,θ,z)dθ.
    $
    (3)

    Note that the zero convergence of $ {\mathcal L}[\rho] $ as $ t \to \infty $ implies the asymptotic formation of phase concentration in probability sense. This can be seen easily from Chebyshev inequality as follows:

    $ L[ρ](t,z)ε2|θθρ,c(t,z)|>ερdθ=ε2P[θθρ,c(t,z)|>ε].
    $

    This implies

    $ \lim\limits_{t \to \infty} \mathbb{P}[ |\theta - \theta_{\rho,c}(t,z)| > \varepsilon] \leq \frac{1}{\varepsilon^2} \lim\limits_{t \to \infty} {\mathcal L}[\rho](t,z) = 0. $

    Under suitable conditions on initial data and system parameters, we will show that there exists random functions $ C(z) $ and $ \Lambda(z) $ such that

    $ {\mathcal L}[|\partial_z \rho|](t,z) \le C(z) e^{-\Lambda(z) t}, \quad t \geq 0. $

    (see Theorem 4.3 for details). However, higher-order sensitivity analysis for $ {\mathcal L}[|\partial^{\alpha}_z \rho|] (t,z) $ with $ \alpha \gg 1 $ might lead to the exponential growth (see Remark 4), a phenomenon not observed in earlier works in this direction.

    The rest of this paper is organized as follows. In Section 2, we briefly introduce the random kinetic Kuramoto equation and discuss its basic properties. In Section 3, we study a pathwise well-posedness of the RKKE by providing a priori estimates such as boundedness of $ H_\pi^l(L_{\theta,\nu}^\infty) $ in any finite time interval, and provide the stability estimates for the RKKE. In Section 4, we perform a local sensitivity analysis for a Lyapunov functional (3). Finally, Section 5 is devoted to a brief summary of our main results and some remaining issues to be explored in future. In Appendices A and B we provide the proof for Theorem 3.2 and Lemma 4.2, respectively.

    Gallery of Notation: Throughout the paper, we use the following notation:

    Let $ \pi: \Omega \to \mathbb R_+ \cup \{0\} $ be a nonnegative p.d.f. function, and let $ y = y(z) $ be a scalar-valued random function defined on $ \Omega $. Then, we define the expected value as

    $ \mathbb{E}[\varphi] : = \int_\Omega \varphi(z) \pi(z) dz, $

    and a weighted $ L^2 $-space:

    $ L_\pi^2(\Omega) : = \{ y:\; \Omega \to \mathbb{R}\; |\; \int_{\Omega} |y(z)|^2 \pi(z) dz < \infty \}, $

    with an inner product and norm:

    $ \langle y_1, y_2 \rangle_{L^2_\pi(\Omega)} : = \int_{\Omega} y_1(z) y_2(z) \pi(z) dz, \qquad \|y\|_{L^2_\pi(\Omega)} : = \Big( \int_{\Omega} |y(z)|^2 \pi(z) dz \Big)^{\frac{1}{2}}. $

    For $ k \in \mathbb Z_+ \cup \{0 \} $, set

    $ \|y\|_{H^l_\pi(\Omega)} : = \Big( \sum\limits_{\ell = 0}^{k} \|\partial_z^{\ell} y\|^2_{L^2_\pi(\Omega)} \Big)^{\frac{1}{2}}, \quad k \geq 1, \qquad \|y\|_{H_\pi^0(\Omega)} : = \|y\|_{L_\pi^2(\Omega)}. $

    Let $ h = h(\theta, \nu,z) $ be scalar-valued random function defined on the extended phase space $ \mathbb T \times \mathbb R \times \Omega $. For such $ h $, we define a Sobolev norm $ W^{k,\infty}_{\theta, \nu} $ and a mixed norm $ H_\pi^l(L_{\theta,\nu}^{\infty}) $ as follows.

    $ h(z)Wk,θ,ν:=0α+βkαθβνh(z)L(T×R),h2Hlπ(Lθ,ν):=|α|lαzh2L2π(Ω;L(T×R)).
    $

    Moreover, as long as there is no confusion, we suppress $ \pi $ and $ \Omega $ dependence in $ L^2_\pi(\Omega) $-norm and $ H^k_\pi(\Omega) $-norm:

    $ \|y\|_{L_z^2} : = \|y\|_{L^2_\pi(\Omega)}, \quad \|y\|_{H_z^k} : = \|y\|_{H^k_\pi(\Omega)}. $

    In this section, we briefly introduce the RKKE and study its basic properties. Let $ \theta_i = \theta_i(t, z) $ be a random phase process of the $ i $-th Kuramoto oscillator whose dynamics is governed by the following system of random ordinary differential equations (ODEs):

    $ tθi(t,z)=νi+1NNj=1κij(z)sin(θj(t,z)θi(t,z)),t>0,1iN.
    $
    (4)

    Here, the coupling matrix $ (\kappa_{ij}(z)) $ is assumed to be a symmetric random matrix. In literature [1,13,21] on the Kuramoto model, the randomness in natural frequencies is assumed to be time-independent and quenched so that $ \nu_i $ is a constant parameter. It will be interesting to see how random natural frequency $ \nu_i $ and random coupling strength $ \kappa_{ij}(z) $ interplay in the synchronization process of (4). This uncertain quantification (UQ) question for (4) was addressed in authors' recent work [18].

    Next, we consider a situation where the number of oscillators tend to infinity and the coupling strengths $ \kappa_{ij}(z) $ are uniform and identical:

    $ N,κij(z)=κ(z),1i,jN.
    $
    (5)

    For the derivation of the mean-field model associated with (1) and (5), we refer to [27,31] for details. It is more convenient to rewrite system (4) with (5) as a dynamical system on the extended phase space $ \mathbb T \times \mathbb R $:

    $ {tθi(t,z)=νi+κ(z)NNj=1sin(θj(t,z)θi(t,z)),t>0,tνi=0.
    $
    (6)

    Next, we return to the pathwise mean-field limit of (6) as $ N \to \infty $. The formal mean-field limit equation can be easily identified using the standard BBGKY hierarchy based on the formal weak limit of marginal distribution functions and molecular chaos assumption (we refer to [22,24] for a brief introduction of BBGKY hierarchy). Note that for a frozen $ z \in \Omega $, the vector field generated by system (6) is bounded and Lipschitz continuous, so it does satisfy Neunzert's framework in [31] based on particle-in-cell method and measure-valued solutions. In fact, this has been rigorously done in [27] in any finite-time interval for any initial data. Recently, for an augmented Kuramoto model uniform-in-time mean-field limit is also derived by combining uniform stability analysis and finite-in-time mean-field limit in [19]. Thus, for each $ z \in \Omega $, we can perform the same argument as the deterministic case to derive the kinetic equation with random inputs:

    $ tf+θ(ω[f]f)=0,(θ,ν)T×R,t>0,ω[f](t,θ,ν,z)=νκ(z)T×Rsin(θθ)f(t,θ,ν,z)dνdθ
    $
    (7)

    subject to initial data:

    $ f(0,θ,ν,z)=f0(θ,ν,z).
    $
    (8)

    Here the initial datum $ f^0 $ is assumed to satisfy the following constraints:

    $ f0(θ,ν,z)=f0(θ+2π,ν,z),(ν,z)R×Ω,Tf0(θ,ν,z)dθ=g(ν,z),T×Rf0(θ,ν,z)dνdθ=1.
    $
    (9)

    Lemma 2.1. Let $ f $ be a smooth solution to (7), (8) and (9) satisfying additional periodic boundary condition:

    $ f(t, 0, \nu, z) = f(t, 2\pi, \nu, z), \quad (\nu, z) \in \mathbb R \times \Omega,\; \; t > 0. $

    Then, for each $ z \in \Omega $ and $ t >0 $, we have

    $ (i)T×Rf(t,θ,ν,z)dνdθ=T×Rf0(θ,ν,z)dνdθ.(ii)T×Rνf(t,θ,ν,z)dνdθ=T×Rνf0(θ,ν,z)dνdθ.
    $

    Proof. We multiply $ 1 $ and $ \nu $ to $ (7)_1 $ and integrate the resulting relation over $ \mathbb T \times \mathbb R $ to obtain the desired estimates:

    $ 0=tT×Rfdνdθ+T×Rθ(ω[f]f)dνdθ=tT×Rfdνdθ,0=tT×Rνfdνdθ+T×Rνθ(ω[f]f)dνdθ=tT×Rνfdνdθ.
    $

    Remark 1. Note that $ f $ satisfies

    $ f(t, \theta, \nu, z) = f(t, \theta + 2\pi, \nu, z), \quad \int_{ \mathbb T} f(t, \theta, \nu, z)d \theta = g(\nu), \quad \int_{ \mathbb T \times \mathbb R} f d\nu d\theta = 1. $

    Since the $ \nu $-variable is not a dynamic variable as can be seen from $ (6)_2 $, thus the $ \nu $-support of $ f $ will not be changed along the random Kuramoto flow. For a later use, we present the above argument in the following lemma.

