1 Introduction
Convex integration methods, introduced by Nash [Reference Nash40] in the context of the
$C^1$
isometric embedding problem and subsequently refined by Gromov in his work on flexible geometric PDEs and by Müller and Šverák [Reference Müller and Šverák39] in their theory of differential inclusions, have experimented an extraordinary development in connection with the study of weak solutions of the incompressible Euler equations. This system reads:

where the time-dependent vector field v is the velocity of the fluid and the scalar function p is the hydrodynamic pressure. One typically considers the Euler equations either on the whole space
$\mathbb {R}^3$
, or on the torus
$\mathbb {T}^3:=(\mathbb {R}/\mathbb {Z})^3$
, or on a bounded domain
$\Omega \subset \mathbb {R}^3$
with smooth boundary (where additional complications may arise).
The motivation to consider weak solutions in this setting is twofold. First, the 3d Euler equations are expected to dynamically produce singularities from smooth initial conditions [Reference Hou34, Reference Wang, Lai, Gómez-Serrano and Buckmaster51]. Second, weak solutions are necessary to describe some of the phenomena that appear in turbulence, such as the energy dissipation in nonsmooth Euler flows famously conjectured by Onsager in 1949 [Reference Onsager42]. Roughly speaking, Onsager’s conjecture asserts that weak solutions that are Hölder continuous in space with exponent greater than
$1/3$
must conserve energy, while for any smaller exponent there should be weak solutions that do not.
The rigidity part of Onsager’s conjecture was proved by Constantin, E and Titi [Reference Constantin, W and Titi18] after a partial result of Eyink [Reference Eyink29]. The endpoint case was addressed in [Reference Cheskidov, Constantin, Friedlander and Shvydkoy13]. Concerning the flexible part of the conjecture, following the construction of
$L^2$
solutions with compact support in space and time due to Scheffer [Reference Scheffer44] and Shnirelman [Reference Shnirelman46], a systematic approach was introduced in the seminal work of De Lellis and Székelyhidi, who introduced
$L^\infty $
-convex integration and
$C^0$
-Nash iteration schemes in this setting [Reference De Lellis and Székelyhidi23, Reference De Lellis and Székelyhidi24]. After a series of significant intermediate results [Reference Daneri and Székelyhidi21, Reference Buckmaster, De Lellis, Isett and Székelyhidi6, Reference Isett and Oh36], the flexible part of Onsager’s conjecture was finally established by Isett [Reference Isett and Oh36], and further refined by Buckmaster, De Lellis, Székelyhidi, and Vicol [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7] to construct solutions for which the kinetic energy is strictly decreasing. In addition to the classical Hölder-based approach, the so-called intermittent
$L^p$
-based flavor of convex integration, introduced by Buckmaster and Vicol [Reference Buckmaster and Vicol10] to prove the nonuniqueness of weak solutions of the 3d Navier–Stokes equations, has also attracted much attention, as it can effectively capture new aspects of Kolmogorov’s theory of turbulence. For detailed expositions of these and other results on various models in fluid mechanics, we refer the reader to the surveys [Reference Buckmaster and Vicol11, Reference De Lellis and Székelyhidi22] and the papers [Reference Albritton, Brué and Colombo1, Reference Buckmaster, Masmoudi, Novack and Vicol8, Reference Buckmaster, Shkoller and Vicol9, Reference Castro, Córdoba and Faraco12, Reference Cheskidov and Shvydkoy15, Reference Cheskidov and Luo14, Reference Cheskidov and Shvydkoy16, Reference Faraco, Lindberg and Székelyhidi30, Reference Faraco, Lindberg and Székelyhidi31, Reference Novack and Vicol41, Reference Shvydkoy47].
A key property of the solutions that one constructs using convex integration techniques is their flexibility. This refers to the fact that, at a certain regularity level, the equations are no longer predictive: there exist infinitely many solutions, in stark contrast to what happens in the case of smooth solutions. Three possible formulations of this property are as follows; as discussed in [Reference Novack and Vicol41, Remark 1.3], once one of them has been established within a certain functional framework, it is usually straightforward to pass to another formulation using techniques that are now standard.
Restricting to the case of
$\mathbb {T}^3$
for concreteness, let us denote by
$\mathcal V\subset L^2(\mathbb {T}^3)$
some suitable function space, which in our case will be some Hölder space
$C^\beta (\mathbb {T}^3)$
. Three standard ways of stating the flexibility of weak solutions in this regularity class are as follows:
-
1. Solutions of compact time support: Given any positive constants
$E,T$ , there exists a weak solution
$v\in C([-T,T], \mathcal V)$ whose time support is contained in
$(-T,T)$ and such that
$\|v(0)\|_{L^2(\mathbb {T}^3)}>E$ .
-
2. Solutions with fixed energy profile: Given any smooth positive function
$e:[0,T]\to (0,\infty )$ , there exists a weak solution
$v\in C([0,T],\mathcal V)$ such that
$\|v(t)\|_{L^2(\mathbb {T}^3)}=e(t)$ for all
$t\in [0,T]$ .
-
3. Arbitrary initial and final states: Given any divergence-free vector fields
$v_{\mathrm {start}},v_{\mathrm {end}}\in \mathcal V$ with the same mean, any
$T>0$ and any
$\epsilon>0$ , there exists a weak solution
$v\in C([0,T],\mathcal V)$ such that
(1.1)$$ \begin{align} \|v(0)-v_{\mathrm{start}}\|_{L^2(\mathbb{T}^3)}+\|v(T)-v_{\mathrm{end}}\|_{L^2(\mathbb{T}^3)}<\epsilon \,. \end{align} $$
(If
$v_{\mathrm {start}},v_{\mathrm {end}}$ are smooth, one can take
$\epsilon =0$ by gluing in time.)
1.1 Main result
Our objective in this paper is to prove an extension theorem for local solutions of the 3d incompressible Euler equations. Roughly speaking, we prove that if a smooth vector field satisfies the Euler equations in a spacetime region
$\Omega \times (0,T)$
(so it is a “local” solution of Euler), one can choose a weak solution on
$\mathbb R^3\times (0,\infty )$
of class
$C^\beta $
for any
$\beta <1/3$
(which is the sharp Hölder regularity) such that both fields coincide on
$\Omega \times (0,T)$
. Moreover, one controls the spatial support of the “global solution” which extends the local one.
This property is very different from the approximation theorems that one can prove for smooth solutions of various classes of linear PDEs [Reference Enciso and Peralta-Salas26, Reference Enciso and Peralta-Salas27, Reference Enciso, García-Ferrero and Peralta-Salas25, Reference Enciso and Peralta-Salas28], and also from the fact (often known as h-principle) that weak solutions of certain regularity can approximate, in Sobolev spaces of negative index, any given subsolution of the Euler equations.
Before presenting this result, let us recall the definition of weak solution (which, as we will be dealing with continuous functions exclusively, is just the distributional one). More precisely, given some
$T\in (0,\infty ]$
and some open set
$\Omega \subseteq \mathbb {R}^3$
with smooth boundary, we will say that a vector field
$v\in C(\Omega \times [0,T),\mathbb {R}^3)$
is a weak solution of the Euler equations on
$\Omega \times (0,T)$
if

for all divergence-free
$\varphi \in C^\infty _c(\Omega \times (0,T),\mathbb {R}^3)$
, and
$\operatorname {\mathrm {div}} v=0$
in the sense of distributions.
The main result of this paper can then be stated as follows:
Theorem 1.1. Fix some
$T>0$
and a bounded open set
$\Omega \subset \mathbb {R}^3$
with smooth boundary and with a finite number of connected components. Assume that
$v_0\in C^{\infty }(\overline {\Omega }\times [0,T],\mathbb {R}^3)$
is a solution of the Euler equations on the spacetime region
$\Omega \times (0,T)$
. Then, for any
$0<\beta <1/3$
, there exists an admissible weak solution
$v \in C^\beta (\mathbb {R}^3\times [0,T])$
of the Euler equations such that
$v|_{\overline {\Omega }\times [0,T]}=v_0$
if and only if

for all
$a\in \mathbb {R}^3$
, all
$t\in [0,T]$
and all connected components
$\Sigma $
of
$\partial \Omega $
. These conditions are automatically satisfied if
$\partial \Omega $
is connected. Furthermore, there exists
$e_0>0$
such that we may prescribe any energy profile
$e\in C^\infty ([0,T],[e_0,+\infty ))$
, that is,
$\|v(t)\|_{L^2(\mathbb {R}^3)}=e(t)$
. In addition, given any open set
$\Omega '\supset \overline \Omega $
, one can in fact assume that the spatial support of v is contained in this region.
Remark 1.2. In fact, one can obtain a global weak solution
$v\in C^\beta (\mathbb {R}^3\times [0,+\infty ))$
such that
$v|_{\mathbb {R}^3\times [0,T]}$
satisfies the claimed properties. Its temporal support may be assumed to be contained in
$[0,T']$
for any
$T'>T$
. We cannot then prescribe the energy profile for
$t>T$
, but we can still choose v so that it remains admissible, that is,

Remark 1.3. It can be proved [Reference Constantin17] that the pressure function associated to this weak solution is in
$ L^\infty _tC^{2\beta }_x\cap C^{2\beta -\delta }_{xt}$
for any
$\delta>0$
.
Before moving on to discuss some applications, let us provide some intuition about the compatibility conditions (1.2). When
$\partial \Omega $
is connected, it is easy to see that any smooth Euler flow
$v_0$
on
$\Omega $
satisfies this condition. Indeed, these two conditions are respectively obtained by integrating over the domain
$\Omega $
the incompressibility condition
$\operatorname {\mathrm {div}} v_0=0$
and the projected Euler equation
$a\cdot (\partial _t v_0+\operatorname {\mathrm {div}}(v_0\otimes v_0)+\nabla p_0)=0$
. If
$v_0$
is the restriction to
$\Omega $
of a global Euler flow, one can refine the argument to show that these conditions must hold on each connected component
$\Sigma $
of the boundary
$\partial \Omega $
, and not every field satisfying the Euler equations on
$\Omega $
will satisfy them. Details are given in Lemma 2.11.
1.2 Applications
We shall next present two applications of the above extension result to the analysis of weak solutions of the 3d Euler equations. These applications do not follow directly from our main theorem, but they use it in an essential way.
Specifically, for these applications we consider subsolutions that are not smooth up to the endpoints of the interval
$(0,T)$
, which implies a lack of uniform-in-time bounds. Thus the scheme does not work as is because the available bounds are not uniform, but we will show in Section 9 that one can tweak the construction in many interesting situations.
The first application we consider concerns the case of the standard vortex sheet
$u_0$
, which we can define as the periodic extension to
$\mathbb T^3$
of:

It follows from the classical local existence results and from the weak-strong uniqueness property [Reference Brenier, De Lellis and Székelyhidi5, Reference Wiedemann52] that wild initial data must be somewhat irregular. However, until the publication of [Reference Székelyhidi49] it was not known how irregular they must be. In that paper it was proved that the vortex sheet
$u_0$
is a wild initial data but the constructed solutions are only in
$L^\infty $
. Results for nonflat vortex sheets have been recently established in [Reference Mengual and Székelyhidi38].
One can use a suitable modification of our main theorem to extend this result to solutions of class
$C^\beta _{\text {loc}}$
:
Theorem 1.4. Let
$0<\beta <1/3$
and let
$T>0$
. There exist infinitely many admissible weak solutions of the Euler equations
$v \in C^\beta _{\mathrm {loc}}(\mathbb {T}^3\times (0,T))$
with initial datum
$u_0$
. For all
$t\in (0,T)$
,
$v(x,t)$
coincides with
$u_0(x)$
outside a “turbulent” zone of size
$O(t)$
.
The second application we will present concerns the existence of a wealth of reasonably well behaved solutions that blow up on a set of maximal Hausdorff dimension. To make this precise, let us say that a point
$(x_0,t_0)$
in spacetime is in the singular set of v, which we will denote by
$\mathscr {S}^\infty _v$
, if
$v\not \in L^\infty ((t_0-\delta ,t_0+\delta )\times B)$
for any ball
$B\ni x_0$
and any
$\delta>0$
. More generally, the q-singular set of v,
$\mathscr S_v^q$
, consists of the spacetime points
$(x_0,t_0)$
such that
$v\not \in L^\infty ((t_0-\delta ,t_0+\delta ), L^q( B))$
for any ball B and any
$\delta>0$
as above. Clearly
$\mathscr S_v^{q}\subset \mathscr S_v^{q'}$
if
$q<q'$
and
$\mathscr S_v^q$
is a closed set.
We are now ready to state the result. Basically, the theorem says that, given any smooth solution
$v_0$
on
$\Omega \times (0,T)$
and any open set
$U\subset \Omega $
, there is an admissible weak solution v which coincides with
$v_0$
outside U and which is uniformly close to
$v_0$
at time
$0$
, yet blows up dramatically on a subset of
$U\times (0,T)$
of full dimension. Interestingly, smooth stationary Euler flows with compact support [Reference Gavrilov32, Reference Constantin, La and Vicol19] are very useful as building blocks in the construction of these solutions.
Theorem 1.5. Consider some
$0<\beta <1/3$
and some
$q>2$
. Let
$T>0$
and let
$\Omega $
be
$\mathbb {T}^3$
or an open subset of
$\mathbb {R}^3$
. Fix some open set U whose closure is contained in
$\Omega $
. Let
$v_0$
be a smooth solution of the Euler equations in
$\Omega \times (0,T)$
. For any
$\varepsilon>0$
there exists a weak solution
$v\in L^2(\Omega \times (0,T))$
of the Euler equations whose q-singular set
$\mathscr {S}_v^q$
is contained in
$ U\times (0,T]$
and has Hausdorff dimension
$4$
. Furthermore, v coincides with
$v_0$
on
$(\Omega \backslash U)\times [0,T]$
and satisfies

Moreover,
$v\in C^\beta _{\text {loc}}((\Omega \times [0,T])\backslash \mathscr {S}_v^q)$
and the energy profile
$\int _\Omega \left \vert v\right \vert {}^2\,dx$
can be chosen to be nonincreasing.
1.3 Strategy of the proof
We prove Theorem 1.1 in two stages: first we extend the field to
$\mathbb {R}^3\times [0,T]$
as a smooth subsolution (see Definition 2.1 in the main text), and then we use a Nash iteration to perturb it into a weak solution. These stages are interrelated in that tools and ideas that we develop to manipulate subsolutions also play a fundamental role in our convex integration scheme.
Concerning the extension of a local smooth solution of the Euler equations as a subsolution, the key result we prove is the following. In view of future applications of this result, which will appear elsewhere, we are stating these results for the Euler equations in any spatial dimension
$n\geq 2$
.
Theorem 1.6. Let
$\Omega _0\subset \mathbb {R}^n,\;n\geq 2$
be a bounded open set with smooth boundary and finitely many connected components and let
$I\subset \mathbb {R}$
be a closed and bounded interval. Let
$(v_0,p_0,\mathring {R}_0)\in C^\infty (\overline {\Omega }_0\times I)$
be a subsolution in
$\Omega _0\times I$
. Let
$\Omega $
be a neighborhood of
$\overline {\Omega }_0$
. There exists a subsolution
$(v,p,\mathring {R})\in C^\infty (\mathbb {R}^n\times I)$
that extends
$(v_0,p_0,\mathring {R}_0)$
and such that
$\operatorname {\mathrm {supp}} (v,p,\mathring {R})(\cdot ,t)\subset \Omega $
for all
$t\in I$
if and only if for each connected component
$\Sigma $
of
$\partial \Omega _0$
and all times
$t\in I$
we have

These conditions are automatically satisfied if
$\partial \Omega _0$
is connected.
Regarding the convex integration scheme, we start off with the strategy from [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7], which we implement in the context of solutions with compact support. The main issue we have to address is that, as we want the resulting solution to coincide with
$v_0$
in
$\Omega \times [0,T]$
, we must ensure that the scheme does not modify the subsolution in that region.
The result of our construction is:
Theorem 1.7. Fix some
$T>0$
and let
$\Omega \subset \mathbb {R}^3$
be a bounded open set with smooth boundary and with a finite number of connected components. Let
$(v_0,p_0,\mathring {R}_0)\in C^\infty (\mathbb {R}^3\times [0,T])$
be a subsolution of the Euler equations. Suppose that
$\operatorname {\mathrm {supp}}\mathring {R}_0\subset \overline {\Omega }\times [0,T]$
. Let
$0<\beta <1/3$
and let
$e\in C^\infty ([0,T],(0,\infty ))$
be an energy profile such that

for all
$0\leq t\leq T$
. Then, there exists a weak solution of the Euler equations,
$v\in C_c^\beta (\mathbb {R}^3\times [0,\infty ))$
, such that
$v=v_0$
in
$(\mathbb {R}^3\backslash \Omega )\times [0,T]$
and

for all
$t\in [0,T]$
.
1.4 Organization of the paper
In Section 2 we develop a set of tools to handle the construction, extension, and gluing of subsolutions that will be used throughout the paper; in particular we prove Theorem 1.6. In Section 3 we present the iterative process used to prove Theorem 1.7, which is carried out in a number of stages. The technical details of each stage of the construction are discussed in detail in Sections 4 to 7. The very short Section 8 shows how to pass from Theorems 1.6 and 1.7 to Theorem 1.1 and Remark 1.2. The modification of this scheme to account for the lack of uniform-in-time bounds is carried out in Section 9. The applications concerning vortex sheets and blowup, cf. Theorem 1.4 and Theorem 1.5, are discussed in Sections 10 and 11, respectively. The paper concludes with two appendices, one about Hölder norms and another with some auxiliary estimates.
2 Extension of subsolutions
The goal of this section is to prove the extension theorem of smooth subsolutions stated in Theorem 1.6. This is a key ingredient to prove our main theorem on the extension of weak solutions of the Euler equations. In Subsection 2.1 we sketch the strategy to prove Theorem 1.6. Some instrumental tools from Hodge theory are presented in Subsection 2.2, and in Subsection 2.3 we show how to construct compactly supported solutions to the key divergence equation. Finally, the proof of Theorem 1.6 is completed in Subsection 2.4.
In addition to constructing the desired solutions, we must estimate their derivatives. We refer to Appendix A for the definition of the Hölder norms used. Specifically, we warn the reader that when dealing with time-dependent functions, we consider the supremum in time of the corresponding norms in space. Nevertheless, obtaining bounds on the derivatives of the solutions to certain differential equations is not enough for our construction. As we will see, we also need to control their
$C^0$
norm. This can be achieved if we work with Besov spaces, which are defined in Appendix A.
Throughout this section, we denote the space of
$n\times n$
symmetric matrices as
$\mathcal {S}^n$
and the space of
$n\times n$
skew-symmetric matrices as
$\mathcal {A}^n$
. Unless otherwise stated, the dimension is
$n\geq 2$
. We define the divergence of a matrix
$M\in C^\infty (\mathbb {R}^n\times \mathbb {R},\mathbb {R}^{n\times n})$
as the vector field whose coordinates are given by

where the derivatives are taken only with respect to the spatial variables. More generally, partial derivatives with Latin subscripts denote partial derivatives in the spatial coordinates, whereas temporal partial derivatives are always denoted by
$\partial _t$
.
We will repeatedly use Einstein’s summation convention: when an index appears twice in an expression, it is implicitly summed over its range. Indices that appear only once in an expression are free indices and are not summed over.
Let us now recall the definition of subsolution of the Euler equations:
Definition 2.1. Let
$V\subset \mathbb {R}^n\times \mathbb {R}$
be an open set. We will say that a triplet
$(v,p,\mathring {R})\in C^\infty (V, \mathbb {R}^n\times \mathbb {R}\times \mathcal {S}^n)$
is a subsolution of the Euler equations if

The symmetric matrix
$\mathring {R}$
is known as the Reynolds stress and it measures the deviation from being a solution of the Euler equations. It is customary to also impose that

All along this article,
$\circ $
above a symmetric matrix will indicate that it is trace-free.
Finally, let us fix some notation that will be used all along this section. We introduce the following norms in the space of
$n\times n$
matrices:

Unless otherwise stated, we will always use the operator norm
$\left \vert \cdot \right \vert $
. However, in some parts of the article we will exploit the elementary property that
$\left \vert \cdot \right \vert {}_2^2$
depends smoothly on the matrix entries. Note that
$\left \vert M\right \vert {}_2^2=\text {tr}(M^tM)$
, which is invariant under orthogonal transformations. Hence, in the case of a symmetric matrix
$S\in \mathcal {S}^n$
, we have

which can be easily deduced by using an orthonormal basis of eigenvectors.
2.1 General strategy
Our techniques for extending subsolutions and performing convex integration in the nonperiodic setting rely on obtaining compactly supported solutions to the (matrix) divergence equation when the source term is compactly supported. Let us illustrate the key ideas behind our method with the following toy problem:
Problem 2.2. Given
$\rho \in C^\infty _c(\mathbb {R}^3)$
such that
$\int \rho =0$
, find
$v\in C^\infty _c(\mathbb {R}^3,\mathbb {R}^3)$
such that
$\operatorname {\mathrm {div}} v=\rho $
.
It is easy to see that
$v_0=\nabla \,\Delta ^{-1}\rho $
solves our equation, but in general it is not compactly supported. To fix this, let B be a ball containing the support of
$\rho $
, so that
$v_0$
is divergence-free outside B. In addition, it follows from the divergence theorem that

This ensures that in
$\mathbb {R}^3\backslash B$
the divergence-free field
$v_0$
can be written as
$v_0=\operatorname {\mathrm {curl}} w_0$
for some smooth field
$w_0$
. We extend
$w_0$
to a smooth field
$w\in C^\infty (\mathbb {R}^3,\mathbb {R}^3)$
and we define

Since w extends
$w_0$
, we see that v vanishes outside B. Furthermore,
$\operatorname {\mathrm {div}} v=\rho $
because
$\operatorname {\mathrm {div}}\operatorname {\mathrm {curl}}\equiv 0$
. Therefore, v is the sought field, which is clearly not unique.
Our approach to solving the divergence equation in the matrix case is the same: the potential-theoretic solution of the equation is not compactly supported. However, far from the support of the source our matrix will be the image of certain differential operator
$\mathscr {L}$
applied to a smooth potential, which is in the kernel of the divergence. We will extend the potential to the whole space and then subtract it from the potential-theoretic solution, obtaining a compactly supported solution.
Just like in the vector case, we will have to impose certain integrability conditions on the source term for this to be possible. As we will see, these conditions are related to the classical conservation laws of linear and angular momentum in the Euler equations.
A totally different method to construct compactly supported solutions to the divergence equation in the (symmetric) matrix case was developed by Isett and Oh in [Reference Isett and Oh36]. Their theorem is stated in a very different setting and adapting it to what we need would require certain work. On the other hand, it will be relatively easy to deduce our result as a consequence of our analysis of the operator
$\mathscr {L}$
introduced below, which is necessary for our result on the extension of subsolutions. Hence, we have preferred to take this path, which we believe is simpler (partly because it has a nice interpretation in terms of the elementary operations of vector calculus).
2.2 Basic tools
The tools that we will need come from the Hodge decomposition theorem for manifolds with boundary. A good reference for this topic is [Reference Schwarz45]. Nevertheless, we do not need the full generality of these results, as we will work in bounded domains of
$\mathbb {R}^n$
. Let us summarize the notation and main definitions that we will need.
Let
$\Omega \subset \mathbb {R}^n$
be a bounded domain with smooth boundary. We denote by
$\Lambda ^k$
the vector space of skew-symmetric k-forms over
$\mathbb {R}^n$
, for
$0\leq k\leq n$
. In this setting, differential k-forms are maps
$\omega \in C^\infty (\overline {\Omega },\Lambda ^k)$
. They form vector spaces in which we have two differential operators: the exterior derivative
$d:C^\infty (\overline {\Omega },\Lambda ^k)\to C^\infty (\overline {\Omega },\Lambda ^{k+1})$
and the codifferential
$\delta :C^\infty (\overline {\Omega },\Lambda ^{k+1})\to C^\infty (\overline {\Omega },\Lambda ^k)$
. The Euclidean product induces a natural scalar product
$(\cdot ,\cdot )$
in
$C^\infty (\overline {\Omega },\Lambda ^k)$
. The tangential part of a differential form is
, where
$j:\partial \Omega \hookrightarrow \overline {\Omega }$
is the natural inclusion, and
$j^*$
is the pushforward. We define the Dirichlet harmonic k-forms as:

Finally, to obtain quantitative estimates we will need to work in Hölder spaces. We refer to Appendix A for the definition of these norms. We also recommend to take a look at the appendix to check our convention of Hölder norms when the field is time-dependent.
With this notation, the first basic lemma that we shall use is:
Lemma 2.3. Let
$\Omega \subset \mathbb {R}^n$
be a bounded domain with smooth boundary and let
$\rho \in C^\infty (\overline {\Omega },\Lambda ^k)$
. The boundary value problem

is solvable if and only if

In that case, the solution is unique and we have

for any
$N\geq 0$
,
$\alpha \in (0,1)$
and certain constants
$C\equiv C(N,\alpha ,\Omega )$
.
For a proof, see [Reference Csató, Dacorogna and Kneuss20, Theorem 7.2] and [Reference Schwarz45, Corollary 3.2.4]. In [Reference Schwarz45, Theorem 3.2.5] we can find the general case of a Riemannian manifold with boundary, but it does not include estimates for Hölder norms, only for Sobolev norms. We recall that, as usual,
$\star $
is the Hodge star operator acting on differential forms and an
$(n-k)$
-cycle is an
$(n-k)$
-chain (in the sense of algebraic topology) whose boundary is zero.
We are mainly interested in the problem
$\delta \omega =\rho $
, but we have to add the other conditions to select a single solution. This allows us to obtain a time-dependent version of Lemma 2.3:
Lemma 2.4. Let
$\Omega \subset \mathbb {R}^n$
be a bounded domain with smooth boundary and let
$I\subset \mathbb {R}$
be a closed and bounded interval. Let
$\rho \in C^\infty (\overline {\Omega }\times I,\Lambda ^k)$
. There exists a differential form
$\omega \in C^\infty (\overline {\Omega }\times I,\Lambda ^{k+1})$
solving the boundary value problem
$(P)$
at each
$t\in I$
if and only if

In that case, the solution is unique and we have

for any
$N\geq 0$
,
$\alpha \in (0,1)$
and certain constants
$C\equiv C(N,\alpha ,\Omega )$
.
Proof. Given a time-dependent differential form, we denote by a subscript the differential form at a given time. By Lemma 2.3, the necessity of the conditions is clear. To prove that they are also sufficient, let us suppose that
$\rho _t$
satisfies the conditions of Lemma 2.3 at all times
$t\in I$
. Hence, applying Lemma 2.3 at each time, we see that there exists a time-dependent
$(k+1)$
-form
$\omega $
solving
$(P)$
at each
$t\in I$
. The question is whether
$\omega $
depends smoothly on t.
Since
$\partial _t \rho $
also satisfies the hypotheses of Lemma 2.3 at all times
$t\in I$
, there exists a
$(k+1)$
-form
$\widetilde {\omega }$
solving
$(P)$
with data
$\partial _t \rho $
. For a fixed
$t_0\in I$
and
$h\neq 0$
small we see that
$h^{-1}(\omega _{t_0+h}-\omega _{t_0})-\widetilde {\omega }_{t_0}$
is the unique solution of
$(P)$
with data
$h^{-1}(\rho _{t_0+h}-\rho _{t_0})-(\partial _t \rho )_{t_0}$
. Therefore,

We deduce that
$\widetilde {\omega }$
is the partial derivative with respect to time of
$\omega $
. Iterating this argument, we conclude that
$\omega $
depends smoothly on time. The claimed estimates are easily obtained by taking the supremum on
$t\in I$
in the estimates of Lemma 2.3.
Remark 2.5. If
$\delta \rho =0$
, the integral of
$\star \rho $
on an
$(n-k)$
-cycle depends only on the homology class of the cycle. Indeed, if C and
$C'$
are two
$(n-k)$
-cycles in
$\overline {\Omega }$
that are the boundary of an
$(n-k+1)$
-chain, by Stokes’ theorem we have

The machinery of differential geometry is quite powerful, but we are interested in the simpler setting of bounded domains
$\Omega \subset \mathbb {R}^n$
. Taking advantage of the canonical basis of Euclidean space, we may forget about differential forms and work with simpler objects. Indeed, there is a natural correspondence between 1-forms and vector fields and between 2-forms and skew-symmetric matrices
$\mathcal {A}^n$
:

Here we have used that
$dx_i\wedge dx_j=-dx_j\wedge dx_i$
. Using the canonical base of
$1$
-forms, the action of the codifferential can be summarized as

where f is any smooth function and
$1\leq i<j\leq n$
. One can then check that the following diagram commutes:

This allows us to write everything in terms of matrices and to simplify the notation. Using this correspondence and Remark 2.5, we may formulate a particular case of Lemma 2.4 as follows:
Lemma 2.6. Let
$\Omega \subset \mathbb {R}^n$
be a bounded domain with smooth boundary and let
$I\subset \mathbb {R}$
be a closed and bounded interval. Let
$v\in C^\infty (\overline {\Omega }\times I,\mathbb {R}^n)$
. The following are equivalent:
-
1. there exists
$A\in C^\infty (\overline {\Omega }\times I,\mathcal {A}^n)$ such that
$\operatorname {\mathrm {div}} A=v$ ,
-
2.
$\operatorname {\mathrm {div}} v=0$ and
$\int _\Sigma v\cdot \nu =0$ for any connected component
$\Sigma $ of
$\partial \Omega $ and any fixed
$t\in I$ ,
-
3.
$\int _{C_{n-1}} v\cdot \nu =0$ for any
$(n-1)$ -cycle
$C_{n-1}$ and any
$t\in I$ .
In that case,
$A\in C^\infty (\overline {\Omega }\times I,\mathcal {A}^n)$
may be chosen so that

for any
$N\geq 0$
,
$\alpha \in (0,1)$
and certain constants
$C\equiv C(N,\alpha ,\Omega )$
.
The proof is straightforward taking into account that the codifferential
$\delta $
becomes the operator
$\operatorname {\mathrm {div}}$
and the integral on cycles becomes the flux of the corresponding vector field across a closed surface. By Remark 2.5, the integral only depends on the homology class, so we can choose to integrate on the connected component of
$\partial \Omega $
belonging to each class.
2.3 The divergence equation
After collecting some basic tools from Hodge theory in the previous subsection, we will now show how to obtain compactly supported solutions to the divergence equation. We begin by introducing some potential-theoretic solutions, which we will later modify in order to fix the support. Let us consider the following differential operator that maps smooth vector fields (with bounded derivatives) to
$C^\infty (\mathbb {R}^n,\mathcal {S}^n)$
:

Here
$\Delta ^{-1}$
refers to the potential-theoretic solution of the Poisson equation, that is, the (spatial) convolution of the source term with the fundamental solution of the Laplace equation in
$\mathbb R^n$
. We remind the reader that partial derivatives with Latin subscripts denote partial derivatives in the spatial coordinates, whereas temporal partial derivatives are always denoted by
$\partial _t$
.
A direct calculation shows that
$\operatorname {\mathrm {div}}\mathscr {R}f=f$
. We notice that
$\mathscr {R}$
is not trace-free. This is not an issue in our proofs, because our constructions with potentials do not preserve being trace-free, so we will usually absorb the trace into the pressure at the end.
Let us now derive a very useful identity. Let
$\Omega \subset \mathbb {R}^n$
be a bounded open set with smooth boundary. Let
$v\in C^\infty (\overline {\Omega },\mathbb {R}^n)$
and
$S\in C^\infty (\overline {\Omega },\mathcal {S}^n)$
. Integrating by parts, we have

where
$\nu $
is the unitary normal vector associated to the exterior orientation and the operator
$\nabla _{\text {sym}}$
is given by:

Its kernel are the so-called Killing vector fields. It is a finite-dimensional vector space that plays an important role in Riemannian geometry. It is well known (see [Reference Petersen43, page 52]) that in
$\mathbb {R}^n$
a basis of this vector space is given by

where

Next, we introduce two vector spaces and a differential operator that will be very important in our construction:
Definition 2.7. Let
$\Omega \subset \mathbb {R}^n$
be a bounded domain and let
$I\subset \mathbb {R}$
be a closed and bounded interval. We define two vector spaces:

and we consider the differential operator

This operator already appeared in the context of convex integration in the original article by De Lellis and Székelyhidi [Reference De Lellis and Székelyhidi23], who noticed that the image of the operator
$\mathscr {L}$
is contained in the space of divergence-free matrices.
For our purposes, this operator can be regarded as a matrix analog of the
$\operatorname {\mathrm {curl}}$
operator in arbitrary dimension. In order to perform the construction sketched in Subsection 2.1, the next step is to understand how to invert this operator (under the appropriate boundary conditions). The following lemma is the key to our approach to solve the divergence equation:
Lemma 2.8. Let
$\Omega \subset \mathbb {R}^n$
be a bounded domain with smooth boundary and let
$I\subset \mathbb {R}$
be a closed and bounded interval. Then
$\mathcal {G}(\overline {\Omega }\times I)$
is the image of the differential operator
$\mathscr {L}$
. Furthermore, given
$S\in \mathcal {G}(\overline {\Omega }\times I)$
, there exists
$A\in \mathcal {P}(\overline {\Omega }\times I)$
such that
$S=\mathscr {L}(A)$
and for any
$\alpha \in (0,1)$
we have

for all
$N\geq 0$
and certain constants
$C\equiv C(N,\alpha ,\Omega )$
.
Proof. First, we prove that the image of
$\mathscr {L}$
is contained in
$\mathcal {G}(\overline {\Omega }\times I)$
. Fix an arbitrary
$A\in \mathcal {P}(\overline {\Omega }\times I)$
. It is clear from the definition that
$\mathscr {L}(A)$
is symmetric. The fact that it is divergence-free follows from the skew-symmetric properties of A:

Next, we fix an arbitrary Killing vector field
$\xi $
and a connected component
$\Sigma $
of
$\partial \Omega $
. We choose a smooth cut-off function
$\varphi $
that vanishes in a neighborhood of
$\Sigma $
and it is identically 1 in a neighborhood of the other connected components of
$\partial \Omega $
. By the choice of
$\varphi $
we have

By our previous discussion,
$\mathscr {L}(A)$
and
$\mathscr {L}(\varphi A)$
are symmetric and divergence-free. In addition,
$\xi $
is a Killing vector, so
$\nabla _{\text {sym}} w=0$
. Thus, from (2.6) we deduce that the term on the left-hand side of the previous equation vanishes and so does the first term on the right-hand side. Therefore, we see that
$\int _\Sigma \xi ^t\mathscr {L}(A)\,n=0$
and, since A,
$\xi $
, and
$\Sigma $
are arbitrary, we conclude that the image of
$\mathscr {L}$
is contained in
$\mathcal {G}(\overline {\Omega }\times I)$
.
Now we will prove the other inclusion and the stated estimate. We fix an arbitrary
$S\in \mathcal {G}(\overline {\Omega }\times I)$
. By definition, when choosing the canonical basis of
$\mathbb {R}^n$
as Killing vectors, we obtain

for any
$i=1, \dots , m$
. If we fix i, we may apply Lemma 2.6 to conclude that there exists
$B^i\equiv B^i_{jl}\in C^\infty (\overline {\Omega }\times I,\mathcal {A}^n)$
such that
$\partial _l B^i_{jl}=S_{ij}$
.
Next we fix a Killing field
$\xi $
of the form
$\xi _i=R_{ik}x_k$
, where
$R\in \mathcal {A}^n$
. For
$j=1, \dots m$
we compute

Note that, since
$S\in \mathcal {G}(\overline {\Omega }\times I)$
,

Regarding the left-hand side term, we define the forms

Since
$B^i_{jl}$
is skew-symmetric in the lower indices, we see that
$\delta \eta =\omega $
. Using the properties of the Hodge star operator and the codifferential, we have
$\star \delta \eta =d\star \eta $
. These forms allow us to rewrite the integral on an
$(n-1)$
-cycle at any
$t\in I$
as:

where
$\tilde {\mu }$
is the measure induced by the standard measure in
$\mathbb {R}^n$
. We have used Stokes’ theorem and the fact that
$(n-1)$
-cycles have no boundary. We conclude that for any
$(n-1)$
-cycle

Choosing
$R=e_{i_0}\otimes e_{l_0}-e_{l_0}\otimes e_{i_0}$
, that is, choosing
$\xi $
as
$\xi _{l_0i_0}$
, we see that for any
$i,l=1, \dots m$
we have:

Applying again Lemma 2.6, we obtain
$A^{ik}_{jl}$
skew-symmetric in
$j, k$
such that

Therefore,

where we have used that B is skew-symmetric in the lower indices and the symmetry of S:
$\partial _l B^i_{jl}=S_{ij}=S_{ji}=\partial _l B^j_{il}$
. In summary, we have found
$A\in C^\infty \left (\overline {\Omega }\times I,\mathbb {R}^{n^4}\right )$
such that:
-
(i)
$A^{ik}_{jl}=-A^{ij}_{kl}$ ,
-
(ii)
$\partial _k A^{ik}_{jl}=-\partial _k A^{lk}_{ji}$ ,
-
(iii)
$\frac {1}{2}\partial _{kl} \left (A^{ik}_{jl}+A^{jk}_{il}\right )=S_{ij}$ .
We define

It is clear that
$\widetilde {A}^{ik}_{jl}$
is skew-symmetric in
$i,k$
. In addition, it is skew-symmetric in
$j,l$
by (i). Furthermore,

Hence, using (iii) we conclude

Therefore,
$\widetilde {A}\in \mathcal {P}(\overline {\Omega }\times I)$
and
$\mathscr {L}(\widetilde {A})=S$
, as we wanted. The estimates for
$\widetilde {A}$
follow from applying twice the estimates from Lemma 2.6.
Finally, we are ready to prove the main result of this subsection, which establishes the existence of compactly supported solutions to the divergence equation. In a different setting, a related class of compactly supported solutions to the divergence equation were constructed by Isett and Oh [Reference Isett and Oh36, Theorem 10.1]. Our approach is based on the operator
$\mathscr {L}$
, which will be essential for the extension of subsolutions in Lemma 2.15. We observe that the compatibility conditions (2.9) in Lemma 2.9 are precisely the conditions (202) in [Reference Isett and Oh36].
We recall that the Besov norms that we use are defined in Appendix A. We need to work with these norms because they are necessary to derive estimates for the
$C^\alpha $
norm of the resulting matrix. This will be essential in the proof of Theorem 1.7.
Lemma 2.9. Let
$\Omega \subset \mathbb {R}^n$
be a bounded domain with smooth boundary and let
$I\subset \mathbb {R}$
be a closed and bounded interval. Let
$f\in C^\infty (\mathbb {R}^n\times I,\mathbb {R}^n)$
such that
$\operatorname {\mathrm {supp}} f(\cdot ,t)\subset \Omega $
for all
$t\in I$
. Then, there exists
$S\in C^\infty (\mathbb {R}^n\times I,\mathcal {S}^n)$
such that
$\operatorname {\mathrm {div}} S=f$
and
$\operatorname {\mathrm {supp}} S(\cdot ,t)\subset \Omega $
for all
$t\in I$
if and only if

In that case, we may choose S so that for all
$N\geq 0$
and any
$\alpha \in (0,1)$
we have

for certain constants
$C=C(\Omega ,N,\alpha )$
.
Proof. First of all, we show that the integrability condition (2.9) is necessary. Let us suppose that such an S exists. We fix a ball
$B\supset \overline {\Omega }$
and use the identity (2.6) to obtain

Let us show that condition (2.9) is also sufficient. The field
$S_0\in C^\infty (\mathbb {R}^n\times I,\mathcal {S}^n)$
given by
solves the equation
$\operatorname {\mathrm {div}} S_0=f$
, where
$\mathscr {R}$
was defined in (2.5). It is easy to check that it satisfies the estimates

for certain constants
$C=C(N,\alpha )$
because
$\mathscr {R}$
is an operator of order
$-1$
. However, it is not compactly supported, in general. We must modify it far from the support of f.
We begin by studying the boundary conditions. Let
$\Sigma _i$
be a connected component of
$\partial \Omega $
and let
$U_i$
be the domain bounded by it. We claim that

Indeed, since
$\Omega $
is a bounded domain,
$U_i$
must be either the complement of the unbounded connected component of
$\mathbb {R}^3\backslash \overline {\Omega }$
or one of the bounded connected components of
$\mathbb {R}^3\backslash \overline {\Omega }$
(if there are any). In the first case, (2.11) follows from the integrability condition (2.9) because
$f(\cdot ,t)\subset \Omega \subset U_i$
. In the second case, (2.11) is trivial because
$f(\cdot ,t)$
vanishes on
$U_i\subset \mathbb {R}^3\backslash \overline {\Omega }$
. Thus, applying the identity (2.6) to each
$U_i$
we obtain

Next, note that for sufficiently small
$r>0$
the boundary of the open set

has twice as many connected components as
$\partial \Omega $
. Furthermore, the boundary of each connected component
$G_i$
of G consists of exactly two hypersurfaces, which we denote as
$\Sigma _i$
and
$\Sigma ^{\prime }_i$
, and we have
$\Sigma _i\subset \partial \Omega $
. By further reducing
$r>0$
, we may assume that
$f(\cdot ,t)$
vanishes on G at all times
$t\in I$
. Then, it follows from (2.12) and the identity (2.6) that

for any
$\xi \in \ker \nabla _{\text {sym}}$
. Next, we fix a ball
$B\supset \overline {\Omega }$
and we consider the domain

We see that

Again, it follows from (2.6) and the integrability condition (2.9) that

We conclude that
$S_0$
is divergence-free on U and in each connected component
$\Sigma $
of
$\partial U$
we have

that is,
$S_0\in \mathcal {G}(\overline {U}\times I)$
. By Lemma 2.8 there exists
$A_0\in \mathcal {P}(\overline {U}\times I)$
such that
$S_0(x,t)=\mathscr {L}(A_0)(x,t)$
for all
$x\in \overline {G}$
and
$t\in I$
. Furthermore, for any
$N\geq 0$
and
$\alpha \in (0,1)$
we have

The constants depend on U, which depends not only on the geometry of
$\Omega $
but also on the minimum distance between the support of
$f(\cdot ,t)$
and
$\partial \Omega $
through the parameter r. However, since U tends to
$B\backslash \overline {\Omega }$
as
$r\to 0$
, the constants remain uniformly bounded, so they ultimately depend only on
$\Omega $
. For this, smoothness of
$\partial \Omega $
is essential, as it allows us to choose parametrizations of
$\overline {U}$
converging to parametrizations of
$\overline {B}\backslash \Omega $
in any Hölder norm as
$r\to 0$
.
Applying Theorem B.3 and antisymmetrizing, we see that there exists a map
$A\in C^\infty (\mathbb {R}^n\times I,\mathbb {R}^{n^4})$
such that
$A^{ik}_{jl}=-A^{ki}_{jl},\, A^{ik}_{jl}=-A^{ik}_{lj}$
that extends
$A_0$
outside
$\overline {U}\times I$
. Furthermore, for any
$N\geq 0$
and
$\alpha \in (0,1)$
, we have

Again, since U tends to
$B\backslash \overline {\Omega }$
as
$r\to 0$
in a suitable manner, the constants ultimately depend only on
$\Omega $
, N, and
$\alpha $
, since they will be uniformly bounded on
$r\in (0,1)$
.
Finally, for
$x\in B$
and
$t\in I$
we define

Since the image of
$\mathscr {L}$
is contained in the kernel of the divergence, we see that
$\operatorname {\mathrm {div}} S=f$
. By construction A extends
$A_0$
, so
$\mathscr {L}(A)=\mathscr {L}(A_0)=S_0$
on
$\overline {U}\times I$
. Therefore,
$S(\cdot ,t)$
vanishes in U, so we may extend it by 0 to
$\mathbb {R}^n\times I$
.
In conclusion, we have constructed
$S\in C^\infty (\mathbb {R}^n\times I,\mathcal {S}^n)$
such that
$\operatorname {\mathrm {div}} S=f$
and
$\operatorname {\mathrm {supp}} S(\cdot ,t)\subset \Omega $
for all
$t\in I$
. Furthermore, the desired estimate follows from (2.10) and (2.13) because
$\mathscr {L}$
is a second-order differential operator.
2.4 Subsolutions and proof of Theorem 1.6
In this subsection we use Lemma 2.8 and Lemma 2.9 to glue and extend subsolutions, which will yield the proof of Theorem 1.6. It should be apparent by now that controlling the
$L^2$
-product with the Killing fields is very important in these constructions. It is not difficult to construct
$f\in C^\infty _c(\mathbb {R}^n,\mathbb {R}^n)$
with the desired
$L^2$
-product with the Killing fields. However, when working with subsolutions we will also need that f be divergence-free. In addition, in our constructions we will work in domains of a certain form. Our approach is based on the following:
Lemma 2.10. Let
$\Omega \subset \mathbb {R}^n$
be a bounded domain with smooth boundary and let
$I\subset \mathbb {R}$
be a closed and bounded interval. Let
$r>0$
and let
$L_{ij}\in C^\infty (I)$
for
$1\leq i<j\leq n$
. There exists a divergence-free field
$w\in C^\infty (\mathbb {R}^n\times I,\mathbb {R}^n)$
such that the support of
$w(\cdot ,t)$
is contained in
$\{x\in \mathbb {R}^n:0<\operatorname {\mathrm {dist}}(x,\Omega )<r\}$
and

for all
$t\in I$
and
$1\leq i<j\leq n$
, where
$\xi _{ij}$
is given by (2.8). Furthermore, for any
$N\geq 0$
we have

Proof. We will construct our field as
$w=\operatorname {\mathrm {div}} A$
for some
$A\in C^\infty _c(\mathbb {R}^n\times I,\mathcal {A}^n)$
that we will choose later. Since A is compactly supported, it follows from the divergence theorem that
$\int a\cdot w=0$
for any
$a\in \mathbb {R}^n$
. Furthermore, for any
$1\leq i<j\leq n$
we have

because
$\partial _l(\xi _{ij})_k=\delta _{il}\delta _{jk}-\delta _{jl}\delta _{ik}$
. Here we have denoted by
$(\xi _{ij})_k$
the k-th component of the vector
$\xi _{ij}$
and we have used Einstein’s summation convention when summing over k and l. By Lemma B.1 we may choose a nonnegative cutoff function
$\varphi \in C^\infty _c(\Omega +B(0,r))$
that is identically 1 in a neighborhood of
$\Omega $
and such that

We define

Since
$\varphi $
is constant in a neighborhood of
$\Omega $
and its support is contained in
$\Omega +B(0,r)$
, we see that the support of
$w(\cdot ,t)$
is contained in
$\{x\in \mathbb {R}^n:0<\operatorname {\mathrm {dist}}(x,\Omega )<r\}$
. By construction

so it follows from Equation (2.14) that
$\int \xi _{ij}\cdot w=L_{ij}$
. Finally, the claimed estimates follow at once from the bounds for
$\varphi $
and the fact that
$\int \varphi \geq \left \vert \Omega \right \vert $
.
Now we have all the ingredients that we need to glue subsolutions in space. The following lemma is the key tool in this section. It will be used not only in the proof of Theorem 1.6, but also in the convex integration scheme. We use skew-symmetric matrices instead of potential vectors because the lemma is stated in any dimension
$n\geq 2$
.
Lemma 2.11. Let
$T>0$
and let
$\Omega _1\Subset \Omega _2\subset \mathbb {R}^n$
be bounded domains with smooth boundary. Let
$(v_i,p_i,\mathring {R}_i)\in C^\infty (\Omega _2\times [0,T])$
be subsolutions for
$i=1,2$
. Let
$r>0$
be sufficiently small. There exists a subsolution
$(v,p,\mathring {R})\in C^\infty (\Omega _2\times [0,T])$
such that

if and only if for each connected component
$\Sigma $
of
$\partial \Omega _1$
, and all times
$t\in [0,T]$
, we have


Suppose that, in addition, we have
$v_1=\operatorname {\mathrm {div}} A_1$
and
$v_2=\operatorname {\mathrm {div}} A_2$
for some potentials
$A_i\in C^\infty (\Omega _2\times I,\mathcal {A}^n)$
. Then, there exists
$A\in C^\infty (\Omega _2\times I,\mathcal {A}^n)$
such that
$v=\operatorname {\mathrm {div}} A$
and
$A(x,t)=A_2(x,t)$
if
$\operatorname {\mathrm {dist}}(x,\Omega _1)\geq r$
.
Remark 2.12. The compatibility conditions (2.16) and (2.17) are automatically satisfied if
$\partial \Omega _1$
is connected or
$\Omega _2=\mathbb {R}^n$
. This will be explained in the proof of the lemma.
Remark 2.13. The subsolution
$(v_1,p_1,\mathring {R}_1)$
need not be defined in all of
$\Omega _2$
and the subsolution
$(v_2,p_2,\mathring {R}_2)$
need not be defined in
$\Omega _1$
. We have assumed this to simplify slightly the statement of the lemma.
Proof. First of all, note that a subsolution
$(v_0,p_0,\mathring {R}_0)$
in a bounded domain G with smooth boundary satisfies


where we have used the divergence theorem, identity (2.6) and the fact that
$\operatorname {\mathrm {div}}[(a\cdot x)\partial _tv_0]=a\cdot \partial _t v_0$
because
$\partial _t v_0$
is divergence-free.
From these equations we readily deduce that the compatibility conditions (2.16) and (2.17) are automatically satisfied if
$\partial \Omega _1$
is connected, as both integrals vanish for each field. In the case
$\Omega _2=\mathbb {R}^n$
, we apply Equations (2.18) and (2.19) to the domain bounded by each connected component of
$\partial \Omega _1$
. We conclude that both integrals vanish for each field in each connected component of
$\partial \Omega _1$
.
Next, we check that the conditions are necessary; we study (2.16) because the expressions are shorter, but the argument for (2.17) is exactly the same. First, if
$\partial \Omega _1$
is connected, it readily follows from (2.18) that (2.16) must be satisfied. Hence, we focus on bounded domains
$\Omega _1$
whose boundary is not connected. In that case,
$\mathbb {R}^n \backslash \Omega _1$
must have at least one bounded connected component. Given a bounded connected component of
$\mathbb {R}^n\backslash \Omega _1$
, we define G to be its intersection with
$\Omega _2$
. Then,
$\partial G$
is composed of a connected component
$\Sigma $
of
$\partial \Omega _1$
and (possibly) some connected components
$\Sigma ^{\prime }_1, \dots , \Sigma ^{\prime }_m$
of
$\partial \Omega _2$
. Since v equals
$v_1$
on
$\Sigma $
and
$v_2$
on the other connected components of
$\partial G$
, it follows from (2.18) that:

On the other hand, applying (2.18) to
$v_2$
on G, we have

which, combined with the previous equation, yields

Since this applies to any bounded connected component of
$\mathbb {R}^n\backslash \Omega _1$
, we can combine it with (2.18) with
$G=\Omega _1$
to obtain an analogous identity for the remaining connected component of
$\partial \Omega _1$
, that is, the boundary of the unbounded connected component of
$\mathbb {R}^n\backslash \Omega _1$
. We conclude (2.16).
Let us now prove that the compatibility conditions (2.16) and (2.17) are also sufficient. Let
$r>0$
be small enough so that
$\{x\in \Omega _2:\operatorname {\mathrm {dist}}(x,\Omega _1)=r\}$
is diffeomorphic to
$\partial \Omega _1$
. We define
. Then, the condition (2.16) ensures that there exists
$A_{12}\in C^\infty (\overline {U}\times [0,T],\mathcal {A}^n)$
such that
$v_2-v_1=\operatorname {\mathrm {div}} A_{12}$
in
$U\times [0,T]$
. Indeed, let
$U_i$
be a connected component of U and let
$\Sigma _i$
and
$\Sigma ^{\prime }_i$
be the connected components of
$\partial U_i$
, where
$\Sigma _i\subset \partial \Omega _1$
. Using (2.16) and the fact that
$v_2-v_1$
is divergence-free:

Hence, the flux of
$v_2-v_1$
through each connected component of U vanishes, so by Lemma 2.6 there exists
$A_{12}\in C^\infty (\overline {U}\times [0,T],\mathcal {A}^n)$
such that
$v_2-v_1=\operatorname {\mathrm {div}} A_{12}$
.
Next, using Lemma B.1 we choose a cutoff function
$\varphi \in C^\infty _c(\Omega _1+B(0,r))$
that equals 1 in a neighborhood of
$\Omega _1$
. We define

where so that
$\varphi \, v_1+(1-\varphi )v_2+w_c$
is divergence-free. The additional correction
$w_L$
is a divergence-free field supported within U that will be defined later. Its purpose is to cancel the angular momentum so that the gluing can be performed in the interior of U. After a tedious computation we obtain

where


Let and
$\rho =\widetilde {\rho }+\partial _t w_L$
. Our goal is to find
$S_2\in C^\infty (\Omega _2\times [0,T],\mathcal {S}^n)$
supported on U for all
$t\in [0,T]$
and such that
$\operatorname {\mathrm {div}} S_2=\rho $
. Thus, we may set
$R=\varphi \,\mathring {R}_1+(1-\varphi )\mathring {R}_2+S_1+S_2$
and absorb the trace into the pressure, obtaining
$\mathring {R}$
and the final pressure p. To do so, first we must check that
$\rho $
satisfies the compatibility conditions (2.9).
Note that
$\widetilde {\rho }=\operatorname {\mathrm {div}}(\varphi M)$
because
$\operatorname {\mathrm {div}} M=0$
, since
$(v_i,p_i,\mathring {R}_i)$
are subsolutions. Hence, by the divergence theorem for any
$a\in \mathbb {R}^n$
we have

Note that

Therefore, combining Equations (2.17), (2.23) and (2.24) we conclude that
$\int _U a\cdot \widetilde {\rho }=0$
for all
$a\in \mathbb {R}^n$
. Since
$\partial _t w_L$
is divergence-free and its support is contained in U, the same holds for
$\partial _t w_L$
, so
$\int _U a\cdot \rho =0$
for all
$a\in \mathbb {R}^n$
.
Next, we study the product with nonconstant Killing fields. For each pair
$1\leq i<j\leq n$
we define

where
$\xi _{ij}$
are the elements of the basis of Killing fields defined in (2.8). It will be useful later on to write the coefficients as:

where we have used the fact that Killing fields are divergence-free as well as the values of
$\varphi $
on
$\partial \Omega _1$
and
$\partial \Omega _2$
. We define

We then define the correction
$w_L$
to be the divergence-free field obtained by applying Lemma 2.10 to the domain
$\Omega _1$
with coefficients
$L_{ij}\in C^\infty ([0,T])$
. Thus, we have

We conclude that
$\int _U\xi \cdot \rho =0$
for any Killing field
$\xi $
, as we wanted. Therefore, by Lemma 2.9 there exists
$S_2\in C^\infty (\Omega _2\times [0,T],\mathcal {S}^n)$
supported on U for all
$t\in [0,T]$
and such that
$\operatorname {\mathrm {div}} S_2=\rho $
. We define the final pressure and the Reynolds stress as

It follows from Equation (2.20) that the resulting triplet
$(v,p,\mathring {R})$
is a subsolution and it satisfies (2.15) because
$S_1$
and
$S_2$
are supported in U for all
$t\in [0,T]$
.
Finally, let us consider that the velocity fields are given by
$v_i=\operatorname {\mathrm {div}} A_i$
. In that case, we may simply take
$A_{12}=A_2-A_1$
instead of constructing a suitable potential using Lemma 2.6. We see that

Inspecting Lemma 2.10 leads us to define

Hence, we have
$v=\operatorname {\mathrm {div}} A$
and we see that A equals
$A_2$
in
$\{\operatorname {\mathrm {dist}}(x,\Omega _1)\geq r\}$
because
$\varphi $
vanishes in a neighborhood of this set.
In the convex integration scheme we will need estimates of the glued subsolution. For the sake of clarity, we keep them separate in a different lemma:
Lemma 2.14. Let
$\alpha \in (0,1)$
. In the conditions of Lemma 2.11 and using the notation of its proof, the new subsolution satisfies:



In addition, if
$v_1=\operatorname {\mathrm {div}} A_1$
and
$v_2=\operatorname {\mathrm {div}} A_2$
, the potential A satisfies


The implicit constants in these inequalities depend on
$\Omega _1$
, N, and
$\alpha $
.
Proof. We begin by estimating
$w_c=A_{12}\cdot \nabla \varphi $
. Since
$\varphi $
satisfies
$\left \|\varphi \right \|_N\lesssim r^{-N}$
and it is independent of time, it is clear that

because the support of
$\nabla \varphi $
is contained in U. Regarding
$w_L$
, it follows from (2.25) that
$\left \vert l_{ij}\right \vert \lesssim \left \|M\right \|_{0; U}$
, so
$\left \vert L_{ij}\right \vert \lesssim T\left \|M\right \|_{0; U}$
. Hence, by Lemma 2.10 we have the bounds

The claimed estimates for
$v-(\varphi v_1+(1-\varphi )v_2)=w_c+w_L$
follow at once. Let us now focus on the Reynolds stress. Using the assumption
$\left \|w\right \|_0\leq \left \|v_1\right \|_0+\left \|v_2\right \|_0$
, we deduce from the definition (2.21) that

Concerning
$S_2$
, let us first estimate
$\rho $
:

Since the support of
$\rho (\cdot ,t)$
is contained in
$\{x\in \mathbb {R}^n:0<\operatorname {\mathrm {dist}}(x,\Omega _1)<r\}$
, we may apply Lemma B.4, obtaining

Hence, it follows from the estimates in Lemma 2.9 that

Since

the claimed bound follows.
Finally, let us estimate A in the case that the velocities are given by
$v_i=\operatorname {\mathrm {div}} A_2$
. By (2.27) we have

The claimed bounds then follow from the estimates derived in Lemma 2.10.
Lemma 2.11 is almost what we want, but it can be made a bit sharper. In particular, in Theorem 1.6 we do not want to assume that the subsolution
$(v_0,p_0,\mathring {R}_0)$
is defined in a neighborhood of
$\overline {\Omega }_0$
. Fortunately, it turns out that all subsolutions can be extended, at least a little bit. Our operator
$\mathscr {L}$
is essential for this:
Lemma 2.15. Let
$\Omega _0\subset \mathbb {R}^n$
be a bounded domain with smooth boundary and let
$I\subset \mathbb {R}$
be a closed and bounded interval. Let
$(v_0,p_0,\mathring {R}_0)\in C^\infty (\overline {\Omega }_0\times I)$
be a subsolution in
$\Omega _0\times I$
. Let
$\Omega $
be a sufficiently small open neighborhood of
$\overline {\Omega }_0$
. Then, there exists a subsolution
$(v,p,\mathring {R})\in C^\infty (\Omega \times I)$
that extends
$(v_0,p_0,\mathring {R}_0)$
.
Remark 2.16. The open neighborhood
$\Omega $
need not be very small. It only needs to be bounded and such that each connected component of
$\mathbb {R}^n\backslash \overline {\Omega }_0$
has nonempty intersection with
$\mathbb {R}^n\backslash \overline {\Omega }$
.
Proof. We begin by constructing the velocity field v. We choose
$\rho \in C^\infty _c(\mathbb {R}^n\times I,\mathbb {R})$
such that
$\operatorname {\mathrm {supp}} \rho (\cdot ,t)$
is contained in
$\mathbb {R}^n\backslash \overline {\Omega }$
for all
$t\in I$
and such that

for each bounded connected component G of
$\mathbb {R}^n\backslash \overline {\Omega }_0$
, whose boundary we have oriented with the outer normal with respect to G. This can be done if
$\Omega $
is a sufficiently small neighborhood of
$\overline {\Omega }_0$
so that the intersection of G with
$\mathbb {R}^n\backslash \overline {\Omega }$
is nonempty.
Next, let so that
$\widetilde {v}\in C^\infty (\mathbb {R}^n\times I,\mathbb {R}^n)$
and
$\operatorname {\mathrm {div}} \widetilde {v}=\rho $
for all
$t\in I$
. By the divergence theorem we have

in each bounded connected component G of
$\mathbb {R}^n\backslash \overline {\Omega }_0$
and for all
$t\in I_0$
. In addition,
$\widetilde {v}-v_0$
is divergence-free in
$\Omega _0\times I$
because
$\rho $
vanishes in this set by construction. In particular, by the divergence theorem we have
$\int _{\partial \Omega _0}(\widetilde {v}-v_0)\cdot \nu =0$
, from which we conclude

for all connected components
$\Sigma $
of
$\partial \Omega _0$
. Therefore, by Lemma 2.6 there exists
$A\in C^\infty (\overline {\Omega }_0\times I,\mathcal {A}^n)$
such that
$\operatorname {\mathrm {div}} A=\widetilde {v}-v_0$
in
$\Omega _0\times I$
. We choose a smooth extension
$\widetilde {A}\in C^\infty (\mathbb {R}^n\times I,\mathbb {R}^{n\times n})$
and then we take the skew-symmetric part, so that
$\widetilde {A}\in C^\infty (\mathbb {R}^n\times \mathbb {R},\mathcal {A}^n)$
. We define:

Since the support of
$\rho (\cdot ,t)$
is contained in
$\mathbb {R}^n\backslash \overline {\Omega }$
and the second term is divergence-free, v is divergence-free in
$\Omega \times I$
. In addition, our choice of
$\widetilde {A}$
ensures that the restriction of v to
$\overline {\Omega }_0\times I$
is
$v_0$
, as we wanted.
Next, we will extend in a similar manner. First, we choose
$f\in C^\infty _c(\mathbb {R}^n\times I,\mathbb {R}^n)$
such that
$\operatorname {\mathrm {supp}} f(\cdot ,t)$
is contained in
$\mathbb {R}^n\backslash \overline {\Omega }$
for all
$t\in I$
and such that

for all bounded connected components
$G_i$
of
$\mathbb {R}^n\backslash \overline {\Omega }_0$
, whose boundary we have oriented with the outer normal with respect to
$G_i$
. To find such an f, we fix a nonnegative radial function
$\psi \in C^\infty _c(B(0,1))$
and a ball
$\overline {B}(x_i,r_i)\subset G_i$
. Due to symmetry, for any two elements
$w_j\neq w_k$
of the basis
$\mathcal {B}$
defined in (2.7) we have

Therefore, it suffices to choose

for the appropriate coefficients
$c_{ij}$
. Then, noticing that
$\partial _t v$
is a smooth vector field with bounded derivatives, we define
so that
$\widetilde {S}\in C^\infty (\mathbb {R}^n\times I,\mathcal {S}^n)$
, and

Since v extends
$v_0$
and
$(v_0,p_0,\mathring {R}_0)$
is a subsolution in
$\Omega _0\times I$
, we have

because f vanishes in that set. In addition, using (2.6) it follows from (2.34) and (2.33) that

for all bounded connected components G of
$\mathbb {R}^m\backslash \overline {\Omega }_0$
. Due to (2.6) and (2.35), this integral also vanishes for the remaining connected component of
$\partial \Omega _0$
, that is, the connected component that separates
$\Omega _0$
and the unbounded connected component of
$\mathbb {R}^n\backslash \Omega _0$
.
We conclude that
$\widetilde {S}-S_0$
is in
$\mathcal {G}(\overline {\Omega }_0\times I)$
. Therefore, by Lemma 2.8 there exists
$\widetilde {E}\in \mathcal {P}(\overline {\Omega }_0\times I)$
such that
$\mathscr {L}(\widetilde {E})=\widetilde {S}-S_0$
in
$\Omega _0\times I$
. We choose a smooth extension
$E\in C^\infty (\mathbb {R}^n\times I,\mathbb {R}^{n^4})$
and then we make the appropriate antisymmetrization. We define

By construction of E, we see that S extends
$S_0$
. Additionally, we have

because the image of
$\mathscr {L}$
is contained in the kernel of the divergence. We define

Since
$f(\cdot ,t)$
vanishes in
$\Omega $
for all
$t\in I$
, we conclude that
$(v,p,\mathring {R})$
is a subsolution in
$\Omega \times I$
that extends
$(v_0,p_0,\mathring {R}_0)$
.
Combining Lemma 2.11 and Lemma 2.15 we can finally prove Theorem 1.6:
Proof of Theorem 1.6
Working with each connected component of
$\Omega _0$
, we may assume that both
$\Omega _0$
and
$\Omega $
are connected (i.e., domains). Then, reducing
$\Omega $
if necessary, by Lemma 2.15 we may assume that
$(v_0,p_0,\mathring {R}_0)$
is a subsolution in
$\Omega \times I$
. The result then follows by applying Lemma 2.11 with the domains
$\Omega _0, \Omega $
and subsolutions
$(v_0,p_0,\mathring {R}_0)$
and
$(0,0,0)$
.
3 Proof of Theorem 1.7
The construction of a weak solution to the Euler equations stated in Theorem 1.7 consists in an iterative argument which is presented in Subsection 3.1, cf. Proposition 3.2. This proposition together with Lemma 3.3 allow us to prove the theorem in Subsection 3.2. We want to remark that most of this article, that is, Sections 4 to 7, is devoted to prove Proposition 3.2, which is the key result for our convex integration scheme.
Section 3.1 and Section 3.2 follow the general outline of Sections 2.1 and 2.2 in [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7] but with two important differences: here the initial subsolution will be nontrivial, that is, different from
$(0,0,0)$
, and the perturbations will be supported in a subset of
$\Omega $
instead of the whole space. Regarding the first issue, we use Lemma 3.3 to help start the iterative process. Concerning the second point, we introduce suitable sets related to the distance to
$\partial \Omega $
and the size of
$\mathring {R}_0$
. The perturbations will be localized to these sets, which is summarized in an additional inductive hypothesis, (3.13).
3.1 The iterative process
Let us assume all along this subsection that the subsolution
$(v_0,p_0)(\cdot ,t)$
is compactly supported for each time
$t\in [0,T]$
. We will construct the desired weak solution of the Euler equations as the limit of a sequence of subsolutions, that is, at a given step
$q\geq 0$
we have
$(v_q,p_q,\mathring {R}_q)\in C^\infty (\mathbb {R}^3\times [0,T])$
solving the Euler-Reynolds system:

to which we add the constraint that

The matrix
$\mathring {R}_q$
measures the deviation from being a solution of the Euler equations. The goal of the process is to make
$\mathring {R}_q$
vanish at the limit
$q\to +\infty $
, so that the limit field is a weak solution of the Euler equations.
Assume we are given the initial subsolution
$(v_0,p_0,\mathring {R}_0)\in C^\infty (\mathbb {R}^3\times [0,T])$
. Let us then show how to construct the rest of the terms iteratively. To construct the subsolution at step q from the one in step
$q-1$
, we will add an oscillatory perturbation with frequency
$\lambda _q$
. Meanwhile, the size of the Reynolds stress will be measured by an amplitude
$\delta _q$
. These parameters are given by


where
$\lceil x\rceil $
denotes the ceiling, that is, the smallest integer
$n\geq x$
. The parameters
$a,b>1$
are very large and very close to 1, respectively. They will be chosen depending on the exponent
$0<\beta <1/3$
that appears in Theorem 1.7, on
$\Omega $
and on the initial subsolution. We introduce another parameter
$\alpha>0$
that will be very small. The necessary size of all of the parameters will be discovered in the proof.
Throughout the process we will also try to achieve a given energy profile
$e\in C^\infty ([0,T])$
, which must satisfy the inequality (1.3). We will also assume

We will see that this can be assumed without losing generality.
Unlike the construction on the torus in [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7], it is essential that we only perturb the field in the region where the Reynolds stress is nonzero. Hence, we have to pay special attention to the support of the fields.
Since the map
$(v_0,p_0,\mathring {R}_0)(\cdot ,t)$
is assumed to be compactly supported at each time
$t\in [0,T]$
, we shall see that with a suitable rescaling we may assume that its support and
$\Omega $
are contained in
$(0,1)^3$
. This is useful because sometimes it will be convenient to consider that we are working with periodic boundary conditions (that is, in
$\mathbb {T}^3$
) to reuse the results in [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7]. On the other hand,
$\mathring {R}_0(\cdot ,t)$
is supported in a potentially smaller domain
$\overline {\Omega }$
. In our construction we must ensure that we do not perturb the subsolution outside of this set.
It will be convenient to do an additional rescaling in our problem. In the rescaled problem the initial subsolution will depend on a, but we assume that nevertheless there exists a sequence
$\{y_N\}_{N=0}^\infty $
independent of the parameters such that



Since the initial Reynolds stress
$\mathring {R}_0$
and its derivatives vanish at
$\partial \Omega \times [0,T]$
, for any
$k\in \mathbb {N}$
there exists a constant
$C_k$
such that for any
$x\in \Omega $
we have

The constants
$C_k$
are independent of a by (3.8). We define

Hence, we have

At step q the perturbation will be localized in a central region

so that
$(v_q,p_q,\mathring {R}_q)$
equals the initial subsolution in
$(\mathbb {R}^3\backslash A_q)\times [0,T]$
. Note that
$d_q\to 0$
as
$q\to \infty $
because so does
$\delta _{q+2}$
. Therefore, in the limit the perturbation covers all of the region where
$\mathring {R}_0$
is nonzero. However, the velocity is not modified outside this set.
As we have mentioned, the error introduced in the gluing step of [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7] is spread throughout the whole space. To avoid this, we introduce an additional gluing in space, which we will explain in more detail in the following sections.
The complete list of inductive estimates is the following:





where M is a geometric constant that depends on
$\Omega $
and is fixed throughout the iterative process.
Remark 3.1. If
$\Omega $
has several connected components
$\Omega ^j$
, we may fix an energy profile
$e^j$
in each of them. In that case, (3.17) would have to be replaced by

Since the construction does not differ much, for simplicity we will assume that
$\Omega $
is connected.
The following proposition is the key result to prove Theorem 1.7, because it establishes the existence of the iterative scheme in the convex integration process.
Proposition 3.2. Let
$T>0$
and let
$\Omega \subset (0,1)^3\subset \mathbb {R}^3$
be an open set with smooth boundary and with a finite number of connected components. Let
$(v_0,p_0,\mathring {R}_0)\in C^\infty (\mathbb {R}^3\times [0,T])$
be a subsolution whose support is contained in
$(0,1)^3\times [0,T]$
and such that
$\operatorname {\mathrm {supp}}\mathring {R}_0\subset \overline {\Omega }\times [0,T]$
. Furthermore, assume that (3.6)
$-$
(3.8) are satisfied for some sequence of positive numbers
$\{y_N\}_{N=0}^\infty $
. There exists a constant M depending only on
$\Omega $
with the following property: Assume
$0<\beta <1/3$
and

Then there exists an
$\alpha _0$
depending on
$\beta $
and b such that for any
$0<\alpha <\alpha _0$
there is an
$a_0$
depending on
$\beta $
, b,
$\alpha $
,
$\Omega $
, and
$\{y_N\}_{N=0}^\infty $
such that for any
$a\geq a_0$
the following holds: Given a strictly positive energy profile satisfying (3.5) and a subsolution
$(v_q,p_q,\mathring {R}_q)$
satisfying (3.13)
$-$
(3.17), there exists a subsolution
$(v_{q+1},p_{q+1},\mathring {R}_{q+1})$
satisfying the same equations (3.13)
$-$
(3.17) with q replaced by
$q+1$
. Furthermore, we have the estimate

We wish to iterate this result to construct a sequence of subsolutions whose limit will be the desired weak solution. However, in order to start the process, the first term in the sequence must satisfy the inductive hypotheses (3.13)
$-$
(3.17). Since we do not assume any bounds on
$(v_0,p_0,\mathring {R}_0)$
, these hypotheses will not be satisfied by the initial subsolution, in general. Although a time dilation would almost solve the problem, we need the following lemma to fully prepare the initial subsolution:
Lemma 3.3. Let
$T>0$
and let
$\Omega \subset (0,1)^3\subset \mathbb {R}^3$
be an open set with smooth boundary and with a finite number of connected components. Let
$(v_0,p_0,\mathring {R}_0)\in C^\infty (\mathbb {R}^3\times [0,T])$
be a subsolution whose support is contained in
$(0,1)^3\times [0,T]$
and such that
$\operatorname {\mathrm {supp}}\mathring {R}_0\subset \overline {\Omega }\times [0,T]$
. Let
$\lambda>0$
be a sufficiently large constant. There exists a subsolution
$(v,p,\mathring {R})\in C^\infty (\mathbb {R}^3\times [0,T])$
such that for any
$N\geq 0$
we have

where the implicit constants are independent of
$\lambda $
. In addition, the energy satisfies

Furthermore,
$(v,p,\mathring {R})$
equals the initial subsolution outside the set

While necessary, this result is nothing new and one could easily obtain it by combining [Reference De Lellis and Székelyhidi24] and Lemma 2.9, or using the ideas in [Reference Isett and Oh36]. For completeness, we sketch its proof at the end of Section 7, considering a simplified version of the preceding construction.
3.2 Proof of Theorem 1.7
We first prove the theorem under the assumption that the subsolution
$(v_0,p_0,\mathring {R}_0)(\cdot ,t)$
is compactly supported for each time
$t\in [0,T]$
. This assumption will be relaxed in next subsection to a condition on the support of
$\mathring {R}_0$
. We fix
$0<\beta <1/3$
, we choose b satisfying (3.18) and
$\alpha $
smaller than the threshold given by Proposition 3.2. Next, we use the scale invariance of the Euler equations and subsolutions

to assume that
$\Omega \subset (0,1)^3$
and
$\operatorname {\mathrm {supp}} (v_0,p_0,\mathring {R}_0)\subset (0,1)^3\times [0,T]$
. The desired energy profile must also be modified:
$e(t) \mapsto \rho ^{-3}e(t)$
. Note that this preserves (1.3). For convenience, we denote the rescaled subsolution like the original. It then suffices to construct the desired solution in this case and then reverse the change of variables.
Next, we use Lemma 3.3 with
$\lambda =\lambda _1^{12\alpha }$
to obtain a subsolution
$(v_1,p_1,\mathring {R}_1)$
that equals the initial subsolution outside the set

and satisfies the estimates

where the constants
$c_N$
are independent of
$\lambda _1$
but they will depend on
$\Omega $
and the initial subsolution. In addition, by (1.3) we have

Next, we use another scale invariance of the Euler equations (and the definition of subsolution):

We choose

and we begin to work in this rescaled setting, which we will indicate with a superscript r. We are thus working in the interval
$[0,\widetilde {T}]$
, where
, and we try to prescribe the energy profile
. By construction of
$\Gamma $
we have

and

It follows from the definition of
$\Gamma $
that if a is sufficiently large we have

so (3.17) holds. On the other hand,

because
$\Gamma $
becomes arbitrarily small by increasing a.
Next, we observe that
$(v_0^r,p_0^r,\mathring {R}{}_0^r)$
still satisfies (3.6)
$-$
(3.8) with the same sequence
$\{y_N\}_{N=0}^\infty $
. Regarding
$(v_1^r,p_1^r,\mathring {R}{}_1^r)$
, it follows from the definition of the rescaling that

On the other hand, since the constants
$c_N$
are independent of
$\lambda _1$
, for sufficiently large a we have

Finally,
$(v_1^r,p_1^r,\mathring {R}{}_1^r)=(v_0^r,p_0^r,\mathring {R}{}_0^r)$
outside

Let us consider the sets
$A_q$
defined in (3.12). We see that
$(v_1^r,p_1^r,\mathring {R}{}_1^r)=(v_0^r,p_0^r,\mathring {R}{}_0^r)$
outside
$A_1\times [0,\widetilde {T}]$
for sufficiently small
$\alpha $
and sufficiently large a. Remember that
$\delta _q$
and
$\lambda _q$
depend on a through the expressions (3.4) and (3.3), respectively.
From now on we assume that we are working in this rescaled problem and we omit the superscript r. Once we obtain the desired weak solution in this setting, to obtain the solution to the original problem it suffices to undo the scaling. By the previous discussion, the energy profile satisfies the inductive hypotheses (3.5) and the subsolution
$(v_1,p_1,\mathring {R}_1)$
satisfies (3.13)
$-$
(3.17). In addition, the initial subsolution
$(v_0,p_0,\mathring {R}_0)$
satisfies (3.6)
$-$
(3.8) for some sequence
$\{y_N\}_{N=0}^\infty $
that does not depend on a. Therefore, we may apply Proposition 3.2 iteratively, obtaining a sequence of compactly supported smooth subsolutions
$\{(v_q,p_q,\mathring {R}_q)\}_{q=1}^\infty $
.
It follows from (3.19) that
$v_q$
converges uniformly to some continuous map v. On the other hand, note that the pressure
$p_q$
is the only compactly supported solution of

Therefore,
$p_q$
also converges to some pressure
$p\in L^r(\mathbb {R}^3)$
for any
$1\leq r<\infty $
. Since
$\mathring {R}_q$
converges uniformly to
$0$
, we conclude that the pair
$(v,p)$
is a weak solution of the Euler equations.
Furthermore, using (3.19) we obtain

so
$\{v_q\}_{q=1}^\infty $
is uniformly bounded in
$C^0_t C^{\beta '}_x$
for all
$\beta '<\beta $
. To recover the time regularity, see [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7].
Next, note that
$(v_q,p_q,\mathring {R}_q)=(v_0,p_0,0)$
in
$(\mathbb {R}^3\backslash \Omega )\times [0,T]$
for all q by (3.13) and the definition of
$A_q$
. Hence, we have
$(v,p)=(v_0,p_0)$
in
$(\mathbb {R}^3\backslash \Omega )\times [0,T]$
.
Finally, it follows from (3.17) and the fact that
$\delta _{q+1}\to 0$
as
$q\to \infty $
that
$\int _\Omega \left \vert v(x,t)\right \vert {}^2dx=e(t)$
, as we wanted. This completes the proof of the theorem for the case that the initial subsolution is compactly supported for all time.
3.3 Dropping the compact support condition
Once we have proved Theorem 1.7 for the case that
$(v_0,p_0,\mathring {R}_0)(\cdot ,t)$
is compactly supported, it is easy to relax this condition and show that it suffices that
$\mathring {R}_0(\cdot ,t)$
is compactly supported.
Indeed, let us choose a bounded domain with smooth boundary
$U\Supset \Omega $
. As we mentioned in Lemma 2.11 and Remark 2.12, any subsolution defined in all
$\mathbb {R}^3$
automatically satisfies the conditions in Theorem 1.6. Hence, there exists a subsolution
$(\widetilde {v}_0,\widetilde {p}_0,\mathring {\widetilde {R}}_0)\in C^\infty (\mathbb {R}^3\times [0,T])$
that extends
$(v_0,p_0,\mathring {R}_0)$
outside
$\overline {U}$
and that is compactly supported in each time slice. Since it is an extension, we see that the Reynolds stress vanishes in
$\overline {U}\backslash \Omega $
.
We can now apply Theorem 1.7 in the case that the subsolution is compactly supported for each time slice, thus obtaining a weak solution
$(v,p)$
with the appropriate regularity and such that
$(v,p)=(v_0,p_0)$
in
$(\overline {U}\backslash \Omega )\times [0,T]$
. Since
$\Omega \Subset U$
, we may glue this region back into
$(v_0,p_0)$
, obtaining a weak solution that equals
$(v_0,p_0)$
outside
$\Omega \times [0,T]$
.
Finally, since
$\Omega $
and
$\mathbb {R}^3\backslash \overline {U}$
are disjoint, when we apply Theorem 1.7 we may fix the energy profile in each region independently. Therefore, we can prescribe the energy profile in
$\Omega $
of the final weak solution to be any smooth function satisfying condition (1.3).
4 Proof of Proposition 3.2
The different steps in the proof of Proposition 3.2 are presented in Subsection 4.1. These steps are elaborated in the next subsections, and we complete the proof of the proposition in Subsection 4.6. Roughly speaking, our proof adapts the arguments of [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7] to the nonperiodic setting. To do so, we introduce an additional step in the iteration: a gluing in space that ensures that the error does not spread out to the whole space when we glue in time.
4.1 Stages of the proof
-
1. Preparing the subsolution. We mollify our subsolution
$(v_q,p_q,\mathring {R}_q)$ to avoid the loss of derivatives problem, obtaining a new subsolution
$(v_\ell ,p_\ell ,\mathring {R}_\ell )$ . It is convenient to glue it in space to the original subsolution far from the turbulent zone.
-
2. Gluing in space. We pick a collection of times
$\{t_i\}\subset [0,T]$ and we consider the solutions
$(v_i,p_i)$ of the Euler equations with initial data
$\widetilde {v}_\ell (t_i)$ . We glue them in space to
$(\widetilde {v}_\ell ,\widetilde {p}_\ell ,\mathring {\widetilde {R}}_\ell )$ , obtaining new subsolutions
$(\widetilde {v}_i,\widetilde {p}_i,\mathring {\widetilde {R}}_i)$ . The error
$\mathring {\widetilde {R}}_i$ is small and these subsolutions equal
$(v_0,p_0,\mathring {R}_0)$ near
$\partial \Omega $ .
-
3. Gluing in time. We glue together the subsolutions
$(\widetilde {v}_i,\widetilde {p}_i,\mathring {\widetilde {R}}_i)$ , obtaining a new subsolution
$(\overline {v}_q,\overline {p}_q,\mathring {\overline {R}}_q)$ in which most of the error is concentrated in temporally disjoint regions. The error remains localized within
$\Omega $ owing to the fact that the differences between the subsolutions
$(\widetilde {v}_i,\widetilde {p}_i,\mathring {\widetilde {R}}_i)$ vanish near
$\partial \Omega $ .
-
4. Perturbation. We add a highly oscillatory perturbation to reduce the error. In fact, we add many corrections, each of them reducing the error in one of the temporally disjoint regions. These perturbations do not interact with each other, which yields the optimal estimates for the new subsolution
$(v_{q+1},p_{q+1},\mathring {R}_{q+1})$ .
Throughout the iterative process we will use the notation
$x\lesssim y$
to denote
$x\leq C y$
for a sufficiently large constant
$C>0$
that is independent of a, b, and q. However, the constant is allowed to depend on
$\alpha $
,
$\beta $
,
$\Omega $
, and
$\{y_N\}_{N=0}^\infty $
and it may change from line to line.
4.2 Preparing the subsolution
The first step consists in mollifying the field in order to avoid the loss of derivatives problem, which is typical of convex integration. The problem is the following: to control a Hölder norm of
$(v_{q+1},p_{q+1},\mathring {R}_{q+1})$
we need estimates of higher-order Hölder norms of
$(v_q,p_q,\mathring {R}_q)$
. As the iterative process goes on, we need to estimate higher and higher Hölder norms of the initial terms of the sequence to control just the first few Hölder norms of the subsolution. However, if we mollify the subsolution
$(v_q,p_q,\mathring {R}_q)$
, we can control all the derivatives in terms of the first few Hölder norms and the mollification parameter.
Note that this process changes the subsolution in the whole space, yet mollification is only strictly necessary in
$A_q\times [0,T]$
, as
$(v_q,p_q,\mathring {R}_q)$
equals
$(v_0,p_0,\mathring {R}_0)$
outside of this set. Furthermore, it will be convenient for later estimates that the resulting subsolution equals the initial subsolution far from the turbulent zone, as this is a property that we want to impose unto
$(v_{q+1},p_{q+1},\mathring {R}_{q+1})$
. Hence, our approach consists in gluing the mollified subsolution to the initial subsolution, which is not quite demanding because both subsolutions are very close far from the turbulent zone.
We begin by fixing a mollification kernel in space
$\psi \in C^\infty _c(\mathbb {R}^3)$
and we introduce the mollification parameter

Since
$(\delta _{q+1}/\delta _q)^{1/2}=\lambda _q^{-\beta (b-1)}$
and our assumption (3.18) implies that
$\beta (b-1)<1/2$
, we see that we may choose
$\alpha $
sufficiently small and a sufficiently large so as to have

Note that our definition of
$\ell $
differs from the one in [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7] by a factor of
$\lambda _q^{-3\alpha /2}$
. The coefficients of
$\alpha $
used in [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7] are fine-tuned to their proof. Since we do some things differently, it is not surprising that we have to change some of these coefficients. In general the factor
$\lambda _q^{-3\alpha /2}$
is quite harmless, as only simple relationships like (4.2) are used throughout most of the paper, and these are the same here and in [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7]. The actual value of
$\ell $
is only used at the very end, when fine relationships between the parameters are needed to estimate
$\mathring {R}_{q+1}$
. We will study these situations when they arise, but in any case we will see that the extra factor is essentially irrelevant. Indeed,
$\alpha $
is assumed to be so small that in those inequalities the term containing
$\alpha $
is negligible. Our definition of
$\ell $
leads to a different coefficient multiplying
$\alpha $
, but this only changes how small
$\alpha $
has to be chosen, so it is not important.
After this brief digression, we define

where the convolution with
$\psi _\ell $
is in space only and
$f\mathring {\otimes }g$
denotes the traceless part of the tensor
$f\otimes g$
. It is easy to check that the triplet
$(v_\ell ,p_\ell ,\mathring {R}_\ell )$
is a subsolution and by [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7, Proposition 2.2] we have the following estimates:

Note that we have an extra factor
$\ell ^{2\alpha }$
in the estimate for the Reynolds stress in comparison to [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7]. This comes from the extra factor
$\lambda _q^{-3\alpha }$
in
$\|\mathring {R}_q\|_0$
and from our definition of
$\ell $
. They cause an extra factor
$\lambda _q^{-3\alpha }$
to appear in the estimate, which yields an extra factor
$\ell ^{2\alpha }$
by (4.2).
Obtaining an extra factor
$\ell ^{2\alpha }$
(actually
$\ell ^\alpha $
would suffice) is our reason for modifying the inductive estimate of the Reynolds stress and the definition of
$\ell $
. We will use this extra factor to compensate a suboptimal estimate that we will be forced to use in Section 5.
Once we have mollified the subsolution, we will glue it to the initial subsolution far from the turbulent zone. It will be convenient to divide
$A_{q+1}\backslash A_q$
into several pieces because we will have to do several constructions in this region. We define

where
$d_q$
was defined in (3.10). Using the elementary inequalities

we deduce
$\delta _{q+2}\gtrsim \lambda _q^{-2\beta b^2}$
. By (3.18) we have
$b^2<5/4$
, so for sufficiently small
$\alpha $
we have

Therefore,

In particular,
$\sigma \gg \ell $
. For
$1\leq j\leq 5$
we define

By hypothesis,
$(v_q,p_q,\mathring {R}_q)$
equals
$(v_0,p_0,\mathring {R}_0)$
outside
$A_q\times [0,T]$
. Hence, it follows from (3.6)
$-$
(3.8) that



Thus, both subsolutions are very close in this region, which makes gluing them much easier. Taking into account (4.6), it follows from Lemma 2.6 that there exists a potential
$A\in C^\infty (B_1\times [0,T],\mathcal {A}^3)$
such that
$\operatorname {\mathrm {div}} A=v_\ell -v_0$
in
$B_1\times [0,T]$
and

Therefore, the matrices
$S_1$
and M that appear in Lemma 2.11 satisfy the estimates
$\left \|S_1\right \|_{N+\alpha }+\left \|M\right \|_{N+\alpha }\lesssim \ell ^2$
. We introduce this estimates into Lemma 2.11 to perform a gluing in the region
$B_1\backslash B_2$
. Since
$\sigma \gg \ell $
we may essentially ignore any factor coming from the derivatives of the cutoff by absorbing it into the
$\ell $
factor. Carrying out the rest of the construction of Lemma 2.11, we conclude that there exists a smooth subsolution such that

and satisfies the estimates




That is, we have obtained a new subsolution satisfying the same estimates as
$(v_\ell ,p_\ell ,\mathring {R}_\ell )$
but with the additional property of being equal to the initial subsolution far from the turbulent zone.
Although many constructions in [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7] work in
$\mathbb {R}^3$
with very little or no modification, it is more convenient to work with periodic fields so that we may use results from [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7] directly. Note that the subsolution that we have just obtained is supported in
$(0,1)^3\times [0,T]$
because it equals the initial subsolution far form the turbulent zone. This allows us to consider its periodic extension to
, which we denote the same. From now on, we consider that we are working in this setting.
4.3 Overview of the gluing in space
The key idea introduced by Isett in [Reference Isett35] is to glue in time exact solutions of the Euler equations to obtain a new subsolution
$(\overline {v}_q,\overline {p}_q,\mathring {\overline {R}}_q)$
such that
$\overline {v}_q$
is close to
$v_q$
but
$\mathring {\overline {R}}_q$
is supported in a series of disjoint temporal regions of the appropriate length. This allows the use of Mikado flows, leading to the optimal regularity
$C^\beta $
for any
$\beta <1/3$
(as in Onsager’s conjecture).
More specifically, we define the length

and we consider the smooth solutions of the Euler equations

where . We will see that they are defined in the time interval
$[t_i-\tau _q,t_i+\tau _q]$
. The pressure is recovered as the unique solution to the equation
$-\Delta p_i=\operatorname {\mathrm {tr}}(\nabla v_i\nabla v_i)$
with the normalization

We will see that these solutions remain sufficiently close to
$\widetilde {v}_\ell $
in their respective intervals. Following Isett’s ideas, we would like to glue them in time to obtain a subsolution in the whole interval
$[0,T]$
. The velocity field would remain sufficiently close to
$\widetilde {v}_\ell $
while the Reynolds stress would be localized to the intersection of the consecutive time intervals.
However, even if
$\mathring {R}_\ell $
is well localized, the solutions
$v_i$
will immediately differ from
$\widetilde {v}_\ell $
in the whole space. Furthermore, different solutions
$v_i$
,
$v_{i+1}$
will also differ in the whole space during the intersection of their temporal domains. If we tried to apply Isett’s procedure to them we would obtain a Reynolds stress that spreads throughout the whole space. This is not suitable for our purposes, so we must modify Isett’s approach.
What we will do is to glue in space the exact solutions
$v_i$
to
$\widetilde {v}_\ell $
in the region where
$\mathring {\widetilde {R}}_\ell $
is small, obtaining subsolutions
$(\widetilde {v}_i,\widetilde {p}_i,\mathring {\widetilde {R}}_i)$
. The Reynolds stress will no longer be 0, but it will be so small that we may ignore it in the current iteration. Since these subsolutions will coincide far from the turbulent zone, we will be able to glue them in time while keeping the Reynolds stress localized.
The actual process is not so simple because the difference between the exact solutions
$(v_i,p_i)$
and the subsolution
$(\widetilde {v}_\ell ,\widetilde {p}_\ell ,\mathring {\widetilde {R}}_\ell )$
is too big, leading to an unacceptably large error if we try to glue them. Our approach consists in producing a series of intermediate subsolutions that act as a sort of interpolation between them. Instead of a single gluing we perform a big number of them, going from
$v_i$
to
$\widetilde {v}_\ell $
far from the turbulent zone. The difference between two consecutive intermediate subsolutions will be very small so that the error introduced in each of these middle gluings is small.
At the end of this process we will obtain subsolutions
$(\widetilde {v}_i,\widetilde {p}_i,\mathring {\widetilde {R}}_i)$
such that


In addition, for
$\left \vert t-t_i\right \vert \leq \tau _q$
and any
$N\geq 0$
we will have the following estimates:



where we write

for the transport derivative. Furthermore, there exist smooth vector potentials
$\widetilde {z}_i$
defined in
$[t_i-\tau _q,t_i+\tau _q]$
such that
$\widetilde {v}_i=\operatorname {\mathrm {curl}} \widetilde {z}_i$
and at the intersection of two intervals we have:


These estimates are completely analogous to the ones in [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7, Proposition 3.3, Proposition 3.4] but we have the additional benefit of the fields being equal to the initial subsolution far from the turbulent zone. We do have to pay a price because now we have subsolutions instead of solutions of the Euler equations. Nevertheless, the errors
$\mathring {\widetilde {R}}_i$
are so small that we may ignore them until the
$(q+1)$
-th iteration.
All of the steps summarized here will be discussed in full detail in Section 5, where we derive the claimed estimates.
4.4 Overview of the gluing in time
Once we have our subsolutions
$(\widetilde {v}_i,\widetilde {p}_i,\mathring {\widetilde {R}}_i)$
we will glue them in time. The goal is to obtain a subsolution defined in all
$[0,T]$
that remains close to
$\widetilde {v}_\ell $
but in which most of the Reynolds stress is localized to temporally disjoint regions of the appropriate length. This will allow us to correct the error in each region separately using Mikado flows. Let


Note that
$\{I_j,J_i\}$
is a pairwise disjoint decomposition of
$[0,T]$
. We choose a smooth partition of unity
$\{\chi _i\}$
such that:
-
•
$\sum _i \chi _i=1$ .
-
•
$\operatorname {\mathrm {supp}} \chi _i\cap \operatorname {\mathrm {supp}} \chi _{i+2}=\varnothing $ . Furthermore,
(4.24)$$ \begin{align} \begin{aligned} \operatorname{\mathrm{supp}} \chi_i\subset \left(t_i-\frac{2}{3}\tau_q,t_i+\frac{2}{3}\tau_q\right),\\ \chi_i(t)=1 \qquad \forall t\in J_i. \end{aligned} \end{align} $$
-
• For any i and
$N\geq 0$ we have
(4.25)$$ \begin{align} \left\|\partial_t^N\chi_i\right\|_0\lesssim \tau_q^{-N}. \end{align} $$
We define

It is clear that
$\overline {v}_q$
is divergence-free and it equals
$v_0$
in
$B_3\times [0,T]$
because of (4.17). In addition, it inherits the estimates of
$\widetilde {v}_i$
:
Proposition 4.1. The velocity field
$\overline {v}_q$
satisfies the following estimates:




for all
$N\geq 0$
.
The proof is exactly the same as the proof of [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7, Proposition 4.3, Proposition 4.5] because our fields
$\widetilde {v}_i$
satisfy completely analogous estimates to the solutions
$v_i$
in [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7].
We conclude that the new velocity
$\overline {v}_q$
equals
$v_0$
far from the turbulent zone, it is close to
$\widetilde {v}_\ell $
and satisfies suitable bounds. Nevertheless, we must check if it leads to a subsolution. If
$t\in J_i$
, then in a neighborhood of t we have
$\chi _i=1$
while the rest of the cutoffs vanish. Thus, for all
$t\in J_i$
we have

Since
$(\widetilde {v}_i,\widetilde {p}_i,\mathring {\widetilde {R}}_i)$
is a subsolution, we have

On the other hand, if
$t\in I_i$
, then
$\chi _j=0$
for
$j\neq i,i+1$
and
$\chi _i+\chi _{i+1}=1$
. Hence, on
$I_i$
we have

After a tedious computation we obtain

where we have used the fact that
$(\widetilde {v}_i,\widetilde {p}_i,\mathring {\widetilde {R}}_i)$
are subsolutions.
In conclusion, for
$t\in J_i$
the triplet
$(\overline {v}_q,\overline {p}_q^{(1)},\overline {\mathcal R}^{(1)}_q)$
is trivially a subsolution, whereas for
$t\in I_i$
it suffices to express the right-hand side of Equation (4.30) as the divergence of a symmetric matrix. However, we must do this carefully because it is very important to keep under control the spatial support of the Reynolds stress. Indeed, since
$\overline {\mathcal R}^{(1)}_q$
equals
$\mathring {R}_0$
outside
$A_{q+1}\times [0,T]$
, any perturbation in the Reynolds stress must be contained within
$A_{q+1}\times [0,T]$
in order to satisfy the inductive property (3.13). Fortunately, we will be able to achieve this because the right-hand side of Equation (4.30) is supported in
$A_q+B(0,3\sigma )$
due to (4.17). We see that performing the gluing in space in the previous stage is crucial.
The logical approach would be to apply Lemma 2.9 to the right-hand side of (4.30) and define the new Reynolds stress as the sum of the obtained matrix and
$\overline {\mathcal R}^{(1)}_q$
. Unfortunately, we cannot simply do that because we cannot obtain the necessary estimates for the transport derivative from Lemma 2.9.
Nevertheless, this difficulty can be solved. In Section 6 we will find smooth symmetric matrices
$\overline {\mathcal R}^{(2)}_q,\overline {\mathcal R}^{(3)}_q$
solving the equation

for
$t\in I_i$
. We set
$\overline {\mathcal R}^{(2)}_q,\overline {\mathcal R}^{(3)}_q=0$
for
$t\notin \bigcup _{i}I_i$
. Since the source term is supported in
$A_q+B(0,3\sigma )$
due to (4.17), we will be able to choose
$\overline {\mathcal R}^{(2)}_q,\overline {\mathcal R}^{(3)}_q$
supported in
$A_q+B(0,4\sigma )$
. The motivation for each matrix is the following:
-
•
$\overline {\mathcal R}^{(2)}_q$ solves the equation except for a small error. We have good bounds for the derivative of the material derivative.
-
•
$\overline {\mathcal R}^{(3)}_q$ corrects the errors introduced when fixing the support of
$\overline {\mathcal R}^{(2)}_q$ . We do not have good bounds for its material derivative, but its
$C^0$ -norm is very small.
Using these auxiliary matrices, we define




By construction of
$\overline {\mathcal R}^{(2)}_q$
and
$\overline {\mathcal R}^{(3)}_q$
, we see that
$(\overline {v}_q,\overline {p}_q,\mathring {\,\overline {\!{R}}}_q)$
is a smooth subsolution. Furthermore, we have

because of (4.17) and the fact that
$\overline {\mathcal R}^{(2)}_q, \overline {\mathcal R}^{(3)}_q$
are supported in
$A_q+B(0,4\sigma )$
. We emphasize again the importance of performing a gluing in space to use
$\widetilde {v}_i$
instead of the solutions
$v_i$
. Otherwise, we would have no control on the Reynolds stress because
$v_i-v_{i+1}$
will in general be spread throughout the whole space.
We summarize the facts about the new Reynolds stress that we will prove in Section 6:
Proposition 4.2. The smooth symmetric matrices
$\mathring {\,\overline {\!{R}}}_q\strut {}^{(1)},\mathring {\,\overline {\!{R}}}_q\strut {}^{(2)}$
satisfy




Comparing (4.37) with the analogous estimate in [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7], we see that we estimate the
$C^N$
-norm instead of the
$C^{N+\alpha }$
-norm. This difference is immaterial because they merely use the
$C^{N+\alpha }$
-norm to estimate the
$C^N$
-norm, which leads to the same bound as (4.37).
In conclusion,
$\mathring {\,\overline {\!{R}}}_q\strut {}^{(1)}$
is so small that we may ignore it for the present iteration, whereas
$\mathring {\,\overline {\!{R}}}_q\strut {}^{(2)}$
is big but it is supported in temporally disjoint regions and it satisfies good estimates. The next stage of the process is aimed at correcting
$\mathring {\,\overline {\!{R}}}_q\strut {}^{(2)}$
by means of highly oscillatory perturbations.
All of the steps summarized here will be discussed in full detail in Section 6, where we derive the claimed estimates.
4.5 Overview of the perturbation step
We have localized most of the Reynolds stress, that is,
$\mathring {\,\overline {\!{R}}}_q\strut {}^{(2)}$
, to small disjoint temporal regions but to reduce it we must resort to convex integration.
First of all, we fix a cutoff
$\phi _q\in C^\infty (\mathbb {R}^3,[0,1])$
that equals
$1$
in
$A_q+\overline {B}(0,4\sigma )$
and whose support is contained in
$A_q+B(0,5\sigma )$
. In particular,
$\phi _q\equiv 1$
on the support of
$\mathring {\,\overline {\!{R}}}_q\strut {}^{(2)}(\cdot ,t)$
for all
$t\in [0,T]$
.
We follow the construction of [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7] with Mikado flows, but there are some differences:
-
• We control the support of the perturbation by introducing the cutoff
$\phi _q$ .
-
• To obtain the desired energy, we must use a slightly different normalization coefficient for the perturbation to account for the presence of the cutoff
$\phi _q$ when integrating.
-
• We construct the new Reynolds stress using Lemma 2.9 so that we have control on its support. To apply this lemma we must introduce a minor correction
$w_L$ to ensure that the perturbation has vanishing angular momentum.
Since
$\left \|\phi _q\right \|_N$
are much smaller than the
$C^N$
-norms of the other maps involved, the presence of the cutoff does not affect the estimates. In addition,
$w_L$
will be negligible because its size is determined by an integral quantity, which is very small for a highly oscillating perturbation.
The form of the perturbation
$\widetilde {w}_{q+1}= v_{q+1}-v_q$
is
$\widetilde {w}_{q+1}=w_0+w_c+w_L$
, where
$w_0$
is the main perturbation term and it is used to cancel the Reynolds stress,
$w_c$
is a small correction to ensure that the perturbation is divergence-free and
$w_L$
is a tiny correction that ensures that the perturbation has vanishing angular momentum. This term is not present in [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7] because they do not control the support of the Reynolds stress, so it suffices to use
.
At the end of the process we will obtain a new subsolution
$(v_{q+1},p_{q+1},\mathring {R}_{q+1})$
that equals
$(v_0,p_0,\mathring {R}_0)$
outside
$A_{q+1}\times [0,T]$
and satisfying the following estimates:



4.6 Proof of Proposition 3.2
Finally, we are ready to complete the proof of the proposition. The estimate (3.19) is a consequence of (4.10), (4.11), (4.26), (4.28) and (4.39):

where the constant C depends on
$\alpha $
,
$\beta $
,
$\Omega $
and
$(v_0,p_0,\mathring {R}_0)$
, but not on a, b or q. Thus, for any fixed b (3.19) holds for sufficiently large a. Regarding (3.15), we use the inequality at level q to get

Hence, if we choose a large enough we obtain (3.15). Finally, (3.16) follows from

The inductive hypotheses (3.13), (3.14) and (3.17) were obtained in the perturbation step, and we are done.
5 Gluing in space
In this section we develop the second step of the proof of Proposition 3.2, which was summarized in Section 4.3. We do it in three subsections.
5.1 Interpolating sequence
We begin by describing the intermediate subsolutions that we will use. First of all, we recall the following local existence result. It is standard, but we provide a proof for the sake of completeness.
Proposition 5.1. For any
$\alpha>0$
there exists a constant
$c(\alpha )>0$
with the following property. Given any
$C^\infty $
initial data
$u_0\in H^3(\mathbb {R}^3)$
, any
$C^\infty $
force
$f\in L^1_{\mathrm {loc}}(\mathbb {R},H^3(\mathbb {R}^3))$
, let us fix a constant
$T>0$
such that

Then there exists a unique solution
$u\in C^\infty (\mathbb {R}^3\times [-T,T])\cap C([-T,T],H^3(\mathbb {R}^3))$
to the Euler equation

Moreover, u obeys the bounds

for all
$N\geq 1$
, where the implicit constant depends on N and
$\alpha>0$
.
Proof. It is classical [Reference Swann48, Reference Kato37, Reference Temam50] that the 3d Euler equation is locally well-posed on
$H^3(\mathbb {R}^3)$
, and that the solution, which is defined a priori for some time
$T=T(\|u_0\|_{H^3(\mathbb {R}^3)})>0$
, can be continued (and stay smooth, provided that
$u_0\in C^\infty $
) as long as the norm
$\|u(t)\|_{H^3(\mathbb {R}^3)}$
remains bounded. Furthermore, the weak Beale–Kato–Majda criterion shows [Reference Beale, Kato and Majda3] that this norm is controlled by
$\|\nabla u\|_{L^1L^\infty }$
. More precisely, one has

Therefore, we only need to provide a uniform a priori estimate for
$\|\nabla u\|_{L^\infty }$
for all times
$t< T$
, where the maximum value
$T>0$
is yet to be determined. To this end, note that for any multi-index
$\theta $
with
$\left \vert \theta \right \vert =N$
we have

Since the pressure satisfies the equation
$-\Delta p=\operatorname {\mathrm {tr}}(\nabla u\nabla u)-\operatorname {\mathrm {div}} f$
, it follows that

Hence, we have

Thus, by Lemma B.2:

Specializing to the case
$N=1$
, we note that for all
$|t|<T$
one has

for some constant
$\tilde {c}(\alpha )>0$
. We define the constant
$c(\alpha )$
that appears in the statement of the proposition as
and then assume that
$T>0$
satisfies (5.1). Let
$y_0$
be the first term on the right-hand side of (5.3) and let
$y(t)$
be the solution to the ODE
$y'= \tilde {c}(\alpha ) y^2$
with
$y(0)=y_0$
. One finds

Since
$y(t)\lesssim 1$
for all
$|t|<T$
, by our choice of T, we conclude that the solution is well defined for
$|t|< T$
. Furthermore, inserting this estimate in (5.2) and applying Grönwall’s inequality, we obtain the desired Hölder bounds for
$N>1$
.
To construct the desired sequence of subsolutions, we define


For
$i\geq 0$
and
$0\leq k\leq m$
we consider the smooth solutions of the forced Euler equations

Thus, the triplet
$(v_{i}{}^{k},p_i^k,{\mathring {R}}_i^k)$
is a subsolution. The pressure is recovered as the unique solution to the equation

with the normalization

Note that (in their common interval of existence) the pair
$(v_i^0,p^0_i)$
equals the solutions
$(v_i,p_i)$
considered in (4.15), while the pair
$(v_i^m,p^m_i)$
equals
$(\widetilde {v}_\ell ,\widetilde {p}_\ell )$
. The rest of the subsolutions form a sort of interpolating sequence between them.
We claim that
$(v_i^k,p^k_i)$
are defined in the time interval
$[t_i-\tau _q,t_i+\tau _q]$
, where
$\tau _q$
was defined in (4.14). Indeed, it follows from (4.11) that

On the other hand, by (4.12) we have

Hence, for sufficiently large a we have

By Proposition 5.1, we conclude that solutions
$(v^k_i,p^k_i)$
of the corresponding forced Euler equations exist in the claimed interval and they are unique. Furthermore, we have the following bounds:
Corollary 5.2. If a is sufficiently large, for
$\left \vert t-t_i\right \vert \leq \tau _q$
and any
$N\geq 1$
we have

Proof. It follows from Proposition 5.1 and estimates (4.11) and (4.12) that

Using definitions (4.1) and (4.14) and the comparison (4.2), we see that

from which the first inequality follows. For the second one, we use (4.14) again.
5.2 Estimates for the interpolating sequence
Now that we have defined our subsolutions, we must ensure that they remain close to
$\widetilde {v}_\ell $
and to each other. Taking into account that
$(\widetilde {v}_\ell ,\widetilde {p}_\ell ,\mathring {\widetilde {R}}_\ell )$
and
$(v_i^k,p_i^k,\mathring {R}_i^k)$
are subsolutions, we see that the difference satisfies the following transport equation:

Hence, the difference
$\widetilde {v}_\ell -v_i^k$
satisfies the same equation as
$v_\ell -v_i$
in [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7] except for a factor multiplying
$\operatorname {\mathrm {div}}\mathring {\widetilde {R}}_\ell $
, but it is less than or equal to 1. Furthermore, by (4.11) and (5.7) the fields
$\widetilde {v}_\ell $
and
$v_i^k$
satisfy the same estimates as
$v_\ell $
and
$v_i$
. Therefore, we may argue as in [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7, Proposition 3.3], obtaining:
Proposition 5.3. If a is sufficiently large, for
$\left \vert t-t_i\right \vert \leq \tau _q$
and any
$N\geq 0$
we have



where we write

for the transport derivative.
Note that our extra factor
$\ell ^{2\alpha }$
in (4.12) is inherited by these estimates. Next, let us consider the vector potential associated to the field
$v_i^k$
:

where
$\mathcal {B}$
is the Biot-Savart operator. We have

Since
$(v_i^k,p_i^k,\frac {k}{m}\mathring {\widetilde {R}}_\ell )$
is a subsolution, we have

On the other hand, the average of
$v_i^k$
at time
$t=t_i$
is the average of
$\widetilde {v}_\ell $
at that time, which vanishes because
$\widetilde {v}_\ell (\cdot ,t_i)$
is divergence-free and its support is contained in
$(0,1)^3$
. Therefore,
$v_i^k$
has zero mean and we have
$\operatorname {\mathrm {curl}} z_i^k=v_i^k$
.
Since our fields satisfy the same estimates as the fields in [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7], we can again argue as in [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7, Proposition 3.4], obtaining:
Proposition 5.4. For
$\left \vert t_i-\tau _q\right \vert \leq \tau _q$
and any
$N\geq 0$
we have


where
$D_{t,\ell }=\partial _t+\widetilde {v}_\ell \cdot \nabla $
.
In summary, the difference
$v_i^k-\widetilde {v}_\ell $
satisfies the same stability estimates as the ones in [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7] plus an additional factor
$\ell ^{2\alpha }$
due to the difference in the definition of
$\ell $
and in the inductive estimate (3.14). Furthermore, we will see that the difference between consecutive fields is much smaller, which will allow us to obtain suitable bounds for the gluing. We argue as in [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7, Proposition 3.3]. Subtracting the equation for each field and rearranging we obtain

Taking the divergence, we have

Since is a Calderón-Zygmund operator, we obtain

where we have used (4.12) and (5.7). The additional factor
$\ell ^{2\alpha }$
is not needed here, so we just omit it. Inserting this estimate into Equation (5.13) and using (4.12) and (5.7) again, we obtain

Applying Lemma B.2 yields

Using Grönwall’s inequality and the assumption
$\left \vert t-t_i\right \vert \leq \tau _q$
we conclude

If we carry on arguing as in [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7, Proposition 3.3] we obtain the following higher-order estimates:

Let us use this bound to estimate the other fields. We may rewrite the equation for the pressure as

because

Interpolating between (3.14) and (3.15) and between (4.10) and (4.11) we have

For the higher-order bounds we use Corollary 5.2. Therefore, from (5.14) we conclude

Introducing this estimate and (4.12) in Equation (5.15), we finally obtain

Let us now estimate the difference in the vector potentials. We recall the identity and that
$\operatorname {\mathrm {div}} z_i^k=0$
. Hence, taking the
$\operatorname {\mathrm {curl}}$
in Equation (5.13) and rearranging we arrive at

Reasoning as in the case of the pressure, we conclude

Since
$z_i^k(\cdot ,t_i)=z_i^{k+1}(\cdot ,t_i)=\mathcal {B}\,\widetilde {v}_\ell (\cdot ,t_i)$
, the difference vanishes at
$t=t_i$
. Using the assumption
$\left \vert t-t_i\right \vert \leq \tau _q$
we deduce

5.3 Gluing the interpolating sequence
Now that we have the appropriate estimates, we will start gluing the subsolutions
$(v_i^k,p_i^k,\mathring {R}_i^k)$
to one another to construct a subsolution that equals
$(v_i,p_i,0)$
in
$A_q+B(0,2\sigma )$
and
$(v_0,p_0,\mathring {R}_0)$
in
$B_3$
. For the sake of clarity, we will do it inductively. Let

and for
$k\geq 0$
consider the sets

and we fix smooth cutoff functions
$\varphi ^k\in C^\infty (\Omega ^{k+1},[0,1])$
that equal 1 in a neighborhood of
$\Omega ^k$
. By Lemma B.1, we may assume the bounds
$\|\varphi ^k\|_N\lesssim r^{-N}$
.
We will construct a sequence of subsolutions
$(\widetilde {v}_i^{ k},\widetilde {p}_i^{ k},\mathring {\widetilde {R}}_i\strut {}^{k})$
and potentials
$\widetilde {z}_i^{ k}$
with
$\widetilde {v}_i^{ k}=\operatorname {\mathrm {curl}} \widetilde {z}_i^{ k}$
such that


Furthermore, for
$\left \vert t-t_i\right \vert \leq \tau _q$
and any
$N\geq 0$
they satisfy





where we write

for the transport derivative.
If we could construct such a sequence, setting would yield the subsolution and potential claimed in Section 4.3. Indeed, (4.18) and (4.19) are just (5.22) and (5.23). The estimates (4.20) and (4.21) follow from (5.24) and (5.25) by interpolation, that is, we use (A.3) to estimate the
$C^{N+\alpha }$
seminorm using the
$C^0$
and
$C^{N+1}$
norms:

This, combined with the estimate for the
$C^N$
norm, yields (4.20). Note that here we have obtained an extra factor
$\ell ^\alpha $
, but we discard it because it will not be necessary. Estimate (4.19) follows from (5.25) by an analogous argument. Meanwhile, to obtain (4.22) and (4.23) from (5.26) and (5.27), we also need to apply the triangle inequality and use the fact that
$z_{i+1}^m-z_i^m=0$
. Recall that both vector potentials are just the restriction of
$\mathcal {B} \widetilde {v}_\ell $
to their respective intervals. Finally, (4.17) follows from (5.21) because
$mr\lesssim \lambda _q^{-1/10}$
, so by (4.5) it must be smaller than
$\sigma $
for sufficiently large a.
Let us then construct this sequence. We define the initial term as

It follows from Corollary 5.2, Proposition 5.3 and Proposition 5.4 that this term satisfies Equations (5.21) to (5.27). Next, let us suppose that we have defined the k-th term
$(\widetilde {v}_i^{ k},\widetilde {p}_i^{ k},\mathring {\widetilde {R}}_i\strut {}^{ k},\widetilde {z}_i^{ k})$
satisfying these inductive hypotheses. We will construct the
$(k+1)$
-th term satisfying them, too. To do so, we will apply Lemma 2.11 to glue
$(\widetilde {v}_i^{ k},\widetilde {p}_i^{ k},\mathring {\widetilde {R}}_i\strut {}^{ k})$
and
$(v_i^{k+1},p_i^{k+1},\mathring {R}_i^{k+1})$
in the region
$U^k$
. Since these subsolutions are defined in the whole space, by Remark 2.12 we do not need to check the compatibility conditions (2.16) and (2.17).
Note that in Lemma 2.11 we use skew-symmetric matrices instead of potential vectors because it is stated in any dimension
$n\geq 2$
. However, it is completely equivalent: given a potential vector z, we simply define
$A_{ij}=\varepsilon _{ijk}z_k$
, where
$\varepsilon _{ijk}$
is the usual Levi-Civita symbol. It is easy to check that A is skew-symmetric and
$\operatorname {\mathrm {curl}} z=\operatorname {\mathrm {div}} A$
.
Hence, applying Lemma 2.11 we obtain a subsolution satisfying:

for
$\left \vert t-t_i\right \vert \leq \tau _q$
. In addition, there exists a smooth vector potential
$\widetilde {z}_i^{ k+1}$
with
$\widetilde {v}_i^{ k+1}=\operatorname {\mathrm {curl}} \widetilde {z}_i^{ k+1}$
and such that
$\widetilde {z}_i^{ k+1}(\cdot ,t)\equiv z_i^{k+1}(\cdot ,t)$
outside
$\Omega ^{k+1}$
. Thus, (5.21) is satisfied.
Furthermore, by definition of
$v_i^{k+1}$
and the inductive hypothesis (5.22) we have

In addition, by (5.21) we know that
$\widetilde {z}_i^{ k}$
equals
$z_i^{k}$
outside
$\Omega ^k$
. Since
$z_i^{k}$
and
$z_i^{k+1}$
both equal
$\mathcal {B}\,\widetilde {v}_\ell $
at time
$t=t_i$
, we see that the difference
$z_i^{k+1}-\widetilde {z}_i^{ k}$
vanishes outside
$\Omega ^k$
at
$t=t_i$
. Inspecting Lemma 2.11 and replacing skew-symmetric matrices by the equivalent potential vectors, we see that

Since only the time derivative of
$w_L$
matters, we may assume that it vanishes at
$t=t_i$
, that is, we start the integration in (2.26) at
$t=t_i$
. Hence, (5.22) holds.
Concerning the estimates, we see in Lemma 2.11 that we only need to consider the bounds in
$U^k$
, so by (5.21) we only need to study the difference between
$(v_i^k,p_i^k,\mathring {R}_i^k,z_i^k)$
and
$(v_i^{k+1},p_i^{k+1},\mathring {R}_i^{k+1},z_i^{k+1})$
. Since

it follows from Equations (5.16), (5.17) and (5.19) that the matrix M that appears in Lemma 2.11 satisfies

With this bound we may use the estimates from Lemma 2.11. Let us focus on the vector potential:


To use these estimates, the following inequality will be useful:

where we have used that the exponent in the middle term is greater than
$1/20$
for any
$0<\beta <1/3$
and
$1<b<11/10$
. We compute

where we have used (5.30) and we have assumed
$\alpha $
to be sufficiently small. Hence, (5.26) holds for the
$(k+1)$
-th term. In addition, (5.24) clearly follows from it.
Note that we have used (5.11) to estimate
$\left \|\cdot \right \|_N$
, that is, we lose an
$\alpha $
. This is clearly not optimal, but it cannot be avoided for
$N=0$
. Thus, we pay the price of losing a factor
$\ell ^\alpha $
in the estimates for
$\left \|\cdot \right \|_{N+\alpha }$
. To compensate this, we gain an extra factor
$\ell ^{2\alpha }$
from (3.14) and (4.1), which are different from their counterparts in [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7].
Let us now estimate the material derivative. By the triangle inequality:

The last two terms can be estimated in the same manner, so we just study one:

where we have used
$\left \|\varphi ^k\right \|_N\lesssim r^{-N}\lesssim \ell ^{-N}$
and
$\tau _qr^{-1}\lesssim 1$
. Next, taking into account that
$\left \|\widetilde {v}_\ell \right \|_N\lesssim \ell ^{-N}$
, we have

where we have used
$\tau _qr^{-1}\lesssim 1$
, the inequality (5.30) and we have assumed
$\alpha $
to be sufficiently small. Using (5.29) along with the same tricks, we obtain the same estimates for the first term in Equation (5.31), and we conclude that (5.27) holds for the
$(k+1)$
-th term.
To obtain (5.25), we estimate the commutator
$[D_{t,\ell },\operatorname {\mathrm {curl}}]$
. We fix an arbitrary vector field u and we compute

Therefore, we have

where we have used (4.11) and that
$\delta _q^{1/2}\lambda _q\tau _q=\ell ^{2\alpha }\lesssim 1$
by definition of
$\tau _q$
. Hence,

Finally, let us consider the size of the new Reynolds stress. Taking into account that
$\widetilde {z}_i^{ k}$
equals
$z_i^k$
in
$U^k$
, it follows from (5.20) and Lemma 2.11 that

where we have used
$\tau _qr^{-1}\lesssim 1$
. Taking into account that
$\widetilde {v}_i^{ k}$
equals
$v_i^k$
in
$U^k$
, it follows from (5.14) and Lemma 2.11 that

Taking into account (5.30), we see that for sufficiently small
$\alpha $
and sufficiently large a we have

If
$x\in \Omega ^k$
, then
$\varphi ^k(x)=1$
and by (5.23) we have

On the other hand, if
$x\notin \Omega ^k$
, then
$\mathring {\widetilde {R}}_i\strut {}^{ k}(x,t)=\frac {k}{m}\mathring {\widetilde {R}}_\ell (x,t)$
by (4.17). Furthermore, we also have
$\mathring {\widetilde {R}}_\ell (x,t)=\mathring {R}_0(x,t)$
because
$\operatorname {\mathrm {dist}}(x,A_q)\geq 2\sigma $
. Since
$x\notin A_q$
, it follows from (3.11) that
$|\mathring {R}_0(x,t)|\leq \frac {1}{4}\delta _{q+2}\lambda _{q+1}^{-6\alpha }$
, so

for
$x\notin \Omega ^k$
. We conclude that (5.23) holds for the
$(k+1)$
-th term.
6 Gluing in time
In this short section we develop the third step of the proof of Proposition 3.2, which was summarized in Section 4.4.
By [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7, Proposition 4.4] the matrix satisfies the following bounds for any
$N\geq 0$
:

We define

and we fix smooth cutoff functions
$\theta _j$
such that
-
•
$\theta _j\equiv 1$ in a neighborhood of
$A_q+B(0,3\sigma +(j-1)r)$ ,
-
• the support of
$\theta _j$ is contained in
$A_q+B(0,3\sigma +jr)$
for
$1\leq j\leq m$
. We define

Note that
$\operatorname {\mathrm {supp}} \overline {\mathcal R}^{(2)}_q(\cdot ,t)$
is contained in
$A_q+B(0,4\sigma )$
because
$mr<\sigma $
for a sufficiently large. Since
$r^{-1}\lesssim \ell ^{-1}$
, we have

Let us estimate the material derivative. We compute

Regarding the second term, it follows from (4.27) and
$\left \|\widetilde {v}_\ell \right \|_N\lesssim \ell ^{-N}$
that the new field also satisfies
$\left \|\overline {v}_q\right \|_N\lesssim \ell ^{-N}$
. Thus,

Since the support of the
$\nabla \theta _j$
are pairwise disjoint, we have

We conclude that the matrix
$\overline {\mathcal R}^{(2)}_q$
satisfies the right estimates but we have changed the equation:

where we have used that the support of
$\operatorname {\mathrm {div}} S$
is contained in
$A_q+B(0,3\sigma )$
by (5.21). Let us correct this. We define
, whose support is contained in

Therefore, by Lemma B.4 we have

We wish to apply Lemma 2.9, so we have to check the compatibility conditions. We fix a Killing field
$\xi $
and we compute

where we have used (2.6) and the fact that
$\theta _j=1$
on the support of
$\operatorname {\mathrm {div}} S$
. To show that this integral vanishes, we first compute

because of (2.6) and the fact that the matrix in parentheses is compactly supported due to (4.9) and (4.17). Since
$\widetilde {v}_i(\cdot ,t_i)=\widetilde {v}_\ell (\cdot ,t_i)$
, we see that
$\int \xi \cdot (\widetilde {v}_i-\widetilde {v}_\ell )=0$
. Repeating this for
$\widetilde {v}_{i+1}$
and subtracting, we conclude

for any Killing field
$\xi $
. Therefore, by Lemma 2.9 there exists a smooth symmetric matrix
$M_j$
such that
$\operatorname {\mathrm {div}} M_j=\rho _j$
and whose support is contained in

Furthermore, we have the estimate

We define

By construction of the
$M_j$
we have

as we wanted. Since the supports of the
$M_j$
are pairwise disjoint, we see that

It follows from our assumption
$b-1<1/10$
that

Therefore, for a is sufficiently large and
$\alpha $
sufficiently small the matrix
$\mathring {\,\overline {\!{R}}}_q\strut {}^{ (1)}$
defined in (4.31) satisfies

where we have used that

due to (5.23). Thus,
$\mathring {\,\overline {\!{R}}}_q\strut {\!}^{ (1)}$
is so small that it may be ignored until the next iteration.
Finally, let us conclude the estimates for
$\mathring {\,\overline {\!{R}}}_q\strut {}^{ (2)}$
, defined in (4.32). It follows from [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7, Proposition 4.4] that

Combining this with (6.1) and (6.2), we infer (4.37) and (4.38).
7 The perturbation step
In this section we complete the final step in the proof of Proposition 3.2, which is done in Subsections 7.1 to 7.5. The last part, Subsection 7.6, contains the proof of Lemma 3.3.
For simplicity, we will assume that
$\Omega $
is connected. If it had several connected components
$\Omega ^j$
and we wanted to fix an energy profile
$e^j$
in each of them, we would simply carry out the construction of this section in each connected component, taking into account Remark 3.1. Note that
$\phi _q$
can be split into cutoffs
$\phi _q^j$
associated to each
$\Omega ^j$
.
7.1 Squiggling stripes and the stress
$\widetilde {R}_{q,i}$
Before we can define the perturbation, we need to introduce several objects. By [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7, Subsection 5.2] there exist nonnegative cutoff functions
$\eta _i$
with the following properties:
-
(i)
$\eta _i\in C^\infty (\mathbb {T}^3\times [0,T], [0,1])$ .
-
(ii)
$\operatorname {\mathrm {supp}}\eta _i\cap \operatorname {\mathrm {supp}}\eta _j=\varnothing $ .
-
(iii)
$\eta _i(x,t)=1$ for any x and
$t\in I_i$ .
-
(iv)
$\operatorname {\mathrm {supp}}\eta _i\subset \mathbb {T}^3\times (t_i-\frac {1}{3}\tau _q,t_{i+1}+\frac {1}{3}\tau _q)\cap [0,T]$ .
-
(v) There exists a positive geometric constant
$c_0>0$ such that for any
$t\in [0,T]$
$$\begin{align*}\sum_i \int_{\mathbb{T}^3}\eta_i(x,t)^2\,dx\geq c_0.\end{align*}$$
-
(vi) For any
$k,N\geq 0$ there exists constants depending on
$k,N$ such that
$$\begin{align*}\left\|\partial_t^k\eta_i\right\|_N\lesssim \tau_q^{-N}.\end{align*}$$
We replace
$\eta _i(x,t)$
by
$\eta _i(mx,t)$
for sufficiently large
$m\in \mathbb {N}$
, so that we may assume that the cutoffs
$\eta _i$
satisfy

for some constant
$\widetilde {c}_0$
depending on
$\Omega $
. This can be done because
$A_0$
will contain one of the cubes of a grid of sidelength
$m^{-1}$
for sufficiently large m. This settles the first inequality. Regarding the second inequality, note that at most 2 of the cutoffs are nonzero at any given time. In addition, the new cutoffs will still satisfy
$(\text {i})-(\text {vi})$
but the constants that appear in
$(\text {vi})$
will now depend on
$\Omega $
, too.
We proceed analogously to [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7], defining the amplitudes


Note that our definition of
$\rho _{q,i}$
differs from the one in [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7] in the normalization. Next, we define the backwards flows
$\Phi _i$
for the velocity field
$\overline {v}_q$
as the solution of the transport equation

Finally, we define


It follows from properties
$(\text {i})-(\text {iv})$
of
$\eta _i$
that
-
•
$\operatorname {\mathrm {supp}} R_{q,i}\subset \operatorname {\mathrm {supp}} \eta _i$ ,
-
• we have
$\sum _i\eta _i^2=1$ on
$\operatorname {\mathrm {supp}} \mathring {\,\overline {\!{R}}}_q\strut {}^{ (2)}$ ,
-
•
$\operatorname {\mathrm {supp}}\widetilde {R}_{q,i}\subset \mathbb {T}^3\times (t_i-\frac {1}{3}\tau _q,t_{i+1}+\frac {1}{3}\tau _q)$ ,
-
•
$\operatorname {\mathrm {supp}} \widetilde {R}_{q,i}\cap \operatorname {\mathrm {supp}} \widetilde {R}_{q,j}=\varnothing $ for all
$i\neq j$ .
We collect some other useful estimates:
Lemma 7.1. For
$a\gg 1$
sufficiently large we have

Furthermore, for any
$N\geq 0$







where . Moreover, for all
$(x,t)$
we have
$\widetilde {R}_{q,i}(x,t)\in B\left (\operatorname {\mathrm {Id}},\frac {1}{2}\right )\subset \mathcal {S}^3$
.
Proof. Since our subsolution
$(\overline {v}_q,\overline {p}_q,\mathring {\,\overline {\!{R}}}_q)$
satisfies analogous estimates to the ones in [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7], we may use the same argument to infer (7.6), (7.7) and (7.10). Estimate (7.8) then follows from (7.7) and (7.1). In fact, since our
$\rho _{q,i}$
only differs from the one in [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7] in a time-dependent normalization coefficient that is bounded above and below, the bounds for
$\left \|\rho _{q,i}\right \|_N$
are the same except for the implicit constant. Next, it follows from the property
$(\text {vi})$
of
$\eta _i$
that

Using this estimate and arguing as in [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7] we obtain (7.11). Finally, taking into account these estimates for
$\rho _{q,i}$
and the bounds (4.37) and (4.38), the facts regarding
$\widetilde {R}_{q,i}$
follow as in [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7].
7.2 Definition of the perturbation
The building blocks of the perturbation are Mikado flows, introduced in [Reference Daneri and Székelyhidi21, Lemma 2.3]:
Lemma 7.2. For any compact subset of positive-definite matrices
$\mathcal {N}\subset \mathcal {S}^3$
there exists a smooth vector field

such that, for every
$R\in \mathcal {N}$

and

Since
$W(R,\cdot )$
is
$\mathbb {T}^3$
-periodic and has zero mean, we may write

for some
$a_k\in C^\infty (\mathcal {N},\mathbb {R}^3)$
. Similarly, for some
$C_k\in C^\infty (\mathcal {N},\mathcal {S}^3)$
we have

It follows from (7.14) that

In addition, because of the smoothness we have

In the construction these estimates are used with a particular choice of
$\mathcal {N}$
(namely
$B(\operatorname {\mathrm {Id}},1/2)\subset \mathcal {S}^3$
) and m. This choice determines the constant M appearing in Proposition 3.2.
With these building blocks, we define the main perturbation term
$w_0$
as

Recall that in Lemma 7.1 we saw that
$\widetilde {R}_{q,i}$
takes values in the compact subset of positive-definite matrices
. Therefore, the previous expression is well-defined. To shorten the notation, we set

Thus, using (7.15) we may write

Although Mikado flows are divergence-free, the perturbation
$w_0$
will not be solenoidal, in general, due to the other factor. We must add a small correction term
$w_c$
so that
is divergence-free. We set

It is clear that
$w_{q+1}$
is divergence-free and the correction
$w_{q+1}-w_0$
can be seen to be a lower-order term in
$\lambda _{q+1}$
.
Unlike in [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7],
$w_{q+1}$
is not the final correction. We must add another small perturbation
$w_L$
to ensure the final correction
$\widetilde {w}_{q+1}=w_{q+1}+w_L$
has no angular momentum. We define
$L\in C^\infty ([0,T],\mathbb {R}^3)$
as

where the integration is undestood to be component-wise. We fix a ball
$B\subset A_0$
and we choose
$\psi \in C^\infty _c(B,\mathbb {R})$
such that
$\int \psi =1$
. We define the correction
$w_L$
as

By construction, the correction
$\widetilde {w}_{q+1}=w_{q+1}+w_L$
is given by
$\widetilde {w}_{q+1}=\operatorname {\mathrm {curl}} z$
for a potential vector z such that
$\int z\,dx=0$
component-wise. Using the vector identity
$\operatorname {\mathrm {div}}(\xi \times z)=z\cdot \operatorname {\mathrm {curl}} \xi -\xi \cdot \operatorname {\mathrm {curl}} z$
, we have

Since the curl of any Killing field is constant, we conclude

7.3 Estimates on the perturbation
Aside from
$w_L$
, our perturbation differs from the one in [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7] in the presence of the factor
$\phi _q$
. Nevertheless, the factor
$\phi _q$
appears multiplying
$a_k(\widetilde {R}_{q,i})$
. Thus, if we show that
$\phi _q a_k(\widetilde {R}_{q,i})$
satisfies the same bounds as
$a_k(\widetilde {R}_{q,i})$
, the same estimates that are derived in [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7] will apply here. It follows from (7.18), (7.12) and (7.13) that

Since
$\left \|\phi _q\right \|_N\lesssim \lambda _q^{-N/10} \lesssim \tau _q^{-N}\lesssim \ell ^{-N}$
we also have

We conclude that all of the estimates in [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7] are also valid here. In particular, we have
Lemma 7.3. Assuming a is sufficiently large, the perturbations
$w_0$
,
$w_c$
and
$w_q$
satisfy the following estimates:

where the constant M depends solely on the constant
$\widetilde {c}_0$
in (7.1).
We carry out the estimates for
$w_L$
with more detail. First, due to (7.18) we have
$\left \|b_{i,k}\right \|_0\lesssim \left \|\rho _{i,q}\right \|_0^{1/2}\left \vert k\right \vert {}^{-6}\lesssim \delta _{q+1}^{1/2}\left \vert k\right \vert {}^{-6}$
. Next, it follows from (7.6) that
$\left \|\nabla \Phi \right \|_0\lesssim 1$
. Introducing these estimates in the definition of L we obtain

where we have used that at most two of the
$b_{i,q}$
are nonzero at any given time. Since
$\psi $
is fixed throughout the iterative process, we conclude

Therefore, the correction
$w_L$
is really small. We see that for sufficiently large a the perturbation
$\widetilde {w}_{q+1}$
then satisfies

Regarding
$\partial _t w_L$
, we must first estimate

Next, we compute

where we have computed the material derivative of
$\nabla \Phi _i$
taking into account that
$\Phi _i$
solves the transport equation. Since
$\left \|\nabla \overline {v}_q\right \|_0\lesssim \delta _q^{1/2}\lambda _q$
by (4.28), we have

because
$\tau _q^{-1}=\delta _q^{1/2}\lambda _q\ell ^{-2\alpha }\lesssim \delta _q^{1/2}\lambda _q\lambda _{q+1}^{-3\alpha }$
by (4.2). Since
$\psi $
is fixed, we conclude

7.4 The final Reynolds stress
Taking into account that
$\phi _q$
and
$\sum _i\eta _i^2$
equal
$1$
on
$\operatorname {\mathrm {supp}} \mathring {\,\overline {\!{R}}}_q\strut {}^{(2)}$
, it follows from the definition of
$R_{q,i}$
that

Using this along with the fact that
$\left (\overline {v}_q,\overline {p}_q, \mathring {\,\overline {\!{R}}}_q\strut {\!}^{ (1)}+\mathring {\,\overline {\!{R}}}_q\strut {}^{(2)}\right )$
is a subsolution, we obtain:

Hence, we conclude that we may construct a new subsolution
$(v_{q+1},p_{q+1},\mathring {R}_{q+1})$
by setting

where the smooth symmetric matrix S satisfies

and the support of
$S(\cdot ,t)$
is contained in
$A_q+B(0,5\sigma )$
for all
$t\in [0,T]$
. If such a matrix existed, the new subsolution
$(v_{q+1},p_{q+1},\mathring {R}_{q+1})$
would equal
$(\overline {v}_q,\overline {p}_q,\mathring {\,\overline {\!{R}}}_q)$
in
$A_{q+1}\times [0,T]$
because the support of
$\phi _q$
and
$\widetilde {w}_{q+1}(\cdot ,t)$
is contained in
$A_q+B(0,5\sigma )$
. Therefore, it equals
$(v_0,p_0,\mathring {R}_0)$
in
$A_{q+1}\times [0,T]$
.
We will show that it is possible to construct S and we will derive the necessary estimates. Let . Note that the support of
$f(\cdot ,t)$
is contained in
$A_q+B(0,5\sigma )$
for all
$t\in [0,T]$
because so is the support of
$\phi _q$
and the perturbation. Next, we see that for any Killing field
$\xi $
we have

because of (7.19) and (2.6). Therefore, by Lemma 2.9 there exists a symmetric matrix
$S\in C^\infty (\mathbb {R}^3\times [0,T],\mathcal {S}^3)$
such that
$\operatorname {\mathrm {div}} S=f$
and with the stated support. We will now estimate the
$C^{\alpha }$
-norm of the potential theoretic solution of the equation. Arguing as in Lemma 2.9, this yields a bound for the
$C^0$
-norm of S.
We begin by studying the first term in f. It follows from (7.21) and the fact that
$\mathscr {R}$
is bounded on Hölder spaces that

The remaining three error terms are analogous to the ones in [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7]. Since our fields satisfy the same estimates as in [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7], the estimates for the potential-theoretic solution are completely analogous. We do have to take into account that the oscillation error has a slightly different expression than the one in [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7]:

The second term
$\mathcal {O}_2$
is the same as in [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7]. Regarding
$\mathcal {O}_1$
, it follows from the fact that the cutoffs
$\eta _i$
have pairwise disjoint support that

We use the definition of
$w_{0,i}$
and (7.16) to write

On the other hand, it follows from (7.17) that

so

which is the same as in [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7] except for the presence of
$\phi _q^2$
. Nevertheless, it is easy to check, as in Section 7.3, that
$\phi _q^2C_k(\widetilde {R}_{q,i})$
satisfies the same estimates as
$C_k(\widetilde {R}_{q,i})$
. Hence, we obtain the same bounds for
$\mathcal {O}_1$
.
Aside from the presence of the cutoff, there is another subtle difference that we have to take into account. The proof in [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7] uses that the following inequality holds for a suitable choice of the parameters:

Since our definition of
$\ell $
is slightly different, we must check that this inequality holds. Remember that
$\lambda _{q+1}\gtrsim \lambda _q^b$
by (4.4). Hence, we have

Note that
$(b-1)(1-\beta )>0$
so by choosing
$\alpha $
sufficiently small we can ensure that the coefficient multiplying N is positive. Thus, for sufficiently large N the exponent is positive and choosing a sufficiently large beats any geometrical constant. We conclude that with this choice of parameters the claimed inequality holds.
A similar argument is used several times, for instance, to obtain the inequality
$\lambda _{q+1}\ell \geq 1$
. Nevertheless, in all of them the difference in the definition of
$\ell $
is quite harmless and it only leads to choosing a smaller
$\alpha $
and a slightly larger N.
In conclusion, the estimates from [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7] apply to our case. Combining them with (7.22) we obtain

Since
$\int \xi \cdot f\,dx=0$
for all
$t\in [0,T]$
and any Killing field
$\xi $
, we may use the construction of Lemma 2.9 to modify
$\mathscr {R}\,f$
into a smooth symmetric matrix S such that
-
•
$\operatorname {\mathrm {div}} S=f$ ,
-
•
$\operatorname {\mathrm {supp}} S(\cdot ,t)\subset A_q+B(0,5\sigma )$ for all
$t\in [0,T]$ ,
-
•
$\left \|S\right \|_0\lesssim \delta _{q+1}^{1/2}\delta _q^{1/2}\lambda _q\lambda _{q+1}^{-(1-4\alpha )}$ .
Combining this with (7.20) we have

We claim that with a suitable choice of the parameters

In that case, (4.35) and (7.24) would yield (4.40) for sufficiently large a, as we wanted. To prove the claimed inequality, we compute

The condition (3.18) ensures that

so the exponent is positive for sufficiently small
$\alpha>0$
. Thus, choosing a sufficiently large beats any numerical constant and (7.25) follows.
7.5 The new energy profile
By definition

Note that in the last two terms we can integrate on the whole
$\mathbb {T}^3$
because the perturbation is supported in
$\Omega $
. Arguing as in [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7] yields the estimate

On the other hand, any term containing
$w_L$
will be smaller than this bound by (7.20). The remaining term is

where we have used (7.23). The second term can be estimated as in [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7] because
$\phi _q^2 C_k(\widetilde {R}_{q,i})$
satisfies the same estimates as
$C_k(\widetilde {R}_{q,i})$
, as argued several times. Regarding the first term:

We conclude that

which yields (4.41) for sufficiently large a.
7.6 Proof of Lemma 3.3
This lemma is just a simplified version of the construction presented in the previous subsections, so we will just sketch its proof. Since the initial Reynolds stress
$\mathring {R}_0$
and its derivatives vanish at
$\partial \Omega \times [0,T]$
, for any
$k\in \mathbb {N}$
there exist a constant
$C_k$
such that for any
$x\in \Omega $
we have

Therefore, if
$\operatorname {\mathrm {dist}}(x,\partial \Omega )<3\lambda ^{-1/12}$
we have

We fix a smooth cutoff function
$\phi \in C^\infty _c(\Omega ,[0,1])$
such that

It can be chosen so that
$\left \|\phi \right \|_N\lesssim \lambda ^{N/12}$
. This function will control the support of the perturbation.
Next, we define the backwards flows
$\Phi $
for the velocity field
$v_0$
as the solution of the transport equation

and we define

Let
$\mathcal {N}\subset \mathcal {S}^3$
be the compact subset of positive definite matrices whose eigenvalues take values between
$1/2$
and
$3/2$
. We see that
$\widetilde {R}$
takes values in
$\mathcal {N}$
. Thus, we may apply Lemma 7.2 and define

with . We also have

Next, the correction
$w_c$
is then defined so that
is divergence-free:

Regarding the angular momentum, we define

We fix a ball
$B\Subset \Omega $
and we choose
$\psi \in C^\infty _c(B)$
such that
$\int \psi =1$
. We add the correction
$w_L$
so that the perturbation has vanishing angular momentum:

If
$\Omega $
has several connected components
$\Omega ^j$
, we will have to consider the partial angular momentum
$L^j$
. We will need to add one such vortex to each
$\Omega ^j$
to cancel the angular momentum in each connected component of
$\Omega $
. We still denote the total correction as
$w_L$
.
The new velocity field is . Note that by taking a larger
$\lambda $
we may force
$B\subset \operatorname {\mathrm {supp}}\phi $
, so the perturbation vanishes for
$\operatorname {\mathrm {dist}}(x,\partial \Omega )\leq 2\lambda ^{-1/12}$
.
Since
$\left \|\phi \right \|_N\lesssim \lambda ^{N/12}$
, the dominant term is the exponential. Hence, from (7.27) and the definition of
$L(t)$
we conclude

Let us denote . Since
$D_t\Phi =0$
, we see that
$\left \vert L'(t)\right \vert \lesssim \lambda ^{-(1-1/12)}$
. We conclude

Finally, we define

where the smooth symmetric matrix S satisfies

It is easy to check that
$(v,p,\mathring {R})$
is a subsolution. Furthermore, the fact that the perturbation has vanishing angular momentum ensures that we may choose S with support contained in
$A_*$
by using Lemma 2.9. Therefore, the
$(v,p,\mathring {R})$
equals the initial subsolution outside
$A_*$
.
Regarding the estimates, by (7.26) we may write the term in brackets in (7.28) as

for certain vectors
$c_k$
such that
$\left \|c_k\right \|_N\lesssim \left \vert k\right \vert {}^{-6}\lambda ^{(N+1)/12}$
. The standard stationary phase lemma (see [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7]) yields

Continuing the construction of Lemma 2.9, the claimed bound for
$\mathring {R}$
follows.
Regarding the energy, (7.26) and a standard stationary phase lemma (see [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7]) yield:

Since
$\mathring {R}_0$
is trace-free,
$\operatorname {\mathrm {tr}}{\widetilde {R}}=3$
. We conclude that (3.20) holds for sufficiently large
$\lambda $
.
8 Proof of Theorem 1.1
We are ready to prove our main theorem using Theorems 1.6 and 1.7. First, we show that the conditions are necessary. Suppose that such a weak solution
$(v,p)$
exists. We fix a connected component
$\Sigma $
of
$\partial \Omega $
and
$a\in \mathbb {R}^3$
. In order to give us some room to mollify the subsolution, we also fix a smooth surface
$\Sigma '\subset \Omega $
that will be used to approximate
$\Sigma $
from the inside of
$\Omega $
. Given
$\varepsilon>0$
, if follows from the smoothness of
$(v_0,p_0)$
that
$\Sigma '$
can be chosen sufficiently close to
$\Sigma $
so that

for any
$t\in [0,T]$
.
Next, we fix a mollification kernel
$\psi \in C^\infty _c(\mathbb {R}^3\times \mathbb {R})$
whose support is contained in the unit ball and for
$0<\ell <\varepsilon $
we define

where
$f\mathring {\otimes }g$
denotes the traceless part of the tensor
$f\otimes g$
. Since
$(v,p)$
is a weak solution, it is easy to see that
$(v_\ell ,p_\ell ,\mathring {R}_\ell )$
is a smooth subsolution in
$\mathbb {R}^3\times (\varepsilon ,T-\varepsilon )$
. On the other hand, the values of
$(v_\ell ,p_\ell ,\mathring {R}_\ell )$
on
$\Sigma '$
depend only on
$(v_0,p_0)$
for
$\ell <\operatorname {\mathrm {dist}}(\Sigma ',\partial \Omega )$
because
$(v,p)$
equals
$(v_0,p_0)$
on
$\overline {\Omega }\times [0,T]$
. In particular, we have,

Hence, for sufficiently small
$\ell $
, for any
$t\in [\varepsilon ,T-\varepsilon ]$
we have

However, these integrals vanish:

because
$(v_\ell ,p_\ell ,\mathring {R}_\ell )$
is a smooth subsolution (see the proof of Lemma 2.11, equations (2.18) and (2.19)). We conclude that for all
$t\in (\varepsilon ,T-\varepsilon )$
we have

Since
$\varepsilon>0$
is arbitrary and
$(v_0,p_0)$
is smooth up to the endpoints of the interval, we deduce

Taking into account that
$a\in \mathbb {R}^3$
and the connected component
$\Sigma $
of
$\partial \Omega $
are arbitrary, we see that the conditions in Theorem 1.1 are, indeed, necessary.
Next, we prove that the conditions are also sufficient. First of all, we may assume that
$\Omega '\supset \overline {\Omega }$
is a bounded set with smooth boundary and with a finite number of connected components. Next, by Theorem 1.6 there exists a subsolution
$(\widetilde {v}_0,\widetilde {p}_0,\mathring {\widetilde {R}}_0)\in C^\infty (\mathbb {R}^3\times [0,T])$
that extends
$(v_0,p_0,0)$
outside
$\overline {\Omega }$
and whose spatial support is contained in
$\Omega '$
. In particular,
$\operatorname {\mathrm {supp}} \mathring {\widetilde {R}}_0(\cdot ,t)$
will be contained in
$\overline {\Omega }'\backslash \Omega $
. Applying Theorem 1.7, we obtain a weak solution of the Euler equations
$(v,p)$
that equals
$(\widetilde {v}_0,\widetilde {p}_0)$
outside
$\overline {\Omega }'\backslash \Omega $
. Therefore, its support is contained in
$\overline {\Omega }'$
and it extends
$(v_0,p_0)$
. On the other hand, when applying Theorem 1.7 we may prescribe any energy profile
$e\in C^\infty ([0,T])$
such that

Hence, we can define the constant
$e_0$
that appears in the statement of Theorem 1.1 as any number greater than the right-hand side of the previous inequality. Finally,
$v\in C^{\beta }(\mathbb {R}^3\times [0,T])$
, as we wanted, thus completing the proof of the theorem.
Sketch of the proof of Remark 1.2.
If we construct the spatial extension
$(\widetilde {v}_0,\widetilde {p}_0,\mathring {\widetilde {R}}_0)\in C^\infty (\mathbb {R}^3\times [0,T])$
so that
$\widetilde {v}_0$
has vanishing total angular momentum, we can easily extend in time to a subsolution
$(\hat {v}_0,\hat {p}_0,\mathring {\hat {R}}_0)\in C^\infty (\mathbb {R}^3\times [0,+\infty )$
whose temporal support is contained in
$[0,T')$
. By taking the energy profile e larger, if necessary, we may extend it to
$[0,T']$
maintaining an analogue of the condition (1.3).
We then carry out the same construction as in Theorem 1.7 with some minor modifications:
-
• in Lemma 3.3 and in the perturbation step of Proposition 3.2 we introduce a temporal cutoff so that we do not modify the subsolution at the times when the Reynolds stress is identically
$0$ and
-
• in the perturbation step of Proposition 3.2 we define
instead of (7.3) if the interval
$(t_i-\frac {1}{3}\tau _q,t_{i+1}+\frac {1}{3}\tau _q)$ is disjoint from
$[0,T]$ .
With such an scheme we only prescribe the energy profile in
$[0,T]$
. However, the energy profile in
$[T,T']$
does not differ much (depending on the initial Reynolds stress) from
$\int \left \vert \hat {v}_0(x,t)\right \vert {}^2dx$
. Hence, if we choose
$e(0)$
sufficiently large, the final weak solution will be admissible.
9 Open time interval
In this section we study what happens when the fields are defined in an open interval
$(0,T)$
and there is some singular behavior at the endpoints of the interval. So far we have always considered the supremum in time of the spatial Hölder norms of our fields. This is not an option for the situation that we have in mind, which is the setting for our applications.
9.1 Main result
Our approach consists in decomposing
$(0,T)$
as a countable union of closed intervals
$\{\mathcal {I}_k\}_{k=0}^\infty $
meeting only at their endpoints and trying to work in each of them independently. While some parts of the iterative scheme that we have discussed in the previous sections depend only on what is happening at the current time (solving the symmetric divergence equation, for instance), many others do not (whenever we have dealt with transport). Therefore, if the Reynolds stress is nonzero at the endpoints of the intervals, when we try to correct it the subsolution in one interval will affect its neighbor. This propagates the bad estimates from near
$t=0$
and
$t=T$
to any
$\mathcal {I}_k$
after enough iterations of the scheme.
Hence, if we want to isolate the closed intervals and work in each of them independently, we must ensure that the Reynolds stress vanishes identically at their endpoints:
Lemma 9.1. Let
$(v,p,R)\in C^\infty (\mathbb {R}^3\times [0,T])$
be a subsolution of the Euler equations. Let
$t_0\in (0,T)$
and let
$s>0$
be sufficiently small. Suppose that the support of
$R(\cdot ,t_0)$
is contained in an open set
$\Omega $
and that
. Then, there exists a smooth subsolution
$(\widetilde {v},\widetilde {p},\widetilde {R})$
such that
$\widetilde {R}(\cdot ,t_0)\equiv 0$
and such that
$(\widetilde {v},\widetilde {p},\widetilde {R})=(v,p,R)$
outside
$\Omega \times (t_0-s,t_0+s)$
. Furthermore, we have the following estimates:

for some constants independent of s and
$(v,p,R)$
.
Proof. We fix a smooth cutoff function
$\chi \in C^\infty _c((-1,1),\mathbb {R})$
that equals
$1$
in a neighborhood of the origin. Consider the field

The condition ensures that
$\widetilde {v}$
is divergence-free. By definition of
$\chi $
, we see that w vanishes unless
$\left \vert t-t_0\right \vert <s$
so we deduce
$\left \|w\right \|_N\leq C\,s\left \|R(\cdot ,t_0)\right \|_{C^{N+1}}$
. Next, we define

It is easy to see that
$(\widetilde {v},p,\mathring {R})$
is a subsolution. Furthermore, it follows from our choice of
$\chi $
and the fact that w vanishes at
$t=t_0$
that
$\mathring {R}(\cdot ,t_0)$
is identically
$0$
. On the other hand, the bound for
$\left \|w\right \|_0$
yields the claimed estimate for
$\widetilde {R}$
.
On the other hand, since the support of
$\chi $
is contained in
$(-1,1)$
and the support of
$R(\cdot ,t_0)$
is contained in
$\Omega $
, we see that the support of
$\widetilde {v}-v$
and
$\widetilde {R}-R$
is contained in
$\Omega \times (t_0-s,t_0+s)$
.
Finally, we absorb the trace of
$\mathring {R}$
into the pressure, which preserves the other properties that we have discussed.
Remark 9.2. The condition is quite restrictive, but it can be removed if one is willing to relinquish spatial control of the velocity field. Indeed, we may decompose
$\operatorname {\mathrm {div}} R(\cdot ,t_0)$
as the sum of a divergence-free field and a gradient, which we absorb into the pressure. The divergence-free part is canceled using the previous lemma. The issue is that the divergence-free component of
$\operatorname {\mathrm {div}} R(\cdot ,t_0)$
does not have compact support, in general. Thus, we modify the subsolution outside
$\operatorname {\mathrm {supp}} R$
and we loose the spatial control, which has been our main concern so far. This approach could yield interesting applications in
$\mathbb {T}^3$
, nevertheless. However, in
$\mathbb {R}^3$
we would have to modify the construction to address the fact that we have to add perturbations in the whole space. We do not pursue this path here.
Since our construction relies on keeping the velocity fixed at the endpoints of the intervals
$\mathcal {I}_k$
, we cannot expect to obtain a nonincreasing energy profile for the final weak solution. Indeed, by weak-strong uniqueness it should equal the smooth solution with that initial data (in its domain of definition) and that is not what will will obtain with the convex integration scheme.
Thus, instead of trying to fix the energy with this construction, we will focus on keeping the changes small after each iteration. Our goal is to ensure that the energy profile can be extended to a continuous function in
$[0,T]$
. In that case, we may use Theorem 1.7 to add a (nonsingular) perturbation elsewhere so that the total energy achieves the desired profile.
Hence, the main result that we will prove in this section, which is a nontrivial variation of Theorem 1.7, is:
Theorem 9.3. Let
$0<\beta <1/3$
. Let
$T>0$
and let
$\{\mathcal {I}_k\}_{k=0}^\infty $
be a sequence of closed intervals meeting only at their endpoints and such that
$(0,T)=\bigcup _{k\geq 0}\mathcal {I}_k$
. Let
$\Omega _k\Subset (0,1)^3$
be a bounded domain with smooth boundary for
$k\geq 0$
. Let
$(v_0,p_0,\mathring {R}_0)\in C^\infty (\mathbb {R}^3\times (0,T))$
be a subsolution of the Euler equations such that
$\operatorname {\mathrm {supp}}\mathring {R}_0(\cdot ,t)\subset \overline {\Omega }_k$
for
$t\in \mathcal {I}_k$
. In addition, assume that
vanishes. Then, there exists a weak solution of the Euler equations
$v \in C^\beta _{\text {loc}}(\mathbb {R}^3\times (0,T))$
that equals
$v_0$
in
$(\mathbb {R}^3 \backslash \Omega _k)\times \mathcal {I}_k$
for any
$k\geq 0$
. In addition,
$v=v_0$
at the endpoints of the intervals
$\mathcal {I}_k$
. Furthermore,

for some universal constant C.
9.2 The iterative process
As in Proposition 3.2, we are given an initial subsolution
$(v_0,p_0,\mathring {R}_0)\in C^\infty (\mathbb {R}^3\times [0,T])$
and we will iteratively construct a sequence of subsolutions
$\{(v_q,p_q,\mathring {R}_q)\}_{q=0}^\infty $
whose limit will be the desired weak solution. To construct the subsolution at step q from the one in step
$q-1$
, we will add an oscillatory perturbation with frequency
$\lambda _q$
. Meanwhile, the size of the Reynolds stress will be measured by an amplitude
$\delta _q$
. These parameters are given by


The parameters
$a,b>1$
are very large and very close to 1, respectively. They will be chosen depending on the exponent
$0<\beta <1/3$
that appears in Theorem 9.3, on
$\Omega $
and on the initial subsolution. We introduce another parameter
$\alpha>0$
that will be very small. The necessary size of all the parameters will be discovered in the proof.
We will assume that the support of
$(v_0,p_0,\mathring {R}_0)(\cdot ,t)$
is contained in
$(0,1)^3$
. Meanwhile, the support of
$\mathring {R}$
is contained in
$\overline {\Omega }\times [0,T]$
for a potentially smaller domain
$\Omega $
with smooth boundary. The main difference with respect to Proposition 3.2 is that we also assume that
$\mathring {R}_0(\cdot ,t)$
vanishes for
$t=0$
and
$t=T$
.
It will be convenient to do an additional rescaling in our problem. In the rescaled problem the initial subsolution will depend on a, but we assume that nevertheless there exists a sequence
$\{y_N\}_{N=0}^\infty $
independent of the parameters such that



Since the initial Reynolds stress
$\mathring {R}_0$
and its derivatives vanish at
$\partial \Omega \times [0,T]$
, for any
$k\in \mathbb {N}$
there exist a constant
$C_k$
such that for any
$x\in \Omega $
we have

Note that the constants
$C_k$
are independent of a by (9.5). We define


Hence, we have

On the other hand, since
$\mathring {R}_0(\cdot ,t)$
vanishes for
$t=0$
and
$t=T$
there exists a constant
$C_t$
depending on
$\|\partial _t \mathring {R}_0\|_0$
such that

Again, the constant
$C_t$
does not depend on a because of (9.5). It depends only on the initial subsolution (before the rescaling). We define


so that

At step q the perturbation will be localized in
$A_q\times \mathcal {J}_q$
so that
$(v_q,p_q,\mathring {R}_q)$
equals the initial subsolution in
$(\mathbb {R}^3\backslash A_q)\times \left ([0,T]\backslash \mathcal {J}_q\right )$
. In this region the Reynolds stress is so small that we will ignore it. We will focus on reducing the error in
$A_q\times \mathcal {J}_q$
.
The complete list of inductive estimates is the following:




where M is a geometric constant that depends on
$\Omega $
and is fixed throughout the iterative process. The following instrumental result is key to the proof of Theorem 9.3, and is analogous to Proposition 3.2:
Proposition 9.4. There exists a universal constant M with the following property: Let
$T\geq 1$
and let
$\Omega \subset (0,1)^3\subset \mathbb {R}^3$
be a bounded domain with smooth boundary. Let
$(v_0,p_0,\mathring {R}_0)\in C^\infty (\mathbb {R}^3\times [0,T])$
be a subsolution whose support is contained in
$(0,1)^3\times [0,T]$
and such that
$\operatorname {\mathrm {supp}}\mathring {R}_0\subset \overline {\Omega }\times [0,T]$
. Furthermore, assume that (9.3)
$-$
(9.5) are satisfied for some sequence of positive numbers
$\{y_N\}_{N=0}^\infty $
. Let
$0<\beta <1/3$
and

Then, there exists an
$\alpha _0$
depending on
$\beta $
and b such that for any
$0<\alpha <\alpha _0$
there exists an
$a_0$
depending on
$\beta $
, b,
$\alpha $
,
$\Omega $
and
$\{y_N\}_{N=0}^\infty $
such that for any
$a\geq a_0$
the following holds: Given a subsolution
$(v_q,p_q,\mathring {R}_q)$
satisfying (9.11)
$-$
(9.14), there exists a subsolution
$(v_{q+1},p_{q+1},\mathring {R}_{q+1})$
satisfying (9.11)
$-$
(9.14) with q replaced by
$q+1$
. Furthermore, we have

As in Theorem 1.7, we need an auxiliary lemma to start the iterative process. This is the analogue of Lemma 3.3:
Lemma 9.5. Let
$T>0$
and let
$\Omega \subset (0,1)^3\subset \mathbb {R}^3$
be a bounded domain with smooth boundary. Let
$(v_0,p_0,\mathring {R}_0)\in C^\infty (\mathbb {R}^3\times [0,T])$
be a subsolution whose support is contained in
$(0,1)^3\times [0,T]$
and such that
$\operatorname {\mathrm {supp}}\mathring {R}_0\subset \overline {\Omega }\times [0,T]$
. Let
$\lambda>0$
be sufficiently large. There exists a subsolution
$(v,p,\mathring {R})\in C^\infty (\mathbb {R}^3\times [0,T])$
that equals the initial subsolution outside the set

and such that

where the implicit constants are independent of
$\lambda $
. Furthermore, there exist geometric constants
$K_1$
,
$K_2$
such that if
$K_1T\left \|v_0\right \|_1\leq 1$
, then

9.3 Proof of Proposition 9.4
The proof is very similar to the proof of Proposition 3.2, but we will only perturb the subsolution at times
$t\in \mathcal {J}_{q+1}$
. It will be convenient to define

The parameters
$\ell $
and
$\tau _q$
are defined as in Proposition 3.2. Let us compare the time parameters:

Note that the exponent of
$\lambda _q$
is greater than
$0$
by our assumption on b. Since we may assume that
$2d_{q+1}\leq d_q$
, we conclude that for sufficiently small
$\alpha $
and sufficiently large a we have

Hence, the temporal cutoffs that we will need to use will not be too sharp.
After these definitions, we can prove the result in a similar manner to Proposition 3.2. As before, we divide the proof in four steps:
-
1. Preparing the subsolution. The beginning of the iterative process is identical to the one in Proposition 3.2: we use a convolution kernel in space
$\psi _\ell $ to mollify
$(v_q,p_q,\mathring {R}_q)$ into
$(v_\ell ,p_\ell ,\mathring {R}_\ell )$ and then we glue it in space to
$(v_0,p_0,\mathring {R}_0)$ , obtaining a subsolution
$(\widetilde {v}_\ell ,\widetilde {p}_\ell ,\mathring {\widetilde {R}}_\ell )$ that equals
$(v_0,p_0,\mathring {R}_0)$ in
$B_2$ .
We must, however, introduce a minor modification: we add a correction
$w_L$ to ensure that
$\widetilde {v}_\ell +w_L$ has the same angular momentum as
$v_0$ . Note that
$v_q$ has the same angular momentum as
$v_0$ because they are equal at
$t=0$ and subsolutions preserve angular momentum, as argued several times. In addition, it is easy to check that mollifying does not change the total angular momentum, so
$v_\ell -v_0$ has
$0$ total angular momentum. This may not be true for
$\widetilde {v}_\ell -v_0$ . Nevertheless, since we are gluing in a (small) region where
$v_\ell $ is very similar to
$v_0$ (in terms of the parameters), the change in the angular momentum will be very small. Therefore, we may add a small correction
$w_L$ to
$\widetilde {v}_\ell $ in
$A_q$ while keeping the desired estimates. Of course, we modify the pressure and the Reynolds stress accordingly to obtain a subsolution, which we still denote as
$(\widetilde {v}_\ell ,\widetilde {p}_\ell ,\mathring {\widetilde {R}}_\ell )$ , for simplicity.
-
2. Gluing in space. In the intervals such that
$[t_i-\tau _q,t_i+\tau _q]\cap \mathcal {J}_q \neq \varnothing $ the process is exactly the same as in Proposition 3.2. In the rest of the intervals it suffices to take
$(\widetilde {v}_i,\widetilde {p}_i,\mathring {\widetilde {R}}_i)=(v_0,p_0,\mathring {R}_0)$ because outside
$\mathcal {J}_q$ we have
$|\mathring {R}|_0\leq \frac {1}{4}\delta _{q+2}\lambda _{q+1}^{-6\alpha }$ , as required by (4.19).
Regarding the estimates (4.20)
$-$ (4.23), remember that
$(v_q,p_q,\mathring {R}_q)$ equals the initial subsolution for
$t\notin \mathcal {J}_q$ . Therefore, it follows from (9.3)
$-$ (9.5) and standard estimates for mollifiers that for
$t\notin \mathcal {J}_q$ we have
$$ \begin{align*} \left\|(v_\ell-v_0)(\cdot,t)\right\|_{C^N}+\left\|(\partial_t v_\ell-\partial_t v_0)(\cdot,t)\right\|_{C^N}&\lesssim \ell^2, \\ \left\|(p_\ell-p_0)(\cdot,t)\right\|_{C^N}&\lesssim \ell^2, \\ \|(\mathring{R}_\ell-\mathring{R}_0)(\cdot,t)\|_{C^N}&\lesssim \ell^2. \end{align*} $$
The same bounds hold for
$(\widetilde {v}_\ell ,\widetilde {p}_\ell ,\mathring {\widetilde {R}}_\ell )$ . Therefore, this choice of
$(\widetilde {v}_i,\widetilde {p}_i,\mathring {\widetilde {R}}_i)$ for the rest of the intervals satisfies (4.20)
$-$ (4.23). Although (4.18) is not satisfied, its only use in the following step is to ensure that
$\widetilde {v}_i$ and
$\widetilde {v}_\ell $ have the same angular momentum. This is exactly what we did at the end of the previous step.
-
3. Gluing in time. The construction is the same as in Proposition 3.2 but due to our choice of
$(\widetilde {v}_i,\widetilde {p}_i,\mathring {\widetilde {R}}_i)$ we are actually gluing only in
$\widetilde {\mathcal {J}}_q$ (remember that
$\tau _q\ll \gamma $ ). We see that
$(\overline {v}_q,\overline {v}_q,\mathring {\,\overline {\!{R}}}_q)$ equals
$(v_0,p_0,\mathring {R}_0)$ for
$t\notin \widetilde {\mathcal {J}}_q$ and the support of
$\mathring {\,\overline {\!{R}}}_q\strut {}^{ (2)}$ is contained in
$[A_q+B(0,4\sigma )]\times \widetilde {\mathcal {J}}_q$ .
-
4. Perturbation. There are only two changes with respect to Proposition 3.2:
-
– We define the amplitudes
$\rho _{q,i}$ as
.
-
– Instead of the cutoff
$\phi _q$ we use
for some smooth cutoff
$\theta _q\in C^\infty _c(I_{q+1})$ that equals
$1$ in
$\widetilde {\mathcal {J}}_q$ . Hence,
$\widetilde {\phi }_q=1$ in the support of
$\mathring {\,\overline {\!{R}}}_q\strut {}^{ (2)}$ .
The amplitudes
$\rho _{q,i}(x,t)$ clearly satisfy the same estimates as in Proposition 3.2. In particular,
$\mathring {R}_{q,i}$ takes values in
$B(\operatorname {\mathrm {Id}},1/2)$ . On the other hand,
$\widetilde {\phi }_q a(\mathring {R}_{q,i})$ will satisfy the same estimates as
$a(\mathring {R}_{q,i})$ because we may choose
$\theta _q$ so that
$\left \vert \partial _t\theta _q\right \vert \lesssim \gamma ^{-1}\leq \tau _q^{-1}$ .
We conclude that we may carry out the same construction as in the perturbation step of Proposition 3.2 (except for fixing the energy). The new subsolution will satisfy (9.12)
$-$ (9.16). In addition, the cutoff
$\widetilde {\phi }_q$ ensures that the support of the perturbation is contained in
$A_{q+1}\times \mathcal {J}_{q+1}$ , as required by the inductive hypothesis (9.11).
Finally, we emphasize an important difference: the constant M in this case is universal. In Proposition 3.2 it depended on
$\Omega $ because so did the amplitudes
$\rho _{q,i}$ . Since here they are independent of
$\Omega $ , arguing as in [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7] yields a universal M.
-
9.4 Proof of Lemma 9.5
We fix a cutoff
$\theta \in C^\infty _c((0,T))$
such that

Since
$\mathring {R}_0$
vanishes at
$t=0$
and
$t=T$
, it is clear that
$|(1-\theta )\mathring {R}_0|\lesssim \lambda ^{-1/2}$
.
We then carry out the same construction as in Lemma 3.3 but replacing the cutoff
$\phi (x)$
by
. This ensures that the perturbation vanishes if
$t\notin (\lambda ^{-1/3},T-\lambda ^{-1/3})$
. It does worsen the estimates, but it is not catastrophic. The most significant change is that when we write the term in brackets in (7.28) as
$\sum _{k\neq 0}c_k e^{i\lambda k\cdot \Phi }$
, the vectors
$c_k$
now satisfy the estimates

Applying the stationary phase lemma now yields

which, continuing the construction in Lemma 3.3, yields the desired bound for
$\mathring {R}$
. Since the exponent is actually smaller than
$-1/2$
, we can expend the extra factor in beating any geometric constant.
Finally, let us derive a precise estimate for
$v-v_0$
. Recall that

The last term can be bounded by a geometric constant depending on the compact subset
$\mathcal {N}$
used when applying Lemma 7.2. On the other hand, by standard estimates for the transport equation there exists a geometric constant
$K_1$
such that

By our assumption on
$T\left \|v_0\right \|_1$
, we have
$\left \|\nabla \Phi -\operatorname {\mathrm {Id}}\right \|_0\leq 1/2$
, so
$\|(\nabla \Phi )^{-1}\|_0\leq 2$
. We conclude that

for some geometric constant
$K_2$
. Since
$w_c+w_L=O(\lambda ^{-1})$
, it is clear that the required estimate holds for sufficiently large
$\lambda $
.
9.5 Proof of Theorem 9.3
To simplify the notation of this proof, given a map f defined in
$\mathbb {R}^3\times I$
for some interval I, we will write

By dividing in half the intervals
$\mathcal {I}_k$
as many times as necessary, we may assume that

where
$\left \vert \mathcal {I}_k\right \vert $
is the length of the interval
$\mathcal {I}_k$
and
$K_1$
is the constant that appears in Lemma 9.5.
We apply Lemma 9.1 at the endpoints of each interval
$\mathcal {I}_k$
, obtaining a new subsolution
$(\widetilde {v}_0,\widetilde {p}_0,\mathring {\widetilde {R}}_0)$
in which the Reynolds stress vanishes at the endpoints of all the intervals
$\mathcal {I}_k$
. In addition, we have
$(\widetilde {v}_0,\widetilde {p}_0,\mathring {\widetilde {R}}_0)=(v_0,p_0,\mathring {R}_0)$
outside
$\Omega _k$
for
$t\in \mathcal {I}_k$
. By taking s sufficiently small in each application of Lemma 9.1, we may assume that

Once we have a subsolution in which the Reynolds stress vanishes at the endpoints of the intervals
$\mathcal {I}_k$
, we may work in each of them independently. Indeed, our constructions keep the subsolution fixed near the endpoints of the intervals.
We fix
$k\geq 0$
, and we fix a sequence of positive numbers
$\{y_{N,k}\}_{N=0}^\infty $
such that



We choose b satisfying (9.15) and we choose
$\alpha $
smaller than the threshold given by Proposition 9.4. Let
$a_{0,k}$
be the threshold given by Proposition 9.4 when applied to
$\Omega _k$
and our sequence
$\{y_{N,k}\}_{N=0}^\infty $
. For
$a_k\geq a_{0,k}$
we consider the parameters
$\lambda _{q,k}$
and
$\delta _{q,k}$
defined as in (9.1) and (9.2).
Taking
$a_k$
larger if necessary, we apply Lemma 9.5 with the parameter
$\lambda _{1,k}^{12\alpha }$
, obtaining a subsolution
$(v_1,p_1,\mathring {R}_1)$
that equals
$(\widetilde {v}_0,\widetilde {p}_0,\mathring {\widetilde {R}}_0)$
outside

and satisfying the following estimates:

where the constants
$C_{N,k}$
are independent of
$\lambda _{1,k}$
but they will depend on
$\Omega _k$
,
$\mathcal {I}_k$
and the initial subsolution. Furthermore, we have

where we have used that
$K_1\left \vert \mathcal {I}_k\right \vert \left \|\widetilde {v}_0\right \|_{1,\mathcal {I}_k}\leq 1$
.
Next, we consider the scale invariance of the Euler equations and subsolutions:

We choose
$\Gamma = \delta _{2,k}^{1/2}$
and we begin to work in this rescaled setting, which we will indicate with a superscript r. Note that
$(\widetilde {v}_0^r,\widetilde {p}_0^r,\mathring {\widetilde {R}}{}^r_0)$
still satisfies (9.3)
$-$
(9.5) with the same sequence
$\{y_{N,k}\}_{N=0}^\infty $
. Regarding
$(v_1^r,p_1^r,\mathring {R}_1^r)$
, it follows from the definition of the rescaling that

On the other hand, since the constants
$C_{N,k}$
are independent of
$\lambda _{1,k}$
, for sufficiently large
$a_k$
we have

Finally,
$(v_1^r,p_1^r,\mathring {R}_1^r)=(\widetilde {v}_0^r,\widetilde {p}_0^r,\mathring {\widetilde {R}}{}^r_0)$
outside

Let us consider the sets
$A_{q,k}$
and
$\mathcal {J}_{q,k}$
that we obtain when we apply the definition of (9.7) and (9.9) to
$\Omega _k$
and
$\mathcal {I}_k^r$
. We see that
$(v_1^r,p_1^r,\mathring {R}_1^r)=(\widetilde {v}_0^r,\widetilde {p}_0^r,\mathring {\widetilde {R}}{}^r_0)$
outside
$A_{1,k}\times \mathcal {J}_{1,k}$
for sufficiently small
$\alpha $
and sufficiently large
$a_k$
.
We conclude that
$(v_1^r,p_1^r,\mathring {R}_1^r)$
satisfies the inductive hypotheses (9.11)
$-$
(9.14) in the interval
$\mathcal {I}_k^r$
with initial subsolution
$(\widetilde {v}_0^r,\widetilde {p}_0^r,\mathring {\widetilde {R}}{}^r_0)$
. In addition, for
$t\in \mathcal {I}_k^r$
the initial subsolution satisfies (9.3)
$-$
(9.5) with the sequence
$\{y_{N,k}\}_{N=0}^\infty $
and the support of
$\mathring {\widetilde {R}}{}^r_0(\cdot ,t)$
is contained in
$\overline {\Omega }_k$
. Finally, by taking
$a_k$
even larger, we may assume that
$\left \vert \mathcal {I}_k^r\right \vert \geq 1$
.
Applying Proposition 9.4 in each interval
$\mathcal {I}_k^r$
and undoing the scaling, we obtain a sequence of subsolutions
$\{(v_q,p_q,\mathring {R}_q)\}_{q=1}^\infty \in C^\infty (\mathbb {R}^3\times (0,T))$
that equal
$(v_0,p_0,0)$
in
$(\mathbb {R}^3\backslash \Omega _k)\times \mathcal {I}_k$
for any
$k\geq 0$
. In addition, for
$t\in \mathcal {I}_k$
we have:


We see that
$v_q$
converges uniformly in compact subsets of
$\mathbb {R}^3\times (0,T)$
to some continuous map v. On the other hand, note that the pressure is the only compactly supported solution of

Therefore,
$p_q$
also converges to some pressure
$p\in L^{s}(\mathbb {R}^3)$
for any
$1\leq s<\infty $
. Since
$\mathring {R}_q$
converges to
$0$
uniformly in compact subsets of
$\mathbb {R}^3\times (0,T)$
, we conclude that
$(v,p)$
is a weak solution of the Euler equations in
$\mathbb {R}^3\times (0,T)$
.
Furthermore, using (9.21) we obtain

so
$\{v_q\}_{q=1}^\infty $
is uniformly bounded in
$C^0_tC^{\beta '}_x$
in any compact subset
$I\subset (0,T)$
for all
$\beta '<\beta $
. Arguing as in [Reference Buckmaster, De Lellis, Székelyhidi and Vicol7] we obtain (local) time regularity. We conclude that
$v\in C^{\beta "}_{\text {loc}}(\mathbb {R}^3\times (0,T))$
, with
$\beta "<\beta '<1/3$
arbitrary.
Finally, we compute the difference between
$v_0$
and v. We write
$b=1+\gamma $
, so that
$b^q-1\geq \gamma q$
. Taking
$a_k$
sufficiently large, we have

Thus, by further increasing
$a_k$
, we have
$\left \|v-v_1\right \|_{0,\mathcal {I}_k}\leq \|\mathring {R}_0\|_{0,\mathcal {I}_k}$
. We conclude

10 Vortex sheet
This section is devoted to the proof of Theorem 1.4. The key ingredient is Theorem 9.3, which is applied to an initial subsolution with the appropriate behavior. We fix a parameter
$0<\lambda <(4T)^{-1}$
. We choose an even function
$f\in C^\infty _c((-1,1))$
such that
$\int f=2$
and we define

Let
$v_0$
and
$R_0$
be the periodic extension of

Direct computation shows that
$\partial _t v_0=\operatorname {\mathrm {div}} \mathring {R}_0$
. It is also clear that
$v_0\cdot \nabla v_0=0$
. Therefore, the triplet
$(v_0,0,\mathring {R}_0)$
is a subsolution.
Since f is even, the support of G is contained in
$(-1,1)$
. For
$k\geq 0$
let us define
and

We see that
$(v_0,0,\mathring {R}_0)$
equals
$(u_0,0,0)$
outside
$\Omega _k$
for
$t\in \mathcal {I}_k$
. Since
, we may apply Theorem 9.3, obtaining a weak solution of the Euler equations
$v\in C^\beta _{\text {loc}}(\mathbb {T}^3\times (0,T))$
that equals
$(u_0,0,0)$
outside
$\Omega _k$
for
$t\in \mathcal I_k$
(in fact, for
$t\in (0,2^{-k}T)$
because
$\Omega _j\subset \Omega _k$
for all
$j\leq k$
). In particular, the initial datum is
$u_0$
. Furthermore, there exists a universal constant C such that for any
$k\geq 0$
we have

Let us estimate the energy. First, we define

Note that
$\delta>0$
because the continuous function F takes values in
$[0,2]$
and it satisfies
$F(0)=1$
, since f is even. We compute

where we have used that
$v_0=u_0$
except close to
$x_3=1/4$
and
$x_3=3/4$
. Next, we write

Let us choose
$\lambda $
sufficiently small so that
$\lambda < (32C')^{-1}\delta $
, where
$C'$
is the constant in (10.1). Hence, for
$t\in \mathcal {I}_k$
we have

where we have used that
$v=v_0$
outside
$\Omega _k$
for
$t\in \mathcal {I}_k$
. By definition of
$\mathcal {I}_k$
, we see that
$2^{-(k+1)}T\leq t$
for any
$t\in \mathcal {I}_k$
. We conclude that

Therefore, the weak solution v is admissible. In addition, we see that we obtain a sequence of different solutions
$\{v_i\}_{i=i_0}^\infty $
by repeating the construction with
$\lambda _i=4^{-i}$
for sufficiently large
$i_0$
.
11 Blowup
In this final section we prove Theorem 1.5 on the existence of Hölder continuous weak solutions of the 3d Euler equations that exhibit a singular set of maximal dimension. The proof makes use of some technical lemmas that are presented in Subsections 11.1 and 11.2.
11.1 Building blocks
The fundamental element in our construction is the following simple blowup, whose proof is postponed to Subsection 11.2:
Lemma 11.1. Let
$0<\beta <1/3$
and let
$q>2$
. Let
$a\in C^\infty (\mathbb {R},\mathbb {R}^3)$
be a bounded map. Given
$\varepsilon>0$
, there exists a weak solution of the Euler equations
$(v_\varepsilon ,p_\varepsilon )$
in
$\mathbb {R}^3\times \mathbb {R}$
such that:
-
•
$(v_\varepsilon ,p_\varepsilon )= (a,-\partial _t a\cdot x)$ outside
$B(0,\varepsilon )\times (0,\varepsilon )$ ,
-
• the q-singular set of
$v_\varepsilon $ is
$\mathscr {S}^q_{v_\varepsilon }=\{(0,\varepsilon )\}$ ,
-
•
$v_\varepsilon \in C^\beta _{\text {loc}}(\mathbb {R}^3\times \mathbb {R}\backslash \mathscr {S}^q_{v_\varepsilon })$ ,
-
• the relative energy
is continuous and so is the map
$t\mapsto \int a\cdot (v_\varepsilon -a)dx$ .
Furthermore, there exists a constant
$C>0$
depending on
$\left \|a\right \|_{L^\infty }<\infty $
but not on
$\varepsilon $
such that

Once we known how to construct a single blowup, as stated in the previous lemma, we will use the following result to glue many of them together. Its proof is completely independent of Lemma 11.1.
Lemma 11.2. Let
$a\in C^\infty (\mathbb {R},\mathbb {R}^3)$
. Let
$\{(v_i,p_i)\}_{i=0}^\infty $
be a sequence of weak solutions of the Euler equations in
$\mathbb {R}^3\times \mathbb {R}$
. Suppose that the sets
$F_i$
given by the closure of

are pairwise disjoint and that

Then, the pair
$(v,p)$
given by

is a weak solution of the Euler equations.
Proof. The hypotheses readily yield
$v\in L^2_{\text {loc}}(\mathbb {R}^4)$
and
$p\in L^1_{\text {loc}}(\mathbb {R}^4)$
. Furthermore, it is easy to see that any partial sum

is a subsolution. Indeed, fix
$\chi _i\in C^\infty (\mathbb {R}^4)$
such that
$\chi _i^{-1}(\{0\})=F_i$
and consider
and
. They are well defined because
$F_1\cap F_2=\varnothing $
. We see that
$\theta _1$
vanishes in
$F_2$
and
$\theta _2$
vanishes in
$\theta _1$
, so for any
$\phi \in C^\infty _c(\mathbb {R}^4,\mathbb {R}^3)$
we have

which follows from the definition of
$(v_i,p_i)$
being a weak solution of the Euler equations using the test-function
$\theta _i\phi $
. An analogous argument proves that
$\widetilde {v}_2$
is weakly divergence-free. Therefore,
$(\widetilde {v}_2,\widetilde {p}_2)$
is a weak solution of the Euler equations. Furthermore, iterating this argument we conclude that
$(\widetilde {v}_k,\widetilde {p}_k)$
is a weak solution of the Euler equations for any
$k\geq 1$
. Since
$\widetilde {v}_k$
converges to v in
$L^2(K)$
and
$\widetilde {p}_k$
converges to p in
$L^1(K)$
for any compact subset
$K\subset \mathbb {R}^4$
, it is easy to see that
$(v,p)$
is a weak solution of the Euler equations.
11.2 Proof of Lemma 11.1
Note it suffices to prove the result for
$\varepsilon =1$
because for
$\varepsilon \neq 1$
we could simply take

with a rescaled accordingly. It is clear that this scaling preserves the q-singular set, that is,

Furthermore, since
$p_1+\partial _t a\cdot x$
will be the only compactly supported solution of

standard Calderón-Zygmund estimates yield

Thus, to prove (11.1) it suffices to show that
$(v_1-a)\in L^\infty _tL^2_x$
.
From now on, we will assume
$\varepsilon =1$
and drop the subscript
$1$
for simplicity. By [Reference Gavrilov32] there exists a nontrivial steady solution of the Euler equations with compact support
$(u,\pi )\in C^\infty _c(\mathbb {R}^3)$
. By rescaling, we may assume that its support is contained in the ball
$B(0,1/4)$
. We define

Therefore,
$(U,P)\in C^\infty _c(B(0,1))$
is a nontrivial steady solution of the Euler equations such that

Hence, by Lemma 2.9 there exists
$S_0\in C^\infty _c(B(0,1),\mathcal {S}^3)$
such that
$\operatorname {\mathrm {div}} S_0=U$
. We introduce a parameter
$\alpha \in (-1,0)$
that will be fixed later and we define

Since
$\operatorname {\mathrm {div}}(x\otimes U+U\otimes x)=4U+x\cdot \nabla U$
and U is divergence-free, we have

Furthermore, we see that the triplet
$(U,P,S)$
satisfies

Consider the self-similar ansatz

It follows from Equation (11.3) that the triplet
$(\widetilde {v}_0,\widetilde {p}_0,\widetilde {R}_0)$
is a subsolution of the Euler equations in
$\mathbb {R}^3\times [0,1)$
(in which the Reynolds stress is not normalized to be trace-free, yet). Regarding the scaling, we see that

We choose

which ensures that

Next, we fix
$\chi \in C^\infty ([0,1])$
that vanishes in a neighborhood of
$0$
and is identically
$1$
in a neighborhood of
$1$
. Since
$(\widetilde {v}_0,\widetilde {p}_0,\widetilde {R}_0)$
is a subsolution, is is easy to see that the triplet
$(v_0,p_0,R_0)$
given by

is also a subsolution and that vanishes. Let
and
. We see that the support of
$R_0(\cdot ,t)$
is contained in
$\Omega _k$
for
$t\in \mathcal {I}_k$
. In fact, so is the support of
$(v_0,p_0,R_0)(\cdot ,t)$
. In addition, we see that

where the implicit constant depends on
$\left \|a\right \|_{L^\infty }$
but not on k.
We would like to apply Theorem 9.3. While our current situation does not exactly meet the hypotheses of Theorem 9.3, this is not really an issue. In spite of the fact that our subsolution is not compactly supported, what matters is the support of the Reynolds stress, as argued in Subsection 3.3. Furthermore, although
$R_0$
is not trace-free, this can be easily solved in the proof of Theorem 9.3. After applying Lemma 9.1, one simply has to replace
$(\widetilde {v}_0,\widetilde {p}_0,\widetilde {R}_0)$
by

Therefore, there exists a weak solution
$(v,p)$
of the Euler equations in
$\mathbb {R}^3\times (0,1)$
with
$v\in C^\beta _{\text {loc}}(\mathbb {R}^3\times (0,1))$
and
$(v,p)=(a,0)$
in
$(\mathbb {R}^3\backslash \Omega _k)\times \mathcal {I}_k$
for
$k\geq 0$
. Furthermore, since
$(v_0,p_0,R_0)=(a,0,0)$
for t sufficiently close to
$0$
, a careful revision of Theorem 9.3 will convince us that
$(v,p)=(a,0)$
in a neighborhood of
$t=0$
. Thus, we may extend
$(v,p)$
to the interval
$(-\infty ,1)$
.
On the other hand, for
$t\in \mathcal {I}_k$
we have

where the constant changes after each inequality and it is allowed to depend on
$\left \|a\right \|_{L^\infty }$
but not on k. Hence,

Our choice of
$\alpha $
ensures that both quantities go to
$0$
when
$k\to \infty $
. We conclude that the maps
$t\mapsto \left \|v(\cdot ,t)-a\right \|^2_{L^2(\mathbb {R}^3)}$
and
$t\mapsto \int a\cdot (v_\varepsilon -a)dx$
can be extended by 0 to a continuous function in the whole
$\mathbb {R}$
.
Next, we show that we may extend our weak solution to
$\mathbb {R}^3\times \mathbb {R}$
by setting
$(v,p)=(a,0)$
for
$t\geq 1$
. For simplicity, we still denote this extension as
$(v,p)$
. It will be a weak solution in
$\mathbb {R}^3\times \mathbb {R}$
if and only if for any solenoidal test-function
$\phi \in C^\infty _c(\mathbb {R}^3\times \mathbb {R})$
we have

We split the integral:

where we have used that, for a fixed time t,
$a(t)$
is just a constant vector and
$\partial _t\phi (\cdot ,t)$
is a compactly supported divergence-free. Thus, the spatial integral vanishes for each time t. We conclude that
$(v,p)$
will be a weak solution in
$\mathbb {R}^3\times \mathbb {R}$
if and only if

We choose a cutoff function
$\theta \in C^\infty (\mathbb {R})$
that equals
$0$
if
$t\leq 0$
and equals
$1$
if
$t\geq 1/2$
. For
$j\geq 1$
consider
. Since
$(1-\theta _j)\phi \in C^\infty _c(\mathbb {R}^3\times (-\infty ,1))$
is solenoidal and
$(v,p)$
is a weak solution in
$\mathbb {R}^3\times (-\infty ,1)$
, we see that

Let us study the second term on the right-hand side. We fix
$t\in \mathcal {I}_k$
and we write

Taking into account that

and (11.4), we surmise that for
$t\in \mathcal {I}_k$

Therefore,

because
$3+5\alpha>0$
by our choice of
$\alpha $
. Regarding the first term on the right-hand side of (11.6), we split the integral into

Again, the first term vanishes because if we keep t fixed
$a(t)$
is just a constant vector and
$\partial _t(\theta _j\phi )(\cdot ,t)$
is a compactly supported divergence-free field. Thus, the spatial integral vanishes for each time t. Concerning the second term, we estimate

Hence,

where the constant may depend on a but it is independent of j. Since
$3+4\alpha>0$
and
$j\geq 1$
is arbitrary, we conclude that (11.5) holds, so
$(v,p)$
is a weak solution in
$\mathbb {R}^3\times \mathbb {R}$
.
Regarding the q-singular set, fix a spatial ball
$B(0,r)$
and let
for
$k\geq 1$
. For sufficiently large k we have
$\chi (t_k)=1$
and, since the velocity is unchanged at the endpoints of the intervals
$\mathcal {I}_k$
, we have

We compute

where we have used that for sufficiently large k the ball
$B(0,2^{(1+\alpha )k}r)$
contains the support of U. By our choice of
$\alpha $
we have
$\alpha +3(1+\alpha )/q<0$
, so

Since the ball
$B(0,r)$
is arbitrary, we see that
$(0,1)\in \mathscr {S}^q_v$
.
Finally, since
$v\in C^\beta _{\text {loc}}(\mathbb {R}^3\times (-\infty ,1))$
and
$v=a$
in
$(\mathbb {R}^3\backslash B(0,2^{-k}))\times (1-2^{-k},1)$
for any
$k\geq 0$
, we conclude that
$v\in C^\beta _{\text {loc}}(\mathbb {R}^3\times \mathbb {R}\backslash \{(0,1)\})$
. In particular, the q-singular set reduces to
$\{(0,1)\}$
. This completes the proof of the lemma.
11.3 Proof of Theorem 1.5
After a temporal rescaling, we may assume that
$T=1$
. After a translation, we may assume that
$\overline {B}(0,4\rho )$
is contained in U for sufficiently small
$\rho>0$
. By subtracting a time dependent constant, we may assume that
$p_0(0,t)=0$
. Let
$a(t)=v_0(0,t)$
. We glue
$(v_0,p_0)$
and
$(a,-\partial _ta\cdot x)$
using Lemma 2.11, obtaining a subsolution
$(v_1,p_1,\mathring {R}_1)$
such that

It is not difficult to deduce from Lemma 2.14 that by reducing
$\rho $
we can obtain
$\left \|v_1-v_0\right \|_0$
and
$\rho ^3\|\mathring {R}_1\|_0^{1/2}$
arbitrarily small.
We apply Theorem 1.7 to obtain a weak solution of the Euler equations
$(v_2,p_2)$
that equals
$(v_0,p_0)$
outside
$B(0,4\rho )\times [0,1]$
and
$(a,-\partial _t a\cdot x)$
in
$\overline {B}(0,3\rho )\times [0,1]$
. Note that we may choose a nonincreasing energy profile that is arbitrarily close to the original one but still satisfies (1.3) because
$\rho ^3\|\mathring {R}_1\|_0^{1/2}$
is arbitrarily small. Since
$\left \|v_1-v_0\right \|_0$
is arbitrarily small, we conclude that
$\left \|v_2-v_0\right \|_0$
may be chosen to be arbitrarily small.
Next, we construct the blowup in
$B(0,\rho )\times (0,1]$
. By [Reference Besicovitch and Ursell4], for any
$k\geq 1$
there exists a function
$f^k\in C^{2^{-k}}(B(0,\rho ))$
taking values in
$[1-2\cdot 4^{-k},1-4^{-k}]$
and whose graph
$G^k$
has Hausdorff dimension
$4-2^{-k}$
. We want the q-singular set
$\mathscr {S}^q$
of the final solution to contain all of these
$G^k$
so that its Hausdorff dimension is
$4$
. To do that, we choose a sequence
$\{x_i\}_{i=1}^\infty $
dense in
$B(0,\rho )$
and we denote
. We see that
$\{(x_i,\tau _i^k)\}_{i=1}^\infty $
is dense in
$G^k$
. Using Lemma 11.1 we will construct a sequence of blowups converging to each of the points
$(x_i,\tau _i^k)$
. Thus, these blowups would accumulate at every point in
$\bigcup _{k\geq 1}G^k$
, which means that this set will be contained in the singular set, as we wanted.
In order to glue the blowups given by Lemma 11.1 using Lemma 11.2, they must have disjoint supports. Hence, we have study the geometry of the situation. Let

so that the sequence
$\{t_{ij}^k\}_{j=1}^\infty $
is contained in
$(1-4^{-(k-1)},1-4^{-k})$
and converges to
$\tau _i^k$
. We want to apply Lemma 11.1 to construct a blowup in

It is clear that choosing
$\varepsilon _{ij}^k$
sufficiently small ensures that the sets
$\overline {U}{}_{ij}^k$
are disjoint for a fixed i and k, but it is not so clear if i is not kept fixed.
We will try to isolate the sequence corresponding to a fixed i. Let and consider the sets

By the definition of
$L^k$
, for any
$i\neq i'\geq 1$
we have
$(x_{i},\tau _{i}^k)\notin \mathcal {C}_{i'}^k$
. We define
$j_0(1)=1$
and for
$i>1$
we define
$j_0(i)$
to be the minimum
$j_0\geq 1$
such that

As we have mentioned,
$(x_{i},\tau _{i}^k)\notin \mathcal {C}_{i'}^k$
for
$i'<i$
, so there exists a neighborhood of
$(x_{i},\tau _{i}^k)$
disjoint from the union of these sets. Since
$\mathcal {C}^k_i\cap [\mathbb {R}^3\times (t^k_{ij_0},\tau ^k_i)]$
will be contained in this neighborhood of
$(x_{i},\tau _{i}^k)$
for sufficiently large
$j_0$
,
$j_0(i)$
is well defined. The point is that we will only add blowups for
$j\geq j_0(i)$
.
Next, let us define

where
$0<\delta \leq 1$
will be chosen later. Since
$\varepsilon _{ij}^k\leq 4^{-(k+j+1)}$
, we see that the
$\{U^k_{ij}\}_{j=1}^\infty $
are pairwise disjoint for fixed
$k,i$
. Furthermore, this definition ensures that

which means that
$U_{ij}^k\subset \mathcal {C}_i^k$
. Therefore, it follows from the definition of
$j_0(i)$
that the sets

are pairwise disjoint, as claimed.
Let
$(v_{ij}^k,p_{ij}^k)$
be the weak solution of the Euler equations given by Lemma 11.1 using the parameter
$\varepsilon ^k_{ij}$
. After a translation, we may assume that the set where
$(v_{ij}^k,p_{ij}^k)\neq (a,-\partial _t a\cdot x)$
is contained in
$U^k_{ij}$
. We define

By Lemma 11.2, the pair
$(v_3,p_3)$
is a weak solution of the Euler equations in
$\mathbb {R}^3\times \mathbb {R}$
. Note that any point
$(x,t)$
not in the closure of
$\bigcup _{k\geq 1}G^k$
has a neighborhood V that intersects only a finite number of the
$U^k_{ij}$
. We conclude that the q-singular set of the weak solution is the closure of the union of the q-singular sets of the
$v_{ij}^k$
, that is,

and
$v\in C^\beta _{\text {loc}}(\mathbb {R}^4\backslash \mathscr {S}^q)$
. Regarding the energy, let us consider the partial sums

Taking into account the identity
$v^2=(v-a)^2+a^2+2a\cdot (v-a)$
, we may write

because the support of the
$(v_{i'j'}^{k'}-a)$
are pairwise disjoint. We see that
$e_{ij}^k$
is continuous by Lemma 11.1. Furthermore, by Equation (11.1) and our choice of
$\varepsilon _{ij}^k$
it converges uniformly to
$\int _{B(0,2\rho )}\left \vert v_4(x,t)\right \vert {}^2dx$
which is, therefore, continuous. In addition, it can get arbitrarily close to
$\frac {32}{3}\pi \rho ^3a(t)^2$
by reducing
$\delta>0$
.
Let us glue this blowup to
$(v_2,p_2)$
. Since
$x_i\in B(0,\rho )$
and we may assume that
$\varepsilon _{ij}^k\leq \rho $
by further reducing
$\delta $
, we see that
$U_{ij}^k\subset B(0,2\rho )\times (0,1)$
, so
$(v_4,p_4)=(a,-\partial _t a\cdot x)$
outside
$B(0,2\rho )\times (0,1)$
. Hence, it glues well with
$(v_2,p_2)$
. We conclude that there exists a weak solution of the Euler equations
$(v_4,p_4)$
that equals
$(v_0,p_0)$
outside
$B(0,4\rho )\times (0,1)$
and has a q-singular set
$\mathscr {S}^q\subset \overline {B}(0,\rho )\times (0,1]$
with Hausdorff dimension
$4$
. In addition,
$v\in C^\beta _{\text {loc}}((\mathbb {R}^3\times [0,1])\backslash \mathscr {S}^q)$
. Furthermore,
$\left \|v_4(\cdot ,0)-v_0(\cdot ,0)\right \|_{C^0}$
is arbitrarily small and
$t\mapsto \int \left \vert v_4(x,t)\right \vert {}^2dx$
is continuous and arbitrarily close to
$t\mapsto \int \left \vert v_0(x,t)\right \vert {}^2dx$
.
To complete the proof it suffices to modify
$(v_4,p_4)$
in
$(B(0,3\rho )\backslash \overline {B}(0,2\rho ))\times [0,1]$
to ensure that the energy profile is nonincreasing. Let
and fix a nonincreasing function
$e(t)>e_4(t)$
. It is easy to obtain a sequence of strictly positive smooth functions
$\{\delta _k\}_{k=1}^\infty $
whose sum is
$e(t)-\widetilde {e}(t)$
and such that
$\left \|\delta _k\right \|_{L^\infty }\lesssim 2^{-k}$
. Let
. By reducing
$r_0$
if necessary, we can find a sequence of pairwise disjoint balls
$\overline {B}(x_k,r_k)\subset B(0,3\rho )\backslash \overline {B}(0,2\rho )$
.
Fix
$k\geq 1$
and let
. We use Theorem 1.7 to construct a weak solution of the Euler equations
$(u_k,\pi _k)$
that equals
$(a(2^{-k/3}t),-(\partial _t a)(2^{-k/3}t)\cdot x)$
for
$x\notin B(0,r_0\rho )$
and such that
$\int _{B(0,r_0\rho )}\left \vert u_k\right \vert {}^2dx=e_k(t)$
. In addition,
$v\in C^\beta $
.
Finally, we consider

The pair
$(v_5,p_5)$
is a weak solution of the Euler equations that equals
$(a,-\partial _t a\cdot x)$
for
$x\notin B(0,3\rho )\backslash \overline {B}(0,2\rho )$
and
$v_5\in C^\beta $
. Regarding the energy

We glue
$(v_5,p_5)$
to
$(v_4,p_4)$
, obtaining the desired weak solution of the Euler equations
$(v_6,p_6)$
. Note that
$\left \|(v_6-v_0)(\cdot ,0)\right \|_{C^0}$
can be taken to be arbitrarily small because we can do so with
$\left \|v_2-v_0\right \|_0$
and
$e-\widetilde {e}$
. This completes the proof.
A Hölder and Besov spaces
Let
$\Omega $
be an open subset of Euclidean space. We denote by
$C^0(\Omega )$
the set of bounded continuous functions on
$\Omega $
, which we equip with the supremum norm, denoted by
. More generally, for any
$N\geq 0$
we define the space
$C^N(\Omega )$
as the set of functions that have bounded continuous derivatives of any order
$k\leq N$
. On this space, we define the following seminorms and norms, respectively:

Here
$\beta $
denotes a multi-index and
$\left \vert \beta \right \vert $
denotes its length. Given
$N\geq 0$
and
$\alpha \in (0,1)$
, we define the Hölder space
$C^{N+\alpha }(\Omega )$
as the set of functions
$f\in C^N(\Omega )$
such that the following quantity is finite:

This set becomes a Banach space when equipped with the following norm:

When we work in a subset
$E\subset \Omega $
instead of the whole
$\Omega $
, we will write
$\left \|\cdot \right \|_{N;E}$
. When the functions also depend on time, we also take the supremum in
$t\in [0,T]$
.
The Hölder norms satisfy the following inequalities:

for
$0\leq s \leq t$
and all
$\varepsilon>0$
, and

for
$0\leq s\leq 1$
. The constant C only depends on the Hölder exponents involved and on the domain
$\Omega $
. For the applications in this article, since f will be a compactly supported function defined on the whole
$\mathbb R^n$
, C will just depend on the Hölder exponents. From (A.1) with
$\varepsilon =\left \|f\right \|_0^{1/r}[f]_r^{-1/r}$
we obtain the following interpolation inequality:

Let
$\beta $
by a multi-index. By induction on
$\left \vert \beta \right \vert $
and the rule for the first derivative of a product, one easily deduces

from which it immediately follows that

The following proposition is standard:
Proposition A.1. Let
$N\in \mathbb {N}$
and
$\alpha \in (0,1)$
. Let
$f\in C^{ N,\alpha }_c(\mathbb {R}^m,\mathbb {R})$
and
$F\in C^{ N,\alpha }_c(\mathbb {R}^m,\mathbb {R}^m)$
. There exists a constant
$C=C(N,\alpha )$
such that the potential-theoretic solutions of

satisfy

Note that in the previous proposition, one does not get any information about the
$C^\alpha $
norm of the solution, aside from estimating it by higher-order norms. For this, we will need to introduce negative regularity spaces. Let us consider a Littlewood–Payley decomposition, for example, as in [Reference Bahouri, Chemin and Danchin2, Section 2.2]. For this, we take smooth radial functions
$\chi ,\varphi :\mathbb {R}^3\to [0,1]$
, whose supports are contained in the ball
$B(0,\frac 43)$
and in the annulus
$\{\frac 34 <|\xi |<\frac 83\}$
respectively, with the property that

for all
$\xi \in \mathbb {R}^3$
. In terms of the Fourier multipliers
$P_{<}:=\chi (D)$
and
$P_N:=\varphi (2^{-N}D)$
, the Besov norm
$B^s_{\infty ,\infty }$
(which is equivalent to the Hölder norm
$C^s$
if
$s\in \mathbb {R}^+\backslash \mathbb {N}$
, and strictly weaker if
$s\in \mathbb {N}$
) can be written as

Here
$s\in \mathbb {R}$
is any real number. Again, when dealing with time-dependent functions, we consider the supremum in time of
$\|f(t)\|_{B^{s}_{\infty ,\infty }}$
.
B Some auxiliary estimates
The first lemma of this appendix shows that we can find a cutoff function with well-behaved bounds on its derivatives:
Lemma B.1. Let
$A\subset \mathbb {R}^n$
be a measurable set and let
$r>0$
. There exists a cutoff function
$\varphi _r\in C^\infty (\mathbb {R}^n,[0,1])$
whose support is contained in
$A+B(0,r)$
and such that
$\varphi _r\equiv 1$
in a neighborhood of A. Furthermore, for any
$N\geq 0$
we have

Proof. We choose a nonnegative function
$\psi \in C^\infty _c(B(0,1))$
such that
$\int \psi =1$
. For
$\varepsilon>0$
we denote

Note that
$\int \psi _\varepsilon =1$
. The desired cutoff function is:

It is easy to see that its support is contained in
$A+B(0,r)$
and that
$\varphi _r\equiv 1$
in a neighborhood of A. Furthermore, it is smooth and

Hence,

from which the result follows.
The second instrumental lemma provides a bound for the solution to a transport equation. The proof is standard, see, for example, [Reference Buckmaster, De Lellis, Isett and Székelyhidi6].
Lemma B.2. Let
$f\in C^\infty (\mathbb {R}^3\times \mathbb {R})$
be the solution of the transport equation

for some vector field
$v\in C^\infty (\mathbb {R}^3\times \mathbb {R},\mathbb {R}^3)$
and
$g\in C^\infty (\mathbb {R}^3\times \mathbb {R})$
. Then, for
$0\leq \alpha \leq 1$
and
$\left \vert t\right \vert \left \|v\right \|_1\leq 1$
we have

In this article we also need and extension theorem for Hölder continuous functions defined on a domain
$\Omega $
:
Theorem B.3. Let
$N\geq 0$
and
$\alpha \in (0,1)$
. Let
$\Omega \subset \mathbb {R}^n$
be a domain with smooth boundary. Then, there exists a linear map
$T:C^{N+\alpha }(\Omega )\to C^{N+\alpha }(\mathbb {R}^n)$
such that
-
•
$Tf=f$ on
$\Omega $ for each
$f\in C^{N+\alpha }(\Omega )$ and
-
• the norm of T is bounded by a constant depending only on
$\Omega $ and N.
Proof. The proof is essentially [Reference Gilbarg and Trudinger33, Lemma 6.37] and is based on a rectification of the boundary and a reflection. We must, however, make some remarks. In [Reference Gilbarg and Trudinger33] they assume
$N\geq 1$
because they are considering sets with less regular boundaries. In the case of a smooth boundary, the result also holds for
$C^\alpha $
functions. We must warn the reader that they are using the notation
$C^{N+\alpha }\left (\overline {\Omega }\right )$
to denote Hölder continuous functions because they use
$C^{N+\alpha }\left (\Omega \right )$
to denote locally Hölder continuous functions. Finally, it is also interesting to note that the form of operator T itself does not depend on N and
$\alpha $
, although its norm as an operator
$C^{N+\alpha }(\Omega )\to C^{N+\alpha }(\mathbb {R}^n)$
does, of course. This holds provided that we follow the construction of [Reference Gilbarg and Trudinger33] but use smooth parametrizations of
$\Omega $
, instead of less regular ones.
The last result of this appendix is a lemma that establishes a bound for a Besov norm of functions that are compactly supported in a collared neighborhood of an
$(n-1)$
-dimensional surface. This can be seen as a Poincaré inequality with negative regularity.
Lemma B.4. Let
$\Omega \subset \mathbb {R}^n$
,
$n\geq 2$
, be a bounded domain with smooth boundary. Let
$\alpha \in (0,1)$
and let
$r>0$
be sufficiently small (depending on
$\Omega $
). Consider a function
$f\in C_c^\infty (\mathbb {R}^n)$
supported in
$\{x\in \mathbb {R}^n:0<\operatorname {\mathrm {dist}}(x,\Omega )<r\}$
. We have

Proof. Let us denote . We begin by computing

where we have used that
$|h(x)|\lesssim \langle x\rangle ^{-8n}$
because h is in the Schwartz class. Here
$P_Nf$
stands for the nonhomogeneous dyadic blocks in the Littlewood-Paley decomposition of f, and h is the corresponding convolution kernel.
We claim that the following estimate holds:

Using this, the bound for f readily follows:

Therefore, the problem is reduced to estimating the integral above.
Let

For each point
$x\in \partial \Omega $
and for each small enough
$R>0$
, there is a boundary normal chart

where
$Q_R:=[0,R)\times (-R,R)^{n-1}$
, such that
$\Psi _{x,R}(0)=x$
and

Recall that, by the definition of boundary normal coordinates, the pullback
$X_{x,R}^*g_0$
of the Euclidean metric
$g_0:=(\delta _{ij})$
to this chart satisfies
$X_{x,R}^*g_0(0)=g_0$
. One can also assume that
$\nabla X_{x,R}(0)$
is the identity matrix, and that

for all
$y\in X_{x,R}(Q_{R/2})$
and all
$x\not \in X_{x,R}(Q_{R/2})\cup \Omega $
.
Since
$\partial \Omega $
is a smooth compact hypersurface of
$\mathbb {R}^n$
, it is standard that there is some small enough
$R>0$
and a finite collection of charts
$\{ X^j:=X_{x_j,R}\}_{j=1}^J$
as above such that
$\{U^j:=X^j(Q_{R/2})\}_{j=1}^J$
is a cover of the set
$\overline {U_{R/2}}$
and which satisfy

and

for all
$z,\tilde z\in Q_R$
. Moreover, the distance between the point
$X^j(z)$
and the boundary is comparable to its first coordinate, in the sense that

for all
$z\in Q_R$
.
Let us now estimate the integral (B.1) over the subset
$U^j\cap U$
, with
$x\in U$
. If
$x\not \in X^j(Q_{R})$
, by (B.2), one immediately has

where we have also used that
$|U|\lesssim r$
. If
$x\in X^j(Q_{R})$
, one can write
$x=X^j(\tilde z)$
for some
$\tilde z\in Q_R$
. By (B.4), one then has

Here we have used the notation
$z=(z_1,z')\in [0,R)\times (-R,R)^{n-1}$
and the fact that the Jacobian
$J_{X^j}$
is bounded by (B.3).
We can now carry out the integrations in
$z_1$
and
$z'$
separately. Since

the estimate follows from the fact that

The estimate (B.1) then follows by summing over j.
Competing interest
The authors have no competing interests to declare.
Financial support
The authors are very grateful to the reviewers for their suggestions and corrections on a previous version of the manuscript. This work has received funding from the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation programme through the grant agreement 862342 (A.E.). It is partially supported by the grants CEX2023-001347-S, RED2022-134301-T and PID2022-136795NB-I00 (A.E. and D.P.-S.) funded by MCIN/AEI/10.13039 /501100011033, and Ayudas Fundación BBVA a Proyectos de Investigación Científica 2021 (D.P.-S.). J.P.-T. was partially supported by an FPI grant CEX2019-000904-S-21-4 funded by MICIU/AEI and by FSE+.