ebook img

A Convergent Finite Difference Scheme for the Variational Heat Equation PDF

1.2 MB·
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 A Convergent Finite Difference Scheme for the Variational Heat Equation

A CONVERGENT FINITE DIFFERENCE SCHEME FOR THE VARIATIONAL HEAT EQUATION G. M. COCLITE, J. RIDDER, AND N. H. RISEBRO 7 1 0 2 Abstract. Thevariationalheatequationisanonlinear,parabolicequa- n tionnotindivergenceformthatarisesasamodelforthedynamicsofthe a J director field in a nematic liquid crystal. We present a finite difference 5 scheme for a transformed, possibly degenerate version of this equation and prove that a subsequence of the numerical solutions converges to a ] weak solution. This result is supplemented by numerical examples that A show that weak solutions are not unique and give some intuition about N how to obtain the physically relevant solution. . h 1. Introduction t a m In this paper we investigate the Cauchy problem [ (cid:40) (1) ut = c(u)(c(u)ux)x, x ∈ Ω, t > 0 1 u(x,0) = u (x), x ∈ Ω, v 0 5 where Ω = R or Ω = [0,1] with periodic boundary conditions. We assume 6 that 2 1 (H.1) c ∈ C2(R), c ≥ 0, |{ξ|c(ξ) = 0}| < ∞, and, w.l.o.g., c ≤ 1, 0 (H.2) u ∈ W1,1(Ω)∩W1,∞(Ω), u ∈ BV(Ω). . 0 0,x 1 We call (1) the “variational heat equation”, because it can be derived from 0 7 a variational principle, similar to the variational wave equation [13, 21, 11, 1 4, 5], see (3) below. : v The variational heat equation arises in the context of the continuum the- i ory for nematic liquid crystals as a model for the dynamics of the director X field. Liquid crystals are materials in a state of matter between the solid r a and the liquid state. In the case of uniaxial nematic liquid crystals, this means that the elongated molecules can move freely like in a fluid, but tend to align along the same direction like in a crystal. On a macroscopic scale such a state can be described by two vector fields, the velocity field and the so-called director field, which are governed by the Ericksen-Leslie equa- tions [22, 24, 8, 16, 17, 18, 9]. The director field is a unit vector field that gives the average direction of the molecules at each point. Date: January 6, 2017. G. M. Coclite is member of the Gruppo Nazionale per l’Analisi Matematica, la Prob- abilita` e le loro Applicazioni (GNAMPA) of the Istituto Nazionale di Alta Matematica (INdAM). 1 2 G.M.COCLITE,J.RIDDER,ANDN.H.RISEBRO To arrive at equation (1), we assume the simplified setting of a uniaxial nematic with no flow and a director field n that lies in the x − y plane and varies only in x-direction. Then the director can be described by an angle u as n = (cos(u),sin(u),0). The Oseen-Frank energy, which models the tendency of the director to align along the same direction everywhere, reduces to ˆ E = (c(u))2(u )2dx, x where (cid:112) (2) c(u) = k cos(u)2+k sin(u)2, 1 2 and k and k are the Oseen-Frank elastic constants corresponding to bend 1 2 andsplaydeformations[22,8,24,19,10]. Inaddition,thedirectorissubject to the dissipation ˆ D = κ (u )2dx, x where κ is the rotational viscosity coefficient. Together, a variational prin- ciple applied to the energy law d E = D, dt and scaling κ = 1 gives (1), see [2, 1]. A similar model is the variational wave equation [13, 21], (3) u = c(u)(c(u)u ) , tt x x which is derived in the same way from the Oseen-Frank energy, but ne- glecting dissipation and instead including inertia in the form of the kinetic energy ˆ σ(u )2dx, x where σ is the rotational inertia of the director, scaled to 1 in (3). Typical valuesfortheelasticconstantsk andk in(2)areoforder10−11–10−12, the 1 2 dissipation κ is of order 10−1–10−3, and the rotational inertia σ is of order 10−13, [22, 25]. On small length scales, the term from the elastic energy and the dissipation can be of the same order. The inertia term however is usually dominated by the dissipation [2], therefore (1) is a more suitable model than (3) in most physical settings. From a mathematical point of view, if k and k are strictly positive, i.e., 1 2 c > 0, equation (1) is a nonlinear, uniformly parabolic equation. While (3), and also the combination of (1) and (3) where both u and u are included, t tt does not possess a unique classical solution [11, 4, 5], standard theory of nonlinear parabolic equations guarantees well-posedness of (1), see [14]. Wearethereforeinterestedinthedegeneratecaseof (1)wherecisallowed to vanish at some points, i.e., if c is given by (2), in the case that k = 0 1 or k = 0. Solutions of degenerate parabolic equations are not necessarily 2 smooth or unique, therefore new concepts of solutions, e.g., weak solutions, A CONVERGENT SCHEME FOR THE VARIATIONAL HEAT EQUATION 3 entropy solutions, or viscosity solutions are required. In the case of (1), a formal calculation shows that there is no maximum principle for u , but x for c(u)u (see Section 3). At points where c(u) vanishes, this allows for x gradient blow-up. The goal of this paper is to design a convergent numerical scheme for (1). The form of the right-hand side and the resulting lack of a gradient bound suggests that one should transform (1) first. One possibility to do this is to define ˆ u 1 (4) v = k (u) = dξ. v c(ξ) Then (1) becomes (5) v = (c2(k¯ (v))v ) , t v x x where k¯ is the inverse of k (u). For this equation it is straightforward to v v obtain an L2 bound and one can also show uniqueness of weak solutions. If we assume c > 0, a simple finite difference scheme based on central differ- ences and averages in space can be shown to converge to a weak solution using Aubin-Lions lemma, see also [15, 20] for examples in a similar setting. If however c = 0 for some u, then (4) is not necessarily finite and a bound on v does not follow directly from the L2 bound. x An alternative transformation of (1) is ˆ u (6) w = k (u) = c(ξ)dξ, w so w satisfies (7) w = c2(k¯ (w))w . t w xx The transformation k and its inverse k¯ are well-defined for any c ≥ 0 if c w w vanishesonlyonsinglepoints. Itisalsopossibletoshowa priori boundsfor both w and w in L∞ and BV (functions of bounded total variation), see x Section 3. However, (7) does not guarantee uniqueness of solutions. Indeed, Ughi et al. [23, 7, 3] showed that for the special case where c2(k¯ (w)) = w, w weak solutions of (7) (defined in a standard way, see Section 2) are not unique. To choose the physically relevant solution, they define “viscosity solutions” which are obtained by taking the limit of classical solutions of the equation with c > 0 or suitable initial data. In the setting of (2), these viscosity solutions correspond to sending k or k to 0 or choosing 1 2 the solution that corresponds to a solution of (5). Ughi et al.’s concept of viscosity solutions is not generally the same as Lions’ theory of viscosity solutions for degenerate parabolic equations [6, 3]. The uniqueness theory ofthelatterisnotapplicablehere, becausetheright-handsideof (7), or(1), is not proper. The scheme that we will present in this paper discretizes (7). Based on discrete versions of the L∞ and BV bounds on w and w , we use Kol- x mogorov’s compactness theorem to show that the numerical approximations 4 G.M.COCLITE,J.RIDDER,ANDN.H.RISEBRO for both w and w converge strongly in L1(Ω). The strong convergence of x the derivative is important, because the weak formulation of (7) includes nonlinear terms in w . Passing to the limit in the definition of the scheme, x we prove that a subsequence of the numerical solutions converges to a weak solution as ∆x,∆t → 0. Our numerical experiments confirm the nonunqiueness properties dis- cussed above. If k = 0 in (2) and the grid is chosen such that c(u (x)) 1 0 is positive at every grid point, then the numerical solutions converge to Ughi et al.’s viscosity solution. This solution is the same as the one ob- tained by a method based on (5) and as the limit k → 0 of solutions of the 1 w-based scheme for any set of grid points. If however one of the grid points coincides with a zero of c(u (x)), we get another solution which corresponds 0 to a classical solution of (7), “glued together” at the zeros of c(u (x)) with 0 Dirichlet boundary conditions. Interpreted as solutions of (1), this type of solutions shows clearly that the gradient is unbounded. The rest of this paper is structured as follows: In Section 2 we will de- fine the scheme for (7), introduce the notion of weak solutions, and state our convergence result. Section 3 contains discrete a priori bounds, which are based on Harten’s lemma and motivated by formal calculations in the continuous case. Time continuity is shown in Section 4 and the convergence proof is carried out in Section 5. In Section 6 we present a series of nu- merical experiments that confirm the convergence result and highlight the nonuniqueness properties of (7). 2. A numerical scheme for w and the main result To be precise, let us restate (7) in the form that will be the basis of our scheme. Assume that (H.3) B ∈ C2(R) and 0 ≤ B ≤ 1, (H.4) w ∈ W1,1(Ω)∩W1,∞(Ω), w ∈ BV(Ω). 0 0,x Then we want to solve (cid:40) w = B(w)w , t > 0,x ∈ Ω, t xx (8) w(x,0) = w (x), x ∈ Ω, 0 on Ω = R or [0,1] with periodic boundary conditions. Equation (1) can be transformed to (8) by defining w as in (6). If u sat- 0 isfies (H.2), then w will satisfy (H.4), but not vice versa. Similarly, (H.3) 0 follows from (H.1). As an example, if we choose c according to (2) with k = 0 and k = 1, then k (u) = |sin(u)| and B(w) = 1 − w2, see also 1 2 w Section 6. To define the scheme, let Ω be discretized by the equidistant grid points x = j∆x, j = 0,...,N, and let tn = n∆t denote the time steps. If j Ω = [0,1], we set periodic boundary conditions. We will implicitly assume that all functions are periodically extended outside of the domain, so that no boundary terms occur. A CONVERGENT SCHEME FOR THE VARIATIONAL HEAT EQUATION 5 A straightforward discretization of (8) is (9) D+wn = B(wn+θ)D2wn+θ, t j j j where we used the difference quotients 1 1 D a = (a −a ), D a = (a −a ), + j j+1 j − j j j−1 ∆x ∆x 1 D+an = (an+1−an), D2a = D D a , t ∆t j + − j and the convex combination wn+θ = θwn+1+(1−θ)wn, where θ ∈ [0,1]. j j j For θ = 0, the scheme is explicit, for θ = 1, it is fully implicit, and for θ = 1 2 we have the Crank-Nicholson time discretization. In the fully implicit case of θ = 1, the scheme is unconditionally stable. Otherwise, we require that the time step ∆t and grid size ∆x satisfy the CFL condition ∆t 1 (10) λ = < . (∆x)2 2(1−θ) For the discrete derivatives zn = D wn and yn = D zn, the scheme defined j + j j − j by (9) becomes (11) D+zn = D (B(wn+θ)D zn+θ), t j + j − j (12) D+yn = D2(B(wn+θ)yn+θ). t j j j We will use these forms below to get a priori bounds on wn. j For given initial data w ∈ W2,1∩W1,∞, define the discrete initial data 0 ˆ (13) w0 = 1 xj+12 w (x)dx. j ∆x 0 xj−12 To get from the discrete approximations wn back to continuous functions, j we use the piecewise linear and piecewise constant interpolations x −x x−x (14) w∆t(x,t) = j+1 wn+ jwn , ∆x j ∆x j+1 for x ∈ [x ,x ), t ∈ [tn,tn+1), j j+1 (15) w∆t(x,t) = wn, j for x ∈ [x ,x ), t ∈ [tn,tn+1), j−1 j+1 2 2 (16) z∆t(x,t) = w∆t(x,t) = D wn = zn, x + j j for x ∈ [x ,x ), t ∈ [tn,tn+1). j j+1 Our main result is the convergence of the numerical scheme. Since B(w) is allowed to vanish, equation (8) is a degenerate parabolic equation and solutions are not necessarily smooth. In particular, the derivative of w may not be defined at every point. We will therefore prove convergence to weak solutions of (8). 6 G.M.COCLITE,J.RIDDER,ANDN.H.RISEBRO Definition2.1(Weaksolutionsof (8)). Afunctionw ∈ L∞(0,∞;H1(Ω))× L∞(Ω×(0,∞)) is a weak solution of (8) if it satisfies ˆ ˆ ˆ ∞ (17) wφ −B(w)w φ −B(cid:48)(w)(w )2φdxdt+ w (x)φ(x,0)dx = 0, t x x x 0 0 Ω Ω for all φ ∈ C∞(Ω×[0,∞)). c The convergence result, which we will prove in Section 5, reads as follows. Theorem 2.1. A subsequence of the interpolations w of the solutions of ∆t the scheme defined by (9) converges in C([0,∞),W1,1(Ω)) to a weak solution of (8) as defined in Definition 2.1. Note that only a subsequence of w converges, because weak solutions ∆t of (8) are not unique. We will comment more on this in Section 6. For the a priori bounds in the next section, we will use the discrete norms (cid:88) (cid:88) (cid:107)an(cid:107) = sup|an|, (cid:107)an(cid:107) = ∆x |an|, |an| = |an−an |, ∞ j 1 j BV j j−1 j j j 3. A priori bounds In the following, we will show discrete maximum principles and BV bounds for wn and zn = D wn. Here, note that the original equation (1) j j + j only possesses a maximum principle for u, but not for u , since in x 1 u = (c(u))2u +4c(u)c(cid:48)(u)u u + (c2(u))(cid:48)(cid:48)(u )3, tx xxx x xx x 2 the third term can lead to growth of local maxima in u . Our numerical x examples in Section 6 confirm this. One advantage of the transformation to w is that for equation (8) both w and z = w are bounded in L∞. x The BV bound for z will be important in the convergence proof, because strong convergence for both w and its first derivative is needed to pass to the limit in the third term of the weak formulation (17). Before turning to the discrete setting, let us show formally how L1 bounds for z and y = z x (i.e., BV bounds for w and z) can be obtained in the continuous case. For z, multiply z = (Bz ) t x x by η(cid:48)(z), where η is some convex smooth function, and integrate in space to get ˆ ˆ d η(z)dx = − B(w)(z )2η(cid:48)(cid:48)(z)dx ≤ 0. x dt Ω Ω Letting η → |·|, we get an L1 bound for z. For y, the formal continuous equivalent of equation (12) is (18) y = (B(w)y) . t xx A CONVERGENT SCHEME FOR THE VARIATIONAL HEAT EQUATION 7 Again, let η ∈ C2(R) be convex and multiply (18) by η(cid:48)(y). Then, η(y) = (By +2B y +B y)η(cid:48)(y) t xx x x xx ≤ (y )2Bη(cid:48)(cid:48)(y)+By η(cid:48)(y)+2B η(y) +B yη(cid:48)(y) x xx x x xx = Bη(y) +2B η(y) +B yη(cid:48)(y) xx x x xx = (Bη(y) ) +B η(y) +B yη(cid:48)(y) x x x x xx = (Bη(y) ) +(B η(y)) −B η(y)+B yη(cid:48)(y) x x x x xx xx = (Bη(y)) +B (η(cid:48)(y)y−η). xx xx Integrating over Ω and taking η(y) = |y| such that it converges to |y| as (cid:15) (cid:15) → 0, we get ˆ d |y| ≤ 0. dt Ω In the discrete case, we will base our proofs on an extended version of Harten’s Lemma [12, p. 118]. Lemma 3.1. Let v be given by j (19) v = u −A ∆ u +B ∆ u −C ∆ v +D ∆ v , j j j−1/2 − j j+1/2 + j j−1/2 − j j+1/2 + j where ∆ u = ±(u −u ). ± j j±1 j (i) If A , B , C , and D are nonnegative for all j, and j+1/2 j+1/2 j+1/2 j+1/2 A +B ≤ 1 for all j, then j+1/2 j+1/2 |v| ≤ |u| . BV BV (ii) If A , B , C , and D are nonnegative for all j, and j+1/2 j+1/2 j+1/2 j+1/2 A +B ≤ 1 for all j, then j−1/2 j+1/2 minu ≤ v ≤ maxu i j i i i Proof. From (19), we get (1+C +D )∆ v = (1−A −B )∆ u j+1/2 j+1/2 + j j+1/2 j+1/2 + j +A ∆ u +B ∆ u j−1/2 − j j+3/2 + j+1 +C ∆ v +D ∆ v . j−1/2 − j j+3/2 + j+1 Hence, under the assumptions of (i), (cid:88) (cid:88) (1+C +D )|∆ v | ≤ (1−A −B )|∆ u | j+1/2 j+1/2 + j j+1/2 j+1/2 + j j j (cid:88) + A |∆ u |+B |∆ u | j−1/2 − j j+3/2 + j+1 j (cid:88) + C |∆ v |+D |∆ v | j−1/2 − j j+3/2 + j+1 j (cid:88) = |∆ u |+(C +D )|∆ v |, + j j+1/2 j+1/2 + j j 8 G.M.COCLITE,J.RIDDER,ANDN.H.RISEBRO from which the BV bound follows. For the maximum principle, we can write (19) as (1+C +D )v = (1−A −B )u +A u +B u j−1/2 j+1/2 j j−1/2 j+1/2 j j−1/2 j−1 j+1/2 j+1 +C v +D v . j−1/2 j−1 j+1/2 j+1 Thus, if the assumptions of (ii) hold, v = max v satisfies j(cid:48) i i (1+C +D )v ≤ (1−A −B )maxu j(cid:48)−1/2 j(cid:48)+1/2 j(cid:48) j(cid:48)−1/2 j(cid:48)+1/2 i i +A maxu +B maxu j(cid:48)−1/2 i j(cid:48)+1/2 i i i +C v +D v , j(cid:48)−1/2 j(cid:48) j(cid:48)+1/2 j(cid:48) and hence, max v = v ≤ max u . Similarly, min v ≥ min u , which i i j(cid:48) i i i i i i concludes the proof. (cid:3) The L∞ and BV bound for wn and zn follow directly from the above j j lemma. Lemma 3.2. Let wn be the solution of (9) and zn = D wn. Then j j + j minw0 ≤ wn ≤ maxw0, |wn| ≤ |w0| , i j i BV BV i i minz0 ≤ zn ≤ maxz0, |zn| ≤ |z0| . i j i BV BV i i Proof. Rewriting (9), we get wn+1 = wn+(1−θ)∆tB(wn+θ)D2wn+θ∆tB(wn+θ)D2wn+1. j j j j j j To apply Harten’s lemma, set v = wn+1, u = wn, and j j j j A = (1−θ)λB(wn+θ), C = θλB(wn+θ), j−1/2 j j−1/2 j B = (1−θ)λB(wn+θ), D = θλB(wn+θ), j+1/2 j j+1/2 j where λ = ∆t/(∆x)2. Because λ satisfies the CFL condition (10) and θ and B(w) take values in [0,1], the assumptions of Harten’s lemma hold and we get the maximum and BV bound for wn. j For z, write (11) as zn+1 = zn+(1−θ)∆tD (B(wn+θ)D zn)+θ∆tD (B(wn+θ)D zn+1). j j + j − j + j − j Set v = zn+1, u = zn, and j j j j A = (1−θ)λB(wn+θ), C = θλB(wn+θ), j−1/2 j j−1/2 j B = (1−θ)λB(wn+θ), D = θλB(wn+θ), j+1/2 j+1 j+1/2 j+1 in Harten’s lemma. Again, due to the CFL condition and the bounds on B, the conditions are satisfied and the claim follows. (cid:3) A CONVERGENT SCHEME FOR THE VARIATIONAL HEAT EQUATION 9 4. Continuity in time In order to show compactness, we will need continuity in time of both w∆t and z∆t. For w∆t this follows directly from the definition of the scheme and the BV bound for z above. Lemma 4.1. Let w∆t be the interpolation (14) of the solutions wn of (9). j Then, for any t,t+τ ≥ 0, ˆ |w∆t(x,t+τ)−w∆t(x,t)|dx ≤ (τ +O(∆t))|z0| +O(∆x)|w0| . BV BV Ω Proof. Using the piecewise constant interpolation w∆t, we get ˆ ˆ |w∆t(x,t+τ)−w∆t(x,t)|dx ≤ |w∆t(x,t+τ)−w∆t(x,t+τ)| Ω Ω (20) +|w∆t(x,t)−w∆t(x,t)| +|w∆t(x,t+τ)−w∆t(x,t)|dx. Regarding the first two terms on the right-hand side, note that for t ∈ [tn,tn+1), ˆ ˆ (cid:88) xj |w∆t(x,t)−w∆t(x,t)|dx = |(x −x)D wn|dx j − j Ω j xj−ˆ12 + xj+12|(x−x )D wn|dx j + j (21) xj (∆x)2 (cid:88) = |D wn| 4 + j j ∆x ∆x = |wn| ≤ |w0| , 4 BV 4 BV where the last inequality is due to Lemma 3.2. For the last term in (20), let m, n be such that t+τ ∈ [tn,tn+1) and t ∈ [tm,tm+1). Using the BV bound on z from Lemma 3.2, we get ˆ ˆ |w∆t(x,t+τ)−w∆t(x,t)|dx = (cid:88) xj+21|wn−wm| j j Ω j xj−12 n−1 (cid:88) (cid:88) = ∆x ∆t|D+wk| t j j k=m n−1 (cid:88) (cid:88) = ∆x∆t |B(wk+θ)D zk+θ| j − j j k=m ≤ ∆t(n−m)|z0| = (τ +O(∆t))|z0| , BV BV and the claim follows. (cid:3) 10 G.M.COCLITE,J.RIDDER,ANDN.H.RISEBRO Forz∆t,wewilluseaversionofKruˇzkov’sinterpolationlemma[12,p.208, Lemma 4.11], which gives continuity in time if for all t ,t ≥ 0 and φ ∈ 1 2 C∞(B ), where B = [−r,r]∩Ω, 0 r r (22ˆ) (cid:12) (cid:12) (cid:12) ((z∆t(x,t )−z∆t(x,t ))φ(x)dx(cid:12) ≤ C (cid:107)φ(cid:48)(cid:107) (|t −t |+O(∆t)), (cid:12) 2 1 (cid:12) r L∞(Br) 2 1 Br in addition to the L∞ and BV bound from Lemma 3.2. Lemma 4.2. Let z∆t be the piecewise constant interpolation of zn = D wn, j + j where wn is the solution of (9). Then z∆t satisfies for any t,t+τ ≥ 0, r > 0, j ˆ (cid:112) ∆t |z∆t(x,t+τ)−z∆t(x,t)| ≤ C max(|z0| ,1)( |τ|+ ), r BV (cid:112) |τ| Br where B = [−r,r]∩Ω. r Proof. ToapplyKruˇzkov’sinterpolationlemma,weneedtoshow(22). First, note that for any time step n, ˆ ˆ (cid:12) (cid:12) (cid:12)(cid:88) xj+1 (cid:12) (cid:12) (z∆t(x,tn+1)−z∆t(x,tn))φdx(cid:12) = (cid:12) (zn+1−zn) φdx(cid:12) (cid:12) (cid:12) (cid:12) j j (cid:12) Ω j xj ˆ (cid:12)(cid:88) xj+1 (cid:12) = (cid:12) ∆tD (B(wn+θ)D zn+θ) φdx(cid:12) (cid:12) + j − j (cid:12) j ˆ xj (cid:12)(cid:88) 1 xj+1 (cid:12) = (cid:12) ∆tB(wn+θ)D zn+θ φ(x)−φ(x−∆x)dx(cid:12) (cid:12) j − j ∆x (cid:12) j xj (cid:88) ≤ ∆t∆x|D zn+θ|(cid:107)φ(cid:48)(cid:107) − j L∞(Ω) j ≤ ∆t(cid:107)φ(cid:48)(cid:107) |z0| . L∞(Ω) BV For given t ,t > 0, let n,m be such that t ∈ [tn,tn+1) and t ∈ [tm,tm+1). 1 2 1 2 The above estimate yields ˆ (cid:12) (cid:12) (cid:12) (z∆t(x,t )−z∆t(x,t ))φdx(cid:12) ≤ (cid:107)φ(cid:48)(cid:107) |z0| (tm−tn) (cid:12) 2 1 (cid:12) L∞(Ω) BV Ω ≤ (cid:107)φ(cid:48)(cid:107) |z0| (t −t +2∆t). L∞(Ω) BV 2 1 Kruˇzkov’s interpolation lemma [12, p. 208, Lemma 4.11] then implies ˆ |τ|+2∆t |z∆t(x,t+τ)−z∆tx,t|dx ≤ C ((cid:15)+(cid:15)|z0| +|z0| ), r BV BV (cid:15) Br for any (cid:15) > 0. Choosing (cid:15) = (cid:112)|τ|, we arrive at the claim. (cid:3) 5. Convergence Finally, we are able to prove the convergence of the scheme, Theorem 2.1.

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.