    Lemma 2.2. Let $ f $ be a $ \mathcal{C}^1 $-regular process to (7), and suppose that the initial process $ f^0(z) $ has a compact support in $ \nu $ for each $ z \in \Omega $: there exists a positive random function $ M : = M(z)>0 $ such that

    $ \sup\limits_{\theta \in \mathbb T} |f^0(\theta, \nu, z)| = 0, \quad for \ all\quad |\nu| \ge M(z), \quad for \ each\quad z \in \Omega. $

    Then for each $ z \in \Omega $ and $ t >0 $, we have

    $ \sup\limits_{\theta \in \mathbb T}|f(t,\theta, \nu, z)| = 0, \quad for \ all\quad |\nu| \ge M(z), \quad for \ each\quad z \in \Omega. $

    Before we leave this section, we state a Grönwall-type lemma to be used in later sections.

    Lemma 2.3. Let $ y: {\mathbb R}_+ \cup \{0 \} \to {\mathbb R}_+ \cup \{0 \} $ be a differentiable function satisfying

    $ yαy+Ceβt,t>0,y(0)=y0,
    $
    (10)

    where $ \alpha $, $ \beta $ and $ C $ are non-negative constants $ \alpha \neq \beta $. Then $ y $ satisfies

    $ y(t) \leq y^0 e^{-\alpha t} + \frac{C}{\alpha-\beta}(e^{-\beta t} - e^{-\alpha t}) $

    Proof. We multiply (10) by $ e^{\alpha t} $ and integrate it over $ (0,t] $ to give

    $ y(t)e^{\alpha t} \le y^0 + \frac{C}{\alpha-\beta}(e^{(\alpha-\beta)t} - 1). $

    This yields the desired estimate.

    In this section, we present a pathwise well-posedness of (2) and propagation of $ H_\pi^l(L_{\theta,\nu}^\infty) $-norm of $ f $. First, we study a priori bound of $ W^{k, \infty}_{\theta,\nu} $-norm.

    For each $ l \in \mathbb N \cup \{0 \} $, it is easy to check that the $ z $-variations $ \{ \partial_z^{l} f \} $ satisfy a hierarchical system:

    $ {tf+ω[f]θf=(θω[f])f,l=0,t(lzf)+ω[f]θ(lzf)=(θω[f])(lzf)lr=1(lr)θ[(rzω[f])(lrzf)]L.O.T.,l1.
    $
    (11)

    Note that the L.H.S. for $ f $ and its $ z $-variations $ \partial_z^l f $ have the same transport structure, while the R.H.S. for $ (11)_2 $ has a lower-order $ z $-variation terms. Thus, the characteristics for $ \partial_z^l f $ will be the same as that of $ f $ and hence, independent of $ l $. More precisely, for a given $ (\theta, \nu) \in \mathbb T \times \mathbb R $ and $ z \in \Omega $, we define a random forward characteristics

    $ (\theta(t, z), \nu(t,z)) : = (\theta(t;0,\theta,\nu, z), \nu(t;0,\theta,\nu,z)) $

    as a solution to (11):

    $ {θ(s,z)s=ω[f](θ(s,z),ν(s,z)),ν(s,z)s=0,s>0,(θ(0,z),ν(0,z))=(θ,ν).
    $
    (12)

    Now, we define the $ \nu $-support of the process $ \partial_z^l f $ and its diameter as follows:

    $ Vl(t,z):=¯{νR | supθT|lzf(θ,ν,t,z)|0},D(Vl)(t,z):=sup{|ν1ν2| | ν1, ν2Vl(t,z)}.
    $
    (13)

    For $ l = 0 $, we set

    $ \mathcal{V}^0(t,z) = : \mathcal{V}(t,z) \quad \mbox{and} \quad D(\mathcal{V}^0)(t,z) = : D(\mathcal{V})(t,z). $

    Note that the $ \nu $-support of the process $ \partial_z^l f $ is a subset of the $ \nu $-support of $ f $. Then, since the $ \nu $-support of $ f $ does not change along the dynamics of $ (11)_1 $ by Lemma 2.2, we have

    $ \mathcal{V}^l(t,z) \subseteq \mathcal{V}(t,z) = \mathcal{V}(z) \quad \mbox{and} \quad D(\mathcal{V}^l)(t,z) \le D(\mathcal{V})(z) \quad l \geq 0,\; \; t \geq 0. $

    In this subsection, we study the propagation of $ W^{k, \infty}_{\theta,\nu} $-regularity of $ z $-variations $ \partial_z^l f : = \partial_z^l f(t,z) $ whose dynamics are given by the hierarchical system (11). For notational simplicity, we set

    $ \| h \|_{L^{\infty}} : = \| h \|_{L_{\theta, \nu}^{\infty}}, \qquad \| h \|_{W^{k,\infty}} : = \| h \|_{W_{\theta, \nu}^{k,\infty}}, $

    and we denote a generic non-negative random function by $ C(z, T) $ which depends on $ T $ and $ z $ and it may differ from line to line.

    Proposition 1. For $ k \in \mathbb N $ and $ z \in \Omega $, let $ f^0 : = f^0(z) $ be the initial process satisfying

    $ \|f^0(z)\|_{W^{k,\infty}} < \infty, \quad D(\mathcal{V}(z)) < \infty, \quad for \ each\ \ \ \ \ z \in \Omega. $

    Then, for $ T \in (0, \infty) $, there exists a unique $ W^{k,\infty} $-regular solution process $ f: = f(t,z) $ such that

    $ supt[0,T)f(t,z)Wk,C(z,T)f0(z)Wk,,
    $
    (14)

    where $ C(z,T) $ is a nonnegative random function $ C(z, T) $.

    Proof. The existence and uniqueness of the solution can be found in [27]. So we only provide a priori estimate for the solution process $ f $. We apply $ \partial_\theta^\alpha \partial_\nu^\beta $ to $ (2)_1 $ with $ 0 \leq \alpha + \beta \leq k $ to get

    $ t(αθβνf)+α+1θβν(ω[f]f)=0.
    $
    (15)

    Next, we split our estimate into two parts

    $ \alpha + \beta = 0, \quad 1 \leq \alpha + \beta \leq k. $

    Let $ z \in \Omega $ be fixed.

    ● Case A $ (\alpha + \beta = 0) $: Note that

    $ tf+ω[f]θf=θω[f]f.
    $
    (16)

    We use the method of characteristics to obtain

    $ f(t,θ(t,z),ν(t,z),z)=f0(θ,ν,z)t0(θω[f]f)(s,θ(s,z),ν(s,z),z)ds.
    $
    (17)

    Since

    $ |\partial_\theta \omega[f](s, \theta(s,z), \nu(s,z), z)| \leq \kappa(z)\left( \int_{ \mathbb T \times \mathbb R} |\cos(\theta_* - \theta)| f(s, \theta_*,\nu_*,z) d\nu_* d\theta_* \right)\leq \kappa(z), $

    it follows from (17) that

    $ f(t,z)Lf0(z)L+κ(z)t0f(τ,z)Ldτ.
    $
    (18)

    ● Case B $ (1\leq \alpha + \beta \leq k) $: Next, we consider the cases:

    $ \beta = 0 \quad \mbox{and} \quad \beta \geq 1. $

    ● Case B-1 ($ \beta = 0 $ case): In this case, we use the same argument as in Case A to see that $ |\partial_\theta^\alpha f| $ satisfies

    $ t(αθf)+ω[f]θ(αθf)=α+1μ=1(α+1μ)(μθω[f])(α+1μθf)=:R1.
    $
    (19)

    On the other hand, since

    $ |\partial_\theta^\mu \omega[f]| \le \kappa(z) \int_{ \mathbb T \times \mathbb R} f(\theta_*,\nu_*,z)d\nu_*d\theta_* = \kappa(z), $

    The R.H.S. of (19) can be estimated as

    $ |R1|C(α,z)α+1μ=1α+1μθfLC(α,z)fWα,.
    $
    (20)

    Now, we integrate relation (19) along the characteristics and use (20) to get

    $ αθf(t,z)Lαθf0(z)L+C(α,z)t0f(τ,z)Wα,dτ.
    $
    (21)

    $ \diamond $ Case B-2 ($ \beta \ge 1 $ case): It follows from (15) that

    $ t(αθβνf)+ω[f]θ(αθβνf)+μ+λ0(α+1μ)(βλ)(μθλνω[f])(α+1μθβλνf)=0.
    $
    (22)

    Since

    $ \partial_\nu \omega[f] = 1, \quad \partial^{\lambda}_\nu \omega[f] = 0, \quad \lambda \geq 2, $

    relation (22) can be simplified as

    $ t(αθβνf)+ω[f]α+1θβνf=βα+1θβ1νfα+1μ=1(α+1μ)(μθω[f])(α+1μθβνf)=:R2.
    $
    (23)

    We integrate the above relation along the characteristics and use the estimate

    $ \Big| {\mathcal R}_2 \Big| \leq C(\alpha, \beta, z) \left( \|\partial_\theta^{\alpha+1}\partial_\nu^{\beta-1} f\|_{L^\infty} + \sum\limits_{\mu = 0}^{\alpha}\|\partial_\theta^\mu \partial_\nu^\beta f\|_{L^\infty}\right) \le C(\alpha, \beta, z)\|f\|_{W^{\alpha+\beta,\infty}} $

    to obtain

    $ αθβνf(t,z)Lαθβνf0(z)L+C(α,β,z)t0f(τ,z)Wα+β,dτ.
    $
    (24)

    Therefore, we obtain relations (21) and (24), sum the resulting relation over all $ 1 \le \alpha +\beta \le k $ and add (18) to obtain a Gronwall's inequality:

    $ \|f(t,z)\|_{W^{k,\infty}} \le \|f^0(z)\|_{W^{k,\infty}} + C(\alpha, \beta, z) \int_0^t \|f(\tau,z)\|_{W^{k,\infty}}d\tau. $

    Then Grönwall's lemma yields the desired estimate (14).

