Small scale formations in the incompressible porous media equation
Abstract.
We construct examples of solutions to the incompressible porous media (IPM) equation that must exhibit infinite in time growth of derivatives provided they remain smooth. As an application, this allows us to obtain nonlinear instability for a class of stratified steady states of IPM.
Key words and phrases:
IPM equation, two-dimensional incompressible flow, small scale creation, derivatives growth, nonlinear instability2010 Mathematics Subject Classification:
35Q35,76B031. Introduction
In this paper, we consider the 2D incompressible porous media (IPM) equation. The equation describes evolution of density carried by the flow of incompressible fluid that is determined via Darcy’s law in the field of gravity:
(1.1) |
Here is the transported density, is the vector field describing the fluid motion, and is the pressure. Throughout this paper, we consider the spatial domain to be one of the following: the whole space , the torus , or the bounded strip that is periodic in . In the last case, due to the presence of boundaries, also satisfies for . In all the three cases, one can obtain a more explicit Biot-Savart law for :
Here , and the inverse Laplacian for and will be specified in Section 2.
There have been many recent papers analyzing the well-posedness questions for the IPM equation and its variants [1, 4, 5, 8, 19], lack of uniqueness of weak solutions [3, 18], and questions of long time dynamics [7, 1]. Viewed as an active scalar, the IPM equation is less regular than the 2D Euler equation in vorticity form, and has the same level of regularity as the SQG equation. Local well-posedness for sufficiently regular initial data has been proved in [5] for , and [1] for the strip . The argument can be adapted to the periodic setting ; we will sketch a simple proof in Section 2.2. The question of global regularity vs finite-time blow up is open for the IPM equation, similarly to the SQG equation case. Moreover, to the best of our knowledge, there are not even examples of smooth solutions to the IPM equation that have infinite growth of derivatives. There are plenty of such examples for the 2D Euler equation, going back to work of Yudovich [12, 20] (see e.g. [6], [13] for more recent examples and further references). However, for the more singular SQG equation case, such examples have been established only recently [10]. The reason for such delay is that an example of infinite in time creation of small scales requires sufficiently strong control of the solution, which is not easily achieved when the drift is more singular. The example of [10] is based on the insight gained in the constructions for the 2D Euler case [13, 21], and is based on a hyperbolic point scenario controlled by odd-odd symmetry of the active scalar. It is tempting to use a similar idea for the IPM equation, but its structure is different - in particular, odd symmetry in but even symmetry of in is conserved instead of the odd-odd symmetry for the SQG equation. This, and the more detailed structure of the Biot-Savart law, appear to be significant obstacles in extending ideas of [10, 13, 21] to the IPM equation setting. In this paper, we construct examples of infinite growth of derivatives in smooth solutions of the IPM equation using a different idea, exploiting existence of monotone quantity which corresponds to the potential energy of the fluid. All our estimates below assume that the solutions remain smooth; more specifically, the arguments work if and are at least Lipschitz. If this regularity fails in finite time, we already have an even more dramatic effect than what we are trying to establish.
1.1. Small scale formation in IPM
In this paper, we consider the following three scenarios:
(S1) Let . Assume is odd in , and in .
(S2) Let be the 2D torus. Assume is odd in .
(S3) Let be a bounded strip that is periodic in . Assume .
Our first result shows that in the scenario (S1), must have infinite-in-time growth in norm for any , if it remains regular for all times. Note that is the sharp threshold, since for we know does not grow in time.
Theorem 1.1.
For , let satisfy the scenario (S1). Assuming that there is a global-in-time smooth solution to (1.1) with initial data , we have
(1.2) |
which implies
(1.3) |
The next result concerns the torus scenario (S2), where we prove infinite-in-time growth under some additional symmetry and positivity assumptions on . As we will see in the remark afterwards, the same result also holds for the bounded strip scenario (S3).
Theorem 1.2.
For , let satisfy the scenario (S2). In addition, assume that is even in , for , and in . Assuming that there is a global-in-time smooth solution to (1.1) with initial data , we have
(1.4) |
which implies
(1.5) |
Remark 1.3.
Observe that the solution in from Theorem 1.2 is automatically a solution in the bounded strip , with satisfying the no-flow condition on the top and bottom boundaries. (This is because is odd in and has period in for all times. Thus is also odd about , implying for all times on .) Therefore, the growth results of Theorem 1.2 directly hold in scenario (S3). We note that the local well-posedness for the scenario (S3) has been established in [1], which in particular ensures the uniqueness of solution while it remains regular.
1.2. Nonlinear instability in IPM
One can easily check that any horizontal stratified state is a stationary solution of (1.1) in , or , since . (As we will see in Lemma 3.1, all smooth stationary solutions in are of the form . However, in and there are other smooth stationary solutions, e.g. any vertical stratified state is also stationary; see also [2, Section 5] for smooth stationary solutions in supported in an infinite slanted strip.)
Below we briefly summarize the previous stability results for the horizontal stratified state . Denoting and plugging it into (1.1), satisfies
(1.6) |
with . For small, the linearized equation is , which can be written as
Since is a negative operator, one would expect the equation to be linearly stable if is uniformly negative (i.e. lighter density on top, heavier on the bottom).
For the stratified state , the asymptotic stability of the nonlinear equation (1.6) has been rigorously established by Elgindi [7] in and Castro–Córdoba–Lear [1] in , which also implies the global well-posedness of (1.1) for initial data close to in certain Sobolev spaces. More precisely, for in , if for , [7, Theorem 1.3] proved that remains regular for all time and satisfies for all . [7] also obtained asymptotic stability results for periodic perturbation , where is still in the whole plane. In [1], the authors proved that for the stratified state in is asymptotically stable in for , although it may converge to a slightly different stratified state from as .
In this paper, we aim to prove two nonlinear instability results for the horizontal stratified steady state in and respectively. What sets our approach apart is that we are not following the common path of converting linear instability into a nonlinear one. Rather, we use the monotone quantity - potential energy - to prove infinite-in-time growth of Sobolev norms and then leverage these results to conclude the nonlinear instability. Our first instability result shows that in , any horizontal stratified steady state that is odd in is nonlinearly unstable, and the instability can grow “infinitely in time”. Namely, for any arbitrarily large , one can construct an initial data that is arbitrarily close to in , such that for all .
Theorem 1.4.
Let be any horizontal stratified state (i.e. ) that is odd in . For any and any , there exists an initial data satisfying
such that the solution to (1.1) with initial data (provided it remains smooth for all times) satisfies
(1.7) |
Finally, we prove an instability result in the bounded strip for any stratified steady state , including those monotone stratified states that are linearly stable such as . Namely, we can construct a smooth perturbation small in norm for any , such that for any .
Theorem 1.5.
Let be any stationary solution. For any , there exists an initial data satisfying
(1.8) |
such that the solution to (1.1) with initial data (provided it remains smooth for all times) satisfies
(1.9) |
Remark 1.6.
It is a natural question whether the perturbation can be made arbitrarily small in higher Sobolev spaces. While it is unclear to us whether is the sharp threshold, we know that the exponent cannot exceed 10: for , if the initial perturbation is small in or above, [1] showed remains uniformly bounded in time.
1.3. Organization of the paper
In Section 2 we discuss some preliminaries and the local well-posedness results in the scenarios (S1)–(S3). In Section 3 we show the monotonicity of the potential energy in the three scenarios, and use it to prove the infinite-in-time growth results in Theorems 1.1–1.2. We take a brief detour in Section 4 to derive some infinite-in-time growth results for less restrictive initial data, which we call the “bubble” solution and the “layered” solution. This will enable us to obtain nonlinear instability results in Section 5 for initial data close to stratified steady states, where we prove Theorems 1.4–1.5.
2. Preliminaries on problem setting and local well-posedness
In this section, we discuss some preliminaries such as the Sobolev spaces for the spatial domains , and respectively, as well as the local-wellposedness results for the IPM equation (1.1). For the whole space and strip case, the local-wellposedness theory have already been established in [4] and [1] respectively. For the torus case we are unable to locate a local-wellposedness result, so we give a short proof in Section 2.2.
2.1. Sobolev norms and local well-posedness in
For any , its Fourier transform is defined as usual as
and the Plancherel theorem yields As usual, we define
and
For (1.1) in , Córdoba–Gancedo–Orive [5, Theorem 3.2] proved local-wellposedness for initial data with . They also established a regularity criteria, showing that remains regular as long as . They also obtained another regularity criteria with a geometric flavor, and we refer the reader to [5, Theorem 3.4] for details.
2.2. Sobolev norms and local well-posedness in
For any , let us denote its Fourier series as
(2.1) |
where the Fourier coefficient for is given by
(2.2) |
By Parseval’s theorem, for any we have which in particular implies
(2.3) |
For , throughout this paper, is defined by
(2.4) |
Finally, for mean-zero (in particular this is the case for ), its inverse Laplacian is given by .
Below we sketch a-priori estimates that can be used to establish local regularity as well as conditional criteria for global regularity. With these estimates, a fully rigorous argument can be given in a standard way, using either smooth mollifier approximations like in [14] or Galerkin approximations.
Suppose that is a smooth solution of where Observe that all the norms of are conserved by evolution. Multiplying the equation by and integrating we obtain
Here is an integer. In the integral above, we can expand the power of the Laplacian and integrate by parts, then use the periodicity to transfer exactly half of the derivatives on . What we get is a sum of terms of the form
where stands for some partial derivative of order Next we apply Leibniz rule to open up the derivative falling on Note that when all derivatives fall on we get
due to incompressibility of Therefore we obtain a sum of the terms of the form
where Let us apply Hölder inequality to control such integral by
where satisfy Let us recall a particular case of Gagliardo-Nirenberg inequality [9, 15]
(2.5) |
for any where in the 2D case The inequality is valid for While (2.5) is usually stated in an extension to is straightforward. Taking now and and applying (2.5), we get
where in the first step we used bound on singular integral operators for Similarly,
Therefore, for all we have
and hence
(2.6) |
Such inequality can be used to show local well-posedness in provided that
To obtain a criteria for blow up, we can run a similar calculation but using and instead of and with This way instead of (2.6) we obtain the differential inequality
We can conclude control of up to any time provided that remains finite. Let us state a proposition summarizing the observations of this section.
Proposition 2.1.
Consider the IPM equation (1.1) with the initial data an integer. Then there exists a time such that for all there exists a unique solution Moreover, the solution blows up at time if and only if
when
Remarks. 1. Uniqueness of the solution can be shown in a standard way; blow up is understood in the sense of leaving the class
2. The Proposition can certainly be improved in terms of the condition on and the regularity criterion, but we do not pursue it in this paper.
2.3. Sobolev norms and local well-posedness in a strip
When the domain is a bounded strip , due to the presence of the top and bottom boundaries, the functional spaces and the local-wellposedness results are more involved than the periodic case. Below we briefly describe the results by Castro–Córdoba–Lear [1], and we refer the readers to the paper for more details.
Biot-Savart law and functional space. In the strip case, the velocity field is given by , where the stream function solves the Poisson’s equation with zero boundary condition (see [1, Section 2.2] for a derivation):
(2.7) |
so that . One can check that the operator (with zero boundary condition) is a positive self-adjoint operator, and it has a family of eigenfunctions that form an orthonormal basis for , given by
where
and
The eigenfunction expansion allows us to define for any and . We can then define the homogenous Sobolev norm as
(2.8) |
and one can check that the spaces and are dual with respect to the norm (see e.g. [17] for the general construction of a scale of Sobolev spaces associated with a positive self-adjoint operator).
For , let us define as usual. For , the above definition of is comparable to if for all even with , where is any partial derivative of order .
Local/global well-posedness results in the strip. Since the goal of [1] was to establish stability results near the steady state , (1.1) was written into an equivalent equation (1.6) (with ) describing the evolution of . Here can be expressed in terms of similarly to (2.7), except that the right hand side is replaced by (using that has zero contribution to since it is a steady state).
When the initial data belongs to the functional space , given by
(which in fact coincides with defined above), the authors proved local-wellposedness of (1.6) for for any [1, Theorem 4.1], and gave a regularity criteria showing that remains in as long as . As we discussed in the introduction, for , they proved the asymptotic stability of (which implies global regularity) for with .
3. Infinite-in-time growth in the IPM
In this section we aim to prove Theorems 1.1 and 1.2. Throughout this paper, the evolution of the potential energy
plays a key role. Let us first prove a simple lemma showing that in each of the scenarios (S1)–(S3), is monotone decreasing in time, and its time derivative is integrable in .
Lemma 3.1.
Assume that and satisfy one of the scenarios (S1)–(S3). Assuming that there is a global-in-time smooth solution to (1.1) with initial data , we have
In addition, we have .
Proof.
A direct computation gives
(3.1) |
Here the last inequality is due to the divergence theorem, where the boundary integral vanishes in all the three scenarios (S1)–(S3): In (S1), the boundary integral (at infinity) vanishes since for all time. In (S2), the boundary integral due to periodicity, and since on (which follows from the facts that is odd in , and periodic in ). In (S3), again due to periodicity in , whereas due to on .
By (3.1) and the Biot-Savart law , in (S1)–(S3), we get
(3.2) |
thus is monotone decreasing. Note that in the strip case , the last identity follows from the definition of the norm in as in (2.8).
Moreover, is uniformly bounded below for all times. In (S1), the assumptions that is odd in and in yield that in , thus for all times. In (S2) and (S3), since a smooth solution of (1.1) has its norm invariant in time, we have that
Hence in all three cases (S1)–(S3), we have
(3.3) |
finishing the proof. ∎
Remark 3.2.
When the equation is set in with decaying sufficiently fast, monotonicity of has been derived in [7, Corollary 1.2]. For the Muskat equation (which can be seen as a “patch” solution of IPM) with surface tension, [11] uses the gradient flow structure to construct weak solutions, where the energy functional is the potential energy plus the surface area of the free boundary.
Proof of Theorem 1.1.
Due to Lemma 3.1, satisfies . Denoting , we have for all , since the norm of is invariant in time for all . Let us define Note that directly implies . We claim that for any ,
(3.4) |
Once we prove (3.4), plugging it into and using the fact that , we have
Here in the last inequality we used that for any , is bounded below by a positive constant , which follows from the elementary interpolation inequality , as well as the fact that and are invariant in time. This finishes the proof of (1.2). Combining (1.2) with the fact that gives (1.3) as a direct consequence.
In the rest we aim to prove (3.4) for any fixed , and we will drop the dependence in and below for notational simplicity. Defining
we observe that
This gives , and combining it with yields . Note that consists of two symmetric cones containing the axis, and it can be expressed in polar coordinates as }.
Clearly, . Let be such that , which will be estimated momentarily. Such definition gives
immediately leading to
(3.5) |
It remains to estimate . Denoting , we know consists of two identical triangles with height and base . Thus
where in the last inequality we used and , due to . Therefore . Plugging it into (3.5) yields (3.4), finishing the proof. ∎
Proof of Theorem 1.2.
Since is even in and odd in , due to the Biot-Savart law , the even-odd symmetry of remains true for all times. In particular, it implies that on the boundary of the smaller square , we have , and combining it with on gives on for all times. In addition, note that on and the fact that imply for all and .
For any , has Fourier series (2.1)–(2.2) (with replaced by ), and the Fourier coefficient for can be written as
(3.6) |
where the last identity follows from the oddness of in .
Let us take a closer look at the function in the last line of (3.6) (where we set ). It satisfies the following properties for all :
-
(a)
for all and , and is even in .
-
(b)
for all .
-
(c)
for all , where only depends on .
Here properties (a, b) follow from the facts that is even in , nonnegative on , and for all times. For property (c), note that Hölder’s inequality and the fact that in yield that
where is a universal constant. Since is advected by a divergence-free flow with , one can easily check that for some , finishing the proof of property (c).
Let us define as the Fourier coefficient of , given by
(3.7) |
Comparing (3.7) with (3.6) directly yields
(3.8) |
Using the functions and , we can estimate from below as
(3.9) |
where is the average of in . By property (c), we have for all . Intuitively, if is small, must be very close to in . With pinned down at zero (by property (b)), and being uniformly positive, must have order 1 oscillations in a small neighborhood near , suggesting it should have a large norm for . This estimate will be made rigorous in Lemma 3.3 right after the proof. Applying Lemma 3.3 to , we have
(3.10) |
Note that
where the last inequality follows by just looking at the part of the sum for the last norm taken on Fourier side and using Setting and applying (3.10), we have
Plugging this inequality into implies (1.4), and combining (1.4) with the fact that gives (1.5) as a direct consequence. ∎
Now we state and prove the lemma used in the proof of Theorem 1.2.
Lemma 3.3.
If satisfies , and (where ), then
(3.11) |
Proof.
Note that has mean zero in , and . By the assumption , there exists some such that . This implies
Applying the Sobolev embedding theorem, we have
and setting gives (3.11) for , where we also use that . For , we can apply the interpolation inequality to obtain
thus we can conclude. ∎
4. Bubble and layer solutions
Previously, we have proved infinite time growth results in Theorem 1.2 and Remark 1.3 for scenarios (S2) and (S3), under some additional assumption on . In this section we aim to work with less restrictive initial data – in particular, the assumption that for can now be dropped. This will enable us to obtain instability results in Section 5 for initial data close to stratified steady states. However, as the proof is done by a different approach, the set of Sobolev exponents with norm growth (as well as the growth rate in time) is not as good as Theorem 1.2 and Remark 1.3.
We first consider the initial data of “bubble” type, that is, its level sets have a connected component enclosing a simply-connected region, and does not vanish on (see Figure 1 for an illustration). Intuitively, since the topological structure of all level sets is preserved under the evolution, the presence of the “bubble” prevents the solution from aligning into a perfectly stratified form where may increasingly vanish. In the next result we rigorously justify this by showing that for all times, and as we will see, this leads to infinite-in-time growth in certain Sobolev norms.
Proposition 4.1.
Let , and assume satisfies the scenario (S2). Suppose there exists a simple closed curve enclosing a simply-connected domain , and satisfies and 111Observe that by Sard’s theorem [16], since , the set of such that contains a critical point has Lebesgue measure zero. . Assuming that there is a global-in-time smooth solution to (1.1) with initial data , we have
(4.1) |
which implies
(4.2) |
Proof.
Since with , is uniformly positive in some open neighborhood of . Combining this with , for any sufficiently close to , the level set has a connected component that is a simple closed curve near . Since encloses a simply-connected region , there exists a simple closed curve such that . Denote by the region enclosed by , which is also simply-connected. See Figure 1 for an illustration of the curves and the domains .

Let us as usual define the trajectories of the flow by
(4.3) |
While the solution remains smooth, the flow map is a measure-preserving smooth mapping. Thus and remain simply-connected in , and they satisfy for all . Denoting by the Lebesgue measure of a set (which is preserved by ), we have for all . In addition, since is advected by (1.1), we have as usual that for all and , thus and for all .
Let us denote the projection map onto the variable, i.e. for any , . Using , we have
Since and are simply-connected domains enclosed by boundaries and respectively, the above becomes
(4.4) |
Using , we have for all . Finally, defining , which is a subset in , we have shown that and for all .
By definition of and (4.4), for any and , has a non-empty intersection with both and . Since and , it implies
(4.5) |
Integrating this in and using , we have for all thus Cauchy-Schwartz yields
(4.6) |
Applying the interpolation inequality for with , we have
(4.7) |
Plugging (4.7), (4.6) into (3.3) we obtain (4.1). Finally, combining (4.1) with the fact that gives (4.2) as a direct consequence. ∎
The growth for “bubble” solutions can be easily adapted to the bounded strip case as follows.
Corollary 4.2.
Let , and assume satisfies scenario (S3). Suppose there exists a simple closed curve enclosing a simply-connected domain , and satisfies and . Assuming that there is a global-in-time smooth solution to (1.1) with initial data , we have
(4.8) |
which implies
(4.9) |
Proof.
The proof is almost identical to the proof of Proposition 4.1. Again, there exists enclosing a simply-connected domain , such that , and . Defining (which is now a subset in ), the same argument gives that and (4.5). Thus (4.6) still holds (except that the integral now takes place in instead of ), and the rest of the proof remains unchanged. Note that the interpolation inequality holds in as well by a straightforward argument using the eigenfunction expansion similar to the standard Fourier argument in ∎
Our next result concerns “layered” initial data, which we define below.
Definition 4.3.
For , we say it has a layered structure if there exists a measure-preserving smooth diffeomorphism that satisfies , such that is a stratified solution, i.e. only depends on . In this case, we call the stratified state corresponding to .
Note that any layered initial data has a unique corresponding stratified state . (Even though the mapping is not unique, e.g. one can shift by any ). To see this, take any curve such that with , and denote by the region bounded between and . Since is measure-preserving and , we know must have the same area as , leading to . Since (thus is also smooth), Sard’s theorem allows us to run this argument for a.e. , which defines uniquely for all . See Figure 2 for an illustration of a layered initial data and its corresponding stratified state .

Clearly, Theorem 1.2 cannot be applied to any layered since , and Proposition 4.1 fails too since there is no level set enclosing a simply-connected region. Despite these difficulties, we will show that small scale formation can still happen to , as long as its potential energy is strictly lower than that of .
Proposition 4.4.
Remark 4.5.

Proof.
To begin with, we will show that for any , also has a layered structure with corresponding stratified state being . The assumption on gives for some measure-preserving diffeomorphism . Combining this with (where is the flow trajectory given by (4.3)), we have for all times. Here is a measure-preserving diffeomorphism, and it keeps the set invariant, since both and have this property: has this property due to Definition 4.3, whereas has this property since on for all times.
Let us denote where the strict positivity is due to (4.10). By Lemma 3.1, is non-increasing in time, thus
(4.11) |
Since is the only stratified state that is topologically reachable from (among all measure-preserving diffeomorphisms that keeps invariant), the fact that has a potential energy strictly less than (with the gap being at least ) intuitively suggests that cannot have all level sets very close to horizontal. Below we will show that this is indeed true, in the sense that is bounded below by a positive constant for all times.
By (4.11) and the definition of potential energy , we have
(4.12) |
and let us take a closer look at the integrand. In the first paragraph of the proof we showed , where is a measure-preserving diffeomorphism that keeps invariant. As a result, for any and , must have a non-empty intersection with , i.e. there exists and depending on and , such that . Combining this with the fact that is a function of only, we have
for any . Plugging this into (4.12) gives
leading to for all . Now that we have a positive lower bound on , the rest of the argument can proceed the same way as in (4.6) and (4.7) in the proof of Proposition 4.1, allowing us to obtain the same estimates (4.1)–(4.2). ∎
5. Instability of horizontally stratified steady states
In this section we aim to prove the two nonlinear instability results Theorems 1.4 and 1.5 in and respectively. We start with Theorem 1.4, which shows that any stratified steady state in that is odd in is nonlinearly unstable in an arbitrarily high Sobolev space. The idea is to locate a point where locally has heavier density on top of lighter one, then use a circular flow to slightly perturb near to construct a “layered” initial data that satisfies the assumption of Proposition 4.4.
Proof of Theorem 1.4.
We claim that for any and , we can construct a satisfying all the following:
-
(a)
is odd in , and has a layered structure in the sense of Definition 4.3 with corresponding stratified state .
-
(b)
.
-
(c)
.
Once these are shown to be true, a direct application of Proposition 4.4 immediately yields the infinite-in-time growth results (4.1) and (4.2). Since , (4.2) directly implies (1.7), finishing the proof.
In the rest of the proof we aim to construct and prove the claim. We will focus on the construction of in the upper half of torus , and at the end we will extend it to the lower part by an odd extension.
Recall that is a smooth stratified state that is odd in . Thus is odd and smooth in . Such cannot be monotone, so there exists some such that . For to be fixed later, let be non-negative and supported on . Let be the velocity field of an incompressible circular flow around , given by
(5.1) |
For any , let be the solution to
(5.2) |
with initial data . Since is supported in a small annulus , clearly outside the annulus. Intuitively, since has heavier density on top of lighter density locally near , we formally expect that should have lower potential energy than for a short time. (Here the “time” is the perturbation parameter, and has nothing to do with the actual time in (1.1)). Below we will rigorously show
(5.3) |
Since the integral can be reduced to the set , it is convenient to write it in polar coordinates centered at the point . Using the change of variables , we have
(5.4) | |||||
where the second identity follows from the facts that is transported by with initial data , and is a circular flow with angular velocity along . We can rewrite as
where the second integral is constant in using the substitution . Thus taking the derivative gives
leading to
(5.5) |
since the integrand is odd about . Taking one more derivative and setting , we have
Since is chosen such that , for all sufficiently small we have
Plugging (5.5) and the above into (5.4) gives that and . Combining these with gives (5.3).
Finally, we use odd reflection to extend to , which is equivalent with simutaneously applying a circular flow to near in the opposite direction as . We then set with sufficiently small, and let us check that it satisfies the claim (a,b,c): (a) is a direct consequence from the definition, since can be reached from by an explicit measure-preserving smooth flow that is only non-zero near . Also, since is transported from with a smooth velocity field , for any , we have as , thus property (b) is satisfied. As for the potential energy, note that (5.3) gives that when is sufficiently small, finishing the proof of (c). ∎
Finally, we are ready to prove Theorem 1.5 which deals with the instability on the strip. The idea is to perturb the steady state to make a small “bubble” localized near one point, then apply Corollary 4.2.
Proof of Theorem 1.5.
First note that any stationary solution must be stratified of the form , since only in this case it satisfies by Lemma 3.1.
Let be a nonnegative function supported in with . For , let
where . Clearly, , and in .
Let us first check that (1.8) is satisfied for with . A simple scaling argument yields that is invariant in (where is any partial derivative of order 2), thus is uniformly bounded for all . Combining this with , we have for all (where only depends on ), where the right hand side can be made arbitrarily small for .
We claim that for any , satisfies the assumption of a “bubble solution” in Corollary 4.2. To see this, note that the definitions of and yields
whereas
Applying Sard’s theorem [16] to , for almost every , we know has a connected component in on which never vanishes. Naming any such connected component , we then have that satisfies the assumption in Corollary 4.2. As a result we have the estimate (4.9). Using that , (4.9) directly implies (1.9), thus finishes the proof. ∎
Remark 5.1.
For a stratified solution that does not satisfy , the perturbation can be made small in higher Sobolev spaces. Namely, if there exists such that , one can proceed as in the proof of Proposition 4.4 to construct a “layered” initial data close to in norm for arbitrarily large .
Acknowledgement. The authors acknowledge partial support of the NSF-DMS grants 1715418, 1846745 and 2006372. AK has been partially supported by Simons Foundation. YY has been partially supported by the Sloan Research Fellowship. This paper has been initiated at the AIM Square, and the authors thank AIM for support and collaborative opportunity.
References
- [1] A. Castro, D. Córdoba and D. Lear, Global existence of quasi-stratified solutions for the confined IPM equation, Arch. Ration. Mech. Anal. 232 (2019), no. 1, 437–471
- [2] P. Constantin, J. La and V. Vicol, Remarks on a paper by Gavrilov: Grad-Shafranov equations, steady solutions of the three dimensional incompressible Euler equations with compactly supported velocities, and applications, Geom. Funct. Anal. 29 (2019) 1773–1793.
- [3] D. Córdoba, D. Faraco and F. Gancedo, Lack of uniqueness for weak solutions of the incompressible porous media equation, Arch. Ration. Mech. Anal. 200 (2011), no. 3, 725–746
- [4] D. Córdoba and F. Gancedo, Contour dynamics of incompressible 3-D fluids in a porous medium with different densities, Comm. Math. Phys. 273 (2007), no. 2, 445–471
- [5] D. Córdoba, F. Gancedo and R. Orive, Analytical behavior of two-dimensional incompressible flow in porous media, J. Math. Phys. 48 (2007), no. 6, 065206, 19 pp
- [6] S. Denisov, Infinite superlinear growth of the gradient for the two-dimensional Euler equation, Discrete Contin. Dyn. Syst. A, 23 (2009), no. 3, 755–764
- [7] T. Elgindi, On the asymptotic stability of stationary solutions of the inviscid incompressible porous medium equation, Arch. Ration. Mech. Anal. 225 (2017), no. 2, 573–599
- [8] S. Friedlander, F. Gancedo, W. Sun and V. Vicol, On a singular incompressible porous media equation, J. Math. Phys. 53 (2012), no. 11, 115602, 20 pp
- [9] E. Gagliardo, Ulteriori propietádi alcune classi di funzioni on più variabli, Ric. Math. 8 (1959), 24–51
- [10] S. He and A. Kiselev, Small scale creation for solutions of the SQG equation, preprint arXiv:1903.07485, to appear at Duke Math. J.
- [11] M. Jacobs, I. Kim and A.R. Mészáros, Weak Solutions to the Muskat Problem with Surface Tension Via Optimal Transport, Arch. Ration. Mech. Anal. 239 (2021), 389–430.
- [12] V. I. Judovic, The loss of smoothness of the solutions of Euler equations with time (Russian), Dinamika Splosn. Sredy, Vyp. 16, Nestacionarnye Problemy Gidrodinamiki, (1974), 71–78, 121
- [13] A. Kiselev and V. Sverak, Small scale creation for solutions of the incompressible two dimensional Euler equation, Annals of Math. 180 (2014), 1205–1220
- [14] A. Majda and A. Bertozzi, Vorticity and Incompressible Flow, Cambridge University Press, 2002
- [15] L. Nirenberg, On elliptic partial differential equations: Lecture II, Ann. Sc. Norm. Super. Pisa, S. 3, 13 (1959), 115–162
- [16] A. Sard, The measure of the critical values of differentiable maps, Bulletin of the American Mathematical Society, 48 (1942), 883–890
- [17] B. Simon, Spectral analysis of rank one perturbations and applications. Mathematical quantum theory. II. Schrödinger operators (Vancouver, BC, 1993), 109–149, CRM Proc. Lecture Notes, 8, Amer. Math. Soc., Providence, RI, 1995.
- [18] L. Székelyhidi, Jr. Relaxation of the incompressible porous media equation, Ann. Sci. de l’Ecole Norm. Superieure (4) 45 (2012), no. 3, 491–509
- [19] B. Yuan and J. Yuan, Global well-posedness of incompressible flow in porous media with critical diffusion in Besov spaces, J. Differential Equations 246 (2009), no. 11, 4405–4422
- [20] V. I. Yudovich, On the loss of smoothness of the solutions of the Euler equations and the inherent instability of ows of an ideal fluid, Chaos, 10 (2000), 705–719
- [21] A. Zlatos, Exponential growth of the vorticity gradient for the Euler equation on the torus, Adv. Math. 268 (2015), 396–403