In this paper, we consider the incompressible Euler and Navier-Stokes equations in R2. It is well known that the Euler and Navier-Stokes equations are globally well-posed for initial data in Hs(s>2). The main purpose of the present paper is to consider the case s=1. We prove that, for initial data in H1, the Euler and Navier-Stokes equations both have global solutions, and the solutions are uniformly bounded with respect to time. Moreover, the solution for the Navier-Stokes equations is unique. We also prove that, as the viscosity tends to zero, the solution of the Navier-Stokes equations converges to the one of the Euler equations.
Citation: Shaoliang Yuan, Lin Cheng, Liangyong Lin. Existence and uniqueness of solutions for the two-dimensional Euler and Navier-Stokes equations with initial data in H1[J]. AIMS Mathematics, 2025, 10(4): 9310-9321. doi: 10.3934/math.2025428
[1] | Kaile Chen, Yunyun Liang, Nengqiu Zhang . Global existence of strong solutions to compressible Navier-Stokes-Korteweg equations with external potential force. AIMS Mathematics, 2023, 8(11): 27712-27724. doi: 10.3934/math.20231418 |
[2] | Xiaoxia Wang, Jinping Jiang . The pullback attractor for the 2D g-Navier-Stokes equation with nonlinear damping and time delay. AIMS Mathematics, 2023, 8(11): 26650-26664. doi: 10.3934/math.20231363 |
[3] | Jianxia He, Ming Li . Existence of global solution to 3D density-dependent incompressible Navier-Stokes equations. AIMS Mathematics, 2024, 9(3): 7728-7750. doi: 10.3934/math.2024375 |
[4] | Yina Lin, Qian Zhang, Meng Zhou . Global existence of classical solutions for the 2D chemotaxis-fluid system with logistic source. AIMS Mathematics, 2022, 7(4): 7212-7233. doi: 10.3934/math.2022403 |
[5] | Hui Fang, Yihan Fan, Yanping Zhou . Energy equality for the compressible Navier-Stokes-Korteweg equations. AIMS Mathematics, 2022, 7(4): 5808-5820. doi: 10.3934/math.2022321 |
[6] | Xiaoguang You . Vanishing viscosity limit of incompressible flow around a small obstacle: A special case. AIMS Mathematics, 2023, 8(2): 2611-2621. doi: 10.3934/math.2023135 |
[7] | Abdelkader Moumen, Ramsha Shafqat, Azmat Ullah Khan Niazi, Nuttapol Pakkaranang, Mdi Begum Jeelani, Kiran Saleem . A study of the time fractional Navier-Stokes equations for vertical flow. AIMS Mathematics, 2023, 8(4): 8702-8730. doi: 10.3934/math.2023437 |
[8] | Zihang Cai, Chao Jiang, Yuzhu Lei, Zuhan Liu . Global classical solution of the fractional Nernst-Planck-Poisson-Navier- Stokes system in R3. AIMS Mathematics, 2024, 9(7): 17359-17385. doi: 10.3934/math.2024844 |
[9] | Kunquan Li . Analytical solutions and asymptotic behaviors to the vacuum free boundary problem for 2D Navier-Stokes equations with degenerate viscosity. AIMS Mathematics, 2024, 9(5): 12412-12432. doi: 10.3934/math.2024607 |
[10] | Benedetta Ferrario, Christian Olivera . Lp-solutions of the Navier-Stokes equation with fractional Brownian noise. AIMS Mathematics, 2018, 3(4): 539-553. doi: 10.3934/Math.2018.4.539 |
In this paper, we consider the incompressible Euler and Navier-Stokes equations in R2. It is well known that the Euler and Navier-Stokes equations are globally well-posed for initial data in Hs(s>2). The main purpose of the present paper is to consider the case s=1. We prove that, for initial data in H1, the Euler and Navier-Stokes equations both have global solutions, and the solutions are uniformly bounded with respect to time. Moreover, the solution for the Navier-Stokes equations is unique. We also prove that, as the viscosity tends to zero, the solution of the Navier-Stokes equations converges to the one of the Euler equations.
In this paper, we focus on the two-dimensional incompressible Euler and Navier-Stokes equations. The incompressible Euler equations in R2 read as
{∂tu+u⋅∇u+∇p=0,divu=0, | (1.1) |
for (t,x)∈R+×R2, where the unknowns u=(u1,u2) and p=p(t,x) represent the velocity and the pressure of the fluid, respectively.
As is well known, system (1.1) is globally well-posed when the initial velocity lies in Hs(s>2) (see, for example, [14]), while the existence problem of solutions for the cases s=2 and s=0 are unknown. Indeed, it was shown in [1] that, if the initial data that belongs to H2-space is perturbed, then the system is ill-posed. Recently, Elgindi and Masmoudi [4] showed ill-posedness in C1∩L2 space. This is partially due to the fact that, by elementary energy methods, we only obtain the following a priori estimate:
ddt‖u‖2Hs≤‖∇u‖L∞‖u‖2Hs, | (1.2) |
which implies that the Sobolev order s should be greater than 2 to close the energy inequality.
To consider the existence problem with initial data of low regularity, we generally resort to the vorticity-stream formulation of the Euler equations, which can be written as
{∂tw+u⋅∇w=0,u=K[w], | (1.3) |
where K:=12π(−x2|x|2,x1|x|2) is the kernel of the Biot-Savart law, and w is the vorticity of the fluid. There is a number of literatures about the existence and uniqueness of weak solutions of system (1.3). When the initial vorticity w0∈L1∩L∞, Yudovich [18] proved the global existence and uniqueness of weak solutions. Later, Vishik [17] and Yudovich [19] extended these results into a space slightly larger than L∞. If w0 belongs to L1∩Lp with p>1, the global existence result was obtained by Diperna and Majda in [3]. We also refer to [2,5,12,13] for the case that w0 is a finite Randon measure. It is worth noting that the L1 condition on w0 stems from the second subequation of system (1.3), from which we bound the L∞-norm of u as:
‖u‖L∞≤C(‖w‖L1+‖w‖L∞). | (1.4) |
We want to emphasize that the above works are either based on system (1.1) to establish smooth solutions or based on system (1.3) to obtain weak ones. The present paper plans to utilize these two systems together to retrieve the L1 restriction on the initial vorticity and lower the Sobolev index s to s=1 for the initial velocity. More precisely, let us consider that u is a smooth solution of system (1.1). Then, by multiplying the first subequation with u and integrating over [0,t)×R2, we have that
‖u‖L2≤‖u0‖L2. | (1.5) |
On the other hand, as u also satisfies system (1.3), it can be checked that
‖∇u‖L2≤C‖w‖L2≤C‖w0‖L2≤C‖u0‖H1. | (1.6) |
Collecting (1.5) and (1.6) gives a uniform bound of the H1-norm of u in time, which brings out the H1 solution by using approximation and compactness arguments.
We are now in the position to state the following result:
Theorem 1.1. Suppose that the initial velocity u0∈H1(R2) is divergence-free. Then the Euler equations (1.1) have a weak solution u∈L∞([0,∞);H1(R2)) with the estimate
supt∈[0,∞)‖u‖H1≤C‖u0‖H1. | (1.7) |
Another interest of the present work is to consider the Cauchy problem of the Navier-Stokes equations, which can be written as:
{∂tuν+uν⋅∇uν+∇pν=νΔuν,divuν=0,uν|t=0=uν0, | (1.8) |
where uν0 is the given initial velocity and ν>0 is the viscosity. The unknowns uν=(uν1,uν2) and pν=pν(t,x) represent the velocity and the pressure of the fluid, respectively.
With initial data in L2, Leray [9] proved that the Navier-Stokes equations have a global weak solution in Rd(d = 2, 3), and Hopf [6] obtained the existence of a global weak solution in domains with boundaries. Since then, many mathematicians studied the uniqueness and regularity of Leray-Hopf solutions. It was proved that for the two-dimensional case, the solution is unique and regular; see [8,10,11,16]. However, the problem of uniqueness of Leray-Hopf solutions for the three-dimensional case is still open.
As is well known, for the case that the initial data lies in Hs(s>1), Eq (1.8) has a unique solution (see, for example, [7,14]). Moreover, the solution is global when s>2. We consider in the present work the borderline case: The problem of existence and uniqueness of solutions for (1.8) when the initial data lies in H1. We find that, when we analyze (1.8) and its vorticitiy-stream formulation together, the obstacle that is caused by the convection term can be bypassed, and for any s≥0, the following interesting a priori estimate holds:
‖uν(t)‖2Hs+ν∫t0‖∇uν(s)‖2Hsds≤C(ν,uν0), | (1.9) |
where C≡C(ν,uν0) is a constant that depends on ν and the Hs-norm of the initial data uν0, while does not depend on t. It is worth mentioning that for s=1, the constant C is independent of ν.
With the uniform bound (1.9) in hand, we can establish the global existence and uniqueness of solutions, and we have the following result:
Theorem 1.2. Let s≥0 be a given integer. Suppose that the initial velocity uν0∈Hs(R2) is divergence-free. Then the Navier-Stokes equations (1.8) have a unique solution uν∈L∞([0,∞);Hs(R2)) with the estimate (1.9).
Remark 1.3. It worth noting that for those s>2 that are not integers, the above conclusion also holds.
As we know, the vanishing viscosity limit problem is one of the most fundamental problems in fluid mechanics. Masmoudi [15] verified the inviscid limit for initial data in Hs(s>2). From the above uniform bound (1.9), we may treat the vanishing viscosity limit problem for initial data only in H1.
Corollary 1.4. Suppose that uν0 is divergence-free and converges to u0 in H1(R2) as ν tends to zero. Then, as ν→0, the solution uν of (1.8) converges to a solution u of (1.1) with the estimate
supt∈[0,∞)‖u‖H1≤C‖u0‖H1. | (1.10) |
The remainder of this article is divided into three sections. In Section 2, we establish the global existence of solutions for the Euler equations (1.1). In Section 3, we prove the global well-posedness of the Navier-Stokes equations, i.e., Theorem 1.2. In the last section, we consider the vanishing viscosity limit problem, and prove Corollary 1.4.
In this section, we prove the global existence of solutions for the Euler equations (1.1) with initial data in H1, i.e., Theorem 1.1. First, we show that the solutions uR of some smoothed version of the equations exist and are uniformly bounded in H1. We then show that, as R→∞, the limit u of uR satisfies the original equations.
We first define a Fourier truncation SR as
^SRf(ξ):=1Rˆf(ξ), | (2.1) |
where ˆf:=F[f] denotes the Fourier transform of f. We construct the approximate Euler equations on the whole plane as:
{∂tuR+SRP[uR⋅∇uR]=0,divuR=0,uR|t=0=SRu0, | (2.2) |
where P=I+∇(−Δ)−1div. It can be checked that uR lies in the space
VR:={f∈L2(R2):ˆf is supported in B(0,R)}. | (2.3) |
We then define
F(uR,vR):=SRP[uR⋅∇vR]. | (2.4) |
It can be checked that F is Lipschitz on the space VR. Indeed, let uR1,vR1,uR2,vR2∈VR, by the definition of F, we have
‖F(uR1,vR1)−F(uR2,vR2)‖L2=‖PSR[uR1⋅∇vR1−uR2⋅∇vR2]‖L2, | (2.5) |
it follows that
‖F(uR1,vR1)−F(uR2,vR2)‖L2≤‖uR1⋅∇vR1−uR2⋅∇vR2‖L2. | (2.6) |
Notice that
‖uR1⋅∇vR1−uR2⋅∇vR2‖L2≤‖(uR1−uR2)⋅∇vR1‖L2+‖uR2⋅∇(vR1−vR2)‖L2≤CR2(‖uR1−uR2‖L2‖vR1‖L2+‖u2‖L2‖vR1−vR2‖L2), | (2.7) |
where we have used the Bernstein's Lemma. From (2.5)–(2.7), we conclude that F is Lipschitz on the space VR. Hence, by Picard's theorem for infinite-dimensional ordinary differential equations (see, for example, Theorem 3.1 in [14]), there exists a global smooth solution uR in VR.
We then show that uR is uniformly bounded in H1 with respect to R. Indeed, from Eq (2.2), the L2 bound of uR can be established.
Lemma 2.1. Suppose that u0∈L2(R2) is divergence-free. Then, we have
‖uR‖L2=‖uR0‖L2≤C‖u0‖L2. | (2.8) |
Proof. The L2 inner product of the first subequation of (2.2) with uR gives:
(∂tuR,uR)+(SRP[uR⋅∇uR],uR)=0. |
Observing that uR is divergence-free and lies in VR, we can deduce from the above identity that
ddt‖uR‖2L2=0, | (2.9) |
which implies (2.8) immediately.
To establish estimates of higher-order derivatives of uR, we resort to the vorticity-stream formulation:
{∂twR+SR[uR⋅∇wR]=0,uR=K[wR], | (2.10) |
where K:=12π(−x2|x|2,x1|x|2) is the kernel of the Biot-Savart law, and wR is the approximate vorticity.
Lemma 2.2. Suppose that u0∈H1(R2) is divergence-free. Then, we have
‖∇uR‖L2≤C‖∇u0‖L2, | (2.11) |
where C is a constant that is independent of time.
Proof. The L2 inner product of the first subequation of (2.10) with wR gives
(∂twR,wR)+(SR[uR⋅∇wR],wR)=0. |
Observing that uR and wR are divergence-free and lie in VR, therefore by using integration by parts, we arrive at
ddt‖wR‖2L2=0, | (2.12) |
which implies
‖wR‖L2=‖wR0‖L2≤‖∇u0‖L2. | (2.13) |
On the other hand, by the Calderón-Zygmund theorem, we can deduce from the second subequation of (2.10) that
‖∇uR‖L2≤C‖wR‖L2. | (2.14) |
By collecting (2.13) and (2.14), we obtain (2.11).
Remark 2.3. We observe that the kernel K of the Biot-Savart law decays at infinity like 1/r, which is not square integrable. This is the cause that we concern only the estimate of ∇uR from (2.10).
Collecting the above two lemmas, we immediately have the following result:
Proposition 2.4. Suppose that u0∈H1(R2) is divergence-free. Then, we have
‖uR‖H1≤C‖u0‖H1. | (2.15) |
Here C is a constant that does not depend on R.
With the uniform bound (2.15), we now proceed with the proof of Theorem 1.1. Indeed, from the Banach-Alaoglu theorem, we find that there exists a subsequence of uR(still denoted by uR) and some u such that
uR∗⇀u in L∞([0,∞);H1(R2)), | (2.16) |
and u satisfies (1.7).
On the other hand, by the Aubin-Lions lemma and Cantor's diagonal process, there exists a subsequence(still denoted by uR) that, for any bounded open subset K and T>0,
uR→u in L2(0,T;L2(K)). | (2.17) |
Furthermore, since uR satisfies (2.2), it follows that, for any test function φ∈C0([0,∞);C∞0,σ(R2)), we have
∫∞0(uR,∂tφ)−∫∞0(SRP[uR⊗uR],∇φ)=(uR0,φ(0)), | (2.18) |
Collecting (2.16) and (2.17) and letting R→∞, we conclude that u is a global solution of Eq (1.1). This completes the proof of Theorem 1.1.
In this section, we first establish some a priori estimates of uν. Then, we prove Theorem 1.2 by using approximation and compactness arguments. Finally, we show that the solution uν is unique.
Let us denote by Λs the fractional derivative operators defined in terms of Fourier transforms as follows:
F[Λsf](ξ)=|ξ|sˆf(ξ). | (3.1) |
We now prove the a priori estimate (1.9). Indeed, from the Navier-Stokes equations (1.8), a priori bound of ‖uν‖L2 can be established, and we have the following result:
Lemma 3.1. Suppose that uν is a solution of (1.8) with smooth initial data uν0. Then, we have
‖uν‖2L2+2ν∫t0‖∇uν‖2L2=‖uν0‖2L2. | (3.2) |
Proof. Multiplying the first subequation of (1.8) with uν, we have that
(∂tuν,uν)+(uν⋅∇uν,uν)+(∇p,uν)=ν(Δuν,uν). |
Observing that uν is divergence-free, we use integration by parts to deduce that
12ddt‖uν‖2L2+ν‖∇uν‖2L2=0, | (3.3) |
which implies (3.2) immediately.
To consider the case s=1, we should resort to the vorticity-stream formulation, which can be written as:
{∂twν+uν⋅∇wν+∇pν=νΔwν,uν=K∗wν, | (3.4) |
where K(x):=12π(−x2|x|2,x1|x|2) is the kernel of the Biot-Savart law, and wν:=∇⊥⋅uν is called the vorticity of the fluid.
From Eq (3.4), we can establish the following result:
Lemma 3.2. Suppose that uν is a solution of (3.4) with smooth initial data uν0. Then, we have
‖∇uν‖2L2+ν∫t0‖D2uν‖2L2≤c‖∇uν0‖2L2. | (3.5) |
Proof. By multiplying the first subequation of (3.4) with wν, we arrive at
(∂twν,wν)+(uν⋅∇wν,wν)=ν(Δwν,wν). |
Observing that uν is divergence-free, therefore, by using integration by parts, it follows that
ddt‖wν‖2L2=ν‖∇wν‖2L2. | (3.6) |
From the above identity, we deduce that
‖wν‖2L2+2ν∫t0‖∇wν‖2L2=‖w0‖2L2≤C‖∇uν0‖2L2. | (3.7) |
On the other hand, from the second subequation of (3.4) and the Calderón-Zygmund theorem, we find that, for s=0,1,
‖∇uν‖Hs≤C‖wν‖Hs. | (3.8) |
By collecting (3.7) and (3.8), we conclude that inequality (3.5) holds.
With the above two results, we establish the following proposition.
Proposition 3.3. Suppose that uν is a solution of (1.8) with smooth initial data uν0. Then uν satisfies the estimate
‖uν‖2Hs+ν∫t0‖∇uν‖2Hsds≤C1‖uν0‖2HsexpC2t, | (3.9) |
where C2≡C2(ν,u0) is a constant that depends on ν and the H2-norm of uν0, while does not depend on t. Moreover, the Sobolev order s is not necessarily an integer when s>2.
Proof. We first prove for the case s=2. We take the derivative Dα, |α|=2 of the Eq (1.8) and then take L2 inner product with Dαuν:
(∂tDαuν,Dαuν)−ν(DαΔuν,Dαuν)=−(Dα[uν⋅∇uν],Dαuν). | (3.10) |
It is easy to see that
(∂tDαuν,Dαuν)−ν(DαΔuν,Dαuν)=12ddt‖Dαuν‖L2+ν‖Dα∇uν‖L2. | (3.11) |
To handle the term on the right hand side of (3.10), we rewrite Dα[uν⋅∇uν] as
Dα[uν⋅∇uν]=∑β≤α,|β|>0Dβuν⋅∇[Dα−βuν]+uν⋅∇Dαuν. | (3.12) |
We use the Gagliardo-Nirenberg inequality to deduce that
|(∑β≤α,|β|>0Dβuν⋅∇[Dα−βuν],Dαuν)|≤C‖∇uν‖L2‖D2uν‖2L4≤C‖∇uν‖L2‖D2uν‖L2‖D3uν‖L2. | (3.13) |
On the other hand, observing that u is divergence-free, it follows
(uν⋅∇[Dαuν],Dαuν)=0. | (3.14) |
Collecting the above inequalities and using Young's inequality gives that
ddt‖D2uν‖L2+ν‖D3uν‖L2≤Cν‖∇uν‖2L2‖D2uν‖2L2. | (3.15) |
Since ‖uν‖H1 is uniformly bounded in time, it follows from the Gronwall's inequality that (3.9) holds for s=2.
We next prove the case s>2, where s is not necessarily an integer. Indeed, we apply the Λs operator to the first subequation of (1.8) and then take the L2 inner product with Λsu:
12ddt‖Λsuν‖2L2+ν‖Λs+1uν‖2L2=−(Λs[uν⋅∇uν],Λsuν). | (3.16) |
We observe that
‖Λs[uν⋅∇uν]‖L2≤‖Λs+1[uν⊗uν]‖L2, | (3.17) |
from Lemma 3.4 of [14], the above inequality yields
‖Λs[uν⋅∇uν]‖L2≤C‖uν‖L∞‖uν‖Hs+1. | (3.18) |
As H2(R2) is embedded in L∞(R2), therefore
|(Λs[uν⋅∇uν],Λsuν)|≤C‖uν‖H2‖uν‖Hs‖uν‖Hs+1. | (3.19) |
Then, by adding (3.2) and (3.16) together, we find that
12ddt‖uν‖2Hs+ν‖uν‖2Hs+1≤C‖uν‖H2‖uν‖Hs‖uν‖Hs+1. | (3.20) |
Then, the Holder's inequality implies
ddt‖uν‖2Hs+ν‖uν‖2Hs+1≤Cν‖uν‖2H2‖uν‖2Hs. | (3.21) |
Observing that ‖uν‖H2 is uniformly bounded, it follows from the Gronwall's inequality that ‖uν‖Hs is also uniformly bounded in time. That is, the inequality (3.9) holds. This completes the proof of Proposition 3.3.
We are now in the position to prove Theorem 1.2. We first consider the truncated Navier-Stokes equations on the whole plane:
{∂tuν,R+SRP[uν,R⋅∇uν,R]=νΔuν,R,divuν,R=0,uν,R|t=0=SRuν0, | (3.22) |
where P=I+∇(−Δ)−1div. It can be checked that uν,R lies in the space VR. We then define
F(uν,R,vν,R):=SRP[uν,R⋅∇vν,R]−νΔuν,R. | (3.23) |
It can be checked that F is Lipschitz on the space VR. Indeed, let uν,R1,vν,R1,uν,R2,vν,R2∈VR, by the definition of F, we have
‖F(uν,R1,vν,R1)−F(uν,R2,vν,R2)‖L2=‖PSR[uν,R1⋅∇vν,R1−uν,R2⋅∇vν,R2]−νPSR[Δuν,R1−Δuν,R2]‖L2, | (3.24) |
it follows that
‖F(uν,R1,vν,R1)−F(uν,R2,vν,R2)‖L2≤‖uν,R1⋅∇vν,R1−uν,R2⋅∇vν,R2‖L2+ν‖Δuν,R1−Δuν,R2‖L2. | (3.25) |
Notice that the first term on the right hand-side of the above inequality satisfies
‖uν,R1⋅∇vν,R1−uν,R2⋅∇vν,R2‖L2≤‖(uν,R1−uν,R2)⋅∇vν,R1‖L2+‖uν,R2⋅∇(vν,R1−vν,R2)‖L2≤CR2(‖uν,R1−uν,R2‖L2‖vν,R1‖L2+‖uν,R2‖L2‖vν,R1−vν,R2‖L2), | (3.26) |
where we have used the Bernstein's Lemma. Similarly, we can obtain that
‖Δuν,R1−Δuν,R2‖L2≤CR2‖uν,R1−uν,R2‖L2. | (3.27) |
From (3.24)–(3.27), we conclude that F is Lipschitz on the space VR. Hence, by Picard's theorem, there exists a local solution u in VR. Moreover, it can be checked that uν,R satisfies the estimate (1.9), thus the solution exists globally.
By the Banach-Alaoglu theorem, we find that there exists a subsequence of uν,R(still denoted by uν,R) and some uν such that
uν,R∗⇀uν in L∞([0,∞);Hs(R2)), | (3.28) |
and u satisfies (1.9).
On the other hand, by the Aubin-Lions lemma and Cantor's diagonal process, there exists a subsequence (still denoted by uν,R) that, for any bounded open subset K and T>0,
uν,R→uν in C([0,T];L2(K)). | (3.29) |
Collecting (3.28) and (3.29), we conclude that uν is a global solution of Eq (1.8).
We here prove for the case s=1, as it is well-known that the solution is unique for s=0. Let uν1,uν2 be two solutions of Eq (1.8). We subtract the first subequation of (1.8) for uν2 from the one for uν1, then multiply with uν1−uν2 and integrate over R2 to obtain that
12ddt‖uν1−uν2‖2L2+ν‖∇(uν1−uν2)‖2L2≤C‖∇uν1‖L2‖uν1−uν2‖L2‖∇(uν1−uν2)‖L2, | (3.30) |
where we have used the Gagliardo–Nirenberg interpolation inequality. Using Young's inequality, we get that
ddt‖uν1−uν2‖2L2≤Cν‖∇uν1‖2L2‖uν1−uν2‖2L2, | (3.31) |
then the Grönwall inequality implies that uν1≡uν2. We thus conclude that the solution is unique.
In this section, we prove Corollary 1.4. Let u0∈H1(R2) be divergence-free. Assume that uν0 is divergence-free and converges to u0 in H1 as ν→0. We now prove that, as ν→0, the solution uν of (1.8) with initial data uν0 converges to a solution u of (1.1) with initial data u0.
We observe that, for arbitrary ν>0, Eq (1.8) with initial data uν0 has a unique solution uν, and uν satisfies that
supt∈[0,∞)‖uν(t)‖2H1+ν∫∞0‖∇uν(t)‖2H1dt≤C‖uν0‖2H1≤C‖u0‖2H1. | (4.1) |
By the Banach-Alaoglu theorem, we find that there exists a subsequence of uν(still denoted by uν) and some u such that
uν∗⇀u in L∞([0,∞);H1(R2)), | (4.2) |
and u satisfies
supt∈[0,∞)‖u(t)‖H1≤C‖u0‖H1. | (4.3) |
On the other hand, by the Aubin-Lions lemma and Cantor's diagonal process, there exists a subsequence (still denoted by uν) that, for any bounded open subset K and T>0,
uν→u in C([0,T];L2(K)). | (4.4) |
Collecting (4.2) and (4.4), we conclude that u is a global weak solution of the Euler equations (1.1). This completes the proof of Corollary 1.4.
The incompressible Euler and Navier-Stokes equations in R2 are studied in this paper. We obtain global existence of weak solutions for initial data in H1. Moreover, it is proved that, as the viscosity tends to zero, the solution of the Navier-Stokes equations converges to the one of the Euler equations.
Shaoliang Yuan: Conceptualization, writing original draft, writing-review and editing; Lin Cheng: Investigation, writing-review and editing; Liangyong Lin: Investigation, writing-review and editing. All authors have read and agreed to the published version of the manuscript.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
S. Yuan was partially supported by the Doctoral Scientific Startup Fund of Fujian Polytechnic Normal University (No.404086).
The authors declare no conflict of interest.
[1] |
J. Bourgain, D. Li, Strong ill-posedness of the incompressible Euler equation in borderline Sobolev spaces, Invent. Math., 201 (2007), 97–157. http://dx.doi.org/10.1007/s00222-014-0548-6 doi: 10.1007/s00222-014-0548-6
![]() |
[2] |
J. M. Delort, Existence de nappes de tourbillon en dimension deux, J. Amer. Math. Soc., 4 (1991), 553–586. http://dx.doi.org/10.2307/2939269 doi: 10.2307/2939269
![]() |
[3] |
R. J. DiPerna, A. J. Majda, Concentrations in regularizations for 2-D incompressible flow, Comm. Pure Appl. Math., 40 (1987), 301–345. http://dx.doi.org/10.1002/cpa.3160400304 doi: 10.1002/cpa.3160400304
![]() |
[4] |
T. M. Elgindi, N. Masmoudi, L∞ ill-posedness for a class of equations arising in hydrodynamics, Arch. Rational. Mech. Anal., 235 (2020), 1979–2025. http://dx.doi.org/10.1007/s00205-019-01457-7 doi: 10.1007/s00205-019-01457-7
![]() |
[5] |
L. C. Evans, S. Müller, Hardy spaces and the two-dimensional Euler equations with nonnegative vorticity, J. Amer. Math. Soc., 7 (1994), 199–219. http://dx.doi.org/10.1090/S0894-0347-1994-1220787-3 doi: 10.1090/S0894-0347-1994-1220787-3
![]() |
[6] |
E. Hopf, Über die Anfangswertaufgabe für die hydrodynamischen Grundgleichungen. Erhard Schmidt zu seinem 75. Geburtstag gewidmet, Math. Nachr., 4 (1950), 213–231. http://dx.doi.org/10.1002/mana.3210040121 doi: 10.1002/mana.3210040121
![]() |
[7] |
T. Kato, G. Ponce, Commutator estimates and the Euler and Navier-Stokes equations, Comm. Pure Appl. Math., 41 (1988), 891–907. http://dx.doi.org/10.1002/cpa.3160410704 doi: 10.1002/cpa.3160410704
![]() |
[8] | O. A. Ladyzhenskaya, The mathematical theory of viscous incompressible flow, New York: Gordon and Breach, 1969. http://dx.doi.org/10.1137/1013008 |
[9] |
J. Leray, Sur le mouvement d'un liquide visqueux emplissant l'espace, Acta Math., 63 (1934), 193–248. http://dx.doi.org/10.1007/BF02547354 doi: 10.1007/BF02547354
![]() |
[10] | J. L. Lions, Un théorème d'existence et unicité dans les équations de Navier-Stokes en dimension 2, C. R. Math. Acad. Sci. Paris, 248 (1959), 3519–3521. |
[11] | J. L. Lions, Quelques méthodes de résolution des problèmes aux limites Non-Linéaires, Dunod, 1969. |
[12] |
J. Liu, Z. Xin, Convergence of vortex methods for weak solutions to the 2-d Euler equations with vortex sheet data, Comm. Pure Appl. Math., 48 (1995), 611–628. http://dx.doi.org/10.1002/cpa.3160480603 doi: 10.1002/cpa.3160480603
![]() |
[13] |
A. J. Majda, Remarks on weak solutions for vortex sheets with a distinguished sign, Indiana Univ. Math. J., 42 (1993), 921–939. http://dx.doi.org/10.1512/iumj.1993.42.42043 doi: 10.1512/iumj.1993.42.42043
![]() |
[14] |
A. J. Majda, A. L. Bertozzi, A. Ogawa, Vorticity and incompressible flow. Cambridge texts in applied mathematics, Appl. Mech. Rev., 55 (2001), 87–135. http://dx.doi.org/10.1115/1.1483363 doi: 10.1115/1.1483363
![]() |
[15] |
N. Masmoudi, Remarks about the inviscid limit of the Navier-Stokes system, Comm. Math. Phys., 270 (2007), 777–788. http://dx.doi.org/10.1007/s00220-006-0171-5 doi: 10.1007/s00220-006-0171-5
![]() |
[16] |
J. Serrin, On the interior regularity of weak solutions of the Navier-Stokes equations, Arch. Ration. Mech. Anal., 9 (1962), 187–195. http://dx.doi.org/10.1007/BF00253344 doi: 10.1007/BF00253344
![]() |
[17] |
M. Vishik, Incompressible flows of an ideal fluid with vorticity in borderline spaces of Besov type, Ann. Sci. Éc. Norm. Supér., 32 (1999), 769–812. http://dx.doi.org/10.1016/S0012-9593(00)87718-6 doi: 10.1016/S0012-9593(00)87718-6
![]() |
[18] |
V. I. Yudovich, Non-stationary flows of an ideal incompressible liquid, Zh. Vychisl. Mat. Mat. Fiz., 3 (1963), 1032–1066. http://dx.doi.org/10.1016/0041-5553(63)90247-7 doi: 10.1016/0041-5553(63)90247-7
![]() |
[19] |
V. I. Yudovich, Uniqueness theorem for the basic nonstationary problem in the dynamics of an ideal incompressible fluid, Math. Res. Lett., 2 (1995), 27–38. http://dx.doi.org/10.4310/mrl.1995.v2.n1.a4 doi: 10.4310/mrl.1995.v2.n1.a4
![]() |