    Remark 2. Note that the Sobolev embedding theorem yields that for $ k >2 $, $ W_{\theta, \nu}^{k,\infty} $-solution is $ {\mathcal C}^1( \mathbb T \times \mathbb R) $ along the path. Thus, it is a classical solution to (2) along the path.

    Now, we provide a lemma regarding the estimates of the frequency $ \omega[f] $.

    Lemma 3.1. For $ l \leq k \in \mathbb N $ and $ T \in (0, \infty) $, suppose that the two solution processes $ \partial_z^l f: = \partial_z^l f(t,z) $ and $ \partial_z^l \tilde{f} : = \partial_z^l \tilde{f}(t,z) $ to (11) satisfy the following conditions: for each $ z \in \Omega $ and $ t \in [0, T) $,

    $ \sum\limits_{l = 0}^k \left(\|\partial_z^l f(t,z)\|_{L^{\infty}} + \| \partial_z^l \tilde{f}(t,z) \|_{L^{\infty}}\right) < \infty, \quad D(\mathcal{V})(z) + D(\tilde{\mathcal{V}})(z) < \infty. $

    Then, for $ t \in [0, T) $ and $ z \in \Omega $ there exists a nonnegative random function $ C(z) $ such that

    $ (i)|αθkz(ω[f])(t,z)|C(z)kl=0lzf(t,z)L,(ii)|αθkz(ω[f]ω[˜f])(t,z)|C(z)kl=0lz(f˜f)(t,z)L,(iii)νω[f]=1,ανω[f]=0,α2.
    $

    Proof. We first recall the relation:

    $ \omega[f](t,\theta, \nu, z) = \nu - \kappa(z) \int_{ \mathbb T \times \mathbb R} \sin(\theta-\theta_*) f(t, \theta_*, \nu_*, z) d\nu_* d\theta_*. $

    (ⅰ) It follows from (2) that

    $ |αθkzω[f]|=|kl=0(kl)klzκ(z)T×Rαθ{sin(θθ)}lzf(θ,ν,z)dθdν|2πkl=0(kl)D(V)(z)|klzκ(z)|lzfLC(z)kl=0lzfL,
    $

    where we used $ \binom{k}{l} \leq 2^k $ and $ C(z) $ is given by

    $ C(z) : = 2^{k+1} \pi D(\mathcal{V})(z) \max\limits_{0 \leq m \leq k} |\partial_z^{m} \kappa(z)| . $

    (ⅱ) Similar to (ⅰ), we have

    $ |\partial_\theta^\alpha \partial_z^k (\omega[f] - \omega[\tilde{f}]) | \le C(z) \sum\limits_{l = 0}^k \|\partial_z^l (f-\tilde{f})\|_{L^\infty}, $

    where $ C(z) $ is given by

    $ C(z) : = 2^{k+1} \pi \max \{ D(\mathcal{V})(z), D(\mathcal{{\tilde V}})(z) \} \max\limits_{0 \leq m \leq k} |\partial_z^{m} \kappa(z)|. $

    (ⅲ) The third estimate follows from the defining relation of $ \omega[f] $ in (2).

    Now, we are ready to provide well-posedness of the process $ \partial_z^l f $ for every $ l \in \mathbb N $.

    Theorem 3.2. For $ k \in \mathbb N $ and $ T \in (0, \infty) $, suppose that the initial process $ f^0 $ satisfies the following conditions: for each $ z \in \Omega $ and $ l = 0, \ \cdots, k $

    $ \| \partial_z^l f^0(z) \|_{W^{k-l, \infty}} < \infty, \quad D(\mathcal{V})(z) < \infty. $

    Then, there exists a unique $ W^{k-l} $-regular process $ \partial_z^l f $ to (11) satisfying

    $ \sup\limits_{t \in [0,T)}\|\partial_z^l f(t,z)\|_{W^{k-l,\infty}} \le C(z,T), \quad for \ each\ z \in \Omega. $

    Proof. Since the proof is lengthy and tedious, we postpone its detailed proof in Appendix A. Here we briefly explain why one has a lower $ W^{k-l, \infty} $-regularity for higher $ z $-variation $ \partial_z^l f $. In Proposition 1, we have $ \|f(t,z) \|_{W^{k, \infty}} \leq C(z, T). $ Then, it follows from (11) that

    $ \partial_t (\partial_z f) + \omega[f] \partial_\theta (\partial_z f) = -(\partial_z \omega[f] )(\partial_\theta f) + \cdots. $

    Since R.H.S. of the above relation has a term $ (\partial_\theta f) $, the above relation yields $ W^{k-1, \infty} $ estimate for $ \partial_z f $. Similiarly, $ \partial_z^2 f $ satisfies

    $ \partial_t (\partial^2_z f) + \omega[f] \partial_\theta (\partial^2_z f) = -2 (\partial_z \omega[f]) \partial_\theta (\partial_z f) - (\partial_z^2 \omega[f]) (\partial_\theta f) + \cdots. $

    Hence, the term $ \partial_\theta (\partial_z^2 f) $ has $ W^{k-2, \infty} $-estimate. Thus, we can get at most $ W^{k-2, \infty} $ estimate for $ \partial^2_z f $. Inductively, we can get $ W^{k-l, \infty} $-estimate for $ \partial_z^l f $. This is why we have lower-order regularity for $ \partial_z^l f $.

    Next, we provide the boundedness of the solution process in $ H_\pi^l (L_{\theta,\nu}^\infty) $-norm.

    Theorem 3.3. For $ k \in \mathbb N $ and $ T \in (0, \infty) $, suppose that the initial process $ f^0 $ and coupling strength satisfy the following conditions:

    $ \sum\limits_{l = 0}^k \sup\limits_{z \in \Omega} \| \partial_z^l f^0(z) \|_{W^{k-l, \infty}} < \infty, \quad \sup\limits_{z \in \Omega} D(\mathcal{V})(z) < \infty, \quad \sum\limits_{l = 0}^k \sup\limits_{z \in \Omega}|\partial_z^l \kappa (z)| < \infty. $

    Then, for $ T \in (0,\infty) $ we have

    $ \|f(t)\|_{H_\pi^k(L_{\theta,\nu}^\infty)} \le C(T) \|f^0\|_{H_\pi^k(L_{\theta,\nu}^\infty)}, \quad t \in (0,T). $

    Proof. The proof is almost similar to that of Proposition 1. Thus, we briefly outline the proof here. By the same argument as in the proof of Proposition 1, we have

    $ \sum\limits_{l = 0}^k \|\partial_z^l f(t,z)\|_{L^\infty} \le C(T) \sum\limits_{l = 0}^k \left( \|\partial_z^l f^0(z)\|_{L^\infty} + \int_0^t \|\partial_z^l f (\tau,z)\|_{L^\infty}d \tau \right). $

    We use Grönwall's lemma to obtain

    $ kl=0lzf(t,z)LC(T)kl=0lzf0(z)L.
    $
    (25)

    Finally, we square both sides in (25), multiply by $ \pi(z) $ and integrate over $ \Omega $ to obtain the desired estimate:

    $ \|f(t)\|_{H_\pi^k(L_{\theta,\nu}^\infty)}^2 \le C(T) \|f^0\|_{H_\pi^k(L_{\theta,\nu}^\infty)}^2. $

    In this subsection, we provide a local-in-time $ W^{k,\infty} $-stability estimate for the RKKE. More precisely, we derive pathwise stability estimate of (2) with respect to initial data.

    Proposition 2. For $ k \in \mathbb N,\; z \in \Omega $ and $ T\in (0,\infty) $, let $ f $ and $ {\tilde f} $ be two $ W^{k+1,\infty} $-processes to (2) with the initial process $ f^0 $ and $ {\tilde f}^0 $ satisfying the following conditions: for each $ t \in (0,T) $ and $ z \in \Omega $,

    $ \Big(\|f^0(z)\|_{W^{k+1,\infty}} + \|{\tilde f}^0(z)\|_{W^{k+1,\infty}} \Big) < \infty, \qquad \left( D(\mathcal{V})(z) + D(\mathcal{{\tilde V}})(z) \right) < \infty. $

    Then, there exists a positive random function $ C(z,T) $ such that for each $ z \in \Omega $,

    $ \sup\limits_{0 \leq t < T} \|(f- \tilde{f})(t,z)\|_{W^{k,\infty}} \le C(z,T) \|(f^0 - \tilde{f}^0)(z)\|_{W^{k,\infty}}. $

    Proof. We use a similar argument as in Proposition 1 to derive the estimate

    $ αθβν(f˜f)Lαθβν(f0˜f0)L+C(z,T)t0(f˜f)(τ,z)Wα+β,dτ,
    $
    (26)

    where $ 0 \le \alpha+\beta \le k $. Finally, we sum the relation (26) over all $ 0 \le \alpha+\beta \le k $ to derive

    $ (f˜f)(t,z)Wk,(f0˜f0)(z)Wk,+C(z,T)t0(f˜f)(τ,z)Wk,dτ.
    $
    (27)

    Therefore, we use Grönwall's inequality on (27) to obtain the desired estimate.

    As an application of the arguments in Theorem 3.2 and Proposition 2, we get the local-in-time stability estimate of variations $ \partial^l_z f $ in $ W^{k-l, \infty} $-norm.

    Theorem 3.4. For $ k \in \mathbb N $ and $ T \in (0, \infty) $, suppose that two initial data $ f^0 $ and $ {\tilde f}^0 $ satisfy the following conditions: for each $ z \in \Omega $,

    $ \sum\limits_{l = 0}^k \left(\|\partial_z^l f^0(z)\|_{W^{k-l+1,\infty}} + \| \partial_z^l \tilde{f}^0(z) \|_{W^{k-l+1,\infty}}\right) < \infty, \quad D(\mathcal{V})(z) + D(\tilde{\mathcal{V}})(z) < \infty, $

    and let $ f: = f(t,z) $ and $ \tilde{f}: = \tilde{f}(t,z) $ be two $ W^{k+1,\infty} $-regular solution processes to (11) with initial data $ f^0(z) $ and $ \tilde{f}^{0}(z) $, respectively. Then, there exists a positive random function $ C(T, z) $ such that

    $ \sup\limits_{0 \leq t < T} \sum\limits_{l = 0}^k \|\partial_z^l(f-\tilde{f})(t,z)\|_{W^{k-l, \infty}} \le C(z,T) \sum\limits_{l = 0}^k \|\partial_z^l (f^0-\tilde{f}^0)(z)\|_{W^{k-l,\infty}}. $

    Proof. We basically follow the arguments in Theorem 3.2 and Proposition 2. Thus, we omit the details.

    In this section, we provide a local sensitivity analysis for the phase concentration that emerges in (2). Since the kinetic equation (2) has been derived from the first-order model, it is not easy to see how frequency synchronization emerges from (2). However, for the kinetic Kuramoto equation with $ g(\nu) = \delta_0(\nu) $, we can study the phase synchronization using a Lyapunov functional approach. First, we set:

    $ f(\theta,\nu,t,z) : = \rho(\theta,t,z) \delta_0(\nu). $

    We substitute this ansatz into (2) to obtain an equation for $ \rho $:

    $ tρ+θ(˜ω[ρ]ρ)=0,θT,t>0,˜ω[ρ](θ,t,z)=κ(z)2π0sin(θθ)ρ(θ,t,z)dθ,
    $
    (28)

    Recall a Lyapunov functional $ {\mathcal L} $ measuring the concentration of phases defined in (3):

    $ {\mathcal L}[\rho] : = \int_ \mathbb T |\theta - \theta_{\rho,c}|^2 \rho(\theta) d\theta, \qquad \theta_{\rho,c} : = \int_ \mathbb T \theta \rho(\theta) d\theta. $

    As discussed in Introduction, if $ {\mathcal L}[\rho] $ goes to zero, then $ \rho $ tends to $ \delta_{\theta_{\rho,c}} $ in probability. Next, we define several notation regarding the $ \theta $-support of $ \rho $:

    $ \label{E-3}(suppθρ)(t,z):=¯{θT | ρ(θ,t,z)0},Dθ(ρ)(t,z):=sup{|θθ| | θ,θsuppθρ(t,z)}.
    $

    If there is no confusion, we set $ \theta_c : = \theta_{\rho,c} $, where $ \rho $ is the solution process to (28). At this point, we would like to see the basic properties of the solution process (28).

    Proposition 3. Let $ \rho : = \rho(t,z) $ be a $ \mathcal{C}^1 $-regular process to (28). Then for each $ z \in \Omega $, we have

    $ (i)2π0ρ(t,θ,z)dθ=2π0ρ0(θ,z)dθ,2π0θρ(t,θ,z)dθ=2π0θρ0(θ,z)dθ.(ii)infθTρ(t,θ,z)0,ifinfθTρ0(θ,z)0.
    $

    Proof. (ⅰ) The conservation of total phase can be followed by the direct integration of (28) using the periodic boundary condition in $ \theta $-variable. For the second relation, we multiply $ \theta $ to $ (28)_1 $ to get

    $ \partial_t \int_0^{2\pi} \theta \rho d\theta = \int_0^{2\pi} \tilde{w}[\rho]\rho d\theta = \kappa(z)\int_{ \mathbb T^2} \sin(\theta_* - \theta) \rho(\theta_*,z)\rho(\theta,z)d\theta_* d\theta = 0, $

    where the last equality follows from the antisymmetry of the integrand.

    (ⅱ) For this, we consider the following characteristic:

    $ \frac{\partial}{\partial s} \tilde{\theta}(s;0,\theta,z) = \tilde{\omega}[\rho](\tilde{\theta}(s;0,\theta,z),s,z), \quad \tilde{\theta}(0;0,\theta,z) = \theta. $

    Now, we integrate (28) along the characteristic curve to yield

    $ \rho(\theta,t,z) = \rho^0(\tilde{\theta}(0;0,\theta,z),z) \exp \left(\int_0^t -\tilde{\omega}[\rho](\tilde{\theta}(s;0,\theta,z),s,z) ds\right). $

    From this, we can deduce the non-negativity of the process $ \rho $.

    Remark 3. Proposition 3 yields that $ \rho_c : = \int_ \mathbb T \rho d\theta $ and $ \theta_{\rho, c} $ are constants. Moreover, as long as the initial process is nonnegative, $ \rho $ is also nonnegative. Hence, without loss of generality, we may assume

    $ ρc(z)1,θρ,c(z)0,infθTρ0(θ,z)0,for anyzΩ.
    $
    (29)

    Under the above setting, we have

    $ \partial_z^l {\mathcal L}[\rho] = \partial_z^l \int_ \mathbb T |\theta|^2 \rho(\theta,z) d\theta = {\mathcal L}[\partial_z^l \rho(z)]. $

    Thus, for the local sensitivity analysis of $ {\mathcal L}[\rho] $, we just need to consider $ {\mathcal L}[\partial_z^l \rho] $, which is also bounded by $ {\mathcal L}[|\partial_z^l \rho|]. $

    Now we provide the contraction property of the $ \theta $-support of $ \rho $ in the following lemma.

    Lemma 4.1. Suppose that the $ \theta $-support of the initial process $ \rho^0 : = \rho^0(z) $ satisfies the following compactness condition:

    $ 0 < D_\theta(\rho^0)(z) < \pi, \quad for \ each\quad z \in \Omega. $

    Then, for $ \mathcal{C}^1 $-regular solution process $ \rho: = \rho(t,z) $ to (28), we have

    $ D_\theta(\rho)(t,z) \le D_\theta(\rho^0)(z), \quad for \ each\quad z \in \Omega. $

    Proof. Consider a forward characteristic $ \tilde{\theta} : = \tilde{\theta}(s;t,\theta,z) $ which is defined as a solution to the following equation:

    $ \frac{\partial}{\partial s} \tilde{\theta}(s;t,\theta,z) = \tilde{\omega}[\rho](\tilde{\theta}(s;t,\theta,z),t,z), \quad \tilde{\theta}(t;t,\theta,z) = \theta. $

    First we consider characteristic curve starting from the maximal point $ \theta_M(t,z) : = \sup\{ \theta \ | \ \theta \in \mbox{supp}_\theta \rho(t,z) \} $. Then, it is easy to see that it is nonincreasing:

    $ s˜θ(s;t,θM,z)|s=t+=κ(z)Tsin(θθM)ρ(θ,z)dθ0.
    $
    (30)

    Similarly, the characteristics curve starting from the minimal point $ \theta_m(t,z) : = \inf\{ \theta \ | \ \theta \in \mbox{supp}_\theta \rho(t,z) \} $ is nondecreasing:

    $ s˜θ(s;t,θm,z)|s=t+0.
    $
    (31)

    Thus, we can deduce from (30) and (31) that $ D_\theta (\rho)(t,z) $ does not increase at every time $ t $. This implies our desired result.

    Proposition 4. Let $ \rho : = \rho(t,z) $ be a $ \mathcal{C}^1 $-regular solution process satisfying the following condition:

    $ 0 < D_\theta(\rho^0)(z) < \pi, \quad for \ each\quad z \in \Omega $

    Then, $ {\mathcal L}[\rho](t,z) $ decays exponentially fast along the sample path:

    $ {\mathcal L}[\rho](t,z) \le {\mathcal L}[\rho^0](z)e^{-2 \kappa(z)R_0(z) t}, \quad t \geq 0, $

    where $ R_0(z) $ is defined by the following relation:

    $ R0(z):=sinDθ(ρ0)(z)Dθ(ρ0)(z).
    $
    (32)

    Proof. Under the setting (29), the functional $ {\mathcal L}[\rho] $ becomes

    $ {\mathcal L}[\rho] : = \int_ \mathbb T |\theta|^2 \rho(\theta) d\theta. $

    Then, it follows from (28) that

    $ tL[ρ](t,z)=2Tθ˜ω[ρ]ρdθ=2κ(z)T2θsin(θθ)ρ(θ,z)ρ(θ,z)dθdθ=κ(z)T2(θθ)sin(θθ)ρ(θ,z)ρ(θ,z)dθdθ,
    $

    where we used the change of variable $ \theta \leftrightarrow \theta_* $.

    On the other hand, by assumption and Lemma 4.1,

    $ |θθ|Dθ(ρ0)(z)<π,θ,θ(suppθρ)(t,z).
    $
    (33)

    Since $ \sin x \geq R_0(z) x, \quad x \in [0, D_\theta(\rho^0)], $ we have the following relation:

    $ xsinxR0(z)x2,x[Dθ(ρ0),Dθ(ρ0)],Dθ(ρ0)(0,π).
    $
    (34)

    We use (33) and (34) to yield

    $ tL[ρ](t,z)κ(z)R0(z)T2|θθ|2ρ(θ,z)ρ(θ,z)dθdθκ(z)R0(z)T2(θ2+θ2)ρ(θ,z)ρ(θ,z)dθdθ=2κ(z)R0(z)L[ρ](t,z),
    $
    (35)

    where we used

    $ |\theta - \theta_*|^2 = |\theta - \theta_c|^2 + |\theta_* - \theta_c|^2 +2(\theta-\theta_c)(\theta_c - \theta_*) $

    and Proposition 4.1 (ⅰ).

    Finally, we use Grönwall's lemma on (35) to obtain the desired result.

    As a direct corollary of Proposition 4, we have the following exponential decay of $ \mathbb{E}[{\mathcal L}[\rho]](t) $.

    Corollary 1. Suppose that the initial data and coupling strength satisfy the following conditions:

    $ 0 \leq \sup\limits_{z \in \Omega} D_\theta(\rho^0)(z) < \pi - \varepsilon \quad and \quad \inf\limits_{z \in \Omega} \kappa(z) \ge \eta > 0 $

    where $ \varepsilon $ is a positive constant. Then, for any $ \mathcal{C}^1 $-regular solution process $ \rho : = \rho(t,z) $ to (28), there exists a positive constant $ C = C(\varepsilon, \eta) $ such that

    $ \mathbb{E}[{\mathcal L}[\rho]](t) \le \mathbb{E}[{\mathcal L}[\rho^0]]e^{-2 C t}. $

    Before we consider the local sensitivity analysis about phase concentration, we first provide the following technical lemma.

    Lemma 4.2. For $ k >1 $, let $ \rho $ be a $ \mathcal{C}^k $-regular solution process to (28) satisfying

    $ D_\theta(\rho^0)(z) < \frac{\pi}{2}, \quad for \ each\quad z \in \Omega. $

    Then for $ \alpha < k $, we have the following upper and lower bound estimates:

    $ (i)tL[|αθρ|](t,z)2κ(z)R0(z)(L[|αθρ|](t,z)+αθρ(t,z)L1θL[ρ](t,z))+2κ(z)αθρL1θL[ρ]+ακL[|αθρ|](t,z)+κ(z)α+1μ=2(α+1μ)L[|α+1μθρ|](t,z),(ii)tL[|αθρ|](t,z)2κ(z)(L[|αθρ|](t,z)+αθρ(t,z)L1θL[ρ](t,z))2κ(z)αθρL1θL[ρ]+ακcosDθ(ρ0)(z)L[|αθρ|](t,z)κ(z)α+1μ=2(α+1μ)L[|α+1μθρ|](t,z),
    $

    where $ R_0(z) $ is defined in (32).

    Proof. Since the proof is rather lengthy, we leave its proof to Appendix B.

    Now, we are ready to provide the local sensitivity analysis for phase concentration. Note that for $ l \in \mathbb N $, $ \partial_z^l \rho $ satisfies

    $ t(lzρ)+˜ω[ρ]θ(lzρ)=(θ˜ω[ρ])lzρθ[lr=1(lr)rz˜ω[ρ]lrzρ].
    $
    (36)

    Note that the $ \theta $-support of $ \partial_z^l \rho $ is a subset of the $ \theta $-support of $ \rho $, which is contractive. Now, we would like to analyze the first-order derivative of the functional $ \Lambda[\rho] $ with respect to $ z $-variable.

    Theorem 4.3. Suppose that initial density $ \rho^0 $ satisfies

    $ 0 < D_\theta(\rho^0)(z) < \frac{\pi}{2}, \quad R_0(z) = \frac{\sin D_\theta (\rho^0)(z)}{D_\theta(\rho^0)(z)} > \frac{1}{2}, \quad for \ each\quad z \in \Omega, $

    and the coupling strength $ \kappa $ is continuously differentiable and bounded. Let $ \rho : = \rho(t,z) $ be a $ \mathcal{C}^2 $-regular global solution to (28) satisfying a priori condition:

    $ J[\rho](z) : = \sup\limits_{t \ge 0}(\|\partial_z \rho(z)\|_{L_\theta^1} + \|\partial_\theta \rho(z)\|_{L_\theta^1}) < \infty, \quad for \ each\quad z \in \Omega. $

    Then we have

    $ {\mathcal L}[|\partial_z \rho|](t,z) \le {\mathcal L}[|\partial_z \rho^0|](z)e^{-2 \kappa(z)R_0(z) t} + \frac{\mathcal{D}(z)}{\kappa(z)}e^{-\kappa(z)(2R_0(z) - 1)t}, $

    where the non-negative random function $ \mathcal{D} $ is given by

    $ D(z):=2κ(z)J[ρ]L[ρ0]+(|zκ(z)|+κ(z)J[ρ])(L[ρ0]+L[|θρ0|]+2J[ρ]L[ρ0]+L[ρ0]1R0(z)).
    $

    Proof. We differentiate (28) with respect to $ z $ to yield

    $ t(zρ)+θ(z˜ω[ρ]ρ+˜ω[ρ]zρ)=0.
    $
    (37)

    We multiply (37) by $ \mbox{sgn}(\partial_z \rho) $ to yield

    $ t|zρ|=θ(˜ω[ρ]|zρ|)θ(z˜ω[ρ]ρ)sgn(zρ).
    $
    (38)

    Then, one can use (38) to obtain

    $ tL[|zρ|]=Tθ2{θ(˜ω[ρ]|zρ|)+θ(z˜ω[ρ]ρ)sgn(zρ)}dθ=:I11+I12.
    $
    (39)

    Next, we estimate $ \mathcal{I}_{11} $ and $ \mathcal{I}_{12} $ as follows:

    ● Step A (Estimates for $ \mathcal{I}_{11} $): By direct calculation, one has

    $ I11=2Tθ˜ω[ρ]|zρ|dθ=2κT2θsin(θθ)ρ(θ)|zρ(θ)|dθdθ=κT2sin(θθ)(θρ(θ)|zρ(θ)|θρ(θ)|zρ(θ)|)dθdθ=κT2(θθ)sin(θθ)(ρ(θ)|zρ(θ)|+ρ(θ)|zρ(θ)|)dθdθ+κT2sin(θθ)(θρ(θ)|zρ(θ)|θρ(θ)|zρ(θ)|)dθdθ=:I111+I112.
    $
    (40)

    $ \diamond $ (Estimate of $ \mathcal{I}_{111} $): We use relation (34) and similar argument as in $ \mathcal{I}_{311} $ of Appendix B to yield

    $ I1112κR0(L[|zρ|]+zρL1θL[ρ]).
    $
    (41)

    $ \diamond $ (Estimate of $ \mathcal{I}_{112} $): Again similar to $ \mathcal{I}_{312} $ in Appendix B, we use Proposition 4 to find

    $ I1122κzρL1θL[ρ]2κzρL1θL[ρ0]eκR0t.
    $
    (42)

    In (40), we combine (41) and (42) to obtain

    $ I112κR0(L[|zρ|]+zρL1θL[ρ])+2κzρL1θL[ρ0]eκR0t.
    $
    (43)

    ● Step B (Estimates for $ \mathcal{I}_{12} $): Note that

    $ I12=Tθ2(θz˜ω[ρ]ρ+z˜ω[ρ]θρ)sgn(zρ) dθ=:I121+I122.
    $
    (44)

    $ \diamond $ (Estimate of $ \mathcal{I}_{121} $): In this case, one gets

    $ I121=T2θ2cos(θθ)(zκρ(θ)+κzρ(θ))ρ(θ)sgn(zρ(θ))dθdθ(|zκ|+κzρL1θ)L[ρ].
    $
    (45)

    $ \diamond $ (Estimate of $ \mathcal{I}_{122} $): Note that

    $ I122=T2θ2cos(θθ)(zκρ(θ)+κ(z)zρ(θ))θρ(θ)sgn(zρ(θ))dθdθ(|zκ|+κzρL1θ)L[|θρ|].
    $
    (46)

    It follows from Proposition 4 and Lemma 4.2 that

    $ tL[|θρ|](t,z)2κR0(L[|θρ|]+θρL1θL[ρ])+2κθρL1θL[ρ]+κL[|θρ|]+κL[ρ]κ(2R01)L[|θρ|]+κ(2θρL1θL[ρ]+L[ρ])κ(2R01)L[|θρ|]+κ(2J[ρ]L[ρ0]+L[ρ0])eκR0t,
    $
    (47)

    where $ J: = J[\rho](z) $ denotes the following random functional:

    $ J[\rho](z) : = \sup\limits_{t \ge 0}\left(\|\partial_z \rho\|_{L_\theta^1} + \|\partial_\theta \rho\|_{L_\theta^1} \right). $

    Thus we use the Grönwall-type inequality in Lemma 2.3 on (47) to yield

    $ {\mathcal L}[|\partial_\theta \rho|] \le \left( {\mathcal L}[|\partial_\theta \rho^0|] + \frac{2J[\rho]\sqrt{ {\mathcal L}[\rho^0]} + {\mathcal L}[\rho^0]}{1-R_0}\right)e^{-\kappa(2R_0-1)t}. $

    Therefore, we can obtain

    $ I12(|zκ|+κzρL1θ)L[ρ]+(|zκ|+κzρL1θ)L[|θρ|](|zκ|+κJ[ρ])L[ρ0]e2κR0t+(|zκ|+κJ[ρ])(L[|θρ0|]+2J[ρ]L[ρ0]+L[ρ0]1R0)eκ(2R01)t(|zκ|+κJ[ρ])(L[ρ0]+L[|θρ0|]+2J[ρ]L[ρ0]+L[ρ0]1R0)eκ(2R01)t.
    $
    (48)

    We combine (43) and (48) to yield

    $ tL[|zρ|]2κR0(L[|zρ|]+zρL1θL[ρ])+2κzρL1θL[ρ0]eκR0t+(|zκ|+κJ[ρ])(L[ρ0]+L[|θρ0|]+2J[ρ]L[ρ0]+L[ρ0]1R0)eκ(2R01)t2κR0L[|zρ|]+Deκ(2R01)t,
    $

    where the random function $ \mathcal{D} : = \mathcal{D}(z) $ was given by

    $ D(z):=2κJ[ρ]L[ρ0]+(|zκ|+κJ[ρ])(L[ρ0]+L[|θρ0|]+2J[ρ]L[ρ0]+L[ρ0]1R0).
    $

    Hence, we use the Grönwall type inequality to obtain

    $ {\mathcal L}[|\partial_z \rho|] \le {\mathcal L}[|\partial_z \rho^0|]e^{-2 \kappa R_0 t} + \frac{\mathcal{D}}{\kappa}e^{-\kappa (2R_0 - 1)t}. $

    This yields our desired result.

    Remark 4. In this remark, we discuss our results about local sensitivity analysis for the functional $ \mathcal{L}[\rho] $.

    $ 1. $ Note that in the proof of Theorem 4.3, we needed the temporal decay of $ {\mathcal L}[|\partial_\theta \rho|](t,z) $ to derive the temporal decay of $ {\mathcal L}[|\partial_z \rho|](t,z) $. Similarly, we needed the decay of $ {\mathcal L}[|\partial_\theta^2 \rho|](t,z) $ to get the decay of $ {\mathcal L}[|\partial^2_z \rho|](t,z) $. Recall that the estimate (i) in Lemma 4.2 yields that for $ \alpha = 2 $,

    $ \partial_t {\mathcal L}[|\partial_\theta^2 \rho|](t,z) \le 2\kappa(z) \left(1- R_0(z) \right) {\mathcal L}[|\partial_\theta^2 \rho|](t,z) + \mbox{L.O.T.} $

    Since the coefficient $ 2\kappa(z) \left(1- R_0(z) \right) > 0 $, the above differential inequality does not yield the time-decay of $ {\mathcal L}[|\partial_\theta^2 \rho|] $. Thus, it prohibits our local sensitivity analysis for the phase concentration for higher order derivatives.

    $ 2. $ On the other hand, it follows from Lemma 4.2 that

    $ \partial_t {\mathcal L}[|\partial_\theta^\alpha \rho|](t,z) \ge \kappa(z) \Big(\alpha \kappa \cos D_\theta (\rho^0)(z) - 2 \Big) {\mathcal L}[|\partial_\theta^\alpha \rho|](t,z) + \mbox{L.O.T.} $

    Thus, if $ \alpha $ is large enough to satisfy

    $ \alpha \cos D_\theta (\rho^0)(z) -2 > 0 \quad \mbox{for some} \ \ z \in \Omega, $

    then we can deduce the exponential growth of $ {\mathcal L}[|\partial_\theta^{\alpha'} \rho|(z)] $ for $ \alpha' \ge \alpha $ and such $ z \in \Omega $. Thus, it might lead the exponential growth for $ {\mathcal L}[|\partial_z^{\alpha} \rho|] $ for $ \alpha \gg 1 $.

    $ 3. $ We may provide some heuristic reason for the above observation; By Proposition 4, we have

    $ \rho(t,\theta, z) \to \delta_{\theta_c(z)}(\theta) \quad \mbox{in probability,} \quad \mbox{as} \quad t \to \infty. $

    Since $ \rho $ converges to a Dirac measure concentrated at a single phase $ \theta_c(z) $, the graph of $ \rho $ is expected to show a steep slope near $ \theta = \theta_c(z) $, which implies a rapid growth of $ |\partial_\theta^\alpha \rho| $ near $ \theta = \theta_c(z) $. However, since the $ \theta $-support of $ \rho $ also shrinks to the phase $ \theta_c(z) $, the term $ |\theta-\theta_c(z)|^2 $ in the integrand of $ \mathcal{L}[|\partial_\theta^\alpha \rho|] $ weakens the growth effect from $ |\partial_\theta^\alpha \rho| $. Thus, the growth and decay of $ \mathcal{L}[|\partial_\theta \rho|] $ will be determined by the dominance of the decay term $ |\theta-\theta_c(z)|^2 $ over the growth term $ |\partial_\theta \rho| $. Hence, the possibie exponential growth of $ \mathcal{L}[|\partial_\theta^\alpha \rho|] $ can be due to the dominance of the rapid growth of $ |\partial_\theta^\alpha \rho| $ over the decay from $ |\theta-\theta_c(z)|^2 $.

    In this paper, we provided a local sensitivity analysis for the kinetic Kuramoto equation with random inputs in a large coupling regime. In the absence of random inputs, it is well known that the kinetic Kuramoto model exhibits a phase concentration phenomena in the large coupling regime. In authors' earlier series of works, we have begun a systematic local sensitivity analysis for the Kuramoto model with random inputs. For the Kuramoto model, we provided a sufficient framework for the local sensitivity analysis on the asymptotic dynamics of solution process. In this work, we have not only shown the well-posedness of the uncertain problem and stability under random perturbation, but also conducted local sensitivity analysis regarding the phase concentration that could be observed in the Kuramoto model for identical oscillators. For this, we have considered a Lyapunov functional measuring the phase concentration and performed a local sensitivity analysis on the functional. In summary, we found two interesting effects due to uncertainties for (2):

    ● (Decreasing $ (\theta, \nu) $-regularity for higher-order $ z $-regularity estimate for $ f $): Propagation of high-order regularity in $ z $-variable is measured in low-order Sobolev norm in the sense that for $ T \in (0, \infty),\; z \in \Omega \subset \mathbb R,\; \; l \leq k $,

    $ \|\partial_z^l f^0(z)\|_{W^{k-l,\infty}_{\theta, \nu}} < \infty \quad \Longrightarrow \quad \sup\limits_{0 \leq t < T}\|\partial_z^l f(t,z)\|_{W^{k-l,\infty}_{\theta, \nu}} \le C(z,T). $

    ● (Formation of zero $ \theta $-variance and unbounded variance of $ \partial_\theta^\alpha \rho $): The $ \theta $-variances for $ \partial_\theta^\alpha \rho $ with $ |\alpha| \leq 1 $ tend to zero exponentially fast, whereas variances of higher order quantity $ \partial_\theta^{\alpha} \rho $ with $ \alpha \gg 1 $ grows exponentially fast.

    Of course, there are several interesting remaining issues on the role of random inputs. For example, in a small and critical coupling regimes, it is not known how the random inputs affects the overall collective dynamics or not. This will be addressed in future works.

    Since the local existence of regular solutions can be done using the standard argument based on contraction mapping theorem, we only provide a priori estimates to conclude the global-in-time existence of regular solutions in any finite-time interval. A priori estimates can be done inductively.

    Recall that $ \partial_z^l f $ satisfies

    $ tf+ω[f]θf+(θω[f])f=0,l=0,t(lzf)+ω[f]θ(lzf)+(θω[f])(lzf)=lr=1(lr)θ[(rzω[f])(lrzf)],l1.
    $
    (49)

    ● (Initial step): For $ l = 0 $, Proposition 1 yields

    $ \sup\limits_{t \in [0,T)}\|f(t,z)\|_{W^{k,\infty}} \le C(z, T) \|f^0(z)\|_{W^{k,\infty}}. $

    ● (Inductive step): Suppose that $ l \geq 1 $ and for $ 0 \le r < l $, there exists a unique $ W^{k-r,\infty} $-regular process $ \partial_z^r f $ to (11) for each $ z \in \Omega $ satisfying the following finiteness condition:

    $ supt[0,T)rzf(t,z)Wkr,C(z,T),0rl1.
    $
    (50)

    In order to get the desired estimate, we need to estimate $ \|\partial_\theta^{\alpha} \partial_\nu^{\beta} f \|_{L^{\infty}} $. For this, we consider the following two cases:

    $ (\alpha, \beta) = (0, 0), \qquad 1 \leq \alpha + \beta \leq k-l. $

    $ \diamond $ Case A ($ (\alpha, \beta) = (0, 0) $): It follows from (11) that

    $ t(lzf)+ω[f]θ(lzf)=(θω[f])(lzf)lr=1(lr)θrzω[f]lrzflr=1(lr)rzω[f]θlrzf=:R3.
    $
    (51)

    We use Lemma 3.1, (50), $ |\partial_\theta \omega[f] | \leq \kappa(z) $ and $ \binom{l}{r} \leq 2^l $ to estimate the R.H.S. of (51) as follows.

    $ |(θω[f])(lzf)|κ(z)lzfL,|lr=1(lr)θrzω[f]lrzf|2lC(z)lr=1(rp=0pzfL)lrzfLC(z,T),|lr=1(lr)rzω[f]θlrzf|2lC(z)C(z,T)lr=1(rp=0pzfL)C(z,T).
    $

    Thus, we have

    $ |R3|C(z,T)(lzfL+1).
    $
    (52)

    Next, we integrate the above relation (51) along the characteristics using the relation (52) to get

    $ lzf(t,z)Llzf0(z)L+C(z,T)(1+t0lzf(τ,z)Ldτ).
    $
    (53)

    $ \diamond $ Case B $ (1 \leq \alpha + \beta \leq k) $: For $ 1 \le \alpha +\beta \le k-l $, we differentiate (2) with respect to $ \theta $ and $ \nu $ for $ \alpha $- and $ \beta $-times respectively, to get

    $ t(αθβνlzf)+α+1μ=0βλ=0lr=0(α+1μ)(βλ)(lr)μθλνrzω[f]α+1μθβλνlrzf=0.
    $
    (54)

    $ \diamond $ Case B.1 $ (\beta \geq 1) $: In this case,

    $ t(αθβνlzf)+ω[f]θ(αθβνlzf)=βα+1θβ1νlzfα+1μ=1(α+1μ)μθω[f]α+1μθβνlzfα+1μ=0lr=1(α+1μ)(lr)μθrzω[f]α+1μθβνlrzf=:I21+I22+I23.
    $
    (55)

    Below, we separately estimate $ \mathcal{I}_{2n} $'s as follows.

    $ \diamond $ (Estimate of $ \mathcal{I}_{21} $): It is easy to see that

    $ |I21|βlzfWα+β,.
    $
    (56)

    $ \diamond $ (Estimate of $ \mathcal{I}_{22} $): We use $ |\partial_\theta^\mu \omega[f] | \leq \kappa(z) $ to get

    $ |I22|α+1μ=1(α+1μ)|μθω[f]α+1μθβνlzf|C(z)α+1μ=1α+1μθβνlzfLC(z)lzfWα+β,.
    $
    (57)

    $ \diamond $ (Estimate of $ \mathcal{I}_{23} $): We use Lemma 3.1 to obtain

    $ |I23|α+1μ=0lr=1(α+1μ)(lr)|μθrzω[f]α+1μθβνlrzf|2α+12lC(z)α+1μ=0lr=1(rp=0pzfL)α+1μθβνlrzfL.
    $
    (58)

    Now, we combine all estimates (56), (57), (58) and use the induction assumption to get

    $ |I21+I22+I23|C(z,T)(lzfWα+β,+α+1μ=0lr=1α+1μθβνlrzfL)C(z,T)(lzfWα+β,+1),
    $
    (59)

    Next, we integrate (55) along the characteristics and use (59) to yield

    $ αθβνlzf(t,z)Lαθβνlzf0(z)L+C(z,T)(1+t0lzf(τ,z)Wα+β,dτ).
    $
    (60)

    $ \diamond $ Case B.2 $ (\beta = 0) $: Similar to Case B.1, we get

    $ αθlzf(t,z)Lαθlzf0(z)L+C(z,T)(1+t0lzf(τ,z)Wα,dτ).
    $
    (61)

    Finally, we gather the results in (60) and (61), sum those over all $ 1 \le \alpha+\beta \le k-l $ and add the result in (53) to obtain

    $ \|\partial_z^l f(t,z)\|_{W^{k-l,\infty}} \le \|\partial_z^l f^0(z)\|_{W^{k-l,\infty}} + C(z,T)\left(1 + \int_0^t \|\partial_z^l f(\tau,z)\|_{W^{k-l,\infty}} d\tau \right). $

    Finally, we use Grönwall's lemma to derive the desired estimate.

    In this section, we present a proof of Lemma 4.2. It follows from (28) that

    $ t(αθρ)+α+1θ(˜ω[ρ]ρ)=0,
    $
    (62)

    or equivalently,

    $ t(αθρ)+θ(˜ω[ρ]αθρ)+αθ˜ω[ρ]αθρ+α+1μ=2(α+1μ)μθ˜ω[ρ]α+1μθρ=0.
    $
    (63)

    We multiply (63) by $ \mbox{sgn}(\partial_\theta^\alpha \rho) $ to yield

    $ tL[|αθρ|]=Tθ2t|αθρ|dθ=2Tθ˜ω[ρ]|αθρ|dθαTθ2θ˜ω[ρ]|αθρ|dθα+1μ=2(α+1μ)Tsgn(αθρ)θ2μθ˜ω[ρ]α+1μθρ dθ.=:I31+I32+I33
    $
    (64)

    (ⅰ) (An upper bound estimate): we separately estimate $ \mathcal{I}_{3n} $'s as follows:

    ● Case A (Estimates for $ \mathcal{I}_{31} $): In this case, one can yield

    $ I31=2κ(z)T2θsin(θθ)ρ(θ)|αθρ(θ)|dθdθ=κ(z)T2sin(θθ)(θρ(θ)|αθ(θ)|θρ(θ)|αθ(θ)|)dθdθ=κ(z)T2(θθ)sin(θθ)(ρ(θ)|αθρ(θ)|+ρ(θ)|αθρ(θ)|)dθdθ+κ(z)T2sin(θθ)(θρ(θ)|αθρ(θ)|θρ(θ)|αθρ(θ)|)dθdθ.=:I311+I312.
    $
    (65)

    $ \diamond $ (Estimate of $ \mathcal{I}_{311} $): Direct calculation leads to

    $ I311˜κ(z)T2|θθ|2(ρ(θ)|αθρ(θ)|+ρ(θ)|αθρ(θ)|dθdθ=~κ(z)T2(θ2+θ2)(ρ(θ)|αθρ(θ)|+ρ(θ)|αθρ(θ)|dθdθ=2˜κ(L[|αθρ|]+αθρL1θL[ρ]).
    $
    (66)

    $ \diamond $ (Estimate of $ \mathcal{I}_{312} $): Here, one obtains

    $ I312κ(z)T2|θ|ρ(θ)|αθρ(θ)|+|θ|ρ(θ)|αθρ(θ)|dθdθ=2κ(z)αθρL1θT|θ|ρ(θ)dθ2κ(z)αθρL1θL[ρ],
    $
    (67)

    where we used Cauchy-Schwarz inequality. In (65), we combine (66) and (67) to obtain

    $ I312˜κ(L[|αθρ|]+αθρL1θL[ρ])+2κ(z)αθρL1θL[ρ].
    $
    (68)

    ● Case B (Estimates for $ \mathcal{I}_{32} $): Note that

    $ I32=ακ(z)T2θ2cos(θθ)ρ(θ)|αθρ(θ)|dθdθακ(z)Tθ2|αθρ(θ)|dθ=ακ(z)L[|αθρ|].
    $
    (69)

    ● Case C (Estimates for $ \mathcal{I}_{33} $): One can get easily that

    $ I33α+1μ=2(α+1μ)Tθ2|α+1μθρ||μθ˜ω[ρ]|dθκ(z)α+1μ=2(α+1μ)L[|α+1μθρ|],
    $
    (70)

    where we used

    $ |\partial_\theta^\mu \tilde{w}[\rho]| \le \kappa(z)\int_{ \mathbb T} \rho(\theta_*)d\theta_* = \kappa(z). $

    Therefore, we combine all the estimates for $ \mathcal{I}_{3n} $'s to obtain

    $ tL[|αθρ|]2˜κ(L[|αθρ|]+αθρL1θL[ρ])+2κ(z)αθρL1θL[ρ]+ακL[|αθρ|]+κα+1μ=2(α+1μ)L[|α+1μθρ|],
    $

    which is our desired upper-bound estimate.

