Review Recurring Topics

The Function of Sleep

  • The importance of sleep can be ascertained by noting the effects of its loss, which tends to be chronic and partial, on cognition, mood, alertness, and overall health. Many theories have been put forth to explain the function of sleep in humans, including proposals based on energy conservation, ecological adaptations, neurocognitive function, neural plasticity, nervous system and physical health, and performance. Most account for only a portion of sleep behavior and few are based on strong experimental support. In this review, we present theories proposing why sleep is necessary and supporting data demonstrating the effects of inadequate sleep, with the intention of gleaning further information as to its necessity, which remains one of the most perplexing mysteries in biology.

    Citation: Daniel A. Barone, Ana C. Krieger. The Function of Sleep[J]. AIMS Neuroscience, 2015, 2(2): 71-90. doi: 10.3934/Neuroscience.2015.2.71

    Related Papers:

    [1] Fei Cao, Sebastien Motsch, Alexander Reamy, Ryan Theisen . Asymptotic flocking for the three-zone model. Mathematical Biosciences and Engineering, 2020, 17(6): 7692-7707. doi: 10.3934/mbe.2020391
    [2] Hyunjin Ahn, Myeongju Kang . Emergent dynamics of various Cucker–Smale type models with a fractional derivative. Mathematical Biosciences and Engineering, 2023, 20(10): 17949-17985. doi: 10.3934/mbe.2023798
    [3] Le Li, Lihong Huang, Jianhong Wu . Flocking and invariance of velocity angles. Mathematical Biosciences and Engineering, 2016, 13(2): 369-380. doi: 10.3934/mbe.2015007
    [4] Huazong Zhang, Sumin Yang, Rundong Zhao, Qiming Liu . Finite-time flocking with collision-avoiding problem of a modified Cucker-Smale model. Mathematical Biosciences and Engineering, 2022, 19(10): 10332-10343. doi: 10.3934/mbe.2022483
    [5] Jingyi He, Changchun Bao, Le Li, Xianhui Zhang, Chuangxia Huang . Flocking dynamics and pattern motion for the Cucker-Smale system with distributed delays. Mathematical Biosciences and Engineering, 2023, 20(1): 1505-1518. doi: 10.3934/mbe.2023068
    [6] Jan Haskovec, Ioannis Markou . Exponential asymptotic flocking in the Cucker-Smale model with distributed reaction delays. Mathematical Biosciences and Engineering, 2020, 17(5): 5651-5671. doi: 10.3934/mbe.2020304
    [7] Nadia Loy, Andrea Tosin . A viral load-based model for epidemic spread on spatial networks. Mathematical Biosciences and Engineering, 2021, 18(5): 5635-5663. doi: 10.3934/mbe.2021285
    [8] Gheorghe Craciun, Matthew D. Johnston, Gábor Szederkényi, Elisa Tonello, János Tóth, Polly Y. Yu . Realizations of kinetic differential equations. Mathematical Biosciences and Engineering, 2020, 17(1): 862-892. doi: 10.3934/mbe.2020046
    [9] Katrine O. Bangsgaard, Morten Andersen, James G. Heaf, Johnny T. Ottesen . Bayesian parameter estimation for phosphate dynamics during hemodialysis. Mathematical Biosciences and Engineering, 2023, 20(3): 4455-4492. doi: 10.3934/mbe.2023207
    [10] Giulia Bertaglia, Walter Boscheri, Giacomo Dimarco, Lorenzo Pareschi . Spatial spread of COVID-19 outbreak in Italy using multiscale kinetic transport equations with uncertainty. Mathematical Biosciences and Engineering, 2021, 18(5): 7028-7059. doi: 10.3934/mbe.2021350
  • The importance of sleep can be ascertained by noting the effects of its loss, which tends to be chronic and partial, on cognition, mood, alertness, and overall health. Many theories have been put forth to explain the function of sleep in humans, including proposals based on energy conservation, ecological adaptations, neurocognitive function, neural plasticity, nervous system and physical health, and performance. Most account for only a portion of sleep behavior and few are based on strong experimental support. In this review, we present theories proposing why sleep is necessary and supporting data demonstrating the effects of inadequate sleep, with the intention of gleaning further information as to its necessity, which remains one of the most perplexing mysteries in biology.


    Flocking is observed in large populations of social agents such as birds [2], fish [3] or insects [4]. It refers to the emergence of large scale spatio-temporal structures which are not directly encoded in the individual agents' behavior. Understanding how large scale structures appear from individual behavior has sparked a huge literature in the recent years concerned with both modelling [5,6,7,8,9,10] and experiments [11,12]. We refer the reader to the reviews [13,14,15]. Yet this phenomenon is still poorly understood. The modelling of individual behavior at the microscopic scale requires to resolve the motion of each individual in the course of time. This leads to so-called Individual-Based Models (IBM) (or particle models) consisting of a huge number of coupled ordinary or stochastic differential equations. By contrast, given the very large number of agents involved, the modelling of the macroscopic scale is best done through Continuum Models (CM). They describe the system as a fluid through average quantities such as the density or mean velocity of the agents. At the mesoscopic scale, kinetic models (KM) are intermediates ketween IBM and CM. They describe the agents dynamics through the statistical distribution of their positions and velocities. In the literature, IBM, KM or CM are often chosen according to the authors' preferences rather than following an explicit rationale. However, understanding flocking requires understanding how the individuals' microscale impacts on the system's macroscale and consequently demands the use of a consistent sequence of IBM, KM and CM models. This consistency can only be guaranteed if the passage from IBM to KM and from KM to CM can be systematically established. But this issue is seldom considered. One goal of this paper is to establish the consistency between a KM and a CM of flocking based on a variant of the celebrated Cucker-Smale model of consensus (Eqs (1.1) and (1.2) below). KM have been introduced in the last years for the mesoscopic description of collective behavior of agents/particles with applications in collective behavior of cell and animal populations, see [1,16,17,18] and the references therein for a general overview on this active field. These models usually include alignment, attraction and repulsion as basic bricks of interactions between individuals.

    In this paper, we will discard the microscopic scale, i.e. the IBM. We will directly consider the mescoscale, i.e. the KM and will focus on the derivation of the CM from the KM. Of course, our KM has an underlying IBM. However the techniques involved in the passage from IBM to KM are quite different from those needed to pass from KM to CM. For this reason, considering the IBM would have brought us beyond the scope of the present paper. We refer to [19,20,21,22,23,24,25,26,27] and references therein for a derivation of KM from IBM. The derivation of the CM from the KM requires a spatio-temporal rescaling. Indeed, the KM still describes-although in a statistical way compared to the IBM---the microscopic dynamics of the particles. In particular, it is written in space and time units that are of the order of the particle interaction distances and times. To describe the macroscopic scale, one needs to introduce a change of variables by which the space and time units become of the order of the system scale, which is much wider that the particle-related scales. This rescaling introduces a small parameter, the ratio of the microscopic to the macroscopic space units. The derivation of the CM from the KM consists in finding the limit of the KM when this small parameter tends to zero. This is the goal of this paper. We will see that the CM (Eqs (1.5) and (1.6) below) corresponding to the considered modified Cucker-Smale model is a system consisting of equations for the density and mean orientation of the particles as functions of space and time. This system, referred to as the `Self-Organized Hydrodynamics (SOH)' bears analogies with the Euler equations of isothermal compressible gas dynamics, with the important difference that the average velocity is replaced by the average orientation, i.e. a vector of norm one. The SOH appears in a variety of contexts related to alignment interactions, such as repulsion [28], nematic alignment [29], suspensions [30], solid orientation [31], and can be seen as a basic CM of collective dynamics.

    Here, we focus on the derivation of macroscopic equations (SOH) for the collective motion of self-propelled particles with alignment and noise when a cruise speed for individuals is imposed asymptotically for large times as in [32,33,34,35,36,37,38]. More precisely, in the presence of friction and self-propulsion and the absence of other interactions, individuals/particles accelerate or break to achieve a cruise speed exponentially fast in time. The alignment between particles is imposed via localized versions of the Cucker-Smale or Motsch-Tadmor reorientation procedure [17,39,40,41,42,43] leading to relaxation terms to the mean velocity modulated or not by the density of particles. By scaling the relaxation time towards the asymptotic cruise speed, or equivalently, penalizing the balance between friction and self-propulsion, this alignment interaction leads asymptotically to variations of the classical kinetic Vicsek-Fokker-Planck equation with velocities on the sphere, see [1,10,44,45,46,47,48]. It was shown in [37,38] that particular versions of the localized kinetic Cucker-Smale model can lead to phase transitions driven by noise. Moreover, these phase transitions are numerically stable in this asymptotic limit converging towards the phase transitions of the limiting versions of the corresponding kinetic Vicsek-Fokker-Planck equation.

    In this work, we choose a localized and normalized version of the Cucker-Smale model not showing phase transition. More precisely, let us denote by $ f = f(t, x, v) \geq 0 $ the particle density in the phase space $ (x, v) \in \mathbb{R}^d \times \mathbb{R}^d $, with $ d \geq 2 $. The standard self-propulsion/friction mechanism leading to the cruise speed of the particles in the absence of alignment is given by the term $ {\rm div}_v \{f (\alpha - \beta |v|^2)v\} $ with $ \alpha, \beta > 0 $, and the relaxation toward the normalized mean velocity writes $ {\rm div}_v \{ f(v - \Omega[f])\} $. Here, for any particle density $ f(x, v) $, the notation $ \Omega[f] $ stands for the orientation of the mean velocity

    $ Ω[f]:={Rdf(,v)vdv|Rdf(,v)vdv|, if Rdf(,v)vdvRd{0},0, if Rdf(,v)vdv=0.
    $

    Notice that we always have

    $ \rho u[f] : = \int_{\mathbb{R} ^d} \!{f(\cdot, v)v} \;{\rm d}v = \left | \int_{\mathbb{R} ^d} \!{f(\cdot, v)v} \;{\rm d}v \right | \Omega [f] \qquad\mbox{with } \rho: = \int_{\mathbb{R} ^d} \!{f(\cdot, v)} \;{\rm d}v. $

    Let us remark that the standard localized Cucker-Smale model would lead to $ \rho {\rm div}_v \{ f(v -u[f])\} $ while the localized Motsch-Tadmor model would lead to $ {\rm div}_v \{ f(v -u[f])\} $. Our relaxation term towards the normalized local velocity $ \Omega [f] $ does not give rise to phase transition in the homogeneous setting on the limiting Vicsek-Fokker-Planck-type model on the sphere according to [45] and it produces a competition to the cruise speed term comprising a tendency towards unit speed. Including random Brownian fluctuations in the velocity variable leads to the kinetic Fokker-Planck type equation

    $ \partial_t f + v \cdot \nabla_x f + {\rm div}_v \{f ( \alpha - \beta |v|^2)v\} = {\rm div}_v \{ \sigma \nabla_v f + f ( v - \Omega[f]) \}, \;\;(t, x, v) \in \mathbb{R}_+ \times \mathbb{R}^d \times \mathbb{R}^d \, . $

    We include this equation in a more general family of equations written in a compact form as

    $ tf+vxf=Q(f),
    $
    (1.1)

    where

    $ Q(f)=divv{σvf+f(vΩ[f])+ηfvV},
    $
    (1.2)

    for any density distribution $ f $ with $ V $ a general confining potential in the velocity variables and $ \eta > 0 $ (see Lemma 2.1 for more information on the type of potentials that we consider). In the particular example considered above we take $ V = V_{\alpha, \beta} (|v|): = \beta \frac{|v|^4}{4} - \alpha \frac{|v|^2}{2}\; $.

    We investigate the large time and space scale regimes of the kinetic tranport Eq (1.1) with collision operator given by (1.2). Namely, we study the asymptotic behavior when $ \varepsilon \to 0 $ of

    $ tfε+vxfε=1εQ(fε),
    $
    (1.3)

    supplemented with the initial condition

    $ fε(0,x,v)=fin(x,v),(x,v)Rd×Rd.
    $

    The rescaling taken in the kinetic transport Eq (1.3) with confining potential $ V_{ \alpha, \beta} $ can be seen as an intermediate scaling between the ones proposed in [48] and [37]. The difference being that we have a relaxation towards the normalized mean velocity $ \Omega[f] $ rather than the mean velocity $ u[f] $ as in [37,48]. This difference is important since in the first case there is no phase transition in the homogeneous limiting setting on the sphere as we mentioned above, while in the second there is, see [38,45,48]. In fact, in [48] the scaling corresponds to $ \eta = 1 / \varepsilon $ in (1.3), that is the relaxation to the cruise speed is penalized with a term of the order of $ 1 / \varepsilon^2 $. Whereas in [37] the scaling correponds to $ \eta = \varepsilon $, that is the cruise speed is not penalized at all.

    The methodology followed in [48] lies within the context of measure solutions by introducing a projection operator onto the set of measures supported in the sphere whose radius is the critical speed $ r = \sqrt{\alpha/\beta} $. These technicalities are needed because the zeroth order expansion of $ f_\varepsilon $ lives on the sphere. This construction followed closely the average method in gyro-kinetic theory [49,50,51].

    However, in our present case we will show in contrast to [37,48] that there are no phase transitions which is in accordance with the results obtained in [46] for the kinetic Vicsek-Fokker-Planck equation with analogous alignment operator on the sphere. A modified version of (1.1) and (1.2) in which phase-transitions occur was studied in [38] whose analysis is postponed to a future work to focus here on the mathematical difficulties of the asymptotic analysis. Another difference in the present case is that the zeroth order expansion of $ f_\varepsilon $ will be parameterized by Von Mises-Fisher distributions in the whole velocity space, that is $ f(t, x, v) = \rho (t, x) M_{\Omega(t, x)}(v) $, with $ \rho $ and $ \Omega $ being, respectively, the density and the mean orientation of the particles. And where for any $ \Omega \in \mathbb{S}^{d-1} $ we define (see section 2)

    $ MΩ(v)=1ZΩexp(ΦΩ(v)σ), with ZΩ=Rdexp(ΦΩ(v)σ)dv
    $
    (1.4)

    and

    $ \Phi ( v) = \frac{ | v - \Omega|^2}{ 2} + V( | v|). $

    The main result of this paper is the asymptotic analysis of the singularly perturbed kinetic transport equation of Cucker-Smale type (1.3). The particle density $ \rho $ and the orientation $ \Omega $ obey the hydrodynamic type equations given in the following result.

    Theorem 1.1. Let $ f ^{{\rm in}} \geq 0 $ be a smooth initial particle density with nonvanishing orientation at any $ x \in \mathbb{R}^d $. For any $ \varepsilon > 0 $ we consider the problem

    $ \partial_t f ^\varepsilon + v \cdot \nabla_x f ^\varepsilon = \frac{1}{\varepsilon} {\rm div}_v \{ \sigma \nabla_v f ^\varepsilon + f ^\varepsilon ( v - \Omega[f ^\varepsilon] ) + f ^\varepsilon \nabla_v V(|v|)\}, \;\;(t, x, v) \in \mathbb{R}_+ \times \mathbb{R}^d \times \mathbb{R}^d \, , $

    with initial condition

    $ f ^\varepsilon (0, x, v) = f ^{{\rm in}} (x, v), \;\;(x, v) \in \mathbb{R}^d \times \mathbb{R}^d \, . $

    At any $ (t, x) \in \mathbb{R}_+ \times \mathbb{R}^d $ the leading order term in the Hilbert expansion $ f_\varepsilon = f + {f_1}_{\varepsilon} + \ldots $ is an equilibrium distribution of $ Q $, that is $ f(t, x, v) = \rho (t, x) M_{\Omega(t, x)}(v) $ with $ M_{\Omega(t, x)}(v) $ defined in (1.4), where the concentration $ \rho $ and the orientation $ \Omega $ satisfy

    $ tρ+divx(ρc1Ω)=0,(t,x)R+×Rd,
    $
    (1.5)
    $ tΩ+c2(Ωx)Ω+σ(IdΩΩ)xρρ=0,(t,x)R+×Rd,
    $
    (1.6)

    with initial conditions

    $ \rho (0, x) = \int_{\mathbb{R} ^d} \!{f ^{{\rm in}} (x, v)} \;{\rm d}v, \;\;\Omega (0, x) = \frac{\int_{\mathbb{R} ^d} \!{f ^{{\rm in}} (x, v)v} \;{\rm d}v}{\left | \int_{\mathbb{R} ^d} \!{f ^{{\rm in}} (x, v)v} \;{\rm d}v \right |}, \;\;x\in \mathbb{R}^d \, . $

    The constants $ c_1, c_2 $ are given by

    $ c1=R+rdπ0cosθe(cosθ,r)sind2θdθdrR+rd1π0e(cosθ,r)sind2θdθdr,
    $
    $ c2=R+rd+1π0cosθχ(cosθ,r)e(cosθ,r)sind1θdθdrR+rdπ0χ(cosθ,r)e(cosθ,r)sind1θdθdr,
    $

    and the function $ \chi $ solves

    $ σc[rd3(1c2)d12e(c,r)cχ]σr[rd1(1c2)d32e(c,r)rχ]+σ(d2)rd3(1c2)d52eχ=rd(1c2)d22e(c,r),
    $

    where $ e(c, r) = \exp(rc/\sigma) \exp(- (r^2 + 1)/(2\sigma) - V(r)/\sigma) $.

    Our article is organized as follows. First, in section 2 we state auxiliary results allowing us to discuss the kernel of the collision operator. Then in section 3 we concentrate on the characterization of the collision invariants. We prove that the generalized collision invariants introduced in [1] coincide with the kernel of a suitable linearised collision operator. We explicitly describe the collision invariants in section 4 and investigate their symmetries. Finally, the limit fluid model is determined in section 5 and we analyse its main properties.

    Plugging into (1.3) the Hilbert expansion

    $ f ^\varepsilon = f + \varepsilon f ^{1} + \ldots \, , $

    we obtain at the leading order

    $ Q(f)=0,
    $
    (2.1)

    whereas to the next order we get

    $ tf+vxf=limε01ε{Q(fε)Q(f)}=dQf(f1)=:Lf(f1),
    $
    (2.2)

    where $ \, \text{d} Q_f $ denotes the first variation of $ Q $ with respect to $ f $. The constraint (2.1) leads immediately to the equilibrium

    $ M_{\Omega} (v) = \frac{1}{Z_\Omega}\exp \left (- \frac{ \Phi_\Omega(v)}{\sigma} \right ) \, , \quad \text{ with } Z_\Omega = \int_{\mathbb{R} ^d} \!p \;{\rm d}v{\exp \left ( - \frac{ \Phi_\Omega(v^{\prime})}{\sigma} \right )} \, , $

    where

    $ ΦΩ(v)=|vΩ|22+V(|v|).
    $
    (2.3)

    Indeed, by using the identity

    $ vMΩ=MΩ(v)σvΦΩ=MΩ(v)σ(vΩ+vV(|v|)),
    $
    (2.4)

    we can recast the operator $ Q $ as

    $ Q(f)=divv(σvf+fvΦΩ[f])=σdivv[MΩfv(fMΩf)].
    $

    We denote by $ \mathbb{S}^{d-1} $ the set of unit vectors in $ \mathbb{R}^d $. For any $ \Omega \in \mathbb{S}^{d-1} $, we consider the weighted spaces

    $ L^2_{M_{\Omega}} = \bigg\{ \chi : \mathbb{R}^d \to \mathbb{R} \mbox{ measurable }, \;\int_{\mathbb{R} ^d} \!{(\chi (v))^2 M_{\Omega}(v)} \;{\rm d}v \lt \infty \bigg\} \, , $

    and

    $ H^1_{M_{\Omega}} = \bigg\{ \chi : \mathbb{R}^d \to \mathbb{R} \mbox{ measurable }, \;\int_{\mathbb{R} ^d} \!{[\;(\chi (v))^2 + |\nabla_v \chi |^2\;]M_{\Omega}(v)} \;{\rm d}v \lt \infty \bigg\} \, . $

    The nonlinear operator $ Q $ should be understood in the distributional sense, and is defined for any particle density $ f = f(v) $ in the domain

    $ D(Q)={f:RdR+ measurable ,f/MΩfH1MΩf}={f:RdR+ measurable ,Rd{(fMΩf)2+|v(fMΩf)|2}MΩf(v)dv<}.
    $

    We introduce the usual scalar products

    $ (\chi, \theta)_{M_{\Omega}} = \int_{\mathbb{R} ^d} \!{\chi (v) \theta (v) M_{\Omega} (v)} \;{\rm d}v, \;\;\chi, \theta \in L^2_{M_{\Omega}} \, , $
    $ ((\chi, \theta))_{M_{\Omega}} = \int_{\mathbb{R} ^d} \!{( \chi (v) \theta (v) + \nabla_v \chi \cdot \nabla_v \theta) M_{\Omega} (v)} \;{\rm d}v, \;\;\chi, \theta \in H^1_{M_{\Omega}} \, , $

    and we denote by $ |\cdot |_{M_{\Omega}}, \|\cdot \|_{M_{\Omega}} $ the associated norms. We make the following hypotheses on the potential $ V $. We assume that for any $ \Omega \in \mathbb{S}^{d-1} $ we have

    $ ZΩ=Rdexp(1σ[|vΩ|22+V(|v|)])dv<.
    $
    (2.5)

    Clearly (2.5) holds true for the potentials $ V_{\alpha, \beta} $. Notice that in that case $ 1 \in L^2_{M_{\Omega}} $ and $ |1|_{M_{\Omega}} = 1 $ for any $ \Omega \in \mathbb{S}^{d-1} $. Moreover, we need a Poincaré inequality, that is, for any $ \Omega \in \mathbb{S}^{d-1} $ there is $ \lambda_\Omega > 0 $ such that for all $ \chi \in H^1_{M_{\Omega}} $ we have

    $ σRd|vχ|2MΩ(v)dvλΩRd|χ(v)Rdχ(v)MΩ(v)dv|2MΩ(v)dv.
    $
    (2.6)

    A sufficient condition for (2.6) to hold comes from the well-known equivalence between the Fokker-Planck and Schrödinger operators (see for instance [52]). Namely, for any $ \Omega \in \mathbb{S}^{d-1} $ we have

    $ - \frac{\sigma}{\sqrt{M_{\Omega}}} {\rm div}_v\left ( M_{\Omega} \nabla_v \left ( \frac{u}{\sqrt{M_{\Omega}}}\right ) \right ) = - \sigma \Delta_v u + \left [ \frac{1}{4\sigma} |\nabla_v \Phi_\Omega |^2 - \frac{1}{2} \Delta_v \Phi_\Omega \right ] u \, . $

    The operator $ \mathcal{H}_\Omega = - \sigma \Delta_v + \left [\frac{1}{4\sigma} |\nabla_v \Phi_\Omega |^2 - \frac{1}{2} \Delta_v \Phi_\Omega \right] $ is defined in the domain

    $ D(\mathcal{H}_\Omega) = \left \{u \in L^2 (\mathbb{R}^d), \;\left [ \frac{1}{4\sigma} |\nabla_v \Phi_\Omega |^2 - \frac{1}{2} \Delta_v \Phi_\Omega \right ]u \in L^2 (\mathbb{R}^d) \right \} \, . $

    Using classical results for Schrödinger operators (see for instance Theorem ⅩⅢ.67 in [53]), we have a spectral decomposition of the operator $ \mathcal{H}_\Omega $ under suitable confining assumptions.

    Lemma 2.1. Assume that for $ \Phi_\Omega $ defined in (2.3) the function $ v \to \frac{1}{4\sigma} |\nabla_v \Phi_\Omega |^2 - \frac{1}{2}\Delta_v \Phi_\Omega $ satisfies the following:

    a) it belongs to $ L^1_{{\rm loc}}(\mathbb{R}^d) $,

    b) it is bounded from below,

    c)

    $ \lim\limits_{|v| \to \infty} \left [ \frac{1}{4\sigma} |\nabla_v \Phi_\Omega |^2 - \frac{1}{2}\Delta_v \Phi_\Omega \right ] = \infty \, . $

    Then $ \mathcal{H}_\Omega ^{-1} $ is a self adjoint compact operator in $ L^2(\mathbb{R}^d) $ and $ \mathcal{H}_\Omega $ admits a spectral decomposition, that is a nondecreasing sequence of real numbers $ (\lambda_\Omega ^n)_{n\in \mathbb{N}} $, $ \lim_{n \to \infty} \lambda_\Omega ^n = \infty $, and a $ L^2(\mathbb{R}^d) $-orthonormal basis $ (\psi_\Omega ^n)_{n \in \mathbb{N}} $ such that $ \mathcal{H}_\Omega \psi_\Omega ^n = \lambda_\Omega ^n \psi_\Omega ^n, n \in \mathbb{N} $, $ \lambda_\Omega ^0 = 0 $, $ \lambda_\Omega ^1 > 0 $.

    Let us note that the spectral gap of the Schrödinger operator $ \mathcal{H}_\Omega $ is the Poincaré constant in the Poincaré inequality (2.6). Notice also that the hypotheses in Lemma 2.1 are satisfied by the potentials $ V_{\alpha, \beta} $, and therefore (2.6) holds true in that case. It is easily seen that the set of equilibrium distributions of $ Q $ is parametrized by $ d $ parameters as stated in the following result.

    Lemma 2.2. Let $ f = f(v)\geq 0 $ be a function in $ D(Q) $. Then $ f $ is an equilibrium for $ Q $ if and only if there are $ (\rho, \Omega) \in \mathbb{R}_+ \times \left (\mathbb{S}^{d-1} \cup \{0\} \right) $ such that $ f = \rho M_{\Omega} $. Moreover we have $ \rho = \rho[f]: = \int_{\mathbb{R} ^d} \!{f(v)} \; {\rm d}v $ and $ \Omega = \Omega[f] $.

    Proof. If $ f $ is an equilibrium for $ Q $, we have

    $ \sigma \int_{\mathbb{R} ^d} \!{\left |\nabla_v \left ( \frac{f}{M_{\Omega}f} \right )\right | ^2 M_{\Omega}f (v)} \;{\rm d}v = 0 \, , $

    and therefore there is $ \rho \in \mathbb{R} $ such that $ f = \rho M_{\Omega}f $. Obviously $ \rho = \int_{\mathbb{R} ^d} \!{f(v)} \; {\rm d}v\geq 0 $ and $ \Omega [f] \in \mathbb{S}^{d-1} \cup \{0\} $. Conversely, we claim that for any $ (\rho, \Omega)\in \mathbb{R}_+ \times (\mathbb{S}^{d-1} \cup \{0\}) $, the particle density $ f = \rho M_{\Omega} $ is an equilibrium for $ Q $. Indeed, we have

    $ \sigma \nabla_v (\rho M_{\Omega} ) + \rho M_{\Omega} (v - \Omega + \nabla_v V ) = \rho ( \sigma \nabla_v M_{\Omega} + M_{\Omega} \nabla_v \Phi_\Omega ) = 0 \, . $

    We are done if we prove that $ \Omega [f] = \Omega $. If $ \Omega = 0 $, it is easily seen that

    $ \int_{\mathbb{R} ^d} \!{f(v)v} \;{\rm d}v = \rho\int_{\mathbb{R} ^d} \!{\frac{1}{Z_0} \exp \left ( - \frac{\Phi_0(v)}{\sigma}\right ) v} \;{\rm d}v = \frac{\rho}{Z_0} \int_{\mathbb{R} ^d} \!{\exp\left (- \frac{1}{\sigma}\left ( \frac{|v|^2}{2} + V(|v|)\right) \right )v} \;{\rm d}v = 0 \, , $

    implying $ \Omega[f] = 0 = \Omega $. Assume now that $ \Omega \in \mathbb{S}^{d-1} $. For any $ \xi \in \mathbb{S}^{d-1} $, $ \xi \cdot \Omega = 0 $, we consider the orthogonal transformation $ \mathcal{O}_\xi = I_d - 2\xi \otimes \xi $. Thanks to the change of variable $ v = \mathcal{O}_\xi v^{\prime} $, we write

    $ Rd(vξ)f(v)dv=ρZΩRd(vξ)MΩ(v)dv=ρZΩRdpdv(Oξvξ)MΩ(Oξv)=ρZΩRdpdv(vξ)MΩ(v)=Rdpdv(vξ)f(v),
    $

    where we have used the radial symmetry of $ V $, $ \mathcal{O}_\xi \xi = - \xi $ and $ \mathcal{O}_\xi \Omega = \Omega $. We deduce that $ \int_{\mathbb{R} ^d} \!{f(v)v} \; {\rm d}v = \int_{\mathbb{R} ^d} \!{(v\cdot \Omega) f(v)} \; {\rm d}v\; \Omega $. We claim that $ \int_{\mathbb{R} ^d} \!{(v\cdot \Omega) f(v)} \; {\rm d}v > 0 $. Indeed we have

    $ Rd(vΩ)f(v)dv=ρZΩvΩ>0(vΩ)MΩ(v)dv+ρZΩvΩ<0(vΩ)MΩ(v)dv=ρZΩvΩ>0(vΩ)[exp(ΦΩ(v)σ)exp(ΦΩ(v)σ)]dv.
    $

    Obviously, we have for any $ v \in \mathbb{R}^d $ such that $ v \cdot \Omega > 0 $

    $ - \frac{ \Phi_\Omega (v)}{\sigma} + \frac{ \Phi_\Omega (-v)}{\sigma} = - \frac{|v - \Omega|^2}{2\sigma} + \frac{|-v - \Omega|^2}{2\sigma} = 2 \frac{v \cdot \Omega}{\sigma} \gt 0 \, , $

    implying that $ \int_{\mathbb{R} ^d} \!{(v\cdot \Omega) f(v)} \; {\rm d}v > 0 $ and

    $ \Omega[f] = \frac{\int_{\mathbb{R} ^d} \!{f(v)v} \;{\rm d}v}{\left | \int_{\mathbb{R} ^d} \!{f(v)v} \;{\rm d}v \right |} = \frac{\int_{\mathbb{R} ^d} \!{\;(v \cdot \Omega) f(v)} \;{\rm d}v\;\Omega}{\left | \int_{\mathbb{R} ^d} \!{\;(v \cdot \Omega)f(v)} \;{\rm d}v\;\Omega \right |} = \Omega \, . $

    In [1], the following notion of generalized collision invariant (GCI) has been introduced.

    Definition 3.1. (GCI)

    Let $ \Omega \in \mathbb{S}^{d-1} $ be a fixed orientation. A function $ \psi = \psi (v) $ is called a generalized collision invariant of $ Q $ associated to $ \Omega $, if and only if

    $ \int_{\mathbb{R} ^d} \!{Q(f)(v) \psi (v) } \;{\rm d}v = 0 \, , $

    for all $ f $ such that $ (I_d - \Omega \otimes \Omega) \int_{\mathbb{R} ^d} \!{f(v)v} \; {\rm d}v = 0 $, that is such that $ \int_{\mathbb{R} ^d} \!{f(v)v } \; {\rm d}v \in \mathbb{R} \Omega $.

    In order to obtain the hydrodynamic limit of (1.3), for any fixed $ (t, x) \in \mathbb{R}_+ \times \mathbb{R}^d $, we multiply (2.2) by a function $ v \to \psi_{t, x} (v) $ and integrate with respect to $ v $ yielding

    $ Rdtf(t,x,v)ψt,x(v)dv+Rdvxf(t,x,v)ψt,x(v)dv=RdLf(t,x,)(f1(t,x,))ψt,x(v)dv=Rdf1(t,x,v)(Lf(t,x,)ψt,x)(v)dv.
    $
    (3.1)

    The above computation leads naturally to the following extension of the notion of collision invariant, see also [48].

    Definition 3.2. Let $ f = f(v)\geq 0 $ be an equilibrium of $ Q $. A function $ \psi = \psi (v) $ is called a collision invariant for $ Q $ associated to the equilibrium $ f $, if and only if $ \mathcal{L}_f ^\star \psi = 0 $, that is

    $ \int_{\mathbb{R} ^d} \!{(\mathcal{L}_f g )(v) \psi (v) } \;{\rm d}v = 0\;\; for~ any ~function ~g = g(v) \, . $

    We are looking for a good characterization of the linearized collision operator $ \mathcal{L}_f $ and its adjoint with respect to the leading order particle density $ f $. Motivated by (2.1), we need to determine the structure of the equilibria of $ Q $ which are given by Lemma 2.2.

    By Lemma 2.2, we know that for any $ (t, x) \in \mathbb{R}_+ \times \mathbb{R}^d $, there are $ (\rho (t, x), \Omega (t, x)) \in \mathbb{R}_+ \times (\mathbb{S}^{d-1} \cup \{0\}) $ such that $ f(t, x, \cdot) = \rho (t, x) M_{\Omega (t, x)} $, where

    $ \rho (t, x) = \rho [f (t, x, \cdot)] \, \, \text{ and } \, \, \Omega (t, x) = \Omega[f(t, x, \cdot)] \, . $

    The evolution of the macroscopic quantities $ \rho $ and $ \Omega $ follows from (2.2) and (3.1), by appealing to the moment method [54,55,56,57,58,59]. Next, we explicitly determine the linearization of the collision operator $ Q $ around its equilibrium distributions. For any orientation $ \Omega \in \mathbb{S}^{d-1} \cup \{0\} $ we introduce the pressure tensor

    $ {\mathcal{M}}_{\Omega} : = \int_{\mathbb{R} ^d} \!{(v-\Omega) \otimes (I_d - \Omega \otimes \Omega) (v - \Omega) M_{\Omega} (v) } \;{\rm d}v\, , $

    and the quantity

    $ c_1 : = \int_{\mathbb{R} ^d} \!{\;(v \cdot \Omega) M_{\Omega} (v)} \;{\rm d}v \gt 0\, . $

    We will check later, see Lemma 3.2, that the pressure tensor $ {\mathcal{M}}_{\Omega} $ is symmetric.

    Proposition 3.1. Let $ f = f(v) \geq 0 $ be an equlibrium distribution of $ Q $ with nonvanishing orientation, that is

    $ f = \rho M_{\Omega}, \;\; \mathit{\text{where}} \; \; \rho = \rho[f], \;\; \mathit{\text{and}} \; \; \Omega = \Omega[f] \in \mathbb{S}^{d-1} \, . $

    (1) The linearization $ \mathcal{L}_f = \, \mathit{\text{d}} Q_f $ is given by

    $ \mathcal{L}_f g = {\rm div}_v \left \{ \sigma \nabla_v g + g \nabla_v \Phi_\Omega - \frac{f}{\int_{\mathbb{R} ^d} \!{\;(v\cdot \Omega) f (v)} \;{\rm d}v} P_f \int_{\mathbb{R} ^d} \!{g(v) v } \;{\rm d}v \right \} \, , $

    where $ P_f : = I_d - \Omega[f] \otimes \Omega [f] $ is the orthogonal projection onto $ \{\xi \in \mathbb{R}^d : \; \xi \cdot \Omega[f] = 0\} $. In particular $ \mathcal{L}_{\rho M_{\Omega}} = \mathcal{L}_{M_{\Omega}} $.

    (2) The formal adjoint of $ \mathcal{L}_f $ is given by

    $ Lfψ=σdivv(MΩvψ)MΩ+PfvW[ψ],W[ψ]:=RdMΩ(v)vψdvRd(vΩ)MΩ(v)dv.
    $
    (3.2)

    (3) We have the identity

    $ \mathcal{L}_f (f(v-\Omega)) = \sigma \nabla_v f - {\rm div}_v \left ( f \frac{{\mathcal{M}}_{\Omega}}{c_1}\right ). $

    Note that $ {\rm div}_v $ refers to the divergence operator acting on matrices defined as applying the divergence operator over rows.

    Proof.

    (1) By standard computations we have

    $ \mathcal{L}_f g = \frac{\, \text{d}}{\, \text{d} s} \Big|_{s = 0} Q(f+sg) = {\rm div}_v \left \{\sigma \nabla_v g + g (v - \Omega[f] + \nabla_v V) - f \frac{\, \text{d}}{\, \text{d} s}\Big|_{s = 0} \Omega[f+sg] \right \} \, , $

    and

    $ \frac{\, \text{d}}{\, \text{d} s}\Big|_{s = 0}\Omega [f+sg] = \frac{(I_d - \Omega[f] \otimes \Omega[f])}{\left | \int_{\mathbb{R} ^d} \!{f(v) v } \;{\rm d}v\right | } \int_{\mathbb{R} ^d} \!{ g(v)v} \;{\rm d}v \, . $

    Therefore we obtain

    $ \mathcal{L}_f g = {\rm div}_v \left \{ \sigma \nabla_v g + g \nabla_v \Phi_\Omega - \frac{f}{\int_{\mathbb{R} ^d} \!{\;(v\cdot \Omega) f (v)} \;{\rm d}v} P_f \int_{\mathbb{R} ^d} \!{g(v) v } \;{\rm d}v \right \} \, . $

    (2) We have

    $ Rd(Lfg)(v)ψ(v)dv=Rd{σvg+gvΦΩfRd(vΩ)f(v)dvPfRdg(v)vdv}vψdv=Rdg[σdivvvψvψvΦΩ]dv+Rdpdvg(v)PfvRdf(v)vψ(v)dvRd(vΩ)f(v)dv,
    $

    implying

    $ \mathcal{L}_f ^\star \psi = \sigma \frac{{\rm div}_v (M_{\Omega} \nabla_v \psi)}{M_{\Omega}} + P_f v \cdot W[\psi] \, . $

    (3) For any $ i \in \{1, ..., d\} $ we have

    $ Lf(f(vΩ)i)=divv{(vΩ)i(σvf+fvΦΩ)+σfeifRdpdvMΩ(v)(vΩ)iPfvRdpdv(vΩ)MΩ(v)},
    $

    and therefore, since $ f = \rho M_\Omega $ satisfies $ \sigma \nabla_v f + f \nabla_v \Phi_\Omega = 0 $, we get

    $ Lf(f(vΩ))=σvfdivv(fRdpdvMΩ(v)(vΩ)PfvRdpdv(vΩ)MΩ(v))=σvfdivv(fMΩc1).
    $

    Notice that at any $ (t, x) \in \mathbb{R}_+ \times \mathbb{R}^d $, the function $ h = 1 $ is a collision invariant for $ Q $, associated to $ f(t, x, \cdot) $. Indeed, for any $ g = g(v) $ we have

    $ \int_{\mathbb{R} ^d} \!{Q(f(t, x, \cdot) + sg)} \;{\rm d}v = 0 \, , $

    implying that $ \int_{\mathbb{R} ^d} \!{\; (\mathcal{L}_{f(t, x, \cdot)}g)(v)} \; {\rm d}v = 0 $ and therefore $ \mathcal{L}_{f(t, x, \cdot)} ^\star 1 = 0 $, for all $ (t, x) \in \mathbb{R}_+ \times \mathbb{R}^d $. Once we have determined a collision invariant $ \psi = \psi (t, x, v) $ at any $ (t, x) \in \mathbb{R}_+ \times \mathbb{R}^d $, we deduce, thanks to (3.1), a balance for the macroscopic quantities $ \rho (t, x) = \rho[f(t, x, \cdot)] $ and $ \Omega (t, x) = \Omega[f(t, x, \cdot)] $, given by the relationship

    $ Rdt(ρMΩ(t,x))ψ(t,x,v)dv+Rdvx(ρMΩ(t,x))ψ(t,x,v)dv=0.
    $
    (3.3)

    When taking as collision invariant the function $ h(t, x, v) = 1 $, we obtain the local mass conservation equation

    $ tρ+divx(ρRd(vΩ(t,x))MΩ(t,x)(v)dvΩ)=0.
    $
    (3.4)

    As usual, we are looking also for the conservation of the total momentum, however, the nonlinear operator $ Q $ does not preserve momentum. In other words, $ v $ is not a collision invariant. Indeed, if $ f = \rho M_{\Omega} $ is an equilibrium with nonvanishing orientation, we have

    $ \mathcal{L}_f ^\star v = \sigma \frac{\nabla_v M_{\Omega}}{M_{\Omega}} + \frac{P_f v}{\int_{\mathbb{R} ^d} \!p \;{\rm d}v{\;(v^{\prime} \cdot \Omega) M_{\Omega} (v^{\prime})}} = - \nabla_v \Phi_\Omega + \frac{P_f v}{\int_{\mathbb{R} ^d} \!p \;{\rm d}v{\;(v^{\prime} \cdot \Omega) M_{\Omega} (v^{\prime})}} \, , $

    and therefore $ v $ is not a collision invariant.

    We concentrate next on the resolution of (3.2). We will use the notation $ \partial_v \xi = \left(\tfrac{\partial \xi_i}{\partial v_j}\right) $ for the Jacobian matrix of a vector field $ \xi $ and $ {\rm div}_v $ for the divergence operator in $ v $ of both vectors and matrices with the convention of taking the divergence over the rows of the matrix. With this convention, we have

    $ RdgdivvAdv=RdAvgdvandRdξdivvηdv=Rdvξηdv
    $
    (3.5)

    for all smooth functions $ g $, vector fields $ \xi, \eta $, and matrices $ A $. We now focus in finding a parameterization of the kernel of the operator $ \mathcal{L}_f ^\star $.

    Lemma 3.1. Let $ f = \rho M_{\Omega} $ be an equilibrium of $ Q $ with nonvanishing orientation. The following two statements are equivalent:

    (1) $ \psi = \psi (v) $ is a collision invariant for $ Q $ associated to $ f $.

    (2) $ \psi $ satisfies

    $ σdivv(MΩvψ)MΩ+PfvW=0,
    $
    (3.6)

    for some vector $ W \in \ker ({\mathcal{M}}_{\Omega} - \sigma c_1 I_d) $.

    Moreover, the linear map $ W : \ker (\mathcal{L}_f ^\star) \to \ker ({\mathcal{M}}_{\Omega} - \sigma c_1 I_d) $, with $ W[\psi] : = \int_{\mathbb{R} ^d} \!{M_{\Omega} (v) \nabla_v \psi} \; {\rm d}v / c_1 $ induces an isomorphism between the vector spaces $ \ker (\mathcal{L}_f ^\star)/\ker W $ and $ \ker ({\mathcal{M}}_{\Omega} - \sigma c_1 I_d) $, where $ \ker W $ is the set of the constant functions.

    Proof.

    (1)$ \implies $(2) Since $ \psi $ is a collision invariant associated to $ f $, i.e. $ \mathcal{L}_f ^\star \psi = 0 $, and by the third statement in Proposition 3.1 we deduce (using also the first formula in (3.5) with $ f {\mathcal{M}}_{\Omega} /c_1 $ and $ \psi $)

    $ 0=RdLfψf(vΩ)dv=Rdψ(v)Lf(f(vΩ))dv=Rdψ(v)[σvfdivv(fMΩc1)]dv=σRdf(v)vψdv+MΩRdf(v)vψc1dv=ρσc1W[ψ]+ρMΩW[ψ].
    $

    Note that if $ \rho = 0 $ then $ f = 0 $ and $ \int v f \, \text{d} v = 0 $, implying that $ \Omega [f] = 0 $. Hence, since $ \Omega \neq 0 $, we have $ \rho > 0 $ and thus $ W[\psi] \in \ker ({\mathcal{M}}_{\Omega} - \sigma c_1 I_d) $, saying that (3.6) holds true with $ W = W[\psi] \in \ker({\mathcal{M}}_{\Omega} - \sigma c_1 I_d) $.

    (2)$ \implies $(1) Let $ \psi $ be a function satisfying (3.6) for some vector $ W \in \ker ({\mathcal{M}}_{\Omega} - \sigma c_1 I_d) $. Multiplying (3.6) by $ f(v-\Omega) $ and integrating with respect to $ v $ yields (thanks to the second formula in (3.5))

    $ - \sigma \rho\int_{\mathbb{R} ^d} \!{\partial_v (v - \Omega) \nabla_v \psi M_{\Omega} (v) } \;{\rm d}v + \rho {\mathcal{M}}_{\Omega} W = 0 \, , $

    which implies $ W[\psi] = W $ since $ {\mathcal{M}}_{\Omega} W = \sigma c_1 W $ by the assumption $ W \in \ker ({\mathcal{M}}_{\Omega} - \sigma c_1 I_d) $. Therefore $ \psi $ is a collision invariant for $ Q $, associated to $ f $

    $ \mathcal{L}_f ^\star \psi = \sigma \frac{{\rm div}_v (M_{\Omega} \nabla_v \psi)}{M_{\Omega}} + P_f v \cdot W[\psi] = \sigma \frac{{\rm div}_v (M_{\Omega} \nabla_v \psi)}{M_{\Omega}} + P_f v \cdot W = 0 \, . $

    Remark 3.1. For any non negative measurable function $ \chi = \chi (c, r) :]-1, +1[\times] 0, \infty[\to \mathbb{R} $ and any $ \Omega \in \mathbb{S}^{d-1} $, for $ d \geq 2 $, we have

    $ Rdχ(vΩ|v|,|v|)dv=|Sd2|R+π0χ(cosθ,r)rd1sind2θdθdr,
    $

    where $ |\mathbb{S}^{d-2}| $ is the surface of the unit sphere in $ \mathbb{R}^{d-1} $, for $ d \geq 3 $, and $ |\mathbb{S}^0| = 2 $ for $ d = 2 $. In particular we have the formula

    $ Rdχ(vΩ|v|,|v|)MΩ(v)dv=R+rd1π0χ(cosθ,r)e(cosθ,r)sind2θdθdrR+rd1π0e(cosθ,r)sind2θdθdr=R+rd1+11χ(c,r)e(c,r)(1c2)d32dcdrR+rd1+11e(c,r)(1c2)d32dcdr,
    $
    (3.7)

    where $ e(c, r) = \exp(rc/\sigma) \exp (- (r^2 + 1)/(2\sigma) - V(r)/\sigma) $.

    Notice that thanks to (3.7) the coefficient $ c_1 $ does not depend upon $ \Omega \in \mathbb{S}^{d-1} $

    $ c1=Rd(vΩ)exp(|vΩ|22σV(|v|)σ)dvRdexp(|vΩ|22σV(|v|)σ)dv=R+rdπ0cosθe(cosθ,r)sind2θdθdrR+rd1π0e(cosθ,r)sind2θdθdr.
    $

    In order to determine all the collision invariants, we focus on the spectral decomposition of the pressure tensor $ {\mathcal{M}}_{\Omega} $ for any $ \Omega \in \mathbb{S}^{d-1} $. In particular, the next lemma will imply the symmetry of the pressure tensor.

    Lemma 3.2. (Spectral decomposition of $ {\mathcal{M}}_{\Omega} $) For any $ \Omega \in \mathbb{S}^{d-1} $ we have $ {\mathcal{M}}_{\Omega} = \sigma c_1 (I_d - \Omega \otimes \Omega) $. In particular we have $ \ker ({\mathcal{M}}_{\Omega} - \sigma c_1 I_d) = (\mathbb{R} \Omega)^\perp $ and thus $ {\rm dim} \left (\ker (\mathcal{L}_f ^\star) /\ker W \right) = {\rm dim} \ker ({\mathcal{M}}_{\Omega} - \sigma c_1 I_d) = d-1 $, cf. Lemma 3.1.

    Proof. Let us consider $ \{E_1, \ldots, E_{d-1}\} $ an orthonormal basis of $ (\mathbb{R}\Omega) ^\perp $. By using the decomposition

    $ v - \Omega = ( \Omega \otimes \Omega ) (v-\Omega) + \sum\limits_{i = 1}^{d-1} ( E_i \otimes E_i ) (v-\Omega) = ( \Omega \otimes \Omega ) (v-\Omega) + \sum\limits_{i = 1}^{d-1} ( E_i \otimes E_i ) v \, , $

    one gets

    $ MΩ=Rd[(ΩΩ)(vΩ)+d1i=1(EiEi)v][d1j=1(EjEj)v]MΩ(v)dv.
    $
    (3.8)

    We claim that the following equalities hold true

    $ Rd[Ω(vΩ)](Ejv)MΩ(v)dv=0,1jd1,
    $
    (3.9)
    $ Rd(Eiv)(Ejv)MΩ(v)dv=δijRd|v|2(vΩ)2d1MΩ(v)dv,1i,jd1.
    $
    (3.10)

    Formula (3.9) is obtained by using the change of variable $ v = (I_d - 2 E_j \otimes E_j) v^{\prime} $. It is easily seen that

    $ \Omega \cdot ( v - \Omega) = \Omega \cdot ( v^{\prime} - \Omega), \;\;E_j \cdot v = - E_j \cdot v^{\prime}, \;\;M_{\Omega} (v) = M_{\Omega} (v^{\prime}), \;\;1\leq j \leq d-1 \, , $

    and therefore we have

    $ \int_{\mathbb{R} ^d} \!{[\Omega \cdot (v - \Omega)] (E_j \cdot v) M_{\Omega} (v) } \;{\rm d}v = - \int_{\mathbb{R} ^d} \!p \;{\rm d}v{[\Omega \cdot (v^{\prime} - \Omega)] (E_j \cdot v^{\prime}) M_{\Omega} (v^{\prime}) } $

    which implies (3.9). For the formulae (3.10) with $ i \neq j $, we appeal to the orthogonal transformation

    $ v = \mathcal{O}_{ij} v^{\prime}, \;\;\mathcal{O}_{ij} = \Omega \otimes \Omega + \sum\limits_{k \notin \{i, j\}} E_k \otimes E_k + E_i \otimes E_j - E_j \otimes E_i \, . $

    Notice that $ \mathcal{O}_{ij} \xi = \xi $, for all $ \xi \in ({\rm span}\{E_i, E_j\}) ^\perp $, $ \mathcal{O}_{ij} E_i = - E_j $, $ \mathcal{O}_{ij} E_j = E_i $ and therefore

    $ (E_i \cdot v) (E_j \cdot v) = - (E_j \cdot v^{\prime}) (E_i \cdot v^{\prime}) \, . $

    After this change of variable we deduce that

    $ \int_{\mathbb{R} ^d} \!{(E_i \cdot v) (E_j \cdot v ) M_{\Omega} (v) } \;{\rm d}v = 0, \;\;1\leq i, j \leq d-1, \;\;i \neq j \, , $

    and also

    $ \int_{\mathbb{R} ^d} \!{(E_i \cdot v)^2 M_{\Omega} (v) } \;{\rm d}v = \int_{\mathbb{R} ^d} \!{(E_j \cdot v)^2 M_{\Omega} (v) } \;{\rm d}v, \;\;1\leq i, j \leq d-1 \, . $

    Thanks to the equality $ \sum_{i = 1}^{d-1} (E_i \cdot v)^2 = |v|^2 - (v \cdot \Omega)^2 $, one gets

    $ \int_{\mathbb{R} ^d} \!{(E_i \cdot v) (E_j \cdot v) M_{\Omega} (v) } \;{\rm d}v = \delta_{ij} \int_{\mathbb{R} ^d} \!{\frac{|v|^2 - (v \cdot \Omega)^2}{d-1} M_{\Omega} (v)} \;{\rm d}v, \;\;1\leq i, j \leq d-1 \, . $

    Coming back to (3.8) we obtain

    $ MΩ=d1i=1(Rd(Eiv)2MΩ(v)dv)EiEi=Rd|v|2(vΩ)2d1MΩ(v)dv(IdΩΩ).
    $

    We are done if we prove that

    $ Rd|v|2(vΩ)2d1MΩ(v)dv=σc1.
    $

    Notice that, using (2.4):

    $ \big( ( |v|^2 I_d - v \otimes v ) \Omega \big) \cdot \nabla_v M_{\Omega} = \frac{|v|^2 - (v \cdot \Omega)^2}{\sigma} M_{\Omega} (v) \, , $

    and therefore

    $ Rd|v|2(vΩ)2σMΩ(v)dv=Rddivv[(|v|2Idvv)Ω]MΩ(v)dv=Rd[2(vΩ)d(vΩ)(vΩ)]MΩ(v)dv=(d1)c1.
    $

    By Lemmas 3.1 and 3.2 the computation of the collision invariants is reduced to the resolution of (3.6) for any $ W \in (\mathbb{R} \Omega) ^\perp $. Hence, if we denote by $ E_1, E_2, \ldots, E_{d-1} $ any orthonormal basis of $ (\mathbb{R} \Omega)^{\perp} $, we obtain a set of $ d-1 $ collision invariants $ \psi_{E_1}, \psi_{E_2}, \ldots, \psi_{E_{d-1}} $ for $ Q $ associated to the equilibrium distribution $ f $ such that $ E_i = W[\psi_{E_i}] $, $ i = 1, \dots, d-1 $. This set of collision invariants forms a basis for the $ \ker(\mathcal{L}_f ^\star) $. In the next section we will characterize this set of collision invariants and provide and easy manner to compute them (see Lemma 4.1).

    We conclude this section by showing that in our case the set of all GCIs of the operator $ Q $ coincide with the kernel of the operator $ \mathcal{L}_f ^\star $.

    Theorem 3.1. Let $ M_{\Omega} $ be an equilibrium of $ Q $ with nonvanishing orientation $ \Omega \in \mathbb{S}^{d-1} $. The set of collision invariants of $ Q $ associated to $ M_{\Omega} $ coincides with the set of the generalized collision invariants of $ Q $ associated to $ \Omega $.

    Proof. Let $ \psi = \psi (v) $ be a generalized collision invariant of $ Q $ associated to $ \Omega $. We denote by $ \{e_1, \ldots, e_d\} $ the canonical basis of $ \mathbb{R}^d $. For any $ f = f(v) $ satisfying $ (I_d - \Omega \otimes \Omega) e_i \cdot \int_{\mathbb{R} ^d} \!{f(v) v } \; {\rm d}v = 0, 1\leq i \leq d $, we have

    $ \int_{\mathbb{R} ^d} \!{f(\sigma \Delta_v \psi - \nabla_v \Phi_\Omega \cdot \nabla_v \psi )} \;{\rm d}v = \int_{\mathbb{R} ^d} \!{Q(f)(v) \psi (v)} \;{\rm d}v = 0 \, . $

    Therefore the linear form $ f \to \int_{\mathbb{R} ^d} \!{f(\sigma \Delta_v \psi - \nabla_v \Phi_\Omega \cdot \nabla_v \psi)} \; {\rm d}v $ is a linear combination of the linear forms $ f \to (I_d - \Omega \otimes \Omega) e_i \cdot \int_{\mathbb{R} ^d} \!{f(v) v } \; {\rm d}v $. We deduce that there is a vector $ \tilde{W} = (\tilde{W}_1, \ldots, \tilde{W}_d) \in \mathbb{R}^d $ such that

    $ \int_{\mathbb{R} ^d} \!{f(\sigma \Delta_v \psi - \nabla_v \Phi_\Omega \cdot \nabla_v \psi )} \;{\rm d}v + (I_d - \Omega \otimes \Omega) \tilde{W} \cdot \int_{\mathbb{R} ^d} \!{f(v) v } \;{\rm d}v = 0 \, , $

    for any $ f $ and thus

    $ \sigma \Delta_v \psi - \nabla_v \Phi_\Omega \cdot \nabla_v \psi + (I_d - \Omega \otimes \Omega)v\cdot \tilde{W} = 0 \, , $

    implying that $ \psi $ satisfies (3.6) with the vector $ W = (I_d - \Omega \otimes \Omega)\tilde{W} \in (\mathbb{R} \Omega) ^\perp $, that is, $ \psi $ is a collision invariant of $ Q $ associated to $ M_{\Omega} $.

    Conversely, let $ \psi = \psi (v) $ be a collision invariant of $ Q $ associated to $ M_{\Omega} $. By Lemmas 3.1 and 3.2 we know that there is $ W \in (\mathbb{R} \Omega)^\perp $ such that

    $ \sigma \Delta_v \psi - \nabla_v \Phi_\Omega \cdot \nabla_v \psi + v \cdot W = 0 \, . $

    Multiplying by any function $ f $ such that $ (I_d - \Omega \otimes \Omega) \int_{\mathbb{R} ^d} \!{f(v)v} \; {\rm d}v = 0 $ one gets

    $ \int_{\mathbb{R} ^d} \!{Q(f) (v) \psi (v)} \;{\rm d}v = \int_{\mathbb{R} ^d} \!{f(v) ( \sigma \Delta_v \psi - \nabla_v \Phi_\Omega \cdot \nabla_v \psi)} \;{\rm d}v = - \int_{\mathbb{R} ^d} \!{f(v)v} \;{\rm d}v \cdot W = 0 \, , $

    implying that $ \psi $ is a generalized collision invariant of $ Q $ associated to $ \Omega $.

    In this section we investigate the structure of the collision invariants of $ Q $ associated to an equilibrium distribution $ f = \rho M_{\Omega} $. By Lemmas 3.1 and 3.2, we need to solve the elliptic problem

    $ σdivv(MΩvψ)=(vW)MΩ(v),vRd,
    $
    (4.1)

    for any $ W \in (\mathbb{R}\Omega)^\perp $. We appeal to a variational formulation by considering the continuous bilinear symmetric form $ a_\Omega: H^1_{M_{\Omega}} \times H^1_{M_{\Omega}} \to \mathbb{R} $ defined as

    $ a_\Omega (\chi, \theta) = \sigma \int_{\mathbb{R} ^d} \!{\nabla_v \chi \cdot \nabla_v \theta\; M_{\Omega} (v)} \;{\rm d}v, \;\;\chi, \theta \in H^1_{M_{\Omega}} \, , $

    and the linear form $ l : H^1_{M_{\Omega}} \to \mathbb{R} $, $ l (\theta) = \int_{\mathbb{R} ^d} \!{\theta(v) (v \cdot W) M_{\Omega} (v)} \; {\rm d}v, \theta \in H^1_{M_{\Omega}} $. Notice that $ l $ is well defined and bounded on $ H^1_{M_{\Omega}} $ provided that the additional hypothesis $ |v| \in L^2_{M_{\Omega}} $ holds true, that is

    $ Rd|v|2exp(|vΩ|22σV(|v|)σ)dv<.
    $

    The above hypothesis is obviously satisfied by the potentials $ V_{\alpha, \beta} $. We say that $ \psi \in H^1_{M_{\Omega}} $ is a variational solution of (4.1) if and only if

    $ aΩ(ψ,θ)=l(θ) for any θH1MΩ.
    $
    (4.2)

    Proposition 4.1. A necessary and sufficient condition for the existence and uniqueness of variational solution to (4.1) is

    $ Rd(vW)MΩ(v)dv=0.
    $
    (4.3)

    Proof. The necessary condition for the solvability of (4.1) is obtained by taking $ \theta = 1 $ (which belongs to $ H^1_{M_{\Omega}} $ thanks to (2.5)) in (4.2) leading to (4.3). This condition is satisfied for any $ W \in (\mathbb{R}\Omega)^\perp $ since we have

    $ \int_{\mathbb{R} ^d} \!{(v \cdot W)M_{\Omega} (v)} \;{\rm d}v = \int_{\mathbb{R} ^d} \!{(v \cdot \Omega)M_{\Omega} (v)} \;{\rm d}v\;(\Omega \cdot W) = 0 \, . $

    The condition (4.3) also guarantees the solvability of (4.1). Indeed, under the hypotheses (2.5) and (2.6), the bilinear form $ a_\Omega $ is coercive on the Hilbert space $ \tilde{H}^1_{M_{\Omega}} : = \{\chi \in H^1_{M_{\Omega}}\; :\; ((\chi, 1))_{M_{\Omega}} = 0\} $, i.e. we have:

    $ a_\Omega (\chi, \chi) \geq \frac{\min\{\sigma, \lambda_\Omega\}}{2} \|\chi \|^2_{M_{\Omega}}, \;\;\chi \in \tilde{H}^1_{M_{\Omega}} \, . $

    Applying the Lax-Milgram lemma to (4.2) with $ \psi $, $ \theta \in \tilde{H}^1_{M_{\Omega}} $ yields a unique function $ \psi \in \tilde{H}^1_{M_{\Omega}} $ such that

    $ aΩ(ψ,˜θ)=l(˜θ) for any ˜θ˜H1MΩ.
    $
    (4.4)

    Actually, the compatibility condition $ l(1) = 0 $ allows us to extend (4.4) to $ H^1_{M_{\Omega}} $. This follows by applying (4.4) with $ \tilde{\theta} = \theta - ((\theta, 1))_{M_{\Omega}} $, for $ \theta \in H^1_{M_{\Omega}} $. Moreover, the uniqueness of the solution for the problem on $ \tilde{H}^1_{M_{\Omega}} $ implies the uniqueness, up to a constant, of the solution for the problem on $ H^1_{M_{\Omega}} $.

    As observed in (3.1), the fluid equations for $ \rho $ and $ \Omega $ will follow by appealing to the collision invariants associated to the orientation $ \Omega \in \mathbb{S}^{d-1} $, for any $ W \in (\mathbb{R} \Omega)^\perp $. When $ W = 0 $, the solutions of (4.1) are all the constants, and we obtain the particle number balance (3.4). Consider now $ W \in (\mathbb{R} \Omega)^\perp \setminus \{0\} $ and $ \psi_W $ a solution of (4.1). Obviously we have $ \psi_W = \tilde{\psi}_W + \int_{\mathbb{R} ^d} \!{\psi_W (v)M_{\Omega} (v)} \; {\rm d}v $, where $ \tilde{\psi}_W $ is the unique solution of (4.1) in $ \tilde{H}^1_{M_{\Omega}} $. It is easily seen, thanks to (3.4) and the linearity of (3.3), that the balances corresponding to $ \psi_W $ and $ \tilde{\psi}_W $ are equivalent. Therefore for any $ W \in (\mathbb{R} \Omega)^\perp $ it is enough to consider only the solution of (4.1) in $ \tilde{H}^1_{M_{\Omega}} $. From now on, for any $ W \in (\mathbb{R} \Omega)^\perp $, we denote by $ \psi_W $ the unique variational solution of (4.1) verifying $ \int_{\mathbb{R} ^d} \!{\psi_W (v) M_{\Omega} (v)} \; {\rm d}v = 0 $. The structure of the solutions $ \psi_W, W \in (\mathbb{R} \Omega)^\perp\setminus \{0\} $ comes by the symmetry of the equilibrium $ M_{\Omega} $. Analyzing the rotations leaving invariant the vector $ \Omega $, we prove as in [48] the following result.

    Proposition 4.2. Consider $ W \in (\mathbb{R} \Omega)^\perp\setminus \{0\} $. For any orthogonal transformation $ \mathcal{O} $ of $ \mathbb{R}^d $ leaving $ \Omega $ invariant, that is $ \mathcal{O} \Omega = \Omega $, we have

    $ \psi_W (\mathcal{O} v) = \psi_{^t \mathcal{O} W } (v), \;\;v \in \mathbb{R}^d \, , $

    where $ ^t \mathcal{O} $ denotes the transpose of the matrix $ \mathcal{O} $.

    Proof. First of all notice that $ ^t \mathcal{O} W \in (\mathbb{R} \Omega) ^\perp \setminus \{0\} $. We know that $ \psi_W $ is the minimum point of the functional

    $ J_W (z) = \frac{\sigma}{2}\int_{\mathbb{R} ^d} \!{|\nabla_v z|^2 M_{\Omega} (v)} \;{\rm d}v - \int_{\mathbb{R} ^d} \!{(v \cdot W) z(v) M_{\Omega} (v)} \;{\rm d}v, z\in \tilde{H}^1_{M_{\Omega}} \, . $

    It is easily seen that, for any orthogonal transformation $ \mathcal{O} $ of $ \mathbb{R}^d $ leaving the orientation $ \Omega $ invariant, and any function $ z \in \tilde{H}^1_{M_{\Omega}} $, we have, by defining $ z_{\mathcal{O}}: = z\circ \mathcal{O} \in \tilde{H}^1_{M_{\Omega}} $

    $ M_{\Omega} \circ \mathcal{O} = M_{\Omega}, \;\;\nabla z_\mathcal{O} = \;^t \mathcal{O} (\nabla z)\circ \mathcal{O} \, . $

    Moreover, we obtain with the change of variables $ v' = \mathcal{O} v $ and using that $ M_\Omega (v) = M_\Omega (v') $:

    $ JtOW(zO)=σ2Rd|tO(z)(Ov)|2MΩ(v)dvRd(vtOW)z(Ov)MΩ(v)dv=σ2Rdpdv|(z)(v)|2MΩ(v)Rdpdv(vW)z(v)MΩ(v)=JW(z).
    $

    Finally, we deduce that

    $ \psi_W \circ \mathcal{O}\in \tilde{H}^1_{M_{\Omega}}, \;\; J_{^t \mathcal{O} W} (\psi_W \circ \mathcal{O}) = J_W (\psi_W ) \leq J_W(z \circ \;^t\mathcal{O}) = J_{^t \mathcal{O} W} (z) \, , $

    for any $ z \in \tilde{H}^1_{M_{\Omega}} $, implying that $ \psi_W \circ \mathcal{O} = \psi_{^t \mathcal{O} W} $.

    The computation of the collision invariants $ \{\psi_W\; :\; W \in (\mathbb{R} \Omega) ^\perp \setminus \{0\}\} $ can be reduced to the computation of one scalar function. For any orthonormal basis $ \{E_1, \ldots, E_{d-1}\} $ of $ (\mathbb{R} \Omega) ^\perp $ we define the vector field $ F = \sum_{i = 1}^{d-1} \psi_{E_i} E_i $. This vector field does not depend upon the basis $ \{E_1, \ldots, E_{d-1}\} $ and has the following properties, see [48].

    Lemma 4.1. The vector field $ F $ does not depend on the orthonormal basis $ \{ E_1, \ldots, E_{d-1}\} $ of $ (\mathbb{R}\Omega)^\perp $ and for any orthogonal transformation $ \mathcal{O} $ of $ \mathbb{R}^d $, preserving $ \Omega $, we have $ F \circ \mathcal{O} = \mathcal{O} F $. There is a function $ \chi $ such that

    $ F(v) = \chi \left ( \frac{v \cdot \Omega}{|v|}, |v| \right ) \frac{(I_d - \Omega \otimes \Omega)(v)}{\sqrt{|v|^2 - (v \cdot \Omega)^2}}, \;\;v\in \mathbb{R} ^d \setminus (\mathbb{R} \Omega) \, , $

    and thus, for any $ i \in \{1, ..., d-1\} $, we have

    $ ψEi(v)=F(v)Ei=χ(vΩ|v|,|v|)vEi|v|2(vΩ)2,vRd(RΩ).
    $
    (4.5)

    Proof. Let $ \{F_1, \ldots, F_{d-1}\} $ be another orthonormal basis of $ (\mathbb{R} \Omega)^\perp $. The following identities hold

    $ E_1 \otimes E_1 +\ldots +E_{d-1} \otimes E_{d-1} + \Omega \otimes \Omega = I_d, \;\;F_1 \otimes F_1 + \ldots+F_{d-1} \otimes F_{d-1} + \Omega \otimes \Omega = I_d \, , $

    and therefore

    $ d1i=1ψEiEi=d1i=1ψd1j=1(EiFj)FjEi=d1i=1d1j=1(EiFj)ψFjEi=d1j=1ψFjd1i=1(EiFj)Ei=d1j=1ψFjFj.
    $

    For any orthogonal transformation of $ \mathbb{R}^d $ such that $ \mathcal{O} \Omega = \Omega $ we obtain thanks to Proposition 4.2

    $ F\circ \mathcal{O} = \sum\limits_{i = 1} ^{d-1} ( \psi_{E_i} \circ \mathcal{O} ) \;E_i = \sum\limits_{i = 1} ^{d-1}\psi_{^t \mathcal{O} E_i} \;E_i = \mathcal{O} \sum\limits_{i = 1} ^{d-1}\psi_{^t \mathcal{O} E_i } \;^t \mathcal{O} E_i = \mathcal{O} F \, , $

    where, the last equality holds true since $ \{^t \mathcal{O} E_1, \ldots, \; ^t\mathcal{O} E_{d-1}\} $ is an orthonormal basis of $ (\mathbb{R} \Omega) ^\perp $. Let $ v \in \mathbb{R}^d \setminus (\mathbb{R} \Omega) $ and consider

    $ E ( v) = \frac{(Id- \Omega \otimes \Omega)v}{\sqrt{|v|^2 - (\Omega \cdot v)^2}} \, . $

    Notice that $ E \cdot \Omega = 0, |E| = 1 $. When $ d = 2 $, since the vector $ F(v) $ is orthogonal to $ \Omega $, there exists a function $ \Lambda = \Lambda (v) $ such that

    $ F(v) = \Lambda (v) E = \Lambda (v) \frac{(I_2- \Omega \otimes \Omega)v}{\sqrt{|v|^2 - (\Omega \cdot v)^2}}, \;\;v \in \mathbb{R}^2 \setminus ( \mathbb{R} \Omega) \, . $

    If $ d \geq 3 $, let us denote by $ ^\bot E $, any unitary vector orthogonal to $ E $ and $ \Omega $. Introducing the orthogonal matrix $ \mathcal{O} = I_d - 2 \; ^\bot E \otimes \; ^\bot E $ (which leaves $ \Omega $ invariant), we obtain $ F \circ \mathcal{O} = \mathcal{O} F $. Observe that

    $ 0 = \;^\bot E \cdot E = \;^\bot E\cdot \frac{v - (v \cdot \Omega)\Omega}{\sqrt{|v|^2 - (v \cdot \Omega) ^2}} = \frac{\;^\bot E\cdot v}{\sqrt{|v|^2 - (v \cdot \Omega) ^2}}, \;\;\mathcal{O} v = v \, , $

    and thus

    $ F(v) = F (\mathcal{O} v ) = \mathcal{O} F(v) = (I_d - 2 \;^\bot E \otimes \;^\bot E) F(v) = F(v) - 2 ( \;^\bot E \cdot F(v)) \;^\bot E \, , $

    from which it follows that $ ^\bot E \cdot F(v) = 0 $, for any vector $ ^ \bot E $ orthogonal to $ E $ and $ \Omega $. Hence, there exists a function $ \Lambda (v) $ such that

    $ F(v) = \Lambda (v) E (v) = \Lambda (v) \frac{(I_d - \Omega \otimes \Omega)v}{\sqrt{|v|^2 - (v \cdot \Omega)^2}}, \;\;v \in \mathbb{R} ^d \setminus (\mathbb{R} \Omega) \, . $

    We will show that $ \Lambda (v) $ depends only on $ v\cdot \Omega/|v| $ and $ |v| $. Indeed, for any $ d\geq 2 $, and any orthogonal transformation $ \mathcal{O} $ such that $ \mathcal{O} \Omega = \Omega $ we have $ F(\mathcal{O} v) = \mathcal{O} F(v) $, $ E (\mathcal{O} v) = \mathcal{O} E (v) $ because

    $ (I_d - \Omega \otimes \Omega) \mathcal{O} v = \mathcal{O} v - (\Omega \cdot \mathcal{O} v ) \Omega = \mathcal{O} (I_d - \Omega \otimes \Omega) v, $
    $ \sqrt{|\mathcal{O} v |^2 - (\mathcal{O} v \cdot \Omega) ^2 } = \sqrt{|v|^2 - (v \cdot \Omega) ^2} \, , $

    implying that $ \Lambda (\mathcal{O} v) = \Lambda (v) $, for any $ v \in \mathbb{R}^d \setminus (\mathbb{R} \Omega) $. We are done if we prove that $ \Lambda (v) = \Lambda (v^{\prime}) $ for any $ v, v^{\prime} \in \mathbb{R}^d \setminus (\mathbb{R} \Omega) $ such that $ v\cdot \Omega /|v| = v^{\prime} \cdot \Omega /|v^{\prime}|, |v| = |v^{\prime}|, v \neq v^{\prime} $. It is enough to consider the rotation $ \mathcal{O} $ such that

    $ \mathcal{O} E = E{\;^\prime}, \;\;(\mathcal{O} - I_d)_{{\rm span}\{E, E^{\;\prime}\}^\perp} = 0, \;\;E = \frac{(I_d - \Omega \otimes \Omega)v}{\sqrt{|v|^2 - (v \cdot \Omega)^2}}, \;E^{\;\prime} = \frac{(I_d - \Omega \otimes \Omega)v^{\prime}}{\sqrt{|v^{\prime}|^2 - (v^{\prime} \cdot \Omega)^2}} \, . $

    The equality $ \mathcal{O} E = E ^{\; \prime} $ implies that $ \mathcal{O} v = v^{\prime} $ and therefore $ \Lambda (v^{\prime}) = \Lambda (\mathcal{O} v) = \Lambda (v) $, showing that there exists a function $ \chi $ such that $ \Lambda (v) = \chi (v \cdot \Omega/|v|, |v|), v \in \mathbb{R}^d \setminus \{0\} $.

    In the last part of this section we concentrate on the elliptic problem satisfied by the function $ (c, r) \to \chi (c, r) $ introduced in Lemma 4.1. Even if $ \psi_{E_i} $ are eventually singular on $ \mathbb{R}\Omega $, it will be no difficulty to define a Hilbert space on which solving for the profile $ \chi $. We again proceed using the minimization of quadratic functionals.

    Proposition 4.3. The function $ \chi $ constructed in Lemma 4.1 solves the problem

    $ σc{rd3(1c2)d12e(c,r)cχ}σr{rd1(1c2)d32e(c,r)rχ}+σ(d2)rd3(1c2)d52eχ=rd(1c2)d22e(c,r),(c,r)]1,+1[×]0,[.
    $
    (4.6)

    Proof. For any $ i \in \{1, \ldots, d-1\} $, let us consider $ \psi_{E_i, h} (v) = h (v \cdot \Omega /|v|, |v|) \frac{v \cdot E_i}{\sqrt{|v|^2 - (v \cdot \Omega)^2}} $ where $ v \to h (v \cdot \Omega /|v|, |v|) $ is a function such that $ \psi_{E_i, h} \in H^1_{M_{\Omega}} $. Observe that if $ h = \chi $, then $ \psi_{E_i, h} $ coincides with $ \psi_{E_i} $. Note that generally $ \psi_{E_i, h} $ are not collision invariants, but perturbations of them, corresponding to profiles $ h $. In this way, the minimization problem (4.7) will lead to a minimization problem on $ h $, whose solution will be $ \chi $. Notice that once that $ \psi_{E_i, h} \in H^1_{M_{\Omega}} $, then $ \int_{\mathbb{R} ^d} \!{\psi_{E_i, h}M_{\Omega} (v)} \; {\rm d}v = 0 $, saying that $ \psi_{E_i, h} \in \tilde{H}^1_{M_{\Omega}} $. We know that $ \psi_{E_i} $ is the minimum point of $ J_{E_i} $ on $ \tilde{H}^1_{M_{\Omega}} $ and therefore

    $ JEi(ψEi)JEi(ψEi,h).
    $
    (4.7)

    A straightforward computation shows that

    $ vψEi,h=vEi|v|2(vΩ)2[chIdvv|v|2|v|Ω+rhv|v|]+h(vΩ|v|,|v|)[Id(v(vΩ)Ω)v|v|2(vΩ)2]Ei|v|2(vΩ)2,
    $

    and

    $ |vψEi,h|2=(vEi)2|v|4(ch)2+(vEi)2|v|2(vΩ)2(rh)2+|v|2(vΩ)2(vEi)2(|v|2(vΩ)2)2h2(vΩ|v|,|v|).
    $

    The condition $ \psi_{E_i, h} \in H^1_{M_{\Omega}} $ writes

    $ \int_{\mathbb{R} ^d} \!{(\psi_{E_i, h})^2 M_{\Omega} (v)} \;{\rm d}v \lt \infty, \;\;\int_{\mathbb{R} ^d} \!{|\nabla_v \psi_{E_i, h}|^2 M_{\Omega} (v)} \;{\rm d}v \lt \infty \, , $

    which is equivalent, thanks to the Poincaré inequality (2.6) to

    $ \int_{\mathbb{R} ^d} \!{|\nabla_v \psi_{E_i, h}|^2 M_{\Omega} (v)} \;{\rm d}v \lt \infty. $

    Based on formula (3.7), we have

    $ Rd|vψEi,h|2MΩ(v)dv=Rd[|v|2(vΩ)2(d1)|v|4(ch)2+(rh)2d1+(d2)h2(d1)(|v|2(vΩ)2)]MΩdv=R+rd1+11[1c2r2(ch)2+(rh)2+(d2)h2r2(1c2)]e(c,r)(1c2)d32dcdr(d1)R+rd1+11e(c,r)(1c2)d32dcdr}
    $

    and therefore we consider the Hilbert space $ H_d = \{h :\; ]-1, +1[\times] 0, \infty[\to \mathbb{R}, \|h \|_d ^2 < \infty \} $, endowed with the scalar product

    $ (g, h)_d = \int_{\mathbb{R}_+} r ^{d-1} \int_{-1} ^{+1}\left [\frac{1-c^2}{r^2} \partial_c g \partial_c h + \partial_r g \partial_r h + \frac{(d-2)gh}{r^2 (1- c^2)} \right ]e(c, r) (1- c^2) ^{\frac{d-3}{2}} \, \text{d} c\, \text{d} r $

    for $ g $ and $ h $ in $ H_d $ and the norm given by

    $ \| h \|_{d} = \sqrt{ ( h, h)_d} \, . $

    The expression $ J_{E_i}(\psi_{E_i, h}) $ writes as functional of $ h $

    $ JEi(ψEi,h)=σ2Rd|vψEi,h|2MΩ(v)dvRd(vEi)2h(vΩ|v|,|v|)|v|2(vΩ)2MΩ(v)dv=σ2Rd|vψEi,h|2MΩ(v)dvRdh(vΩ|v|,|v|)|v|2(vΩ)2d1MΩ(v)dv=J(h)(d1)R+rd1+11e(c,r)(1c2)d32dcdr,
    $

    where

    $ J(h)=σ2R+rd1+11[1c2r2(ch)2+(rh)2+(d2)h2r2(1c2)]e(c,r)(1c2)d32dcdrR+rd1+11h(c,r)r1c2e(c,r)(1c2)d32dcdr.
    $

    Coming back to (4.7) and using (4.5), we deduce that

    $ \chi \in H_d\;\;\mbox{ and } \;\;J(\chi) \leq J(h)\;\;\mbox{ for any }\;\; h \in H_d \, . $

    Therefore, by the Lax-Milgram lemma, we deduce that $ \chi $ solves the problem (4.6).

    After identifying the collision invariants, we determine the fluid equations satisfied by the macroscopic quantities entering the dominant particle density $ f(t, x, v) = \rho (t, x) M_{\Omega (t, x)} (v) $. As seen before the balance for the particle density follows thanks to the collision invariant $ \psi = 1 $. The other balances follow by appealing to the vector field $ F $ (cf. Lemma 4.1) and the details are given in section 5.1.

    Applying (3.1) with $ \psi = 1 $ leads to (3.4). For any $ (t, x) \in \mathbb{R}_+ \times \mathbb{R}^d $ we consider the vector field

    $ v \to F(t, x, v) = \chi \left (\frac{v \cdot \Omega (t, x)}{|v|}, |v| \right ) \frac{(I_d - \Omega (t, x) \otimes \Omega (t, x))v}{\sqrt{|v|^2 - (v \cdot \Omega (t, x))^2}} \, . $

    By the definition of $ F(t, x, \cdot) $, we know that

    $ \mathcal{L}_{f(t, x, \cdot)} ^\star F(t, x, \cdot) = 0, \;\;(t, x) \in \mathbb{R}_+ \times \mathbb{R}^d \, , $

    and therefore (3.1) implies

    $ RdtfF(t,x,v)dv+RdvxfF(t,x,v)dv=0,(t,x)R+×Rd.
    $
    (5.1)

    It remains to compute $ \int_{\mathbb{R} ^d} \!{\partial_t f F(t, x, v)} \; {\rm d}v $ and $ \int_{\mathbb{R} ^d} \!{v \cdot \nabla_x f F(t, x, v) } \; {\rm d}v $ in terms of $ \rho (t, x) $ and $ \Omega (t, x) $. By a direct computation we obtain

    $ \partial_t f = \partial_t \rho M_{\Omega} + \rho \frac{M_{\Omega} }{\sigma} (v - \Omega) \cdot \partial_t \Omega \, , $

    implying, thanks to the equalities $ \partial_t \Omega \cdot \Omega = 0 $ and $ v \cdot \partial_t \Omega = P_f v \cdot \partial_t \Omega $,

    $ RdtfFdv=Rd(tρ+ρσ(vΩ)tΩ)χ(vΩ|v|,|v|)MΩ(v)Pfv|Pfv|dv=tρRdχ(vΩ|v|,|v|)MΩ(v)Pfv|Pfv|dv+ρσRdχ(vΩ|v|,|v|)MΩ(v)PfvPfv|Pfv|dvtΩ.
    $
    (5.2)

    It is an easy exercise to show that the integral $ \int_{\mathbb{R} ^d} \!{\chi \left (\frac{v \cdot \Omega}{|v|}, |v|\right)M_{\Omega} (v) \frac{P_f v}{|P_f v |}} \; {\rm d}v $ vanishes and that the following relationship holds

    $ Rdχ(vΩ|v|,|v|)MΩ(v)PfvPfv|Pfv|dv=Rdχ(vΩ|v|,|v|)|v|2(vΩ)2d1MΩ(v)dv(IdΩΩ).
    $

    Therefore, by taking into account that $ \partial_t \Omega \cdot \Omega = 0 $, the equality (5.2) becomes

    $ RdtfFdv=˜c1ρσtΩ,˜c1=Rdχ(vΩ|v|,|v|)|v|2(vΩ)2d1MΩ(v)dv.
    $
    (5.3)

    Similarly we write (for any smooth vector field $ \xi (x) $, the notation $ \partial_x \xi $ stands for the Jacobian matrix of $ \xi $, i.e. $ (\partial_x \xi)_{ i, j} = \partial_{x_j} \xi_i $)

    $ v \cdot \nabla_x f = (v \cdot \nabla_x \rho ) M_{\Omega} + \frac{\rho}{\sigma} v \cdot ( \;^t \partial_x \Omega (v-\Omega)) M_{\Omega} \, , $

    implying

    $ Rd(vxf)Fdv=Rdχ(vΩ|v|,|v|)MΩ(v)PfvPfv|Pfv|dvxρ+ρσRd(vΩ)χ(vΩ|v|,|v|)MΩ(v)PfvPfv|Pfv|dv(xΩ)Ω=˜c1(IdΩΩ)xρ+ρσ˜c2xΩΩ,
    $
    (5.4)

    where

    $ \tilde{c}_2 = \int_{\mathbb{R} ^d} \!{(v \cdot \Omega) \chi \left ( \frac{v \cdot \Omega}{|v|}, |v|\right )\frac{\sqrt{|v|^2 - (v \cdot \Omega)^2}}{d-1}M_{\Omega} (v)} \;{\rm d}v \, . $

    Notice that in the above computations we have used $ (^t \partial_x \Omega) \Omega = 0 $ and

    $ \int_{\mathbb{R} ^d} \!{(v \cdot E_i) (v \cdot E_j) (v \cdot E_k) \chi \left ( \frac{v \cdot \Omega}{|v|}, |v|\right )M_{\Omega} (v)} \;{\rm d}v = 0 \, , $

    for any $ i, j, k \in \{1, \ldots, d-1\} $. Combining (5.1), (5.3) and (5.4) yields

    $ \tilde{c}_1 \frac{\rho}{\sigma} \partial_t \Omega + \frac{\rho}{\sigma} \tilde{c}_2 \;\partial_x \Omega \;\Omega + \tilde{c_1} (I_d - \Omega \otimes \Omega) \nabla_x \rho = 0 \, , $

    or equivalently

    $ tΩ+c2xΩΩ+σ(IdΩΩ)xρρ=0,
    $
    (5.5)

    where

    $ c2=˜c2˜c1=Rd(vΩ)χ(vΩ|v|,|v|)|v|2(vΩ)2MΩ(v)dvRdχ(vΩ|v|,|v|)|v|2(vΩ)2MΩ(v)dv=R+rd+1π0cosθχ(cosθ,r)e(cosθ,r)sind1θdθdrR+rdπ0χ(cosθ,r)e(cosθ,r)sind1θdθdr.
    $

    Let us start by noticing that the system (1.5) and (1.6) is hyperbolic as a consequence of Theorem 4.1 in [60]. On the other hand, the orientation balance Eq (5.5) propagates the constraint $ |\Omega| = 1 $. Indeed, let $ \Omega = \Omega (t, x) $ be a smooth solution of (5.5), satisfying $ |\Omega (0, \cdot)| = 1 $. Multiplying by $ \Omega (t, x) $ we obtain

    $ \frac{1}{2} \partial_t |\Omega |^2 + \frac{c_2}{2} ( \Omega (t, x) \cdot \nabla_x ) |\Omega |^2 = 0 \, , $

    implying that $ |\Omega| $ is constant along the characteristics of the vector field $ c_2 \Omega (t, x) \cdot \nabla_x $ and thus $ |\Omega (t, \cdot)| = |\Omega (0, \cdot)| = 1 $, for all $ t \geq 0 $.

    The rescaled Eq (1.3) can be considered as an intermediate model between the equations introduced in [37] and [48] when there are no phase transitions. In [48], the authors considered a strong relaxation towards the `terminal speed' (or cruise speed). Whereas in [37] the authors do not impose a penalization on the self-propelled/friction term. Our result could be applied to obtain the results in [48] without resorting to measures supported on the sphere by doing a double passage to the limit. First, passing to the limit $ \varepsilon \rightarrow 0 $ in (1.3), taking $ V = V_{\alpha, \beta} $, we obtain (1.5) and (1.6). Afterwards, we rescale $ V $ to $ \lambda \widetilde{V} $ in the system (1.5) and (1.6) and study the limit when $ \lambda \rightarrow \infty $. This amounts to study the behavior of the coefficients

    $ c_{ 1, \lambda} = \frac{\int_{\mathbb{R}_+}r^d \int_0 ^{\pi} {\cos \theta \, e_\lambda (\cos \theta, r) \sin ^{d-2} \theta} \;{\rm d}\theta\, \text{d} r}{\int_{\mathbb{R}_+}r^{d-1} \int_0 ^{\pi} {e_\lambda (\cos \theta, r) \sin ^{d-2} \theta} \;{\rm d}\theta\, \text{d} r} \, , $

    and

    $ c_{2, \lambda} = \frac{\int_{\mathbb{R}_+}r^{d+1} \int_0 ^{\pi} {\cos \theta \, \chi_\lambda (\cos \theta, r) \, e_\lambda (\cos \theta, r) \sin ^{d-1} \theta} \;{\rm d}\theta\, \text{d} r}{\int_{\mathbb{R}_+}r^{d} \int_0 ^{\pi} {\chi_\lambda (\cos \theta, r) \, e_\lambda (\cos \theta, r) \sin ^{d-1} \theta} \;{\rm d}\theta\, \text{d} r} \, , $

    when $ \lambda \rightarrow \infty $, where the function $ \chi_\lambda $ solves the elliptic problem

    $ σc[rd3(1c2)d12eλ(c,r)cχλ]σr[rd1(1c2)d32eλ(c,r)rχλ]+σ(d2)rd3(1c2)d52eλχλ=rd(1c2)d22eλ(c,r),
    $

    and $ e_\lambda (c, r) = \exp(rc/\sigma) \exp(- (r^2 + 1)/(2\sigma) - \lambda \widetilde{V} (r)/\sigma) $.

    In order to analyse the asymptotic behavior of $ c_{1, \lambda} $ we introduce the following result.

    Lemma 5.1. Let $ \varphi \in C^2 ((0, \infty); \mathbb{R}) $ and $ g \in C^0 ((0, \infty) \times (0, \pi); \mathbb{R}) $. Let us assume that

    i) $ \int_{\mathbb{R}_+} \int_0^\pi \exp (\lambda \varphi (r)) | g (r, \theta)| \, \mathit{\text{d}} \theta \, \mathit{\text{d}} r < \infty $,

    ii) The function $ \varphi $ has a unique global maximum at an interior point $ r_0 $,

    iii) $ \int_0^\pi g (r_0, \theta) \, \mathit{\text{d}} \theta \neq 0 $.

    Then the function $ G(\lambda) $ defined as

    $ G( \lambda) = \int_{\mathbb{R}_+} \int_0^\pi \exp ( \lambda \varphi ( r)) g ( r, \theta) \, \mathit{\text{d}} \theta \, \mathit{\text{d}} r \, , $

    has the the following asymptotic behavior

    $ G( \lambda) \sim \sqrt{ \frac{ 2 \pi }{ | \varphi'' ( r_0)|}} \frac{\exp ( \lambda \varphi ( r_0))}{ \sqrt{ \lambda}} \int_0^\pi g( r_0, \theta) \, \mathit{\text{d}} \theta \, , $

    as $ \lambda \rightarrow \infty $.

    The proof of this result is a direct application of the Laplace method, see for instance [61]. As an immediate consequence of Lemma 5.1 we obtain

    $ \lim\limits_{ \lambda \rightarrow \infty} c_{ 1, \lambda} = \frac{ r_0 \int_0^\pi \cos \theta \exp ( r_0 \cos \theta / \sigma) \sin^{d-2} \theta \, \text{d} \theta}{ \int_0^\pi \exp ( r_0 \cos \theta / \sigma) \sin^{d-2} \theta \, \text{d} \theta} \, , $

    where $ r_0 $ is the minimum of the potential function $ V_{\alpha, \beta} (r) $. Let us note that the asymptotic study of the coefficient $ c_{1, \lambda} $ when $ \lambda \rightarrow \infty $ can also be performed using Lemma 5.1 for more general potentials than $ V_{\alpha, \beta} (r) $. In particular, we could also consider smooth potentials $ V (|v|) $ having a unique global minimum $ r_0 $ such that $ V' (r) < 0 $, for $ 0 < r < r_0 $, and $ V' (r) > 0 $, for $ r > r_0 $. On the other hand, the asymptotic study of $ c_{2, \lambda} $ could be addressed following similar techniques as in [60], however, we do not dwell upon this matter here and leave it for a future work.

    In this paper, we have considered a flocking model consisting of a modified Cucker-Smale involving noise, alignment and self-propulsion. We have investigated its macroscopic limit when the time and space variables are rescaled to the macroscopic scale. In this limit, the velocity distribution converges to an equilibrium which depends on the local density and local mean orientation of the particles. The density and mean orientation evolve in space and time according to a hyperbolic system of equations named the Self-Organized Hydrodynamics. This system is akin to the system of isothermal compressible gas dynamics with the important difference that the velocity is a vector of unit norm. It is structurally identical with that obtained from the Vicsek dynamics, with differences only in the values of the coefficients. An interesting open problem would be to investigate whether the coefficients of the two models coincide in the limit where the Cucker-Smale model converges to the Vicsek model. Future work will investigate different variants in the Cucker-Smale model, possibly involving phase transitions (which have been avoided in the present work by an appropriate expression of the alignment force).

    PAS and PD acknowledge support by the Engineering and Physical Sciences Research Council (EPSRC) under grants no. EP/M006883/1 and EP/P013651/1. This work was accomplished during the visit of MB to Imperial College London with an ICL-CNRS Fellowship. JAC was partially supported by the EPSRC grant number EP/P031587/1. PD also acknowledges support by the Royal Society and the Wolfson Foundation through a Royal Society Wolfson Research Merit Award no. WM130048 and by the National Science Foundation (NSF) under grant no. RNMS11-07444 (KI-Net). PD is on leave from CNRS, Institut de Mathématiques de Toulouse, France. PAS and PD would like to thank the Isaac Newton Institute for Mathematical Sciences for support and hospitality during the programme ``Growth form and self-organisation'' when part of this work was carried over which was supported by: EPSRC grant numbers EP/K032208/1 and EP/R014604/1.

    Data statement: No new data were collected in the course of this research.

    The authors declare there is no conflict of interest.

    [1] Banks S, Dinges DF (2007) Behavioral and physiological consequences of sleep restriction. J Clin Sleep Med 3: 519-528.
    [2] Van Dongen HP, Maislin G, Mullington JM, et al. (2003) The cumulative cost of additional wakefulness: dose-response effects on neurobehavioral functions and sleep physiology from chronic sleep restriction and total sleep deprivation. Sleep 26: 117-126.
    [3] Knutson KL, Spiegel K, Penev P, et al. (2007) The metabolic consequences of sleep deprivation. Sleep Med Rev 11: 163-178. doi: 10.1016/j.smrv.2007.01.002
    [4] Dixit A, Mittal T (2015) Executive Functions are not Affected by 24 Hours of Sleep Deprivation: A Color-Word Stroop Task Study. Indian J Psychol Med 37: 165-168. doi: 10.4103/0253-7176.155615
    [5] Mignot E (2008) Why we sleep: the temporal organization of recovery. PLoS Biol 6: e106. doi: 10.1371/journal.pbio.0060106
    [6] McCormick DA, Bal T (1997) Sleep and arousal: thalamocortical mechanisms. Annu Rev Neurosci 20: 185-215. doi: 10.1146/annurev.neuro.20.1.185
    [7] Pace-Schott EF, Hobson JA (2002) The neurobiology of sleep: genetics, cellular physiology and subcortical networks. Nat Rev Neurosci 3: 591-605. doi: 10.1038/nrn895
    [8] Villablanca JR (2004) Counterpointing the functional role of the forebrain and of the brainstem in the control of the sleep-waking system. J Sleep Res 13: 179-208. doi: 10.1111/j.1365-2869.2004.00412.x
    [9] Krueger J CL, Rector D (2009) Cytokines and other neuromodulators. Stickgold R, Walker M (eds) The neuroscience of sleep.
    [10] Porkka-Heiskanen T (2011) Methylxanthines and sleep. Fredholm BB (ed) Methylxanthines, handbook of experimental pharmacology 331-348.
    [11] Krueger JM, Rector DM, Roy S, et al. (2008) Sleep as a fundamental property of neuronal assemblies. Nat Rev Neurosci 9: 910-919. doi: 10.1038/nrn2521
    [12] Vyazovskiy VV, Olcese U, Hanlon EC, et al. (2011) Local sleep in awake rats. Nature 472: 443-447. doi: 10.1038/nature10009
    [13] Llinas RR, Steriade M (2006) Bursting of thalamic neurons and states of vigilance. J Neurophysiol 95: 3297-3308. doi: 10.1152/jn.00166.2006
    [14] Coulon P, Budde T, Pape HC (2012) The sleep relay--the role of the thalamus in central and decentral sleep regulation. Pflugers Arch 463: 53-71. doi: 10.1007/s00424-011-1014-6
    [15] Steriade M MR (1990) Brainstem control of wakefulness and sleep. Plenum, New York.
    [16] Haas HL L, JS (2012) Waking with the hypothalamus. Pflugers Arch 463: 31-42. doi: 10.1007/s00424-011-0996-4
    [17] Basheer R SR, Thakkar MM, McCarley RW (2004) Adenosine and sleep-wake regulation. Progr Neurobiol 73: 379-396. doi: 10.1016/j.pneurobio.2004.06.004
    [18] Obal F, Jr., Krueger JM (2003) Biochemical regulation of non-rapid-eye-movement sleep. Front Biosci 8: d520-550. doi: 10.2741/1033
    [19] De Sarro G, Gareri P, Sinopoli VA, et al. (1997) Comparative, behavioural and electrocortical effects of tumor necrosis factor-alpha and interleukin-1 microinjected into the locus coeruleus of rat. Life Sci 60: 555-564. doi: 10.1016/S0024-3205(96)00692-3
    [20] Manfridi A, Brambilla D, Bianchi S, et al. (2003) Interleukin-1beta enhances non-rapid eye movement sleep when microinjected into the dorsal raphe nucleus and inhibits serotonergic neurons in vitro. Eur J Neurosci 18: 1041-1049. doi: 10.1046/j.1460-9568.2003.02836.x
    [21] De A, Churchill L, Obal F, Jr., et al. (2002) GHRH and IL1beta increase cytoplasmic Ca(2+) levels in cultured hypothalamic GABAergic neurons. Brain Res 949: 209-212. doi: 10.1016/S0006-8993(02)03157-8
    [22] Huber R, Tononi G, Cirelli C (2007) Exploratory behavior, cortical BDNF expression, and sleep homeostasis. Sleep 30: 129-139.
    [23] Porkka-Heiskanen T, Alanko L, Kalinchuk A, et al. (2002) Adenosine and sleep. Sleep Med Rev 6: 321-332. doi: 10.1053/smrv.2001.0201
    [24] Oishi Y, Huang ZL, Fredholm BB, et al. (2008) Adenosine in the tuberomammillary nucleus inhibits the histaminergic system via A1 receptors and promotes non-rapid eye movement sleep. Proc Natl Acad Sci U S A 105: 19992-19997. doi: 10.1073/pnas.0810926105
    [25] Rosenberg PA, Li Y, Le M, et al. (2000) Nitric oxide-stimulated increase in extracellular adenosine accumulation in rat forebrain neurons in culture is associated with ATP hydrolysis and inhibition of adenosine kinase activity. J Neurosci 20: 6294-6301.
    [26] MM H (2011) Thalamocortical dynamics of sleep: roles of purinergic neuromodulation. Semin Cell Dev Biol 22: 245-251. doi: 10.1016/j.semcdb.2011.02.008
    [27] Steriade M, McCormick DA, Sejnowski TJ (1993) Thalamocortical oscillations in the sleeping and aroused brain. Science 262: 679-685. doi: 10.1126/science.8235588
    [28] Brown RE, Basheer R, McKenna JT, et al. (2012) Control of sleep and wakefulness. Physiol Rev 92: 1087-1187. doi: 10.1152/physrev.00032.2011
    [29] Rechtschaffen A KA (1968) A manual of Standardized Terminology, Techniques and Scoring System for Sleep Stages of Human Subjects. National Institue of Health Publication Washington, DC: NIH US Goverment Printing Office
    [30] Jeanneret PR, Webb WB (1963) Strength of grip on arousal from full night's sleep. Percept Mot Skills 17: 759-761. doi: 10.2466/pms.1963.17.3.759
    [31] Tassi P, Muzet A (2000) Sleep inertia. Sleep Med Rev 4: 341-353. doi: 10.1053/smrv.2000.0098
    [32] Klemm WR (2011) Why does rem sleep occur? A wake-up hypothesis. Front Syst Neurosci 5: 73.
    [33] Roffwarg HP, Muzio JN, Dement WC (1966) Ontogenetic development of the human sleep-dream cycle. Science 152: 604-619. doi: 10.1126/science.152.3722.604
    [34] Endo T, Roth C, Landolt HP, et al. (1998) Selective REM sleep deprivation in humans: effects on sleep and sleep EEG. Am J Physiol 274: R1186-1194.
    [35] Barone DA, Krieger AC (2014) Muscle Tone Control of REM Sleep. REM Sleep: Characteristics, Disorders, and Physiological Effects / Editor: Chelsea L Saylor
    [36] Kamphuis J, Lancel M, Koolhaas JM, et al. (2015) Deep sleep after social stress: NREM sleep slow-wave activity is enhanced in both winners and losers of a conflict. Brain Behav Immun 47:149-54. doi: 10.1016/j.bbi.2014.12.022
    [37] Benington JH, Heller HC (1995) Restoration of brain energy metabolism as the function of sleep. Prog Neurobiol 45: 347-360. doi: 10.1016/0301-0082(94)00057-O
    [38] H. Blake RWG (1937) Brain potentials during sleep. Am J Physiol 119 692-703.
    [39] Friedman L, Bergmann BM, Rechtschaffen A (1979) Effects of sleep deprivation on sleepiness, sleep intensity, and subsequent sleep in the rat. Sleep 1: 369-391.
    [40] Tobler I, Borbely AA (1986) Sleep EEG in the rat as a function of prior waking. Electroencephalogr Clin Neurophysiol 64: 74-76. doi: 10.1016/0013-4694(86)90044-1
    [41] Dijk DJ, Beersma DG, Daan S (1987) EEG power density during nap sleep: reflection of an hourglass measuring the duration of prior wakefulness. J Biol Rhythms 2: 207-219. doi: 10.1177/074873048700200304
    [42] Franken P, Tobler I, Borbely AA (1991) Sleep homeostasis in the rat: simulation of the time course of EEG slow-wave activity. Neurosci Lett 130: 141-144. doi: 10.1016/0304-3940(91)90382-4
    [43] Lancel M, van Riezen H, Glatt A (1992) The time course of sigma activity and slow-wave activity during NREMS in cortical and thalamic EEG of the cat during baseline and after 12 hours of wakefulness. Brain Res 596: 285-295. doi: 10.1016/0006-8993(92)91559-W
    [44] Huber R, Deboer T, Tobler I (2000) Effects of sleep deprivation on sleep and sleep EEG in three mouse strains: empirical data and simulations. Brain Res 857: 8-19. doi: 10.1016/S0006-8993(99)02248-9
    [45] Sanchez-Vives MV, Mattia M (2014) Slow wave activity as the default mode of the cerebral cortex. Arch Ital Biol 152: 147-155.
    [46] Meerlo P, Pragt BJ, Daan S (1997) Social stress induces high intensity sleep in rats. Neurosci Lett 225: 41-44. doi: 10.1016/S0304-3940(97)00180-8
    [47] Meerlo P, de Bruin EA, Strijkstra AM, et al. (2001) A social conflict increases EEG slow-wave activity during subsequent sleep. Physiol Behav 73: 331-335. doi: 10.1016/S0031-9384(01)00451-6
    [48] Berger RJ, Phillips NH (1995) Energy conservation and sleep. Behav Brain Res 69: 65-73. doi: 10.1016/0166-4328(95)00002-B
    [49] Berger RJ (1984) Slow wave sleep, shallow torpor and hibernation: homologous states of diminished metabolism and body temperature. Biol Psychol 19: 305-326. doi: 10.1016/0301-0511(84)90045-0
    [50] Siegel JM (2009) Sleep viewed as a state of adaptive inactivity. Nat Rev Neurosci 10: 747-753. doi: 10.1038/nrn2697
    [51] Rechtschaffen A (1998) Current perspectives on the function of sleep. Perspect Biol Med 41: 359-390. doi: 10.1353/pbm.1998.0051
    [52] Zepelin H, Rechtschaffen A (1974) Mammalian sleep, longevity, and energy metabolism. Brain Behav Evol 10: 425-470. doi: 10.1159/000124330
    [53] Horne J (2002) Why sleep? Biologist (London) 49: 213-216.
    [54] Schmidt MH (2014) The energy allocation function of sleep: A unifying theory of sleep, torpor, and continuous wakefulness. Neurosci Biobehav Rev 47c: 122-153.
    [55] Adam K (1980) Sleep as a restorative process and a theory to explain why. Prog Brain Res 53: 289-305. doi: 10.1016/S0079-6123(08)60070-9
    [56] Oswald I (1980) Sleep as restorative process: human clues. Prog Brain Res 53: 279-288. doi: 10.1016/S0079-6123(08)60069-2
    [57] Landgraf D, Shostak A, Oster H (2012) Clock genes and sleep. Pflugers Arch 463: 3-14. doi: 10.1007/s00424-011-1003-9
    [58] Wisor JP (2012) A metabolic-transcriptional network links sleep and cellular energetics in the brain. Pflugers Arch 463: 15-22. doi: 10.1007/s00424-011-1030-6
    [59] Cirelli C, Gutierrez CM, Tononi G (2004) Extensive and divergent effects of sleep and wakefulness on brain gene expression. Neuron 41: 35-43. doi: 10.1016/S0896-6273(03)00814-6
    [60] Mackiewicz M, Shockley KR, Romer MA, et al. (2007) Macromolecule biosynthesis: a key function of sleep. Physiol Genomics 31: 441-457. doi: 10.1152/physiolgenomics.00275.2006
    [61] Cirelli C (2006) Cellular consequences of sleep deprivation in the brain. Sleep Med Rev 10: 307-321. doi: 10.1016/j.smrv.2006.04.001
    [62] Clugston GA, Garlick PJ (1982) The response of protein and energy metabolism to food intake in lean and obese man. Hum Nutr Clin Nutr 36c: 57-70.
    [63] Clugston GA, Garlick PJ (1982) The response of whole-body protein turnover to feeding in obese subjects given a protein-free, low-energy diet for three weeks. Hum Nutr Clin Nutr 36: 391-397.
    [64] Golden MH, Waterlow JC (1977) Total protein synthesis in elderly people: a comparison of results with [15N]glycine and [14C]leucine. Clin Sci Mol Med 53: 277-288.
    [65] Horne JA (1980) Sleep and body restitution. Experientia 36: 11-13. doi: 10.1007/BF02003942
    [66] Meddis R (1975) On the function of sleep. Anim Behav 23: 676-691. doi: 10.1016/0003-3472(75)90144-X
    [67] Rial RV, Nicolau MC, Gamundi A, et al. (2007) The trivial function of sleep. Sleep Med Rev 11: 311-325. doi: 10.1016/j.smrv.2007.03.001
    [68] Webb WB (1974) Sleep as an adaptive response. Percept Mot Skills 38: 1023-1027. doi: 10.2466/pms.1974.38.3c.1023
    [69] Villafuerte G, Miguel-Puga A, Rodriguez EM, et al. (2015) Sleep deprivation and oxidative stress in animal models: a systematic review. Oxid Med Cell Longev 2015: 234952.
    [70] Komoda Y, Honda K, Inoue S (1990) SPS-B, a physiological sleep regulator, from the brainstems of sleep-deprived rats, identified as oxidized glutathione. Chem Pharm Bull (Tokyo) 38: 2057-2059. doi: 10.1248/cpb.38.2057
    [71] Honda K, Komoda Y, Inoue S (1994) Oxidized glutathione regulates physiological sleep in unrestrained rats. Brain Res 636: 253-258. doi: 10.1016/0006-8993(94)91024-3
    [72] Kimura M, Kapas L, Krueger JM (1998) Oxidized glutathione promotes sleep in rabbits. Brain Res Bull 45: 545-548. doi: 10.1016/S0361-9230(97)00441-3
    [73] Krueger JM, Obal F, Jr., Fang J (1999) Why we sleep: a theoretical view of sleep function. Sleep Med Rev 3: 119-129. doi: 10.1016/S1087-0792(99)90019-9
    [74] Basner M, Rao H, Goel N, et al. (2013) Sleep deprivation and neurobehavioral dynamics. Curr Opin Neurobiol 23: 854-863. doi: 10.1016/j.conb.2013.02.008
    [75] Hennevin E, Huetz C, Edeline JM (2007) Neural representations during sleep: from sensory processing to memory traces. Neurobiol Learn Mem 87: 416-440. doi: 10.1016/j.nlm.2006.10.006
    [76] Tononi G, Cirelli C (2014) Sleep and the price of plasticity: from synaptic and cellular homeostasis to memory consolidation and integration. Neuron 81: 12-34. doi: 10.1016/j.neuron.2013.12.025
    [77] Xie L, Kang H, Xu Q, et al. (2013) Sleep drives metabolite clearance from the adult brain. Science 342: 373-377. doi: 10.1126/science.1241224
    [78] Mendelsohn AR, Larrick JW (2013) Sleep facilitates clearance of metabolites from the brain: glymphatic function in aging and neurodegenerative diseases. Rejuvenation Res 16: 518-523. doi: 10.1089/rej.2013.1530
    [79] Spira AP, Gamaldo AA, An Y, et al. (2013) Self-reported sleep and beta-amyloid deposition in community-dwelling older adults. JAMA Neurol 70: 1537-1543.
    [80] Hahn EA, Wang HX, Andel R, et al. (2014) A change in sleep pattern may predict Alzheimer disease. Am J Geriatr Psychiatry 22: 1262-1271. doi: 10.1016/j.jagp.2013.04.015
    [81] Lim AS, Yu L, Kowgier M, et al. (2013) Modification of the relationship of the apolipoprotein E epsilon4 allele to the risk of Alzheimer disease and neurofibrillary tangle density by sleep. JAMA Neurol 70: 1544-1551. doi: 10.1001/jamaneurol.2013.4215
    [82] Ambrosini MV, Giuditta A (2001) Learning and sleep: the sequential hypothesis. Sleep Med Rev 5: 477-490. doi: 10.1053/smrv.2001.0180
    [83] Ribeiro S, Mello CV, Velho T, et al. (2002) Induction of hippocampal long-term potentiation during waking leads to increased extrahippocampal zif-268 expression during ensuing rapid-eye-movement sleep. J Neurosci 22: 10914-10923.
    [84] Ribeiro S, Gervasoni D, Soares ES, et al. (2004) Long-lasting novelty-induced neuronal reverberation during slow-wave sleep in multiple forebrain areas. PLoS Biol 2: E24. doi: 10.1371/journal.pbio.0020024
    [85] Huber R, Ghilardi MF, Massimini M, et al. (2004) Local sleep and learning. Nature 430: 78-81. doi: 10.1038/nature02663
    [86] Walker MP, Stickgold R (2004) Sleep-dependent learning and memory consolidation. Neuron 44: 121-133. doi: 10.1016/j.neuron.2004.08.031
    [87] Stickgold R, Walker MP (2005) Memory consolidation and reconsolidation: what is the role of sleep? Trends Neurosci 28: 408-415. doi: 10.1016/j.tins.2005.06.004
    [88] Maquet P, Schwartz S, Passingham R, et al. (2003) Sleep-related consolidation of a visuomotor skill: brain mechanisms as assessed by functional magnetic resonance imaging. J Neurosci 23: 1432-1440.
    [89] Tononi G, Cirelli C (2006) Sleep function and synaptic homeostasis. Sleep Med Rev 10: 49-62. doi: 10.1016/j.smrv.2005.05.002
    [90] Kavanau JL (1997) Memory, sleep and the evolution of mechanisms of synaptic efficacy maintenance. Neuroscience 79: 7-44. doi: 10.1016/S0306-4522(96)00610-0
    [91] Stickgold R (2006) Neuroscience: a memory boost while you sleep. Nature 444: 559-560. doi: 10.1038/nature05309
    [92] Eichenbaum H (2007) To sleep, perchance to integrate. Proc Natl Acad Sci U S A 104: 7317-7318. doi: 10.1073/pnas.0702503104
    [93] Fenn KM, Nusbaum HC, Margoliash D (2003) Consolidation during sleep of perceptual learning of spoken language. Nature 425: 614-616. doi: 10.1038/nature01951
    [94] Ferrara M, Iaria G, De Gennaro L, et al. (2006) The role of sleep in the consolidation of route learning in humans: a behavioural study. Brain Res Bull 71: 4-9. doi: 10.1016/j.brainresbull.2006.07.015
    [95] Peigneux P, Laureys S, Fuchs S, et al. (2004) Are spatial memories strengthened in the human hippocampus during slow wave sleep? Neuron 44: 535-545. doi: 10.1016/j.neuron.2004.10.007
    [96] Gottselig JM, Hofer-Tinguely G, Borbely AA, et al. (2004) Sleep and rest facilitate auditory learning. Neuroscience 127: 557-561. doi: 10.1016/j.neuroscience.2004.05.053
    [97] Peters KR, Smith V, Smith CT (2007) Changes in sleep architecture following motor learning depend on initial skill level. J Cogn Neurosci 19: 817-829. doi: 10.1162/jocn.2007.19.5.817
    [98] Ellenbogen JM, Payne JD, Stickgold R (2006) The role of sleep in declarative memory consolidation: passive, permissive, active or none? Curr Opin Neurobiol 16: 716-722. doi: 10.1016/j.conb.2006.10.006
    [99] Ellenbogen JM, Hulbert JC, Stickgold R, et al. (2006) Interfering with theories of sleep and memory: sleep, declarative memory, and associative interference. Curr Biol 16: 1290-1294. doi: 10.1016/j.cub.2006.05.024
    [100] Roth TC, 2nd, Rattenborg NC, Pravosudov VV (2010) The ecological relevance of sleep: the trade-off between sleep, memory and energy conservation. Philos Trans R Soc Lond B Biol Sci 365: 945-959. doi: 10.1098/rstb.2009.0209
    [101] Fogel SMS, C. T. (2006) Declarative learningdependent changes in theta power during REM sleep. Sleep: A375-A375.
    [102] Born J, Rasch B, Gais S (2006) Sleep to remember. Neuroscientist 12: 410-424. doi: 10.1177/1073858406292647
    [103] Wyatt RJ, Fram DH, Kupfer DJ, et al. (1971) Total prolonged drug-induced REM sleep suppression in anxious-depressed patients. Arch Gen Psychiatry 24: 145-155. doi: 10.1001/archpsyc.1971.01750080049007
    [104] Siegel JM (2001) The REM sleep-memory consolidation hypothesis. Science 294: 1058-1063. doi: 10.1126/science.1063049
    [105] Rasch B, Pommer J, Diekelmann S, et al. (2009) Pharmacological REM sleep suppression paradoxically improves rather than impairs skill memory. Nat Neurosci 12: 396-397. doi: 10.1038/nn.2206
    [106] Irwin MR (2015) Why sleep is important for health: a psychoneuroimmunology perspective. Annu Rev Psychol 66: 143-172. doi: 10.1146/annurev-psych-010213-115205
    [107] Baglioni C, Battagliese G, Feige B, et al. (2011) Insomnia as a predictor of depression: a meta-analytic evaluation of longitudinal epidemiological studies. J Affect Disord 135: 10-19. doi: 10.1016/j.jad.2011.01.011
    [108] Dryman A, Eaton WW (1991) Affective symptoms associated with the onset of major depression in the community: findings from the US National Institute of Mental Health Epidemiologic Catchment Area Program. Acta Psychiatr Scand 84: 1-5. doi: 10.1111/j.1600-0447.1991.tb01410.x
    [109] Lee E, Cho HJ, Olmstead R, et al. (2013) Persistent sleep disturbance: a risk factor for recurrent depression in community-dwelling older adults. Sleep 36: 1685-1691.
    [110] Cho HJ, Lavretsky H, Olmstead R, et al. (2008) Sleep disturbance and depression recurrence in community-dwelling older adults: a prospective study. Am J Psychiatry 165: 1543-1550. doi: 10.1176/appi.ajp.2008.07121882
    [111] Jaussent I, Bouyer J, Ancelin ML, et al. (2011) Insomnia and daytime sleepiness are risk factors for depressive symptoms in the elderly. Sleep 34: 1103-1110.
    [112] Manber R, Edinger JD, Gress JL, et al. (2008) Cognitive behavioral therapy for insomnia enhances depression outcome in patients with comorbid major depressive disorder and insomnia. Sleep 31: 489-495.
    [113] Giedke H, Schwarzler F (2002) Therapeutic use of sleep deprivation in depression. Sleep Med Rev 6: 361-377. doi: 10.1016/S1087-0792(02)90235-2
    [114] Grozinger M, Kogel P, Roschke J (2002) Effects of REM sleep awakenings and related wakening paradigms on the ultradian sleep cycle and the symptoms in depression. J Psychiatr Res 36: 299-308. doi: 10.1016/S0022-3956(02)00022-5
    [115] Killgore WD, Kamimori GH, Balkin TJ (2011) Caffeine protects against increased risk-taking propensity during severe sleep deprivation. J Sleep Res 20: 395-403. doi: 10.1111/j.1365-2869.2010.00893.x
    [116] McKenna BS, Dickinson DL, Orff HJ, et al. (2007) The effects of one night of sleep deprivation on known-risk and ambiguous-risk decisions. J Sleep Res 16: 245-252. doi: 10.1111/j.1365-2869.2007.00591.x
    [117] Venkatraman V, Chuah YM, Huettel SA, et al. (2007) Sleep deprivation elevates expectation of gains and attenuates response to losses following risky decisions. Sleep 30: 603-609.
    [118] Killgore WD, Killgore DB, Day LM, et al. (2007) The effects of 53 hours of sleep deprivation on moral judgment. Sleep 30: 345-352.
    [119] Trinder J, Waloszek J, Woods MJ, et al. (2012) Sleep and cardiovascular regulation. Pflugers Arch 463: 161-168. doi: 10.1007/s00424-011-1041-3
    [120] Barone DA, Krieger AC (2013) Stroke and obstructive sleep apnea: a review. Curr Atheroscler Rep 15: 334. doi: 10.1007/s11883-013-0334-8
    [121] Mullington JM, Haack M, Toth M, et al. (2009) Cardiovascular, inflammatory, and metabolic consequences of sleep deprivation. Prog Cardiovasc Dis 51: 294-302. doi: 10.1016/j.pcad.2008.10.003
    [122] Vgontzas AN, Fernandez-Mendoza J, Liao D, et al. (2013) Insomnia with objective short sleep duration: the most biologically severe phenotype of the disorder. Sleep Med Rev 17: 241-254. doi: 10.1016/j.smrv.2012.09.005
    [123] Meng L, Zheng Y, Hui R (2013) The relationship of sleep duration and insomnia to risk of hypertension incidence: a meta-analysis of prospective cohort studies. Hypertens Res 36: 985-995. doi: 10.1038/hr.2013.70
    [124] Palagini L, Bruno RM, Gemignani A, et al. (2013) Sleep loss and hypertension: a systematic review. Curr Pharm Des 19: 2409-2419. doi: 10.2174/1381612811319130009
    [125] Suka M, Yoshida K, Sugimori H (2003) Persistent insomnia is a predictor of hypertension in Japanese male workers. J Occup Health 45: 344-350. doi: 10.1539/joh.45.344
    [126] Vgontzas AN, Liao D, Bixler EO, et al. (2009) Insomnia with objective short sleep duration is associated with a high risk for hypertension. Sleep 32: 491-497.
    [127] Fernandez-Mendoza J, Vgontzas AN, Liao D, et al. (2012) Insomnia with objective short sleep duration and incident hypertension: the Penn State Cohort. Hypertension 60: 929-935. doi: 10.1161/HYPERTENSIONAHA.112.193268
    [128] Chung WS, Lin CL, Chen YF, et al. (2013) Sleep disorders and increased risk of subsequent acute coronary syndrome in individuals without sleep apnea: a nationwide population-based cohort study. Sleep 36: 1963-1968.
    [129] Vozoris NT (2013) The relationship between insomnia symptoms and hypertension using United States population-level data. J Hypertens 31: 663-671. doi: 10.1097/HJH.0b013e32835ed5d0
    [130] Phillips B, Buzkova P, Enright P (2009) Insomnia did not predict incident hypertension in older adults in the cardiovascular health study. Sleep 32: 65-72.
    [131] Phillips B, Mannino DM (2007) Do insomnia complaints cause hypertension or cardiovascular disease? J Clin Sleep Med 3: 489-494.
    [132] Ayas NT, White DP, Manson JE, et al. (2003) A prospective study of sleep duration and coronary heart disease in women. Arch Intern Med 163: 205-209. doi: 10.1001/archinte.163.2.205
    [133] Hoevenaar-Blom MP, Spijkerman AM, Kromhout D, et al. (2011) Sleep duration and sleep quality in relation to 12-year cardiovascular disease incidence: the MORGEN study. Sleep 34: 1487-1492.
    [134] Mallon L, Broman JE, Hetta J (2002) Sleep complaints predict coronary artery disease mortality in males: a 12-year follow-up study of a middle-aged Swedish population. J Intern Med 251: 207-216. doi: 10.1046/j.1365-2796.2002.00941.x
    [135] Wang Q, Xi B, Liu M, et al. (2012) Short sleep duration is associated with hypertension risk among adults: a systematic review and meta-analysis. Hypertens Res 35: 1012-1018. doi: 10.1038/hr.2012.91
    [136] Cappuccio FP, Cooper D, D'Elia L, et al. (2011) Sleep duration predicts cardiovascular outcomes: a systematic review and meta-analysis of prospective studies. Eur Heart J 32: 1484-1492. doi: 10.1093/eurheartj/ehr007
    [137] Sabanayagam C, Shankar A, Buchwald D, et al. (2011) Insomnia symptoms and cardiovascular disease among older American Indians: the Native Elder Care Study. J Environ Public Health 2011: 964617.
    [138] Heslop P, Smith GD, Metcalfe C, et al. (2002) Sleep duration and mortality: The effect of short or long sleep duration on cardiovascular and all-cause mortality in working men and women. Sleep Med 3: 305-314. doi: 10.1016/S1389-9457(02)00016-3
    [139] Kronholm E, Laatikainen T, Peltonen M, et al. (2011) Self-reported sleep duration, all-cause mortality, cardiovascular mortality and morbidity in Finland. Sleep Med 12: 215-221. doi: 10.1016/j.sleep.2010.07.021
    [140] Ikehara S, Iso H, Date C, et al. (2009) Association of sleep duration with mortality from cardiovascular disease and other causes for Japanese men and women: the JACC study. Sleep 32: 295-301.
    [141] Suzuki E, Yorifuji T, Ueshima K, et al. (2009) Sleep duration, sleep quality and cardiovascular disease mortality among the elderly: a population-based cohort study. Prev Med 49: 135-141. doi: 10.1016/j.ypmed.2009.06.016
    [142] Dew MA, Hoch CC, Buysse DJ, et al. (2003) Healthy older adults' sleep predicts all-cause mortality at 4 to 19 years of follow-up. Psychosom Med 65: 63-73. doi: 10.1097/01.PSY.0000039756.23250.7C
    [143] Kripke DF, Garfinkel L, Wingard DL, et al. (2002) Mortality associated with sleep duration and insomnia. Arch Gen Psychiatry 59: 131-136. doi: 10.1001/archpsyc.59.2.131
    [144] LeBlanc M, Merette C, Savard J, et al. (2009) Incidence and risk factors of insomnia in a population-based sample. Sleep 32: 1027-1037.
    [145] Morin CM, LeBlanc M, Daley M, et al. (2006) Epidemiology of insomnia: prevalence, self-help treatments, consultations, and determinants of help-seeking behaviors. Sleep Med 7: 123-130. doi: 10.1016/j.sleep.2005.08.008
    [146] Ohayon M (1996) Epidemiological study on insomnia in the general population. Sleep 19: S7-15.
    [147] Ohayon MM (2002) Epidemiology of insomnia: what we know and what we still need to learn. Sleep Med Rev 6: 97-111. doi: 10.1053/smrv.2002.0186
    [148] Buysse DJ (2014) Sleep health: can we define it? Does it matter? Sleep 37: 9-17.
    [149] Besedovsky L, Lange T, Born J (2012) Sleep and immune function. Pflugers Arch 463: 121-137. doi: 10.1007/s00424-011-1044-0
    [150] Redwine L, Hauger RL, Gillin JC, et al. (2000) Effects of sleep and sleep deprivation on interleukin-6, growth hormone, cortisol, and melatonin levels in humans. J Clin Endocrinol Metab 85: 3597-3603.
    [151] Meier-Ewert HK, Ridker PM, Rifai N, et al. (2004) Effect of sleep loss on C-reactive protein, an inflammatory marker of cardiovascular risk. J Am Coll Cardiol 43: 678-683. doi: 10.1016/j.jacc.2003.07.050
    [152] Haack M, Sanchez E, Mullington JM (2007) Elevated inflammatory markers in response to prolonged sleep restriction are associated with increased pain experience in healthy volunteers. Sleep 30: 1145-1152.
    [153] Vgontzas AN, Zoumakis E, Bixler EO, et al. (2004) Adverse effects of modest sleep restriction on sleepiness, performance, and inflammatory cytokines. J Clin Endocrinol Metab 89: 2119-2126. doi: 10.1210/jc.2003-031562
    [154] van Leeuwen WM, Lehto M, Karisola P, et al. (2009) Sleep restriction increases the risk of developing cardiovascular diseases by augmenting proinflammatory responses through IL-17 and CRP. PLoS One 4: e4589. doi: 10.1371/journal.pone.0004589
    [155] Abedelmalek S, Chtourou H, Aloui A, et al. (2013) Effect of time of day and partial sleep deprivation on plasma concentrations of IL-6 during a short-term maximal performance. Eur J Appl Physiol 113: 241-248. doi: 10.1007/s00421-012-2432-7
    [156] Schmid SM, Hallschmid M, Jauch-Chara K, et al. (2011) Disturbed glucoregulatory response to food intake after moderate sleep restriction. Sleep 34: 371-377.
    [157] Stamatakis KA, Punjabi NM (2010) Effects of sleep fragmentation on glucose metabolism in normal subjects. Chest 137: 95-101. doi: 10.1378/chest.09-0791
    [158] Faraut B, Boudjeltia KZ, Dyzma M, et al. (2011) Benefits of napping and an extended duration of recovery sleep on alertness and immune cells after acute sleep restriction. Brain Behav Immun 25: 16-24. doi: 10.1016/j.bbi.2010.08.001
    [159] Shearer WT, Reuben JM, Mullington JM, et al. (2001) Soluble TNF-alpha receptor 1 and IL-6 plasma levels in humans subjected to the sleep deprivation model of spaceflight. J Allergy Clin Immunol 107: 165-170. doi: 10.1067/mai.2001.112270
    [160] Irwin M, Rinetti G, Redwine L, et al. (2004) Nocturnal proinflammatory cytokine-associated sleep disturbances in abstinent African American alcoholics. Brain Behav Immun 18: 349-360. doi: 10.1016/j.bbi.2004.02.001
    [161] Irwin M, Mascovich A, Gillin JC, et al. (1994) Partial sleep deprivation reduces natural killer cell activity in humans. Psychosom Med 56: 493-498. doi: 10.1097/00006842-199411000-00004
    [162] Irwin M, McClintick J, Costlow C, et al. (1996) Partial night sleep deprivation reduces natural killer and cellular immune responses in humans. Faseb j 10: 643-653.
    [163] Vgontzas AN, Pejovic S, Zoumakis E, et al. (2007) Daytime napping after a night of sleep loss decreases sleepiness, improves performance, and causes beneficial changes in cortisol and interleukin-6 secretion. Am J Physiol Endocrinol Metab 292: E253-261.
    [164] Faraut B, Nakib S, Drogou C, et al. (2015) Napping reverses the salivary interleukin-6 and urinary norepinephrine changes induced by sleep restriction. J Clin Endocrinol Metab 100: E416-426. doi: 10.1210/jc.2014-2566
    [165] Chaput JP, Despres JP, Bouchard C, et al. (2008) The association between sleep duration and weight gain in adults: a 6-year prospective study from the Quebec Family Study. Sleep 31: 517-523.
    [166] Patel SR, Malhotra A, White DP, et al. (2006) Association between reduced sleep and weight gain in women. Am J Epidemiol 164: 947-954. doi: 10.1093/aje/kwj280
    [167] Cappuccio FP, D'Elia L, Strazzullo P, et al. (2010) Quantity and quality of sleep and incidence of type 2 diabetes: a systematic review and meta-analysis. Diabetes Care 33: 414-420. doi: 10.2337/dc09-1124
    [168] Cooper AJ, Westgate K, Brage S, et al. (2015) Sleep duration and cardiometabolic risk factors among individuals with type 2 diabetes. Sleep Med 16: 119-125. doi: 10.1016/j.sleep.2014.10.006
    [169] Lou P, Qin Y, Zhang P, et al. (2015) Association of sleep quality and quality of life in type 2 diabetes mellitus: A cross-sectional study in China. Diabetes Res Clin Pract 107: 69-76. doi: 10.1016/j.diabres.2014.09.060
    [170] Spiegel K, Tasali E, Penev P, et al. (2004) Brief communication: Sleep curtailment in healthy young men is associated with decreased leptin levels, elevated ghrelin levels, and increased hunger and appetite. Ann Intern Med 141: 846-850. doi: 10.7326/0003-4819-141-11-200412070-00008
    [171] Spiegel K, Leproult R, Van Cauter E (1999) Impact of sleep debt on metabolic and endocrine function. Lancet 354: 1435-1439. doi: 10.1016/S0140-6736(99)01376-8
    [172] Conlon M, Lightfoot N, Kreiger N (2007) Rotating shift work and risk of prostate cancer. Epidemiology 18: 182-183. doi: 10.1097/01.ede.0000249519.33978.31
    [173] Kubo T, Ozasa K, Mikami K, et al. (2006) Prospective cohort study of the risk of prostate cancer among rotating-shift workers: findings from the Japan collaborative cohort study. Am J Epidemiol 164: 549-555. doi: 10.1093/aje/kwj232
    [174] Kubo T, Oyama I, Nakamura T, et al. (2011) Retrospective cohort study of the risk of obesity among shift workers: findings from the Industry-based Shift Workers' Health study, Japan. Occup Environ Med 68: 327-331. doi: 10.1136/oem.2009.054445
    [175] Parent ME, El-Zein M, Rousseau MC, et al. (2012) Night work and the risk of cancer among men. Am J Epidemiol 176: 751-759. doi: 10.1093/aje/kws318
    [176] Schwartzbaum J, Ahlbom A, Feychting M (2007) Cohort study of cancer risk among male and female shift workers. Scand J Work Environ Health 33: 336-343. doi: 10.5271/sjweh.1150
    [177] Haus EL, Smolensky MH (2013) Shift work and cancer risk: potential mechanistic roles of circadian disruption, light at night, and sleep deprivation. Sleep Med Rev 17: 273-284. doi: 10.1016/j.smrv.2012.08.003
    [178] von Ruesten A, Weikert C, Fietze I, et al. (2012) Association of sleep duration with chronic diseases in the European Prospective Investigation into Cancer and Nutrition (EPIC)-Potsdam study. PLoS One 7: e30972. doi: 10.1371/journal.pone.0030972
    [179] Jiao L, Duan Z, Sangi-Haghpeykar H, et al. (2013) Sleep duration and incidence of colorectal cancer in postmenopausal women. Br J Cancer 108: 213-221. doi: 10.1038/bjc.2012.561
    [180] Zhang X, Giovannucci EL, Wu K, et al. (2013) Associations of self-reported sleep duration and snoring with colorectal cancer risk in men and women. Sleep 36: 681-688.
    [181] Bishop D (2008) An applied research model for the sport sciences. Sports Med 38: 253-263. doi: 10.2165/00007256-200838030-00005
    [182] Drust B, Waterhouse J, Atkinson G, et al. (2005) Circadian rhythms in sports performance--an update. Chronobiol Int 22: 21-44. doi: 10.1081/CBI-200041039
    [183] Fullagar HH, Skorski S, Duffield R, et al. (2015) Sleep and Athletic Performance: The Effects of Sleep Loss on Exercise Performance, and Physiological and Cognitive Responses to Exercise. Sports Med 45(2):161-86.
    [184] Hausswirth C, Louis J, Aubry A, et al. (2014) Evidence of disturbed sleep and increased illness in overreached endurance athletes. Med Sci Sports Exerc 46: 1036-1045. doi: 10.1249/MSS.0000000000000177
    [185] Gleeson M (2007) Immune function in sport and exercise. J Appl Physiol (1985) 103: 693-699. doi: 10.1152/japplphysiol.00008.2007
    [186] Samuels C (2008) Sleep, recovery, and performance: the new frontier in high-performance athletics. Neurol Clin 26: 169-180; ix-x. doi: 10.1016/j.ncl.2007.11.012
    [187] Durmer JS, Dinges DF (2005) Neurocognitive consequences of sleep deprivation. Semin Neurol 25: 117-129. doi: 10.1055/s-2005-867080
    [188] Venter RE (2014) Perceptions of team athletes on the importance of recovery modalities. Eur J Sport Sci 14 Suppl 1: S69-76.
    [189] Erlacher D, Ehrlenspiel F, Adegbesan OA, et al. (2011) Sleep habits in German athletes before important competitions or games. J Sports Sci 29: 859-866. doi: 10.1080/02640414.2011.565782
    [190] Juliff LE, Halson SL, Peiffer JJ (2015) Understanding sleep disturbance in athletes prior to important competitions. J Sci Med Sport 18: 13-18. doi: 10.1016/j.jsams.2014.02.007
    [191] Hanton S, Fletcher D, Coughlan G (2005) Stress in elite sport performers: a comparative study of competitive and organizational stressors. J Sports Sci 23: 1129-1141. doi: 10.1080/02640410500131480
    [192] Chen JC, Brunner RL, Ren H, et al. (2008) Sleep duration and risk of ischemic stroke in postmenopausal women. Stroke 39: 3185-3192. doi: 10.1161/STROKEAHA.108.521773
    [193] Chien KL, Chen PC, Hsu HC, et al. (2010) Habitual sleep duration and insomnia and the risk of cardiovascular events and all-cause death: report from a community-based cohort. Sleep 33: 177-184.
    [194] Gangwisch JE, Malaspina D, Boden-Albala B, et al. (2005) Inadequate sleep as a risk factor for obesity: analyses of the NHANES I. Sleep 28: 1289-1296.
    [195] Cappuccio FP, D'Elia L, Strazzullo P, et al. (2010) Sleep duration and all-cause mortality: a systematic review and meta-analysis of prospective studies. Sleep 33: 585-592.
    [196] Machado RM, Koike MK (2014) Circadian rhythm, sleep pattern, and metabolic consequences: an overview on cardiovascular risk factors. Horm Mol Biol Clin Investig 18: 47-52.
    [197] Petrov ME, Lichstein KL (2015) Differences in sleep between black and white adults: an update and future directions. Sleep Med Jan 23, in press.
  • This article has been cited by:

    1. P. Aceves-Sanchez, P. Degond, E. E. Keaveny, A. Manhart, S. Merino-Aceituno, D. Peurichard, Large-Scale Dynamics of Self-propelled Particles Moving Through Obstacles: Model Derivation and Pattern Formation, 2020, 82, 0092-8240, 10.1007/s11538-020-00805-z
    2. José A. Carrillo, Rishabh S. Gvalani, Phase Transitions for Nonlinear Nonlocal Aggregation-Diffusion Equations, 2021, 382, 0010-3616, 485, 10.1007/s00220-021-03977-4
    3. Mihaï Bostan, José Antonio Carrillo, Fluid models with phase transition for kinetic equations in swarming, 2020, 30, 0218-2025, 2023, 10.1142/S0218202520400163
    4. José Antonio Carrillo, Mattia Zanella, Monte Carlo gPC Methods for Diffusive Kinetic Flocking Models with Uncertainties, 2019, 47, 2305-221X, 931, 10.1007/s10013-019-00374-2
    5. Jennifer Weissen, Simone Göttlich, Dieter Armbruster, Density dependent diffusion models for the interaction of particle ensembles with boundaries, 2021, 14, 1937-5077, 681, 10.3934/krm.2021019
    6. Pierre Degond, Angelika Manhart, Sara Merino-Aceituno, Diane Peurichard, Lorenzo Sala, How environment affects active particle swarms: a case study, 2022, 9, 2054-5703, 10.1098/rsos.220791
    7. Hyunjin Ahn, Seung-Yeal Ha, Myeongju Kang, Woojoo Shim, Emergent behaviors of relativistic flocks on Riemannian manifolds, 2021, 427, 01672789, 133011, 10.1016/j.physd.2021.133011
    8. Yunfei Su, Guochun Wu, Lei Yao, Yinghui Zhang, Hydrodynamic limit for the inhomogeneous incompressible Navier-Stokes-Vlasov equations, 2023, 342, 00220396, 193, 10.1016/j.jde.2022.09.029
    9. Pierre Degond, Antoine Diez, Mingye Na, Bulk Topological States in a New Collective Dynamics Model, 2022, 21, 1536-0040, 1455, 10.1137/21M1393935
    10. Mihaï Bostan, Anh-Tuan VU, Fluid Models for Kinetic Equations in Swarming Preserving Momentum, 2024, 22, 1540-3459, 667, 10.1137/21M145402X
    11. Yong‐Shen Zhou, Teng‐Fei Zhang, Qingxia Ma, Local well‐posedness of classical solutions for the kinetic self‐organized model of Cucker–Smale type, 2024, 47, 0170-4214, 7329, 10.1002/mma.9974
    12. Ning Jiang, Yi-Long Luo, Teng-Fei Zhang, From kinetic flocking model of Cucker–Smale type to self-organized hydrodynamic model, 2024, 34, 0218-2025, 2395, 10.1142/S0218202524500519
    13. Pierre Degond, Amic Frouvelle, Macroscopic limit of a Fokker-Planck model of swarming rigid bodies, 2024, 0956-7925, 1, 10.1017/S0956792524000111
    14. P. Degond, A. Diez, Topological Travelling Waves of a Macroscopic Swarmalator Model in Confined Geometries, 2023, 188, 0167-8019, 10.1007/s10440-023-00628-9
    15. Juan Pablo Pinasco, Nicolas Saintier, Martin Kind, Learning, Mean Field Approximations, and Phase Transitions in Auction Models, 2024, 14, 2153-0785, 396, 10.1007/s13235-023-00508-9
  • Reader Comments
  • © 2015 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(24357) PDF downloads(9457) Cited by(11)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog