Non-uniqueness of the transport equation at high spacial integrability
Abstract
In this paper, we show the non-uniqueness of the weak solution in the class for the transport equation driven by a divergence-free vector field happens in the range with some , as long as , . As a corollary, in time of the density is critical in some sense for the uniqueness of weak solution. Our proof is based on the convex integration method developed in [39, 20].
Keywords: Transport equation; non-uniqueness; convex integration.
MR Subject Classification: 35A02; 35D30; 35Q35.
1 Introduction
This paper deals with the problem of (non-)uniqueness of weak solution to the linear transport equation on the torus written by
(1.1) |
where is the unknown density, is the initial data and is a given divergence-free vector field, i.e. in the sense of distribution. In this case, the weak solution to (1.1) is defined as
Another equivalent definition is (see e.g. [34, p276])
and is continuous in time with in the distributional sense.
Since the transport equation is linear, the existence of weak solutions can be obtained by the method of regularization even for very rough vector fields, see [35, Proposition II.1]. However, to obtain the uniqueness, more complicated discussions will be involved.
1.1 Background
In the smooth setting, the method of characteristics solves first-order PDE by converting the PDE into an appropriate system of ODE [37, Sec.3.2]. In particular, for Lipschitz vector fields , the solution to (1.1) can be given by the Lagrangian representation with the flow map solving the ODEs
(1.2) |
This link results in the existence and uniqueness of the PDE (1.1) by the Cauchy-Lipschitz (Picard-Linderöf) theory for the ODE (1.2) (The Lipschitz condition on the vector fields can be replaced by one-side Lipschitz, -Lipschitz or Osgood condition, we refer to [1] for the elaborate surveys).
For non-Lipschitz vector fields, the link is less obvious and the uniqueness issue of (1.1) becomes subtler. The first breakthrough result traces back to the celebrated work of DiPerna and Lions [35], they proved that for all , the Lagrangian representation and uniqueness hold in the class for any given vector field with , where is the Hölder conjugate exponent of . This result was extended to the case by Ambrosio [6] with deep tools from geometric measure theory. Whereafter, there are abundant researches for the theory of non-smooth vector fields and their applications on non-linear PDEs. For instance, the analogous results of [35, 6] have been established, for vector fields with gradient given by a singular integral in [13] and for nearly incompressible vector fields in [11], and the results in [13] has been adapted to obtain the Lagrangian solutions for the Vlasov-Poisson system with data [12], see also [8]. We refer to the works [22, 23, 38, 25, 18, 19] and the surveys [9, 28, 24, 7] for other important progresses and related results in this direction.
For the non-uniqueness issue of (1.1), two examples were given by DiPerna and Lions [35], for an autonomous vector field but and for an autonomous divergence-free vector field but . Much later, based on the work [2], Depauw [33] constructed an example of a divergence-free vector field but , in the class of bounded densities. See also [10, 42, 5]. In [3, 4], Alberti, Bianchini and Crippa showed an example of an autonomous divergence-free vector field but , in the class of bounded densities. More recently, based on anomalous dissipation and mixing, Drivas, Elgindi, Iyer and Jeong proved non-uniqueness in the class for [36]. Roughly speaking, these results are proved by showing the non-uniqueness of the flow maps (the ODE level), with the violation of either the boundedness of or the spacial regularity of .
Another approach is based on the convex integration technique (for complete summaries on this enormous theory, we refer to the surveys articles [30, 31, 32, 15, 17]), which gives counterexamples of uniqueness directly at the PDE level. The first non-uniqueness result with this technique was obtained by Crippa et al in [26] using the framework of [29], but for vector field was merely bounded. Later, a breakthrough result for the Sobolev vector field was obtained by Modena and Székelyhidi [40]. Based on their work, a series of works for the non-uniqueness to (1.1) have been done recently, we list the main functional setting of them below:
-
(I).
[40](Modena and Székelyhidi): , for , and .
- (II).
-
(III).
[39](Modena and Sattig): , for and .
-
(IV).
[14](Bruè, Colombo and De Lellis): , , for , and .
-
(V).
[20](Cheskidov and Luo): , for , and .
Comments:
Very recently, [19] proves that the uniqueness result holds for with in the class under the assumption that the so called forward-backward integral curves of are trivial. Whereas, the non-uniqueness result of [39] holds for with in the class . Hence, roughly speaking when , the regularity condition with is optimal for the uniqueness of the weak solution in the class .
When , the uniqueness is unclear in the range for , . However, [14, Theorem 1.5] states that the uniqueness result holds for positive density and with , which goes beyond the DiPerna-Lions range.
In [20], the non-uniqueness with the sharp spacial integrability () has been reached but at the expensive of time regularity. They proved the result using the convex integration scheme of [40] combined with temporal intermittency and oscillations in the spirit of [16]. We refer to [16, 20, 17] for a more comprehensive discussion of the temporal intermittency.
A few months after we finished this paper, we learn that Cheskidov and Luo post a extremely nice result [21] on arXiv, in which they show the non-uniqueness result for , , , it also implies that the time-integrability assumption in the uniqueness of the DiPerna-Lions theory is sharp. In view of this surprising result, we have reconsidered the implications of our study, see Remark 1.3.
1.2 Main results
Based on the frameworks of [39, 20], we obtain the following main result in this paper, which indicates that the non-uniqueness happens even in the range , as long as , .
Theorem 1.1.
Let and , are Hölder conjugate exponents of respectively, satisfying and
(1.3) |
with .
Then there exists a divergence-free vector field
such that the uniqueness of (1.1) fails in the class .
Remark 1.1.
On the one hand, Theorem 1.1 means that even for the spacial integrability higher than the DiPerna-Lions regime , non-uniqueness still happens, which might be new in contrast to the previous works [39, 14, 20]. On the other hand, notice for and ,
might be taken greater than 1, hence we improve the result of [20] in two aspects: the restrictions on time integrability () and on spacial integrability ().
Remark 1.2.
After suitable parameters setting, the condition (1.3) might be replaced by
(1.4) |
i.e., we allow the spatial integrability of can be arbitrary . However, since the low time integrability , Cheskidov and Luo’s work [21] fully covers our result in this case. We mention that a comfortable condition on might be . However, it can not be reached in this paper, since the second condition in (1.3) or (1.4) plays a significant role in the proof of Lemma 4.8 (When , we need to set in the proof of Lemma 4.8, then the balance of parameters setting in Sec.4.3 will be upset).
Remark 1.3.
Cheskidov and Luo’s work [21] is very exciting, since they improve extremely the spatial regularity for the non-uniqueness. However, this is achieved by abandoning all the time regularity of the vector field except integrability. A significant difference between Theorem 1.1 and Cheskidov, Luo’s result [21] is that we can provide non-uniqueness with a little higher temporal integrability of (). To the best of the authors’ knowledge, there is no result of the non-uniqueness for , under the range , before Theorem 1.1. Notice also when , [19] has proved the uniqueness for with , under some additional assumptions of integral curves of . Hence it might be an interesting and valuable problem to consider uniqueness/non-uniqueness under high temporal integrability of and high spatial integrability of ().
Remark 1.4.
It seems possible to extend Theorem 1.1 to the border case and to the transport-diffusion equation
by utilizing the technique in [41, 39]. We finally mention that the two dimensional case is not treated in this paper due to the stationary Mikado flow, see Remark 3.4. However, thanks to the recent tricks given by Cheskidov and Luo [21], our result can be extended to the two dimensional case without difficulty.
Notice as long as . Hence, combine the uniqueness result in [35, Corollary II.1], we obtain immediately from Theorem 1.1 that at least in the following sense, the in time of the density is critical for the uniqueness of weak solutions to (1.1).
Corollary 1.2.
We identify with an 1-dimensional torus and the time-periodic function on means for all . Theorem 1.1 follows immediately from the following theorem.
Theorem 1.3.
Let and , are Hölder conjugate exponents of respectively, satisfying and (1.3). For any and any time-periodic with constant mean
there exists a divergence-free vector field and a density such that the following holds.
-
(i).
and .
-
(ii).
is continuous in the distributional sense and for , .
-
(iii).
is a weak solution to (1.1) with initial data .
-
(iv).
The deviation of norm is small on average: .
Proof of Theorem 1.1. Let with . We take with satisfying if and if . We apply Theorem 1.3 with and obtain solving (1.1) with . By the choice of , we claim that cannot have a constant norm and obviously , which implies the non-uniqueness (as well as the existence of non-renormalized solution).
Indeed, assume for some . On the one hand, due to , we have
hence , which implies .
On the other hand,
hence , in contradiction with , we prove the claim. ∎
1.3 Notations
The norms of , , will be denoted standardly as , , or just when there is no confusion. The norm of will be denoted as .
For any , its spacial mean is and denote simply as . We denote as the space of smooth periodic functions with zero mean.
Denote the standard partial differential operators with multiindex as . For any , the obvious facts are for any , and of course as long as .
will be denoted as with some inessential constant . If the constant depends on some quantities, for instance , it will be denoted as , and means .
represents a positive constant that might depend on the old solution and the constant but never on given in Proposition 2.1. may change from line to line.
2 Main Proposition and the Proof of Theorem 1.3
Without loss of generality, in the rest of the paper, we assume and identify the time interval with an 1-dimensional torus.
We follow the framework of [20] (see [40] for the earliest version) to obtain space-time periodic approximate solutions to the transport equation by solving the continuity-defect equation
(2.1) |
where is called the defect field.
For any , denote . To build a iteration scheme for proving Theorem 1.3, we will construct the small perturbations on of to obtain a new solution such that the new defect field has small norm. This is the following main proposition of the paper.
Proposition 2.1.
Let and , are Hölder conjugates of respectively, satisfying and (1.3). There exist a universal constant and a large integer such that the following holds.
Proof of Theorem 1.3. Assume . We will construct a sequence of solutions to (2.1). For , we set
Notice the constant mean assumption on implies zero mean of , hence solves (2.1).
Next we apply Proposition 2.1 inductively to obtain for as follows. Set and choose sequence such that , . Observe that .
Given , we apply Proposition 2.1 with parameters and to obtain a new triple which verifies
When , we have
Clearly there are functions and such that in and in . Moreover, we have and in , and in . Combine the fact for some , we obtain the temporal continuity of in the distributional sense and for , , furthermore is a weak solution to (1.1) with initial data .
Finally, thanks to the choice of , we have
∎
3 Technical tools
In this section, we collect the technical tools prepared in [40, 20] for the proof of the main proposition (Proposition 2.1). We refer to [40, 39, 20] for more details.
3.1 Anti-divergence operators
By the classical Fourier analysis, for any , a unique solution in of the Poisson equation
is given by . Hence the standard anti-divergence operator can be defined as
which satisfies
Obviously, for every and there holds
Further, the first order bilinear anti-divergence operator can be defined as
which satisfies
The bilinear anti-divergence operator has the additional advantage of gaining derivative from when has zero mean and a very small period. See also higher order variants in [39].
Remark 3.1.
Notice the definitions of in [39] are slightly different from the definitions in this paper, which actually are defined as
In this case, must be mean zero.
Lemma 3.1 ([20, Lemma 2.1]).
Let . For every and , the anti-divergence operator is bounded on :
(3.1) |
Moreover for all and , the Calderón-Zygmund inequality holds:
(3.2) |
Lemma 3.2 ([20, Lemma 2.2]).
Let and . Then for any :
Lemma 3.3 ([20, Lemma 2.4]).
Let and . Then for all ,
Remark 3.2.
Lemma 3.4 ([20, Lemma 2.5]).
Let , and . Then for all even
3.2 Mikado flows
Now we define the Mikado flow given in [20], which are a family of periodic stationary solutions to the transport equation (1.1). Where is the Mikado density, is the Mikado filed.
Let . Fix satisfying
Denote and , . We can define a family of non-periodic stationary solutions to (1.1) as
The potential is defined as
The periodic solutions with mutually disjoint supports can be constructed by translation and periodization of the non-periodic flow . This is based on the geometrical fact that along any two directions in for , there exist two disjoint lines.
For instance, notice are cylinders with side length lying at -axis respectively. When , one can move the -th cylinder with length along a direction , such that all cylinders are mutually disjoint and keep lying in . By means of the Possion summation, we define periodic Mikado flow as
and the periodic potential is defined as
For the Mikado flow, we have the following proposition.
Proposition 3.5 ([20, Proposition 4.3, Theorem 4.4]).
Let and . Then the periodic functions , verify the following.
-
(i).
For any , ,
(3.3) -
(ii).
solve
(3.4) -
(iii).
There hold
(3.5)
where is the -th standard Euclidean basis.
Remark 3.4.
The concept of Mikado flow was introduced in [27] firstly and then adapted in [40, 41, 39, 14, 20] for the non-uniqueness results of the transport equation. Notice the construction of stationary Mikado flow requires the dimension is not less than three. An alternative is the space-time Mikado flow introduced in [39] (see also [14]), which can be constructed in . However, the space-time flow brings new difficulty when applying the temporal intermittency.
3.3 Intermittent functions in time
In this subsection, we define the intermittent oscillatory functions and . Take satisfying and
For , define
In this case, we have that be 1-periodic functions and the following facts (similar to Proposition 3.5) hold
(3.6) |
Define , which obviously satisfies
(3.7) |
We write , where is the -th standard Euclidean basis.
Recall the notation for . Define the smooth cutoff functions satisfying
(3.8) |
Where is fixed sufficiently small enough such that
(3.9) |
Notice . By a slight abuse of notation, denote the 1-periodic extension in time of . Define .
4 Proof of Proposition 2.1
In this section, we follow the lines in [20] to construct the perturbations and defect field, and then finish the proof of Proposition 2.1. In particular, we set that the concentration in time is stronger than in space, the oscillation in time is weaker than in space, see Sec. 4.3.
4.1 Constructing perturbations
We first define the principle part of the perturbations by the Mikado flows given in Proposition 3.5. Let
(4.1) |
where
Notice . By [20, Lemma 7.1], we have
and the following estimates hold true:
(4.2) |
Moreover for any , there exists a constant such that
(4.3) |
Notice is not mean zero and is not divergence-free. To make sure the zero mean of and divergence-free , we need the corrections of the perturbations. The correctors are defined by
Notice by Proposition 3.5, has zero mean, we have
hence
Finally, to take advantage of the temporal oscillations, we define the temporal oscillator
Notice by definition, has zero mean.
Now we are able to define the perturbations by
(4.4) |
and are defined by
(4.5) |
4.2 Constructing the defect field
Now we define the new defect field satisfying the continuity-defect equation
We split into four parts
(4.6) |
satisfying
Obviously, can be defined by
Since
with the help of , we define by
Now we consider . We split into three parts
where
Firstly notice
According to , and when , there holds . We have
Hence
Notice
Hence satisfies
Since
can be defined by
4.3 Setting the parameters
There are four controllable parameters . We will set as some positive powers of and then let large enough.
Concentration parameter in space : hereafter, we take with large enough depending on the old solution and given in Proposition 2.1, such that all lemmas below hold. Actually, how large is and what quantities do depend on are inessential as long as is taken to be finite at last, since the most significant matter is that we need to balance the estimates on the perturbations and the new defect field by reasonable setting of . Recall , the following setting works well.
Concentration parameter in time : setting with
(4.7) |
Oscillation parameter in space : setting with
(4.8) |
Oscillation parameter in time : setting .
Finally, we choose . Where is the floor function and is the ceiling function.
Remark 4.1.
4.4 Estimates on the perturbations
Lemma 4.1 (Estimate on ).
In particular, for large enough,
Proof.
Lemma 4.2 (Estimate on ).
For , we have
In particular, for large enough,
Lemma 4.3 (Estimate on ).
In particular, for large enough,
Proof.
Lemma 4.4 (Estimate on with norm).
In particular, for large enough,
Proof.
Lemma 4.5 (Estimate on with norm).
for some . In particular, for large enough, we have
Remark 4.2.
and the first condition of (1.3) are required in this estimate.
Lemma 4.6 (Estimate on ).
for some . In particular, for large enough, we have
Remark 4.3.
and the first condition of (1.3) are required in this estimate.
Proof.
Lemma 4.7.
For and large enough, we have
Moreover .
4.5 Estimates on the new defect field
Lemma 4.8 (Estimate on ).
For large enough, we have
(4.9) |
Remark 4.4.
Proof.
For , by Lemma 3.2 and notice , we have
by (4.3) | |||
by (3.4) | |||
by (3.2) and |
(When under the conditions (1.4), we apply (3.2) to the last inequality with sufficiently close to 1). Hence by (3.6) and Proposition 3.5
by (4.7),(4.8) | |||
where we have used by the second assumption of (1.3). Hence take large enough, we have
(4.10) |
Lemma 4.9 (Estimate on ).
For large enough, we have
Proof.
Lemma 4.10 (Estimate on ).
For large enough, we have
Proof.
Lemma 4.11 (Estimate on ).
For large enough, we have
Proof.
Lemma 4.12 (Estimate on ).
For large enough, we have
Lemma 4.13 (Estimate on ).
4.6 Conclusion
Proof of Proposition 2.1. Combine Lemma 4.1-4.13, there exists large enough depending on the old solution and given in Proposition 2.1, such that for , we have
The inessential constants in the estimates can be taken as a universal constant , that is
Notice (2.2) follows from Lemma 4.7, meanwhile (2.3) follows from the definition (4.1) of and the fact . Proposition 2.1 follows. ∎
Acknowledgments
This work is supported by the National Natural Science Foundation of China (Grant No.11871024).
References
- [1] R. P. Agarwal and V. Lakshmikantham. Uniqueness and nonuniqueness criteria for ordinary differential equations, volume 6 of Series in Real Analysis. World Scientific Publishing Co., Inc., River Edge, NJ, 1993.
- [2] M. Aizenman. On vector fields as generators of flows: a counterexample to Nelson’s conjecture. Ann. of Math. (2), 107(2):287–296, 1978.
- [3] G. Alberti, S. Bianchini, and G. Crippa. Structure of level sets and Sard-type properties of Lipschitz maps. Ann. Sc. Norm. Super. Pisa Cl. Sci. (5), 12(4):863–902, 2013.
- [4] G. Alberti, S. Bianchini, and G. Crippa. A uniqueness result for the continuity equation in two dimensions. J. Eur. Math. Soc. (JEMS), 16(2):201–234, 2014.
- [5] G. Alberti, G. Crippa, and A. L. Mazzucato. Exponential self-similar mixing by incompressible flows. J. Amer. Math. Soc., 32(2):445–490, 2019.
- [6] L. Ambrosio. Transport equation and Cauchy problem for vector fields. Invent. Math., 158(2):227–260, 2004.
- [7] L. Ambrosio. Well posedness of ODE’s and continuity equations with nonsmooth vector fields, and applications. Rev. Mat. Complut., 30(3):427–450, 2017.
- [8] L. Ambrosio, M. Colombo, and A. Figalli. On the Lagrangian structure of transport equations: the Vlasov-Poisson system. Duke Math. J., 166(18):3505–3568, 2017.
- [9] L. Ambrosio and G. Crippa. Existence, uniqueness, stability and differentiability properties of the flow associated to weakly differentiable vector fields. In Transport equations and multi-D hyperbolic conservation laws, volume 5 of Lect. Notes Unione Mat. Ital., pages 3–57. Springer, Berlin, 2008.
- [10] S. Attanasio and F. Flandoli. Zero-noise solutions of linear transport equations without uniqueness: an example. C. R. Math. Acad. Sci. Paris, 347(13-14):753–756, 2009.
- [11] S. Bianchini and P. Bonicatto. A uniqueness result for the decomposition of vector fields in . Invent. Math., 220(1):255–393, 2020.
- [12] A. Bohun, F. Bouchut, and G. Crippa. Lagrangian solutions to the Vlasov-Poisson system with density. J. Differential Equations, 260(4):3576–3597, 2016.
- [13] F. Bouchut and G. Crippa. Lagrangian flows for vector fields with gradient given by a singular integral. J. Hyperbolic Differ. Equ., 10(2):235–282, 2013.
- [14] E. Brué, M. Colombo, and C. De Lellis. Positive solutions of transport equations and classical nonuniqueness of characteristic curves. Arch. Ration. Mech. Anal., 240(2):1055–1090, 2021.
- [15] T. Buckmaster and V. Vicol. Convex integration and phenomenologies in turbulence. EMS Surv. Math. Sci., 6(1-2):173–263, 2019.
- [16] T. Buckmaster and V. Vicol. Nonuniqueness of weak solutions to the Navier-Stokes equation. Ann. of Math. (2), 189(1):101–144, 2019.
- [17] T. Buckmaster and V. Vicol. Convex integration constructions in hydrodynamics. Bull. Amer. Math. Soc. (N.S.), 58(1):1–44, 2021.
- [18] L. Caravenna and G. Crippa. Uniqueness and Lagrangianity for solutions with lack of integrability of the continuity equation. C. R. Math. Acad. Sci. Paris, 354(12):1168–1173, 2016.
- [19] L. Caravenna and G. Crippa. A directional Lipschitz extension lemma, with applications to uniqueness and Lagrangianity for the continuity equation. Comm. Partial Differential Equations, 46(8):1488–1520, 2021.
- [20] A. Cheskidov and X. Luo. Nonuniqueness of weak solutions for the transport equation at critical space regularity. Ann. PDE, 7(1):Paper No. 2, 45, 2021.
- [21] A. Cheskidov and X. Luo. Extreme temporal intermittency in the linear sobolev transport: almost smooth nonunique solutions. on arXiv, 2022.
- [22] F. Colombini and N. Lerner. Uniqueness of continuous solutions for BV vector fields. Duke Math. J., 111(2):357–384, 2002.
- [23] F. Colombini, T. Luo, and J. Rauch. Uniqueness and nonuniqueness for nonsmooth divergence free transport. In Seminaire: Équations aux Dérivées Partielles, 2002–2003, Sémin. Équ. Dériv. Partielles, pages Exp. No. XXII, 21. École Polytech., Palaiseau, 2003.
- [24] G. Crippa. The flow associated to weakly differentiable vector fields, volume 12 of Tesi. Scuola Normale Superiore di Pisa (Nuova Series) [Theses of Scuola Normale Superiore di Pisa (New Series)]. Edizioni della Normale, Pisa, 2009.
- [25] G. Crippa and C. De Lellis. Estimates and regularity results for the DiPerna-Lions flow. J. Reine Angew. Math., 616:15–46, 2008.
- [26] G. Crippa, N. Gusev, S. Spirito, and E. Wiedemann. Non-uniqueness and prescribed energy for the continuity equation. Commun. Math. Sci., 13(7):1937–1947, 2015.
- [27] S. Daneri and L. Székelyhidi, Jr. Non-uniqueness and h-principle for Hölder-continuous weak solutions of the Euler equations. Arch. Ration. Mech. Anal., 224(2):471–514, 2017.
- [28] C. De Lellis. ODEs with Sobolev coefficients: the Eulerian and the Lagrangian approach. Discrete Contin. Dyn. Syst. Ser. S, 1(3):405–426, 2008.
- [29] C. De Lellis and L. Székelyhidi, Jr. The Euler equations as a differential inclusion. Ann. of Math. (2), 170(3):1417–1436, 2009.
- [30] C. De Lellis and L. Székelyhidi, Jr. The -principle and the equations of fluid dynamics. Bull. Amer. Math. Soc. (N.S.), 49(3):347–375, 2012.
- [31] C. De Lellis and L. Székelyhidi, Jr. High dimensionality and h-principle in PDE. Bull. Amer. Math. Soc. (N.S.), 54(2):247–282, 2017.
- [32] C. De Lellis and L. Székelyhidi, Jr. On turbulence and geometry: from Nash to Onsager. Notices Amer. Math. Soc., 66(5):677–685, 2019.
- [33] N. Depauw. Non unicité des solutions bornées pour un champ de vecteurs BV en dehors d’un hyperplan. C. R. Math. Acad. Sci. Paris, 337(4):249–252, 2003.
- [34] R. J. DiPerna and P.-L. Lions. Global weak solutions of kinetic equations. Rend. Sem. Mat. Univ. Politec. Torino, 46(3):259–288 (1990), 1988.
- [35] R. J. DiPerna and P.-L. Lions. Ordinary differential equations, transport theory and Sobolev spaces. Invent. Math., 98(3):511–547, 1989.
- [36] T. D. Drivas, T. M. Elgindi, G. Iyer, and I.-J. Jeong. Anomalous dissipation in passive scalar transport. Arch. Ration. Mech. Anal., 243(3):1151–1180, 2022.
- [37] L. C. Evans. Partial differential equations, volume 19 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, second edition, 2010.
- [38] C. Le Bris and P.-L. Lions. Renormalized solutions of some transport equations with partially velocities and applications. Ann. Mat. Pura Appl. (4), 183(1):97–130, 2004.
- [39] S. Modena and G. Sattig. Convex integration solutions to the transport equation with full dimensional concentration. Ann. Inst. H. Poincaré Anal. Non Linéaire, 37(5):1075–1108, 2020.
- [40] S. Modena and L. Székelyhidi, Jr. Non-uniqueness for the transport equation with Sobolev vector fields. Ann. PDE, 4(2):Paper No. 18, 38, 2018.
- [41] S. Modena and L. Székelyhidi, Jr. Non-renormalized solutions to the continuity equation. Calc. Var. Partial Differential Equations, 58(6):Paper No. 208, 30, 2019.
- [42] Y. Yao and A. Zlatoš. Mixing and un-mixing by incompressible flows. J. Eur. Math. Soc. (JEMS), 19(7):1911–1948, 2017.