    (ⅱ) (A lower bound estimate): we will estimate separately $ \mathcal{I}_{3n} $'s similarly to the upper bound case. Thus we estimate first $ \mathcal{I}_{31k} $'s as follows:

    $ I311κ(z)T2|θθ|2(ρ(θ)|αθρ(θ)|+ρ(θ)|αθρ(θ)|)dθdθκ(z)(L[|αθρ|]+αθρL1θL[ρ]),I3122κ(z)αθρL1θL[ρ].
    $

    For $ \mathcal{I}_{32} $, one gets

    $ \mathcal{I}_{32} \ge \alpha \kappa(z) \cos D_\theta (\rho^0)(z) \int_ \mathbb T \theta^2 |\partial_\theta^\alpha \rho(\theta)|d\theta = \alpha \kappa(z) \cos D_\theta(\rho^0)(z) {\mathcal L}[|\partial_\theta^\alpha \rho|]. $

    For $ \mathcal{I}_{33} $, one can obtain

    $ \mathcal{I}_{33} \ge -\sum\limits_{\mu = 2}^{\alpha+1}\binom{\alpha+1}{\mu} \int_ \mathbb T \theta^2 |\partial_\theta^{\alpha+1-\mu}\rho| |\partial_\theta^\mu \tilde{w}[\rho]| d\theta \ge -\kappa(z) \sum\limits_{\mu = 2}^{\alpha+1} \binom{\alpha+1}{\mu} {\mathcal L}[|\partial_\theta^{\alpha+1-\mu} \rho|]. $

    We combine all estimates for $ \mathcal{I}_{3n} $'s to yield

    $ tL[|αθρ|]2κ(L[|αθρ|]+αθρL1θL[ρ])2κ(z)αθρL1θL[ρ]+ακcosDθ(ρ0)L[|αθρ|]κα+1μ=2(α+1μ)L[|α+1μθρ|],
    $

    which gives our desired lower-bound estimate.

    The work of S.-Y. Ha was supported by National Research Foundation of Korea(NRF-2017R1A2B2001864), and the work of S. Jin was supported by NSFC grant No. 31571071, NSF grants DMS-1522184 and DMS-1107291: RNMS KI-Net, and by the Office of the Vice Chancellor for Research and Graduate Education at the University of Wisconsin. The work of J. Jung is supported by the German Research Foundation (DFG) under the project number IRTG2235.



    [1] The Kuramoto model: A simple paradigm for synchronization phenomena. Rev. Mod. Phys (2005) 77: 137-185.
    [2] Existence of partial entrainment and stability of phase-locking behavior of coupled oscillators. Prog. Theor. Phys. (2004) 112: 921-941.
    [3] G. Albi, L. Pareschi and M. Zanella, Uncertain quantification in control problems for flocking models, Math. Probl. Eng., 2015 (2015), Art. ID 850124, 14pp. doi: 10.1155/2015/850124
    [4] Exponential dephasing of oscillators in the kinetic Kuramoto model. J. Stat. Phys. (2016) 162: 813-823.
    [5] On the complete phase synchronization for the Kuramoto model in the mean-field limit. Commun. Math. Sci. (2015) 13: 1775-1786.
    [6] J. Bronski, L. Deville and M. J. Park, Fully synchronous solutions and the synchronization phase transition for the finite-$N$ Kuramoto model, Chaos, 22 (2012), 033133, 17pp. doi: 10.1063/1.4745197
    [7] J. Buck and E. Buck, Biology of sychronous flashing of fireflies, Nature, 211 (1966), 562.
    [8] Contractivity of transport distances for the kinetic Kuramoto equation. J. Stat. Phys. (2014) 156: 395-415.
    [9] Particle based gPC methods for mean-field models of swarming with uncertainty. Comm. in Comp. Phys. (2019) 25: 508-531.
    [10] Asymptotic formation and orbital stability of phase-locked states for the Kuramoto model. Physica D (2012) 241: 735-754.
    [11] On exponential synchronization of Kuramoto oscillators. IEEE Trans. Automatic Control (2009) 54: 353-357.
    [12] Synchronization analysis of Kuramoto oscillators. Commun. Math. Sci. (2013) 11: 465-480.
    [13] Synchronization in complex networks of phase oscillators: A survey. Automatica (2014) 50: 1539-1564.
    [14] On the critical coupling for Kuramoto oscillators. SIAM. J. Appl. Dyn. Syst. (2011) 10: 1070-1099.
    [15] Synchronization in a pool of mutually coupled oscillators with random frequencies. J. Math. Biol (1985) 22: 1-9.
    [16] Local sensitivity analysis for the Cucker-Smale model with random inputs. Kinetic Relat. Models. (2018) 11: 859-889.
    [17] A local sensitivity analysis for the kinetic Cucker-Smale equation with random inputs. J. Differential Equations (2018) 265: 3618-3649.
    [18] S.-Y. Ha, S. Jin and J. Jung, Local sensitivity analysis for the Kuramoto mdoel with random inputs in a large coupling regime, Submitted.
    [19] Uniform stability and mean-field limit for the augmented Kuramoto model. Netw. Heterog. Media (2018) 13: 297-322.
    [20] Emergence of phase-locked states for the Kuramoto model in a large coupling regime. Commun. Math. Sci. (2016) 4: 1073-1091.
    [21] Collective synchronization of classical and quantum oscillators. EMS Surveys in Mathematical Sciences (2016) 3: 209-267.
    [22] From particle to kinetic and hydrodynamic description of flocking. Kinetic and Related Models (2008) 1: 415-435.
    [23] A. Jadbabaie, N. Motee and M. Barahona, On the stability of the Kuramoto model of coupled nonlinear oscillators, Proceedings of the American Control Conference, (2004), 4296-4301.
    [24] E. H. Kennard, Kinetic theory of gases. McGraw-Hill Book Company, New York and London, 1938.
    [25] Y. Kuramoto, Chemical Oscillations, Waves and Turbulence, Springer-Verlag, Berlin. 1984. doi: 10.1007/978-3-642-69689-3
    [26] Y. Kuramoto, International symposium on mathematical problems in mathematical physics, Lecture Notes in Theoretical Physics, 30 (1975), 420.
    [27] On the vlasov limit for systems of nonlinearly coupled oscillators without noise. Transport Theory and Statistical Physics (2005) 34: 523-535.
    [28] The spectrum of the partially locked state for the Kuramoto model. J. Nonlinear Science (2007) 17: 309-347.
    [29] The spectrum of the locked state for the Kuramoto model of coupled oscillators. Physica D (2005) 205: 249-266.
    [30] Stability of incoherence in a population of coupled oscillators. J. Stat. Phys. (1991) 63: 613-635.
    [31] H. Neunzert, An introduction to the nonlinear Boltzmann-Vlasov equation. In kinetic theories and the Boltzmann equation, Kinetic Theories and the Boltzmann Equation (Montecatini, 1981), 60-110, Lecture Notes in Math., 1048, Springer, Berlin, 1984. doi: 10.1007/BFb0071878
    [32] (2001) Synchronization: A Universal Concept in Nonlinear Sciences. Cambridge: Cambridge University Press.
    [33] From Kuramoto to Crawford: Exploring the onset of synchronization in populations of coupled oscillators. Physica D (2000) 143: 1-20.
    [34] A convergence result for the Kurmoto model with all-to-all couplings. SIAM J. Appl. Dyn. Syst. (2011) 10: 906-920.
    [35] On computing the critical coupling coefficient for the Kuramoto model on a complete bipartite graph. SIAM J. Appl. Dyn. Syst. (2009) 8: 417-453.
    [36] Global phase-locking in finite populations of phase-coupled oscillators. SIAM J. Appl. Dyn. Syst. (2008) 7: 134-160.
    [37] Biological rhythms and the behavior of populations of coupled oscillators. J. Theor. Biol. (1967) 16: 15-42.
  • This article has been cited by:

    1. Seung-Yeal Ha, Shi Jin, Jinwook Jung, Woojoo Shim, A local sensitivity analysis for the hydrodynamic Cucker-Smale model with random inputs, 2020, 268, 00220396, 636, 10.1016/j.jde.2019.08.031
    2. Ning Jiang, Zeng Zhang, Local Well-Posedness and Sensitivity Analysis for the Self-Organized Kinetic Model, 2021, 176, 0167-8019, 10.1007/s10440-021-00457-8
  • Reader Comments
  • © 2019 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(3536) PDF downloads(243) Cited by(2)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog