Next Article in Journal
Linear Convergence of Split Equality Common Null Point Problem with Application to Optimization Problem
Next Article in Special Issue
Topology Optimization of Elastoplastic Behavior Conditions by Selectively Suppressing Plastic Work
Previous Article in Journal
Extreme Value Index Estimation by Means of an Inequality Curve
Previous Article in Special Issue
The Effect of a Hyperbolic Two-Temperature Model with and without Energy Dissipation in a Semiconductor Material
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Note on the Solutions for a Higher-Order Convective Cahn–Hilliard-Type Equation

by
Giuseppe Maria Coclite
1,*,† and
Lorenzo di Ruvo
2,†
1
Dipartimento di Meccanica, Matematica e Management, Politecnico di Bari, 70125 Bari, Italy
2
Dipartimento di Matematica, Università di Bari, 70125 Bari, Italy
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Mathematics 2020, 8(10), 1835; https://doi.org/10.3390/math8101835
Submission received: 1 September 2020 / Revised: 5 October 2020 / Accepted: 7 October 2020 / Published: 19 October 2020
(This article belongs to the Special Issue Applied Mathematics and Solid Mechanics)

Abstract

:
The higher-order convective Cahn-Hilliard equation describes the evolution of crystal surfaces faceting through surface electromigration, the growing surface faceting, and the evolution of dynamics of phase transitions in ternary oil-water-surfactant systems. In this paper, we study the H 3 solutions of the Cauchy problem and prove, under different assumptions on the constants appearing in the equation and on the mean of the initial datum, that they are well-posed.

1. Introduction

In this paper, we study the well-posedness of the Cauchy problem:
t u + κ x u 2 + γ x 2 u β 2 x 6 u + α x 4 u + δ 2 x 4 u 3 = 0 , t > 0 , x R , u ( 0 , x ) = u 0 ( x ) , x R ,
with
κ , γ , β , α , δ R , γ < 0 , α 0 , κ , β , δ 0 , or ,
β , α , δ R , κ = γ = 0 , α 0 , β , δ 0 , or ,
κ , γ , β , α , δ R , β 0 , δ = 0 , or ,
κ , γ , β , α , δ R , β , δ 0 .
On the initial datum, we assume
u 0 H 3 ( R ) , or ,
u 0 H 3 ( R ) L 1 ( R ) , R u 0 ( x ) d x = 0 .
Inspired by [1,2,3,4,5,6,7,8,9,10,11,12], in light of (7), we define the following function:
P 0 ( x ) = x u 0 ( y ) d y ,
on which we assume
P 0 L 2 ( R ) 2 = R x u 0 ( y ) d y 2 d x < .
The equation in (1) is derived in [13] to model the evolution of a crystalline surface with small slopes that undergoes faceting. The unknown u gives the surface slope, the constant κ is proportional to the atomic flux deposition strength and the convective term κ x u 2 arises from the deposited atoms normal impingement. The sixth-order linear term x 6 u regularizes the equation, taking into account the surface curvature and the anisotropy of the surface energy under the surface diffusion.
From a mathematical point of view, the existence and uniqueness of weak solutions of (1) with periodic boundary conditions is proven in [14], under the assumptions κ > 0 and γ = 0 . In the same setting, a similar result is proven in two space dimensions in [15]. In [16], the authors derived the stationary solutions of (1), again assuming κ > 0 and γ = 0 . In [17], the existence of a global-in-time attractor is studied, while the well-posedness of the classical solutions of (1) is proven in [18], requiring (7)–(9), and γ = 0 . In this paper, we will prove that, if (2) or (3) hold, we have the well-posedness of the classical solutions of (1) assuming (6), while if (5) holds, we have the well-posedness of (1) assuming (7)–(9).
Taking κ = 0 , (1) becomes
t u + γ x 2 u β 2 x 6 u + α x 4 u + δ 2 x 4 u 3 = 0 ,
which is a Cahn-Hilliard type equation [19,20,21]. It was deduced in [22] to describe the evolution of crystal surfaces faceting through surface electromigration. It also describes the phase transition development in ternary oil-water-surfactant systems. One part of the surfactant is hydrophilic, and the other one (termed amphiphile) is lipophilic. Oil, water, and microemulsion (i.e., a homogeneous, isotropic mixture of oil and water) can coexist in equilibrium. The unknown u, in (10), gives the local difference between oil and water concentrations.
From a mathematical point of view, in [23] the initial-boundary-value problem for (10) is analyzed, under appropriate assumptions on γ , β , α , δ . In [24], the authors analyze the existence of a global-in-time attractor. The existence of weak solutions for the initial-boundary-value problem for (10) is proven in the case of degenerate mobility in [25]. Finally, in [18], the well-posedness of the classical solution of the Cauchy problem of (10) is proven, assuming (7)–(9), with γ = 0 . In this paper, we will show that the classical solutions of the Cauchy problem of (10) are well-posed, assuming (6), if γ 0 and α > 0 , while in the general case, we will prove the same result assuming (7)–(9).
Observe that in [13], it is proven that, as κ , (1) reduces
t u + x u 2 + a 1 x 2 u b 1 2 x 6 u + c 1 x 4 u = 0 ,
which is known as the Kuramoto-Sivashinsky equation (see [26,27,28]). In Section 4, we will prove the well-posedness of the Cauchy problem for (11), assuming (6).
When β = δ = 0 and α = f 2 0 , (1) reads
t u + κ x u 2 + γ x 2 u + f 2 x 4 u = 0 .
(12) appears in several physical situations; for example, it models long waves on a viscous fluid flowing down an inclined plane [29] and drift waves in a plasma [30]. (12) was also independently deduced by Kuramoto [27,31,32] to describe the phase turbulence in reaction-diffusion systems, and by Sivashinsky [28] to describe plane flame propagation, taking into account the combined influence of diffusion and thermal conduction of the gas on the stability of a plane flame front.
Equation (12) can be used to study incipient instabilities in several physical and chemical systems [33,34,35]. Moreover, (12) is also termed the Benney-Lin equation [36,37], and was deduced by Kuramoto as a model for phase turbulence in the Belousov-Zhabotinsky reaction [38].
The existence and the dynamical properties of the exact solutions for (12) can be found in [39,40,41,42,43,44]. The control problem for (12) with periodic boundary conditions, and on a bounded interval, are studied in [45,46,47]. The problem of global-in-time exponential stabilization of (12) with periodic boundary conditions is analyzed in [48]. A generalization of the optimal control theory for (12) is proposed in [49], while the f global boundary controllability of (12) is considered in [50]. The existence of solitonic solutions for (12) is proven in [51]. The well-posedness of the Cauchy problem for (12) is proven in [52,53,54], using the energy space technique, a priori estimates together with an application of the Cauchy-Kovalevskaya and the fixed point method, respectively. Instead, the initial-boundary value problem for (1) is studied, using a priori estimates together with an application of the Cauchy-Kovalevskaya, and the energy space technique in [55,56,57]. Inspired by [58,59,60], the convergence of the solution of (12) to the unique entropy one of the Burgers equation is proven in [61].
Finally, due to its general structure, we conjecture that (1) can have a possible application in machine learning (see [62,63]).

2. Results and Organization of the Paper

In this paper, we improve and complete the results of [14,16,17,18] working with H 3 initial data and having general assumptions on the constants appearing in (1). The main result of this paper is the following theorem. We prove the global-in-time existence, uniqueness, and stability of the solutions of the Cauchy problem (1).
Theorem 1.
Fix T > 0 . Assuming one of the following
(i)
(2) and (6),
( i i )
(3) and (6),
( i i i )
(4) and (6),
( i v )
(5) and (7),
and (9), there exists a unique solution u of (1) such that
u H 1 ( ( 0 , T ) × R ) L ( 0 , T ; H 3 ( R ) ) .
In particular, under the Assumptions (7) and (9), we have that
R u ( t , x ) d x = 0 .
Moreover, if u 1 and u 2 are two solutions of (1), we have that
u 1 ( t , · ) u 2 ( t , · ) L 2 ( R ) e C ( T ) t u 1 , 0 u 2 , 0 L 2 ( R ) ,
for some suitable C ( T ) > 0 , and every 0 t T .
The well-posedness of (1) is guaranteed for a short time by the Cauchy-Kowaleskaya Theorem [64]. The solutions are indeed global, thanks to suitable a priori estimates.
The paper is organized as follows. In Section 3 we prove Theorem 1, assuming (i) or ( i i ). In Section 4 we prove Theorem 1, assuming ( i i i ). In Section 5 we prove Theorem 1, assuming ( i v ).

3. Proof of Theorem 1 Assuming ( i ) or ( ii )

In this section, we prove Theorem 1, assuming (i) or ( i i ). For the sake of notational simplicity, define
γ = a 2 , α = b 2 ,
and then (1) reads
t u + κ x u 2 a 2 x 2 u β 2 x 6 u + b 2 x 4 u + δ 2 x 4 u 3 = 0 , t > 0 , x R , u ( 0 , x ) = u 0 ( x ) , x R .
Since the short time well-posedness of (17) is guaranteed by the Cauchy-Kowaleskaya Theorem [64], here we need to prove some suitable global a priori estimates.
From now on, we denote with C 0 the constants which depend only on the initial data, and with C ( T ) , the constants which depend also on T.
Following [65] (Lemma 2.2 ), we begin with the following energy estimate in the space L ( 0 , ; H 1 ( R ) ) L 2 ( 0 , ; H 2 ( R ) ) .
Lemma 1.
Assuming (2), for each t > 0 , we have that
β 2 2 x u ( t , · ) L 2 ( R ) 2 + δ 2 4 u ( t , · ) L 4 ( R ) 4 + b 2 2 u ( t , · ) L 2 ( R ) 2 + K 1 2 0 t x 2 u ( s , · ) L 2 ( R ) 2 d s + K 2 2 0 t u ( s , · ) x u ( s , · ) L 2 ( R ) 2 d s + a 2 b 2 0 t x u ( s , · ) L 2 ( R ) 2 d s + 0 t R b 2 x 2 u + δ 2 x 2 u 3 β 2 x 4 u 2 d s d x C 0 ,
where K 1 2 , K 2 2 are two appropriate positive constants.
Assuming (3), for each t > 0 , we have that
β 2 2 x u ( t , · ) L 2 ( R ) 2 + δ 2 4 u ( t , · ) L 4 ( R ) 4 + b 2 2 u ( t , · ) L 2 ( R ) 2 + 0 t R b 2 x 2 u + δ 2 x 2 u 3 β 2 x 4 u 2 d s d x C 0 .
Moreover, there exists C 0 > 0 , independent on κ , a , b , such that, for each t 0 ,
u ( t , · ) L ( R ) C 0 .
Proof. 
We begin by proving (18). Assume (2). Multiplying (17) by β 2 x 2 u + δ 2 u 3 + b 2 u , we have that
β 2 x 2 u + δ 2 u 3 + b 2 u t u + 2 κ β 2 x 2 u + δ 2 u 3 + b 2 u u x u a 2 β 2 x 2 u + δ 2 u 3 + b 2 u x 2 u β 2 β 2 x 2 u + δ 2 u 3 + b 2 u x 6 u + b 2 β 2 x 2 u + δ 2 u 3 + b 2 u x 4 u + δ 2 β 2 x 2 u + δ 2 u 3 + b 2 u x 4 u 3 = 0 .
Performing some integration by parts, we gain
R β 2 x 2 u + δ 2 u 3 + b 2 u t u d x = d d t β 2 2 x u ( t , · ) L 2 ( R ) 2 + δ 2 4 u ( t , · ) L 4 ( R ) 4 + b 2 2 u ( t , · ) L 2 ( R ) 2 , 2 κ R β 2 x 2 u + δ 2 u 3 + b 2 u u x u d x = 2 κ β 2 R u x u x 2 u d x , a 2 R β 2 x 2 u + δ 2 u 3 + b 2 u x 2 u d x = a 2 β 2 x 2 u ( t , · ) L 2 ( R ) 2 + 3 a 2 δ 2 u ( t , · ) x u ( t , · ) L 2 ( R ) 2 + a 2 b 2 x u ( t , · ) L 2 ( R ) 2 , β 2 R β 2 x 2 u + δ 2 u 3 + b 2 u x 6 u d x = β 4 R x 3 u x 5 u d x + β 2 δ 2 R x u 3 x 5 u d x + β 2 b 2 R x u x 5 u d x = β 4 R ( x 4 u ) 4 d x β 2 δ 2 R x 2 u 3 x 4 u d x β 2 a 2 R x 2 u x 4 u d x , b 2 R β 2 x 2 u + δ 2 u 3 + b 2 u x 4 u d x = β 2 b 2 R x 2 u x 4 u d x δ 2 b 2 R x u 3 x 3 u d x b 4 R x u x 3 u d x = β 2 b 2 R x 2 u x 4 u d x + δ 2 b 2 R x 2 u 3 x 2 u d x + b 4 R ( x 2 u ) 2 d x , δ 2 R β 2 x 2 u + δ 2 u 3 + b 2 u x 4 u 3 d x = β 2 δ 2 R x 3 u x 3 u 3 d x δ 4 R x u 3 x 3 u 3 d x b 2 δ 2 R x u x 3 u 3 d x = β 2 δ 2 R x 4 u x 2 u 3 d x + δ 4 R x 2 u 3 2 d x + b 2 δ 2 R x 2 u x 2 u 3 d x .
Therefore, thanks to (22), an integration of (21) on R gives
d d t β 2 2 x u ( t , · ) L 2 ( R ) 2 + δ 2 4 u ( t , · ) L 4 ( R ) 4 + b 2 2 u ( t , · ) L 2 ( R ) 2 + a 2 β 2 x 2 u ( t , · ) L 2 ( R ) 2 + 3 a 2 δ 2 u ( t , · ) x u ( t , · ) L 2 ( R ) 2 + a 2 b 2 x u ( t , · ) L 2 ( R ) 2 + β 4 R ( x 4 u ) 4 d x 2 β 2 δ 2 R x 2 u 3 x 4 u d x 2 β 2 a 2 R x 2 u x 4 u d x + 2 δ 2 b 2 R x 2 u 3 x 2 u d x + b 4 R ( x 2 u ) 2 d x + δ 4 R x 2 u 3 2 d x = 2 κ β 2 R u x u x 2 u d x .
Since
b 4 R ( x 2 u ) 2 d x + δ 4 R x 2 u 3 2 d x + β 4 R ( x 4 u ) 4 d x + 2 δ 2 b 2 R x 2 u 3 x 2 u d x 2 β 2 a 2 R x 2 u x 4 u d x 2 β 2 δ 2 R x 2 u 3 x 4 u d x = R b 2 x 2 u + δ 2 x 2 u 3 β 2 x 4 u 2 d x ,
(23) becomes
d d t β 2 2 x u ( t , · ) L 2 ( R ) 2 + δ 2 4 u ( t , · ) L 4 ( R ) 4 + b 2 2 u ( t , · ) L 2 ( R ) 2 + a 2 β 2 x 2 u ( t , · ) L 2 ( R ) 2 + 3 a 2 δ 2 u ( t , · ) x u ( t , · ) L 2 ( R ) 2 + a 2 b 2 x u ( t , · ) L 2 ( R ) 2 + R b 2 x 2 u + δ 2 x 2 u 3 β 2 x 4 u 2 d x = 2 κ β 2 R u x u x 2 u d x .
Due to the Young inequality,
2 | κ | β 2 R | u | x u | | x 2 u | d x = 2 β 2 R κ u x u D 1 D 1 x 2 u d x β 2 κ 2 D 1 u ( t , · ) x u ( t , · ) L 2 ( R ) 2 + β 2 D 1 x 2 u ( t , · ) L 2 ( R ) 2 ,
where D 1 is a positive constant, which will be specified later. Consequently, by (24),
d d t β 2 2 x u ( t , · ) L 2 ( R ) 2 + δ 2 4 u ( t , · ) L 4 ( R ) 4 + b 2 2 u ( t , · ) L 2 ( R ) 2 + β 2 a 2 D 1 x 2 u ( t , · ) L 2 ( R ) 2 + 3 a 2 δ 2 β 2 κ 2 D 1 u ( t , · ) x u ( t , · ) L 2 ( R ) 2 + a 2 b 2 x u ( t , · ) L 2 ( R ) 2 + R b 2 x 2 u + δ 2 x 2 u 3 β 2 x 4 u 2 d x 0 .
We search D 1 , such that
a 2 D 1 > 0 , 3 a 2 δ 2 β 2 κ 2 D 1 > 0 .
Since, after a rescaling, | a | can be taken very big, D 1 does exist and (26) holds.
Therefore, by (25) and (26),
d d t β 2 2 x u ( t , · ) L 2 ( R ) 2 + δ 2 4 u ( t , · ) L 4 ( R ) 4 + b 2 2 u ( t , · ) L 2 ( R ) 2 + K 1 2 x 2 u ( t , · ) L 2 ( R ) 2 + K 2 2 u ( t , · ) x u ( t , · ) L 2 ( R ) 2 + a 2 b 2 x u ( t , · ) L 2 ( R ) 2 + R b 2 x 2 u + δ 2 x 2 u 3 β 2 x 4 u 2 d x 0 ,
where K 1 2 , K 2 2 are two appropriate positive constants. Integrating on ( 0 , t ) , by (6), we have that
β 2 2 x u ( t , · ) L 2 ( R ) 2 + δ 2 4 u ( t , · ) L 4 ( R ) 4 + b 2 2 u ( t , · ) L 2 ( R ) 2 + K 1 2 0 t x 2 u ( s , · ) L 2 ( R ) 2 d s + K 2 2 0 t u ( s , · ) x u ( s , · ) L 2 ( R ) 2 d s + a 2 b 2 0 t x u ( s , · ) L 2 ( R ) 2 d s + 0 t R b 2 x 2 u + δ 2 x 2 u 3 β 2 x 4 u 2 d s d x C 0 ,
that is, (18).
We continue by proving (19). Assume (3). Since κ = γ = 0 , (17) reads
t u β 2 x 6 u + b 2 x 4 u + δ 2 x 4 u 3 = 0 .
Multiplying (27), by β 2 x 2 u + δ 2 u 3 + b 2 u , arguing as in the previous case, an integration on R gives
d d t β 2 2 x u ( t , · ) L 2 ( R ) 2 + δ 2 4 u ( t , · ) L 4 ( R ) 4 + b 2 2 u ( t , · ) L 2 ( R ) 2 + R b 2 x 2 u + δ 2 x 2 u 3 β 2 x 4 u 2 d x = 0 .
(6) and an integration on ( 0 , t ) give (19).
Finally, we prove (20). Thanks to (18) or (19) and the Hölder inequality,
| u ( t , x ) | 3 = 3 x u 2 x u d y 3 R u 2 | x u | d x 3 u ( t , · ) L 4 ( R ) 2 x u ( t , · ) L 2 ( R ) 2 C 0 .
Hence,
u ( t , · ) L ( R ) 3 C 0 ,
which gives (20). □
We continue by proving an L 2 estimate, which is independent on κ , a , b .
Lemma 2.
Fix T > 0 and assume (2) or (3). There exists a constant C ( T ) > 0 , independent on κ , a , b , such that
u ( t , · ) L 2 ( R ) 2 + 2 a 2 R x u ( s , · ) L 2 ( R ) 2 d s + β 2 R x 3 u ( s , · ) L 2 ( R ) 2 d s + 2 b 2 0 t x 2 u ( s , · ) L 2 ( R ) 2 d s C ( T ) ,
0 t x 2 u ( s , · ) L 2 ( R ) 2 d s C ( T ) ,
0 t u ( s , · ) ( x u ( s , · ) ) 2 L 2 ( R ) 2 d s C ( T ) ,
0 t x u ( s , · ) L 4 ( R ) 4 d s C ( T ) ,
for every 0 t T .
Proof. 
Let 0 t T . We begin by observing that
x u 3 = 3 u 2 x u , x 2 u 3 = 6 u ( x u ) 2 + 3 u 2 x 2 u .
Multiplying (17) by 2 u , an integration on R and several integrations by part give
d d t u ( t , · ) L 2 ( R ) 2 = 2 R u t u d x = 4 κ R u 2 x d x + 2 a 2 R u x 2 u d x + 2 β 2 R u x 6 u d x 2 b 2 R u x 4 u d x 2 δ 2 R u x 4 u 3 d x = 2 a 2 x u ( t , · ) L 2 ( R ) 2 2 β 2 R x u x 5 u d x + 2 b 2 R x u x 3 u d x + 2 δ 2 R x u x 3 u 3 d x = 2 a 2 x u ( t , · ) L 2 ( R ) 2 + 2 β 2 R x 2 u x 4 u d x 2 b 2 x 2 u ( t , · ) L 2 ( R ) 2 2 δ 2 R x 2 u x 2 u 3 d x = 2 a 2 x u ( t , · ) L 2 ( R ) 2 2 β 2 x 3 u ( t , · ) L 2 ( R ) 2 2 b 2 x 2 u ( t , · ) L 2 ( R ) 2 2 δ 2 R x 2 u x 2 u 3 d x ,
that is,
d d t u ( t , · ) L 2 ( R ) 2 + 2 a 2 x u ( t , · ) L 2 ( R ) 2 + 2 β 2 x 3 u ( t , · ) L 2 ( R ) 2 + 2 b 2 x 2 u ( t , · ) L 2 ( R ) 2 = 2 δ 2 R x 2 u x 2 u 3 d x .
Due to the Young inequality, we can estimate the right-hand side of (33), as follows:
2 δ 2 R | x 2 u | | x 2 u 3 | d x x 2 u ( t , · ) L 2 ( R ) 2 + δ 4 R x 2 u 3 2 d x .
Consequently, (33) becomes
d d t u ( t , · ) L 2 ( R ) 2 + 2 a 2 x u ( t , · ) L 2 ( R ) 2 + 2 β 2 x 3 u ( t , · ) L 2 ( R ) 2 + 2 b 2 x 2 u ( t , · ) L 2 ( R ) 2 x 2 u ( t , · ) L 2 ( R ) 2 + δ 4 R x 2 u 3 2 d x .
Observe that, by (32),
R x 2 u 3 2 d x = 36 R u 2 ( x u ) 4 d x + 9 u 2 ( t , · ) x 2 u ( t , · ) L 2 ( R ) 2 + 36 R u 3 ( x u ) 2 x 2 u d x = 36 R u 2 ( x u ) 4 d x + 9 u 2 ( t , · ) x 2 u ( t , · ) L 2 ( R ) 2 36 R u 2 ( x u ) 4 d x = 9 u 2 ( t , · ) x 2 u ( t , · ) L 2 ( R ) 2 = 9 R u 4 ( x 2 u ) 2 d x .
Using (20),
R x 2 u 3 2 d x 9 u ( t , · ) L ( R ) 4 x 2 u ( t , · ) L 2 ( R ) 2 C 0 x 2 u ( t , · ) L 2 ( R ) 2 .
It follows from (34) and (36) that
d d t u ( t , · ) L 2 ( R ) 2 + 2 a 2 x u ( t , · ) L 2 ( R ) 2 + 2 β 2 x 3 u ( t , · ) L 2 ( R ) 2 + 2 b 2 x 2 u ( t , · ) L 2 ( R ) 2 C 0 x 2 u ( t , · ) L 2 ( R ) 2 .
Since
C 0 x 2 u ( t , · ) L 2 ( R ) 2 = C 0 R x 2 u x 2 u d x = C 0 R x u x 3 u d x ,
Lemma 1 and the Young inequality give
C 0 x 2 u ( t , · ) L 2 ( R ) 2 2 R C 0 x u 2 β β x 3 u d x C 0 x u ( t , · ) L 2 ( R ) 2 + β 2 x 3 u ( t , · ) L 2 ( R ) 2 C 0 + β 2 x 3 u ( t , · ) L 2 ( R ) 2 .
Consequently, by (37),
d d t u ( t , · ) L 2 ( R ) 2 + 2 a 2 x u ( t , · ) L 2 ( R ) 2 + β 2 x 3 u ( t , · ) L 2 ( R ) 2 + 2 b 2 x 2 u ( t , · ) L 2 ( R ) 2 C ( T ) .
Integrating on ( 0 , t ) , by (6), we have that
u ( t , · ) L 2 ( R ) 2 + 2 a 2 0 t x u ( s , · ) L 2 ( R ) 2 d s + β 2 0 t x 3 u ( s , · ) L 2 ( R ) 2 d s + 2 b 2 0 t x 2 u ( s , · ) L 2 ( R ) 2 d s C 0 + C ( T ) t C ( T ) ,
which gives (28).
(29) follows from (28), (38) and an integration on ( 0 , t ) .
We prove (30). We begin by observing that, by (35) and (36), we have that
36 R u 2 ( x u ) 4 d x C 0 x 2 u ( t , · ) L 2 ( R ) 2 9 u 2 ( t , · ) x 2 u ( t , · ) L 2 ( R ) 2 36 R u 3 ( x u ) 2 x 2 u d x .
Thanks to the Young inequality,
36 R | u | 3 ( x u ) 2 | x 2 u | d x 18 R u 2 ( x u ) 4 d x + 18 u 2 ( t , · ) x 2 u ( t , · ) L 2 ( R ) 2 .
Consequently, by (39),
18 R u 2 ( x u ) 4 d x C 0 x 2 u ( t , · ) L 2 ( R ) 2 + 9 u 2 ( t , · ) x 2 u ( t , · ) L 2 ( R ) 2 .
Observe that, by (20),
9 u 2 ( t , · ) x 2 u ( t , · ) L 2 ( R ) 2 = 9 R u 4 ( x 2 u ) 2 d x C 0 x 2 u ( t , · ) L 2 ( R ) 2 .
Therefore, by (40),
18 u ( t , · ) ( x u ( t , · ) ) 2 L 2 ( R ) 2 C 0 x 2 u ( t , · ) L 2 ( R ) 2 .
Integrating (41) on ( 0 , t ) , by (29), we have (30).
Finally, we prove (31). We begin by observing that [66] (Lemma 2.3 ) says that
x u ( t , · ) L 4 ( R ) 4 6 u ( t , · ) L 2 ( R ) 2 + x u ( t , · ) L 2 ( R ) 2 x 2 u ( t , · ) L 2 ( R ) 2 .
Therefore, by Lemma 1 and (28), we have that
x u ( t , · ) L 4 ( R ) 4 C ( T ) x 2 u ( t , · ) L 2 ( R ) 2 .
Integrating on ( 0 , t ) , by (29), we have (31). □
We continue with an a priori estimate in the space L 2 ( 0 , ; H 4 ( R ) ) .
Lemma 3.
Fix T > 0 and assume (2) or (3). There exists a constant C ( T ) > 0 , independent on κ , a , b , such that
x u ( t , · ) L 2 ( R ) 2 + 2 a 2 0 t x 2 u ( s , · ) L 2 ( R ) 2 d s + β 2 0 t x 4 u ( s , · ) L 2 ( R ) 2 d s + 2 b 2 0 t x 3 u ( s , · ) L 2 ( R ) 2 d s C ( T )
for every 0 t T .
Proof. 
Let 0 t T . Multiplying (17) by 2 x 2 u , an integration on R gives
d d t x u ( t , · ) L 2 ( R ) 2 = 2 R x 2 u t u d x = 4 κ R u x u x 2 u d x 2 a 2 x 2 u ( t , · ) L 2 ( R ) 2 2 β 2 R x 2 u x 6 u d x + 2 b 2 R x 2 u x 4 u d x + 2 δ 2 R x 2 u x 4 u 3 d x = 2 κ R ( x u ) 3 d x 2 a 2 x 2 u ( t , · ) L 2 ( R ) 2 + 2 β 2 R x 3 u x 5 u d x 2 b 2 x 3 u ( t , · ) L 2 ( R ) 2 2 δ 2 R x 3 u x 3 u 3 d x = 2 κ R ( x u ) 3 d x 2 a 2 x 2 u ( t , · ) L 2 ( R ) 2 2 β 2 x 4 u ( t , · ) L 2 ( R ) 2 2 b 2 x 3 u ( t , · ) L 2 ( R ) 2 + 2 δ 2 R x 4 u x 2 u 3 d x ,
that is,
d d t x u ( t , · ) L 2 ( R ) 2 + 2 a 2 x 2 u ( t , · ) L 2 ( R ) 2 + 2 β 2 x 4 u ( t , · ) L 2 ( R ) 2 + 2 b 2 x 3 u ( t , · ) L 2 ( R ) 2 = 2 κ R ( x u ) 3 d x + 2 δ 2 R x 4 u x 2 u 3 d x .
Due to Lemma 1, (36) and the Young inequality,
2 | κ | R | x u | 3 d x κ 2 x u ( t , · ) L 2 ( R ) 2 + x u ( t , · ) L 4 ( R ) 4 C 0 + x u ( t , · ) L 4 ( R ) 4 , 2 δ 2 R | x 4 u | | x 2 u 3 | d x = 2 R β x 4 u δ 2 x 2 u 3 β d x β 2 x 4 u ( t , · ) L 2 ( R ) 2 + δ 4 β 2 R x 2 u 3 2 d x β 2 x 4 u ( t , · ) L 2 ( R ) 2 + C 0 x 2 u ( t , · ) L 2 ( R ) 2 .
Therefore, by (43),
d d t x u ( t , · ) L 2 ( R ) 2 + 2 a 2 x 2 u ( t , · ) L 2 ( R ) 2 + β 2 x 4 u ( t , · ) L 2 ( R ) 2 + 2 b 2 x 3 u ( t , · ) L 2 ( R ) 2 x u ( t , · ) L 4 ( R ) 4 + C 0 x 2 u ( t , · ) L 2 ( R ) 2 .
Integrating on ( 0 , t ) , by (6), (29) and (31),
x u ( t , · ) L 2 ( R ) 2 + 2 a 2 0 t x 2 u ( s , · ) L 2 ( R ) 2 d s + β 2 0 t x 4 u ( s , · ) L 2 ( R ) 2 d s + 2 b 2 0 t x 3 u ( s , · ) L 2 ( R ) 2 d s C 0 + 0 t x u ( s , · ) L 4 ( R ) 4 d s + C 0 0 t x 2 u ( s , · ) L 2 ( R ) 2 d s C ( T ) ,
which gives (42). □
We continue with an a priori estimate in the space L 2 ( 0 , ; H 5 ( R ) ) .
Lemma 4.
Fix T > 0 and assume (2) or (3). There exists a constant C ( T ) > 0 , independent on κ , a , b ,
x u L ( ( 0 , T ) × R ) C ( T ) ,
x 2 u ( t , · ) L 2 ( R ) 2 + 2 a 2 0 t x 4 u ( s , · ) L 2 ( R ) 2 d s + β 2 0 t x 5 u ( s , · ) L 2 ( R ) 2 d s + 2 b 2 0 t x 4 u ( s , · ) L 2 ( R ) 2 d s C ( T ) ,
for every 0 t T , where C ( T ) is independent on κ , a , b .
Proof. 
Let 0 t T . We begin by observing that, by (32), we have that
x 3 u 3 = 6 ( x u ) 3 + 18 u x u x 2 u + 3 u 2 x 3 u .
Multiplying (17) by 2 x 4 u , thanks to (46), an integration on R gives
d d t x 2 u ( t , · ) L 2 ( R ) 2 = 2 R x 4 u t u d x = 2 κ R u x u x 4 u d x + 2 a 2 R x 2 u x 4 u d x + 2 β 2 R x 4 u x 6 u d x 2 b 2 x 4 u ( t , · ) L 2 ( R ) 2 2 δ 2 R x 4 u x 4 u 3 d x = 2 κ R u x u x 4 u d x 2 a 2 x 3 u ( t , · ) L 2 ( R ) 2 2 β 2 x 5 u ( t , · ) L 2 ( R ) 2 2 b 2 x 4 u ( t , · ) L 2 ( R ) 2 + 2 δ 2 R x 5 u x 3 u 3 d x = 2 κ R u x u x 4 u d x 2 a 2 x 3 u ( t , · ) L 2 ( R ) 2 2 β 2 x 5 u ( t , · ) L 2 ( R ) 2 2 b 2 x 4 u ( t , · ) L 2 ( R ) 2 12 δ 2 R ( x u ) 3 x 5 u d x 36 R u x u x 2 u x 5 u d x 6 R u 2 x 3 u x 5 u d x ,
that is,
d d t x 2 u ( t , · ) L 2 ( R ) 2 + 2 a 2 x 4 u ( t , · ) L 2 ( R ) 2 + 2 β 2 x 5 u ( t , · ) L 2 ( R ) 2 + 2 b 2 x 4 u ( t , · ) L 2 ( R ) 2 = 2 κ R u x u x 4 u d x 12 δ 2 R ( x u ) 3 x 5 u d x 36 δ 2 R u x u x 2 u x 5 u d x 6 δ 2 R u 2 x 3 u x 5 u d x .
Due to Lemma 1 and the Young inequality,
2 | κ | R | u | | x u | | x 4 u | d x κ 2 R u 2 ( x u ) 2 d x + x 4 u ( t , · ) L 2 ( R ) 2 κ 2 u ( t , · ) L ( R ) 2 x u ( t , · ) L 2 ( R ) 2 + x 4 u ( t , · ) L 2 ( R ) 2 C 0 x u ( t , · ) L 2 ( R ) 2 + x 4 u ( t , · ) L 2 ( R ) 2 C 0 + x 4 u ( t , · ) L 2 ( R ) 2 , 12 δ 2 R | x u | 3 | x 5 u | d x = 2 R 6 δ 2 ( x u ) 3 β D 2 β D 2 x 5 u d x 36 δ 4 β 2 D 2 R ( x u ) 6 d x + β 2 D 2 x 5 u ( t , · ) L 2 ( R ) 2 36 δ 4 β 2 D 2 x u L ( ( 0 , T ) × R ) 2 x u ( t , · ) L 4 ( R ) 4 + β 2 D 2 x 5 u ( t , · ) L 2 ( R ) 2 , 36 δ 2 R | u | | x u x 2 u | | x 5 u | d x 36 δ 2 u ( t , · ) L ( R ) R | x u x 2 u | | x 5 u | d x 2 C 0 R | x u x 2 u | | x 5 u | d x = 2 R C 0 x u x 2 u β D 2 β D 2 x 5 u d x C 0 D 2 R ( x u ) 2 ( x 2 u ) 2 d x + β 2 D 2 x 5 u ( t , · ) L 2 ( R ) 2 C 0 D 2 x u L ( ( 0 , T ) × R ) 2 x 2 u ( t , · ) L 2 ( R ) 2 + β 2 D 2 x 5 u ( t , · ) L 2 ( R ) 2 , 6 δ 2 R u 2 | x 3 u | | x 5 u | d x 6 δ 2 u ( t , · ) L ( R ) 2 R | x 3 u | | x 5 u | d x 2 C 0 R | x 3 u | | x 5 u | d x 2 R C 0 x 3 u β D 2 β D 2 x 5 u d x C 0 D 2 x 3 u ( t , · ) L 2 ( R ) 2 + β 2 D 2 x 5 u ( t , · ) L 2 ( R ) 2 ,
where D 2 is a positive constant, which will be specified later. Therefore, by (47),
d d t x 2 u ( t , · ) L 2 ( R ) 2 + 2 a 2 x 4 u ( t , · ) L 2 ( R ) 2 + β 2 2 3 D 2 x 5 u ( t , · ) L 2 ( R ) 2 + 2 b 2 x 4 u ( t , · ) L 2 ( R ) 2 C 0 + x 4 u ( t , · ) L 2 ( R ) 2 + 36 δ 4 β 2 D 2 x u L ( ( 0 , T ) × R ) 2 x u ( t , · ) L 4 ( R ) 4 + C 0 D 2 x u L ( ( 0 , T ) × R ) 2 x 2 u ( t , · ) L 2 ( R ) 2 + C 0 D 2 x 3 u ( t , · ) L 2 ( R ) 2 ,
and taking D 2 = 1 3
d d t x 2 u ( t , · ) L 2 ( R ) 2 + 2 a 2 x 4 u ( t , · ) L 2 ( R ) 2 + β 2 x 5 u ( t , · ) L 2 ( R ) 2 + 2 b 2 x 4 u ( t , · ) L 2 ( R ) 2 C 0 + x 4 u ( t , · ) L 2 ( R ) 2 + 108 δ 4 β 2 x u L ( ( 0 , T ) × R ) 2 x u ( t , · ) L 4 ( R ) 4 + C 0 x u L ( ( 0 , T ) × R ) 2 x 2 u ( t , · ) L 2 ( R ) 2 + C 0 x 3 u ( t , · ) L 2 ( R ) 2 .
Integrating on ( 0 , t ) , by (6), (29), (31) and (42), we obtain that
x 2 u ( t , · ) L 2 ( R ) 2 + 2 a 2 0 t x 4 u ( t , · ) L 2 ( R ) 2 + β 2 0 t x 5 u ( s , · ) L 2 ( R ) 2 d s + 2 b 2 0 t x 4 u ( s , · ) L 2 ( R ) 2 d s C 0 + C 0 t + 0 t x 4 u ( s , · ) L 2 ( R ) 2 d s + 108 δ 4 β 2 x u L ( ( 0 , T ) × R ) 2 0 t x u ( s , · ) L 4 ( R ) 4 d s + C 0 x u L ( ( 0 , T ) × R ) 2 0 t x 2 u ( s , · ) L 2 ( R ) 2 d s + C 0 0 t x 3 u ( s , · ) L 2 ( R ) 2 d s C ( T ) 1 + x u L ( ( 0 , T ) × R ) 2 .
We prove (44). Thanks to Lemma 1, (48) and the Hölder inequality,
( x u ( t , x ) ) 2 = 2 x x u x 2 d x 2 R | x | | x 2 u | d x 2 x u ( t , · ) L 2 ( R ) x 2 u ( t , · ) L 2 ( R ) C ( T ) 1 + x u L ( ( 0 , T ) × R ) 2 .
Hence,
x u L ( ( 0 , T ) × R ) 4 C ( T ) x u L ( ( 0 , T ) × R ) 2 C ( T ) 0 ,
which gives (44).
Finally, (45) follows from (44) and (48). □
We continue with an a priori estimate in the space L 4 ( 0 , ; W 2 , 4 ( R ) ) .
Lemma 5.
Fix T > 0 and assume (2) or (3). There exists a constant C ( T ) > 0 , independent on κ , a , b , such that
0 t x 2 u ( s , · ) L 4 ( R ) 4 d s C ( T ) ,
for every 0 t T .
Proof. 
Let 0 t T . We begin by observing that
R ( x 2 u ) 4 d x = R ( x 2 u ) 3 x 2 u d x = 3 R x u ( x 2 u ) 2 x 3 u d x .
Due to the Young inequality,
3 R | u | x u ( x 2 u ) 2 | x 3 u | d x 9 2 R ( x u ) 2 ( x 3 u ) 2 d x + 1 2 R ( x 2 u ) 4 d x .
It follows from (50) that
R ( x 2 u ) 4 d x 9 R ( x u ) 2 ( x 3 u ) 2 d x .
By (44), we have that
x 2 u ( t , · ) L 4 ( R ) 4 9 x u L ( ( 0 , T ) × R ) 2 x 3 u ( t , · ) L 2 ( R ) 2 C ( T ) x 3 u ( t , · ) L 2 ( R ) 2 .
Integrating on ( 0 , t ) , by (28), we have that
0 t x 2 u ( s , · ) L 4 ( R ) 4 d s C ( T ) 0 t x 3 u ( s , · ) L 2 ( R ) 2 d s C ( T ) ,
which gives (49). □
We continue with an a priori estimate in the space L 2 ( 0 , ; H 6 ( R ) ) .
Lemma 6.
Fix T > 0 and assume (2) or (3). There exists a constant C ( T ) > 0 , independent on κ , a , b , such that
x 3 u ( t , · ) L 2 ( R ) 2 + 2 a 2 0 t x 4 u ( s , · ) L 2 ( R ) 2 d s + β 2 0 t x 6 u ( s , · ) L 2 ( R ) 2 d s + 2 b 2 0 t x 5 u ( s , · ) L 2 ( R ) 2 d s C ( T )
In particular, we have that
x 2 u L ( ( 0 , T ) × R ) C ( T ) ,
where C ( T ) is independent on κ , a , b .
Proof. 
Let 0 t T . We begin by observing that, by (46),
x 4 u 3 = 36 ( x u ) 2 x 2 u + 18 u ( x 2 u ) 2 + 24 u x u x 3 u + 3 u 2 x 4 u
Multiplying (17) by 2 x 6 u , thanks to (46), an integration on R gives
d d t x 3 u ( t , · ) L 2 ( R ) 2 = 2 R x 6 u t u d x = 4 κ R u x u x 6 u d x 2 a 2 R x 2 u x 6 u d x 2 β 2 x 6 u ( t , · ) L 2 ( R ) 2 + 2 b 2 R x 4 u x 6 u d x + 2 δ 2 R x 5 u x 4 u 3 d x = 4 κ R u x u x 6 u d x + 2 a 2 R x 3 u x 5 u d x 2 β 2 x 6 u ( t , · ) L 2 ( R ) 2 2 b 2 x 5 u ( t , · ) L 2 ( R ) 2 + 72 δ 2 R ( x u ) 2 x 2 u x 6 u d x + 36 δ 2 R u ( x 2 u ) 2 x 6 u d x + 48 δ 2 R u x u x 3 u x 6 u d x + 6 δ 2 R u 2 x 4 u x 6 u d x = 4 κ R u x u x 6 u d x 2 a 2 x 4 u ( t , · ) L 2 ( R ) 2 2 β 2 x 6 u ( t , · ) L 2 ( R ) 2 2 b 2 x 5 u ( t , · ) L 2 ( R ) 2 + 72 δ 2 R ( x u ) 2 x 2 u x 6 u d x + 36 δ 2 R u ( x 2 u ) 2 x 6 u d x + 48 δ 2 R u x u x 3 u x 6 u d x + 6 δ 2 R u 2 x 4 u x 6 u d x .
Therefore, we have that
d d t x 3 u ( t , · ) L 2 ( R ) 2 + 2 a 2 x 4 u ( t , · ) L 2 ( R ) 2 + 2 β 2 x 6 u ( t , · ) L 2 ( R ) 2 + 2 b 2 x 5 u ( t , · ) L 2 ( R ) 2 = 4 κ R u x u x 6 u d x + 72 δ 2 R ( x u ) 2 x 2 u x 6 u d x + 36 δ 2 R u ( x 2 u ) 2 x 6 u d x + 48 δ 2 R u x u x 3 u x 6 u d x + 6 δ 2 R u 2 x 4 u x 6 u d x .
Due to Lemma 1, (44), (45) and the Young inequality,
4 | κ | R | u x u | | x 6 u | d x 4 | κ | u ( t , · ) L ( R ) R | x u | | x 6 u | d x 2 C 0 R | x u | | x 6 u | d x = 2 R C 0 x u β D 3 β D 3 x 6 u d x d x C 0 D 3 x u ( t , · ) L 2 ( R ) 2 + β 2 D 3 x 6 u ( t , · ) L 2 ( R ) 2 C 0 D 3 + β 2 D 3 x 6 u ( t , · ) L 2 ( R ) 2 , 72 δ 2 R ( x u ) 2 | x 2 u | | x 6 u | d x 72 δ 2 x u L ( ( 0 , T ) × R ) 2 R | x 2 u | | x 6 u | d x 2 C ( T ) R | x 2 u | | x 6 u | d x = 2 R C ( T ) x 2 u β D 3 β D 3 x 6 u d x C ( T ) D 3 x 2 u ( t , · ) L 2 ( R ) 2 + β 2 D 3 x 6 u ( t , · ) L 2 ( R ) 2 C ( T ) D 3 + β 2 D 3 x 6 u ( t , · ) L 2 ( R ) 2 , 36 δ 2 R | u | ( x 2 u ) 2 | x 6 u | d x = 36 δ 2 u L ( ( 0 , T ) × R ) R ( x 2 u ) 2 | x 6 u | d x 2 C ( T ) R ( x 2 u ) 2 | x 6 u | d x = 2 R C ( T ) ( x 2 u ) 2 β D 3 β D 3 x 6 u d x C ( T ) D 3 x 2 u ( t , · ) L 4 ( R ) 4 + β 2 D 3 x 6 u ( t , · ) L 2 ( R ) 2 , 48 δ 2 R | u | | x u | | x 3 u | | x 6 u | d x 48 δ 2 u ( t , · ) L ( R ) R | x u | | x 3 u | | x 6 u | d x 2 C 0 R | x u | | x 3 u | | x 6 u | d x 2 C 0 x u L ( ( 0 , T ) × R ) R | x 3 u | | x 6 u | d x 2 C ( T ) R | x 3 u | | x 6 u | d x 2 R C ( T ) x 3 u β D 3 β D 3 x 6 u d x C ( T ) D 3 x 3 u ( t , · ) L 2 ( R ) 2 + β 2 D 3 x 6 u ( t , · ) L 2 ( R ) 2 , 6 δ 2 R u 2 | x 4 u | | x 6 u | d x 6 δ 2 u ( t , · ) L ( R ) 2 R | x 4 u | | x 6 u | d x 2 C 0 R | x 4 u | | x 6 u | d x = 2 R C 0 x 4 u β D 3 β D 3 x 6 u d x C 0 D 3 x 4 u ( t , · ) L 2 ( R ) 2 + β 2 D 3 x 6 u ( t , · ) L 2 ( R ) 2 ,
where D 3 is a positive constant, which will be specified later. It follows from (54) that
d d t x 3 u ( t , · ) L 2 ( R ) 2 + 2 a 2 x 4 u ( t , · ) L 2 ( R ) 2 + β 2 2 5 D 3 x 6 u ( t , · ) L 2 ( R ) 2 + 2 b 2 x 5 u ( t , · ) L 2 ( R ) 2 C ( T ) D 3 + C ( T ) D 3 x 2 u ( t , · ) L 4 ( R ) 4 + C ( T ) D 3 x 3 u ( t , · ) L 2 ( R ) 2 + C 0 D 3 x 4 u ( t , · ) L 2 ( R ) 2 .
Taking D 3 = 1 5 , we have that
d d t x 3 u ( t , · ) L 2 ( R ) 2 + 2 a 2 x 4 u ( t , · ) L 2 ( R ) 2 + β 2 x 6 u ( t , · ) L 2 ( R ) 2 + 2 b 2 x 5 u ( t , · ) L 2 ( R ) 2 C ( T ) + C ( T ) x 2 u ( t , · ) L 4 ( R ) 4 + C ( T ) x 3 u ( t , · ) L 2 ( R ) 2 + C 0 x 4 u ( t , · ) L 2 ( R ) 2 .
Integrating on ( 0 , t ) , by (6), (28), (42), (49), we have that
x 3 u ( t , · ) L 2 ( R ) 2 + 2 a 2 0 t x 4 u ( s , · ) L 2 ( R ) 2 d s + β 2 0 t x 6 u ( s , · ) L 2 ( R ) 2 d s + 2 b 2 0 t x 5 u ( s , · ) L 2 ( R ) 2 d s C 0 + C ( T ) 0 t x 2 u ( s , · ) L 4 ( R ) 4 d s + C ( T ) 0 t x 3 u ( s , · ) L 2 ( R ) 2 d s + C 0 0 t x 4 u ( s , · ) d s C ( T ) ,
which gives (51).
Finally, we prove (52). Thanks to (45), (51) and the Hölder inequality,
( x 2 u ( t , x ) ) 2 = 2 x x 2 u x 3 u d y 2 R | x 2 u | | x 3 u | d x 2 x 2 u ( t , · ) L 2 ( R ) x 3 u ( t , · ) L 2 ( R ) C ( T ) .
Hence,
x 2 u L ( ( 0 , T ) × R ) 2 C ( T ) ,
which gives (52). □
We continue with an a priori estimate in the space H 1 ( ( 0 , ) × R ) .
Lemma 7.
Fix T > 0 and assume (2) or (3). There exists a constant C ( T ) > 0 , independent on κ , a , b , such that
a 2 x u ( t , · ) L 2 ( R ) 2 + β 2 x 3 u ( t , · ) L 2 ( R ) 2 + b 2 x 2 u ( t , · ) L 2 ( R ) 2 + 0 t t u ( s , · ) L 2 ( R ) 2 d s C ( T ) ,
for every 0 t T .
Proof. 
Let 0 t T . Multiplying (17) by 2 t u , we have that
2 ( t u ) 2 + 4 κ u x u t u 2 a 2 x 2 u t u 2 β 2 x 6 u t u + 2 b 2 x 4 u t u + 2 δ 2 t u x 4 u 3 = 0 .
Since,
2 a 2 R x 2 u t u = a 2 d d t x u ( t , · ) L 2 ( R ) 2 , 2 β 2 R x 6 u t u d x = β 2 d d t x 3 u ( t , · ) L 2 ( R ) 2 , 2 b 2 R x 4 u t u d x = b 2 d d t x 2 u ( t , · ) L 2 ( R ) 2 ,
thanks to (53), an integration of (56) on R gives
d d t a 2 x u ( t , · ) L 2 ( R ) 2 + β 2 x 3 u ( t , · ) L 2 ( R ) 2 + b 2 d d t x 2 u ( t , · ) L 2 ( R ) 2 + 2 t u ( t , · ) L 2 ( R ) 2 = 4 κ R u x u t u d x 72 δ 2 R ( x u ) 2 x 2 u t u d x + 36 δ 2 R u ( x 2 u ) 2 t u d x 48 δ 2 R u x u x 2 u t u d x 6 δ 2 R u 2 x 4 u t u d x .
Due to Lemma 1, (44), (45), (52) and the Young inequality,
4 | κ | R | u | | x u | | t u | d x 4 | κ | u ( t , · ) L ( R ) R | x u | | t u | d x 2 C 0 R | x u | | t u | d x = 2 R C 0 x u D 3 D 3 t u d x C 0 D 3 x u ( t , · ) L 2 ( R ) 2 + D 3 t u ( t , · ) L 2 ( R ) 2 C 0 D 3 + D 3 t u ( t , · ) L 2 ( R ) 2 , 72 δ 2 R ( x u ) 2 | x 2 u | | t u | d x 72 δ 2 x u L ( ( 0 , T ) × R ) 2 R | x 2 u | | t u | d x 2 C ( T ) R | x 2 u | | t u | d x = 2 R C ( T ) x 2 u D 3 D 3 t u d x C ( T ) D 3 x 2 u ( t , · ) L 2 ( R ) 2 + D 3 t u ( t , · ) L 2 ( R ) 2 C ( T ) D 3 + D 3 t u ( t , · ) L 2 ( R ) 2 , 36 δ 2 R | u | ( x 2 u ) 2 | t u | d x 36 δ 2 u ( t , · ) L ( R ) R ( x 2 u ) 2 | t u | d x 2 C 0 R ( x 2 u ) 2 | t u | d x = 2 C 0 x 2 u L ( ( 0 , T ) × R ) R | x 2 u | | t u | d x 2 C ( T ) R | x 2 u | | t u | d x = 2 R C ( T ) x 2 u D 3 D 3 t u d x C ( T ) D 3 x 2 u ( t , · ) L 2 ( R ) 2 + D 3 t u ( t , · ) L 2 ( R ) 2 C ( T ) D 3 + D 3 t u ( t , · ) L 2 ( R ) 2 , 48 δ 2 R | u | | x u | | x 2 u | | t u | d x 48 δ 2 u ( t , · ) L ( R ) R | x u | | x 2 u | | t u | d x 2 C 0 R | x u | | x 2 u | | t u | d x 2 C 0 x u L ( ( 0 , T ) × R ) R | x 2 u | | t u | d x 2 C ( T ) R | x 2 u | | t u | d x = 2 R C ( T ) x 2 u D 3 D 3 x 2 u d x C ( T ) D 3 x 2 u ( t , · ) L 2 ( R ) 2 + D 3 t u ( t , · ) L 2 ( R ) 2 C ( T ) D 3 + D 3 t u ( t , · ) L 2 ( R ) 2 , 6 δ 2 R u 2 | x 4 u | | t u | d x = 6 δ 2 u ( t , · ) L ( R ) 2 R | x 4 u | | t u | d x 2 C 0 R | x 4 u | | t u | d x = 2 R C 0 x 4 u D 3 D 3 t u d x C 0 D 3 x 4 u ( t , · ) L 2 ( R ) 2 + D 3 t u ( t , · ) L 2 ( R ) 2 ,
where D 3 is a positive constant, which will be specified later. As a consequence, (57) becomes
d d t a 2 x u ( t , · ) L 2 ( R ) 2 + β 2 x 3 u ( t , · ) L 2 ( R ) 2 + b 2 d d t x 2 u ( t , · ) L 2 ( R ) 2 + 2 1 2 D 3 t u ( t , · ) L 2 ( R ) 2 C ( T ) D 3 + C 0 D 3 x 4 u ( t , · ) L 2 ( R ) 2 ,
and taking D 3 = 1 2 ,
d d t a 2 x u ( t , · ) L 2 ( R ) 2 + β 2 x 3 u ( t , · ) L 2 ( R ) 2 + b 2 d d t x 2 u ( t , · ) L 2 ( R ) 2 + t u ( t , · ) L 2 ( R ) 2 C ( T ) + C 0 x 4 u ( t , · ) L 2 ( R ) 2 .
Integrating on ( 0 , t ) , by (6) and (42), we have that
a 2 x u ( t , · ) L 2 ( R ) 2 + β 2 x 3 u ( t , · ) L 2 ( R ) 2 + b 2 x 2 u ( t , · ) L 2 ( R ) 2 + 0 t t u ( s , · ) L 2 ( R ) 2 d s C 0 + C ( T ) t + C 0 0 t x 4 u ( s , · ) L 2 ( R ) 2 d s C ( T ) ,
which gives (55). □
We are finally ready to prove Theorem 1, assuming (i) or ( i i ).
Proof of Theorem 1 assuming ( i ) or ( ii ).
The well-posedness of (1) is guaranteed for a short time by the Cauchy-Kowaleskaya Theorem [64]. Thanks to the a priori estimates proved in Lemmas 1–7, we have that the global-in-time existence of a is the solution of (1) that satisfies (13).
The stability estimates (15) can be proved using the same arguments of [18] (Theorem 1). □

4. Proof of Theorem 1 Assuming ( iii )

In this section, we prove Theorem 1 assuming ( i i i ). Due to (4), here, (1) becomes
t u + κ x u 2 + γ x 2 u β 2 x 6 u + α x 4 u = 0 , t > 0 , x R , u ( 0 , x ) = u 0 ( x ) , x R .
The argument of this section is analogous to that of the previous one. We deduce the local-in-time well-posedness from the Cauchy-Kowaleskaya Theorem [64], and we improve the local-in-time existence to the global-in-time one, proving some suitable a priori estimates on u.
We begin with an energy estimate in the space L l o c ( 0 , ; L 2 ( R ) ) L l o c 2 ( 0 , ; H 3 ( R ) ) .
Lemma 8.
Fix T > 0 . There exists a constant C ( T ) > 0 , such that
u ( t , · ) L 2 ( R ) 2 + β 2 e C 0 t 0 t e C s x 3 u ( s , · ) L 2 ( R ) 2 d s C ( T ) ,
for every 0 t T . In particular, (29) holds. Moreover, we have that
0 t x u ( s , · ) L 2 ( R ) 2 d s C ( T ) ,
for every 0 t T .
Proof. 
Let 0 t T . Multiplying (58) by 2 u , an integration on R gives
d d t u ( t , · ) L 2 ( R ) 2 = 2 R u t u d x = 4 κ R u 2 x u d x 2 γ R u x 2 u d x + 2 β 2 R u x 6 u d x 2 α R u x 4 u d x = 2 γ R u x 2 u d x 2 β 2 R x u x 5 u d x + 2 α R x u x 3 u d x = 2 γ R u x 2 u d x + 2 β 2 R x 2 u x 4 u d x 2 α x 2 u ( t , · ) L 2 ( R ) 2 = 2 γ R u x 2 u d x 2 β 2 x 3 u ( t , · ) L 2 ( R ) 2 2 α x 2 u ( t , · ) L 2 ( R ) 2 ,
that is,
d d t u ( t , · ) L 2 ( R ) 2 + 2 β 2 x 3 u ( t , · ) L 2 ( R ) 2 = 2 γ R u x 2 u d x 2 α x 2 u ( t , · ) L 2 ( R ) 2 .
Since, using the Young inequality,
2 | γ | R | u | | x 2 u | d x γ 2 u ( t , · ) L 2 ( R ) 2 + x 2 u ( t , · ) L 2 ( R ) 2 ,
we can pass from (61) to
d d t u ( t , · ) L 2 ( R ) 2 + 2 β 2 x 3 u ( t , · ) L 2 ( R ) 2 C 0 u ( t , · ) L 2 ( R ) 2 + C 0 x 2 u ( t , · ) L 2 ( R ) 2 .
Observe that
C 0 x 2 u ( t , · ) L 2 ( R ) 2 = C 0 R x 2 u x 2 u d x = C 0 R x u x 3 u d x .
Therefore, by the Young inequality,
C 0 x 2 u ( t , · ) L 2 ( R ) 2 2 R C 0 x u 2 D 4 D 4 x 3 u d x C 0 D 4 x u ( t , · ) L 2 ( R ) 2 + D 4 x 3 u ( t , · ) L 2 ( R ) 2 ,
where D 4 is a positive constant, which will be specified later. Observe again that
C 0 D 4 x u ( t , · ) L 2 ( R ) 2 = C 0 D 4 R x u x u d x = C 0 D 4 R u x 2 u d x .
Consequently, by the Young inequality,
C 0 D 4 x u ( t , · ) L 2 ( R ) 2 2 R C 0 D 4 u 2 D 5 D 5 x 2 u d x C 0 D 4 2 D 5 u ( t , · ) L 2 ( R ) 2 + D 5 x 2 u ( t , · ) L 2 ( R ) 2 ,
where D 5 is a positive constant, which will be specified later. Il follows from (63) and (64) that
C 0 D 5 x 2 u ( t , · ) L 2 ( R ) 2 C 0 D 4 2 D 5 u ( t , · ) L 2 ( R ) 2 + D 4 x 3 u ( t , · ) L 2 ( R ) 2 .
Choosing
D 5 = C 0 2 ,
we have that
C 0 2 x 2 u ( t , · ) L 2 ( R ) 2 2 D 4 2 u ( t , · ) L 2 ( R ) 2 + D 4 x 3 u ( t , · ) L 2 ( R ) 2 ,
that is
C 0 x 2 u ( t , · ) L 2 ( R ) 2 4 D 4 2 u ( t , · ) L 2 ( R ) 2 + 2 D 4 x 3 u ( t , · ) L 2 ( R ) 2
It follows from (62) and (66) that
d d t u ( t , · ) L 2 ( R ) 2 + 2 β 2 D 4 x 3 u ( t , · ) L 2 ( R ) 2 C 0 + 4 D 4 2 u ( t , · ) L 2 ( R ) 2 .
Choosing
D 4 = β 2 2 ,
we have that
d d t u ( t , · ) L 2 ( R ) 2 + β 2 x 3 u ( t , · ) L 2 ( R ) 2 C 0 u ( t , · ) L 2 ( R ) 2 .
By the the Gronwall Lemma and (6), we get
u ( t , · ) L 2 ( R ) 2 + β 2 e C 0 t 0 t e C 0 s x 3 u ( s , · ) L 2 ( R ) 2 d s C 0 e C 0 t C ( T ) ,
which gives (59).
We prove (29). Thanks to (59), (66) and (67),
C 0 x 2 u ( t , · ) L 2 ( R ) 2 C ( T ) + β 2 2 x 3 u ( t , · ) L 2 ( R ) 2 .
Integrating on ( 0 , t ) , by (59), we have that
C 0 0 t x 2 u ( s , · ) L 2 ( R ) 2 d s C ( T ) t + β 2 2 0 t x 3 u ( s , · ) L 2 ( R ) 2 d s C ( T ) ,
which gives (29).
Finally, we prove (60). Thanks to (59), (64) and (67),
C 0 x u ( t , · ) L 2 ( R ) 2 C ( T ) + C 0 x 2 u ( t , · ) L 2 ( R ) 2 .
Integrating on ( 0 , t ) , by (29), we have that
C 0 0 t x u ( s , · ) L 2 ( R ) 2 d s C ( T ) t + C 0 0 t x 2 u ( s , · ) L 2 ( R ) 2 d s C ( T ) ,
which gives (60). □
We continue with an energy estimate in the space L l o c ( 0 , ; H 1 ( R ) ) L l o c 2 ( 0 , ; H 4 ( R ) ) .
Lemma 9.
Fix T > 0 . There exists a constant C ( T ) > 0 , such that
u L ( ( 0 , T ) × R ) C ( T ) ,
x u ( t , · ) L 2 ( R ) 2 + 2 β 2 0 t x 4 u ( s , · ) L 2 ( R ) 2 d s C ( T ) ,
for every 0 t T .
Proof. 
Let 0 t T . Multiplying (58) by 2 x 2 u , an integration on R gives
d d t x u ( t , · ) L 2 ( R ) 2 = 2 R x 2 u t u d x = 4 κ R u x u x 2 u d x + 2 γ x 2 u ( t , · ) L 2 ( R ) 2 2 β 2 R x 6 u x 2 u d x + 2 α R x 4 u x 2 u d x = 4 κ R u x u x 2 u d x + 2 γ x 2 u ( t , · ) L 2 ( R ) 2 + 2 β 2 R x 5 u x 3 u d x 2 α x 3 u ( t , · ) L 2 ( R ) 2 = 4 κ R u x u x 2 u d x + 2 γ x 2 u ( t , · ) L 2 ( R ) 2 2 β 2 x 4 u ( t , · ) L 2 ( R ) 2 2 α x 3 u ( t , · ) L 2 ( R ) 2 .
Consequently, we have that
d d t x u ( t , · ) L 2 ( R ) 2 + 2 β 2 x 4 u ( t , · ) L 2 ( R ) 2 = 4 κ R u x u x 2 u d x + 2 γ x 2 u ( t , · ) L 2 ( R ) 2 2 α x 3 u ( t , · ) L 2 ( R ) 2 .
Due to the Young inequality,
4 κ R | u x u | | x 2 u | d x 2 κ 2 R u 2 ( x u ) 2 d x + 2 x 2 u ( t , · ) L 2 ( R ) 2 2 κ 2 u L ( ( 0 , T ) × R ) 2 x u ( t , · ) L 2 ( R ) 2 + 2 x 2 u ( t , · ) L 2 ( R ) 2 .
Therefore, by (70),
d d t x u ( t , · ) L 2 ( R ) 2 + 2 β 2 x 4 u ( t , · ) L 2 ( R ) 2 2 κ 2 u L ( ( 0 , T ) × R ) 2 x u ( t , · ) L 2 ( R ) 2 + 2 ( 1 + | γ | ) x 2 u ( t , · ) L 2 ( R ) 2 + 2 | α | x 3 u ( t , · ) L 2 ( R ) 2 C 0 u L ( ( 0 , T ) × R ) 2 x u ( t , · ) L 2 ( R ) 2 + C 0 x 2 u ( t , · ) L 2 ( R ) 2 + C 0 x 3 u ( t , · ) L 2 ( R ) 2 .
(6), (29), (59), (60) and an integration on ( 0 , t ) give
x u ( t , · ) L 2 ( R ) 2 + 2 β 2 0 t x 4 u ( s , · ) L 2 ( R ) 2 d s C 0 + C 0 u L ( ( 0 , T ) × R ) 2 0 t x u ( s , · ) L 2 ( R ) 2 d s + C 0 0 t x 2 u ( s , · ) L 2 ( R ) 2 d s + C 0 0 t x 3 u ( s , · ) L 2 ( R ) 2 d s C ( T ) 1 + u L ( ( 0 , T ) × R ) 2 .
We prove (68). Thanks to (59), (69), and the Hölder inequality,
u 2 ( t , x ) = 2 x u x u d y 2 R | u | | x u | d x u ( t , · ) L 2 ( R ) x u ( t , · ) L 2 ( R ) C ( T ) 1 + u L ( ( 0 , T ) × R ) 2 .
Hence,
u L ( ( 0 , T ) × R ) 4 C ( T ) u L ( ( 0 , T ) × R ) 2 C ( T ) 0 ,
which gives (68).
Finally, (69) follows form (68) and (71). □
We continue with an energy estimate in the space L l o c ( 0 , ; H 2 ( R ) ) L l o c 2 ( 0 , ; H 5 ( R ) ) .
Lemma 10.
Fix T > 0 . There exists a constant C ( T ) > 0 , such that
x 2 u ( t , · ) L 2 ( R ) 2 + β 2 R x 5 u ( s , · ) L 2 ( R ) 2 d s C ( T ) ,
for every 0 t T . In particular, (44) holds.
Proof. 
Let 0 t T . Multiplying (58) by 2 x 4 u , an integration on R gives
d d t x 2 u ( t , · ) L 2 ( R ) 2 = 2 R x 4 u t u d x = 4 κ R u x u x 4 u d x 2 γ R x 2 u x 4 u d x + 2 β 2 R x 6 u x 4 u d x 2 α x 4 u ( t , · ) L 2 ( R ) 2 = 4 κ R u x u x 4 u d x + 2 γ x 3 u ( t , · ) L 2 ( R ) 2 2 β 2 x 5 u ( t , · ) L 2 ( R ) 2 2 α x 4 u ( t , · ) L 2 ( R ) 2 .
Therefore, we have that
d d t x 2 u ( t , · ) L 2 ( R ) 2 + 2 β 2 x 5 u ( t , · ) L 2 ( R ) 2 = 4 κ R u x u x 4 u d x + 2 γ x 3 u ( t , · ) L 2 ( R ) 2 2 α x 4 u ( t , · ) L 2 ( R ) 2 .
Due to (68), (69) and the Young inequality,
4 | κ | R | u x u | | x 4 u | d x 4 | κ | u L ( ( 0 , T ) × R ) R | x u | | x 4 u | d x C ( T ) R | x u | | x 4 u | d x C ( T ) x u ( t , · ) L 2 ( R ) 2 + C ( T ) x 4 u ( t , · ) L 2 ( R ) 2 C ( T ) + C ( T ) x 4 u ( t , · ) L 2 ( R ) 2 .
It follows from (73) that
d d t x 2 u ( t , · ) L 2 ( R ) 2 + 2 β 2 x 5 u ( t , · ) L 2 ( R ) 2 C ( T ) + C ( T ) x 4 u ( t , · ) L 2 ( R ) 2 + 2 | γ | x 3 u ( t , · ) L 2 ( R ) 2 + 2 | α | x 4 u ( t , · ) L 2 ( R ) 2 C ( T ) + C ( T ) x 4 u ( t , · ) L 2 ( R ) 2 + C ( T ) x 3 u ( t , · ) L 2 ( R ) 2 .
Integrating on ( 0 , t ) , by (6), (59) and (69), we have
x 2 u ( t , · ) L 2 ( R ) 2 + 2 β 2 0 t x 5 u ( s , · ) L 2 ( R ) 2 d s C 0 + C ( T ) t + C ( T ) 0 t x 4 u ( s , · ) L 2 ( R ) 2 d s + C ( T ) 0 t x 3 u ( s , · ) L 2 ( R ) 2 d s C ( T ) ,
which gives (72).
Finally, by (69), (72) and the Hölder inequality, we have (44). □
We continue with an energy estimate in the space L l o c ( 0 , ; H 3 ( R ) ) L l o c 2 ( 0 , ; H 6 ( R ) ) .
Lemma 11.
Fix T > 0 . There exists a constant C ( T ) > 0 , such that
x 3 u ( t , · ) L 2 ( R ) 2 + β 2 0 t x 6 u ( s , · ) L 2 ( R ) 2 d s C ( T ) ,
for every 0 t T . In particular, (52) holds.
Proof. 
Let 0 t T . Multiplying (58) by 2 x 6 u , an integration on R gives
d d t x 3 u ( t , · ) L 2 ( R ) 2 = 2 R x 6 u t u d x = 4 κ R u x u x 6 u d x + 2 γ R x 6 u x 2 u d x 2 β 2 x 6 u ( t , · ) L 2 ( R ) 2 + 2 α R x 6 u x 4 u d x = 4 κ R u x u x 6 u d x 2 γ R x 5 u x 3 u d x 2 β 2 x 6 u ( t , · ) L 2 ( R ) 2 2 α x 5 u ( t , · ) L 2 ( R ) 2 = 4 κ R u x u x 6 u d x + 2 γ x 4 u ( t , · ) L 2 ( R ) 2 2 β 2 x 6 u ( t , · ) L 2 ( R ) 2 2 α x 5 u ( t , · ) L 2 ( R ) 2 .
Therefore, we have that
d d t x 3 u ( t , · ) L 2 ( R ) 2 + 2 β 2 x 6 u ( t , · ) L 2 ( R ) 2 = 4 κ R u x u x 6 u d x + 2 γ x 4 u ( t , · ) L 2 ( R ) 2 2 α x 5 u ( t , · ) L 2 ( R ) 2 .
Due to (68), (69) and the Young inequality,
4 | κ | R | u | | x u | | x 6 u | d x 2 | κ | u L ( ( 0 , T ) × R ) R | x u | | x 6 u | d x 2 C ( T ) R | x u | | x 6 u | d x = 2 R C ( T ) x u β β x 6 u d x C ( T ) x u ( t , · ) L 2 ( R ) 2 + β 2 x 6 u ( t , · ) L 2 ( R ) 2 C ( T ) + β 2 x 6 u ( t , · ) L 2 ( R ) 2 .
It follows from (75) that
d d t x 3 u ( t , · ) L 2 ( R ) 2 + β 2 x 6 u ( t , · ) L 2 ( R ) 2 C ( T ) + 2 | γ | x 4 u ( t , · ) L 2 ( R ) 2 + 2 | α | x 5 u ( t , · ) L 2 ( R ) 2 C ( T ) + C ( T ) x 4 u ( t , · ) L 2 ( R ) 2 + C ( T ) x 5 u ( t , · ) L 2 ( R ) 2 .
By (6), (69), (72), and an integration on ( 0 , t ) ,
x 3 u ( t , · ) L 2 ( R ) 2 + β 2 0 t x 6 u ( s , · ) L 2 ( R ) 2 d s C 0 + C ( T ) t + C ( T ) 0 t x 4 u ( s , · ) L 2 ( R ) 2 d s + C ( T ) 0 t x 5 u ( s , · ) L 2 ( R ) 2 d s C ( T ) ,
which gives (74).
Finally, arguing as in Lemma 6, we have (52). □
We continue with an energy estimate in the space L l o c ( 0 , ; H 3 ( R ) ) H l o c 1 ( ( 0 , ) × R ) .
Lemma 12.
Fix T > 0 . There exists a constant C ( T ) > 0 , such that
β 2 x 3 u ( t , · ) L 2 ( R ) 2 + 0 t t u ( s , · ) L 2 ( R ) 2 d s C ( T ) .
for every 0 t T .
Proof. 
Let 0 t T . Multiplying (58) by 2 t u , an integration on R gives
d d t β 2 x 3 u ( t , · ) L 2 ( R ) 2 + α x 2 u ( t , · ) L 2 ( R ) 2 γ x u ( t , · ) L 2 ( R ) 2 = 2 β 2 R t u x 6 u d x + 2 α R t u x 4 u d x + 2 γ R t x u x 2 u d x = 2 t u ( t , · ) L 2 ( R ) 2 4 κ R u x u t u d x .
Therefore, we have that
d d t β 2 x 3 u ( t , · ) L 2 ( R ) 2 + α x 2 u ( t , · ) L 2 ( R ) 2 γ x u ( t , · ) L 2 ( R ) 2 + 2 t u ( t , · ) L 2 ( R ) 2 = 4 κ R u x u t u d x .
Due to (68), (69), and the Young inequality,
4 | κ | R | u | | x u | | t u | d x 4 | κ | u L ( ( 0 , T ) × R ) R | x u | | t u | d x 2 C ( T ) R | x u | | t u | d x C ( T ) x u ( t , · ) L 2 ( R ) 2 + t u ( t , · ) L 2 ( R ) 2 C ( T ) + t u ( t , · ) L 2 ( R ) 2 .
Consequently, by (77),
d d t β 2 x 3 u ( t , · ) L 2 ( R ) 2 + α x 2 u ( t , · ) L 2 ( R ) 2 γ x u ( t , · ) L 2 ( R ) 2 + t u ( t , · ) L 2 ( R ) 2 C ( T ) .
(6) and an integration on ( 0 , t ) give
β 2 x 3 u ( t , · ) L 2 ( R ) 2 + α x 2 u ( t , · ) L 2 ( R ) 2 γ x u ( t , · ) L 2 ( R ) 2 + 0 t t u ( s , · ) L 2 ( R ) 2 d s C 0 + C ( T ) t C ( T ) .
Therefore, by (69), (72), we have that
β 2 x 3 u ( t , · ) L 2 ( R ) 2 + 0 t t u ( s , · ) L 2 ( R ) 2 d s C ( T ) + | α | x 2 u ( t , · ) L 2 ( R ) 2 + | γ | x u ( t , · ) L 2 ( R ) 2 C ( T ) ,
which gives (76). □
We are finally ready to prove Theorem 1 assuming ( i i i ).
Proof of Theorem 1 assuming ( iii ). 
The well-posedness of (1) is guaranteed for a short time by the Cauchy-Kowaleskaya Theorem [64]. Thanks to the a priori estimates proved in Lemmas 8–12, we have that the global-in-time existence of a is solution of (1) that satisfies (13).
The stability estimates (15) can be proved using the same arguments of [18] (Theorem 1). □

5. Proof of Theorem 1 Assuming ( iv )

In this final section, we prove Theorem 1 assuming ( i v ).
The argument is again analogous to the one of the previous sectionss. We deduce the local-in-time well-posedness from the Cauchy-Kowaleskaya Theorem [64], and we improve the local-in-time existence to the global-in-time one proving some suitable a priori estimates on u.
We begin with the zero mean estimate.
Lemma 13.
For each t > 0 , we have (14).
Proof. 
Integrating (1) on R , we have that
R t u ( t , x ) d x = d d t R u ( t , x ) d x = 0
(14) follows from (6) and (78). □
Remark 1.
In light of (14), we can consider the following equation:
P ( t , x ) = x u ( t , y ) d y .
Moreover, again by (14), we have that
P ( t , ) = P ( t , ) = 0 .
We continue by proving some energy estimates on the function P.
Lemma 14.
Let T > 0 . There exists a constant C ( T ) > 0 , such that
P ( t , · ) L 2 ( R ) 2 + 2 β 2 e 2 t 0 t e 2 s x 2 u ( s , · ) L 2 ( R ) 2 d s + 2 δ 2 e t 0 2 t e 2 s u ( s , · ) x u ( s , · ) L 2 ( R ) 2 d s C ( T ) + C ( T ) 0 t u ( s , · ) L 4 ( R ) 4 d s + C ( T ) 0 t x u ( s , · ) L 2 ( R ) 2 d s ,
for every 0 t T .
Proof. 
Let 0 t T . Integrating (1) on ( , x ) , we have that
x t u d x + κ u 2 + γ x u β 2 x 5 u + α x 3 u + δ 2 x 3 u 3 = 0 .
Differentiating (79) with respect to t, we obtain that
t P ( t , x ) = d d t x u ( t , y ) d y = x t u ( t , y ) d y .
It follows from (82) and (83) that
t P + κ u 2 + γ x u β 2 x 5 u + α x 3 u + δ 2 x 3 u 3 = 0 .
Arguing as in [18] (Lemma 2), we have that
2 β 2 R P x 5 u d x = 2 β 2 x 2 u ( t , · ) L 2 ( R ) 2 , 2 α R P x 3 u d x = 2 α x u ( t , · ) L 2 ( R ) 2 2 δ 2 R P x 3 u 3 d x = 6 δ 2 u ( t , · ) x u ( t , · ) L 2 ( R ) 2 .
Therefore, multiplying (84) by 2 P , thanks to (85), an integration on R gives
d d t P ( t , · ) L 2 ( R ) 2 + 2 β 2 x 2 u ( t , · ) L 2 ( R ) 2 + 6 δ 2 u ( t , · ) x u ( t , · ) L 2 ( R ) 2 = 2 κ R P u 2 d x 2 γ R P x u d x 2 α x u ( t , · ) L 2 ( R ) 2 .
Due to the Young inequality,
2 | κ | R | P | u 2 d x P ( t , · ) L 2 ( R ) 2 + κ 2 u ( t , · ) L 4 ( R ) 4 , 2 | γ | R | P | | x u | d x P ( t , · ) L 2 ( R ) 2 + γ 2 x u ( t , · ) L 2 ( R ) 2 .
It follows from (86) that
d d t P ( t , · ) L 2 ( R ) 2 + 2 β 2 x 2 u ( t , · ) L 2 ( R ) 2 + 6 δ 2 u ( t , · ) x u ( t , · ) L 2 ( R ) 2 2 P ( t , · ) L 2 ( R ) 2 + C 0 u ( t , · ) L 4 ( R ) 4 + C 0 x u ( t , · ) L 2 ( R ) 2 .
Therefore, by the Gronwall Lemma and (9), we have that
P ( t , · ) L 2 ( R ) 2 + 2 β 2 e 2 t 0 t e 2 s x 2 u ( s , · ) L 2 ( R ) 2 d s + 6 δ 2 e 2 t 0 t e 2 s u ( s , · ) x u ( s , · ) L 2 ( R ) 2 d s C 0 + C 0 e 2 t 0 t e 2 s u ( s , · ) L 4 ( R ) 4 d s + C 0 e 2 t 0 t e 2 s x u ( s , · ) L 2 ( R ) 2 d s C ( T ) + C ( T ) 0 t u ( s , · ) L 4 ( R ) 4 d s + C ( T ) 0 t x u ( s , · ) L 2 ( R ) 2 d s ,
which gives (81). □
Lemma 15.
Let T > 0 . There exists a constant C ( T ) > 0 , such that
β 2 6 x u ( t , · ) L 2 ( R ) 2 + δ 2 4 u ( t , · ) L 4 ( R ) 2 C ( T ) ,
for every 0 t T . In particular, we have (29), (31),
P ( t , · ) L 2 ( R ) C ( T ) , u ( t , · ) L 2 ( R ) C ( T ) , 0 t u ( s , · ) x u ( s , · ) L 2 ( R ) 2 d s C ( T ) , 0 t R β 2 ( x 4 u ) 2 δ 2 x 3 u 3 2 d s d x C ( T ) ,
for every 0 t T . Moreover,
P L ( ( 0 , T ) × R ) C ( T ) u L ( ( 0 , T ) × R ) C ( T ) .
Proof. 
Let 0 t T . Consider an real constant A, which will be specified later. Observe that
δ 2 γ R u 3 x 2 u d x = 3 δ 2 γ u ( t , · ) x u ( t , · ) L 2 ( R ) 2 , 2 A γ R u x 2 u d x = 2 A x u ( t , · ) L 2 ( R ) 2 .
Multiplying (1) by
β 2 x 2 u + δ 2 u 3 + A u ,
thanks to (88) and arguing as in [18] (Lemma 3), an integration on R gives
d d t β 2 2 x u ( t , · ) L 2 ( R ) 2 + δ 2 4 u ( t , · ) L 4 ( R ) 4 + A 2 u ( t , · ) L 2 ( R ) 2 + R β 2 ( x 4 u ) 2 δ 2 x 2 u 3 2 d x = β 2 A + α x 3 u ( t , · ) L 2 ( R ) 2 δ 2 A + α R x 2 u x 2 u 3 d x A α γ β 2 x 2 u ( t , · ) L 2 ( R ) 2 + 3 δ 2 γ u ( t , · ) x u ( t , · ) L 2 ( R ) 2 + 2 β 2 κ R u x u x 2 u d x .
Taking
A = α ,
we have that
d d t β 2 2 x u ( t , · ) L 2 ( R ) 2 + δ 2 4 u ( t , · ) L 4 ( R ) 4 α 2 u ( t , · ) L 2 ( R ) 2 + R β 2 ( x 4 u ) 2 δ 2 x 2 u 3 2 d x = α 2 + γ β 2 x 2 u ( t , · ) L 2 ( R ) 2 + 3 δ 2 γ u ( t , · ) x u ( t , · ) L 2 ( R ) 2 + 2 β 2 κ R u x u x 2 u d x .
Due to the Young inequality,
2 β 2 | κ | R | u x u | | x 2 u | d x β 2 κ 2 u ( t , · ) x u ( t , · ) L 2 ( R ) 2 + β 2 x 2 u ( t , · ) L 2 ( R ) 2 .
It follows from (89) that
d d t β 2 2 x u ( t , · ) L 2 ( R ) 2 + δ 2 4 u ( t , · ) L 4 ( R ) 4 α 2 u ( t , · ) L 2 ( R ) 2 + R β 2 ( x 4 u ) 2 δ 2 x 2 u 3 2 d x C 0 x 2 u ( t , · ) L 2 ( R ) 2 + C 0 u ( t , · ) x u ( t , · ) L 2 ( R ) 2 .
Integrating on ( 0 , t ) , by (7), we have that
β 2 2 x u ( t , · ) L 2 ( R ) 2 + δ 2 4 u ( t , · ) L 4 ( R ) 4 α 2 u ( t , · ) L 2 ( R ) 2 + 0 t R β 2 ( x 4 u ) 2 δ 2 x 2 u 3 2 d s d x C 0 + C 0 0 t x 2 u ( s , · ) L 2 ( R ) 2 d s + C 0 0 t u ( s , · ) x u ( s , · ) L 2 ( R ) 2 d s ,
that is
β 2 2 x u ( t , · ) L 2 ( R ) 2 + δ 2 4 u ( t , · ) L 4 ( R ) 4 + 0 t R β 2 ( x 4 u ) 2 δ 2 x 2 u 3 2 d s d x C 0 + C 0 0 t x 2 u ( s , · ) L 2 ( R ) 2 d s + C 0 0 t u ( s , · ) x u ( s , · ) L 2 ( R ) 2 d s + α 2 u ( t , · ) L 2 ( R ) 2 C 0 + C 0 0 t x 2 u ( s , · ) L 2 ( R ) 2 d s + C 0 0 t u ( s , · ) x u ( s , · ) L 2 ( R ) 2 d s + C 0 u ( t , · ) L 2 ( R ) 2 .
Observe that, by (79) and (80),
C 0 u ( t , · ) L 2 ( R ) 2 = C 0 R u u d x = C 0 R P x u d x .
Therefore, by the Young inequality,
C 0 u ( t , · ) L 2 ( R ) 2 2 R 3 C 0 P 2 β β x u 3 d x C 0 P ( t , · ) L 2 ( R ) 2 + β 2 3 x u ( t , · ) L 2 ( R ) 2 .
It follows from (81), (90) and (91) that
β 2 6 x u ( t , · ) L 2 ( R ) 2 + δ 2 4 u ( t , · ) L 4 ( R ) 4 + 0 t R β 2 ( x 4 u ) 2 δ 2 x 2 u 3 2 d s d x C 0 + C 0 0 t x 2 u ( s , · ) L 2 ( R ) 2 d s + C 0 0 t u ( s , · ) x u ( s , · ) L 2 ( R ) 2 d s + C 0 P ( t , · ) L 2 ( R ) 2 C ( T ) + C ( T ) 0 t u ( s , · ) L 4 ( R ) 4 d s + C ( T ) 0 t x u ( s , · ) L 2 ( R ) 2 d s C ( T ) + C ( T ) β 2 6 0 t x u ( s , · ) L 2 ( R ) 2 d s + δ 2 4 0 t u ( s , · ) L 4 ( R ) 4 d s .
Therefore, we have that
β 2 6 x u ( t , · ) L 2 ( R ) 2 + δ 2 4 u ( t , · ) L 4 ( R ) 4 C ( T ) + C ( T ) β 2 6 0 t x u ( s , · ) L 2 ( R ) 2 d s + δ 2 4 0 t u ( s , · ) L 4 ( R ) 4 d s .
The Gronwall Lemma and (7) give (87).
Finally, arguing as in [18] (Lemma 3), the proof is concluded. □
Arguing as in [18] (Theorem 1), we have Theorem 1.

6. Conclusions

This paper is dedicated to the well-posedness of a solution to the Cauchy problem for a higher-order convective Cahn-Hilliard equation. Such an equation models the evolution of crystal surfaces faceting through surface electromigration, the growing surface faceting, and the evolution of dynamics of phase transitions in ternary oil-water-surfactant systems. The well-posedness of (1) is proved for a short time by the Cauchy-Kowaleskaya Theorem [64]. The global-in-time well-posedness is thus proved, proving several a priori estimates.

Author Contributions

All authors contributed equally to this work. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

The authors are members of the Gruppo Nazionale per l’Analisi Matematica, la Probabilità e le loro Applicazioni (GNAMPA) of the Istituto Nazionale di Alta Matematica (INdAM). GMC has been partially supported by the Research Project of National Relevance “Multiscale Innovative Materials and Structures” granted by the Italian Ministry of Education, University and Research (MIUR Prin 2017, project code 2017J4EAYB and the Italian Ministry of Education, University and Research under the Programme Department of Excellence Legge 232/2016 (Grant No. CUP—D94I18000260001).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Coclite, G.M.; di Ruvo, L. Well-posedness of bounded solutions of the non-homogeneous initial-boundary value problem for the Ostrovsky-Hunter equation. J. Hyperbolic Differ. Equ. 2015, 12, 221–248. [Google Scholar] [CrossRef]
  2. Coclite, G.M.; di Ruvo, L. Oleinik type estimates for the Ostrovsky-Hunter equation. J. Math. Anal. Appl. 2015, 423, 162–190. [Google Scholar] [CrossRef]
  3. Coclite, G.M.; di Ruvo, L. Convergence of the Ostrovsky equation to the Ostrovsky-Hunter one. J. Differ. Equ. 2014, 256, 3245–3277. [Google Scholar] [CrossRef]
  4. Coclite, G.M.; di Ruvo, L. Well-posedness results for the short pulse equation. Z. Angew. Math. Phys. 2015, 66, 1529–1557. [Google Scholar] [CrossRef] [Green Version]
  5. Coclite, G.M.; di Ruvo, L. Wellposedness of bounded solutions of the non-homogeneous initial boundary for the short pulse equation. Boll. Unione Mat. Ital. 2015, 8, 31–44. [Google Scholar] [CrossRef] [Green Version]
  6. Coclite, G.M.; di Ruvo, L. Convergence of the regularized short pulse equation to the short pulse one. Math. Nachr. 2018, 291, 774–792. [Google Scholar] [CrossRef] [Green Version]
  7. Coclite, G.M.; di Ruvo, L. Well-posedness and dispersive/diffusive limit of a generalized Ostrovsky-Hunter equation. Milan J. Math. 2018, 86, 31–51. [Google Scholar] [CrossRef]
  8. Coclite, G.M.; di Ruvo, L. A note on the non-homogeneous initial boundary problem for an Ostrovsky-Hunter type equation. Discret. Contin. Dyn. Syst. Ser. S 2020, 13, 3357–3389. [Google Scholar] [CrossRef] [Green Version]
  9. Coclite, G.M.; di Ruvo, L. A non-local regularization of the short pulse equation. Minimax Theory Appl. 2020, in press. [Google Scholar] [CrossRef]
  10. Coclite, G.M.; di Ruvo, L. A non-local elliptic-hyperbolic system related to the short pulse equation. Nonlinear Anal. 2020, 190, 111606–111628. [Google Scholar] [CrossRef]
  11. Coclite, G.M.; Ruvo, L.D. Well-Posedness Results for the Continuum Spectrum Pulse Equation. Mathematics 2019, 7, 1006. [Google Scholar] [CrossRef] [Green Version]
  12. Coclite, G.M.; di Ruvo, L. On the solutions for an Ostrovsky type equation. Nonlinear Anal. Real World Appl. 2020, 55, 103141. [Google Scholar] [CrossRef]
  13. Savina, T.V.; Golovin, A.A.; Davis, S.H.; Nepomnyashchy, A.A.; Voorhees, P.W. Faceting of a growing crystal surface by surface diffusion. Phys. Rev. E 2003, 67, 021606. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Korzec, M.D.; Rybka, P. On a higher-order convective Cahn-Hilliard-type equation. SIAM J. Appl. Math. 2012, 72, 1343–1360. [Google Scholar] [CrossRef]
  15. Korzec, M.D.; Nayar, P.; Rybka, P. Global attractors of sixth order PDEs describing the faceting of growing surfaces. J. Dyn. Differ. Equ. 2016, 28, 49–67. [Google Scholar] [CrossRef] [Green Version]
  16. Korzec, M.D.; Evans, P.L.; Münch, A.; Wagner, B. Stationary solutions of driven fourth- and sixth-order Cahn-Hilliard-type equations. SIAM J. Appl. Math. 2008, 69, 348–374. [Google Scholar] [CrossRef] [Green Version]
  17. Zhao, X.; Duan, N. Optimal control of the sixth-order convective Cahn-Hilliard equation. Bound. Value Probl. 2014, 2014, 206. [Google Scholar] [CrossRef] [Green Version]
  18. Coclite, G.M.; di Ruvo, L. On the Well-Posedness of A High Order Convective Cahn-Hilliard Type Equations. Algorithms 2020, 13, 170. [Google Scholar] [CrossRef]
  19. Gompper, G.; Goos, J. Fluctuating interfaces in microemulsion and sponge phases. Phys. Rev. E 1994, 50, 1325–1335. [Google Scholar] [CrossRef]
  20. Gompper, G.; Kraus, M. Ginzburg-Landau theory of ternary amphiphilic systems. I. Gaussian interface fluctuations. Phys. Rev. E 1993, 47, 4289–4300. [Google Scholar] [CrossRef]
  21. Gompper, G.; Kraus, M. Ginzburg-Landau theory of ternary amphiphilic systems. II. Monte Carlo simulations. Phys. Rev. E 1993, 47, 4301–4312. [Google Scholar] [CrossRef] [PubMed]
  22. Barakat, F.; Martens, K.; Pierre-Louis, O. Nonlinear Wavelength Selection in Surface Faceting under Electromigration. Phys. Rev. Lett. 2012, 109, 056101. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Pawł ow, I.; Zaja̧czkowski, W.M. A sixth order Cahn-Hilliard type equation arising in oil-water-surfactant mixtures. Commun. Pure Appl. Anal. 2011, 10, 1823–1847. [Google Scholar] [CrossRef]
  24. Wang, Z.; Liu, C. Some properties of solutions for the sixth-order Cahn-Hilliard-type equation. Abstr. Appl. Anal. 2012, 2012, 414590. [Google Scholar] [CrossRef]
  25. Liu, A.; Liu, C. Weak solutions for a sixth order Cahn-Hilliard type equation with degenerate mobility. Abstr. Appl. Anal. 2014, 2014, 407265. [Google Scholar] [CrossRef]
  26. Bohr, T.; Jensen, M.H.; Paladin, G.; Vulpiani, A. Dynamical Systems Approach to Turbulence; Cambridge Nonlinear Science Series; Cambridge University Press: Cambridge, UK, 1998; Volume 8, p. xx+350. [Google Scholar] [CrossRef]
  27. Kuramoto, Y. Diffusion-Induced Chaos in Reaction Systems. Prog. Theor. Phys. Suppl. 1978, 64, 346–367. Available online: https://academic.oup.com/ptps/article-pdf/doi/10.1143/PTPS.64.346/5293041/64-346.pdf (accessed on 19 October 2020). [CrossRef]
  28. Sivashinsky, G. Nonlinear analysis of hydrodynamic instability in laminar flames—I. Derivation of basic equations. Acta Astronaut. 1977, 4, 1177–1206. [Google Scholar] [CrossRef]
  29. Topper, J.; Kawahara, T. Approximate Equations for Long Nonlinear Waves on a Viscous Fluid. J. Phys. Soc. Jpn. 1978, 44, 663–666. [Google Scholar] [CrossRef]
  30. Cohen, B.; Krommes, J.; Tang, W.; Rosenbluth, M. Non-linear saturation of the dissipative trapped-ion mode by mode coupling. Nucl. Fusion 1976, 16, 971–992. [Google Scholar] [CrossRef]
  31. Kuramoto, Y.; Tsuzuki, T. On the Formation of Dissipative Structures in Reaction-Diffusion Systems: Reductive Perturbation Approach. Prog. Theor. Phys. 1975, 54, 687–699. Available online: https://academic.oup.com/ptp/article-pdf/54/3/687/5402406/54-3-687.pdf (accessed on 19 October 2020). [CrossRef] [Green Version]
  32. Kuramoto, Y.; Tsuzuki, T. Persistent Propagation of Concentration Waves in Dissipative Media Far from Thermal Equilibrium. Prog. Theor. Phys. 1976, 55, 356–369. Available online: https://academic.oup.com/ptp/article-pdf/55/2/356/5323559/55-2-356.pdf (accessed on 19 October 2020). [CrossRef] [Green Version]
  33. Chen, L.H.; Chang, H.C. Nonlinear waves on liquid film surfaces—II. Bifurcation analyses of the long-wave equation. Chem. Eng. Sci. 1986, 41, 2477–2486. [Google Scholar] [CrossRef]
  34. Hooper, A.P.; Grimshaw, R. Nonlinear instability at the interface between two viscous fluids. Phys. Fluids 1985, 28, 37–45. [Google Scholar] [CrossRef]
  35. LaQuey, R.E.; Mahajan, S.M.; Rutherford, P.H.; Tang, W.M. Nonlinear Saturation of the Trapped-Ion Mode. Phys. Rev. Lett. 1975, 34, 391–394. [Google Scholar] [CrossRef] [Green Version]
  36. Benney, D.J. Long waves on liquid films. J. Math. Phys. 1966, 45, 150–155. [Google Scholar] [CrossRef]
  37. Lin, S.P. Finite amplitude side-band stability of a viscous film. J. Fluid Mech. 1974, 63, 417–429. [Google Scholar] [CrossRef]
  38. Li, C.; Chen, G.; Zhao, S. Exact travelling wave solutions to the generalized Kuramoto-Sivashinsky equation. Lat. Am. Appl. Res. 2004, 34, 65–68. [Google Scholar]
  39. Foias, C.; Nicolaenko, B.; Sell, G.R.; Temam, R. Inertial manifolds for the Kuramoto-Sivashinsky equation and an estimate of their lowest dimension. J. Math. Pures Appl. 1988, 67, 197–226. [Google Scholar]
  40. Khalique, C. Exact Solutions of the Generalized Kuramoto-Sivashinsky Equation. Casp. J. Math. Sci. CJMS 2012, 1, 109–116. [Google Scholar]
  41. Kudryashov, N.A. Exact solutions of the generalized Kuramoto-Sivashinsky equation. Phys. Lett. A 1990, 147, 287–291. [Google Scholar] [CrossRef]
  42. Nicolaenko, B.; Scheurer, B. Remarks on the Kuramoto-Sivashinsky equation. Phys. D Nonlinear Phenom. 1984, 12, 391–395. [Google Scholar] [CrossRef]
  43. Nicolaenko, B.; Scheurer, B.; Temam, R. Some global dynamical properties of the Kuramoto-Sivashinsky equations: Nonlinear stability and attractors. Phys. D Nonlinear Phenom. 1985, 16, 155–183. [Google Scholar] [CrossRef]
  44. Xie, Y. Solving the generalized Benney equation by a combination method. Int. J. Nonlinear Sci. 2013, 15, 350–354. [Google Scholar]
  45. Armaou, A.; Christofides, P.D. Feedback control of the Kuramoto—Sivashinsky equation. Phys. D Nonlinear Phenom. 2000, 137, 49–61. [Google Scholar] [CrossRef]
  46. Cerpa, E. Null controllability and stabilization of the linear Kuramoto-Sivashinsky equation. Commun. Pure Appl. Anal. 2010, 9, 91–102. [Google Scholar] [CrossRef]
  47. Giacomelli, L.; Otto, F. New bounds for the Kuramoto-Sivashinsky equation. Commun. Pure Appl. Math. 2005, 58, 297–318. [Google Scholar] [CrossRef]
  48. Christofides, P.D.; Armaou, A. Global stabilization of the Kuramoto-Sivashinsky equation via distributed output feedback control. Syst. Control Lett. 2000, 39, 283–294. [Google Scholar] [CrossRef]
  49. Hu, C.; Temam, R. Robust control of the Kuramoto-Sivashinsky equation. Dyn. Contin. Discret. Impuls. Syst. Ser. B Appl. Algorithms 2001, 8, 315–338. [Google Scholar]
  50. Liu, W.J.; Krstić, M. Stability enhancement by boundary control in the Kuramoto-Sivashinsky equation. Nonlinear Anal. 2001, 43, 485–507. [Google Scholar] [CrossRef] [Green Version]
  51. Sajjadian, M. The shock profile wave propagation of Kuramoto-Sivashinsky equation and solitonic solutions of generalized Kuramoto-Sivashinsky equation. Acta Univ. Apulensis Math. Inform. 2014, 38, 163–176. [Google Scholar]
  52. Biagioni, H.A.; Bona, J.L.; Iório, R.J., Jr.; Scialom, M. On the Korteweg-de Vries-Kuramoto-Sivashinsky equation. Adv. Differ. Equ. 1996, 1, 1–20. [Google Scholar]
  53. Coclite, G.M.; di Ruvo, L. On Classical Solutions for A Kuramoto–Sinelshchikov–Velarde-Type Equation. Algorithms 2020, 13, 77. [Google Scholar] [CrossRef] [Green Version]
  54. Tadmor, E. The well-posedness of the Kuramoto-Sivashinsky equation. SIAM J. Math. Anal. 1986, 17, 884–893. [Google Scholar] [CrossRef] [Green Version]
  55. Coclite, G.M.; di Ruvo, L. On the initial-boundary value problem for the Kuramoto-Sinelshchikov type equation. Math. Eng. 2020, 3, 1–43. [Google Scholar] [CrossRef]
  56. Li, J.; Zhang, B.Y.; Zhang, Z. A nonhomogeneous boundary value problem for the Kuramoto-Sivashinsky equation in a quarter plane. Math. Methods Appl. Sci. 2017, 40, 5619–5641. [Google Scholar] [CrossRef] [Green Version]
  57. Li, J.; Zhang, B.Y.; Zhang, Z. A non-homogeneous boundary value problem for the Kuramoto-Sivashinsky equation posed in a finite interval. ESAIM Control Optim. Calc. Var. 2020, 26, 43. [Google Scholar] [CrossRef] [Green Version]
  58. Coclite, G.M.; di Ruvo, L. Dispersive and diffusive limits for Ostrovsky-Hunter type equations. NoDEA Nonlinear Differ. Equ. Appl. 2015, 22, 1733–1763. [Google Scholar] [CrossRef] [Green Version]
  59. LeFloch, P.G.; Natalini, R. Conservation laws with vanishing nonlinear diffusion and dispersion. Nonlinear Anal. 1999, 36, 213–230. [Google Scholar] [CrossRef]
  60. Schonbek, M.E. Convergence of solutions to nonlinear dispersive equations. Commun. Partial Differ. Equ. 1982, 7, 959–1000. [Google Scholar] [CrossRef]
  61. Coclite, G.M.; di Ruvo, L. Convergence of the Kuramoto-Sinelshchikov equation to the Burgers one. Acta Appl. Math. 2016, 145, 89–113. [Google Scholar] [CrossRef]
  62. Camps-Valls, G.; Martino, L.; Svendsen, D.H.; Campos-Taberner, M.; Muñoz-Mar, J.; Laparra, V.; Luengo, D.; Garca-Haro, F.J. Physics-aware Gaussian processes in remote sensing. Appl. Soft Comput. 2018, 68, 69–82. [Google Scholar] [CrossRef]
  63. Grigo, C.; Koutsourelakis, P.S. A physics-aware, probabilistic machine learning framework for coarse-graining high-dimensional systems in the Small Data regime. J. Comput. Phys. 2019, 397, 108842. [Google Scholar] [CrossRef] [Green Version]
  64. Taylor, M.E. Partial Differential Equations I. Basic Theory, 2nd ed.; Applied Mathematical Sciences; Springer: New York, NY, USA, 2011; Volume 115, p. xxii+65. [Google Scholar] [CrossRef]
  65. Coclite, G.M.; di Ruvo, L.; Ernest, J.; Mishra, S. Convergence of vanishing capillarity approximations for scalar conservation laws with discontinuous fluxes. Netw. Heterog. Media 2013, 8, 969–984. [Google Scholar] [CrossRef]
  66. Coclite, G.M.; di Ruvo, L. Existence results for the Kudryashov–Sinelshchikov–Olver equation. Proc. R. Soc. Edinb. Sect. A Math. 2020, 1–26. [Google Scholar] [CrossRef]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Coclite, G.M.; di Ruvo, L. A Note on the Solutions for a Higher-Order Convective Cahn–Hilliard-Type Equation. Mathematics 2020, 8, 1835. https://doi.org/10.3390/math8101835

AMA Style

Coclite GM, di Ruvo L. A Note on the Solutions for a Higher-Order Convective Cahn–Hilliard-Type Equation. Mathematics. 2020; 8(10):1835. https://doi.org/10.3390/math8101835

Chicago/Turabian Style

Coclite, Giuseppe Maria, and Lorenzo di Ruvo. 2020. "A Note on the Solutions for a Higher-Order Convective Cahn–Hilliard-Type Equation" Mathematics 8, no. 10: 1835. https://doi.org/10.3390/math8101835

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop