ebook img

Global regularity of solutions of coupled Navier-Stokes equations and nonlinear Fokker-Planck equations PDF

0.18 MB·English
Save to my drive
Quick download
Download
Most books are stored in the elastic cloud where traffic is expensive. For this reason, we have a limit on daily download.

Preview Global regularity of solutions of coupled Navier-Stokes equations and nonlinear Fokker-Planck equations

Global regularity of solutions of coupled 9 Navier-Stokes equations and nonlinear Fokker 0 0 2 Planck equations n a J Peter Constantin Gregory Seregin 8 ∗ † 2 January 28, 2009 ] P A . h t a m Abstract We provide a proof of global regularity of solutions of coupled [ Navier-Stokes equations and Fokker-Planck equations, in two spatial dimen- 1 sions, in the absence of boundaries. The proof yields a priori estimates for v 2 the growth of spatial gradients. 6 4 4 . 1991 Mathematical subject classification (Amer. Math. Soc.): 35K, 1 0 35Q30, 82C31, 76A05. 9 Key Words: Navier-Stokes equations, nonlinear Fokker-Planck equations, 0 : global existence. v i X r 1 Introduction a We consider a system ∂ u+u u ν∆ u+ p = div σ, t x x x x ·∇ − ∇ div u = 0, (1.1) x  ∂ f +u f +div (Wf) = κ(∆ f +div (f U))  t x g g g g ·∇ ∇ ∗The University of Chicago †Oxford University 1 The functions u = u(x,t) R2, p(x,t) R and f(x,m,t) are the unknown functions with x T2, m∈ M and t∈ 0 independent variables. M is ∈ ∈ ≥ a compact connected Riemannian manifold without boundary of dimension N, with metric g (m). The operations ,div ,∆ are covariant derivative αβ g g g ∇ of scalars, its adjoint and Laplace-Beltrami operator, respectively, i.e. in a local chart m = (m1,...mN), U = (∂ U) g mα α=1,...N ∇ divgW = √1g∂mα √ggαβWβ ∆ f = div f g g∇g (cid:0) (cid:1) where, as customary, gαβ denotes the inverse of g , g its determinant and αβ repeated indices are summed. The second equation in (1.1) is the nonlinear Fokker-Planck equation. Linear Fokker-Planck equations arise naturally as Kolmogorov forward equations for the probability density distributions as- sociated with stochastic differential equations. Such linear equations with irregular coefficients were studied in ([10]). The cotangent field W is given by ∂u (x,t) i W(x,m,t) = u(x,t) : c(m) = c (m), (1.2) x ji ∇ ∂x j where c (m) = (c (m)) are smooth, time independent functions of ji ji;l l=1,...N m. The potential U(x,m,t) is given by U(x,m,t) = k(m,m)f(x,m,t)dm = U[f](x,m,t) (1.3) ′ ′ ′ ZM where dm stands for the Riemannian volume element, √gdm1...dmN. The interaction kernel k is Lipschitz continuous in M and is a given time inde- pendent function. The kernel is a symmetric function k(m,m) = k(m,m) ′ ′ and the operator f U[f] is bounded selfadjoint in L2(M). The cotangent 7→ fields c , the interaction kernel k, the kinematic viscosity ν > 0 and the ji microscopic diffusivity κ > 0 are all the parameters in the problem. The added stress σ is given by them by σ (x,t) = (c U(x,m,t) div c )fdm. (1.4) ij ij g g ij ·∇ − ZM The system (1.1) is a model of complex fluids ([2]) that is a natural gen- eralization of classical models of fluids with rod-like particles suspended in 2 them ([7]). The constitutive relation (1.4), modulo multiples of the identity matrix, was introduced ([3]) in order to have the natural energy balance d 1 u(x,t) 2 + [f](x,t) dx+ ν u 2 +κ [f](x,t) dx = 0, x dt 2| | E |∇ | D ZRd (cid:26) (cid:27) ZRd (cid:8) (cid:9) (1.5) where 1 [f] = f logf + fU[f] dm (1.6) E 2 ZM (cid:18) (cid:19) and [f] = (U[f]+logf) 2fdm. (1.7) g D |∇ | ZM We note that the Fokker-Planck equations satisfies the weak maximum prin- ciple (if u is a given smooth function) and therefore if f(x,m,0) 0, then ≥ f(x,m,t) 0. Moreover the micoscopic density ≥ ρ (x,t) = f(x,m,t)dm M ZM obeys the pure transport equation ∂ ρ +u ρ = 0, t M x M ·∇ and therefore, the region occupied by microscopic corpora is material (car- ried by the flow). In particular, from the fact that u is divergence-free it follows immediately that the density ρ L (dt;L1 L (dx)), if the initial ∞ ∞ ∈ ∩ density ρ (x,0) is bounded and integrable. The fluid density is taken to M be identically 1. We do not use the energy balance (1.5) in this paper, but the fundamental properties used for the proof of existence and regularity of solutions of (1.1) originate from the same source as the energy balance, namely the structure of the equations and the constitutive equation (1.4). In particular, we have, a priori, f 0, f L (dx;L1(M)) and consequently ∞ ≥ ∈ σ L . ∞ ∈ Global regularity for (1.1) was first proved in ([5]). Independently and simultaneously, global regularity for a similar model was proved in ([11]). That model is a version of the FENE model in which the physical gradient of velocityisreplacedbyitsanti-symmetricpart,andtheparticlesarerestricted to the unit disk by a potential that is infinite at the unit circle. Both proofs suffer from the fact that they are based on estimates with loss of regularity, 3 and they are non-quantitative. In particular, there is no a priori bound on the growth rate for the spatial gradients. In this paper we use results from ([6]) and the method of ([3]) and ([4]) to produce quantitative bounds. We use approximations that respect the basic properties of (1.1). We consider a standard mollifier x y J (u)(x) = δ 2 φ − u(y,t)dy δ − δ R2 Z (cid:18) (cid:19) with φ C (R2) and approximate (1.1) by ∈ 0∞ ∂ u+u u ν∆ u+ p = div J (σ), t x x x x δ ·∇ − ∇ div u = 0, (1.8) x  ∂ f +J (u) f +div (J (W)f) = κ(∆ f +div (f U))  t δ x g δ g g g ·∇ ∇ We obtain bounds independent of δ. Existence of solutions of the nonlinear system (1.8) with δ > 0 can be obtained by an implicit iteration scheme, using linear equations in each step of the approximation: ∂ u(n+1) +u(n)) u(n+1) ν∆ u(n+1) + p(n+1) = div J σ(n+1), t x x x x δ ·∇ − ∇ div u(n+1) = 0, x  ∂ f(n+1) +J (u(n)) f(n+1) +div (J (W(n))f(n+1)) =  t δ x g δ  κ(∆ f(n+1) +div (f·∇(n+1) U(n+1))) g g g ∇ (1.9)    The existence of solutions of (1.1) then follows fromthe existence of solutions of the approximate systems (1.8) and uniform bounds. The purpose of this article is to establish these bounds. In what follows all the bounds will be uniform in δ 0, and when δ = 0, then J is taken to be the identity. δ ≥ Definition 1.1. Let q > 2. We will say that (u ,f ) are standard initial 0 0 data if div u = 0, u W2,q(T2), f > 0, f (x,m) W1,q(T2;L2(M)) and x 0 0 0 0 ∈ ∈ f (x,m)dm = 1. M 0 R The Navier-Stokes equation can be written in the form ∂ u ν∆ u+ p = div τ, t − x ∇x x (1.10) div u = 0, x where τ (x,t) = J σ (x,t) u (x,t)u (x,t). (1.11) ij δ ij i j − 4 Taking the divergence of (1.10) we can solve for the pressure p = ( ∆ ) 1∂ ∂ τ . (1.12) x − i j ij − − using periodicboundaryconditions. Theoperatorτ pin(1.12)isbounded → in L2. The solutions of the Navier-Stokes equations discussed in this paper have τ( ,t) L2 and the pressure in the solution of (1.8) is meant to be given · ∈ by (1.12). 2 Statements of the main result and lemmas The main result we prove in this paper is Theorem 2.1. Let q 4, (u ,f ) be standard initial data and let T > 0 be 0 0 ≥ arbitrary. Let p > 2q , α > N +1. There exists a constant K depending only q 2 2 on the norms of the−initial data, T,κ,ν,p,q,α, with K bounded for bounded T, and a unique solution (u,f) of (1.1) with pressure p given by (1.12) and such that u K, (2.1) x x Lp(0,T;Lq(T2)) k∇ ∇ k ≤ sup xu( ,t) L∞ K, (2.2) k∇ · k ≤ t T ≤ and sup f( ,t) K. (2.3) k · kW1,q(T2;H−α(M))) ≤ t T ≤ hold. Remark 2.2. The constant K grows at most like a double exponential of T multiplied by a first order polynomial in T. Let H and V be the completions of the set of all divergence-free vector fields of C (T2;R2) with vanishing mean value on the torus, with respect to the ∞ L2 norm and the Dirichlet integral, respectively. The following result was proved in [6]: Proposition 2.3. Let u L (0,T;H) L2(0,T;V), p L2(0,T;L2(T2)) ∞ ∈ ∩ ∈ be a solution of the initial value problem ∂ u+u u ν∆u+ p = divσ, divu = 0, (2.4) t ·∇ − ∇ u( ,0) = u ( ) H, (2.5) 0 · · ∈ 5 where σ Lr(T2 (0,T);M2 2) with r 4. Then, given s > 0, there exists × ∈ × ≥ a constant C depending only on s, ν, the norm of u in H, the norm of σ s 0 in Lr(T2 (0,T)), such that × u L∞(T2 (s,T)) Cs. (2.6) k k × ≤ Moreover, the function u is H¨older continuous in T2 [s,T] with exponent × γ = 1 4. − r Remark 2.4. The existence and uniqueness of a solution to the initial value problem (2.4) and (2.5) with above properties is well known, see [8]. The proof of the result is based on local iterative estimates for L4 space- time integrals of the velocity, in the spirit of De Giorgi. The fact that u,p solve the Navier-Stokes equation (2.4), with p given in (1.12) with δ = 0, i.e., p = R R (σ u u ), (2.7) i j ij i j − 1 whereRi = ∂i( ∆x)−2 aretheRiesztransforms,isusedtorelatethepressure − to the velocity. The iteration relates integrals on smaller parabolic cubes to integrals on larger ones. For the iterative procedure to succeed, the modulus of absolute continuity of the map Ω T2 (0,T) u(x,t) 4dxdt, ⊂ × 7→ | | ZΩ (cid:8) (cid:9) needs tobecontrolled apriori,to guaranteethat suchanintegralisarbitrarily small, iftheparabolicLebesqguemeasureofΩissmallenough. Thefollowing result was used in ([6]) to control the modulus of absolute continuity. Proposition 2.5. Let u L (0,T;H) L2(0,T;V) be a solution of the 2D ∞ Navier-Stokes equations (∈2.4) with initi∩al data (2.5), u H Lr(T2) and 0 ∈ ∩ σ Lr(T2 (0,T);M2 2) with r 4. There exists a constant K depending × on∈ly on the×norm σ , ν≥,T and the norm of u in H Lr(T2) such Lr(T2 (0,T)) 0 k k × ∩ that sup u( ,t) K. (2.8) Lr(T2) k · k ≤ 0 t T ≤ ≤ Proposition 2.6. Let (u ,f ) be standard initial data. There exists T > 0 0 0 0 and a constant K, depending only on the initial data and the parameters ν,κ,p,q,α, where q > 2,p > 2q ,α > 1 + N, such that a unique solution q 2 2 − 6 of (1.1) exists on the time interval [0,T ], with pressure p given by (1.12), 0 satisfying kukLp(0,T0;Lq(dx))+ sup [ku(·,t)kW1,q(T2)+kf(·,t)kW1,q(T2;H−α(M))] ≤ K. (2.9) 0 t T0 ≤ ≤ Proposition 2.7. Let (u ,f ) be standard initial data and let T > 0 be 0 0 arbitrary. Let q 2, p > 2q , α > N + 1. There exists a constant K ≥ q 2 2 depending only on the initial−data, T,κ,ν,p,q,α, bounded for bounded T, such that, if (u,f) is a solution of (1.8) or a solution of (1.1) with u Lp(0,T;W2,q(T2)), pressure p given by (1.12), a∈nd with f Lp(0,T;W1,q(T2;H α(M))), − ∈ then sup ku(·,t)k2L∞(T2) +kσ(·,t)kL∞(T2) ≤ K (2.10) 0 t T ≤ ≤ h i holds. Lemma 2.8. Let (u ,f ) be standard initial data and let T > 0 be arbitrary. 0 0 Let q 4, p > 2q , α > N +1. There exists a constant K depending only on ≥ q 2 2 the initial data, T−,κ,ν,p,q,α, bounded for bounded T, such that, if (u,f) is a solution of (1.8) or a solution of (1.1) with u Lp(0,T;W2,q(T2)), pressure ∈ p given by (1.12) and f Lp(0,T;W1,q(T2;H α(M))), then − ∈ sup k∇xu(·,t)kL∞(T2) ≤ Klog(2+kfkLp(0,T;W1,q(T2;H−α(M)))). (2.11) 0 t T ≤ ≤ Theorem 2.9. Let T > 0, and u ,f be arbitrary standard initial data. Let 0 0 k 1 q 4 and assume that u Wk+1,q(T2) and f Wk,q(T2;Lq(M)). 0 0 ≥ ≥ ∈ ∈ Then, for any p > 2q , there exist constants K depending only on k,q,p, q 2 ν,κ,T and the norms−of the initial data, such that the solution of (1.1) on [0,T] with pressure p given by (1.12) sstisfies sup u( ,t) + u K (2.12) Wk,q(T2) Lp(0,T;Wk+1,q(T2)) k · k k k ≤ 0 T ≤ and sup f K. (2.13) Wk,q(T2;Lq(M))) k k ≤ t T ≤ 7 3 Proof of Proposition 2.5 We multiply (2.4) by u u r 2 and integrate in space and integrate by parts. − | | We obtain d u(x,t) rdx+ν u(x,t) 2 u(x,t) r 2dx rdt T2 | | T2 |∇x | | | − ≤ C [ σ(x,t) + p(x,t) ] u(x,t) u(x,t) r 2dx. RT2 | | | R| |∇x || | − Writing xu Ru r−2 = xu u r−22 u r−22, using a H¨older inequality with ex- |∇ || | |∇ || | | | ponents r,2, 2r , and then the Schwartz inequality and the viscous term, we r 2 deduce − d u( ,t) r C p( ,t) 2 + σ( ,t) 2 u r 2 dt k · k ≤ k · kLr k · kLr k kL−r Dividing by u r 2 and using th(cid:2)e boundedness of Riesz(cid:3)transforms we obtain k kL−r d u 2 C u 4 + σ 2 dtk kLr ≤ k kL2r k kLr (cid:2) (cid:3) Now we use the inequality u 2 C u + u u (3.14) k kL2r(T2) ≤ k kL2(T2) k∇x kL2(T2) k kLr(T2) (cid:2) (cid:3) to deduce that d u 2 C σ 2 +C u 2 + u 2 u 2 . dtk kLr ≤ k kLr k kL2 k∇ kL2 k kLr (cid:2) (cid:3) Because T u 2 + u 2 dt C k kL2 k∇ kL2 ≤ Z0 (cid:2) (cid:3) the bound (2.8) follows from Gronwall’s inequality. We present below a sketch of the proof of (3.14). We claim first that for any r 2 there exists a ≥ constant C , such that r kfk2L2r(R2) ≤ CrkfkLr(R2)k∇fkL2(R2) (3.15) holds for all f Lr(R2) with f L2(R2). This is a generalization of ∈ ∇ ∈ the well-known Ladyzhenskaya inequality ([8]) corresponding to r = 2. An elemenatry proof of (3.15) was given in ([6]). We give, for the sake of com- pleteness, a generalization and proof in the Appendix. The inequality (3.14) 8 follows by considering the function u as the restriction to [0,2π]2 of a peri- odic function U defined in the whole space R2, and taking a smooth com- pactly supported function φ that is identically 1 on an open neighbourhood of [0,2π]2. The inequality (3.15) holds for f = φU, and in view of the fact that φU C u and similar inequalities, we have Lr(R2) Lr(T2) k k ≤ k k u 2 f 2 k kL2r(T2) ≤ k kL2r(R2) ≤ C (φU) φU L2(R2) Lr(R2) k∇ k k k ≤ C φ U + U φ u L2(R2) L2(R2) Lr(T2) k ∇ k k ∇ k k k ≤ C u L2(T2) + u L2(T2) u Lr(T2). (cid:2)k∇ k k k k k (cid:3) (cid:2) (cid:3) 4 Proof of Proposition 2.6 In this section we will denote by C constants that may depend on ν,κ,T,p,q and α and are locally bounded in T > 0. We will denote by K constants that may depend in addition on standard initial data, and are locally bounded in T and the norms of standard initial data. All the constants are independent of δ 0. W≥e consider the vorticity, ω(x,t) = u(x,t) = ∂u2(x,t) ∂u1(x,t) and, ∇⊥ · ∂x1 − ∂x2 taking the curl of the Navier-Stokes equation, we obtain the vorticity equa- tion ∂ ω +u ω ν∆ ω = div J σ. (4.16) t ·∇x − x ∇⊥x · x δ We multiply this by ω q 2ω and integrate in space: − | | d ω qdx+ν(q 1) ω 2 ω q 2dx qdt T2| | − T2|∇x | | | − q−2 q−2 ≤ (Rq −1) T2|divxJδσ||ωR| 2 |∇xω||ω| 2 dx. h i Using the H¨older inequaRlity with exponents q, 2q ,2 and hiding the term q 2 involving the gradient of ω in the viscous term, w−e obtain qddt T2|ω|qdx+ ν(q2−1) T2|∇xω|2|ω|q−2dx (4.17) ≤ qR2−ν1kdivxJδσk2LqkωkRLq−q2. Integrating and using the well-known fact that the Lq norms of vorticity bound from above the Lq norms of the full gradient of velocity (modulo multiplicative constants), we obtain t u( ,t) 2 K +C div J σ( ,s) 2 ds. (4.18) k∇x · kLq ≤ k x δ · kLq Z0 9 The forces applied by the particles are obtained after f is integrated along with smooth coefficients on M in order to produce σ (1.4). Therefore, only very weak regularity of f with respect to the microscopic variables m is sufficient to control σ. We take advantage of this fact in order to control spatial gradients of f in terms of u L1(L ). We consider the L2(M) x ∞ ∇ ∈ selfadjoint pseudodifferential operator α R = ( ∆g +I)−2 (4.19) − with α > N +1. We differentiate the Fokker-Planck equation 2 ∂ f +J u f +div (J (W)f) = κdiv ( (logf +U[f])) (4.20) t δ x g δ g g ·∇ ∇ in (1.8) with respect to x, apply R, multiply by R f and integrate on M. x ∇ Let us denote by N(x,t)2 = R f(x,m,t) 2dm (4.21) x | ∇ | Z M the square of the L2 norm of R f on M. Note that x ∇ J σ(x,t) CN(x,t) (4.22) x δ |∇ | ≤ holds in view of the definition (1.4). We obtain 1 (∂ +J u )N2 C( J u +κ)N2 +C J u N (4.23) t δ x δ x δ x x 2 ·∇ ≤ | ∇ | | ∇ ∇ | pointwise in (x,t) with an absolute constant C. The proof of this fact ap- pearedinseveral places ([2], [3], [4]) andwillnotbereproducedhere. Nowwe multiply (4.23) by Nq−2, integrate dx, multiply by kNkpL−q(qdx) and use H¨older inequalities in both space and time: d N( ,t) p pdtk · kLq(dx) ≤ ≤ CkJδ∇x∇xu(·,t)kpLq(dx) +C kJδ∇xu(·,t)kL∞(dx) +1 kN(·,t)kpLq(dx). (4.24) (cid:0) (cid:1) In order to proceed we need to use the representation of the gradient of the solution, from the Navier-Stokes equation (1.10): t u(x,t) = eνt∆ u eν(t s)∆∆Hτ( ,s)ds (4.25) x x 0 − ∇ ∇ − · Z 0 10

See more

The list of books you might like

Most books are stored in the elastic cloud where traffic is expensive. For this reason, we have a limit on daily download.