Abstract

We consider a nonlinear 4th-order degenerate parabolic partial differential equation that arises in modelling the dynamics of an incompressible thin liquid film on the outer surface of a rotating horizontal cylinder in the presence of gravity. The parameters involved determine a rich variety of qualitatively different flows. We obtain sufficient conditions for finite speed of support propagation and for waiting time phenomena by application of a new extension of Stampacchia's lemma for a system of functional equations.

1. Introduction

The time evolution of thickness of a viscous liquid film spreading over a solid surface under the action of the surface tension and gravity can be described by lubrication models [1–5]. These models approximate the full Navier-Stokes system that describes the motion of the liquid flow. Thin films play an increasingly important role in a wide range of applications, for example, packaging, barriers, membranes, sensors, semiconductor devices, and medical implants [6–8].

In this paper we consider the dynamics of a viscous incompressible thin fluid film on the outer surface of a horizontal circular cylinder that is rotating around its axis in the presence of a gravitational field. The motion of the liquid film is governed by four physical effects: viscosity, gravity, surface tension, and centrifugal forces. These are reflected in the parameters: 𝑅: the radius of the cylinder, πœ”: its rate of rotation (assumed constant), 𝑔: the acceleration due to gravity, 𝜈: the kinematic viscosity, 𝜌: the fluid's density, and 𝜎: the surface tension. These parameters yield three independent dimensionless numbers: the Reynolds number Re=(𝑅2πœ”)/𝜈, 𝛾=𝑔/(π‘…πœ”2), and the Weber number We=(πœŒπ‘…3πœ”2)/𝜎. The understanding of coating flow dynamics is important for industrial printing process where rotating cylinder transports the coating material in the form of liquid paint. The rotating thin fluid film can exhibit variety of different behaviour including: interesting pattern formations (β€œshark teeth” and β€œduck bill” patterns), fluid curtains, hydroplaning drops, and frontal avalanches [8–10]. As a result, the coating flow has been the subject of continuous study since the pioneering model was derived in 1977 by Moffatt (see [11]): πœ•β„Ž+πœ•πœ•π‘‘ξ‚ƒ1πœ•πœƒπœ”β„Žβˆ’3π‘”β„Žπœˆπ‘…3ξ‚„cosπœƒ=0.(1.1) The surface tension and inertial effects were neglected in (1.1). Here β„Ž(π‘₯,𝑑) is the thickness of the fluid film, πœƒ is a rotation angle, and 𝑑 is a time variable. The linear stability of rigidly rotating films on a rotating circular cylinder under three-dimensional disturbances was examined in [12, 13]. It was shown that the most unstable mode for thin film flows on the surface of a cylinder is the purely axial one that leads to so-called β€œring instabilities”. During the past decade, coating and rimming problems attracted many researchers who analyzed different types of flow regime asymptotically [14–17] and numerically [18–20]. For a detailed review of a growing literature on different types of thin film flows please see [21] and references there in.

The coating flow is generated by viscous forces due to cylinder's surface motion relative to the fluid. There is no temperature gradient, hence the interface does not experience a shear stress. If the cylinder is fully coated there is only one free boundary where the liquid meets the surrounding air. Otherwise, there is also a free boundary (or contact line) where the air and liquid meet the cylinder's surface.

The asymptotic evolution equation for the thickness of the fluid film with the surface tension effect: πœ•β„Ž+πœ•πœ•π‘‘ξ‚Έ1πœ•πœƒπœ”β„Žβˆ’3π‘”β„Žπœˆπ‘…31cosπœƒ+3πœŽπœŒπ‘…4πœˆβ„Ž3ξ‚΅πœ•β„Ž+πœ•πœ•πœƒ3β„Žπœ•πœƒ3ξ‚Άξ‚Ή=0,(1.2) was derived by Pukhnachev [22] in 1977. It is valid under the assumptions that the fluid film is thin β„Žβ‰ͺ𝑅 and its slope is small (1/𝑅)(πœ•β„Ž/πœ•πœƒ)β‰ͺ1. Later in 2009, taking into account inertial effects, Kelmanson [23] presented a more general model: πœ•β„Ž+πœ•πœ•π‘‘ξ‚Έ1πœ•πœƒπœ”β„Žβˆ’3π‘”β„Žπœˆπ‘…31cosπœƒ+3πœŽπœŒπ‘…4πœˆβ„Ž3ξ‚΅πœ•β„Ž+πœ•πœ•πœƒ3β„Žπœ•πœƒ3ξ‚Ά+13πœ”2πœŒβ„Žπœˆπ‘…3πœ•β„Žξ‚Ήπœ•πœƒ=0.(1.3) He analyzed, asymptotically and numerically, diverse effects of inertia in both small- and large-surface-tension limits.

We should mention that all three lubrication approximation models described above were based on the assumption of the no-slip boundary condition. It is well known [24] that the combination of constant viscosity and no-slip boundary conditions at the liquid-solid interface yields a logarithmic divergence in the rate of dissipation at moving contact line, that is, an infinite energy is needed to make the droplet expand. The most common way to overcome this difficulty is to introduce effective slip conditions (see (2.1)) that indeed removes the force singularity at advancing contact lines (see [25]).

The main goal of our paper is to study waiting time phenomenon for the coating flows under an assumption of effective slip conditions, that is, we analyze (2.1) that is a modified version of (1.3). Our approach is based on now well-established nonlinear PDE analysis for degenerate higher order parabolic equations.

The sufficient conditions: β„Ž0(π‘₯)≀𝐴|π‘₯|4/𝑛 for 0<𝑛<2, |β„Ž0π‘₯(π‘₯)|≀𝐡|π‘₯|4/π‘›βˆ’1 for 2≀𝑛<3, (where 𝐴 and 𝐡 are some positive constants) on nonnegative initial data, β„Ž0 for the occurrence of waiting time phenomena were derived by Dal Passo et al. [26] for the classic thin film equation: β„Žπ‘‘+ξ€·||β„Ž||π‘›β„Žπ‘₯π‘₯π‘₯ξ€Έπ‘₯=0.(1.4) These results were based on an energy method developed in [27] for quasilinear parabolic equations. To the best of our knowledge, there is only one publication [28], where the waiting time phenomenon in the classic thin film equation (1.4) was discovered for β„Ž0(π‘₯)∼|π‘₯|𝛼 for 2<𝛼<4/𝑛. The result was obtained by means of matching asymptotic methods and was supported by numerous numerical simulations. For more general nonlinear degenerate parabolic equations with nonlinear lower order terms the waiting time phenomenon was analyzed in [29–31].

It is well known [32] that the similarity solutions of the second order nonlinear parabolic equation: 𝑐𝑑=ξ€·π‘π‘šπ‘π‘₯ξ€Έπ‘₯,π‘š>0,(1.5) subject to prescribing appropriate initial data, demonstrate the existence of a waiting-time phenomena before the free boundary moves. The comparison theorem, that is not applicable in our case, then enabled a number of results to be obtained about the existence and length of waiting times for general initial data. Our approach is completely different and based on local entropy/energy functional estimates.

We also analyze speed of support propagation and obtain an upper bound on it for the modified version of (1.3) (see (2.1)). The first finite speed results for nonnegative generalized solutions of the classic thin film equation (1.4) were obtained in [33, 34] for the case 0<𝑛<2 and 2≀𝑛<3, respectively. For more general types of thin film equations the finite speed of support propagation phenomenon was studied in [35–39] (see also references there in).

The outline of our paper is as follows. We first prove for 𝑛>0 the long-time existence of a generalized weak solution and then prove that it can have an additional regularity in Section 2. In Sections 3 and 4 we show finite speed support propagation in the β€œslow” convection case (𝑛>1): for 1<𝑛<3 and waiting time phenomena for 1<𝑛<2, accordingly. The general strategy is to use an extension of Stampacchia's lemma for a system of functional equations (see Lemma  3.1 [26], where this extension is proved for a single equation and Lemma  A.2 in [37], where this extension is proved for systems in the homogeneous case). This result to our knowledge is new and might be of independent interest. We leave as an open problem the β€œfast” convection case (0<𝑛<1): finite speed of support propagation and sufficient conditions for waiting time phenomenon.

2. Existence and Regularity of Solutions

We are interested in the existence of nonnegative generalized weak solutions to the following initial-boundary value problem: ⎧βŽͺ⎨βŽͺβŽ©β„Ž(𝑃)𝑑+ξ€·π‘“ξ€·π‘Ž(β„Ž)0β„Žπ‘₯π‘₯π‘₯+π‘Ž1β„Žπ‘₯+𝑀π‘₯ξ€Έξ€Έπ‘₯=0in𝑄𝑇,πœ•π‘–β„Žπœ•π‘₯π‘–πœ•(βˆ’π‘Ž,𝑑)=π‘–β„Žπœ•π‘₯𝑖(π‘Ž,𝑑)for𝑑>0,𝑖=0,3,β„Ž(π‘₯,0)=β„Ž0(π‘₯)β©Ύ0,(2.1) where 𝑓(β„Ž)=|β„Ž|𝑛,β€‰β€‰β„Ž=β„Ž(π‘₯,𝑑),  Ω=(βˆ’π‘Ž,π‘Ž),  𝑄𝑇=(0,𝑇)Γ—Ξ©,  𝑛>0,β€‰β€‰π‘Ž0>0,β€‰β€‰π‘Ž1β‰₯0, and 𝑀(π‘₯,𝑑) such that 𝑀(π‘₯,β‹…)βˆˆπ‘Š1∞(0,𝑇)fora.e.π‘₯∈Ω,𝑀(β‹…,𝑑)βˆˆπ‘Š2∞[].(Ξ©)fora.e.π‘‘βˆˆ0,𝑇(2.2) ote that (1.3) is a particular case of (2.1) that corresponds to 𝑛=3 and 𝑀(π‘₯,𝑑)=cos(π‘₯βˆ’πœ”π‘‘).

We consider a generalized weak solution in the following sense [40, 41].

Definition 2.1. A generalized weak solution of problem (𝑃) is a nonnegative function β„Ž satisfying β„ŽβˆˆπΆ1/2,1/8π‘₯,π‘‘ξ‚€π‘„π‘‡ξ‚βˆ©πΏβˆžξ€·0,𝑇;𝐻1ξ€Έ(Ξ©),β„Žπ‘‘βˆˆπΏ2𝐻0,𝑇;1ξ€Έ(Ξ©)ξ…žξ‚,β„ŽβˆˆπΆ4,1π‘₯,π‘‘βˆš(𝒫),ξ€·π‘Žπ‘“(β„Ž)0β„Žπ‘₯π‘₯π‘₯+π‘Ž1β„Žπ‘₯+𝑀π‘₯ξ€ΈβˆˆπΏ2(𝒫),(2.3) where π’«βˆΆ={β„Ž>0}. The solution β„Ž satisfies (2.1) in the following sense: ξ€œπ‘‡0βŸ¨β„Žπ‘‘ξ€(β‹…,𝑑),πœ™βŸ©π‘‘π‘‘βˆ’π’«ξ€·π‘Žπ‘“(β„Ž)0β„Žπ‘₯π‘₯π‘₯+π‘Ž1β„Žπ‘₯+𝑀π‘₯ξ€Έπœ™π‘₯𝑑π‘₯𝑑𝑑=0,(2.4) for all πœ™βˆˆπΆ1(𝑄𝑇)∩𝐢(𝑄𝑇) with πœ™(βˆ’π‘Ž,β‹…)=πœ™(π‘Ž,β‹…); β„Ž(β‹…,𝑑)β†’β„Ž(β‹…,0)=β„Ž0pointwise&stronglyin𝐿2[],πœ•(Ξ©)as𝑑→0,(2.5)β„Ž(βˆ’π‘Ž,𝑑)=β„Ž(π‘Ž,𝑑)βˆ€π‘‘βˆˆ0,π‘‡π‘–β„Žπœ•π‘₯π‘–πœ•(βˆ’π‘Ž,𝑑)=π‘–β„Žπœ•π‘₯𝑖(π‘Ž,𝑑),(2.6) for 𝑖=1,2,3 at all points of the lateral boundary where β„Žβ‰ 0.

Because the second term of (2.4) has an integral over {β„Ž>0} rather than over 𝑄𝑇, the generalized weak solution is β€œweaker” than a standard weak solution. Here, {β„Ž>0} is short hand for {(π‘₯,𝑑)βˆˆπ‘„π‘‡βˆΆβ„Ž(π‘₯,𝑑)>0}. This short hand is used throughout: the time interval included in {β„Ž>0} is to be inferred from the context it appears in.

A key object for proving additional properties of a generalized weak solution is an integral quantity introduced by Bernis and Friedman [42]: the β€œentropy” ∫𝐺0(β„Ž(π‘₯,𝑑))𝑑π‘₯. The function 𝐺0(𝑧) is defined by 𝐺0⎧βŽͺβŽͺ⎨βŽͺβŽͺβŽ©π‘§(𝑧)∢=2βˆ’π‘›π‘§(2βˆ’π‘›)(1βˆ’π‘›)+𝑑𝑧+𝑐if𝑛≠1,2,𝑧lnπ‘§βˆ’π‘§+𝑒if𝑛=1,βˆ’ln𝑧+𝑒+1if𝑛=2,(2.7) where ξ‚»βŽ§βŽͺ⎨βŽͺβŽ©π‘‘=1if1<𝑛<2,0otherwise,𝑐=(π‘›βˆ’1)1/(1βˆ’π‘›)2βˆ’π‘›if1<𝑛<2,0otherwise.(2.8) By construction, 𝐺0 is a nonnegative convex function on [0,∞). For 1≀𝑛≀2, the linear part of 𝐺0 is chosen to ensure that 𝐺0 has a positive lower bound on [0,∞). Also in the statement of Theorem 2.2 we use an β€œπ›Ό-entropy”, ∫𝐺0(𝛼)(β„Ž(π‘₯,𝑑))𝑑π‘₯, where 𝐺0(𝛼)⎧βŽͺβŽͺ⎨βŽͺβŽͺβŽ©π‘§(𝑧)∢=𝑧lnπ‘§βˆ’π‘§+𝑒if𝛼=π‘›βˆ’1,βˆ’ln𝑧+𝑒𝑧+1if𝛼=π‘›βˆ’2,2βˆ’π‘›+π›Όξ‚»βŽ§βŽͺ⎨βŽͺ⎩(2βˆ’π‘›+𝛼)(1βˆ’π‘›+𝛼)+𝑑𝑧+𝑐otherwise,(2.9)𝑑=1ifπ›Όβˆˆ(π‘›βˆ’2,π‘›βˆ’1),0otherwise,𝑐=(π‘›βˆ’1βˆ’π›Ό)1/(1+π›Όβˆ’π‘›)2+π›Όβˆ’π‘›ifπ›Όβˆˆ(π‘›βˆ’2,π‘›βˆ’1),0otherwise.(2.10)𝐺0(𝛼) is a nonnegative convex function on [0,∞). The linear part of 𝐺0(𝛼) is chosen to ensure that 𝐺0(𝛼) has a positive lower bound on [0,∞) if π‘›βˆ’2β‰€π›Όβ‰€π‘›βˆ’1. If 𝛼=0, the 𝛼-entropy is the same as the entropy (2.7).

