Weak and mild solutions to the MHD equations and the viscoelastic Navier–Stokes equations with damping in Wiener amalgam spaces
Abstract
We study the three-dimensional incompressible magnetohydrodynamic (MHD) equations and the incompressible viscoelastic Navier–Stokes equations with damping. Building on techniques developed by Bradshaw, Lai, and Tsai (Math. Ann. 2024), we prove the existence of mild solutions in Wiener amalgam spaces that satisfy the corresponding spacetime integral bounds. In addition, we construct global-in-time local energy weak solutions in these amalgam spaces using the framework introduced by Bradshaw and Tsai (SIAM J. Math. Anal. 2021). As part of this construction, we also establish several properties of local energy solutions with initial data, including initial and eventual regularity as well as small-large uniqueness, extending analogous results obtained for the Navier–Stokes equations by Bradshaw and Tsai (Comm. Partial Differential Equations 2020).
1 Introduction
The incompressible magnetohydrodynamic (MHD) equations describe the interaction of a fluid’s velocity field and a magnetic field within a conducting medium, coupling the incompressible Navier–Stokes equations with Maxwell’s equations of electromagnetism. These fundamental equations are given by
(MHD) |
where is the velocity, the magnetic field, and the pressure. The study of MHD equations has attracted considerable attention. The foundational result of Duvaut and Lions [11] established the global existence of weak solutions with finite energy. Building upon this, Sermange and Temam [38] investigated regularity criteria for weak solutions. Subsequent effort refined these regularity conditions under various assumptions. For instance, Wu [41] and Zhou [44] established Serrin-type criteria and scaling-invariant regularity conditions, while He and Xin [15] and Kang and Lee [19] developed partial regularity results for suitable weak solutions. Further improvements were made via harmonic analysis methods, as in [9], and directionally-restricted criteria, such as those by Cao and Wu [7]. More recently, local regularity theory for MHD has been advanced in parabolic Morrey spaces [8], and Fernández-Dalgo and Jarrín [13] provided weak-strong uniqueness results in weighted spaces, alongside constructions of weak suitable solutions in local Morrey spaces.
On the existence side, Miao, Yuan, and Zhang [35] proved global mild solutions for small data in , and He and Xin [16] constructed self-similar solutions under small homogeneous initial data. Moreover, the existence of forward discretely self-similar and self-similar local Leray solutions is established in the critical space [25] and in the weighted spaces [12]. The criticality of the class was also highlighted in [34], where global regularity of weak solutions was established under this condition. Additional contributions include the construction of global smooth solutions under spectral constraints [30], the use of Morrey spaces to ensure global well-posedness for small data [31], and the construction of forward self-similar solutions via topological and blow-up methods [42].
Complementing the MHD system, the incompressible viscoelastic Navier–Stokes equations with damping (vNSEd) model non-Newtonian fluids with both viscous and elastic characteristics. In the simplified setting where both relaxation and retardation times are infinite, the vNSEd system reads
(1.1) |
with initial data
where is the velocity field, is the local deformation tensor of the fluid, and is the pressure. This model arises from Oldroyd-type theories for viscoelastic fluids and captures the interplay between fluid motion and elastic stresses. The addition of a damping term in the equation for (following Lin–Liu–Zhang [29]) is critical for obtaining global solutions, particularly in the absence of intrinsic dissipative mechanisms. To be more precise, they introduced the following viscoelastic Navier-Stokes equations with damping as a way to approximate solutions of (1.1):
(1.2) |
for a damping parameter . Existence results for smooth solutions under smallness conditions or specific symmetries have been established by various authors [10, 26, 29]. Note that if at some instance of time, then at all later times. In fact, by taking divergence of and using , one have the following equation for :
Hence it is natural to assume
The authors [29] noted that using standard weak convergence methods to pass the limit of solutions to (1.2) as does not yield weak solutions of (1.1). Despite this, system (1.2) remains an interesting subject of study. For instance, Lai, Lin, and Wang [24] established the existence of forward self-similar classical solution to (1.2) for locally Hölder continuous, -homogeneous initial data. Additionally, the existence of forward discretely self-similar and self-similar local Leray solutions in the critical space is established in [25], following the analysis in [2].
Regularity issues for weak solutions of the viscoelastic Navier–Stokes equations with damping have been investigated from several perspectives. Hynd [17] proved a version of the Caffarelli–Kohn–Nirenberg partial regularity theorem adapted to the viscoelastic system with damping, while Kim [22] established Serrin-type regularity criteria in weak- spaces. These results have been further extended in [39], which proved global existence of mild solutions in scaling-invariant spaces for small data and derived various regularity criteria in Lorentz, multiplier, BMO, and Besov spaces. Additional contributions include the construction of global classical solutions with symmetry assumptions in periodic domains by Liu and Lin [32], and refined local energy bounds leading to improved -regularity conditions in the sense of Caffarelli–Kohn–Nirenberg in [43].
Since the damping parameter does not affect our analysis, we set throughout this paper. Then, columnwisely, (1.2) can be rewritten as
(vNSEd) |
where is the -th column vector of .
A central theme in the analysis of both the (MHD) and (vNSEd) systems is the interplay between the nonlinear couplings, scaling symmetries, and the functional framework chosen for solutions. While much progress has been made in classical Lebesgue, Sobolev, and Besov spaces, recent advances have highlighted the utility of Wiener amalgam spaces in studying fluid systems. In this paper, the Wiener amalgam spaces are denoted and defined by the norm
These spaces, which blend local integrability and global decay properties, provide a flexible setting that accommodates non-decaying or large initial data while retaining control over both local and global behaviors. We identify with with the norm . The closure of under the norm is denoted by . Note that for we have the Hölder inequality:
(1.3) |
We will consider two kinds of spacetime integrals: For , , and , define the norms and as follows:
(1.4) |
and
(1.5) |
These norms are different from each other when . By Minkowski’s integral inequality,
(1.6) |
and
(1.7) |
Previous works [4, 1] developed a detailed theory for the incompressible Navier–Stokes equations in Wiener amalgam spaces, establishing mild and weak solutions, spacetime integral bounds, and eventual regularity results for different ranges of the Lebesgue exponent . In this paper, we extend these techniques to the (MHD) and (vNSEd) systems. Specifically, we prove the existence of mild solutions for small data in critical and subcritical Wiener amalgam spaces, as well as global weak solutions under appropriate integrability and decay conditions. Our analysis demonstrates the robustness of the Wiener amalgam framework in addressing the intricate coupling structures and nonlinearities in these models.
In the following subsections, we introduce the definitions of mild and local energy solutions for the systems (MHD) and (vNSEd), and present our main results.
1.1 Mild solutions of MHD equations
A pair of vector fields is called a mild solution to (MHD) if it satisfies
(1.8) |
where is a bilinear operator defined by ,
(1.9) |
in which is the Helmholtz projection operator. More precisely, the vector components of the bilinear operators and can be expressed by
where is the Ossen tensor derived by Oseen in [37]. We refer the readers to [18, Section 2.2] for a brief introduction of the Oseen tensor.
We first consider the case of data with , which we refer to as subcritical, and state the existence of mild solutions in the amalgam spaces in the following theorem.
Theorem 1.1 (Subcritical data (MHD)).
Let and . If are divergence free, then, for any positive time chosen so that
(1.10) |
there exists a unique mild solution to (MHD). Moreover, satisfies
(1.11) |
If , then . If or , then we still have as .
Furthermore, if , then for any and with ,
provided for all .
We now turn to the critical case, i.e., the case when data . When the data is sufficiently small, we have the following existence theorem of mild solutions.
Theorem 1.2 (Critical data I (MHD)).
Let . Fix . There exists such that for all divergence-free with , there exists a mild solution to (MHD) with
The solution is unique in the class
(1.12) |
Furthermore, . We have for and for . If , then we have for any ball and that
(1.13) |
For any and given by , by taking sufficiently small, this solution further satisfies
(1.14) |
The following theorem concerns the critical case with enough decay, .
Theorem 1.3 (Critical data II (MHD)).
Let . For all divergence-free , there exist and a unique mild solution to (MHD) satisfying
with , . For any , , and , there is such that
(1.15) |
Furthermore, there is such that if . If we assume further
(1.16) |
with when , then there exists such that if . Instead of (1.25), if we assume
(1.17) |
then there exists such that if .
1.2 Mild solutions of viscoelastic Navier–Stokes equations with damping
The main results of the mild solutions for the viscoelastic Navier–Stokes equations with damping are stated as follows:
Theorem 1.4 (Subcritical data (vNSEd)).
Let and . If , , where and , , are divergence free, then, for any positive time chosen so that
(1.19) |
there exists a unique mild solution to (vNSEd). Moreover, satisfies
(1.20) |
If , then . If or , then we still have as .
Furthermore, if , then for any and with ,
provided for all .
Theorem 1.5 (Critical data I (vNSEd)).
Let . Fix . There exists such that for all divergence-free , , with , , there exists a mild solution to (vNSEd) with
The solution is unique in the class
(1.21) |
Furthermore, . We have for and for . If , then we have for any ball and that
(1.22) |
For any and given by , by taking sufficiently small, this solution further satisfies
(1.23) |
Theorem 1.6 (Critical data II (vNSEd)).
Let . For all divergence-free , there exist , , and a unique mild solution to (vNSEd) satisfying
with , . For any , , and , there is such that
(1.24) |
Furthermore, there is such that if . If we assume further
(1.25) |
with when , then there exists such that if . Instead of (1.25), if we assume
(1.26) |
then there exists such that if .
1.3 Weak solutions of MHD equations
We first introduce the notion of local energy solutions for the MHD equations, which is consistent with the concept introduced in [4] for the Navier–Stokes equations.
Definition 1.7 (local energy solution (MHD)).
Let . A pair of vector fields , , is a local energy solution to (MHD) with divergence-free initial data , denoted as , if the following hold:
1. There exists such that is a distributional solution to (MHD).
2. For any , satisfies
3. For any , , and , there exists a function of time so that, for every and ,
(1.27) |
in where is the kernel of , , and is the characteristic function of .
4. For all compact subsets of we have and in as .
5. For all cylinders compactly supported in and all nonnegative , we have the local energy inequality
(1.28) |
6. The functions
are continuous in for any compactly supported .
For given divergence-free , let denote the set of all local energy solutions to (MHD) with initial data .
Theorem 1.8 (Eventual regularity in (MHD)).
Assume , where , are divergence free, and . Then has eventual regularity, i.e., there is such that and are regular at whenever , and
for sufficiently large .
Define the local energy
(1.29) |
Theorem 1.9 (Explicit growth rate in (MHD)).
Assume , where , are divergence free, and satisfies, for some ,
Then, for any , with , we have
(1.30) |
for positive constants and independent of and . In particular, if then as .
Theorem 1.10 (Existence in (MHD)).
Assume , where , are divergence free. Then, there exists a time-global local energy solution and associated pressure to (MHD) in with initial data so that, for any ,
(1.31) |
In particular, .
1.4 Weak solutions of viscoelastic Navier–Stokes equations with damping
We now define analogous local energy solutions to the viscoelastic Navier–Stokes equations with damping as follows.
Definition 1.11 (local energy solution (vNSEd)).
Let . A pair of a vector field and a tensor , , , , is a local energy solution to (vNSEd) with initial data , , where are divergence free, denoted as , if the following hold:
1. There exists such that is a distributional solution to (vNSEd).
2. For any , satisfies
3. For any , , and , there exists a function of time so that, for every and ,
(1.32) |
in where is the kernel of , , and is the characteristic function of .
4. For all compact subsets of we have and , , in as .
5. For all cylinders compactly supported in and all nonnegative , we have the local energy inequality
(1.33) |
6. The functions
are continuous in for any compactly supported .
For given divergence-free , , let , , denote the set of all local energy solutions to (vNSEd) with initial data .
The main results concerning weak solutions of the viscoelastic Navier–Stokes equations with damping are stated as follows:
Theorem 1.12 (Eventual regularity in (vNSEd)).
Assume , where , are divergence free, and , . Then has eventual regularity, i.e., there is such that and are regular at whenever , and
for sufficiently large .
Define the local energy
Theorem 1.13 (Explicit growth rate in (vNSEd)).
Assume , where , are divergence free, and , , satisfies, for some ,
Then, for any , with , we have
for positive constants and independent of and . In particular, if then as .
Theorem 1.14 (Existence in (vNSEd)).
Assume , where , are divergence free. Then, there exists a time-global local energy solution and associated pressure to (vNSEd) in with initial data , , so that, for any ,
In particular, .
The remainder of the paper is organized as follows: In Section 2, we construct mild solutions in critical and subcritical spaces and prove Theorems 1.1, 1.2, 1.3, 1.4, 1.5, and 1.6. In Section 3, we examine properties of local energy solutions, including uniqueness and regularity, establish a priori bounds and explicit growth rate, and prove the global existence results: Theorems 1.8, 1.9, 1.10, 1.12, 1.13, and 1.14.
2 Construction of mild solutions in Wiener amalgam spaces
This section is dedicated to proving Theorems 1.1, 1.2, 1.3, 1.4, 1.5, and 1.6. The proof techniques closely follow those outlined in [1, Section 3]. Specifically, we apply the Picard iteration scheme as in [1, Section 3] to the problems (1.8) and (1.18) to construct mild solutions for both the MHD equations (MHD) and the viscoelastic Navier–Stokes equations with damping (vNSEd), respectively. Since the structure of (vNSEd) is analogous to that of (MHD), we prove Theorems 1.1, 1.2, and 1.3 for (MHD). The proofs of Theorems 1.4, 1.5, and 1.6 for (vNSEd) are omitted for brevity.
2.1 Mild solutions in subcritical spaces: Proof of Theorem 1.1
The proof of Theorem 1.1 is an adaption of the proof of [1, Theorem 1.1] for the Navier–Stokes equations to the MHD equations.
Define
By [1, Lemma 2.1], we have the linear estimate
So,
(2.1) |
For bilinear estimate, again by [1, Lemma 2.1], we estimate:
where we used the embedding . Similarly,
Thus, the full bilinear estimate becomes
(2.2) |
We look for a solution of the form
Suppose is small enough so that . Then by the Picard contraction principle, there exists a unique strong mild solution satisfying
(2.3) |
To prove continuity at time zero, assume . Then
Both terms tend to zero as , the latter by [1, Lemma 2.3]. Hence, as . Similarly,
The continuity at can be shown as usual, see e.g., [40, lines 3-8, page 86], including or .
If either or , the semigroup terms no longer vanish, but the bilinear terms still tend to zero:
and
Hence,
as asserted in the theorem.
For uniqueness, let be two mild solutions with initial data . Then, for ,
so that we have
Similarly,
Thus, for small enough , this implies on , and repeating the argument yields uniqueness on .
To obtain the spacetime integral bound, assume , , , and . Consider the Banach space
We can assume since for . From we get . For the linear term, by (2.1) and [1, Lemma 2.4] (which needs and ), we have for a fixed that
For the bilinear term, by (2.2) and [1, Lemma 2.7] with and so that due to , (allowing ),
Since
we derive
Choose small enough such that , the Picard iteration yields a unique strong mild solution . Since , uniqueness implies , and the solution satisfies the desired spacetime integral bound. This completes the proof of Theorem 1.1. ∎
2.2 Mild solutions in critical spaces with small data: Proof of Theorem 1.2
The proof of Theorem 1.2 is an adaption of the proof of [1, Theorem 1.2] for the Navier–Stokes equations to the MHD equations.
Let
and
The inclusion is obvious. We define the spaces
(2.4) |
with corresponding norms
The inclusion follows immediately.
From [1, (2.16)], and choosing from the definition of in [1, Lemma 2.4], we have
(2.5) |
Also, by [1, Lemma 2.1],
(2.6) |
Combining, we obtain
(2.7) |
Using [1, Lemma 2.7] with , , and (so that ), and applying Hölder’s inequality, we estimate the bilinear terms:
(2.8) |
where we used the inclusion in the last inequality. Similarly, we have
(2.9) |
Hence,
(2.10) |
Additionally, applying [1, Lemma 2.1] and Hölder inequality (1.3), we obtain
Similarly,
Hence
Taking small enough, by (2.7), it is possible to make
(2.11) |
The Picard iteration yields a mild solution to (MHD) so that
This solution is unique among all mild solutions with data satisfying . In fact, since we can also apply the Picard contraction to , the solution is also unique in the class .
Next, we show that a solution with small enough initial data also belongs to . Let be the Picard iteration sequence in . By construction,
Note that
By [1, Lemma 2.1],
As is usual in arguments like this, we now seek estimates for in in terms of measurements of and in and . We have by [1, Lemma 2.1] and Hölder inequality (1.3),
(2.12) |
and
(2.13) |
which imply
(2.14) |
Moreover, we have
and
which imply
By switching in the estimates,
We now return to our main objective: proving that the Picard iterates are uniformly bounded in . From the recursive relation, we have
Thus, if , then is uniformly bounded by . To show that the limit , consider the difference of successive iterates:
(2.15) |
This shows the sequence is Cauchy in , and the limit of lies in , since convergence in implies convergence in the full -norm.
Next, we prove convergence to the initial data when . By [1, Lemma 2.3] we have
(2.16) |
whenever . We extend this to the Picard sequence by induction. The base case satisfies (2.16). Suppose the inductive hypothesis holds:
(2.17) |
Then, from the iteration estimate in the class where we are taking , we have
which implies
(2.18) |
Since the Picard sequence converges in , the limit also satisfies
(2.19) |
for sufficiently small. Using (2.14) and (2.19), we find
Combining this with [1, Lemma 2.3], we conclude
If then we have a weaker mode of convergence. Fix a ball . Take large so that . We re-write the bilinear form as
If then we have
by Young’s inequality, so that
For any , the above vanishes as provided .
By taking , we can ensure that for all and we have . Hence, for ,
We now prove continuity at positive times. Let be fixed. We will send . Note that by [1, Lemma 2.3] we have in as . We therefore only need to show . Following [40, p. 86], we take slightly less than so that and write
where , , , or . For the first and second terms, by [1, (2.3)] with , and using the embedding , we have
and
both of which can be made arbitrarily small by taking close to and close to .
For the third term we note that by [1, Lemma 2.3], for each , we have
which follows (even if ) the fact that , which is a consequence of [1, Lemma 2.1]. Additionally,
where integration in is with respect to . So, by Lebesgue’s dominated convergence theorem,
The above show the continuity of at positive times.
To prove the spacetime integral bound (1.23) for , , we work in the spaces with the norms
Note that .
For the linear estimate in , taking so that in [1, Lemma 3.1], we have
(2.20) |
For bilinear estimate, taking , in [1, Lemma 3.2],
By choosing small enough so that , Picard iteration yields a unique mild solution satisfying
Now, we claim that any solution with sufficiently small also belongs to . By [1, (3.21)] of [1, Lemma 3.1], we have
Taking so that in [1, Lemma 3.1], we also obtain
Combining both estimates, we conclude:
(2.21) |
Next, we establish the bilinear estimate:
(2.22) |
Indeed, by applying [1, Lemma 3.2] with and , we obtain
and
Using the same argument as before, we conclude that , possibly after taking a smaller . In particular, , with the norm controlled by . The case in Theorem 1.2 then follows from the embeddings and . ∎
2.3 Mild solutions in critical spaces with enough decay: Proof of Theorem 1.3
The proof of Theorem 1.3 is an adaption of the proof of [1, Theorem 1.3] for the Navier–Stokes equations to the MHD equations.
By [1, Lemma 2.1], we have for ,
(2.23) |
where
(2.24) |
For and , let , be Banach spaces defined as
(2.25) |
and
(2.26) |
with norms
and
respectively. Note that .
By [1, Lemma 2.1] again and Hölder inequality (1.3) using due to ,
(2.27) |
(2.28) |
so that
(2.29) |
Hence,
where is a universal constant.
Concerning the caloric extension of , we have for of any size that
by [1, (2.13)] of [1, Lemma 2.3]. Hence, there exists so that
(2.30) |
If, on the other hand, , then by (2.23), we have (2.30) for . The Picard contraction theorem then guarantees the existence of a mild solution to (MHD) so that
This solution is unique among all mild solutions with data satisfying .
Next, we show that a solution with initial data also belongs to . Let be the Picard iteration sequence in . By construction,
(2.31) |
Note that
We now bound in in terms of and in and . We have by [1, Lemma 2.1] and Hölder inequality (1.3) using ,
(2.32) |
(2.33) |
so that
(2.34) |
By , we have
Based on the above estimates we conclude
(2.37) |
We can now conclude that is Cauchy in by the calculation preceding and including (LABEL:ineq:difference.picard.iterates). However, the smallness of the constant is now provided by (2.30)-(2.31), not by .
We now show continuity. For small data, we can try to inherit continuity from Theorem 1.2. But we will provide a proof valid for general data. We first address convergence to the initial data. By [1, Lemma 2.3] we have
(2.38) |
whenever . By our estimates in the class where we are taking , we have
From this and by induction, for any we have
The limit (2.38), convergence of the Picard iterates in and the above inequality imply that, by taking small, we can make small. To elaborate, we have
(2.39) |
We may choose large so that the first term is small and then make the second term small by taking small. Hence,
(2.40) |
Using (2.34), this implies
This and [1, Lemma 2.3] imply
The proof of continuity for positive times follows from the same argument used in the proof of Theorem 1.2 and is therefore omitted for brevity.
We now prove the spacetime integral bound (1.24) for and . Note that we exclude , i.e., . By imbedding for , we may assume . (We do not take since we need for global existence). Denote the Banach space
For the linear term, by [1, Lemma 2.1] and [1, Lemma 2.4],
(2.41) |
for any and . Note that
(2.42) |
For the bilinear term, by [1, Lemma 2.7] with , , and , so that
(2.43) |
we have
for any and . Note
no matter or . We conclude, also using (2.37),
(2.44) |
By (2.42), we can find so that
Then the Picard sequence satisfies for all , and we get . Thus, satisfies the spacetime integral bound (1.24).
We now establish the global -estimates when is sufficiently small in . To this end, we aim to eliminate the dependence of constants on , i.e., we choose in (2.41) and in (2.44).
Let us analyze the conditions under which in (2.44) is permissible. When , we additionally assume that so that the exponent . In particular, we can take when since the condition required in [1, Lemma 2.7] is satisfied. Note that when , we have , but then .
For , we impose the additional condition , which ensures and hence again allows us to take .
With in (2.41) and (2.44), we obtain a global-in-time estimate for in , provided that the initial data satisfies .
Observe that when , all required conditions–including the upper bound – are satisfied if we take . Once we have established for some , then the inclusion property implies for all . Thus, the condition can be removed entirely.
We now consider the -estimates of , restricting to for simplicity. Fix and , and define by the relation . Using [1, Lemma 2.8] with , and such that , and by applying Hölder inequality, we obtain
The condition is equivalent to . Hence,
(2.45) |
Let denote the set of all for which we can establish . Since for , the set , if nonempty, must be an interval of the form
(2.46) |
for some .
Define by
Given and , and recalling that , we deduce that
(2.47) |
From [1, Lemma 2.4] with , we have
(2.48) |
with constants independent of , provided one of the following holds:
-
A.
, (and if ),
-
E1.
,
-
E2.
and .
If the linear estimate (2.48) holds, and if so that the bilinear estimate (2.45) holds, then the Picard iteration converges in for sufficiently small initial data.
Case A: This case applies as soon as , i.e., . Strict inequality holds when . By (2.47) and since , the value
belongs to both and . Thus, converges in for , or for slightly larger than when and .
Case E1: This case applies when , allowing us to take . It thus covers the parameter range , and .
Case E2: While this case also requires , it does not yield smaller admissible than Case A.
This completes the proof of -estimates, and concludes the proof of Theorem 1.3. ∎
3 Local energy solutions in Wiener amalgam spaces
In this section, we address weak solutions and establish Theorems 1.8, 1.9, 1.10 for the MHD equations (MHD), along with Theorems 1.12, 1.13, 1.14 for the viscoelastic Navier–Stokes equations with damping (vNSEd). Given the similarity between the structures of (vNSEd) and (MHD), we focus on presenting the proofs of Theorems 1.8, 1.9, 1.10 for (MHD). The details of verification of Theorems 1.12, 1.13, 1.14 for (vNSEd) are left to the readers.
Define
and
Lemma 3.1.
Let be divergence free, and assume . For all we have
(3.1) |
(3.2) |
where
and
(3.3) |
for a small universal constant .
Proof.
The proof of the lemma is an adaption of the proof of [3, Lemma 2.1] for the Navier–Stokes equations to the MHD equations.
By Hölder and Young inequalities, for any , we have
and
In addition, applying Sobolev inequality yields
and
where is a positive constant independent of . For the pressure term, using the local pressure expansion (1.27), we obtain
(3.4) |
where
Applying the local energy inequality, we deduce
(3.5) |
for sufficiently small . Therefore, we conclude that
(3.6) |
Making the change of variables and applying Grönwall’s inequality [3, Lemma 2.2], we derive
for where , and
for some small constant independent of and . Note that is nondecreasing in , and its continuity in follows directly from the local energy inequality. With the above estimate for , the lemma follows by the standard continuation in argument, provided that is chosen sufficiently small. ∎
We recall a local regularity criterion for suitable weak solutions that is a replacement for the case of the Navier–Stokes equations proposed in [6] (see also [28, 36]). See [15] for the definition of suitable weak solutions.
Lemma 3.2 (-regularity criterion [34, Theorem 3.1]).
There exists a universal constant such that, if is a suitable weak solution of (MHD) in , , and
then and are Hölder continuous on .
The corresponding -regularity criterion for weak solutions of the viscoelastic Navier–Stokes equations with damping, (vNSEd), is established in [17, Proposition 3.2].
Theorem 3.3 (Initial and eventual regularity).
There is a small positive constant such that the following holds. Assume that are divergence free and that . Let
1. If there exists so that
then has eventual regularity. Moreover, if , then
2. If there exists so that
then has initial regularity. Moreover, if , then
3. If satisfies
then the set of singular times of in is empty. Moreover, for all ,
Proof.
The proof of the theorem is an adaption of the proof of [3, Theorem 1.2] for the Navier–Stokes equations to the MHD equations.
Assume there exists such that for all , we have , where is a small constant to be determined.
Fix and . Define where is the function of from the local pressure expansion (1.27). Then is a suitable weak solution to (MHD) with associated pressure . By the estimate (3.2), we have
Thus, if and , then the right side is bounded by , and we may apply Lemma 3.2. It follows that
and for ,
(3.7) |
Hence, is regular in . Since is arbitrary, we conclude that is regular at every point , with the bound given by (3.7). Note that this threshold depends only on and is uniform for all .
The argument proceeds analogously in the case where ; we omit the details for brevity.
Finally, if for all , then for all . In this case, is regular with the same bound (3.7) throughout the entire space-time domain . ∎
As a consequence of Theorem 3.3, the uniqueness result below can be established by adapting the argument used in the proof of [3, Theorem 1.7].
Theorem 3.4 (Uniqueness for data that is small at high frequencies).
Assume that are divergence free. Let . Then there exist universal constants so that if
for some , then as distributions on , .
Proof.
Assume for some , and either satisfies
(3.8) |
or . Let satisfy . By Theorem 3.3 we have for ,
Using (3.1) we have the estimate
(3.9) |
where is defined in (3.3). Moreover, by item 3 of Theorem 3.3, we have
(3.10) |
Combining (3.9) and (3.10), we deduce that, for ,
(3.11) |
Using the estimates (3.9) for and (3.11) for , we have for all and that
which implies that
(3.12) |
where .
We now check that satisfies the integral formula (1.8), i.e., it is a mild solution, on . If , then this follows from a direct adaption of [20, §8] for the Navier–Stokes equations to the MHD equations. On the other hand, assume satisfies (3.8). By (3.1), we have
Since satisfies (3.8), we have as . So, there exists so that, for all , . We conclude that for any
(3.13) |
where .
Let be defined as
(3.14) |
where is a bilinear operator defined by , where and are given in (1.9). Using [33, (1.8), (1.10)], we have
and
Thus
(3.15) |
Then satisfies
Following the same logic in [20, §8] and adapting it to MHD equations, we have that the mollified and are harmonic in and for fixed
This shows for , for all . Hence and . This shows that any local energy solution with data satisfying the assumptions of Theorem 3.4 is a mild solution.
As a corollary of Theorem 3.4, the following local uniqueness result in can be derived by adapting the proof of [3, Corollary 1.8].
Corollary 3.5 (Local uniqueness in ).
Assume are divergence free. Let and be elements of . Then, there exists so that as distributions on .
Proof.
Assume . Then, in particular, . Let be given. Since , there exists such that
On the other hand, since , their local -norms are uniformly small on sufficiently small balls. That is, there exists such that
Applying Hölder’s inequality, we obtain:
Therefore, by Theorem 3.4, any local energy solution with initial data is unique in the local energy class, at least for a short time. ∎
3.1 Eventual regularity for local energy solutions
Lemma 3.6.
Assume , . Then
Consequently, if , then
Proof.
The proof of the lemma is an adaption of the proof of [4, Lemma 2.2] for the Navier–Stokes equations to the MHD equations.
Let be given. Suppose that for some . Then the norms can be expressed as and , where
For any , we have the estimate
(3.16) |
Now fixe any . Since , we can choose large enough such that and , where
Set and . Following the argument leading to [4, (2.4)] by Hölder’s inequality we obtain, for all ,
To conclude, we first choose small enough so that . Then, for this fixed , we choose sufficiently large to ensure that , which is possible provided .
To prove the final assertions, we begin by noting that . Moreover, if for , then in particular . Therefore,
For the final part, observe that when and , we have . Thus,
∎
3.2 A priori bounds and explicit growth rate
In this section we prove new a priori bounds for data and use it to prove Theorem 1.9.
Lemma 3.7.
Assume for some are divergence free and that satisfies, for some ,
(3.17) |
Then there are positive constants and , both independent of and such that, for all with ,
(3.18) |
where
Furthermore, for all ,
(3.19) |
and
(3.20) |
for all .
Proof.
The proof of the lemma is an adaption of the proof of [4, Lemma 3.1] for the Navier–Stokes equations to the MHD equations.
Let be radial, non-increasing cutoff function such that on , , and , . Let be as in the statement of the lemma, and define the scaled cutoff . Fix .
For each , define the localized energy quantity
We begin by deriving bounds on , which will then be used to control the quantity
in terms of for sufficiently small . By assumption, . To estimate , we apply the local energy inequality (1.33):
We now estimate each term on the right-hand side, beginning with the second term. Using the properties of , we have:
For the cubic terms, we apply the Gagliardo–Nirenberg inequality:
Let . Then, integrating the above estimates over time yields
(3.21) |
where we’ve used in the final step. As a consequence, we have
and
Now, the only remaining term is the pressure term. To estimate it, we use the local pressure expansion (1.27) to write for as
Note that
(3.22) |
for and . This ensures that is well-defined even if and lack decay at infinity.
For the localized term , standard Calderon–Zygmund theory gives
where we used the support and scaling properties of the cutoff functions. Then, applying Hölder’s inequality and the inequality along with the bound from (3.21), we obtain
To estimate , we begin with the following pointwise bound for :
where we used the estimate (3.22), and the convolution is taken over the lattice , with
Assume now that . Using , we estimate
(3.23) |
Note that .
Combining all estimates, we obtain the key bound:
(3.24) |
provided . Note that all constants here are independent of . We now raise both sides of (3.24) to the power and sum over . The left-hand side yields . For the right-hand side:
-
•
The initial data term is controlled by
-
•
The linear term gives
-
•
For the nonlinear cubic term, using for and , we have
-
•
For the convolution term we use Young’s convolution inequality to find
where is bounded independently of .
Now, since , we conclude for and some constant independent of , that
(3.25) |
The right side is finite for by assumption (3.17). It follows from the same argument as in [4, p. 2005], is continuous in . So, from (3.25) we conclude that
provided and , which is achieved if (using )
where . This shows the first estimate (3.18) of Lemma 3.7, for , with . Note that the constants , , and do not depend on and .
For , we replace (3.23) by
Raising both sides of the above inequality to the power and sum over , we get
Above we have used for and . For the convolution term we use Young’s convolution inequality to obtain
where we used the fact that is bounded independently of . We conclude that, for some constant independent of , we have
(3.26) |
The same argument as in [4, p. 2005] shows that is continuous in . Therefore, we conclude from (3.26) and a continuity argument that the estimate (3.18) also holds for .
Proof of Theorem 1.9.
The proof of Theorem 1.9 is an adaption of the proof of [4, Theorem 1.4] for the Navier–Stokes equations to the MHD equations.
Observe that, by the assumption , the estimate (3.16), and Hölder’s inequality, we have
(3.28) |
where and for . Now, applying Lemma 3.7, we obtain the bound
Next, using the definition of from Lemma 3.7 and the estimate (3.28), we find
where . Furthermore, from (3.28), we also have . This yields the desired upper bound in the statement of Theorem 1.9. ∎
Lemma 3.8 (Far-field regularity of local energy solutions with data in ).
Assume for some are divergence free. If is a local energy solution on evolving from with . Then
3.3 Global existence
In this section, we prove Theorem 1.10.
For , we achieve this by considering the perturbed MHD equations
(3.30) |
where and are given divergence-free vector fields. A local energy solution to the perturbed MHD equations, (3.30), is a weak solution satisfying Definition 1.7 with the obvious modifications, namely, and satisfy the perturbed system as distributions and also satisfy the perturbed local energy inequality.
For , we achieve this via the localized and regularized MHD equations:
(3.31) |
where for a spatial mollifier and for a fixed radially decreasing cutoff function satisfying on and .
3.3.1 The case
Lemma 3.9.
Let and be given, let , and let be a spatial mollifier in . Assume that are divergence free and that satisfies and
Then there exist and a mild solution of the integral equation
(3.32) |
for , where
with , and satisfies
(3.33) |
for a universal constant . This is the unique mild solution of (3.32) in the class (3.33). There exists a pressure so that and solve
(3.34) |
in the weak sense on . Finally, and are smooth by the interior regularity of the Stokes equations with smooth coefficients.
Proof.
The proof of the lemma is an adaption of the proof of [4, Lemma 4.1] for the Navier–Stokes equations to the MHD equations.
Note that since . We begin by establishing estimates for the iterates of the Picard scheme. Define the initial iterates as
and for , set
For the initial iterates, it follows from the same approach obtaining [4, (4.7)] that
(3.35) |
For the th iterates, we use the assumption that
We have
(3.36) |
where
The first two terms on the right-hand side of (3.36) have already been estimated in (3.35). For and , using the same technique deriving [4, (4.11), (4.12)], we get
(3.37) |
where
and
(3.38) |
We next estimate the terms and . Using the same argument for [4, (4.13), (4.14)] yields
(3.39) |
(3.40) |
Taking the supremum in time of the left-hand side of (LABEL:eq-4.15-BT-SIMA2021), applying the norm, using Young’s convolution inequality, and raising everything to the power yields
(3.41) |
where is the first part of the norm defined by . So, if is small as determined by , , and (but independently of ), , then the right-hand side of (3.41) is controlled by . This establishes a uniform-in- bound for .
These uniform bounds and the estimation methods above allow us to show the difference estimate
Thus, if is sufficiently small, then is a Cauchy sequence in norm and converges to a limit in the sense that
Uniqueness in the class (3.33) follows from the same difference estimates as before. Suppose and are two mild solutions of (3.32) satisfying (3.33). Then
Hence, if is sufficiently small, we conclude that .
We now recover a pressure associated to . It is known that for some , with . Therefore, the following nonlinear terms belongs to :
Consequently is well-defined. It follows that solves the Stokes system with pressure and forcing term equal to . Adding back the linear term , we see that solve the perturbed, regularized Navier–Stokes equations. The local pressure expansion (1.27) follows from the definition of .
We now establish the estimate
(3.42) |
This follow from the local energy equality satisfied by and the associated pressure , valid for due to the regularity of the solutions due to smoothness and convergence to the data in :
(3.43) |
To estimate the nonlinear terms, we use the bound for . Hence, we have
(3.44) |
and
(3.45) |
From the assumptions on and , we also have
and
These imply the following estimates:
(3.46) |
(3.47) |
(3.48) |
(3.49) |
(3.50) |
and
(3.51) |
The pressure satisfies the local pressure expansion (1.27), which allows us–after incorporating an additive constant–to express it as a sum of two components: , where is a Calderon–Zygmund operator applied to a localized term, and is a nonsingular integral operator acting on data supported away from the ball . Due to the structure of the pressure term in the local energy inequality, the additive constant plays no role and may be disregarded. By applying the Calderon–Zygmund inequality, the contribution from to the local energy inequality can be estimated in the same way as the nonlinear and perturbative terms discussed earlier. Specifically, it is controlled by the right-hand sides of estimates (3.44) through (3.51). We are thus left to estimate only the far-filed component . In ,
Therefore, the contribution of to the local energy equation satisfies
(3.52) |
Lemma 3.10.
Assume that , for some , are divergence free. There exists a small universal constant so that for all and for all divergence free vector fields satisfying
for some and if, additionally, a given local energy solution to the perturbed MHD equations, (3.30), satisfies
then there are positive universal constants and such that
where
Consequently,
Proof.
The proof of the lemma is an adaption of the proof of [4, Lemma 4.2] for the Navier–Stokes equations to the MHD equations.
Once the perturbation terms in the local energy inequality for are estimated, the proof proceeds identically to that of Lemma 3.7 with and .
The linear terms in the perturbed local energy inequality can be estimated as follows:
and
The pressure can be decomposed into local and far-field contributions. The local part contains new terms that are handled exactly as in the previous estimates, once the Calderon–Zygmund inequality is applied. The far-field pressure splits as , where matches the far-field term treated in the proof of Lemma 3.7, and is the remaining contribution. Since the estimate for is already established in the proof of Lemma 3.7, we focus on bounding in . Specifically, we have
where is defined in (3.38), and
This yields the estimate:
Combining the above estimates with the argument in the proof of Lemma 3.7 (see (3.24)), we obtain
(3.53) |
where we are using . The first two lines above coincide exactly with the estimates in the proof of Lemma 3.7, so we focus on the two additional terms in the last line. To control the final term, we apply the norm:
We now choose (which bounds ) sufficiently small so that, after taking the norm of both sides of the inequality, the -weighted terms on the right can be absorbed into the left-hand side. With this absorption, the remaining terms are exactly as in the proof of Lemma 3.7, and the conclusion follows by the same argument. ∎
Lemma 3.11.
Let be given and assume , where is given in Lemma 3.11. Assume that are divergence free and that are divergence free that satisfy
Then there exist and a weak solution and pressure to (3.34) on . Furthermore, we have that and satisfies
for some positive constant independent of and . Here, depends on but not on , , , or .
Proof.
The proof of the lemma is an adaption of the proof of [4, Lemma 4.3] for the Navier–Stokes equations to the MHD equations.
Let be a smooth solution of (3.34) on with initial data. The energy equality for the regularized perturbed problem reads:
By the estimate from [27, p. 217], the right-hand side is uniformly bounded in , implying that for all for some constant independent of . Furthermore, if , then by Lemma 3.10, the local energy estimates extend up to time , yielding for some constant independent of .
Now, let be the time-scale provided in Lemma 3.9 corresponding to initial data of size in . Then, Lemma 3.9 ensures that the solution exists on and belongs to . Since Lemma 3.10 also applies to the regularized system, we further conclude that and hence . In addition, the energy estimate gives . This allows us to restart the solution at any time , and apply Lemma 3.10 again with the same bounds. By uniqueness, the extended solution coincides with the original one, and hence we obtain a solution on that remains in . Repeating this argument and iterating the solution step-by-step, we reach the full time interval . Throughout the iteration, the and norms remain bounded uniformly by and , respectively. Therefore, for each , the solution to the regularized system exists on and satisfies with bounds independent of . ∎
Lemma 3.12.
Proof.
The proof of the lemma is an adaption of the proof of [4, Lemma 4.4] for the Navier–Stokes equations to the MHD equations.
Fix . For each , approximate the data by divergence-free vector fields satisfying . Such approximations can be constructed using the Bogovskii map (see [40]). Let denote the solutions constructed in Lemma 3.12 corresponding to the initial data . By the uniform estimates from Lemma 3.12, we obtain bounds on and in the dual of , which allow us to extract a subsequence of and of , such that, as ,
where for and for some , and is the local pressure expansion for on ball . The limit is a local energy solution to the perturbed MHD equations with initial data . We claim that satisfies the perturbed local energy inequality: for all nonnegative ,
(3.54) |
The first two lines are inherited via standard compactness arguments. We now focus on the convergence of the remaining terms, especially those not involving , which are of higher order. We have
(3.55) |
(3.56) |
(3.57) |
and
(3.58) |
Our goals is to show that the above twelve quantities vanishes as . Let be a ball containing . Then, using Hölder’s inequality and log-convexity of norms, we have
and
as by strong convergence of to in . Next, weak convergence of to in ensures that , , , as , since the products , and all belong to . Finally, for the mollifier terms, we use strong convergence of the mollified quantities in and uniform bounds on , in , to deduce , , , and as . Hence, all error terms vanish in the limit, and (3.54) holds. Moreover, following the argument in [23, (3.28)-(3.29)], we derive the time-slice version of the perturbed local energy inequality: for any nonnegative and any ,
(3.59) |
We now establish the bound for . Let the cutoff function used in the proof of Lemma 3.7, and define for each . Fix a large integer , and restrict to . Applying (3.59) with this choice of , the right-hand side can be approximated by the corresponding terms for the sequence , since all such quantities converge in the limit. In particular, for all , we can ensure that the difference between the terms involving and those for is less than uniformly, provided for some sufficiently large . Taking these approximate terms, applying standard estimates (which can be derived similar to the proof of (3.19))
(3.60) |
Taking the essential supremum in time followed by the -sum over , we obtain a uniform bound
Since this bound holds for all , it follows that , with the norm estimate . ∎
Lemma 3.13.
Proof.
The proof of the lemma is an adaption of the proof of [4, Lemma 5.1] for the Navier–Stokes equations to the MHD equations.
We begin with the special case to highlight the key ideas. Suppose the initial data , and define for . Then,
which shows that with . Let be the solution of the perturbed MHD equations with (so that it is a solution of (MHD)) constructed via Lemma 3.12 with initial data . Then , where is the existence time depending on the size of the initial data in . For almost every , we have that and that . The inclusion follows from Lemma 3.8 and the embedding . Hence, for almost every . In particular, these bounds hold at some time . We now restart the MHD equations at time , treating as new initial data in . By Lemma 3.12, there exists a local energy solution in . By uniqueness of local energy solution with (Corollary 3.5), we have on some short interval . This allows us to glue to , yielding a local energy solution (still denoted ) on that lies in . Hence, . Repeating this argument, we obtain a solution for any , using the uniform-in-time control of .
Now consider the general case . Let , and let be the local energy solution to the perturbed MHD equations, (3.30), given by Lemma 3.12. Assuming , we have for some . We now derive an energy estimate. Using the following bounds:
and
we obtain the energy inequality:
(3.61) |
for . If , the result follows. Otherwise, we choose sufficiently small (depending on ) to absorb the right-hand side, yielding:
This in turn implies
Thus, we obtain uniform-in-time control of and the argument from the case applies to yield the desired result. ∎
Lemma 3.14.
Suppose that are divergence free. Assume also that for a universal constant . Then there exists a second universal constant and and comprising a local energy solution to (MHD) in with initial data so that and are smooth in space and time, , and
Furthermore, if , then on .
Proof.
The proof of the lemma is an adaption of the proof of [4, Lemma A.1] for the Navier–Stokes equations to the MHD equations.
Since , it follows that as . Viewing the MHD equations as a coupled system of inhomogeneous Stokes systems, we apply the linear theory from [21, §5] and follow the argument of Theorem 1.5 therein to construct a global-in-time local energy solution evolving from . We may assume , where is given in Theorem 3.4. Then, for all and ,
Thus, Theorem 3.4 ensures uniqueness of the local energy solution up to some time .
Now consider the mild solution constructed in Theorem 1.1 with , . Since , the closure of in the norm, there exists a time and a unique mild solution . In the construction of this mild solution (see Section 2.1), we may redefine the space with the norm to ensure that the mild solution for all . Then, by adapting the regularity argument from [14, §4], we conclude that is smooth in both space and time for all . By choosing sufficiently small, the existence time for the mild solution in Theorem 1.1 exceeds .
We now verify that defines a local energy solution. By embeddings, the same convergence properties at hold with and replaced by and , respectively. This implies that if is compactly supported, then
Moreover, the maps
are continuous for due to the smoothness of .
To complete the verification, we adapt the pressure construction from [5, Theorem 1.4], as carried out in [5, §6] for the Navier–Stokes equations. This yields a pressure such that satisfies the MHD equations in the distributional sense. The local expansion of also guarantees that . The local energy inequality follows from the space-time smoothness of , and item 2 in the definition of local energy solution is satisfied since . Hence, , and uniqueness implies on . Therefore,
for all .
Proof of Theorem 1.10 for .
The proof of Theorem 1.10 for is an adaption of the proof of [4, Theorem 1.5] for the Navier–Stokes equations to the MHD equations.
Assume that are divergence free. By Lemma 3.12 with , there exists a local energy solution to the MHD equations on such that . Moreover, by Lemma 3.8, we have
Choose a time so that . We aim to construct a local energy solution in with initial data , where is the fixed time-scale in Lemma 3.14. Using the Bogovskii map (see [40] for the details), for any , we can decompose the initial data as
where
By Lemma 3.14, choosing sufficiently small ensures the existence of a local energy solution with pressure , defined on , evolving from the initial data , which is smooth in both space and time and satisfies
Furthermore, by the uniqueness result in Lemma 3.14, this solution coincides with the one given by Lemma 3.12, and hence .
Next, we apply Lemma 3.13 with the perturbation factor , again choosing sufficiently small to ensure that the time-scale it yields is at least . This gives a local energy solution to the the perturbed MHD equation with initial data and associated pressure . Define . This gives a local energy solution on . To very the local energy inequality for , we use approximations , as in the proof of Lemma 3.12, and apply the inequality to .
3.3.2 The case
Now, we consider the case and look at the localized and regularized MHD equations, (3.31). The following lemma corresponds to [1, Lemma 4.6] for the Navier–Stokes equations.
Lemma 3.15.
Let . For each and divergence free with , , if , we can find a unique solution to the integral form of (3.31)
(3.62) |
satisfying
where and are absolute constants.
Proof.
The proof of the lemma is an adaption of the proof of [1, Lemma 4.6] for the Navier–Stokes equations to the MHD equations.
Let denote the mapping defined by the right-hand side of (3.62) for . By [1, Lemma 2.9] and the assumption , we obtain the estimate
Therefore, for some constants ,
To estimate the difference, let . Then
Applying the Picard contraction principle, we see that if the time satisfies , then has a unique fixed point with , solving the integral system (3.62). ∎
Lemma 3.16.
Let , , be divergence free. For each , we can find in and in for some positive which solve the localized and regularized MHD equations, (3.31), in the sense of distributions and in as for any compact subset of .
Proof.
The proof of the lemma is an adaption of the proof of [1, Lemma 4.7] for the Navier–Stokes equations to the MHD equations. We provide the corresponding details by following the same logic used in the proof of [23, Lemma 3.4] for the Navier–Stokes equations, adapting it from the framework to the setting, and from the Navier–Stokes equations to the MHD equations.
By Lemma 3.15, there is a mild solution to (3.62) for . Apparently,
and
where we have used [1, Lemma 2.9] and assumed . Also, for any compact subset of , we have , as goes to by Legesgue’s convergence theorem. Then, it follows that and for any compact subset of .
Note that both and , with , solve the homogeneous heat equation in the distribution sense. Also, using , we can easily see that .
On the other hand, and imply
Hence by the classical theory, and defined by
solves the Stokes system with the source term in the distribution sense. Moreover, solves the forced heat equation with the forcing in the distribution sense. By adding the homogeneous heat equation for with and the Stokes system for and , satisfies
in the sense of distribution. Moreover, by adding the homogeneous heat equation for with and the forced heat equation for , satisfies
in the sense of distribution. ∎
We next show global existence for the localized and regularized MHD equations, (3.31).
Lemma 3.17.
Proof.
The proof of the lemma is an adaption of the proof of [1, Lemma 4.8] for the Navier–Stokes equations to the MHD equations.
We set the radius . By (3.28), we have the bound for all . Define
where the constants and are as in Lemma 3.7. Note that is increasing and as . Now, by the same argument used in the proof of Lemma 3.7, if a solution of (3.31) satisfies , then it satisfies the a priori bounds (3.18) and (3.20) on the interval with radius .
Since the system (3.31) is a coupled system of inhomogeneous Stokes systems with localized and regularized forcing, standard theory guarantees the existence of unique global solution . By uniqueness, this solution agrees with the -solution constructed in Lemma 3.16, and thus for some . Fix . Applying (3.18) with , we obtain . Then, by Lemma 3.16, there exists an -solution on for some . By uniqueness, this solution coincides with , and we conclude that , with the a priori bound (3.18) valid up to time . This argument can be iterated: by repeatedly extending the solution, we obtain for , until . Thus, for each , we obtain , with the a priori bound (3.18) holding for up to time . Since , the lemma follows. ∎
Proof of Theorem 1.10 for .
The proof of Theorem 1.10 for is an adaption of the proof of [1, Theorem 1.4] for the Navier–Stokes equations to the MHD equations. We provide the necessary details by following the same stragegy as in the proof of [3, Theorem 1.5] for the Navier–Stokes equations, adapting the argument from the framework to the setting, and from the Navier–Stokes to the MHD equations.
For , let be the solution of the localized, regularized MHD equations, (3.31), with , given in Lemma 3.17. They share the same a priori bound (3.18) for up to time , thus
Using this a priori bound, we now construct the desired global solutions as the limit of defined in , by induction. Lemma 3.17 implies that are uniformly bounded in the class from inequalities
(3.63) |
(3.64) |
(3.65) |
where is the function of in (3.19) with and , and
(3.66) |
where is the space dual to . Hence there exists a sequence (where the corresponding are denoted by ) that converges to a solution of (MHD) on in the following sense:
By Lemma 3.17 all are also uniformly bounded on for , and, recursively, we can extract subsequences from which converge to a solution of (MHD) on as in the following sense:
Let be the extension by of to . Note that, at each step, agrees with on . Let . Then on for every .
Let on and equal elsewhere. Let denote the corresponding regularization parameter. Then, for every fixed and as ,
(3.67) |
Based on the uniform bounds for the approximates, we have that satisfies (1.30).
From the convergence properties of , it follows that in for all where in which is defined for by
with
We have and
Thus is independent of for .
We now establish the above local pressure expression for all scales. Note that the formula is valid for at all scales, that is, for any , fixed and , we have the following equality in ,
where and . Similarly, let
Fix , and . Choose large enough that . We claim that converges to in . If this is the case, by taking the limit of the weak form of (3.31), we can show that also satisfies (MHD) in . Hence , and we may define
which is hence a function of in that is independent of . This gives the desired local pressure expansion in .
To verify the claim we work term by term. Note that the estimate in [23, (3.26)] shows that
as for every and . This implies
in , and
in for every . For the far-field part, still assuming , we have
where we’ve used the embedding . This can be made arbitrarily small by taking large and noting and are fixed. Consequently, and since the other parts of the pressure converge, we conclude that
(3.68) |
which leads to the desired local pressure expansion. Since was arbitrary, this gives the pressure formula for arbitrarily large times.
At this point we have established items 1.-3. from the definition of local energy solutions. We now check remaining items.
Fix and choose so that . Then (3.67) holds for all with replaced by . Furthermore, the estimates (3.63)–(3.66) and (3.68) are valid in up to a re-definition of . Moreover, we have
(3.69) |
and
(3.70) |
It follows from (3.69) and (3.70) that for every ,
(3.71) |
are continuous in for every . Since was arbitrary, we can extend this to all times. The local energy inequality follows from the local energy equality for and , and (3.67), (3.68) in , (3.69), and for some . Convergence to initial data in follows from (3.71) and the local energy inequality. This confirms that items 4.-6. from the definition of local energy solutions are satisfied and finishes the proof of Theorem 1.10 for . ∎
Acknowledgments
I warmly thank Zachary Bradshaw and Tai-Peng Tsai for helpful comments. The research was partially support by the AMS-Simons Travel Grant and the Simons Foundation Math + X Investigator Award #376319 (Michael I. Weinstein). The author gratefully acknowledges the unwavering financial and emotional support of his wife, Anyi Bao.
References
- [1] Z. Bradshaw, C.-C. Lai, and T.-P. Tsai. Mild solutions and spacetime integral bounds for Stokes and Navier-Stokes flows in Wiener amalgam spaces. Math. Ann., 388(3):3053–3126, 2024.
- [2] Z. Bradshaw and T.-P. Tsai. Forward discretely self-similar solutions of the Navier-Stokes equations II. Ann. Henri Poincaré, 18(3):1095–1119, 2017.
- [3] Z. Bradshaw and T.-P. Tsai. Global existence, regularity, and uniqueness of infinite energy solutions to the Navier-Stokes equations. Comm. Partial Differential Equations, 45(9):1168–1201, 2020.
- [4] Z. Bradshaw and T.-P. Tsai. Local energy solutions to the Navier-Stokes equations in Wiener amalgam spaces. SIAM J. Math. Anal., 53(2):1993–2026, 2021.
- [5] Z. Bradshaw and T.-P. Tsai. On the local pressure expansion for the Navier–Stokes equations. J. Math. Fluid Mech., 24:1–32, 2022.
- [6] L. Caffarelli, R. Kohn, and L. Nirenberg. Partial regularity of suitable weak solutions of the Navier-Stokes equations. Comm. Pure Appl. Math., 35(6):771–831, 1982.
- [7] C. Cao and J. Wu. Two regularity criteria for the 3D MHD equations. J. Differential Equations, 248(9):2263–2274, 2010.
- [8] D. Chamorro, F. Cortez, J. He, and O. Jarrín. On the local regularity theory for the magnetohydrodynamic equations. Doc. Math., 26:125–148, 2021.
- [9] Q. Chen, C. Miao, and Z. Zhang. On the regularity criterion of weak solution for the 3D viscous magneto-hydrodynamics equations. Comm. Math. Phys., 284:919–930, 2008.
- [10] Y. Chen and P. Zhang. The global existence of small solutions to the incompressible viscoelastic fluid system in 2 and 3 space dimensions. Comm. Partial Differential Equations, 31(10-12):1793–1810, 2006.
- [11] G. Duvaut and J.-L. Lions. Inéquations en thermoélasticité et magnétohydrodynamique. Arch. Rational Mech. Anal., 46:241–279, 1972.
- [12] P. G. Fernández-Dalgo and O. Jarrín. Discretely self-similar solutions for 3D MHD equations and global weak solutions in weighted spaces. J. Math. Fluid Mech., 23(1):Paper No. 22, 30, 2021.
- [13] P. G. Fernández-Dalgo and O. Jarrín. Weak-strong uniqueness in weighted L2 spaces and weak suitable solutions in local Morrey spaces for the MHD equations. J. Differential Equations, 271:864–915, 2021.
- [14] Y. Giga, K. Inui, and S. Matsui. On the cauchy problem for the Navier-Stokes equations with nondecaying initial data. Quad. Mat., 4:28–68, 1999.
- [15] C. He and Z. Xin. Partial regularity of suitable weak solutions to the incompressible magnetohydrodynamic equations. J. Funct. Anal., 227(1):113–152, 2005.
- [16] C. He and Z. Xin. On the self-similar solutions of the magneto-hydro-dynamic equations. Acta Math. Sci. Ser. B (Engl. Ed.), 29(3):583–598, 2009.
- [17] R. Hynd. Partial regularity of weak solutions of the viscoelastic Navier-Stokes equations with damping. SIAM J. Math. Anal., 45(2):495–517, 2013.
- [18] K. Kang, B. Lai, C.-C. Lai, and T.-P. Tsai. The Green tensor of the nonstationary Stokes system in the half space. Comm. Math. Phys., 399(2):1291–1372, 2023.
- [19] K. Kang and J. Lee. Interior regularity criteria for suitable weak solutions of the magnetohydrodynamic equations. J. Differential Equations, 247(8):2310–2330, 2009.
- [20] K. Kang, H. Miura, and T.-P. Tsai. Short time regularity of Navier–Stokes flows with locally L3 initial data and applications. Int. Math. Res. Not. IMRN, 2021(11):8763–8805, 2021.
- [21] N. Kikuchi and G. Seregin. Weak solutions to the Cauchy problem for the Navier-Stokes equations satisfying the local energy inequality. In Nonlinear equations and spectral theory, volume 220 of Amer. Math. Soc. Transl. Ser. 2, pages 141–164. Amer. Math. Soc., Providence, RI, 2007.
- [22] J.-M. Kim. On regularity criteria of weak solutions to the 3D viscoelastic Navier-Stokes equations with damping. Appl. Math. Lett., 69:153–160, 2017.
- [23] H. Kwon and T.-P. Tsai. Global Navier–Stokes flows for non-decaying initial data with slowly decaying oscillation. Comm. Math. Phys., 375(3):1665–1715, 2020.
- [24] B. Lai, J. Lin, and C. Wang. Forward self-similar solutions to the viscoelastic Navier-Stokes equation with damping. SIAM J. Math. Anal., 49(1):501–529, 2017.
- [25] C.-C. Lai. Forward discretely self-similar solutions of the MHD equations and the viscoelastic Navier-Stokes equations with damping. J. Math. Fluid Mech., 21(3):Paper No. 38, 28, 2019.
- [26] Z. Lei, C. Liu, and Y. Zhou. Global solutions for incompressible viscoelastic fluids. Arch. Ration. Mech. Anal., 188(3):371–398, 2008.
- [27] P. G. Lemarié-Rieusset. Recent developments in the Navier-Stokes problem, volume 431 of Chapman & Hall/CRC Research Notes in Mathematics. Chapman & Hall/CRC, Boca Raton, FL, 2002.
- [28] F. Lin. A new proof of the Caffarelli-Kohn-Nirenberg theorem. Comm. Pure Appl. Math., 51(3):241–257, 1998.
- [29] F.-H. Lin, C. Liu, and P. Zhang. On hydrodynamics of viscoelastic fluids. Comm. Pure Appl. Math., 58(11):1437–1471, 2005.
- [30] Y. Lin, H. Zhang, and Y. Zhou. Global smooth solutions of MHD equations with large data. J. Differential Equations, 261(1):102–112, 2016.
- [31] F. Liu, S. Xi, Z. Zeng, and S. Zhu. Global mild solutions to three-dimensional magnetohydrodynamic equations in Morrey spaces. J. Differential Equations, 314:752–807, 2022.
- [32] M. Liu and X. Lin. Global classical solutions to the elastodynamic equations with damping. J. Inequal. Appl., 2021(1):88, 2021.
- [33] Y. Maekawa and Y. Terasawa. The Navier-Stokes equations with initial data in uniformly local spaces. Differential Integral Equations, 19(4):369–400, 2006.
- [34] A. Mahalov, B. Nicolaenko, and T. Shilkin. -solutions to the MHD equations. J. Math. Sci, 143:2911–2923, 2007.
- [35] C. Miao, B. Yuan, and B. Zhang. Well-posedness for the incompressible magneto-hydrodynamic system. Math. Methods Appl. Sci., 30(8):961–976, 2007.
- [36] J. Nečas, M. Růžička, and V. Šverák. On Leray’s self-similar solutions of the Navier-Stokes equations. Acta Math., 176(2):283–294, 1996.
- [37] C. W. Oseen. Neuere methoden und ergebnisse in der hydrodynamik, volume 1. Akademische Verlagsgesellschaft, 1927.
- [38] M. Sermange and R. Temam. Some mathematical questions related to the MHD equations. Comm. Pure Appl. Math., 36(5):635–664, 1983.
- [39] Z. Tan, W. Wu, and J. Zhou. Existence of mild solutions and regularity criteria of weak solutions to the viscoelastic Navier–Stokes equation with damping. Commun. Math. Sci., 18(1):205–226, 2020.
- [40] T.-P. Tsai. Lectures on Navier-Stokes equations, volume 192 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, 2018.
- [41] J. Wu. Regularity results for weak solutions of the 3D MHD equations. Discrete Contin. Dyn. Syst, 10(1/2):543–556, 2004.
- [42] Y. Yang. Forward self-similar solutions to the mhd equations in the whole space. arXiv preprint arXiv:2404.02601, 2024.
- [43] G. Yue, Z. Pang, and Y. Wu. On the interior regularity criteria for the viscoelastic fluid system with damping. Adv. Calc. Var., 18(2):323–338, 2025.
- [44] Y. Zhou. Remarks on regularities for the 3D MHD equations. Discrete Contin. Dyn. Syst, 12(5):881–886, 2005.