Time-asymptotic stability of composite weak planar waves for a general multi-D viscous system
Abstract.
We prove the time-asymptotic stability of the superposition of a weak planar viscous 1-shock and either a weak planar n-rarefaction or a weak planar viscous n-shock for a general multi-D viscous system. In 2023, Kang-Vasseur-Wang [11] showed the stability of the superposition of a viscous shock and a rarefaction for 1-D compressible barotropic Navier-Stokes equations and solved a long-standing open problem officially introduced by Matsumura-Nishihara [23] in 1992. Our work is an extension of [11], where a general multi-D viscous system is studied. Same as in [11], we apply the -contraction method with shifts, an energy based method invented by Kang and Vasseur in [9], for both viscous shock and rarefaction at the level of the solution. In such a way, we can work with general perturbations and compositions of waves. Finally, a technique to classify and control higher-order terms is developed to work in multi-D.
1. Introduction
We consider a general multi-D viscous system
(1.1) |
where with an open convex phase space . We assume the periodic boundary condition in the transverse direction . The notation for the physical space is with denoting the transverse direction . The flux functions are assumed to be smooth. In addition, the flux function is assumed to be strictly hyperbolic and genuinely nonlinear. The viscosity coefficient matrix is assumed to be smooth. For any , is assumed to be an positive definite matrix. We assume that the entropy is smooth and strictly convex, and for any , there exists an entropy flux of such that
(1.2) |
for any .
Let us endow the system with an initial value
with fixed end states , i.e.,
(1.3) |
This general framework includes the 3-D barotropic Brenner-Navier-Stokes equations:
(1.4) |
where , , is the density, is the velocity, and is the corrected velocity defined by
Such a system was introduced by Brenner as a correction of the Navier-Stokes equations. In [1], Brenner mentioned some extreme situations where the Navier-Stokes equations did not describe the compressible fluid well, and based on those results, Brenner proposed that the specific momentum density of the fluid equals the corrected velocity (the volume velocity) instead of the velocity (the mass velocity).
In this paper, we study the long-time asymptotic behavior of solutions of (1.1) with initial value satisfying (1.3). This long-time asymptotic behavior is closely related to the following 1-D Riemann problem:
(1.5) |
with an associated initial value
(1.6) |
Elementary solutions of the 1-D Riemann problem. Since is genuinely nonlinear, (1.5) has two elementary solutions: shock wave and rarefaction wave. We denote the eigenvalues of as and the corresponding right eigenvectors as . Let .
For any , there exists an integral curve such that for any that is close enough to , there exists a solution to (1.5) with initial value (1.6) that is defined by
with
for any -Riemann invariant . The existence of -Riemann invariants can be found in [28]. We call the i-rarefaction wave.
We define the shock set
In a neighborhood of , consists of n smooth curves that intersects at . We call the i-shock curve. For any that is close enough to , there exists a solution to (1.5) with initial value (1.6) that is defined by
We call the i-shock wave and the speed of the i-shock wave. In particular, is close to . We denote as .
Stability of elementary solutions for 1-D viscous systems. If the Riemann solution to an inviscid system is a rarefaction wave (respectively a shock wave), then the asymptotic state of the solution to the corresponding viscous system with a perturbed initial value is the rarefaction wave (respectively the viscous shock wave).
While the rarefaction wave spreads much faster than the diffusion process, the jump of the shock is smoothed by the viscosity and becomes a thin transition layer. The viscous i-shock wave refers to the smoothened i-shock wave, and satisfies
(1.7) |
Ilin-Oleinik [7] proved the stability of elementary solutions for the scalar case in 1960, but the maximum principle they used did not work for systems (see [20]). Later, the energy method has become the main tool. When studying the stability of rarefaction waves, the energy method can be applied directly at the level of the solutions. We call such a method the direct energy method. The stability of rarefaction waves was first proved by Matsumura-Nishihara [22, 23] for the 1-D compressible Navier-Stokes equations. Later, Liu-Xin [16] and Nishihara-Yang-Zhao [25] pushed the stability result of rarefaction waves to the Navier-Stokes-Fourier system.
However, the direct energy method fails when working with viscous shock waves. The first results were based on the energy method applied at the level of the antiderivative. Matsumura-Nishihara [21] in 1985, and Goodman [4] in 1986 independently proved the stability of viscous shock waves. Matsumura and Nishihara proved the stability for the 1-D compressible barotropic Navier-Stokes system, while Goodman proved the stability for a general 1-D system with a positive definite viscosity. As both papers worked with the antiderivative, they need to assume the zero mass condition, i.e., that the mass of the initial perturbation is zero.
Two very fruitful approaches removed the stringent zero mass condition. The Green’s function method was started by Liu [15] in 1985. It involves a constant shift on the viscous shock and the introduction of a diffusion wave and a coupled diffusion wave in the transverse characteristic fields. Szepessy-Xin [29] showed the stability of viscous shock waves for a 1-D general system with a nondegenerate artificial viscosity, and Liu-Zeng [17] applied the Green’s function method to 1-D systems with degenerate viscosity. Another approach is the Evans function method. In 2004, Mascia-Zumbrun [19] showed the spectral stability of viscous shock waves for the 1-D compressible Navier-Stokes system. The study of spectral stability is very advanced now. In 2017, Humpherys-Lyng-Zumbrun [6] proved the spectral stability of large-amplitude planar viscous shock waves for the compressible Navier-Stokes equations in multi-D by the numerical Evans function method. Although this technique gives stability only for single elementary waves, the stability result works in the whole space .
Both Green’s function method and Evans function method can deal with more general perturbations and provide pointwise estimates but fail to give global-in-time stability results for composite waves. Assuming the strengths of the two viscous shock waves are suitably small with the same order, Huang-Matsumura [5] showed the stability of two viscous shock waves for the 1-D Navier-Stokes-Fourier equations under a more relaxed condition than the zero mass condition.
Matsumura Conjecture. Even in 1-D, the stability of the composition of a viscous shock wave and a rarefaction wave is very difficult to study. Matsumura-Nishihara [22] mentioned the problem in 1986 and officially introduced it as an open problem in 1992 in [23]. In 2018, Matsumura [20] classified the problem as a very hard open problem. There are two difficulties. First, the direct energy method used to study the stability of rarefaction waves does not match very well with the methods developed for viscous shock waves. At the same time, the rarefaction wave is not an exact solution to the viscous system and any spatial shift of the rarefaction wave has the same asymptotic state. Hence, it is hard to analyze the interaction between the rarefaction wave and the viscous shock wave.
The breakthrough happened in 2023. Using the -contraction method with shifts, Kang-Vasseur-Wang proved the stability of the composition of a viscous shock wave and a rarefaction wave for the 1-D compressible barotropic Navier-Stokes equations in [11] and the stability of the generic Riemann solutions for the 1-D compressible Navier-Stokes-Fourier equations in [12]. The -contraction method with shifts is an energy method that can be applied at the level of the solution for contact waves, rarefaction waves, and viscous shock waves, so it unifies the methods for elementary waves. This paper is an extension of their work [11], in which the stability of a planar viscous 1-shock wave and either a planar n-rarefaction wave or a planar viscous n-shock wave is proved for a general multi-D viscous system. Note that we consider here only extremal waves.
Multi-D results. The stability of planar rarefaction waves for the 3-D compressible Navier-Stokes-Fourier system was shown by Li-Wang-Wang [14] in 2018. In 2023, the stability of planar viscous shock waves for the 3-D compressible Navier-Stokes equations was proved by Wang-Wang [30]. Later in 2024, Kang-Lee [8] generalized Wang-Wang’s stability result to the compositions of planar viscous shocks for the same system. Note that all these multi-D results are based on a-contraction method with shifts and work only with periodic transversal variables and weak elementary waves.
Result of the paper. Let be the smallest integer that is strictly bigger than , i.e.,
(1.8) |
The main result of the paper is the following theorem.
Theorem 1.1.
For any , there exist constants such that the following is true.
Let and or be such that
Let be the viscous 1-shock wave solution to (1.7) with end states and , and be the viscous n-shock wave solution to (1.7) or the n-rarefaction wave solution to (1.5) with end states and . Let be if and be if .
Let be an initial value such that
(1.9) |
where .
Then the viscous system (1.1) has a unique global-in-time solution . Moreover, there exist absolutely continuous shifts for such that
(1.10) | |||
(1.11) |
where
In addition,
(1.12) |
and for any ,
(1.13) |
Remark. Theorem 1.1 shows that if and are connected by a composition of a planar viscous 1-shock wave and either a planar viscous n-shock wave or a planar n-rarefaction wave, then the asymptotic state of the solution to the viscous system (1.1) with a perturbed initial value is the composition with viscous shocks shifted.
Proposition 1.2.
Remark. The detailed transformation is in the appendix. Proposition 1.2 implies that Theorem 1.1 can be applied to the 3-D barotropic Brenner-Navier-Stokes equations.
Structure of the paper. Section 2 discusses the properties of the viscous shock wave and the approximate rarefaction wave and introduces the weight functions, shift functions, and the superposition wave.
In section 3, we show how Theorem 1.1 is proved by local-in-time estimates and a priori estimates. Proposition 3.1 provides local-in-time estimates which could be shown in the same way as previous work, while a priori estimates are given in Proposition 3.3 which is to be shown in sections 4 and 5.
We get the estimates by the -contraction method with shifts in section 4 and go from the estimates to the estimates (a priori estimates) by induction in section 5.
The -contraction method with shifts. The method of -contraction relies on the ad-hoc construction of the shifts for solving special ODEs. For the scalar case, the method can be applied directly on the norm (see [9]). However, it was shown that the result is not true for systems in [27]. To work on systems, we need to introduce weight functions . In this paper, we follow the method of [11] written for the special case of the 1-D compressible barotropic Navier-Stokes equations. Our extension allows to clarify and explain why the method works at a deeper level.
To get the estimates, we study the evolution of a pseudo-distance given by the physical structure of the problem
where the relative entropy is defined by
(1.14) |
We define two bases at the beginning of subsection 4.3. The basis (4.3) (respectively (4.8)) is designed for the wave (respectively ). It contains the special direction (respectively ) of the wave (respectively ) and is orthogonal with respect to the viscous matrix in order to be consistent with the viscosity.
In Lemma 4.3, we apply the relative entropy method introduced by Dafermos [2] and DiPerna [3] and project the perturbation onto the bases. We get
The shift term represents the new terms induced by the shifts for . The viscosity operation (the right-hand side of (1.1)) gives the viscous term . The hyperbolic term comes from the flux functions . The interaction between waves and creates the interaction term .
Lemma 4.3 discusses the hyperbolic “scalarization”. The weight function (respectively ) activates the spectral gap, creates new negative hyperbolic terms, and initiates cancellation for hyperbolic terms corresponding to the perturbation in all directions except the special direction (respectively ) of the wave (respectively ). We get
where is a constant that depends on the strengths of , and represents the constants that depend only on . (respectively ) is the hyperbolic term corresponding to the projection of the perturbation in the special direction (respectively ) of the wave (respectively ), and denotes the sum of the absolute value of the hyperbolic terms corresponding to the other orthogonal directions of the perturbation. We see the hyperbolic “scalarization” scalarizes the problem to the special direction (respectively ) of the wave (respectively ).
The hyperbolic remainder due to the perturbation in the special direction of the rarefaction wave is negative, so the rarefaction wave is contractive at the hyperbolic level. However, the hyperbolic remainder due to the perturbation in the special direction of a viscous shock wave is positive and needs to be depleted using the viscous term . This is done by introducing a Poincaré type inequality in Lemma 4.4. For , we show
The viscous term controls the norm of the derivative, while the strengths of the shifts for are chosen to be big enough that the shift term handles the average with the help of .
In Lemma 4.5, we choose weight functions and shift functions for that work for both the hyperbolic “scalarization” and the Poincaré type inequality. We get
By Lemma 2.2 and Lemma 4.1, we can be bound the interaction term by a small and time integrable function depending on the strengths of the waves . In all, we obtain the estimates at the end of section 4.
Lemma 1.3.
For any satisfying ,
(1.15) |
Notations and a remark. Before we go to the proofs, let us fix some notations. Let the eigenvalues of be . For any , let and be the right and left eigenvectors corresponding to such that is tangent to , , and . We define
(1.16) |
Let denote the positive constants that depend only on . For , we define
(1.17) |
As is symmetry, we know
(1.18) |
2. Preliminaries
2.1. Viscous shock wave
Let . We examine the viscous i-shock wave satisfying
(2.1) |
Recall the Rankine-Hugoniot condition and the Lax inequality
Let the wave strength of be
The proof of the existence of the viscous i-shock wave can be found in [18]. The following results are proved in [13].
Lemma 2.1.
For any , there exist and for any such that the following is true.
For any such that , there exists a unique solution to (2.1) such that
We define
and the projection of onto
(2.2) |
From now on, we call the planar viscous i-shock wave. We know satisfies
(2.3) |
In [13], we show
(2.4) |
and
(2.5) |
In particular, is strictly increasing.
2.2. Construction of approximate rarefaction wave
As in [11], we will consider a smooth approximation of the planar n-rarefaction wave with the help of the smooth solution to the Burgers’ equation
(2.6) |
The smooth approximate planar n-rarefaction wave is defined by
(2.7) | ||||
where is the smooth solution to the Burgers’ equation (2.6) and is any -Riemann invariant to (1.5).
It is easy to check that is the solution to the inviscid system, i.e.,
(2.8) |
We define the wave strength of the rarefaction
and
(2.9) |
The following properties of the approximate planar n-rarefaction wave follow from the properties of the smooth solution to the Burgers’ equations proved in [22].
Lemma 2.2.
The smooth approximate n-rarefaction wave defined in (2.7) satisfies the following properties.
-
1)
and for any .
-
2)
For any and ,
where depends on and depends on .
-
3)
For any ,
-
4)
In particular, is strictly increasing,
(2.10) |
and
(2.11) | |||
For convenience, we call the planar n-rarefaction wave from now on.
2.3. Construction of weight functions, shift functions and the superposition wave
We are ready to introduce the weight functions, the shift functions and the superposition wave.
Let be either the planar viscous n-shock wave or the planar n-rarefaction wave . Recall (2.2) and (2.9). We define the weight functions by
(2.12) | ||||
for large enough constants that depend only on . As are increasing, the weight function is decreasing and the weight function is increasing. Also, we have by taking small enough.
For any function and , we define
We define the superposition wave
(2.13) |
and the weight function
(2.14) |
where are defined in (2.12), and the shift is defined as the solution to the ODE
(2.15) |
for a large enough constant that depends only on for if and for if . As the planar n-rarefaction wave is not shifted, we define if for consistency. The existence and uniqueness of shifts are proved in Proposition 3.1.
(2.16) |
where
if , then
and if , then
The terms are error terms caused by the fact that is not an exact solution of the system (1.1). The term comes from the shifts. We see depend on the choice of , because while the planar viscous shock wave is shifted and is a solution to the viscous system (2.3), the planar rarefaction wave is not shifted and is a solution to the inviscid model (2.8).
3. Proof of Theorem 1.1
First, we introduce local-in-time estimates in Proposition 3.1 and a priori estimates in Proposition 3.3. Then we discuss how the two propositions prove the global existence and the asymptotic behavior results stated in Theorem 1.1.
3.1. Local-in-time estimates and a priori estimates
Proposition 3.1.
For any and any , there exists that depends on such that the following is true.
The local-in-time existence and uniqueness of the solution can be done similarly as in Serre’s paper [26]. Since the viscosity matrices of our system (1.1) are positive definite, the proof will be simpler. The existence and uniqueness of shifts can be shown in the same way as in subsection 3.3 of [11].
Before introducing a priori estimates, we first establish the assumption of a priori estimates. Since sections 4 and 5 give the proof of a priori estimates, the following assumption will be the assumption of all lemmas in both sections.
Assumption 3.2.
Proposition 3.3.
For any , there exist such that the following is true.
3.2. Global existence and estimates
We define
We fix as in Proposition 3.3. We take the strength of the initial perturbation and the wave strength in Theorem 1.1 and in Proposition 3.1 to be small enough such that
(3.7) |
We define
Assume . Then Proposition 3.3 gives
By Proposition 3.1, there exists such that
Contradiction! Hence, we get and (1.10) in Theorem 1.1. Now we can apply Proposition 3.3 on and get
(3.8) |
and for any such that ,
(3.9) | ||||
In addition,
(3.10) |
3.3. Time-asymptotic behavior
Let be such that . We define
We show the classical estimate
(3.11) |
By (3.8), we get
By (3.9), we have
The classical estimate (3.11) gives
The Gagliardo–Nirenberg inequality proved in [24], the periodicity in the transverse direction, and (3.8) give
(3.12) |
By (3.12) and Lemma 2.2, we get (1.11) in Theorem 1.1. By (3.10) and (3.12), we get (1.13) in Theorem 1.1.
4. Energy estimate
We show the estimates by the -contraction method with shifts in this section. Later in section 5, induction will help us get the estimates and finish the proof of Proposition 3.3.
First, we develop tools to handle interaction terms and higher-order terms in subsections 4.1 and 4.2. Then we apply the -contraction method with shifts. The hyperbolic “scalarization” is discussed in subsection 4.3. The positive hyperbolic remainder corresponding to the special direction of the planar viscous shock wave motivates the Poincaré type inequality introduced in subsection 4.4. Finally in subsection 4.5, we choose the constants defining weight functions and shift functions in a way that makes both hyperbolic “scalarization” and Poincaré type inequality work.
4.1. Wave interaction estimates
To control the interaction between waves, the idea is to take the shifts small enough that the main layer regions do not overlap.
4.2. Higher derivatives estimates
Let
(4.1) |
For any , we define
(4.2) |
Lemma 4.2 plays an important role in section 5 when working with higher derivatives. It singles out terms that need to be handled differently and unifies the rest in the same form. The last two inequalities in the lemma will be used in the proof of Lemma 4.3 to evaluate the interaction terms induced by .
Lemma 4.2.
Proof.
For any and any function , we define
We show the first inequality. If we apply the chain rule to
then either all fall on or some fall on . Therefore, we know that
is bounded by the sum of the absolute value of the terms in the following form
where , , , , and is some derivative of . By (3.2), Lemma 2.1, and Lemma 2.2, we have
Note the constants in this proof depend on , but the dependency on does not matter as is fixed and finite.
4.3. Relative entropy method
For each layer corresponding to and , we will construct a basis of the phase space that is well-adapted to both the special direction of the wave (respectively ) and the dissipation matrix . As the dissipation takes place in the so-called entropic variables , we project such quantity onto the bases. Since is small, we know
Hence, the corresponding natural special direction of the wave for the entropic variables is (respectively ). The point is to work with a basis that is orthogonal with respect to the dissipation matrix and contains (respectively ).
When working with the layer of , we complete into an orthogonal basis of with respect to the dissipation. In particular, we choose a basis such that for any and any ,
where . This is always possible thanks to the Gram-Schmidt process. We project the perturbation in entropic variables onto this basis:
(4.3) |
When working with the hyperbolic terms, we will work with the conserved quantity . By Taylor expansion and (3.2), we have
(4.4) |
As is a basis, we have
(4.5) |
for some constants that depend only on .
Let be the projection onto , i.e.,
(4.6) |
For any ,
As , there exists such that
(4.7) |
When working with the layer of , we complete into an orthogonal basis of with respect to the dissipation. In particular, we choose a basis such that for any and any ,
We project the perturbation in entropic variables onto this basis:
(4.8) |
Such projection has similar properties as (4.3).
We choose such bases to make the best use of dissipation in the special directions of the planar shock waves. In the case that is a planar rarefaction wave, we could work with the natural hyperbolic basis since we do not need the Poincaré inequality. However, for the sake of consistency, we will use the basis in the case too.
Relative functions. The relative flux is defined by
(4.9) | ||||
The flux of the relative entropy is defined by
(4.10) |
where , the entropy flux of , is defined in (1.2).
We want to study the evolution of the weighted relative entropy
It involves controlling layer quantities of the form:
In each of these layers, a “scalarization” effect takes place. Such effect damps the perturbation in the transverse direction well-adapted to both the special direction of the wave (respectively ) and the diffusion eigen-direction.
The main lemma of this section is the following. We postpone our choice of to subsection 4.5, where we unify the choice of constants for both hyperbolic “scalarization” and Poincaré inequality.
Lemma 4.3.
For any , any , and any , there exist such that the following is true.
Remark. Recall (1.16). We have
We have four families of terms: the shift family , the viscous family , the hyperbolic family , and the interaction family .
The shift family corresponds to the new terms induced by the shifts. The viscous family comes from the dissipation. The viscous term corresponds to the transverse direction, and the viscous terms correspond to the direction. If we examine in the phase space, corresponds to the special direction (respectively ) of the wave (respectively ), and corresponds to the other orthogonal directions. In Lemma 4.4 in the next section, we see how helps us get the Poincaré type inequality.
The flux functions give the hyperbolic family . The hyperbolic term (respectively ) corresponds to the special direction (respectively ) of the wave (respectively ). The hyperbolic term corresponds to the other orthogonal directions. The hyperbolic “scalarization” happens in all directions except the special direction (respectively ) of the wave (respectively ). More precisely, the coefficient of becomes negative if we choose sufficiently large , while the coefficients of are independent of the choice of . In short, the hyperbolic “scalarization” reduces the problem to the scalar case with hyperbolic remainders in the form of .
Such hyperbolic remainder is negative if it corresponds to a planar rarefaction wave, but it is positive if it corresponds to a planar viscous shock wave. In Lemma 4.4 in the next section, we show in detail how the dissipation , the shift family , and the hyperbolic terms corresponding to the other orthogonal directions control the positive hyperbolic remainder by the Poincaré type inequality. While the dissipation dominates the norm of the derivative, the shifts will be chosen in a way that the shift family controls the average with the help of the hyperbolic terms corresponding to the other orthogonal directions .
Proof.
By the definition of relative entropy function (1.14), (1.1), and (2.16), we have
According to definitions of (4.9) and (4.10), we have
where the shift term is
the dissipation gives
the wave interaction causes
and the flux induces
We are going to analyze those terms and get defined in the lemma.
Hyperbolic parts and dissipation in . As and are planar waves, and do not depend on . Also, we know are positive definite and is strictly convex. Therefore, we have
Dissipation in . By integration by parts, we have
We apply projections (4.3) and (4.8) to . By (3.2) and the fact that is positive definite, we take small enough and get
Compared with , and give extra smallness to .By Young’s inequality, (2.5), and (2.10), we take small enough and get
We have
which implies
As , we get
for some . Thus there exists some constant that depends only on such that
Taking small enough, we get
By the chain rule, Lemma 2.1, Lemma 2.2, and (3.2), we have
Hyperbolic parts in . By the definition of the flux of the relative entropy (4.10), (3.2), (2.5), and (2.10), we take small enough and get
where
We apply (4.4) to . Recall (4.6). Let
By (1.18), (4.4), (4.5), and (4.7), we take small enough and get
where
Similarly, we get
where
As , we get
By the definition of the relative flux (4.9) and (3.2), we take small enough and get
where
We apply (4.4) to . As the hyperbolic “scalarization” will happen in all the directions except the special direction of the wave , we use Young’s inequality to make the hyperbolic term in the direction as small as possible
Similarly, we get
Then we have
In all, we get
Shift terms. We have
By (2.5) and (2.10), we take small enough and get
where
By Taylor expansion and Lemma 2.2, we take small enough and get
In all, we get
4.4. Poincaré type inequality
Lemma 4.4.
For any , any , and any , there exist such that the following is true.
Remark. By taking large enough, the positive remainders are and . We will take large enough to control and small enough to control in Lemma 4.5 in subsection 4.5.
Proof.
First, we fix the time and the transverse direction and compactify the problem by changing variables. Let
As is strictly increasing, we define the change of variable
By (2.4), we can take small enough and get
Lemma 1.3 gives
which implies
where is independent of . Therefore, we have
Using the classical Poincaré inequality, we find
By Young’s inequality, Hölder’s inequality, (2.5), and (3.2), we take small enough and get
where if and if for any .
4.5. estimates
Lemma 4.5.
For any , there exist such that the following is true.
5. Proof of Prop 3.3
5.1. Some estimates
Before proving the estimates, we bound the higher-order terms evaluated in Lemma 4.2.
Recall defined in (1.8). Let be a function. The Gagliardo-Nirenberg inequality for functions defined on is proved in [24]. Together with the periodicity in the transverse direction, we get the following two inequalities. If , then for any , there exists such that
(5.1) |
If , then for any , there exists such that
(5.2) |
Lemma 5.1.
Proof.
If , then
Assume . Without loss of generality, assume . For any , (5.1) and (3.1) give
(5.3) |
for any . We will consider 3 cases.
2) case and . Then and for any . By (5.2), we have
for any . Since and ,
Hence, there exist such that
Then Hölder’s inequality and (5.3) give
3) case and . Then , , and . Hölder’s inequality, Gagliardo-Nirenberg inequality and (3.1) give
∎
5.2. estimates
We prove the estimates by induction.
Lemma 5.2.
For any , there exist such that the following is true.
Proof.
We will prove the lemma by induction. Recall the estimates shown in Lemma 4.5. Let . We assume the inequality is true for , i.e.,
We want to show the inequality is true for , i.e.,
(5.6) |
Note the constants in this proof depend on , but the dependency on does not matter as is fixed and finite.
Recall the definition of in (4.1). Let be such that . We take on both sides of (2.17) and get
Then we multiply both sides by and take integration w.r.t. . Recall (1.17). By integration by parts, we have
where is the multi-index whose j-th component is and all other components are 0.
Let to be chosen later in (5.8). Recall (1.17). By Young’s inequality and integration by parts, we have
where
where
We first evaluate . We have
where
As are positive definite and is strictly convex,
By (3.2) and Young’s inequality, we get
Therefore, we have
If , then Lemma 4.2 gives
If , then Lemma 4.2 gives
for some .
We get
(5.7) | |||
Let
As is arbitrary, we get
where are constants that depend only on . We take
(5.8) |
Finally, we take integration over time. By (2.11) and Lemma 4.1, we get the estimates (5.6). Hence, we prove (5.4). The inequality (5.7) also gives
By taking integration over time, (2.11) and Lemma 4.1 give (5.5). ∎
6. Appendix
6.1. Proof of Proposition 1.2
The entropy of the 3-D barotropic Brenner-Navier-Stokes equations is
We let
We rewrite the 3-D barotropic Brenner-Navier-Stokes equations (1.4) w.r.t. :
As
we get
and for any ,
Hence, for any ,
Since the determinant of is , is positive definite.
References
- [1] H. Brenner. Fluid mechanics revisited. Physica A: Statistical Mechanics and its Applications, 370(2):190–224, 2006.
- [2] C. M. Dafermos. Entropy and the stability of classical solutions of hyperbolic systems of conservation laws. In Recent mathematical methods in nonlinear wave propagation (Montecatini Terme, 1994), volume 1640 of Lecture Notes in Math., pages 48–69. Springer, Berlin, 1996.
- [3] R. J. DiPerna. Uniqueness of solutions to hyperbolic conservation laws. Indiana Univ. Math. J., 28(1):137–188, 1979.
- [4] J. Goodman. Nonlinear asymptotic stability of viscous shock profiles for conservation laws. Arch. Rational Mech. Anal., 95(4):325–344, 1986.
- [5] F. Huang and A. Matsumura. Stability of a composite wave of two viscous shock waves for the full compressible Navier-Stokes equation. Comm. Math. Phys., 289(3):841–861, 2009.
- [6] J. Humpherys, G. Lyng, and K. Zumbrun. Multidimensional stability of large-amplitude Navier-Stokes shocks. Arch. Ration. Mech. Anal., 226(3):923–973, 2017.
- [7] A. M. Ilin and O. A. Oleinik. Asymptotic behavior of solutions of the Cauchy problem for some quasi-linear equations for large values of the time. Mat. Sb. (N.S.), 51(93):191–216, 1960.
- [8] M.-J. Kang and H. Lee. Long-time behavior toward composite wave of shocks for 3D barotropic Navier-Stokes system. arXiv 2406.11215, 2024.
- [9] M.-J. Kang and A. Vasseur. Criteria on contractions for entropic discontinuities of systems of conservation laws. Arch. Ration. Mech. Anal., 222(1):343–391, 2016.
- [10] M.-J. Kang and A. Vasseur. Contraction property for large perturbations of shocks of the barotropic Navier-Stokes system. J. Eur. Math. Soc. (JEMS), 23(2):585–638, 2021.
- [11] M.-J. Kang, A. Vasseur, and Y. Wang. Time-asymptotic stability of composite waves of viscous shock and rarefaction for barotropic Navier-Stokes equations. Adv. Math., 419:Paper No. 108963, 66, 2023.
- [12] M.-J. Kang, A. Vasseur, and Y. Wang. Time-asymptotic stability of generic Riemann solutions for compressible Navier-Stokes-Fourier equations, 2023.
- [13] Y.-S. Kwon, J. Meng, and A. Vasseur. -contraction applied to the time-asymptotic stability of weak shocks for a general 1-d viscous system. In preparation, 2024.
- [14] L.-A. Li, T. Wang, and Y. Wang. Stability of planar rarefaction wave to 3D full compressible Navier-Stokes equations. Arch. Ration. Mech. Anal., 230(3):911–937, 2018.
- [15] T.-P. Liu. Nonlinear stability of shock waves for viscous conservation laws. Mem. Amer. Math. Soc., 56(328):v+108, 1985.
- [16] T.-P. Liu and Z. P. Xin. Nonlinear stability of rarefaction waves for compressible Navier-Stokes equations. Comm. Math. Phys., 118(3):451–465, 1988.
- [17] T.-P. Liu and Y. Zeng. Shock waves in conservation laws with physical viscosity. Mem. Amer. Math. Soc., 234(1105):vi+168, 2015.
- [18] A. Majda and R. L. Pego. Stable viscosity matrices for systems of conservation laws. J. Differential Equations, 56(2):229–262, 1985.
- [19] C. Mascia and K. Zumbrun. Stability of small-amplitude shock profiles of symmetric hyperbolic-parabolic systems. Comm. Pure Appl. Math., 57(7):841–876, 2004.
- [20] A. Matsumura. Waves in compressible fluids: viscous shock, rarefaction, and contact waves. In Handbook of mathematical analysis in mechanics of viscous fluids, pages 2495–2548. Springer, Cham, 2018.
- [21] A. Matsumura and K. Nishihara. On the stability of travelling wave solutions of a one-dimensional model system for compressible viscous gas. Japan J. Appl. Math., 2(1):17–25, 1985.
- [22] A. Matsumura and K. Nishihara. Asymptotics toward the rarefaction waves of the solutions of a one-dimensional model system for compressible viscous gas. Japan J. Appl. Math., 3(1):1–13, 1986.
- [23] A. Matsumura and K. Nishihara. Global stability of the rarefaction wave of a one-dimensional model system for compressible viscous gas. Comm. Math. Phys., 144(2):325–335, 1992.
- [24] L. Nirenberg. On elliptic partial differential equations. Ann. Scuola Norm. Sup. Pisa Cl. Sci. (3), 13:115–162, 1959.
- [25] K. Nishihara, T. Yang, and H. Zhao. Nonlinear stability of strong rarefaction waves for compressible Navier-Stokes equations. SIAM J. Math. Anal., 35(6):1561–1597, 2004.
- [26] D. Serre. Local existence for viscous system of conservation laws: -data with . In Nonlinear partial differential equations and hyperbolic wave phenomena, volume 526 of Contemp. Math., pages 339–358. Amer. Math. Soc., Providence, RI, 2010.
- [27] D. Serre and A. Vasseur. -type contraction for systems of conservation laws. J. Éc. polytech. Math., 1:1–28, 2014.
- [28] J. Smoller. Shock waves and reaction-diffusion equations, volume 258 of Grundlehren der mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]. Springer-Verlag, New York, second edition, 1994.
- [29] A. Szepessy and Z. P. Xin. Nonlinear stability of viscous shock waves. Arch. Rational Mech. Anal., 122(1):53–103, 1993.
- [30] T. Wang and Y. Wang. Nonlinear stability of planar viscous shock wave to three-dimensional compressible Navier-Stokes equations. arXiv 2204.09428, To appear in JEMS, 2022.