Time-asymptotic stability of composite waves of viscous shock and rarefaction for barotropic Navier-Stokes equations
Abstract.
We prove the time-asymptotic stability of composite waves consisting of the superposition of a viscous shock and a rarefaction for the one-dimensional compressible barotropic Navier-Stokes equations. Our result solves a long-standing problem first mentioned in 1986 by Matsumura and Nishihara in [26]. The same authors introduced it officially as an open problem in 1992 in [27] and it was again described as very challenging open problem in 2018 in the survey paper [24]. The main difficulty is due to the incompatibility of the standard anti-derivative method, used to study the stability of viscous shocks, and the energy method used for the stability of rarefactions. Instead of the anti-derivative method, our proof uses the -contraction with shifts theory recently developed by two of the authors. This method is energy based, and can seamlessly handle the superposition of waves of different kinds.
Key words and phrases:
compressible Navier-Stokes equations, viscous shock wave, rarefaction wave, shift, -contraction, stability
1. Introduction
Consider the one-dimensional compressible barotropic Navier-Stokes equations. In the Lagrangian mass coordinates, the system is described as
(1.1) |
where the unknown functions , and represent respectively the specific volume, and the velocity of the gas. The pressure function is given by the well-known law
where are both constants depending on the fluid, and the constant corresponds to the viscosity coefficient. Without loss of generality, we normalize two of the constants as and . The system is then endowed with initial values:
We consider initial values with fixed end states , that is such that
(1.2) |
On top of its physical relevance, system (1.1) can be seen as the typical example of viscous conservation laws involving a physical viscosity. The large-time behavior of solutions to (1.1), with initial values verifying (1.2), is closely related to the Riemann problem of the associated Euler equations:
(1.3) |
with the Riemann initial data
(1.4) |
corresponding to the end states (1.2). In the scalar case (where the system (1.1) is replaced by a single viscous equation), the time-asymptotic stability of the viscous waves, and their link to the inviscid problem was first proved in 1960 by Ilin-Oleinik [9] (see also Sattinger [32]). The case for systems as (1.1) is far more difficult (see [24]). One of the motivation for the study of large-time behavior of solutions to compressible Navier-Stokes equation for Riemann initial data was to obtain insights about inviscid limit to the Euler equation. In 2005 [1] , Bianchini-Bressan showed, for small BV initial values, the convergence at the inviscid limit of solution to parabolic system with “artificial viscosity” to the unique solution of the associated hyperbolic system. However, to this day, the result is still unknown for the physical Navier-Stokes system, even in the barotropic case (1.1).
Riemann problem for the inviscid model: Let us first describe the well-known solution of the Riemann problem for the inviscid model (1.3)-(1.4), first proposed and solved by Riemann [31] in 1860s. This system of conservation laws is strictly hyperbolic. This means that the derivative of the flux function with respect to the conserved variables, about a fixed state :
is diagonalizable with real distinct eigenvalues. Note that this matrix defined the waves generated by the linearization of the system (1.3) about this fixed state . Its eigenvalues and generate both characteristic fields which are genuinely nonlinear. Therefore, the self-similar solution, so called Riemann solution, of the Riemann problem is determined by a combination of at most two elementary solutions from the following four families: 1-rarefaction; 2-rarefaction; 1-shock and 2-shock (see for instance [4]). These families are completely defined through their associated curves in the states plane . For any , the 1-rarefaction curve corresponds to the integral curve of the first eigenvalue , and is defined by
(1.5) |
The 2-rarefaction curve can be defined in the same way from the second eigenvalue . For any initial values of the Riemann problem (1.4) with , , such that , the solution of (1.3) is the 1-rarefaction wave defined as
(1.6) |
together with
(1.7) |
where is called the 1-Riemann invariant to the Euler equation (1.3). The case of 2-rarefaction wave is treated similarly from the second eigenvalue . We can now define the shock curves using the Rankine-Hugoniot condition, as the one-parameter family of all the such that there exists with:
(1.8) |
The general theory shows that this condition defines actually 2 curves that meet at the point , one for the value (the 1-shock curve which corresponds to admissible shocks for ), and one for the value (the 2-shock curve with admissible shocks for ). Whenever , the solution to (1.3)-(1.4) with , , is given by the discontinuous traveling wave defined as
(1.9) |
For the general case of any states , it can be shown that there exists a (unique) intermediate state such that is on a curve of the second families from (either or ), and is on a curve of the first families from (either or ). The solution of (1.3)-(1.4) is then obtained by the juxtaposition of the two associated waves
The wave is 1-rarefaction fan solution to (1.6)-(1.7) if , or 1-shock solution to (1.9) if , with , . The wave is 2-shock solution to (1.9) if , or, 2-rarefaction fan solution if , both with the end states , . Note that the cases of single wave are included as degenerate cases when , or . Previous time-asymptotic results for the viscous model: The time-asymptotic behavior of the viscous solution to (1.1) depends on whether the associated Riemann solution to the associated inviscid model (1.3)-(1.4) involves shock waves or rarefaction waves. In the case where (1.4) is a shock, the viscous counterpart for (1.1), called viscous shock, is the traveling wave defined by the following ODE:
(1.10) |
Matsumura-Nishihara [25] proved the stability of the viscous shock waves (1.10) for the compressible Navier-Stokes equations (1.1). Independently, Goodman showed in [7] the same result of a general system with artificial diffusion. This corresponds to the case where diffusion is added to all the equations of the system. In both papers, the proof were done under the zero mass condition which is crucial for using the so called anti-derivative method. Then Liu [20], Szepessy-Xin [36] and Liu-Zeng [22] removed the crucial zero mass condition in [25, 7] by introducing the constant shift on the viscous shock and the diffusion waves and the coupled diffusion waves in the transverse characteristic fields. Masica-Zumbrun [23] proved the spectral stability of viscous shock to 1D compressible Navier-Stokes system under a spectral condition, which is slightly weaker than the zero mass condition. The case of the superposition of two shocks for the Navier-Stokes-Fourier system was treated by Huang-Matsumura in [8]. Finally, the asymptotic stability of viscous shocks for Navier-stokes systems with degenerated viscosities was studied in Matsumura-Wang [28], and generalized to a larger class of viscosity in [39] using the BD entropy introduced by Bresch-Desjardins in [2]. The treatment of stability of rarefactions is performed with very different techniques based on direct energy methods. Matsumura-Nishihara [26, 27] first proved the time-asymptotic stability of the rarefaction waves for the compressible and isentropic Navier-Stokes equations (1.1). It was later generalized to the Navier-Stokes-Fourier system by Liu-Xin [21] and Nishihara-Yang-Zhao [30]. The case of the juxtaposition of a shock and a rarefaction: However, the time-asymptotic stability of the superposition of a viscous shock wave and a rarefaction wave has been an open problem up to now. The main difficulty is that the classical anti-derivative method used for the stability of shocks does not match well with the energy method classically used for the stability of rarefactions. The problem of the stability of such a superposition of a rarefaction and a viscous shock was first mentioned in 1986 by Matsumura and Nishihara in [26]. The same authors introduced it officially as an open problem in 1992 in [27] and Matsumura described it again as very challenging open problem in 2018 in the survey paper [24]. Our main theorem is proving this conjecture.
Theorem 1.1.
For a given constant state , there exist constants such that the following holds true.
For any and such that
denote the 1-rarefaction solution to (1.3) with end states and , and the 2-viscous shock solution of (1.10) with end states and . Let be any initial data such that
(1.11) |
where .
Then, the compressible Navier-Stokes system (1.1) admits a unique global-in-time solution . Moreover, there exists an absolutely continuous shift such that
(1.12) |
In addition, as ,
(1.13) |
and
(1.14) |
Remark 1.1.
Theorem 1.1 states that if the two far-field states in (1.2) are connected by the superposition of shock and rarefaction waves, then the solution to the compressible Navier-Stokes equations (1.1) tends to the composition wave of the self-similar rarefaction wave and the viscous shock wave with the shift , which solves the open problem proposed by Matsumura-Nishihara [27] since 1992.
Remark 1.2.
Remark 1.3.
Note that our result in Theorem 1.1 also holds true in the case of a single viscous shock, that is, . In this case, Theorem 1.1 provides an alternative proof for stability of a single viscous shock. Our proof is far simpler than the ones of Masica-Zumbrun [23], or Liu-Zeng [22]. Moreover, our approach does not have the disadvantages of the anti-derivative method, as the necessity to consider zero mass initial perturbations for instance. This simplification is what allows us to consider the combination of waves of different kinds. Therefore, our approach follows exactly the suggestion of Matsumura in [24, Section 4.2, page 2540] to find a simpler proof, for the stability of viscous shock, than the ones in [23] or [22], in order to attack many other open problems. Note however, that our simplification comes at the cost of less precise information, especially on the shift .
Remark 1.4.
The extension of Theorem 1.1 to general smooth viscosity function and general pressure function satisfying follows without meaningful added difficulties, since we consider small -perturbations. For the sake of clarity, and to simplify slightly the arguments, we made the choice to write the paper in the physical relevant context of constant viscosities and power pressure laws.
The main new ingredient of our proof is the use of the method of -contraction with shifts [15] to track the stability of the viscous shock. The method is based on the relative entropy introduced by Dafermos [5] and DiPerna [6]. It is energy based, and so meshes seamlessly with the treatment of the rarefaction. The method of -contraction with shifts: The method of -contraction with shifts was developed in [12] (see also [19]) to study the stability of extremal shocks for inviscid system of conservation laws, as for example, the Euler system (1.3). Consider the entropy of the system (which is actually the physical energy) defined for any state as:
We then consider the relative entropy defined in [5] for any two states , :
Note that is convex, and so is nonnegative and equal to zero if and only if . Therefore can be used as a pseudo-distance between and . It can be shown that rarefactions (that is solutions to (1.6)-(1.7)) have a contraction property for this pseudo-metric (see for instance [37]). Indeed, for any weak entropic solution to (1.3), it can be shown that
The contraction property is not true if is a shock (that is traveling waves (1.9) verifying the Rankine-Hugoniot conditions (1.8)). However, the contraction property can be recovered up to a shift, after weighting the relative entropy (see [12]). Indeed, there exists weights (depending only on the shock ) such that for any weak entropic solution of (1.3) (verifying a mild condition called strong trace property) there exists a Lipschitz shift function such that
This was first proved in the scalar case by Leger [18] for . It has been shown in [33] that the contraction with is usually false for most systems. Therefore the weighting via the coefficients is essential. Note that in the case of the full Euler system, the a-contraction property up to shifts is true for all the single wave patterns, including the 1-shocks, 3-shocks (see [38]), and the 2-contact discontinuities (see [34]). Although the -contraction property with shifts holds for general extremal shocks, it is not always true for intermediate shocks (see [10] for instance). The first extension of the method to viscous models was done in the 1D scalar case [13] (see also [11]) and then in the multi-D case [17]. The case of the barotropic Navier-Stokes equation (1.1) was treated in [15]. The -contraction property takes place in variables associated to the BD entropy (see [2]): , where is the effective velocity defined as . In these variables, system (1.1) with is transformed as
(1.15) |
The only nonlinear term of the hyperbolic system (1.3) is the pressure which is a function of . The system (1.15) is then better than (1.1) since the diffusion is in the variable corresponding to the nonlinear term . It was shown in [15] that there exists a monotonic function (with limits at ), depending only on the viscous shock solution to (1.10) (in the variables), such that for any solution to (1.15), there exists a shift function with
The strategy of this paper is to apply the -contraction method to the composite wave made of a shock wave and a rarefaction wave. The weight function and the shift is applied only on the shock wave. The combination of the viscous shock and the rarefaction is not an exact solution to (1.1). This introduces some errors that can be controlled thanks to the separation of the waves. Because of the shift, the separation of the waves is not automatic. We show, however, that it is still true, and that the shock cannot artificially stick to the rarefaction. This provides an ”almost” -contraction in the effective variables . We then recover the classical control on the norm of the perturbation in the classical variables .
The -contraction with shift theory for a small viscous shock: Note that the -contraction result of [15] provides a uniform stability for viscous shocks with respect to the strength of the viscosity. This is used in [16] to obtain the stability of inviscid shocks of (1.3) among any inviscid limits of (1.1). Since the conjecture of Matsumura [24] does not mention the uniform stability with respect to the viscosity, we choose in this paper to restrict ourselves to the classical framework and show the stability with fixed. This allows us to simplify some of the arguments of [15] in this context. The method is even more powerful in this restricted framework and should be developed in the foreseeable future to a large family of problems. Let us describe the fundamental ideas in this context. A Poincaré type inequality and the scalar case: At its core, the method of -contraction with shift in the viscous cases relies on the following Poincaré type inequality (see [15, Lemma 2.9]).
Lemma 1.1.
For any satisfying ,
(1.16) |
The eigenfunctions of the associated Euler-Lagrange equation to this minimization problem are the Legendre polynomials, and their eigenvalues are given explicitly. As a consequence, the inequality is sharp. The weighted norm of this inequality comes naturally when considering the following Burgers equation (see [13]):
(1.17) |
and its viscous shock profile defined as
This shock does not depend on time (it is a stationary wave). Integrating in , and denoting gives
(1.18) |
Consider now the relative entropy associated to the entropy between a generic solution of (1.17) and the shifted shock for an arbitrary shift :
The shifted shock verifies the equation
Multiplying the difference of (1.17) and the shifted shock equation by , we can show that
(1.19) |
Note that, at the final step, we made the change of variable flipping the shift from the shock to the function . We now fix the speed of the shift as
which defines the shift thanks to the Cauchy-Lipschitz theorem. We claim that, for this shift, is non-increasing in time. This statement will be proved, if we can show that for any function :
(1.20) |
where . Indeed, for any fixed time , denote . The inequality (1.20) for this specific function applied to (1.19) shows that at all these times:
Therefore, the contraction up to a shift is reduced to the Poincaré type inequality (1.20). Because , it is equivalent to
Let us rewrite this inequality in the natural variable associated to the shock:
This change of variable is possible since is an increasing function from 0 to 1. We have also
and so (1.20) is equivalent to
But thanks to (1.18), . Hence (1.16) implies (1.20) since . The case of the Navier-Stokes system: If we perform the same idea on the Navier-Stokes system (1.15) in the BD effective variables , but without weight function , we are obtaining (after Taylor expansion, using the smallness of the shock and of the perturbation) the inequality
with
Thanks to the BD effective variables, the first equality is very similar to the scalar case. Especially, the dissipation is in the variable only, as the “bad” quadratic term. However, the term involves now a linear combination of and . Therefore, whatever the choice of , we cannot control any weighted mean value of from this term as in the scalar case. The point of the method is that the flux of the relative entropy (which disappears when integrating in ) is better behaved. On top of a “bad” quadratic term in , it involves a “good” (meaning with a good sign) quadratic term involving a linear combination of and . The weight function is used to activate those flux terms. Note that the linear combination involved in the flux terms is independent of the linear combination involved in the term. Therefore the use of both the weight and the shift allows to control the weighted mean value of needed to use the Poincaré Lemma 1.1. The weight function is chosen such that is proportional to which is the analogue of for the scalar case, and is a natural weight associated to the shock layer. Its strength, however, is enhanced by a factor bigger than the size of the shock , in order to make the relative entropy flux term dominant.
The rest of the paper is organized as follows. We begin with preliminaries in Section 2. It includes known properties on the rarefaction and on the viscous shock, together with simple properties on the behavior of the pressure functional and the relative entropy. The general set up is laid out in Section 3. We introduce an a priori estimates result in Proposition 3.2. Then we show by a continuing argument how this proposition implies Theorem 1.1. The last two sections are dedicated to the proof of Proposition 3.2. The -contraction argument is set up in Section 4 where global a priori estimates are proved in the variables . From these global estimates, we deduce global a priori estimates in the variables in Section 5, concluding the proof of Proposition 3.2.
2. Preliminaries
We gather in this section some well-known results which will be useful in the rest of the paper.
2.1. Relative quantities
As explained in the introduction, the -contraction with shifts theory is based on the relative entropy, and the specific volume variable is of particular importance. For any function defined on , we define the associated relative quantity defined for as
We gather, in the following lemma, useful explicit inequalities on the relative quantities associated to the pressure , and the internal energy . The proofs are simply based on Taylor expansions, and can be found in [15, Lemmas 2.4, 2.5 and 2.6].
Lemma 2.1.
For given constants , and , their exist constants , such that the following holds true.
1) For any such that ,
(2.1) |
(2.2) |
2) For any ,
(2.3) |
3) For any , and for any satisfying , and , the following holds true:
(2.4) |
(2.5) |
(2.6) |
2.2. Rarefaction wave
We now recall important properties of the 1-rarefaction waves. Consider a in (1.2), and . Set , and consider the Riemann problem for the inviscid Burgers equation:
(2.7) |
If , then (2.7) has the rarefaction wave fan given by
(2.8) |
It is easy to check that the 1-rarefaction wave to the Riemann problem (1.3)-(1.4), defined in (1.6)-(1.7), is given explicitly by
(2.9) |
The self-similar 1-rarefaction wave is Lipschitz continuous and satisfies the Euler system a.e. for ,
(2.10) |
Let denote the strength of the rarefaction wave. Notice that by .
2.3. Viscous shock wave
We turn to the 2-viscous shock wave connecting and such that . Recall the Rankine-Hugoniot condition (1.8) and the Lax entropy condition
(2.11) |
The Riemann problem (1.3)-(1.4) admits a unique 2-shock solution
(2.12) |
By (1.8), it holds that
(2.13) |
By introducing a new variable , the the 2-viscous shock wave satisfies the ODE
(2.14) |
The properties of the 2-viscous shock wave can be listed as follows. The proof of this lemma can be found in [25] or [7] (see also [15]).
Lemma 2.2.
For any state , there exists a constant such that the following is true. For any end state such that , there exists a unique solution to (2.14). Let denote the strength of the shock as . It holds that
and
2.4. Composite waves of viscous shock and rarefaction
3. Set-up of the problem, and proof of Theorem 1.1
3.1. Construction of approximate rarefaction wave
As in [26], we will consider a smooth approximate solution of the 1-rarefaction wave, by using the smooth solution to the Burgers equation:
(3.1) |
Then, by the characteristic methods, the solution of the problem (3.1) has the following properties and their proofs can be found in [26].
Lemma 3.1.
Suppose and set . Then the problem (3.1) has a unique smooth global solution such that
(1) for and .
(2) The following estimates hold for all and :
(3) If then it holds that ,
(4) It holds that ,
(5) .
We now construct the smooth approximate 1-rarefaction wave of the 1-rarefaction wave fan by
(3.2) |
where is the smooth solution to the Burgers equation in (3.1). One can easily check that the above approximate rarefaction wave satisfies the system:
(3.3) |
Lemma 3.2.
The smooth approximate 1-rarefaction wave defined in (LABEL:AR) satisfies the following properties. Let denote the rarefaction wave strength as .
(1) and for all and .
(2) The following estimates hold for all and :
(3) For it holds that
(4) For and it holds that
(5) .
3.2. Local in time estimates on the solution
For simplification of our analysis, we rewrite the compressible Navier-Stokes system (1.1) into the following system, based on the change of variable associated to the speed of propagation of the shock :
(3.4) |
We will consider stability of the solution to (3.4) around the superposition wave of the approximate rarefaction wave and the viscous shock wave shifted by (to be defined in (3.8)) :
(3.5) |
For any initial perturbation of the superposition of waves (3.5), there exists a global strong solution to (3.4) (see for instance [29]). We will use a standard method of continuation to show the global in time control of this perturbation. For that, we first recall local in time estimates for strong solutions to (1.1) (and so also for (3.4)). They can be found in [35] (see also [29, Proposition 2.2]).
Proposition 3.1.
Let and be smooth monotone functions such that
(3.6) |
For any constants with and , there exists a constant such that if
then (3.4) has a unique solution on such that
and
Moreover:
(3.7) |
3.3. Construction of shift
For the continuation argument, the main tool is the a priori estimates of Proposition 3.2. These estimates depend on the shift function, and for this reason, we are giving its definition right now. The definition depends on the weight function defined in (4.11). For now, we will only use the fact that . We then define the shift as a solution to the ODE:
(3.8) |
where is the specific constant chosen as with , which will be used in the proof of Lemma 4.5 (see (4.49)).
The following lemma ensures that (3.8) has a unique absolutely continuous solution defined on any interval in time for which (3.7) is verified.
Lemma 3.3.
For any , there exists a constant such that the following is true. For any , and any function verifying
(3.9) |
the ODE (3.8) has a unique absolutely continuous solution on . Moreover,
(3.10) |
Proof.
We will use the following lemma as a simple adaptation of the well-known Cauchy-Lipschitz theorem.
Lemma 3.4.
[3, Lemma A.1] Let and . Suppose that a function satisfies
for some functions and . Then for any , there exists a unique absolutely continuous function satisfying
(3.11) |
3.4. A priori estimates
First, it follows from (3.3) that verifies
(3.13) |
Therefore, using (2.14) and (3.13) we find that the approximated combination of waves defined in (3.5) solves the system:
(3.14) |
where , and
(3.15) |
Note that the shift is performed only in the shock layer. The terms and are the wave interactions due to nonlinearity of the viscosity and the pressure and error terms due to the inviscid rarefaction.
We now state the key step for the proof of Theorem 1.1.
Proposition 3.2.
For a given point , there exist positive constants such that the following holds.
Suppose that is the solution to (3.4) on for some , and is defined in (3.5) with being the absolutely continuous solution to (3.8) with weight function defined in (4.11). Assume that both the rarefaction and shock waves strength satisfy and that
and
(3.16) |
Then, for all ,
(3.17) |
where is independent of and
(3.18) |
In addition, by (3.8),
(3.19) |
3.5. Global in time estimates on the perturbations
We first prove (LABEL:ext-main) from Theorem 1.1 by using Proposition 3.1 and Proposition 3.2 and a continuation argument.
Let us consider the positive constants of Proposition 3.2. The constant control the maximum size of the shock and the rarefaction, and can be chosen even smaller if needed. First, by (3.6) in Proposition 3.1, the smooth and monotone functions especially satisfy for some ,
(3.20) |
This together with Lemmas 2.2 and 3.2 then implies that for some ,
(3.21) |
By smallness of , we observe that for any ,
(3.22) |
Let be the above positive constant:
where note that can be chosen independently on , for example, as .
The specific constants will be used to apply Propositions 3.1 and 3.2 as below.
Consider any initial data verifying the hypothesis (1.11) of Theorem 1.1, that is,
(3.23) |
which together with (LABEL:underini) yields
(3.24) |
Especially, this together with Sobolev embedding implies that
(3.25) |
which together with smallness of implies that
Since satisfies by (3.22), Proposition 3.1 with (LABEL:inie) and (3.25) implies that there exists such that (3.4) has a unique solution on satisfying
(3.26) |
and
Then, using the same argument as in (LABEL:fffest), and then using Lemmas 3.2 and 3.3, we find that for all ,
Indeed, some estimates above are computed as follows:
Using smallness of , and choosing small enough such that , we have
(3.27) |
Therefore, (3.26) and (3.27) imply that
Especially, since is absolutely continuous, and
we have
We now consider the maximal existence time:
If , then the continuation argument implies that
(3.28) |
But, since it follows from (LABEL:fffest) and (LABEL:inie) that
it holds from Proposition 3.2 that
which contradicts the above equality (3.28).
Therefore, , which together with Proposition 3.2 implies
(3.29) |
and
(3.30) |
In addition, since the rarefaction wave is Lipschitz continuous in for all and from Lemma 3.2, we have
Since by (LABEL:fcru), and by Lemma 3.2, we have
which implies the desired result (LABEL:ext-main).
Especially, since the right-hand side of (LABEL:fcru) is small enough, we find that (by Sobolev embedding as before)
(3.31) |
These and the above estimates (LABEL:fcru)-(3.30) are useful to prove the long-time behaviors (1.13)-(1.14) as follows.
3.6. Time-asymptotic behavior, and end of the proof of Theorem 1.1
We now want to prove (1.13) and (1.14). Consider a function defined on by
The aim is to show the classical estimate:
(3.32) |
Since
the uniform bound (3.31) yields
(3.33) |
Thus, it follows from (LABEL:fcru), (3.33) and that
which proves the first part of (3.32).
To show the second part of (3.32), we combine the systems (3.4) and (3.14) as follows:
(3.34) |
Using (LABEL:combt) and the integration by parts, we have
(3.35) |
For the last three terms above, we get further estimates as follows.
Using (3.31) with Lemmas 2.2 and 3.2, one has
Then, using (LABEL:maingood), we have
Using the interpolation inequality and (LABEL:fcru), the last term above is estimated as
Similarly, using Lemmas 2.2 and 3.2 with recalling , we have
and
Notice that the right-hand sides above are finite by Lemma 3.2 and Lemma 4.2.
Thus, the above estimates with (LABEL:fcru) imply the proof of the second part of (3.32).
Therefore, we have (3.32), which implies
This together with the interpolation inequality and (LABEL:fcru) implies
(3.36) |
which together with Lemma 3.2 (5) implies (1.13). In addition, by (3.30) and (3.36), it holds that
(3.37) |
which proves (1.14). Thus we complete the proof of Theorem 1.1.
Hence, the remaining part of this paper is dedicated to the proof of Proposition 3.2.
Notations: In what follows, we use the following notations for notational simplicity.
1. denotes a positive -constant which may change from line to line, but which is independent of the small constants , (to appear in (4.11)) and the time .
2. For any function and any time-dependent shift ,
3. We omit the dependence on for (3.5) as follows:
For simplicity, we also omit the arguments of the waves without confusion: for example,
4. Energy estimates for the system of -variables
We introduce a new effective velocity
(4.1) |
Then, the system (3.4) is transformed into
(4.2) |
We set . Then, it follows from (2.14) that
(4.3) |
Set
(4.4) |
Then, it holds from (3.13) and (4.3) that
(4.5) |
where is defined in (3.15)
(4.6) |
This section is dedicated to the proof of the following lemma.
Lemma 4.1.
Under the hypotheses of Proposition 3.2, there exists (independent of ) such that for all ,
(4.7) |
where , and
(4.8) |
4.1. Wave interaction estimates
We here present useful estimates for the error terms introduced in (3.15) and (4.6). First, we notice that the a priori assumption (3.16) with the Sobolev embedding and (2.3) implies
(4.9) |
This smallness together with (3.8), (3.16) and (2.3) yields that
(4.10) |
This especially proves (3.19), and will be used to get the wave interaction estimates in Lemma 4.2.
Lemma 4.2.
Proof.
First, by (4.10) with (4.9), it holds that
which together with yields
Let us take so small such that the above bound is less than , that is,
Then, since
it holds from Lemma 2.2 that
Likewise, by Lemma 2.2,
On the other hand, since
it holds from Lemma 3.2 that
where note that is -constant, since .
Therefore, using the above estimates together with the bounds: (by Lemmas 2.2 and 3.2)
we have
and
Hence, this with the smallness of implies that
and
∎
4.2. Construction of weight function
We define the weight function by
(4.11) |
where the constant is chosen to be so small but far bigger than such that
(4.12) |
Notice that
(4.13) |
and
(4.14) |
and so,
(4.15) |
4.3. Relative entropy method
We rewrite (4.2) into the viscous hyperbolic system of conservation laws:
(4.16) |
where
Consider the entropy of (4.16), where , i.e., .
To write the above viscous term in terms of the derivative of the entropy:
(4.17) |
we observe that
especially, by ,
Thus, using the nonnegative matrix
the above system (4.16) can be rewritten as
(4.18) |
Let
(4.19) |
Note that (4.5) can be written as
(4.20) |
where are defined in (3.15), (4.6) respectively. Consider the relative entropy functional defined by
(4.21) |
and the relative flux defined by
(4.22) |
Let be the flux of the relative entropy defined by
(4.23) |
where is the entropy flux of , i.e., .
By a straightforward computation, for the system (4.16), we have
(4.24) |
where the relative pressure is defined as
(4.25) |
Below, we will estimate the relative entropy (weighted by defined in (4.11)) of the solution of (4.18) w.r.t. the shifted wave (4.19) as follows:
Lemma 4.3.
Remark 4.1.
Since and by Lemma 3.2, consists of good terms, while consists of bad terms.
Proof.
By the definition of the relative entropy with (4.18) and (4.21), we first have
Using the definitions (4.22) and (4.23) with the same computation as in [37, Lemma 4]) (see also [15, Lemma 2.3]), we have
(4.28) |
Using (LABEL:relative_e) and (4.17), we have
By integration by parts, we have
Using (4.17) and
(4.29) |
we have
and
Especially, since
we use (4.4) and (4.3) to have
Therefore, we have
∎
4.4. Maximization in terms of
On the right-hand side of (4.26), we will use Lemma 1.1 for the diffusion term in order to control the bad terms only related to the perturbation (or ). Therefore, we will rewrite into the maximized representation in terms of in the following lemma.
Lemma 4.4.
Remark 4.2.
Since and , consists of four good terms.
Proof.
Let and be the first terms of and respectively:
Applying the quadratic identity with to the integrands of , we have
Therefore, we have the desired representation (4.30)-(LABEL:badgood). ∎
4.5. Proof of Lemma 4.1
First of all, using Lemma 4.3 and Lemma 4.4 together with a change of variable , we have
(4.32) |
where note from (4.19) that
For the bad terms and good terms, we use the following notations:
(4.33) |
where
and
For notational simplicity in this section, we omit the dependence of the solution on the shift, i.e., .
First, note from (3.16) with the change of variable that
(4.34) |
Since the diffusion term is related to the small perturbation of pressure, we will perform the Taylor expansion near for the leading order terms and then use Lemma 1.1 on the sharp Poincaré inequality in the following lemma.
For , we have from (LABEL:relative_e) and (4.29) that
We decompose the functional as follows:
where
Notice from (3.8) that
(4.35) |
and so,
(4.36) |
4.5.1. Leading order estimates
Lemma 4.5.
There exists such that
Proof.
We first rewrite the main terms in terms of the new variables and :
(4.37) |
and
(4.38) |
Note that
(4.39) |
and the change of variable will be used below.
Note also that and so .
To perform the sharp estimates, we will consider the -constants:
which are indeed independent of the small constants , since .
Note that
(4.40) |
with together with implies
(4.41) |
Estimate on :
First, to estimate the term , we will estimate due to (4.35).
By the change of variable, we have
Using (4.40) and , we have
(4.42) |
For
we observe that since (by considering )
it follows from (4.41) and (4.34) that
This implies
(4.43) |
Therefore, by (4.35), (4.42) and (4.43), we have
which yields
This and the algebraic inequality for all imply
Thus,
(4.44) |
Change of variable for : By the change of variable, we have
which together with (4.40) yields
(4.45) |
For , using and the change of variable, we have
Using (2.4) with (4.34), we have
(4.46) |
which together with (4.40)-(4.41) yields
Change of variable for : For , we first use (2.5) with (4.34) to split it into two parts:
(4.47) |
We only do the change of variable for the good term as follows: by (4.40)-(4.41) and the change of variable,
This and (4.45) yield
(4.48) |
Change of variable for : First, using and the change of variable, we have
Integrating (4.3) over yields
Since
we have
which together with yields
Since and ,
Since the right-hand side above is the same as the one in the proofs of [14, Appendix B] and [15, Lemma 3.1]), we have
In addition, since (4.34) yields and
we have
Since
we have
Conclusion : First, by (4.46), (4.48) and the above estimates, we have
which together with the smallness of yields
Using Lemma 1.1 and the fact that for ,
we have
Since the specific -constant satisfies
(4.49) |
it holds from (4.44) and (4.47) that
which implies the desired estimate. ∎
4.5.2. Proof of Lemma 4.1
First of all, we use (4.32), (LABEL:sbg), (4.36) to have
Using Lemma 4.5 and the Young’s inequality, we find that there exist such that
In what follows, to control the above bad terms, we will use the above good terms and
(4.50) |
Note that from (LABEL:good1) and (4.50), it is obvious that with the change of variables .
Estimate on the cubic term : For simplicity, we use the notation as in (4.37). We first use (4.15) and the interpolation inequality to have
Using (4.12), (3.16) with (2.3), we have
Estimate on the term : Likewise, using (4.15) and the interpolation inequality,
Using (3.16), Lemma 4.2, (4.12) and Young’s inequality, it holds that
For , we first estimate in terms of and (using the definition of in (4.1) and in (4.4)) as follows. Observe that
(4.51) |
which together with the wave interaction estimates in Lemma 4.2 and Lemma 3.2 implies
Then, by using (3.16),
(4.52) |
This together with (3.16) and yields
and so
Estimates on the terms : Using the Young’s inequality, we have
For , we use the facts that
and
Then,
In addition, using Young’s inequality and by Lemma 3.2, we have
Estimates on the terms : We first compute that (using , , and Lemmas (3.2)-(2.2))
(4.53) |
and
(4.54) |
Then,
Using the interpolation inequality and (3.16) with Young’s inequality,
(4.55) |
For , using (2.3), (4.52) and (3.16),
Thus,
Conclusion : From the above estimates, we have
Integrating the above inequality over for any , we have
Notice that by Lemma 3.2,
and
Thus,
(4.56) |
In addition, since it follows from Lemma 4.2 that
(4.57) |
and so,
(4.58) |
we have
This implies the desired estimate (LABEL:esthv) together with the new notations (LABEL:good1), where note that
5. Proof of Proposition 3.2
In this section, we use the original system (3.4) to estimate , and then we complete the proof of Proposition 3.2.
5.1. Estimates for
We first present the zeroth-order energy estimates for the system (3.4).
Lemma 5.1.
Under the hypotheses of Proposition 3.2, there exists (independent of ) such that for all ,
(5.1) |
where are as in (LABEL:good1), and
(5.2) |
Proof.
First of all, as in Section 4.3, we first rewrite (3.4) into the form:
(5.3) |
where
and note that by the entropy of (3.4),
By the above representation, the system (3.14) can be written as
(5.4) |
where are as in (3.15).
Then, applying the equality (LABEL:genrhs) with to the system (5.3), we have
It follows from the above system that
In addition, since , we have .
Since (2.2) and (2.4) yields
we have
To control , we will use the follows estimate: as done in (4.51),
In addition, using the fact that
and so,
we have
Then, using Lemma 4.2 to have
Thus,
For , note first that by . Using Lemma 2.1,
We will use the good terms and to control .
Using and Young’s inequality, we have
For , using (4.54) and (as done in (LABEL:ccomp))
(5.5) |
we have
Using the same estimates as in (4.55) with (3.16), we have
Using (3.16), we have
Therefore, from the above estimates, we find that for some constant ,
Integrating the above inequality over for any , and using (4.56)-(4.58), we have
(5.6) |
Therefore, multiplying (LABEL:f00) by the constant , and then adding the result to (LABEL:esthv), together with the smallness of , we have
(5.7) | ||||
where we have used that (by Lemma 2.1 and (4.34))
Finally, to complete the proof, we will show that
(5.8) |
and
(5.9) |
Using the definition of in (4.1) and in (4.4), we observe that
which yields
This with Lemma 3.2 and Lemma 4.2 implies (5.8).
As in (4.51), we have
which together with Lemmas 2.2 and 3.2 implies (5.9).
Hence, the combination of (5.1),(5.8) and (5.9) implies the desired estimate.
∎
5.2. Estimates for
We here complete the proof of Proposition 3.2, by using the following lemma together with the following two estimates (by using Lemma 2.1) :
Lemma 5.2.
Under the hypotheses of Proposition 3.2, there exist (independent of ) such that for all ,
where are as in (LABEL:good1), and are as in (LABEL:good2), and
Proof.
For notational simplicity, we set . Then, it follows from the second equations of (3.4) and (3.14) that
Multiplying the above equation by and integrating the result w.r.t. , we have
First, we get a good term
from as follows:
We use the good terms and to control the remaining terms as follows.
Using Young’s inequality,
Using , and the interpolation inequality and (3.16), we have
Using (by Lemma 3.2),
Using (LABEL:f_3),
Therefore, we find that for some ,
Integrating the above estimate over for any , and using (4.57) and the fact that (by Lemma 3.2)
we have
Multiplying the above inequality by the constant , and then adding the result to (LABEL:estu0), together with the smallness of , we have
This implies the desired result in Lemma 5.2. ∎
Conflict of Interest: The authors declared that they have no conflicts of interest to this work.
References
- [1] S. Bianchini and A. Bressan. Vanishing viscosity solutions to nolinear hyperbolic systems. Ann. of Math., 166:223–342, 2005.
- [2] D. Bresch and B. Desjardins. On the construction of approximate solutions for the 2D viscous shallow water model and for compressible Navier-Stokes models. J. Math. Pures Appl., 86(9):362–368, 2006.
- [3] K. Choi, M.-J. Kang, Y. Kwon, and A. Vasseur. Contraction for large perturbations of traveling waves in a hyperbolic-parabolic system arising from a chemotaxis model. Math. Models Methods Appl. Sci., 30(2):387–437, 2020.
- [4] C. M. Dafermos. The second law of thermodynamics and stability. Arch. Rational Mech. Anal., 70(2):167–179, 1979.
- [5] 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.
- [6] R. J. DiPerna. Uniqueness of solutions to hyperbolic conservation laws. Indiana Univ. Math. J., 28(1):137–188, 1979.
- [7] J. Goodman. Nonlinear asymptotic stability of viscous shock profiles for conservation laws. Arch. Rational Mech. Anal., 95(4):325–344, 1986.
- [8] 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.
- [9] A. M. Ilin and O. A. Oleĭnik. 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.
- [10] M.-J. Kang. Non-contraction of intermediate admissible discontinuities for 3-D planar isentropic magnetohydrodynamics. Kinet. Relat. Models, 11(1):107–118, 2018.
- [11] M.-J. Kang. -type contraction for shocks of scalar viscous conservation laws with strictly convex flux. J. Math. Pures Appl. (9), 145:1–43, 2021.
- [12] 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.
- [13] M.-J. Kang and A. Vasseur. -contraction for shock waves of scalar viscous conservation laws. Annales de l’Institut Henri Poincaré (C) : Analyse non linéaire, 34(1):139Ð156, 2017.
- [14] M.-J. Kang and A. Vasseur. Well-posedness of the Riemann problem with two shocks for the isentropic euler system in a class of vanishing physical viscosity limits. arXiv:2010.16090, 2020.
- [15] 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.
- [16] M.-J. Kang and A. Vasseur. Uniqueness and stability of entropy shocks to the isentropic Euler system in a class of inviscid limits from a large family of Navier-Stokes systems. Invent. Math., 224(1):55–146, 2021.
- [17] M.-J. Kang, A. Vasseur, and Y. Wang. -contraction for planar shock waves of multi-dimensional scalar viscous conservation laws. J. Differential Equations, 267(5):2737–2791, 2019.
- [18] N. Leger. stability estimates for shock solutions of scalar conservation laws using the relative entropy method. Arch. Ration. Mech. Anal., 199(3):761–778, 2011.
- [19] N. Leger and A. Vasseur. Relative entropy and the stability of shocks and contact discontinuities for systems of conservation laws with non-BV perturbations. Arch. Ration. Mech. Anal., 201(1):271–302, 2011.
- [20] T.-P. Liu. Nonlinear stability of shock waves for viscous conservation laws. Mem. Amer. Math. Soc., 56(328):v+108, 1985.
- [21] 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.
- [22] T.-P. Liu and Y. Zeng. Shock waves in conservation laws with physical viscosity. Mem. Amer. Math. Soc., 234(1105):vi+168, 2015.
- [23] C. Mascia and K. Zumbrun. Stability of small-amplitude shock profiles of symmetric hyperbolic-parabolic systems. Comm. Pure Appl. Math., 57:841–876, 2004.
- [24] 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.
- [25] A. Matsumura and K. Nishihara. On the stability of traveling wave solutions of a one-dimensional model system for compressible viscous gas. Japan J. Appl. Math., 2:17–25, 1985.
- [26] 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., 3:1–13, 1986.
- [27] A. Matsumura and K. Nishihara. Global stability of the rarefaction wave of a one-dimensional model system for compressible viscous gas. Comm. Math. Phys., 144:325–335, 1992.
- [28] A. Matsumura and Yang Wang. Asymptotic stability of viscous shock wave for a one-dimensional isentropic model of viscous gas with density dependent viscosity. Methods Appl. Anal., 17(3):279–290, 2010.
- [29] A. Mellet and A. Vasseur. Existence and uniqueness of global strong solutions for one-dimensional compressible Navier-Stokes equations. SIAM J. Math. Anal., 39(4):1344–1365, 2007/08.
- [30] 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.
- [31] B. Riemann. Uberdie fortpflanzung ebener luftwellen von endlicher schwingungsweite. Abhandlungen der Koniglichen Gesellschaft der Wissenschaften zu Gottingen, 8:43–65, 1860.
- [32] D. H. Sattinger. On the stability of waves of nonlinear parabolic systems. Advances in Math., 22(3):312–355, 1976.
- [33] D. Serre and A. Vasseur. -type contraction for systems of conservation laws. J. Éc. polytech. Math., 1:1–28, 2014.
- [34] D. Serre and A. Vasseur. About the relative entropy method for hyperbolic systems of conservation laws. Contemp. Math. AMS, 658:237–248, 2016.
- [35] V. A. Solonnikov. The solvability of the initial-boundary value problem for the equations of motion of a viscous compressible fluid. Zap. Naun. Sem. Leningrad. Otdel. Mat. Inst. Steklov., 59:128–142, 1976.
- [36] A. Szepessy and Z. P. Xin. Nonlinear stability of viscous shock waves. Arch. Rational Mech. Anal., 122(1):53–103, 1993.
- [37] A. Vasseur. Recent results on hydrodynamic limits. Handb. Differ. Equ. Elsevier/North-Holland, Amsterdam, 2008.
- [38] A. Vasseur. Relative entropy and contraction for extremal shocks of conservation laws up to a shift. In Recent advances in partial differential equations and applications, volume 666 of Contemp. Math., pages 385–404. Amer. Math. Soc., Providence, RI, 2016.
- [39] A. Vasseur and L. Yao. Nonlinear stability of viscous shock wave to one-dimensional compressible isentropic Navier-Stokes equations with density dependent viscous coefficient. Commun. Math. Sci., 14(8):2215–2228, 2016.