Theorem 2.2. (a) (Existence). Let 𝑛>0 and the nonnegative initial data β„Ž0∈𝐻1(Ξ©), β„Ž0(βˆ’π‘Ž)=β„Ž0(π‘Ž) satisfy ξ€œΞ©πΊ0ξ€·β„Ž0𝑑π‘₯<∞.(2.11) Then for any time 0<𝑇<∞ there exists a nonnegative generalized weak solution, β„Ž, on 𝑄𝑇 in the sense of the Definition 2.1. Furthermore, β„ŽβˆˆπΏ2ξ€·0,𝑇;𝐻2ξ€Έ.(Ξ©)(2.12) Let β„°0(1𝑇)∢=2ξ€œΞ©ξ€·π‘Ž0β„Ž2π‘₯(π‘₯,𝑇)βˆ’π‘Ž1β„Ž2(ξ€Έπ‘₯,𝑇)βˆ’2𝑀(π‘₯,𝑇)β„Ž(π‘₯,𝑇)𝑑π‘₯(2.13) then the weak solution satisfies β„°0(𝑇)+{β„Ž>0}β„Žπ‘›ξ€·π‘Ž0β„Žπ‘₯π‘₯π‘₯+π‘Ž1β„Žπ‘₯+𝑀π‘₯ξ€Έ2𝑑π‘₯𝑑𝑑≀ℰ0(0)βˆ’π‘„π‘‡β„Žπ‘€π‘‘π‘‘π‘₯𝑑𝑑.(2.14) (b) (Regularity). If the initial data also satisfies ξ€œΞ©πΊ0(𝛼)ξ€·β„Ž0𝑑π‘₯<∞,(2.15) for some βˆ’1/2<𝛼<1,  𝛼≠0 then the nonnegative generalized weak solution has the extra regularity β„Ž(𝛼+2)/2∈𝐿2(0,𝑇;𝐻2(Ξ©)) and β„Ž(𝛼+2)/4∈𝐿4(0,𝑇;π‘Š14(Ξ©)).

The theorem above was proved earlier in [41] for the case 𝑛=3 only. We note that the analogue of Theorem 4.2 in [42] also holds: there exists a nonnegative weak solution with the integral formulation ξ€œπ‘‡0βŸ¨β„Žπ‘‘(β‹…,𝑑),πœ™βŸ©π‘‘π‘‘+π‘Ž0ξ€π‘„π‘‡ξ€·π‘›β„Žπ‘›βˆ’1β„Žπ‘₯β„Žπ‘₯π‘₯πœ™π‘₯+β„Žπ‘›β„Žπ‘₯π‘₯πœ™π‘₯π‘₯𝑑π‘₯π‘‘π‘‘βˆ’π‘„π‘‡β„Žπ‘›ξ€·π‘Ž1β„Žπ‘₯+𝑀π‘₯ξ€Έπœ™π‘₯𝑑π‘₯𝑑𝑑=0.(2.16) If initial data satisfy finite 𝛼-entropy condition, that is, ∫𝐺0(𝛼)(β„Ž0)𝑑π‘₯<∞ then one can prove existence of nonnegative solutions with some additional regularity properties and use an integral formulation [43] to define them that is similar to that of (2.16) except that the second integral is replaced by the results of one more integration by parts (there are no β„Žπ‘₯π‘₯π‘₯ terms). It is worth to mention that for the case 0<𝑛<2 the finite entropy assumption in Theorem 2.2 can be omitted because it does not impose any restriction on nonnegative initial data. One needs to impose finite entropy and finite 𝛼-entropy conditions on initial data if 𝑛β‰₯2 only.

2.1. Regularized Problem

Given 𝛿,β€‰β€‰πœ€>0, a regularized parabolic problem, similar to one that was studied by Bernis and Friedman [42] can be written as:

(𝑃𝛿,πœ–)β„Žπ‘‘+ξ€·π‘“π›Ώπœ€ξ€·π‘Ž(β„Ž)0β„Žπ‘₯π‘₯π‘₯+π‘Ž1β„Žπ‘₯+𝑀π‘₯ξ€Έξ€Έπ‘₯πœ•=0,(2.17)π‘–β„Žπœ•π‘₯π‘–πœ•(βˆ’π‘Ž,𝑑)=π‘–β„Žπœ•π‘₯𝑖(π‘Ž,𝑑)for𝑑>0,𝑖=0,3,(2.18)β„Ž(π‘₯,0)=β„Ž0,πœ€(π‘₯),(2.19) where π‘“π›Ώπœ€(𝑧)∢=π‘“πœ€(𝑧)+𝛿=|𝑧|4|𝑧|4βˆ’π‘›+πœ€+π›Ώβˆ€π‘§βˆˆβ„1,𝛿>0,πœ€>0.(2.20) The 𝛿>0 in (2.20) makes the problem (2.17) regular (i.e., uniformly parabolic). The parameter πœ€ is an approximating parameter which has the effect of increasing the degeneracy from 𝑓(β„Ž)∼|β„Ž|𝑛 to π‘“πœ€(β„Ž)βˆΌβ„Ž4. The nonnegative initial data, β„Ž0, is approximated via β„Ž0+πœ€πœƒβ©½β„Ž0,πœ€βˆˆπΆ4+𝛾2(Ξ©)forsome0<πœƒ<5,πœ•π‘–β„Ž0,πœ€πœ•π‘₯π‘–πœ•(βˆ’π‘Ž)=π‘–β„Ž0,πœ€πœ•π‘₯𝑖(π‘Ž)for𝑖=β„Ž0,3,0,πœ€β†’β„Ž0stronglyin𝐻1(Ξ©)asπœ€β†’0.(2.21) The πœ€ term in (2.21) β€œlifts” the initial data so that they are smoothing from 𝐻1(Ξ©) to 𝐢4+𝛾(Ξ©). By EΔ­del’man [44, Theorem 6.3, p.302], the regularized problem has a unique classical solution β„Žπ›Ώπœ€βˆˆπΆ4+𝛾,1+𝛾/4π‘₯,𝑑(Ω×[0,πœπ›Ώπœ€]) for some time πœπ›Ώπœ€>0. For any fixed value of 𝛿 and πœ€, by EΔ­del’man [44, Theorem 9.3, p.316] if one can prove a uniform in time a priori bound |β„Žπ›Ώπœ€(π‘₯,𝑑)|β‰€π΄π›Ώπœ€<∞ for some longer time interval [0,𝑇loc,π›Ώπœ€]  (𝑇loc,π›Ώπœ€>πœπ›Ώπœ€) and for all π‘₯∈Ω then Schauder-type interior estimates [44, Corollary 2, p.213] imply that the solution β„Žπ›Ώπœ€ can be continued in time to be in 𝐢4+𝛾,1+𝛾/4π‘₯,𝑑(Ω×[0,𝑇loc,π›Ώπœ€]).

Although the solution β„Žπ›Ώπœ€ is initially positive, there is no guarantee that it will remain nonnegative. The goal is to take 𝛿→0,β€‰β€‰πœ€β†’0 in such a way that (1)  𝑇loc,π›Ώπœ€β†’π‘‡loc>0, (2) the solutions β„Žπ›Ώπœ€ converge to a (nonnegative) limit, β„Ž, which is a generalized weak solution, and (3)β€‰β€‰β„Ž inherits certain a priori bounds. This is done by proving various a priori estimates for β„Žπ›Ώπœ€ that are uniform in 𝛿 and πœ€ and hold on a time interval [0,𝑇loc] that is independent of 𝛿 and πœ€. As a result, {β„Žπ›Ώπœ€} will be a uniformly bounded and equicontinuous (in the 𝐢1/2,1/8π‘₯,𝑑 norm) family of functions in Ω×[0,𝑇loc]. Taking 𝛿→0 will result in a family of functions {β„Žπœ€} that are classical, positive, unique solutions to the regularized problem with 𝛿=0. Taking πœ€β†’0 will then result in the desired generalized weak solution β„Ž. This last step is where the possibility of nonunique weak solutions arise; see [40] for simple examples of how such constructions applied to β„Žπ‘‘=βˆ’(|β„Ž|π‘›β„Žπ‘₯π‘₯π‘₯)π‘₯ can result in two different solutions arising from the same initial data.

2.2. A Priori Estimates

Our first task is to derive a priori estimates for classical solutions of (2.17)–(2.21). The lemmas given in this section are proved in the Section 4.

We use an integral quantity based on a function πΊπ›Ώπœ€ chosen such that πΊξ…žξ…žπ›Ώπœ€1(𝑧)=π‘“π›Ώπœ€(𝑧),πΊπ›Ώπœ€(𝑧)β©Ύ0.(2.22) This is analogous to the β€œentropy” function first introduced by Bernis and Friedman [42].

Lemma 2.3. Let β„Ž0πœ€ satisfy (2.21) and be built from a nonnegative function β„Ž0 that satisfies the hypotheses of Theorem 2.2. Then there exist 𝛿0>0, πœ€0>0 and time 𝑇loc>0 such that if π›Ώβˆˆ[0,𝛿0), πœ€βˆˆ[0,πœ€0), and β„Žπ›Ώπœ€ is a solution of the problem (2.17)–(2.21) with initial data β„Ž0πœ€, then for any π‘‡βˆˆ[0,𝑇loc] the following inequalities: ξ€œΞ©ξ‚»β„Ž2π›Ώπœ€,π‘₯(π‘₯,𝑇)+2𝑐1π‘Ž0πΊπ›Ώπœ€ξ€·β„Žπ›Ώπœ€ξ€Έξ‚Ό(π‘₯,𝑇)𝑑π‘₯+π‘Ž0ξ€π‘„π‘‡π‘“π›Ώπœ€ξ€·β„Žπ›Ώπœ€ξ€Έβ„Ž2π›Ώπœ€,π‘₯π‘₯π‘₯𝑑π‘₯𝑑𝑑≀𝐾1ξ€œ<∞,(2.23)Ξ©πΊπ›Ώπœ€ξ€·β„Žπ›Ώπœ€ξ€Έ(π‘₯,𝑇)𝑑π‘₯+π‘Ž0ξ€π‘„π‘‡β„Ž2π›Ώπœ€,π‘₯π‘₯𝑑π‘₯𝑑𝑑≀𝐾2<∞(2.24) hold. The energy β„°π›Ώπœ€(𝑑) (see (2.13)) satisfies β„°π›Ώπœ€(𝑇)+π‘„π‘‡π‘“π›Ώπœ€ξ€·β„Žπ›Ώπœ€π‘Žξ€Έξ€·0β„Žπ›Ώπœ€,π‘₯π‘₯π‘₯+π‘Ž1β„Žπ›Ώπœ€,π‘₯+𝑀π‘₯ξ€Έ2𝑑π‘₯𝑑𝑑=β„°π›Ώπœ€(0)βˆ’π‘„π‘‡β„Žπ›Ώπœ€π‘€π‘‘π‘‘π‘₯𝑑𝑑.(2.25) The time 𝑇loc and the constants 𝐾𝑖 are independent of 𝛿 and πœ€.

The proof of existence of 𝛿0,β€‰β€‰πœ€0,  𝑇loc,  𝐾1, and 𝐾2 is constructive; how to find them and what quantities determine them are shown with details in Section 4.

Lemma 2.3 yields uniform-in-𝛿-and-πœ€ bounds for βˆ«β„Ž2π›Ώπœ€,π‘₯, βˆ«πΊπ›Ώπœ€(β„Žπ›Ώπœ€), βˆ¬β„Ž2π›Ώπœ€,π‘₯π‘₯, and βˆ¬π‘“π›Ώπœ€(β„Žπ›Ώπœ€)β„Ž2π›Ώπœ€,π‘₯π‘₯π‘₯. However, these bounds are found in a different manner than in earlier work for the equation β„Žπ‘‘=βˆ’(|β„Ž|π‘›β„Žπ‘₯π‘₯π‘₯)π‘₯, for example. Although the inequality (2.24) is unchanged, the inequality (2.23) has an extra term involving πΊπ›Ώπœ€. In the proof, this term was introduced to control additional, lower-order terms. This idea of a β€œblended” β€–β„Žπ‘₯β€–2-entropy bound was first introduced by Shishkov and Taranets for long-wave stable thin film equations with convection [30].

The final a priori bounds for positive, classical solutions use the following functions, parameterized by 𝛼 for π›Όβˆ‰{2,3}, πΊπœ€(𝛼)(𝑧)=𝐺0(𝛼)𝑧(𝑧)+πœ€π›Όβˆ’2⇒𝐺(π›Όβˆ’3)(π›Όβˆ’2)πœ€(𝛼)(𝑧)ξ…žξ…ž=π‘§π›Όπ‘“πœ€,(𝑧)(2.26) where 𝐺0(𝛼) is given by (2.9). In the following lemma, we restrict ourselves to the case π›Όβˆˆ[βˆ’1/2,1]; note that πΊπœ€(𝛼)(𝑧)β‰₯0 for such 𝛼.

Lemma 2.4. Assume πœ€0 and 𝑇loc are from Lemma 2.3, 𝛿=0, and πœ€βˆˆ[0,πœ€0). Assume π›Όβˆˆ[βˆ’1/2,1] and that β„Žπœ€ is a positive, classical solution of the problem (2.17)–(2.21) with initial data β„Ž0,πœ€ satisfying Lemma 2.3. If the initial data β„Ž0,πœ€ is built from β„Ž0 which also satisfies ξ€œΞ©πΊ0(𝛼)ξ€·β„Ž0(ξ€Έπ‘₯)𝑑π‘₯<∞(2.27) then there exists 𝐾4 such that ξ€œΞ©ξ‚†β„Ž2πœ€,π‘₯(π‘₯,𝑇)+πΊπœ€(𝛼)ξ€·β„Žπœ€(π‘₯,𝑇)𝑑π‘₯+π‘„π‘‡ξ€Ίπ›½β„Žπ›Όπœ€β„Ž2πœ€,π‘₯π‘₯+π›Ύβ„Žπœ€π›Όβˆ’2β„Ž4πœ€,π‘₯𝑑π‘₯𝑑𝑑≀𝐾4<∞(2.28) holds for all π‘‡βˆˆ[0,𝑇loc] and 𝐾4 is independent of πœ€ and is determined by 𝛼,β€‰β€‰πœ€0,β€‰β€‰π‘Ž0,β€‰β€‰π‘Ž1,  𝑀π‘₯,  Ω and β„Ž0. Here ξƒ―π‘Žπ›½=0π‘Žif0≀𝛼≀1,01+2𝛼14(1βˆ’π›Ό)ifβˆ’2⎧βŽͺ⎨βŽͺβŽ©π‘Žβ‰€π›Ό<0,𝛾=0𝛼(1βˆ’π›Ό)6π‘Žif0≀𝛼≀1,0(1+2𝛼)(1βˆ’π›Ό)136fβˆ’2≀𝛼<0.(2.29) Furthermore, if π›Όβˆˆ(βˆ’1/2,1)⧡{0} then ξ‚†β„Žπœ€(𝛼+2)/2ξ‚‡πœ€βˆˆ(0,πœ€0)βŠ‚πΏ2ξ€·0,𝑇loc;𝐻2ξ€Έ,ξ‚†β„Ž(Ξ©)πœ€(𝛼+2)/4ξ‚‡πœ€βˆˆ(0,πœ€0)βŠ‚πΏ4ξ€·0,𝑇loc;π‘Š1,4ξ€Έ(Ξ©)(2.30) are uniformly bounded.

The 𝛼-entropy, ∫𝐺0(𝛼)(β„Ž)𝑑π‘₯, was first introduced for 𝛼=βˆ’1/2 in [45] and an a priori bound like that of Lemma 2.4 and regularity results like those of Theorem 2.2 were found simultaneously and independently in [40, 43].

The proof of existence starts from a construction of a classical solution β„Žπ›Ώπœ€ on [0,𝑇loc] that satisfies the hypotheses of Lemma 2.3 if π›Ώβˆˆ(0,𝛿0) and πœ€βˆˆ(0,πœ€0). Taking the regularizing parameter, 𝛿, to zero, one proves that there is a limit β„Žπœ€ and that β„Žπœ€ is a generalized weak solution. After that additional nonlinear estimates are required to analyze properties of the limit β„Žπœ€; specifically to show that it is strictly positive, classical, and unique. Hence, the a priori bounds given by Lemmas 2.3 and 2.4 are applicable to β„Žπœ€. This allows us to take the approximating parameter, πœ€, to zero and to construct the desired nonnegative generalized weak solution of Theorems 2.2 (see, e.g., [41]).

2.3. Long-Time Existence of Solutions

Lemma 2.5. Let β„Ž be a generalized solution of Theorem 2.2. Then π‘Ž04β€–β€–β„Žξ€·β‹…,𝑇locξ€Έβ€–β€–2𝐻1(Ξ©)≀ℰ0(0)+𝐾5+𝐾6𝑇loc,(2.31) where β„°0(0) is defined in (2.13), βˆ«β„Žπ‘€=0, and 𝐾5=β€–π‘€β€–βˆž2βˆšπ‘€+63ξ€·π‘Ž0+π‘Ž1ξ€Έ3/2π‘Ž0𝑀2+π‘Ž0+π‘Ž12𝑀2||Ξ©||,𝐾6=β€–β€–π‘€π‘‘β€–β€–βˆžπ‘€.(2.32)

Proof of Lemma 2.5. By (2.13), π‘Ž02ξ€œΞ©β„Ž2π‘₯(π‘₯,𝑇)𝑑π‘₯≀ℰ0(π‘Žπ‘‡)+12ξ€œΞ©β„Ž2(ξ€œπ‘₯,𝑇)𝑑π‘₯+Ξ©ξ€β„Ž(π‘₯,𝑇)𝑀(π‘₯,𝑇)𝑑π‘₯βˆ’π‘„π‘‡β„Žπ‘€π‘‘π‘‘π‘₯𝑑𝑑.(2.33) The linear-in-time bound (2.14) on β„°0(𝑇loc) then implies π‘Ž02β€–β€–β„Žξ€·β‹…,𝑇locξ€Έβ€–β€–2𝐻1≀ℰ0(π‘Ž0)+0+π‘Ž12ξ€œΞ©β„Ž2𝑑π‘₯+β€–π‘€β€–βˆž+β€–β€–π‘€π‘‘β€–β€–βˆžπ‘‡ξ€Έπ‘€.(2.34) Using the estimate (see [41, Lemma 4.1, page 1837]) β€–β„Žβ€–2𝐿2(Ξ©)≀62/3𝑀4/3ξ‚΅ξ€œΞ©β„Ž2π‘₯𝑑π‘₯1/3+𝑀2||Ξ©||,(2.35) and Young's inequality: π‘Ž0+π‘Ž12ξ€œΞ©β„Ž2π‘Žπ‘‘π‘₯≀0+π‘Ž1262/3𝑀4/3ξ‚΅ξ€œΞ©β„Ž2π‘₯𝑑π‘₯1/3+𝑀2||Ξ©||ξƒͺβ‰€π‘Ž04ξ€œΞ©β„Ž2π‘₯ξ€·π‘₯,𝑇locξ€Έ2βˆšπ‘‘π‘₯+63ξ€·π‘Ž0+π‘Ž1ξ€Έ3/2βˆšπ‘Ž0𝑀2+π‘Ž0+π‘Ž12𝑀2||Ξ©||.(2.36) Using this in (2.34), the desired bound (2.31) follows immediately.

This 𝐻1-estimate will be used to extend the short-time existence of a solution to the long-time existence result of Theorem 2.2 (see [41, Proof of Theorem 3, page 1838]).

3. Finite Speed of Support Propagation

Theorem 3.1. Let 1<𝑛<3. Assume β„Ž0 is nonnegative, β„Ž0∈𝐻1(Ξ©) and suppβ„Ž0βŠ‚(βˆ’π‘Ÿ0,π‘Ÿ0)⋐Ω. Then the solution β„Ž of Theorem 2.2 has finite speed of support propagation, that is, there exists a continuous nondecreasing function Ξ“(𝑇),  Γ(0)=0 such that suppβ„Ž(𝑇,β‹…)βŠ‚(βˆ’π‘Ÿ0βˆ’Ξ“(𝑇),π‘Ÿ0+Ξ“(𝑇))⋐Ω for all 𝑇≀𝑇0∢=Ξ“βˆ’1(π‘Žβˆ’π‘Ÿ0).

In the following theorem, we find the explicit upper bounds of the Ξ“(𝑇) for a solution of the corresponding Cauchy problem with a compactly supported nonnegative initial data β„Ž0∈𝐻1(ℝ1). Note that the definition of generalized weak solution of the Cauchy problem is as Definition 2.1 except that Ξ© is replaced by ℝ1 and the relation (2.6) is dropped. Using Lemma 2.5, we can show that the upper estimate of Ξ“(𝑇) from Theorem 3.1 is independent on Ξ© therefore the solution from Theorem 2.2 can be extended to be identically zero for |π‘₯|>π‘Ÿ0βˆ’Ξ“(𝑇) and thus is a solution on the line for all 𝑇≀𝑇0. Performing a similar procedure in [𝑇0,2𝑇0],…,[π‘šπ‘‡0,(π‘š+1)𝑇0],…, we obtain a compactly supported nonnegative solution of the Cauchy problem for all 𝑇β‰₯0 and Theorem 2.2 holds with Ξ©=ℝ1.

Theorem 3.2. Let 1<𝑛<3. Assume β„Ž0 is nonnegative, β„Ž0∈𝐻1(ℝ1), suppβ„Ž0βŠ‚(βˆ’π‘Ÿ0,π‘Ÿ0) and β„Ž is a solution of the Cauchy problem. Then the following estimates:  Γ(𝑇)≀𝐷1(𝑇1/(𝑛+4)+𝑇5/(𝑛+4)) for all 𝑇>0 if 1<𝑛<2,  Γ(𝑇)≀𝐷2𝑇1/(𝑛+4) for small enough time if 2≀𝑛<3, are valid. Here the constants 𝐷𝑖 depend on the parameters problem and initial data only.

3.1. Proof of Theorem 3.1 for the Case 1<𝑛<2

The following lemma contains the local entropy estimate. The proof of Lemma 3.3 is similar to (A.16), (A.29), therefore it is omitted.

Lemma 3.3. Let 𝜁∈𝐢1,2𝑑,π‘₯(𝑄𝑇) such that suppπœβŠ‚Ξ©, (𝜁4)ξ…ž=0 on πœ•Ξ©, and 𝜁4(βˆ’π‘Ž,𝑑)=𝜁4(π‘Ž,𝑑). Assume that βˆ’1/2<𝛼<1, and 𝛼≠0. Then there exist constants 𝐢𝑖(𝑖=1,2,3) dependent on 𝑛,β€‰β€‰π‘š,  𝛼,β€‰β€‰π‘Ž0, and π‘Ž1, independent of Ξ©, such that for all 0<𝑇<βˆžξ€œΞ©πœ4(π‘₯,𝑇)𝐺0(𝛼)(ξ€β„Ž(π‘₯,𝑇))𝑑π‘₯βˆ’π‘„π‘‡ξ€·πœ4𝑑𝐺0(𝛼)(β„Ž)𝑑π‘₯𝑑𝑑+𝐢1ξ€π‘„π‘‡ξ€·β„Ž(𝛼+2)/2ξ€Έ2π‘₯π‘₯𝜁4ξ€œπ‘‘π‘₯π‘‘π‘‘β‰€Ξ©πœ4(π‘₯,0)𝐺0(𝛼)ξ€·β„Ž0𝑑π‘₯+𝐢2ξ€π‘„π‘‡β„Žπ›Ό+2ξ€·πœ4+𝜁4π‘₯+𝜁2𝜁2π‘₯π‘₯+𝜁2𝜁2π‘₯+𝜁3||𝜁π‘₯π‘₯||𝑑π‘₯𝑑𝑑+𝐢3ξ€π‘„π‘‡β„Žπ›Ό+1ξ€·||𝜁3||||𝜁π‘₯||+𝜁4𝑑π‘₯𝑑𝑑.(3.1)

Let 0<𝑛<2, and let suppβ„Ž0βŠ†(βˆ’π‘Ÿ0,π‘Ÿ0)⋐Ω. For an arbitrary π‘ βˆˆ(0,π‘Žβˆ’π‘Ÿ0) and 𝛿>0 we consider the families of sets Ξ©ξ€·(𝑠)=Ξ©β§΅βˆ’π‘Ÿ0βˆ’π‘ ,π‘Ÿ0ξ€Έ+𝑠,𝑄𝑇(𝑠)=(0,𝑇)Γ—Ξ©(𝑠).(3.2) We introduce a nonnegative cutoff function πœ‚(𝜏) from the space 𝐢2(ℝ1) with the following properties: ⎧βŽͺ⎨βŽͺβŽ©πœπœ‚(𝜏)=0ifπœβ‰€0,2(3βˆ’2𝜏)if0<𝜏<1,1ifπœβ‰€1.(3.3) Next we introduce our main cut-off functions πœ‚π‘ ,𝛿(π‘₯)∈𝐢2(Ξ©) such that 0β‰€πœ‚π‘ ,𝛿(π‘₯)≀1 for all π‘₯∈Ω and possess the following properties: πœ‚π‘ ,π›Ώξƒ©ξ€·π‘Ÿ(π‘₯)=πœ‚|π‘₯|βˆ’0ξ€Έ+𝑠𝛿ξƒͺ=ξ‚»||ξ€·πœ‚1,π‘₯∈Ω(𝑠+𝛿),0,π‘₯∈Ω⧡Ω(𝑠),𝑠,𝛿π‘₯||≀3𝛿,||ξ€·πœ‚π‘ ,𝛿π‘₯π‘₯||≀6𝛿2,(3.4) for all 𝑠>0,  𝛿>0βˆΆπ‘Ÿ0+𝑠+𝛿<π‘Ž. Choosing 𝜁4(π‘₯,𝑑)=πœ‚π‘ ,𝛿(π‘₯)π‘’βˆ’π‘‘/𝑇, from (3.1) we arrive at ξ€œΞ©(𝑠+𝛿)β„Žπ›Όβˆ’π‘›+2(1𝑇)𝑑π‘₯+𝑇𝑄𝑇(𝑠+𝛿)β„Žπ›Όβˆ’π‘›+2𝑑π‘₯𝑑𝑑+𝐢𝑄𝑇(𝑠+𝛿)ξ€·β„Ž(𝛼+2)/2ξ€Έ2π‘₯π‘₯≀𝐢𝑑π‘₯𝑑𝑑𝛿4𝑄𝑇(𝑠)β„Žπ›Ό+2𝐢𝑑π‘₯𝑑𝑑+𝛿𝑄𝑇(𝑠)β„Žπ›Ό+1𝑑π‘₯𝑑𝑑=∢𝐢2𝑖=1π›Ώβˆ’π›Όπ‘–ξ€π‘„π‘‡(𝑠)β„Žπœ‰π‘–,(3.5) for all π‘ βˆˆ(0,π‘Žβˆ’π‘Ÿ0), where (π‘›βˆ’1)+<𝛼<1 and 0<𝛿<1 is enough small. We apply the Nirenberg-Gagliardo interpolation inequality (see Lemma B.2) in the region Ξ©(𝑠+𝛿) to a function π‘£βˆΆ=β„Ž(𝛼+2)/2 with π‘Ž=(2πœ‰π‘–)/(𝛼+2),  𝑏=(2(π›Όβˆ’π‘›+2))/(𝛼+2),  𝑑=2,  𝑖=0,  𝑗=2, and πœƒπ‘–=((𝛼+2)(πœ‰π‘–βˆ’π›Ό+π‘›βˆ’2))/(πœ‰π‘–(4π›Όβˆ’3𝑛+8)) under the conditions: π›Όβˆ’π‘›+2<πœ‰π‘–for𝑖=1,2.(3.6) Integrating the resulted inequalities with respect to time and taking into account (3.5), we arrive at the following relations: 𝑄𝑇(𝑠+𝛿)β„Žπœ‰π‘–β‰€πΆπ‘‡1βˆ’(πœƒπ‘–πœ‰π‘–)/(𝛼+2)2𝑖=1π›Ώβˆ’π›Όπ‘–ξ€π‘„π‘‡(𝑠)β„Žπœ‰π‘–ξƒͺ1+πœˆπ‘–ξƒ©+𝐢𝑇2𝑖=1π›Ώβˆ’π›Όπ‘–ξ€π‘„π‘‡(𝑠)β„Žπœ‰π‘–ξƒͺπœ‰π‘–/(π›Όβˆ’π‘›+2),(3.7) where πœˆπ‘–=(4(πœ‰π‘–βˆ’π›Ό+π‘›βˆ’2))/(4π›Όβˆ’3𝑛+8). These inequalities are true provided that πœƒπ‘–πœ‰π‘–π›Ό+2<1βŸΊπœ‰π‘–<5π›Όβˆ’4𝑛+10for𝑖=1,2.(3.8) Simple calculations show that inequalities (3.6) and (3.8) hold with some (π‘›βˆ’1)+<𝛼<1 if and only if 1<𝑛<2. The finite speed of propagations follows from (3.7) by applying Lemma B.3 with 𝑠1=0. Hence, ξ€·suppβ„Ž(𝑇,β‹…)βŠ‚βˆ’π‘Ÿ0βˆ’Ξ“(𝑇),π‘Ÿ0ξ€Έξ€Ί+Ξ“(𝑇)⋐Ωforallπ‘‡βˆΆπ‘‡βˆˆ0,𝑇0ξ€»,(3.9) where 𝑇0∢=Ξ“βˆ’1(π‘Žβˆ’π‘Ÿ0).

3.2. Proof of Theorem 3.2 for the Case 1<𝑛<2

We can repeat the previous procedure from Section 3.1 for Ξ©(𝑠)=ℝ1⧡(βˆ’π‘Ÿ0βˆ’π‘ ,π‘Ÿ0+𝑠) and we obtain 𝐺𝑖(𝑠+𝛿)∢=𝑄𝑇(𝑠+𝛿)β„Žπœ‰π‘–β‰€πΆπ‘‡1βˆ’(πœƒπ‘–πœ‰π‘–)/(𝛼+2)2𝑖=1π›Ώβˆ’π›Όπ‘–ξ€π‘„π‘‡(𝑠)β„Žπœ‰π‘–ξƒͺ1+πœˆπ‘–,(3.10) instead of (3.7), and 𝑇Γ(𝑇)=𝐢(1βˆ’(πœƒ1πœ‰1)/(𝛼+2))(1+𝜈2)𝑇(1βˆ’(πœƒ2πœ‰2)/(𝛼+2))𝜈1(1+𝜈1)(𝐺(0))𝜈1ξ€Έ1/(4(1+𝜈1)(1+𝜈2))𝑇+𝐢(1βˆ’(πœƒ2πœ‰2)/(𝛼+2))(1+𝜈1)𝑇(1βˆ’(πœƒ1πœ‰1)/(𝛼+2))𝜈2(1+𝜈2)(𝐺(0))𝜈2ξ€Έ1/((1+𝜈1)(1+𝜈2)),(3.11) where 𝑇𝐺(0)=𝐢(1βˆ’(πœƒ1πœ‰1)/(𝛼+2))(1+𝜈2)𝐺2ξ€Έ(0)1+𝜈1+𝑇(1βˆ’(πœƒ2πœ‰2)/(𝛼+2))(1+𝜈1)𝐺1ξ€Έ(0)1+𝜈2.(3.12) Now we need to estimate 𝐺(0). With that end in view, we obtain the following estimates: 𝐺𝑖(0)≀𝐢1𝐢2+𝐢3𝑇(πœ‰π‘–βˆ’1)/(𝛼+5)𝑇1βˆ’(πœ‰π‘–βˆ’1)/(𝛼+5),𝑖=1,2,(3.13) where 1<πœ‰π‘–<𝛼+6, and 𝐢𝑖 depends on initial data only. Really, applying the Nirenberg-Gagliardo interpolation inequality (see Lemma B.2) in Ξ©=ℝ1 to a function π‘£βˆΆ=β„Ž(𝛼+2)/2 with π‘Ž=(2πœ‰π‘–)/(𝛼+2),  𝑏=2/(𝛼+2),  𝑑=2,  𝑖=0,  𝑗=2, and Μƒπœƒπ‘–=((𝛼+2)(πœ‰π‘–βˆ’1))/(πœ‰π‘–(𝛼+5)) under the condition πœ‰π‘–>1, we deduce that ξ€œβ„1β„Žπœ‰π‘–β€–β€–β„Žβ‰€π‘0β€–β€–(2(3πœ‰π‘–1+𝛼+2))/((𝛼+2)(𝛼+5))ξ‚΅ξ€œβ„1ξ€·β„Ž(𝛼+2)/2ξ€Έ2π‘₯π‘₯𝑑π‘₯(πœ‰π‘–βˆ’1)/(𝛼+5).(3.14) Integrating (3.14) with respect to time and taking into account the HΓΆlder inequality ((πœ‰π‘–βˆ’1)/(𝛼+5)<1β‡’πœ‰π‘–<𝛼+6), we arrive at the following relations: ξ€π‘„π‘‡β„Žπœ‰π‘–β€–β€–β„Žβ‰€π‘0β€–β€–(2(3πœ‰π‘–1+𝛼+2))/((𝛼+2)(𝛼+5))𝑇1βˆ’(πœ‰π‘–βˆ’1)/(𝛼+5)ξ‚΅ξ€π‘„π‘‡ξ€·β„Ž(𝛼+2)/2ξ€Έ2π‘₯π‘₯𝑑π‘₯(πœ‰π‘–βˆ’1)/(𝛼+5).(3.15) From (3.15), due to (A.16) (as πœ€β†’0) and (2.31), we find (3.13).

Inserting (3.13) into (3.12), we obtain after straightforward computations that 𝑇Γ(𝑇)≀𝐢1/(𝑛+4)+𝑇5/(𝑛+4)ξ€Έforall𝑇β‰₯0.(3.16)

3.3. Proof of Theorem 3.1 for the Case 4/3<𝑛<3

The following lemma contains the local energy estimate. The proof of Lemma 3.4 is Appendix A.

Lemma 3.4. Let π‘›βˆˆ(1/2,3) and 𝛽>(1βˆ’π‘›)/3. Let 𝜁∈𝐢2(Ξ©) such that supp𝜁 in Ξ© and (𝜁6)ξ…ž=0 on πœ•Ξ©, and 𝜁(βˆ’π‘Ž)=𝜁(π‘Ž). Then there exist constants 𝐢𝑖(𝑖=1,3) dependent on 𝑛,β€‰β€‰π‘š,β€‰β€‰π‘Ž0, and π‘Ž1, independent of Ξ© and πœ€, such that for any 0<𝑇<βˆžξ€œΞ©πœ6β„Ž2π‘₯(ξ€œπ‘₯,𝑇)𝑑π‘₯+Ω𝜁4β„Žπ›½+1(𝑇)𝑑π‘₯+𝐢1ξ€π‘„π‘‡πœ6ξ€·β„Ž(𝑛+2)/2ξ€Έ2π‘₯π‘₯π‘₯β‰€ξ€œπ‘‘π‘₯π‘‘π‘‘Ξ©πœ6β„Ž20ξ€œ(π‘₯)𝑑π‘₯+Ω𝜁4β„Ž0𝛽+1𝑑π‘₯+𝐢2ξ€π‘„π‘‡β„Žπ‘›+2ξ‚€πœ6+𝜁6π‘₯+𝜁3||𝜁π‘₯π‘₯||3𝑑π‘₯𝑑𝑑+𝐢3ξ€π‘„π‘‡ξ€½πœ’{𝜁>0}β„Žπ‘›+3π›½βˆ’1+β„Žπ‘›πœ6ξ€Ύξ€œπ‘‘π‘₯𝑑𝑑,(3.17)Ω𝜁6β„Ž2π‘₯(π‘₯,𝑇)𝑑π‘₯+𝐢1ξ€π‘„π‘‡πœ6ξ€·β„Ž(𝑛+2)/2ξ€Έ2π‘₯π‘₯π‘₯ξ€œπ‘‘π‘₯π‘‘π‘‘β©½Ξ©πœ6β„Ž20(π‘₯)𝑑π‘₯+𝐢2ξ€π‘„π‘‡β„Žπ‘›+2ξ‚€πœ6+𝜁6π‘₯+𝜁3||𝜁π‘₯π‘₯||3𝑑π‘₯𝑑𝑑+𝐢3ξ€π‘„π‘‡β„Žπ‘›πœ6𝑑π‘₯𝑑𝑑.(3.18)

Let πœ‚π‘ ,𝛿(π‘₯) be denoted by (3.4). Setting 𝜁6(π‘₯)=πœ‚π‘ ,𝛿(π‘₯) into (3.17), after simple transformations, we obtain ξ€œΞ©(𝑠+𝛿)β„Ž2π‘₯(ξ€œπ‘₯,𝑇)𝑑π‘₯+Ξ©(𝑠+𝛿)β„Žπ›½+1(𝑇)𝑑π‘₯+𝐢𝑄𝑇(𝑠+𝛿)ξ€·β„Ž(𝑛+2)/2ξ€Έ2π‘₯π‘₯π‘₯≀𝐢𝑑π‘₯𝑑𝑑𝛿6𝑄𝑇(𝑠)β„Žπ‘›+2𝑑π‘₯𝑑𝑑+𝐢𝑄𝑇(𝑠)ξ€½β„Žπ‘›+3π›½βˆ’1+β„Žπ‘›ξ€Ύπ‘‘π‘₯𝑑𝑑=∢𝐢3𝑖=1π›Ώβˆ’π›Όπ‘–ξ€π‘„π‘‡(𝑠)β„Žπœ‰π‘–,(3.19) for all for all π‘ βˆˆ(0,π‘Žβˆ’π‘Ÿ0),𝛿>0βˆΆπ‘Ÿ0+𝑠+𝛿<π‘Ž. We apply the Nirenberg-Gagliardo interpolation inequality (see Lemma B.2) in the region Ξ©(𝑠+𝛿) to a function π‘£βˆΆ=β„Ž(𝑛+2)/2 with π‘Ž=(2πœ‰π‘–)/(𝑛+2),  𝑏=(2(𝛽+1))/(𝑛+2),  𝑑=2,  𝑖=0,  𝑗=3, and πœƒπ‘–=((𝑛+2)(πœ‰π‘–βˆ’π›½βˆ’1))/(πœ‰π‘–(𝑛+5𝛽+7)) under the conditions: 𝛽<πœ‰π‘–βˆ’1for𝑖=1,3.(3.20) Integrating the resulted inequalities with respect to time and taking into account (3.19), we arrive at the following relations: 𝑄𝑇(𝑠+𝛿)β„Žπœ‰π‘–β‰€πΆπ‘‡1βˆ’(πœƒπ‘–πœ‰π‘–)/(𝑛+2)3𝑖=1π›Ώβˆ’π›Όπ‘–ξ€π‘„π‘‡(𝑠)β„Žπœ‰π‘–ξƒͺ1+πœˆπ‘–ξƒ©+𝐢𝑇3𝑖=1π›Ώβˆ’π›Όπ‘–ξ€π‘„π‘‡(𝑠)β„Žπœ‰π‘–ξƒͺπœ‰π‘–/(𝛽+1),(3.21) where πœˆπ‘–=(6(πœ‰π‘–βˆ’π›½βˆ’1))/(𝑛+5𝛽+7). These inequalities are true provided that πœƒπ‘–πœ‰π‘–πœ‰π‘›+2<1βŸΊπ›½>π‘–βˆ’π‘›βˆ’86for𝑖=1,3.(3.22) Simple calculations show that inequalities (3.20) and (3.22) hold with some π›½βˆˆ((2βˆ’π‘›)/2,π‘›βˆ’1) if and only if 4/3<𝑛<3. Since all integrals on the right-hand sides of (3.21) vanish as 𝑇→0, the finite speed of propagations follows from (3.21) by applying Lemma B.3 with 𝑠1=0 and sufficiently small 𝑇. Hence, ξ€·suppβ„Ž(𝑇,β‹…)βŠ‚βˆ’π‘Ÿ0βˆ’Ξ“(𝑇),π‘Ÿ0ξ€Έ+Ξ“(𝑇)⋐Ωforallπ‘‡βˆΆ0≀𝑇≀𝑇0.(3.23)

3.4. Proof of Theorem 3.2 for the Case 4/3<𝑛<3

Suppose that Ξ©(𝑠)=ℝ1⧡{π‘₯∢|π‘₯|<𝑠},  𝑄𝑇(𝑠)=(0,𝑇)Γ—Ξ©(𝑠) for all 𝑠>π‘Ÿ0, suppβ„Ž0βŠ†(βˆ’π‘Ÿ0,π‘Ÿ0), and Ξ“(𝑇)=π‘Ÿ(𝑇)βˆ’π‘Ÿ0. Since the time interval is small, we can assume that π‘Ÿ(𝑇)<2π‘Ÿ0. Hence, for all π‘ βˆˆ(π‘Ÿ0,2π‘Ÿ0), we can take (up to regularization) 𝜁=(|π‘₯|βˆ’π‘ )+ in (3.18). As a result, we obtain 12ξ€œΞ©(𝑠)(|π‘₯|βˆ’π‘ )6+β„Ž2π‘₯𝑑π‘₯+𝛿6𝐢1𝑄𝑇(𝑠+𝛿)ξ€·β„Ž(𝑛+2)/2ξ€Έ2π‘₯π‘₯π‘₯𝑑π‘₯𝑑𝑑≀𝐢4𝑄𝑇(𝑠)ξ€½β„Žπ‘›+2+(π‘Ÿ(𝑇)βˆ’π‘ )6+β„Žπ‘›ξ€Ύπ‘‘π‘₯𝑑𝑑,(3.24) for all 𝑇≀𝑇0,β€‰β€‰π‘ βˆˆ(π‘Ÿ0,2π‘Ÿ0). Using the Hardy type inequality ξ€œΞ©(𝑠)(|π‘₯|βˆ’π‘ )𝛼+𝑓2𝑑π‘₯≀𝐢0ξ€œΞ©(𝑠)(|π‘₯|βˆ’π‘ )+𝛼+2𝑓2π‘₯𝑑π‘₯,(3.25) where 𝐢0=4/(𝛼+1)2 and 𝛼>βˆ’1, we deduce that ξ€œΞ©(𝑠+𝛿)ξ‚΅ξ€œβ„Žπ‘‘π‘₯≀Ω(𝑠+𝛿)(|π‘₯|βˆ’π‘ )4+β„Ž2𝑑π‘₯1/2ξ‚΅ξ€œΞ©(𝑠+𝛿)(|π‘₯|βˆ’π‘ )+βˆ’4𝑑π‘₯1/2≀𝐢03𝛿3ξ‚Ά1/2ξ‚΅ξ€œΞ©(𝑠)(|π‘₯|βˆ’π‘ )6+β„Ž2π‘₯𝑑π‘₯1/2,(3.26) whence ξ‚΅ξ€œΞ©(𝑠+𝛿)ξ‚Άβ„Žπ‘‘π‘₯2≀𝐢03π›Ώβˆ’3ξ€œΞ©(𝑠)(|π‘₯|βˆ’π‘ )6+β„Ž2π‘₯𝑑π‘₯,(3.27) for all 𝛿>0, π‘ βˆˆ(π‘Ÿ0,2π‘Ÿ0). Substituting (3.27) in (3.24), we get 32𝐢0π›Ώβˆ’3supπ‘‘ξ‚΅ξ€œΞ©(𝑠+𝛿)ξ‚Άβ„Žπ‘‘π‘₯2+𝐢1𝑄𝑇(𝑠+𝛿)ξ€·β„Ž(𝑛+2)/2ξ€Έ2π‘₯π‘₯π‘₯≀𝐢𝑑π‘₯𝑑𝑑4𝛿6𝑄𝑇(𝑠)ξ€½β„Žπ‘›+2+Ξ“6(𝑇)β„Žπ‘›ξ€Ύπ‘‘π‘₯𝑑𝑑,(3.28) for all 𝑇≀𝑇0,β€‰β€‰π‘ βˆˆ(π‘Ÿ0,2π‘Ÿ0). By the Nirenberg-Gagliardo, HΓΆlder and Young inequalities, after simple transformations, for πœ–π‘–>0, we have 𝐢4𝛿6𝑄𝑇(𝑠)β„Žπ‘›+2𝑑π‘₯π‘‘π‘‘β‰€πœ–1𝑄𝑇(𝑠)ξ€·β„Ž(𝑛+2)/2ξ€Έ2π‘₯π‘₯π‘₯πΆξ€·πœ–π‘‘π‘₯𝑑𝑑+1𝛿𝑛+7ξ€œπ‘‡0ξ‚΅ξ€œΞ©(𝑠)ξ‚Άβ„Žπ‘‘π‘₯𝑛+2𝐢𝑑𝑑,4Ξ“6(𝑇)𝛿6𝑄𝑇(𝑠)β„Žπ‘›π‘‘π‘₯π‘‘π‘‘β‰€πœ–2𝑄𝑇(𝑠)ξ€·β„Ž(𝑛+2)/2ξ€Έ2π‘₯π‘₯π‘₯ξ€·πœ–π‘‘π‘₯𝑑𝑑+𝐢2ξ€Έξ‚΅Ξ“(𝑇)𝛿3(𝑛+7)/4ξ€œπ‘‡0ξ‚΅ξ€œΞ©(𝑠)ξ‚Άβ„Žπ‘‘π‘₯(3𝑛+1)/4𝑑𝑑.(3.29) Substituting the estimates (3.29) in (3.28) and making the standard iterative procedure for small enough 0<πœ–π‘–<1, we arrive at the inequality 32𝐢0supπ‘‘ξ‚΅ξ€œΞ©(𝑠+𝛿)ξ‚Άβ„Žπ‘‘π‘₯2+𝐢5𝛿3𝑄𝑇(𝑠+𝛿)ξ€·β„Ž(𝑛+2)/2ξ€Έ2π‘₯π‘₯π‘₯𝑑π‘₯𝑑𝑑≀𝐢62𝑖=1𝐺𝑇(𝑖)(𝑠)𝛿𝛼𝑖,(3.30) where    𝛼1=𝑛+4,  𝛼2=(3(𝑛+3))/4,    𝐺𝑇(1)∫(𝑠)∢=𝑇0(∫Ω(𝑠)β„Žπ‘‘π‘₯)𝑛+2𝑑𝑑,    𝐺𝑇(2)(𝑠)∢=Ξ“(3(𝑛+7))/4∫(𝑇)𝑇0(∫Ω(𝑠)β„Žπ‘‘π‘₯)(3𝑛+1)/4𝑑𝑑. Thus, (3.30) yields 𝐺𝑇(𝑖)(𝑠+𝛿)≀𝐢7π‘‡Ξ“πœ‡π‘–ξƒ©(𝑇)2𝑖=1𝐺𝑇(𝑖)(𝑠)𝛿𝛼𝑖ξƒͺ𝛽𝑖,(3.31) for all π‘ βˆˆ(π‘Ÿ0,2π‘Ÿ0) and 0<𝛿<𝑠, where πœ‡1=0,β€‰β€‰πœ‡2=(3(𝑛+7))/4,  𝛽1=(𝑛+2)/2,  𝛽2=(3𝑛+1)/8. By Lemma B.3, from (3.31) we find that 𝐺𝑇(𝑖)(𝑠0)=0, where Ξ“(𝑇)≀𝑠0(𝑇)=𝐢8(𝑇1/𝛼1+𝑇1/𝛼2Ξ“πœ‡2/𝛼2(𝑇)). As πœ‡2/𝛼2=(𝑛+7)/(𝑛+3)>1 for any 𝑇≀𝑇0, we have Ξ“(𝑇)≀𝐢9𝑇1/(𝑛+4).

4. Waiting Time Phenomenon

Let Ξ©(𝑠)={π‘₯∢π‘₯β‰₯𝑠} for all π‘ βˆˆβ„1, and 𝐑0(ξ€œπ‘ )∢=Ξ©(𝑠)β„Ž0π›Όβˆ’π‘›+2(π‘₯)𝑑π‘₯=0βˆ€π‘ β‰₯0,(4.1) where (π‘›βˆ’1)+<𝛼<1. Let us assume that the function 𝐑0(𝑠) satisfies the flatness conditions. Namely, for every π‘ βˆΆπ‘ 0<𝑠<0 the following estimate: 𝐑0ξ€½(𝑠)β‰€πœ’max(βˆ’π‘ )1+(4(π›Όβˆ’π‘›+2))/𝑛,(βˆ’π‘ )1+(4βˆ’3𝑛+4(π›Όβˆ’π‘›+2))/(4(π‘›βˆ’1))ξ€Ύ=⎧βŽͺ⎨βŽͺβŽ©πœ’(βˆ’π‘ )1+(4(π›Όβˆ’π‘›+2))/𝑛4for3≀𝑛<2,πœ’(βˆ’π‘ )1+(4βˆ’3𝑛+4(π›Όβˆ’π‘›+2))/(4(π‘›βˆ’1))4for1<𝑛<3,(4.2) is valid.

Theorem 4.1. Let 1<𝑛<2. Assume β„Ž0 is nonnegative, β„Ž0∈𝐻1(ℝ1) and meas{Ξ©(𝑠)∩suppβ„Ž0}=βˆ… for all 𝑠β‰₯0, that is, the condition (4.1) is valid, and the flatness condition (4.2) also holds.
Then for the solution β„Ž of Theorem 2.2 (with Ξ©=ℝ1) there exists the time π‘‡βˆ—=π‘‡βˆ—(πœ’)>0 depending on the known parameters only such that suppβ„Ž(𝑑,β‹…)∩Ω(0)=βˆ…βˆ€0<π‘‘β‰€π‘‡βˆ—,(4.3) where πœ’ is the constant from the flatness condition. Note, that π‘‡βˆ—β†’+∞ as πœ’β†’0.

Remark 4.2. Let the initial data β„Ž0∈𝐢(ℝ1) satisfy the following properties:(1) if 1<𝑛<4/3 then we suppose that supπ‘₯∈Ω(𝑠)β„Ž0(π‘₯)β‰€πœ’(βˆ’π‘ )(4βˆ’3𝑛+4(π›Όβˆ’π‘›+2))/(4(π‘›βˆ’1)(π›Όβˆ’π‘›+2))ξ€·forsomeπ›Όβˆˆ(π‘›βˆ’1)+ξ€Έ;,1(4.4) (2) if 4/3≀𝑛<2 then we suppose that supπ‘₯∈Ω(𝑠)β„Ž0(π‘₯)β‰€πœ’(βˆ’π‘ )4/𝑛.
These assumptions on the initial data are sufficient for the validity of flatness condition (4.2) and guarantee the appearance of the WTP, that is, the validation of property (4.3).

Remark 4.3. Note that due to Lemma 2.5 we have the estimate β„Žπ›Ό+2(π‘₯,𝑑)≀𝐢(1+𝑑)β„Žπ›Ό+1(π‘₯,𝑑). Therefore, using this inequality in (3.1) with Ξ©=ℝ1, we could also obtain the waiting time phenomenon by the application of Theorem 2.1 from [46] with 𝑀=β„Ž(𝛼+2)/2,  𝑙=π‘˜=𝑝=2,β€‰β€‰π‘ž=(2(π›Όβˆ’π‘›+2))/(𝛼+2), and 𝑠=(2(𝛼+1))/(𝛼+2).

Proof of Theorem 4.1. Similar to (3.10) for Ξ©(𝑠)={π‘₯∢π‘₯β‰₯𝑠} and we obtain 𝐺𝑖(𝑠+𝛿)∢=𝑄𝑇(𝑠+𝛿)β„Žπœ‰π‘–β‰€πΎπ‘‡1βˆ’(πœƒπ‘–πœ‰π‘–)/(𝛼+2)2ξ“π‘˜=1π›Ώβˆ’π›Όπ‘˜πΊπ‘˜(𝑠)+𝐑0ξƒͺ(𝑠)1+πœˆπ‘–.(4.5) Let us check that all conditions of Lemma B.4 are satisfied. We denote by 𝐺max(𝑠)∢=max𝑖=1,2⎧βŽͺ⎨βŽͺβŽ©π‘02𝛽+12ξ“π‘˜=1ξ€·πΊπ‘˜ξ€Έ(𝑠)π›½π‘˜ξƒͺπ›½π‘–βˆ’1⎫βŽͺ⎬βŽͺ⎭(𝑠)1/(𝛼𝑖𝛽),𝑔max(𝑠)∢=max𝑖=1,2⎧βŽͺ⎨βŽͺ⎩2(𝛽+1)/(𝛼𝑖𝛽)22π›½βˆ’1ξ“π‘˜=1𝐾𝑇1βˆ’(πœƒπ‘˜πœ‰π‘˜)/(𝛼+2)ξ€Έπ›½π‘˜ξƒͺ𝛽𝑖/𝛼𝑖𝐑0ξ€Έ(𝑠)(π›½π‘–βˆ’1)/π›Όπ‘–βŽ«βŽͺ⎬βŽͺ⎭,𝑐0=22π›½βˆ’1ξ“π‘˜=1𝐾𝑇1βˆ’(πœƒπ‘˜πœ‰π‘˜)/(𝛼+2)ξ€Έπ›½π‘˜,𝛽𝑖=1+πœˆπ‘–,𝛽=𝛽1𝛽2.(4.6) Taking 𝑠=βˆ’2𝛿 in (4.5) and passing to the limit π›Ώβ†’βˆž, due to the boundedness of functions πΊπ‘˜(𝑠) and 𝐑0(𝑠), we deduce πΊπ‘˜(βˆ’βˆž)≀𝐾𝑇1βˆ’(πœƒπ‘˜πœ‰π‘˜)/(𝛼+2)π‘π›½π‘˜0(βˆ’βˆž).(4.7) This implies that the condition (i) of Lemma B.4 is fulfilled. Because of the assumption (4.2) on the function 𝐑0(𝑠), we can find π‘‡βˆ— such that the condition (ii) of Lemma B.4 is valid for all π‘‡βˆˆ[0,π‘‡βˆ—]. Here π‘‡βˆ—=π‘‡βˆ—(πœ’) goes to infinity as πœ’β†’0. Hence, the application of Lemma B.4 ends the proof.

Appendices

A. Proofs of a Priori Estimates

The first observation is that the periodic boundary conditions imply that classical solutions of (2.17) conserve mass: ξ€œΞ©β„Žπ›Ώπœ€(ξ€œπ‘₯,𝑑)𝑑π‘₯=Ξ©β„Ž0,πœ€(π‘₯)𝑑π‘₯=π‘€πœ€<∞forall𝑑>0.(A.1) Further, (2.21) implies π‘€πœ€βˆ«β„Žβ†’π‘€=0 as πœ€β†’0. Also, we will relate the 𝐿𝑝 norm of β„Ž to the 𝐿𝑝 norm of its zero-mean part as follows: ||||≀||||π‘€β„Ž(π‘₯)β„Ž(π‘₯)βˆ’πœ€Ξ©||||+π‘€πœ€Ξ©β‡’β€–β„Žβ€–π‘π‘β‰€2π‘βˆ’1‖𝑣‖𝑝𝑝+ξ‚΅2||Ξ©||ξ‚Άπ‘βˆ’1π‘€π‘πœ€,(A.2) where π‘£βˆΆ=β„Žβˆ’π‘€πœ€/Ξ©. We will use the PoincarΓ© inequality which holds for any zero-mean function in 𝐻1(Ξ©)‖𝑣‖𝑝𝑝≀𝑏1‖‖𝑣π‘₯‖‖𝑝𝑝1≀𝑝<∞,𝑏1=||Ξ©||𝑝𝑝.(A.3) Also used will be an interpolation inequality [47, Theorem 2.2, page 62] for functions of zero mean in 𝐻1(Ξ©): ‖𝑣‖𝑝𝑝≀𝑏2‖‖𝑣π‘₯β€–β€–2π‘Žπ‘β€–π‘£β€–π‘Ÿ(1βˆ’π‘Ž)𝑝,(A.4) where π‘Ÿβ‰₯1,  𝑝β‰₯π‘Ÿ,β€‰β€‰π‘Ž=(1/π‘Ÿβˆ’1/𝑝)/(1/π‘Ÿ+1/2),  𝑏2=(1+π‘Ÿ/2)π‘Žπ‘. It follows that for any zero-mean function 𝑣 in 𝐻1(Ξ©)‖𝑣‖𝑝𝑝≀𝑏3‖‖𝑣π‘₯‖‖𝑝2,β‡’β€–β„Žβ€–π‘π‘β‰€π‘4β€–β€–β„Žπ‘₯‖‖𝑝2+𝑏5π‘€π‘πœ€,(A.5) where 𝑏3=𝑏1||Ξ©||(2βˆ’π‘)/𝑝𝑏if1≀𝑝≀21(𝑝+2)/2𝑏2𝑏if2<𝑝<∞,4=2π‘βˆ’1𝑏3,𝑏5=ξ‚΅2||Ξ©||ξ‚Άπ‘βˆ’1.(A.6) To see that (A.5) holds, consider two cases. If 1≀𝑝<2, then by (A.3), ‖𝑣‖𝑝 is controlled by ‖𝑣π‘₯‖𝑝. By the HΓΆlder inequality, ‖𝑣π‘₯‖𝑝 is then controlled by ‖𝑣π‘₯β€–2. If 𝑝>2 then by (A.4), ‖𝑣‖𝑝 is controlled by ‖𝑣π‘₯β€–π‘Ž2‖𝑣‖21βˆ’π‘Ž where π‘Ž=1/2βˆ’1/𝑝. By the PoincarΓ© inequality, ‖𝑣‖21βˆ’π‘Ž is controlled by ‖𝑣π‘₯β€–21βˆ’π‘Ž.

Proof of Lemma 2.3. In the following, we denote the solution β„Žπ›Ώπœ€ by β„Ž whenever there is no chance of confusion.
To prove the bound (2.23) one starts by multiplying (2.17) by βˆ’β„Žπ‘₯π‘₯, integrating over 𝑄𝑇, and using the periodic boundary conditions (2.18) yields 12ξ€œΞ©β„Ž2π‘₯(π‘₯,𝑇)𝑑π‘₯+π‘Ž0ξ€π‘„π‘‡π‘“π›Ώπœ€(β„Ž)β„Ž2π‘₯π‘₯π‘₯1𝑑π‘₯𝑑𝑑=2ξ€œΞ©β„Ž20πœ€,π‘₯(π‘₯)𝑑π‘₯βˆ’π‘Ž1ξ€π‘„π‘‡π‘“π›Ώπœ€(β„Ž)β„Žπ‘₯β„Žπ‘₯π‘₯π‘₯𝑑π‘₯π‘‘π‘‘βˆ’π‘„π‘‡π‘“π›Ώπœ€(β„Ž)𝑀π‘₯β„Žπ‘₯π‘₯π‘₯𝑑π‘₯𝑑𝑑.(A.7) By Cauchy and Young inequalities, due to (A.3)–(A.5), it follows from (A.7) that 12ξ€œΞ©β„Ž2π‘₯(π‘Žπ‘₯,𝑇)𝑑π‘₯+02ξ€π‘„π‘‡π‘“π›Ώπœ€(β„Ž)β„Ž2π‘₯π‘₯π‘₯1𝑑π‘₯𝑑𝑑≀2ξ€œΞ©β„Ž20πœ€,π‘₯𝑑π‘₯+𝑐1ξ€π‘„π‘‡β„Ž2π‘₯π‘₯𝑑π‘₯𝑑𝑑+𝑐2ξ€œπ‘‡0ξ‚»ξ‚΅ξ€œmax1,Ξ©β„Ž2π‘₯𝑑π‘₯πœ…1𝑑𝑑,(A.8) where πœ…1=max{𝑛,3},  𝑐1=(π‘Ž21/4π‘Ž0)𝑏2,  𝑐2=(π‘Ž21/2π‘Ž0)𝑏4+(π‘Ž21/2π‘Ž0)𝑏5π‘€πœ€2𝑛+(π‘Ž21/4π‘Ž0)𝑏2+(π‘Ž21/π‘Ž0)𝛿 + sup𝑑≀𝑇((‖𝑀π‘₯β€–22/π‘Ž0)𝛿+(‖𝑀π‘₯β€–2∞/π‘Ž0)𝑏4+(‖𝑀π‘₯β€–2∞/π‘Ž0)𝑏5π‘€π‘›πœ€). Multiplying (2.17) by πΊβ€²π›Ώπœ€(β„Ž), integrating over 𝑄𝑇, and using the periodic boundary conditions (2.18), we obtain ξ€œΞ©πΊπ›Ώπœ€(β„Ž(π‘₯,𝑇))𝑑π‘₯+π‘Ž0ξ€π‘„π‘‡β„Ž2π‘₯π‘₯ξ€œπ‘‘π‘₯π‘‘π‘‘β‰€Ξ©πΊπ›Ώπœ€ξ€·β„Ž0πœ€ξ€Έπ‘‘π‘₯+𝑐3ξ€œπ‘‡0ξ‚»ξ€œmax1,Ξ©β„Ž2π‘₯ξ‚Ό(π‘₯,𝑑)𝑑π‘₯𝑑𝑑,(A.9) where 𝑐3=π‘Ž1+sup𝑑≀𝑇‖𝑀π‘₯β€–2. Further, from (A.8) and (A.9) we find ξ€œΞ©β„Ž2π‘₯𝑑π‘₯+2𝑐1π‘Ž0ξ€œΞ©πΊπ›Ώπœ€(β„Ž(π‘₯,𝑇))𝑑π‘₯+π‘Ž0ξ€π‘„π‘‡π‘“π›Ώπœ€(β„Ž)β„Ž2π‘₯π‘₯π‘₯β‰€ξ€œπ‘‘π‘₯π‘‘π‘‘Ξ©β„Ž20πœ€,π‘₯𝑑π‘₯+2𝑐1π‘Ž0ξ€œΞ©πΊπ›Ώπœ€ξ€·β„Ž0πœ€ξ€Έπ‘‘π‘₯+𝑐4ξ€œπ‘‡0ξ‚»ξ‚΅ξ€œmax1,Ξ©β„Ž2π‘₯ξ‚Ά(π‘₯,𝑑)𝑑π‘₯πœ…1𝑑𝑑,(A.10) where 𝑐4=2𝑐1𝑐3/π‘Ž0+2𝑐2. Applying the nonlinear GrΓΆnwall lemma [48] to 𝑣(𝑇)≀𝑣(0)+𝑐4βˆ«π‘‡0max{1,π‘£πœ…1(𝑑)}𝑑𝑑 with βˆ«π‘£(𝑑)=(β„Ž2π‘₯(π‘₯,𝑑)+2𝑐1/π‘Ž0πΊπ›Ώπœ€(β„Ž(π‘₯,𝑑)))𝑑π‘₯ yields ξ€œΞ©β„Ž2π‘₯(𝑐π‘₯,𝑑)+21π‘Ž0πΊπ›Ώπœ€(β„Ž(π‘₯,𝑑))𝑑π‘₯≀21/(πœ…1βˆ’1)ξ‚»ξ€œmax1,Ξ©ξ‚΅β„Ž20πœ€,π‘₯(π‘₯)+2𝑐1π‘Ž0πΊπ›Ώπœ€ξ€·β„Ž0πœ€ξ€Έξ‚Άξ‚Ό(π‘₯)𝑑π‘₯=πΎπ›Ώπœ€<∞(A.11) for all π‘‘βˆˆ[0,π‘‡π›Ώπœ€,loc], where π‘‡π›Ώπœ€,loc1∢=2𝑐4ξ€·πœ…1ξ€Έξƒ―ξ‚΅ξ€œβˆ’1min1,Ξ©ξ‚΅β„Ž20πœ€,π‘₯(π‘₯)+2𝑐1π‘Ž0πΊπ›Ώπœ€ξ€·β„Ž0πœ€ξ€Έξ‚Άξ‚Ά(π‘₯)𝑑π‘₯βˆ’(πœ…1βˆ’1)ξƒ°.(A.12) Using the 𝛿→0,β€‰β€‰πœ€β†’0 convergence of the initial data and the choice of πœƒβˆˆ(0,2/5) (see (2.21)) as well as the assumption that the initial data β„Ž0 has finite entropy (2.11), the times π‘‡π›Ώπœ€,loc converge to a positive limit and the upper bound 𝐾 in (A.11) can be taken finite and independent of 𝛿 and πœ€ for 𝛿 and πœ€ sufficiently small. Therefore there exists 𝛿0>0 and πœ€0>0 and 𝐾 such that the bound (A.11) holds for all 0≀𝛿<𝛿0 and 0β‰€πœ€<πœ€0 with 𝐾 replacing πΎπ›Ώπœ€ and for all 0≀𝑑≀𝑇loc9∢=10limπœ€β†’0,𝛿→0π‘‡π›Ώπœ€,loc.(A.13)
Using the uniform bound on βˆ«β„Ž2π‘₯ that (A.11) provides, one can find a uniform-in-𝛿-and-πœ€ bound for the right-hand-side of (A.10) yielding the desired a priori bound (2.23). Similarly, one can find a uniform-in-𝛿-and-πœ€ bound for the right-hand-side of (A.9) yielding the desired a priori bound (2.24). The time 𝑇loc and the constant 𝐾 are determined by 𝛿0,β€‰β€‰πœ€0,β€‰β€‰π‘Ž0,β€‰β€‰π‘Ž1,  sup𝑑≀𝑇‖𝑀π‘₯β€–2, sup𝑑≀𝑇‖𝑀π‘₯β€–βˆž, βˆ«β„Ž0, β€–β„Ž0π‘₯β€–2, and ∫𝐺0(β„Ž0).
To prove the bound (2.25), multiply (2.17) by βˆ’π‘Ž0β„Žπ‘₯π‘₯βˆ’π‘Ž1β„Žβˆ’π‘€, integrate over 𝑄𝑇, integrate by parts, use the periodic boundary conditions (2.18) to find (2.25).

Proof of Lemma 2.4. In the following, we denote the positive, classical solution β„Žπœ€ by β„Ž whenever there is no chance of confusion.
Multiplying (2.17) by (πΊπœ€(𝛼)(β„Ž))ξ…ž, integrating over 𝑄𝑇, taking 𝛿→0, and using the periodic boundary conditions (2.18) yield ξ€œΞ©πΊπœ€(𝛼)(β„Ž(π‘₯,𝑇))𝑑π‘₯+π‘Ž0ξ€π‘„π‘‡β„Žπ›Όβ„Ž2π‘₯π‘₯𝑑π‘₯𝑑𝑑+π‘Ž0𝛼(1βˆ’π›Ό)3ξ€π‘„π‘‡β„Žπ›Όβˆ’2β„Ž4π‘₯=ξ€œπ‘‘π‘₯π‘‘π‘‘Ξ©πΊπœ€(𝛼)ξ€·β„Ž0πœ€ξ€Έπ‘‘π‘₯+π‘Ž1ξ€π‘„π‘‡β„Žπ›Όβ„Ž2π‘₯1𝑑π‘₯π‘‘π‘‘βˆ’ξ€π›Ό+1π‘„π‘‡β„Žπ›Ό+1𝑀π‘₯π‘₯𝑑π‘₯𝑑𝑑.(A.14)Case  1  (0<𝛼<1). The coefficient multiplying βˆ¬β„Žπ›Όβˆ’2β„Ž4π‘₯ in (A.14) is positive and can therefore be used to control the term βˆ¬β„Žπ›Όβ„Ž2π‘₯ on the right-hand side of (A.14). Specifically, using the Cauchy-Schwartz inequality and the Cauchy inequality, π‘Ž1ξ€π‘„π‘‡β„Žπ›Όβ„Ž2π‘₯π‘Žπ‘‘π‘₯𝑑𝑑≀0𝛼(1βˆ’π›Ό)6ξ€π‘„π‘‡β„Žπ›Όβˆ’2β„Ž4π‘₯𝑑π‘₯𝑑𝑑+3π‘Ž212π‘Ž0𝛼(1βˆ’π›Ό)π‘„π‘‡β„Žπ›Ό+2𝑑π‘₯𝑑𝑑.(A.15) Using the bound (A.15) in (A.14), due to (A.5), yields ξ€œΞ©πΊπœ€(𝛼)(β„Ž(π‘₯,𝑇))𝑑π‘₯+π‘Ž0ξ€π‘„π‘‡β„Žπ›Όβ„Ž2π‘₯π‘₯𝑑π‘₯𝑑𝑑+π‘Ž0𝛼(1βˆ’π›Ό)6ξ€π‘„π‘‡β„Žπ›Όβˆ’2β„Ž4π‘₯β‰€ξ€œπ‘‘π‘₯π‘‘π‘‘Ξ©πΊπœ€(𝛼)ξ€·β„Ž0πœ€ξ€Έπ‘‘π‘₯+3π‘Ž212π‘Ž0𝛼(1βˆ’π›Ό)π‘„π‘‡β„Žπ›Ό+2𝑑π‘₯𝑑𝑑+sup𝑑≀𝑇‖‖𝑀π‘₯π‘₯β€–β€–βˆžξ€π›Ό+1π‘„π‘‡β„Žπ›Ό+1β‰€ξ€œπ‘‘π‘₯𝑑𝑑.Ξ©πΊπœ€(𝛼)ξ€·β„Ž0πœ€ξ€Έπ‘‘π‘₯+𝑑1ξ€œπ‘‡0ξƒ―ξ‚΅ξ€œmax1,Ξ©β„Ž2π‘₯𝑑π‘₯𝛼/2+1𝑑𝑑,(A.16) where 𝑑1=𝑏4((3π‘Ž21)/(2π‘Ž0𝛼(1βˆ’π›Ό)))+𝑏4((sup𝑑≀𝑇‖𝑀π‘₯π‘₯β€–βˆž)/(1+𝛼))+𝑏5(((3π‘Ž21)/(2π‘Ž0𝛼(1βˆ’π›Ό)))π‘€πœ€π›Ό+2 + ((sup𝑑≀𝑇‖𝑀π‘₯π‘₯β€–βˆž)/(1+𝛼))π‘€πœ€π›Ό+1). Using the Cauchy inequality in (A.7) and taking 𝛿→0, after applying the Cauchy-Schwartz inequality and (A.5), yields ξ€œΞ©β„Ž2π‘₯𝑑π‘₯+π‘Ž0ξ€π‘„π‘‡π‘“πœ€(β„Ž)β„Ž2π‘₯π‘₯π‘₯ξ€œπ‘‘π‘₯π‘‘π‘‘β‰€Ξ©β„Ž20πœ€,π‘₯+𝑑π‘₯2π‘Ž21π‘Ž0ξ€π‘„π‘‡β„Žπ‘›β„Ž2π‘₯𝑑π‘₯𝑑𝑑+2sup𝑑≀𝑇‖‖𝑀π‘₯β€–β€–2βˆžπ‘Ž0ξ€π‘„π‘‡β„Žπ‘›ξ€œπ‘‘π‘₯π‘‘π‘‘β‰€Ξ©β„Ž20πœ€,π‘₯+π‘Žπ‘‘π‘₯0𝛼(1βˆ’π›Ό)6ξ€π‘„π‘‡β„Žπ›Όβˆ’2β„Ž4π‘₯𝑑π‘₯𝑑𝑑+𝑑2ξ€œπ‘‡0ξƒ―ξ‚΅ξ€œmax1,Ξ©β„Ž2π‘₯𝑑π‘₯𝑛+1βˆ’π›Ό/2𝑑𝑑,(A.17) where 𝑑2=((6π‘Ž41)/(π‘Ž30𝛼(1βˆ’π›Ό)))𝑏4+((2sup𝑑≀𝑇‖𝑀π‘₯β€–2∞)/π‘Ž0)𝑏4+𝑏5(((6π‘Ž41)/(π‘Ž30𝛼(1βˆ’π›Ό)))π‘€πœ€2(𝑛+1)βˆ’π›Ό + ((2sup𝑑≀𝑇‖𝑀π‘₯β€–2∞)/π‘Ž0)π‘€π‘›πœ€). Using (A.16) yields ξ€œΞ©β„Ž2π‘₯(ξ€œπ‘₯,𝑇)𝑑π‘₯+Ξ©πΊπœ€(𝛼)(β„Ž(π‘₯,𝑇))𝑑π‘₯+π‘Ž0ξ€π‘„π‘‡π‘“πœ€(β„Ž)β„Ž2π‘₯π‘₯π‘₯β‰€ξ€œπ‘‘π‘₯π‘‘π‘‘Ξ©β„Ž20πœ€,π‘₯ξ€œπ‘‘π‘₯+Ξ©πΊπœ€(𝛼)ξ€·β„Ž0πœ€ξ€Έπ‘‘π‘₯+𝑑3ξ€œπ‘‡0ξƒ―ξ‚΅ξ€œmax1,Ξ©β„Ž2π‘₯𝑑π‘₯𝑛+1βˆ’π›Ό/2ξƒ°,(A.18) where 𝑑3=𝑑1+𝑑2. Applying the nonlinear GrΓΆnwall lemma [48] to 𝑣(𝑇)≀𝑣(0)+𝑑3βˆ«π‘‡0max{1,𝑣𝑛+1βˆ’π›Ό/2(𝑑)}𝑑𝑑 with βˆ«π‘£(𝑇)=(β„Ž2π‘₯(π‘₯,𝑇)+πΊπœ€(𝛼)(β„Ž(π‘₯,𝑇)))𝑑π‘₯ yields ξ€œΞ©ξ‚€β„Ž2π‘₯(π‘₯,𝑇)+πΊπœ€(𝛼)(ξ‚β„Ž(π‘₯,𝑇))𝑑π‘₯≀41/(2π‘›βˆ’π›Ό)ξ‚»ξ€œmax1,Ξ©ξ‚€β„Ž20πœ€,π‘₯(π‘₯)+πΊπœ€(𝛼)ξ€·β„Ž0,πœ€ξ€Έξ‚ξ‚Ό(π‘₯)𝑑π‘₯=πΎπœ€<∞,(A.19) for all 𝑇: 0≀𝑇≀𝑇(𝛼)πœ€,loc1∢=𝑑3ξƒ―ξ‚΅ξ€œ(2π‘›βˆ’π›Ό)min1,Ξ©ξ‚€β„Ž20πœ€,π‘₯+πΊπœ€(𝛼)ξ€·β„Ž0,πœ€ξ€Έξ‚ξ‚Άπ‘‘π‘₯βˆ’(2π‘›βˆ’π›Ό)/2ξƒ°.(A.20) The bound (A.19) holds for all 0β‰€πœ€<πœ€0 where πœ€0 is from Lemma 2.3 and for all 𝑑≀min{𝑇loc,𝑇(𝛼)πœ€,loc} where 𝑇loc is from Lemma 2.3.
Using the πœ€β†’0 convergence of the initial data and the choice of πœƒβˆˆ(0,2/5) (see (2.21)) as well as the assumption that the initial data β„Ž0 has finite 𝛼-entropy (2.27), the times 𝑇(𝛼)πœ€,loc converge to a positive limit and the upper bound πΎπœ€ in (A.19) can be taken finite and independent of πœ€. Therefore there exists πœ€0 and 𝐾 such that the bound (A.19) holds for all 0β‰€πœ€<πœ€0 with 𝐾 replacing πΎπœ€ and for all 0≀𝑑≀𝑇(𝛼)locξ‚»π‘‡βˆΆ=minloc,910limπœ€β†’0𝑇(𝛼)πœ€,locξ‚Ό,(A.21) where 𝑇loc is the time from Lemma 2.3.
Using the uniform bound on βˆ«β„Ž2π‘₯ that (A.19) provides, one can find a uniform-in-πœ€ bound for the right-hand-side of (A.16) yielding the desired bound ξ€œΞ©πΊπœ€(𝛼)(β„Ž(π‘₯,𝑇))𝑑π‘₯+π‘Ž0ξ€π‘„π‘‡β„Žπ›Όβ„Ž2π‘₯π‘₯𝑑π‘₯𝑑𝑑+π‘Ž0𝛼(1βˆ’π›Ό)6ξ€π‘„π‘‡β„Žπ›Όβˆ’2β„Ž4π‘₯𝑑π‘₯𝑑𝑑≀𝐾1,(A.22) which holds for all 0<πœ€<πœ€0 and all 0≀𝑇≀𝑇(𝛼)loc. Note, (A.22) implies that for all 0<πœ€<πœ€0 that β„Žπœ€π›Ό/2+1 and β„Žπœ€π›Ό/4+1/2 are contained in balls in 𝐿2(0,𝑇;𝐻2(Ξ©)) and 𝐿4(0,𝑇;π‘Š14(Ξ©)) respectively, that is, ξ€π‘„π‘‡ξ€·β„Žπœ€π›Ό/2+1ξ€Έ2π‘₯π‘₯𝑑π‘₯𝑑𝑑≀𝐾,π‘„π‘‡ξ€·β„Žπœ€π›Ό/4+1/2ξ€Έ4π‘₯𝑑π‘₯𝑑𝑑≀𝐾.(A.23) From these estimates follows immediately (2.30).
Case  2 (βˆ’1/2<𝛼<0). For 𝛼<0 the coefficient multiplying βˆ¬β„Žπ›Όβˆ’2β„Ž4π‘₯ in (A.14) is negative. However, we will show that if 𝛼>βˆ’1/2 then one can replace this coefficient with a positive coefficient while also controlling the term βˆ¬β„Žπ›Όβ„Ž2π‘₯ on the right-hand side of (A.14). Using the Cauchy-Schwartz inequality, it is easy to show that ξ€π‘„π‘‡β„Žπ›Όβˆ’2β„Ž4π‘₯9𝑑π‘₯𝑑𝑑≀(1βˆ’π›Ό)2ξ€π‘„π‘‡β„Žπ›Όβ„Ž2π‘₯π‘₯𝑑π‘₯π‘‘π‘‘βˆ€π›Ό<1.(A.24) Using (A.24) in (A.14) yields ξ€œΞ©πΊπœ€(𝛼)(β„Ž(π‘₯,𝑇))𝑑π‘₯+π‘Ž01+2𝛼1βˆ’π›Όπ‘„π‘‡β„Žπ›Όβ„Ž2π‘₯π‘₯β‰€ξ€œπ‘‘π‘₯π‘‘π‘‘Ξ©πΊπœ€(𝛼)ξ€·β„Ž0πœ€ξ€Έπ‘‘π‘₯+π‘Ž1ξ€π‘„π‘‡β„Žπ›Όβ„Ž2π‘₯𝑑π‘₯𝑑𝑑+sup𝑑≀𝑇‖‖𝑀π‘₯π‘₯β€–β€–βˆžξ€π›Ό+1π‘„π‘‡β„Žπ›Ό+1𝑑π‘₯𝑑𝑑.(A.25) Note that if 𝛼>βˆ’1/2 then all the terms on the left-hand side of (A.25) are positive. We now control the term βˆ¬β„Žπ›Όβ„Ž2π‘₯ on the right-hand side of (A.25). By integration by parts and the periodic boundary conditions ξ€π‘„π‘‡β„Žπ›Όβ„Ž2π‘₯1𝑑π‘₯𝑑𝑑=βˆ’ξ€1+π›Όπ‘„π‘‡β„Žπ›Ό+1β„Žπ‘₯π‘₯𝑑π‘₯𝑑𝑑.(A.26) Applying the Cauchy inequality to (A.26) yields π‘Ž1ξ€π‘„π‘‡β„Žπ›Όβ„Ž2π‘₯π‘Žπ‘‘π‘₯𝑑𝑑≀0(1+2𝛼)2(1βˆ’π›Ό)π‘„π‘‡β„Žπ›Όβ„Ž2π‘₯π‘₯π‘Žπ‘‘π‘₯𝑑𝑑+21(1βˆ’π›Ό)2π‘Ž0(1+2𝛼)(1+𝛼)2ξ€π‘„π‘‡β„Žπ›Ό+2𝑑π‘₯𝑑𝑑.(A.27) Using inequality (A.27) in (A.25) yields ξ€œΞ©πΊπœ€(𝛼)(β„Ž(π‘₯,𝑇))𝑑π‘₯+π‘Ž01+2𝛼2(1βˆ’π›Ό)π‘„π‘‡β„Žπ›Όβ„Ž2π‘₯π‘₯ξ€œπ‘‘π‘₯π‘‘π‘‘β‰€Ξ©πΊπœ€(𝛼)ξ€·β„Ž0πœ€ξ€Έ+π‘Žπ‘‘π‘₯21(1βˆ’π›Ό)2π‘Ž0(1+2𝛼)(1+𝛼)2ξ€π‘„π‘‡β„Žπ›Ό+2𝑑π‘₯𝑑𝑑+sup𝑑⩽𝑇‖‖𝑀π‘₯π‘₯β€–β€–βˆžξ€π›Ό+1π‘„π‘‡β„Žπ›Ό+1𝑑π‘₯𝑑𝑑.(A.28) Adding ((π‘Ž0∬(1+2𝛼)(1βˆ’π›Ό))/36)π‘„π‘‡β„Žπ›Όβˆ’2β„Ž4π‘₯𝑑π‘₯𝑑𝑑 to both sides of (A.28) and using the inequality (A.24) yields ξ€œΞ©πΊπœ€(𝛼)(β„Ž(π‘₯,𝑇))𝑑π‘₯+π‘Ž0(1+2𝛼)4(1βˆ’π›Ό)π‘„π‘‡β„Žπ›Όβ„Ž2π‘₯π‘₯+π‘Žπ‘‘π‘₯𝑑𝑑0(1+2𝛼)(1βˆ’π›Ό)36π‘„π‘‡β„Žπ›Όβˆ’2β„Ž4π‘₯ξ€œπ‘‘π‘₯π‘‘π‘‘β‰€Ξ©πΊπœ€(𝛼)ξ€·β„Ž0πœ€ξ€Έπ‘‘π‘₯+𝑒1ξ€œπ‘‡0ξƒ―ξ‚΅ξ€œmax1,Ξ©β„Ž2π‘₯𝑑π‘₯𝛼/2+1𝑑𝑑,(A.29) where 𝑒1=((π‘Ž21(1βˆ’π›Ό))/(2π‘Ž0(1+2𝛼)(1+𝛼)2))𝑏4+((sup𝑑≀𝑇‖𝑀π‘₯π‘₯β€–βˆž)/(𝛼+1))𝑏4 +𝑏5(((π‘Ž21(1βˆ’π›Ό))/(2π‘Ž0(1+2𝛼)(1+𝛼)2))π‘€πœ€π›Ό+2+((sup𝑑≀𝑇‖𝑀π‘₯π‘₯β€–βˆž)/(𝛼+1))π‘€πœ€π›Ό+1). Recall the bound (A.17). As before, by the Cauchy inequality, 2π‘Ž21π‘Ž0ξ€π‘„π‘‡β„Žπ‘›β„Ž2π‘₯π‘Žπ‘‘π‘₯𝑑𝑑≀0(1+2𝛼)(1βˆ’π›Ό)36π‘„π‘‡β„Žπ›Όβˆ’2β„Ž4π‘₯+𝑑π‘₯𝑑𝑑36π‘Ž41π‘Ž30(1+2𝛼)(1βˆ’π›Ό)π‘„π‘‡β„Ž2(𝑛+1)βˆ’π›Όπ‘‘π‘₯𝑑𝑑.(A.30) Using (A.30) in (A.17) yields ξ€œΞ©β„Ž2π‘₯𝑑π‘₯+π‘Ž0ξ€π‘„π‘‡π‘“πœ€(β„Ž)β„Ž2π‘₯π‘₯π‘₯ξ€œπ‘‘π‘₯π‘‘π‘‘β‰€Ξ©β„Ž20πœ€,π‘₯+π‘Žπ‘‘π‘₯0(1+2𝛼)(1βˆ’π›Ό)36π‘„π‘‡β„Žπ›Όβˆ’2β„Ž4π‘₯𝑑π‘₯𝑑𝑑+𝑒2ξ€œπ‘‡0ξƒ―ξ‚΅ξ€œmax1,Ξ©β„Ž2π‘₯𝑑π‘₯𝑛+1βˆ’π›Ό/2𝑑𝑑,(A.31) where 𝑒2=((36π‘Ž41)/(π‘Ž30(1+2𝛼)(1βˆ’π›Ό)))𝑏4+((2sup𝑑≀𝑇‖𝑀π‘₯β€–2∞)/π‘Ž0)𝑏4 + 𝑏5(((36π‘Ž41)/(π‘Ž30(1+2𝛼)(1βˆ’π›Ό)))π‘€πœ€2(𝑛+1)βˆ’π›Ό+((2sup𝑑≀𝑇‖𝑀π‘₯β€–2∞)/π‘Ž0)π‘€π‘›πœ€). Using (A.29) yields ξ€œΞ©β„Ž2π‘₯(ξ€œπ‘₯,𝑇)𝑑π‘₯+Ξ©πΊπœ€(𝛼)(β„Ž(π‘₯,𝑇))𝑑π‘₯+π‘Ž0ξ€π‘„π‘‡π‘“πœ€(β„Ž)β„Ž2π‘₯π‘₯π‘₯β‰€ξ€œπ‘‘π‘₯π‘‘π‘‘Ξ©β„Ž20πœ€,π‘₯ξ€œπ‘‘π‘₯+Ξ©πΊπœ€(𝛼)ξ€·β„Ž0πœ€ξ€Έπ‘‘π‘₯+𝑒3ξ€œπ‘‡0ξƒ―ξ‚΅ξ€œmax1,Ξ©β„Ž2π‘₯𝑑π‘₯𝑛+1βˆ’π›Ό/2ξƒ°,(A.32) where 𝑒3=𝑒1+𝑒2. The rest of the proof now continues as in the 0<𝛼<1 case. Specifically, one finds a bound ξ€œΞ©ξ‚€β„Ž2π‘₯(π‘₯,𝑇)+πΊπœ€(𝛼)(ξ‚β„Ž(π‘₯,𝑇))𝑑π‘₯≀41/(2π‘›βˆ’π›Ό)ξ‚»ξ€œmax1,Ξ©ξ‚€β„Ž20πœ€,π‘₯(π‘₯)+πΊπœ€(𝛼)ξ€·β„Ž0πœ€ξ€Έξ‚ξ‚Ό(π‘₯)𝑑π‘₯=πΎπœ€<∞(A.33) for all 𝑇: 0≀𝑇≀𝑇(𝛼)πœ€,loc1∢=𝑒3ξƒ―ξ‚΅ξ€œ(2π‘›βˆ’π›Ό)min1,Ξ©ξ‚€β„Ž20πœ€,π‘₯(π‘₯)+πΊπœ€(𝛼)ξ€·β„Ž0,πœ€ξ€Έξ‚ξ‚Ά(π‘₯)𝑑π‘₯βˆ’(2π‘›βˆ’π›Ό)/2ξƒ°.(A.34) The time 𝑇(𝛼)loc is defined as in (A.21) and the uniform bound (A.33) used to bound the right-hand side of (A.29) yields the desired bound ξ€œΞ©πΊπœ€(𝛼)(π‘Žβ„Ž(π‘₯,𝑇))𝑑π‘₯+0(1+2𝛼)4(1βˆ’π›Ό)π‘„π‘‡β„Žπ›Όβ„Ž2π‘₯π‘₯π‘Žπ‘‘π‘₯𝑑𝑑+0(1+2𝛼)(1βˆ’π›Ό)36π‘„π‘‡β„Žπ›Όβˆ’2β„Ž4π‘₯𝑑π‘₯𝑑𝑑≀𝐾2.(A.35)

Proof of Lemma 3.4. Let πœ™(π‘₯)=𝜁6(π‘₯). Multiplying (2.17) by βˆ’(πœ™(π‘₯)β„Žπ‘₯)π‘₯, and integrating on 𝑄𝑇, yields 12ξ€œΞ©(π‘₯)β„Ž2π‘₯(1π‘₯,𝑇)𝑑π‘₯πœ™βˆ’2ξ€œΞ©πœ™(π‘₯)β„Ž20πœ€,π‘₯(π‘₯)𝑑π‘₯=βˆ’π‘„π‘‡π‘“πœ€ξ€·π‘Ž(β„Ž)0β„Žπ‘₯π‘₯π‘₯+π‘Ž1β„Žπ‘₯+𝑀π‘₯πœ™ξ€Έξ€·π‘₯π‘₯β„Žπ‘₯+2πœ™π‘₯β„Žπ‘₯π‘₯+πœ™β„Žπ‘₯π‘₯π‘₯𝑑π‘₯𝑑𝑑=βˆ’π‘„π‘‡π‘“πœ€ξ€·π‘Ž(β„Ž)0β„Žπ‘₯π‘₯π‘₯+π‘Ž1β„Žπ‘₯ξ€Έπœ™π‘₯π‘₯β„Žπ‘₯𝑑π‘₯π‘‘π‘‘βˆ’2π‘„π‘‡π‘“πœ€ξ€·π‘Ž(β„Ž)0β„Žπ‘₯π‘₯π‘₯+π‘Ž1β„Žπ‘₯ξ€Έπœ™π‘₯β„Žπ‘₯π‘₯βˆ’ξ€π‘‘π‘₯π‘‘π‘‘π‘„π‘‡π‘“πœ€ξ€·π‘Ž(β„Ž)0β„Žπ‘₯π‘₯π‘₯+π‘Ž1β„Žπ‘₯ξ€Έπœ™β„Žπ‘₯π‘₯π‘₯𝑑π‘₯π‘‘π‘‘βˆ’π‘„π‘‡π‘“πœ€(β„Ž)𝑀π‘₯ξ€·πœ™π‘₯π‘₯β„Žπ‘₯+2πœ™π‘₯β„Žπ‘₯π‘₯+πœ™β„Žπ‘₯π‘₯π‘₯𝑑π‘₯π‘‘π‘‘βˆ’π‘Ž0ξ€π‘„π‘‡π‘“πœ€(β„Ž)β„Ž2π‘₯π‘₯π‘₯πœ™π‘‘π‘₯𝑑𝑑=∢𝐼1+𝐼2+𝐼3+𝐼4+𝐼5.(A.36) We now bound the terms 𝐼1,  𝐼2,  𝐼3, and 𝐼4. First, 𝐼1=βˆ’π‘Ž0ξ€π‘„π‘‡πœ™π‘₯π‘₯π‘“πœ€(β„Ž)β„Žπ‘₯π‘₯π‘₯β„Žπ‘₯𝑑π‘₯π‘‘π‘‘βˆ’π‘Ž1ξ€π‘„π‘‡πœ™π‘₯π‘₯π‘“πœ€(β„Ž)β„Ž2π‘₯𝑑π‘₯π‘‘π‘‘β‰€πœ–1ξ€π‘„π‘‡πœ6ξ€½π‘“πœ€(β„Ž)β„Ž2π‘₯π‘₯π‘₯+β„Žπ‘›βˆ’4β„Ž6π‘₯ξ€Ύξ€·πœ–π‘‘π‘₯𝑑𝑑+𝐢1ξ€Έξ€π‘„π‘‡β„Žπ‘›+2ξ‚€πœ6+𝜁6π‘₯+𝜁3||𝜁π‘₯π‘₯||3𝐼𝑑π‘₯𝑑𝑑,(A.37)2=βˆ’2π‘Ž0ξ€π‘„π‘‡πœ™π‘₯π‘“πœ€(β„Ž)β„Žπ‘₯π‘₯π‘₯β„Žπ‘₯π‘₯𝑑π‘₯π‘‘π‘‘βˆ’2π‘Ž1ξ€π‘„π‘‡πœ™π‘₯π‘“πœ€(β„Ž)β„Žπ‘₯π‘₯β„Žπ‘₯𝑑π‘₯π‘‘π‘‘β‰€πœ–2ξ€π‘„π‘‡πœ6ξ‚†π‘“πœ€(β„Ž)β„Ž2π‘₯π‘₯π‘₯+β„Žπ‘›βˆ’2β„Ž2π‘₯β„Ž2π‘₯π‘₯+β„Žπ‘›βˆ’1||β„Žπ‘₯π‘₯||3ξ‚‡ξ€·πœ–π‘‘π‘₯𝑑𝑑+𝐢2ξ€Έξ€π‘„π‘‡β„Žπ‘›+2ξ€·πœ6+𝜁6π‘₯𝐼𝑑π‘₯𝑑𝑑,(A.38)3=βˆ’π‘Ž0ξ€π‘„π‘‡πœ™π‘“πœ€(β„Ž)β„Ž2π‘₯π‘₯π‘₯𝑑π‘₯π‘‘π‘‘βˆ’π‘Ž1ξ€π‘„π‘‡πœ™π‘“πœ€(β„Ž)β„Žπ‘₯π‘₯π‘₯β„Žπ‘₯𝑑π‘₯π‘‘π‘‘β‰€βˆ’π‘Ž0ξ€π‘„π‘‡πœ6π‘“πœ€(β„Ž)β„Ž2π‘₯π‘₯π‘₯𝑑π‘₯𝑑𝑑+πœ–3ξ€π‘„π‘‡πœ6ξ€·π‘“πœ€(β„Ž)β„Ž2π‘₯π‘₯π‘₯+β„Žπ‘›βˆ’4β„Ž6π‘₯ξ€Έξ€·πœ–π‘‘π‘₯𝑑𝑑+𝐢3ξ€Έξ€π‘„π‘‡β„Žπ‘›+2𝜁6𝐼𝑑π‘₯𝑑𝑑,(A.39)4=βˆ’6π‘„π‘‡π‘“πœ€(β„Ž)β„Žπ‘₯𝑀π‘₯𝜁4ξ€·5𝜁2π‘₯+𝜁𝜁π‘₯π‘₯𝑑π‘₯π‘‘π‘‘βˆ’12π‘„π‘‡π‘“πœ€(β„Ž)β„Žπ‘₯π‘₯𝑀π‘₯𝜁5𝜁π‘₯𝑑π‘₯π‘‘π‘‘βˆ’π‘„π‘‡π‘“πœ€(β„Ž)β„Žπ‘₯π‘₯π‘₯𝑀π‘₯𝜁6𝑑π‘₯π‘‘π‘‘β‰€πœ–4ξ€π‘„π‘‡πœ6ξ‚†π‘“πœ€(β„Ž)β„Ž2π‘₯π‘₯π‘₯+β„Žπ‘›βˆ’4β„Ž6π‘₯+β„Žπ‘›βˆ’1||β„Žπ‘₯π‘₯||3ξ‚‡ξ€·πœ–π‘‘π‘₯𝑑𝑑+𝐢4ξ€Έξ€π‘„π‘‡β„Žπ‘›+2ξ€·πœ6π‘₯+𝜁3𝜁3π‘₯π‘₯ξ€Έξ€·πœ–π‘‘π‘₯𝑑𝑑+𝐢4ξ€Έξ€π‘„π‘‡β„Žπ‘›πœ6𝑑π‘₯𝑑𝑑.(A.40)
Now, multiplying (2.17) by 𝜁4(β„Ž+𝛾)𝛽, 𝛽>(1βˆ’π‘›)/3,𝛾>0 and integrating on 𝑄𝑇, using the Young's inequality, letting 𝛾→0, we obtain the following estimate: ξ€œΞ©πœ4β„Žπ›½+1(ξ€œπ‘‡)𝑑π‘₯β‰€Ξ©πœ4β„Žπ›½+10πœ€π‘‘π‘₯+πœ–4ξ€π‘„π‘‡πœ6ξ€½π‘“πœ€(β„Ž)β„Ž2π‘₯π‘₯π‘₯+β„Žπ‘›βˆ’4β„Ž6π‘₯ξ€Ύξ€·πœ–π‘‘π‘₯𝑑𝑑+𝐢4ξ€Έξ€π‘„π‘‡ξ€½πœ’{𝜁>0}β„Žπ‘›+3π›½βˆ’1+β„Žπ‘›+2ξ€·πœ6+𝜁6π‘₯ξ€Έ+β„Žπ‘›πœ6𝑑π‘₯𝑑𝑑,(A.41) where 𝛽>(1βˆ’π‘›)/3. If we now add inequalities (A.36) and (A.41), in view of (A.37)–(A.39), then, applying Lemma B.1, choosing πœ–π‘–>0, and letting πœ€β†’0, we obtain (3.17).

B. Auxiliary Lemmas

Lemma B.1 (see [34, 38]). Let Ξ©βŠ‚β„π‘,𝑁<6, be a bounded convex domain with smooth boundary, and let βˆšπ‘›βˆˆ(2βˆ’1βˆ’π‘/(𝑁+8),3) for 𝑁>1, and 1/2<𝑛<3 for 𝑁=1. Then the following estimates hold for any strictly positive functions π‘£βˆˆπ»2(Ξ©) such that βˆ‡π‘£β‹…β†’π‘›=0 on πœ•Ξ© and βˆ«Ξ©π‘£π‘›|βˆ‡Ξ”π‘£|2<∞: ξ€œΞ©πœ‘6ξ‚†π‘£π‘›βˆ’4||||βˆ‡π‘£6+π‘£π‘›βˆ’2||𝐷2𝑣||2||||βˆ‡π‘£2ξ‚‡ξ‚»ξ€œβ‰€π‘Ξ©πœ‘6𝑣𝑛||||βˆ‡Ξ”π‘£2+ξ€œ{πœ‘>0}𝑣𝑛+2||||βˆ‡πœ‘6ξ‚Ό,ξ€œΞ©πœ‘6||βˆ‡Ξ”π‘£(𝑛+2)/2||2ξ‚»ξ€œβ‰€π‘Ξ©πœ‘6𝑣𝑛||||βˆ‡Ξ”π‘£2+ξ€œ{πœ‘>0}𝑣𝑛+2||||βˆ‡πœ‘6+πœ‘2||𝐷2πœ‘||2||||βˆ‡πœ‘2+πœ‘3||||Ξ”πœ‘3,(B.1) where πœ‘βˆˆπΆ2(Ξ©) is an arbitrary nonnegative function such that the tangential component of βˆ‡πœ‘ is equal to zero on πœ•Ξ©, and the constant 𝑐>0 is independent of 𝑣.

Lemma B.2 (see [49]). If Ξ©βŠ‚β„π‘ is a bounded domain with piecewise-smooth boundary, π‘Ž>1, π‘βˆˆ(0,π‘Ž), 𝑑>1, and 0≀𝑖<𝑗,𝑖,π‘—βˆˆβ„•, then there exist positive constants 𝑑1 and 𝑑2  (𝑑2=0ifΞ© is unbounded) depending only on Ξ©, 𝑑, 𝑗, 𝑏, and 𝑁 such that the following inequality is valid for every 𝑣(π‘₯)βˆˆπ‘Šπ‘—,𝑑(Ξ©)βˆ©πΏπ‘(Ξ©): β€–β€–π·π‘–π‘£β€–β€–πΏπ‘Ž(Ξ©)≀𝑑1‖𝐷𝑗𝑣||πœƒπΏπ‘‘(Ξ©)‖𝑣‖𝐿1βˆ’πœƒπ‘(Ξ©)+𝑑2‖𝑣‖𝐿𝑏(Ξ©),πœƒ=1/𝑏+𝑖/π‘βˆ’1/π‘Žβˆˆξ‚Έπ‘–1/𝑏+𝑗/π‘βˆ’1/𝑑𝑗,1.(B.2)

Lemma B.3 (see [37]). Let (𝛽1,…,π›½π‘š)βˆˆβ„π‘š,π‘šβ‰₯1 and let βˆπ›½=π‘šπ‘—=1𝛽𝑗,𝛽𝑖=𝛽/𝛽𝑖=βˆπ‘šπ‘—=1,𝑗≠𝑖𝛽𝑗. Assume that 𝐺𝑖(𝑠) are nonnegative nonincreasing functions satisfying the conditions: 𝐺𝑖(𝑠+𝛿)β‰€π‘π‘–ξƒ©π‘šξ“π‘–=1𝐺𝑖(𝑠)𝛿𝛼𝑖ξƒͺπ›½π‘–βˆ€π‘ >0,𝛿>0,𝑖=1,π‘š(B.3) with real constants 𝑐𝑖>0,𝛽𝑖>1, and 𝛼𝑖β‰₯0 for 𝑖=1,π‘š, and 𝛼𝑖>0 for 𝑖=1,β„“. Let βˆ‘πΊ(𝑠)=π‘šπ‘–=1(𝑐𝛽𝑖𝑖)(𝐺𝑖(𝑠))𝛽𝑖, and let the function 𝐻(𝑠)=π‘šπ›½βˆ‘π‘šπ‘–=β„“+1𝑐𝛽𝑖𝑖(𝑐𝛽𝑖𝑖)1βˆ’π›½π‘–(𝐺𝑖(𝑠))π›½π‘–βˆ’1 be such that 𝐻(𝑠1)<1 at a some 𝑠1β‰₯0. Then there exists a positive constant 𝑐>1 depending on π‘š,𝛼𝑖,𝛽𝑖,β„“, and 𝐻(𝑠1) such that 𝐺𝑖(𝑠0)≑0 for all 𝑖=1,β„“, where 𝑠0=𝑠1βˆ‘+𝑐ℓ𝑖=1(𝑐𝛽𝑖𝑖(𝑐𝛽𝑖𝑖)1βˆ’π›½π‘–(𝐺(𝑠1))π›½π‘–βˆ’1)1/(𝛼𝑖𝛽). Note, if β„“=π‘š then 𝑠1=0.

Lemma B.4. Let (𝛽1,…,π›½π‘š)βˆˆβ„π‘š,π‘šβ‰₯1, and let βˆπ›½=π‘šπ‘—=1𝛽𝑗, 𝛽𝑖=𝛽/𝛽𝑖=βˆπ‘šπ‘—=1,𝑗≠𝑖𝛽𝑗. Assume that 𝐺𝑖(𝑠), 𝑔(𝑠) are nonnegative nonincreasing functions satisfying the conditions: 𝐺𝑖(𝑠+𝛿)β‰€π‘π‘–ξƒ©π‘šξ“π‘–=1𝐺𝑖(𝑠)𝛿𝛼𝑖ξƒͺ+𝑔(𝑠)π›½π‘–βˆ€π‘ βˆˆβ„1,𝛿>0,𝑖=1,π‘š(B.4) with real constants 𝑐𝑖>0,𝛽𝑖>1, and 𝛼𝑖>0. Let the functions 𝐺max(𝑠)∢=max𝑖=1,π‘šβŽ§βŽͺ⎨βŽͺβŽ©π‘šπ‘02π›½ξƒ©π‘šξ“π‘˜=1ξ€·πΊπ‘˜ξ€Έ(𝑠)π›½π‘˜ξƒͺπ›½π‘–βˆ’1⎫βŽͺ⎬βŽͺ⎭(𝑠)1/𝛼𝑖𝛽,𝑐0=2π‘šπ›½βˆ’1ξ“π‘˜=1ξ€·π‘π‘˜ξ€Έπ›½π‘˜(B.5) and 𝑔max(𝑠)∢=max𝑖=1,π‘š(π‘š2𝛽)1/𝛼𝑖𝛽(2π›½βˆ’1βˆ‘π‘šπ‘˜=1(π‘π‘˜)π›½π‘˜)𝛽𝑖/𝛼𝑖(𝑔(𝑠))(π›½π‘–βˆ’1)/𝛼𝑖 be such that(i) for some 𝑠1∈(βˆ’βˆž,𝑠0) the inequality 𝐺max(𝑠)β‰€π‘˜1𝑔max(𝑠) holds for all 𝑠<𝑠1,(ii)𝑔max(𝑠)β‰€π‘˜2(𝑠0βˆ’π‘ ) for all 𝑠≀𝑠0,  where π‘˜1>(1βˆ’max𝑖=1,π‘š{2βˆ’(π›½π‘–βˆ’1)/(𝛼𝑖𝛽)})βˆ’1 and 0<π‘˜2<π‘˜1βˆ’1(1βˆ’π‘˜1βˆ’1βˆ’max𝑖=1,π‘š{2βˆ’(π›½π‘–βˆ’1)/(𝛼𝑖𝛽)}). Then 𝐺𝑖(𝑠)≑0 for all 𝑠β‰₯𝑠0.

Proof. Let us denote by βˆ‘πΊ(𝑠)∢=π‘šπ‘˜=1(πΊπ‘˜(𝑠))π›½π‘˜. Raising both side of (B.4) to the power 𝛽𝑖 and summing with respect to 𝑖, we deduce 𝐺(𝑠+𝛿)β‰€π‘šξ“π‘˜=1ξ€·π‘π‘˜ξ€Έπ›½π‘˜ξƒ©π‘šξ“π‘–=1𝐺𝑖(𝑠)𝛿𝛼𝑖ξƒͺ+𝑔(𝑠)𝛽≀𝑐02π‘šπ›½βˆ’1𝑖=1𝐺𝛽𝑖(𝑠)𝛿𝛼𝑖𝛽+𝑐0𝑔𝛽(𝑠)≀𝑐02π‘šπ›½βˆ’1𝑖=1𝐺𝛽𝑖(𝑠)𝛿𝛼𝑖𝛽+𝑐0𝑔𝛽(𝑠).(B.6) Choosing βˆ‘π›Ώ=𝛿(𝑠)=π‘šπ‘–=1(π‘šπ‘02π›½πΊπ›½π‘–βˆ’1(𝑠))1/(𝛼𝑖𝛽), we arrive at 1𝐺(𝑠+𝛿(𝑠))≀2𝐺(𝑠)+𝑐0𝑔𝛽(𝑠),(B.7) whence we find that 𝛿(𝑠+𝛿(𝑠))β‰€πœ–π›Ώ(𝑠)+̃𝑔(𝑠),(B.8) where πœ–=max𝑖=1,π‘š{2βˆ’(π›½π‘–βˆ’1)/(𝛼𝑖𝛽)}, βˆ‘Μƒπ‘”(𝑠)∢=π‘šπ‘–=1(π‘šπ‘π›½π‘–02𝛽)1/(𝛼𝑖𝛽)(𝑔(𝑠))(π›½π‘–βˆ’1)/𝛼𝑖. Applying [27, Lemma  4] to 𝛿(𝑠), taking into account the conditions (i) and (ii), we obtain 𝛿(𝑠)=0 for all 𝑠β‰₯𝑠0.

Acknowledgment

The research of Dr. R. Taranets leading to these results has received funding from the European Community’s Seventh Framework Programme (FP7/2007-2013) under Grant Agreement no PIIF-GA-2009-254521β€”[TFE].