Loading [MathJax]/jax/output/SVG/jax.js

On the local and global existence of solutions to 1d transport equations with nonlocal velocity

  • Received: 01 June 2018 Revised: 01 February 2019
  • Primary: 35A01; Secondary: 35D30, 35Q35, 35Q86

  • We consider the 1D transport equation with nonlocal velocity field:

    $ θt+uθx+νΛγθ=0,u=N(θ), $

    where $ \mathcal{N} $ is a nonlocal operator and $ \Lambda^{\gamma} $ is a Fourier multiplier defined by $ \widehat{\Lambda^{\gamma} f}(\xi) = |\xi|^{\gamma}\widehat{f}(\xi) $. In this paper, we show the existence of solutions of this model locally and globally in time for various types of nonlocal operators.

    Citation: Hantaek Bae, Rafael Granero-Belinchón, Omar Lazar. On the local and global existence of solutions to 1d transport equations with nonlocal velocity[J]. Networks and Heterogeneous Media, 2019, 14(3): 471-487. doi: 10.3934/nhm.2019019

    Related Papers:

    [1] Hantaek Bae, Rafael Granero-Belinchón, Omar Lazar . On the local and global existence of solutions to 1d transport equations with nonlocal velocity. Networks and Heterogeneous Media, 2019, 14(3): 471-487. doi: 10.3934/nhm.2019019
    [2] François James, Nicolas Vauchelet . One-dimensional aggregation equation after blow up: Existence, uniqueness and numerical simulation. Networks and Heterogeneous Media, 2016, 11(1): 163-180. doi: 10.3934/nhm.2016.11.163
    [3] Young-Pil Choi, Seok-Bae Yun . A BGK kinetic model with local velocity alignment forces. Networks and Heterogeneous Media, 2020, 15(3): 389-404. doi: 10.3934/nhm.2020024
    [4] Raffaele Scandone . Global, finite energy, weak solutions to a magnetic quantum fluid system. Networks and Heterogeneous Media, 2025, 20(2): 345-355. doi: 10.3934/nhm.2025016
    [5] Hyeontae Jo, Hwijae Son, Hyung Ju Hwang, Eun Heui Kim . Deep neural network approach to forward-inverse problems. Networks and Heterogeneous Media, 2020, 15(2): 247-259. doi: 10.3934/nhm.2020011
    [6] Leonid Berlyand, Mykhailo Potomkin, Volodymyr Rybalko . Sharp interface limit in a phase field model of cell motility. Networks and Heterogeneous Media, 2017, 12(4): 551-590. doi: 10.3934/nhm.2017023
    [7] Hyeong-Ohk Bae, Hyoungsuk So, Yeonghun Youn . Interior regularity to the steady incompressible shear thinning fluids with non-Standard growth. Networks and Heterogeneous Media, 2018, 13(3): 479-491. doi: 10.3934/nhm.2018021
    [8] Verónica Anaya, Mostafa Bendahmane, David Mora, Ricardo Ruiz Baier . On a vorticity-based formulation for reaction-diffusion-Brinkman systems. Networks and Heterogeneous Media, 2018, 13(1): 69-94. doi: 10.3934/nhm.2018004
    [9] José Antonio Carrillo, Yingping Peng, Aneta Wróblewska-Kamińska . Relative entropy method for the relaxation limit of hydrodynamic models. Networks and Heterogeneous Media, 2020, 15(3): 369-387. doi: 10.3934/nhm.2020023
    [10] Young-Pil Choi, Seung-Yeal Ha, Jeongho Kim . Propagation of regularity and finite-time collisions for the thermomechanical Cucker-Smale model with a singular communication. Networks and Heterogeneous Media, 2018, 13(3): 379-407. doi: 10.3934/nhm.2018017
  • We consider the 1D transport equation with nonlocal velocity field:

    $ \begin{equation*} \label{intro eq} \begin{split} &\theta_t+u\theta_x+\nu \Lambda^{\gamma}\theta = 0, \\ & u = \mathcal{N}(\theta), \end{split} \end{equation*} $

    where $ \mathcal{N} $ is a nonlocal operator and $ \Lambda^{\gamma} $ is a Fourier multiplier defined by $ \widehat{\Lambda^{\gamma} f}(\xi) = |\xi|^{\gamma}\widehat{f}(\xi) $. In this paper, we show the existence of solutions of this model locally and globally in time for various types of nonlocal operators.



    In this paper, we study transport equations with nonlocal velocity. One of the most well-known equation is the two dimensional Euler equation in vorticity form,

    $ \omega_{t}+u\cdot \nabla \omega = 0, $

    where the velocity $ u $ is recovered from the vorticity $ \omega $ through

    $ u = \nabla^\perp(-\Delta)^{-1}\omega \quad \text{or equivalently} \quad \widehat{u}(\xi) = \frac{i\xi^{\perp}\;\;}{|\xi|^{2}\;\;}\widehat{\omega}(\xi). $

    Other nonlocal and quadratically nonlinear equations, such as the surface quasi-geostrophic equation, the incompressible porous medium equation, Stokes equations, magneto-geostrophic equation in multi-dimensions, have been studied intensively as one can see in [1,2,5,6,7,8,9,13,14,15,16,19,20,23,24,25] and references therein.

    We here consider the 1D transport equations with nonlocal velocity field of the form

    $ θt+uθx+νΛγθ=0,xR, $ (1a)
    $ u=N(θ), $ (1b)

    where $ \mathcal{N} $ is typically expressed by a Fourier multiplier. The differential operator $ \Lambda^{\gamma} = (\sqrt{-\Delta})^{\gamma} $ is defined by the action of the following kernels [10]:

    $ Λγf(x)=cγp.v.Rf(x)f(y)|xy|1+γdy, $ (2)

    where $ c_{\gamma}>0 $ is a normalized constant. Alternatively, we can define $ \Lambda^{\gamma} = (\sqrt{-\Delta})^{\gamma} $ as a Fourier multiplier: $ \widehat{\Lambda^{\gamma} f}(\xi) = |\xi|^{\gamma}\widehat{f}(\xi) $. The study of 1 is mainly motivated by [11] where Córdoba, Córdoba, and Fontelos proposed the following 1D model

    $ θt+uθx=0, $ (3a)
    $ u=Hθ,(H: the Hilbert transform) $ (3b)

    for the 2D surface quasi-geostrophic equation and proved the finite time blow-up of smooth solutions. In this paper, we deal with 3a-3b and its variations with the following objectives.

    (1) The existence of weak solution with rough initial data. The existence of global-in-time solutions is possible even if strong solutions blow up in finite time, as in the case of the Burgers' equation.

    (2) The existence of strong solution when the velocity $ u $ is more singular than $ \theta $. We intend to see the competitive relationship between nonlinear terms and viscous terms.

    More specifically, the topics covered in this paper can be summarized as follows.

    The model 1: $ \mathcal{N} = -\mathcal{H} $ and $ \nu = 0 $. We first show the existence of local-in-time solution in a critical space under the scaling $ \theta_{0}(x)\mapsto \theta_{0}(\lambda x) $. We then introduce the notion of a weak super-solution and obtain a global-in-time weak super-solution with $ \theta_{0}\in L^{1}\cap L^{\infty} $ and $ \theta_{0}\ge 0 $.

    The model 2: $ \mathcal{N} = -\mathcal{H}(\partial_{xx} )^{-\alpha} $, $ \alpha>0 $, $ \nu = 1 $, and $ \gamma>0 $. This is a regularized version of 3a-3b which is also closely related to many equations as mentioned in [3]. In this case, we show the existence of weak solutions globally in time under weaker conditions on $ \alpha $ and $ \gamma $ compared to [3].

    The model 3: $ \mathcal{N} = -\mathcal{H}(\partial_{xx} )^{\beta} $, $ \beta>0 $, $ \nu = 1 $, and $ \gamma>0 $. Since $ \beta>0 $, the velocity field is more singular than the previous two models. In this case, we show the existence of strong solutions locally in time in two cases: (1) $ 0<\beta\leq \frac{\gamma}{4} $ when $ 0<\gamma<2 $ and (2) $ 0<\beta<1 $ when $ \gamma = 2 $. We also show the existence of strong solutions for $ 0<\beta<\frac{1}{2} $ and $ \gamma = 2 $ with rough initial data. We finally show the existence of strong solutions globally in time with $ 0<\beta<\frac{1}{4} $ and $ \gamma = 2 $.

    We will give detailed statements and proofs of our results in Section 3–5.

    All constants will be denoted by $ C $ that is a generic constant. In a series of inequalities, the value of $ C $ can vary with each inequality. We use following notation: for a Banach space $ X $,

    $ C_{T}X = C([0, T]:X), \quad L^{p}_{T}X = L^{p}(0, T:X). $

    The spatial derivatives are defined as

    $ \partial^{l}f(t, x) = \frac{\partial^{l} f}{\partial x^{l}}(t, x), \quad l\in \mathbb{N}. $

    For $ l = 1, 2, 3 $, we also use the followings:

    $ f_{x}, \quad f_{xx} = \partial_{xx}f, \quad f_{xxx}. $

    We now give some properties of the Hilbert transform and related function spaces. The Hilbert transform is defined by

    $ \mathcal{H}f(x) = \text{p.v.} \int_{\mathbb{R}} \frac{f(y)}{x-y}dy. $

    We will use the BMO space and its dual which is the Hardy space $ \mathcal{H}^1 $ which consists of those $ f $ such that $ f $ and $ \mathcal{H}f $ are integrable [17,Chapter 6]. By the following Cotlar formula [12]

    $ 2H(fHf)=(Hf)2f2, $ (4)

    we have $ f\mathcal{H}f \in \mathcal{H}^1 $ and for any $ f\in L^{2} $,

    $ fHfH1f2L2. $ (5)

    We have $ \Lambda f(x) = \mathcal{H}f_{x}(x) $ by using $ \widehat{Hf}(\xi) = -\text{sgn}(\xi) \widehat{f}(\xi) $, where $ \Lambda $ is defined in 2.

    Since we are dealing with equations on $ \mathbb{R} $, we state some definitions and function spaces on $ \mathbb{R} $.

    Let $ f\in \mathcal{S}' $, a tempered distribution. Then, its Fourier transform is defined by

    $ \widehat{f}(\xi) = \int_{\mathbb{R}}f(x)e^{-2\pi i x\xi}dx. $

    Let $ s\in \mathbb{R} $. The energy space $ H^{s} $ is defined by

    $ H^{s}(\mathbb{R}) = \left\{f\in \mathcal{S}': \left\|f\right\|^{2}_{H^{s}} = \int_{\mathbb{R}}(1+|\xi|^{2})^{s}\left|\widehat{f}(\xi)\right|^{2}d\xi < \infty\right\}. $

    We also define homogeneous spaces:

    $ \dot{H}^{s}(\mathbb{R}) = \left\{f\in \mathcal{S}': \left\|f\right\|^{2}_{\dot{H}^{s}} = \int_{\mathbb{R}}|\xi|^{2s}\left|\widehat{f}(\xi)\right|^{2}d\xi < \infty\right\}. $

    We note that for $ s>0 $ and $ \sigma>0 $,

    $ ˙HsHsσ $ (6)

    because

    $ \left\|f\right\|^{2}_{H^{-(s+\sigma)}} = \int_{\mathbb{R}}\frac{\left|\widehat{f}(\xi)\right|^{2}}{(1+|\xi|^{2})^{s+\sigma}}d\xi = \int_{\mathbb{R}}\frac{|\xi|^{2s}}{(1+|\xi|^{2})^{s+\sigma}}\frac{\left|\widehat{f}(\xi)\right|^{2}}{|\xi|^{2s}}d\xi\leq C \left\|f\right\|^{2}_{\dot{H}^{-s}}. $

    In this paper, we also use two estimations in [18].

    (1) Fractional product rule. For $ s>0 $ and $ p $, $ p_{i} $, and $ q_{i} $ such that

    $ \frac{1}{p} = \frac{1}{p_{1}}+\frac{1}{q_{1}} = \frac{1}{p_{2}}+\frac{1}{q_{2}}, \quad 1\leq p < \infty, \ \ p_{i}, q_{i} \ne 1, $

    we have the following estimations:

    $ Λs(fg)LpC[ΛsfLp1gLq1+fLp2ΛsgLq2], $ (7)

    and

    $ (IΛ)s(fg)LpC[(IΛ)sfLp1gLq1+fLp2(IΛ)sgLq2], $ (8)

    where $ (I-\Delta) $ is defined as the Fourier multiplier whose symbol is $ 1+|\xi|^{2} $.

    (2) Commutator estimate.

    $ |l|2l(fg)flgL2C(fxLgxL2+fxxL2gL). $ (9)

    We here briefly introduce the Littlewood-Paley theory based on [4]. We first provide notation and definitions in the Littlewood-Paley theory. Let $ \mathcal{C} $ be the ring of center 0, of small radius $ \frac{3}{4} $ and great radius $ \frac{8}{3} $. We take smooth radial functions $ (\chi, \phi) $ with values in $ [0, 1] $ that are supported on the ball $ B_{\frac{4}{3}}(0) $ and $ \mathcal{C} $, respectively, and satisfy

    $ χ(ξ)+j=0ϕ(2jξ)=1   ξRd,j=ϕ(2jξ)=1   ξRd{0},|jj|2  supp ϕ(2j)supp ϕ(2j)=,j1  supp χsupp ϕ(2j)=. $ (10)

    From now on, we use the notation

    $ \phi_{j}(\xi) = \phi\left(2^{-j}\xi\right). $

    We define dyadic blocks and lower frequency cut-off functions.

    $ h=F1ϕ,˜h=F1χ,Δjf=ϕj(D)f=2jdRdh(2jy)f(xy)dy,Sjf=χ(2jD)f=2jdRd˜h(2jy)f(xy)dy,Δ1f=χ(D)f=Rd˜h(y)f(xy)dy. $ (11)

    Then, the homogeneous Littlewood-Paley decomposition is given by

    $ f = \sum\limits_{j\in \mathbb{Z}} \Delta_{j}f \ \ \text{in} \ \ \mathcal{S}^{'}_{h}, $

    where $ \mathcal{S}^{'}_{h} $ is the space of tempered distributions $ u\in \mathcal{S}^{'} $ such that

    $ \lim\limits_{j\rightarrow -\infty}S_{j}u = 0\quad \text{in $\mathcal{S}'$}. $

    We now define the homogeneous Besov spaces:

    $ \dot{B}^{s}_{p, q} = \left\{f\in \mathcal{S}^{'}_{h}: \ \left\|f\right\|_{\dot{B}^{s}_{p, q}} = \left\|2^{js}\left\|\Delta_{j}f\right\|_{L^{p}}\right\|_{l^{q}(\mathbb{Z})} < \infty \right\}. $

    We recall Bernstein's inequality in 1D : for $ 1\leq p\leq q \leq \infty $ and $ k\in \mathbb{N} $,

    $ sup|α|=kαΔjfLpC2jkΔjfLp,ΔjfLqC2j(1p1q)ΔjfLp. $ (12)

    Moreover, the Besov spaces enjoy nice scaling properties. Let $ f_{\lambda}(x) = f(\lambda x) $. Then, there exists a constant $ C $ such that

    $ C1fλ˙Bsp,rλs3pf˙Bsp,rCfλ˙Bsp,r. $ (13)

    We also have the following commutator estimate.

    Lemma 2.1 (Commutator estimate). For $ f, g \in \mathcal{S} $ (Schwarz class)

    $ \left\|[f, \Delta_{j}]g_{x}\right\|_{L^{2}}\leq Cc_{j}2^{-\frac{3}{2}j} \left\|f_{x}\right\|_{\dot{B}^{\frac{1}{2}}_{2, 1}} \left\|g\right\|_{\dot{B}^{\frac{3}{2}}_{2, 1}}, \quad \sum\limits^{\infty}_{j = -\infty}c_{j}\leq 1. $

    We finally introduce Simon's compactness.

    Lemma 2.2. [26] Let $ X_{0} $, $ X_{1} $, and $ X_{2} $ be Banach spaces such that $ X_{0} $ is compactly embedded in $ X_{1} $ and $ X_{1} $ is a subset of $ X_{2} $. Then, for $ 1\leq p<\infty $, the set $ \left\{v\in L^{p}_{T}X_{0}: \ \frac{\partial v}{\partial t}\in L^{1}_{T}X_{2}\right\} $ is compactly embedded in $ L^{p}_{T}X_{1} $.

    We now study 1a-1b with $ \mathcal{N} = -\mathcal{H} $ and $ \nu = 0 $ which is nothing but 3a-3b:

    $ θt(Hθ)θx=0, $ (14a)
    $ θ(0,x)=θ0(x). $ (14b)

    The local well-posedness of 14a-14b is established in $ H^{2} $ ([2]) and $ H^{\frac{3}{2}-\gamma} $ with the viscous term $ \Lambda^{\gamma}\theta $ ([14]). To improve these results, we notice that 14a-14b has the following scaling invariant property: if $ \theta(t, x) $ is a solution of 14a-14b, then so is $ \theta_{\lambda}(t, x) = \theta(\lambda t, \lambda x) $. So, we take initial data in a space whose norm is closely invariant under the scaling:

    $ \theta_{0}(x)\mapsto \theta_{\lambda 0}(x) = \theta_{0}(\lambda x). $

    In this paper, we take the space $ \dot{B}^{\frac{3}{2}}_{2, 1} $ because there is a constant $ C $ such that

    $ C^{-1} \left\|\theta_{\lambda 0}\right\|_{\dot{B}^{\frac{3}{2}}_{2, 1}} \leq \|\theta_{0}\|_{\dot{B}^{\frac{3}{2}}_{2, 1}} \leq C \left\|\theta_{\lambda 0}\right\|_{\dot{B}^{\frac{3}{2}}_{2, 1}} $

    by taking $ s = \frac{3}{2} $, $ p = 2 $, and $ r = 1 $ in 13. The first result in this paper is the following theorem.

    Theorem 3.1. For any $ \theta_{0}\in \dot{B}^{\frac{3}{2}}_{2, 1} $, there exists $ T = T(\|\theta_0\|) $ such that a unique solution of 14a-14b exists in $ C_{T}\dot{B}^{\frac{3}{2}}_{2, 1} $.

    Proof. We only provide a priori estimates of $ \theta $ in the space stated in Theorem 3.1. The other parts, including the approximation procedure, are rather standard.

    We apply $ \Delta_{j} $ to 14a, multiply by $ \Delta_{j}\theta $, and integrate the resulting equation over $ \mathbb{R} $ to get

    $ 12ddtΔjθ2L2=RΔj((Hθ)θx)Δjθdx=R((Hθ)Δjθx)Δjθdx+RΔj((Hθ)θx)ΔjθdxR((Hθ)Δjθx)Δjθdx=R((Hθ)Δjθx)Δjθdx+R{[Δj,Hθ]Δjθx}Δjθdx=12R(Hθ)x|Δjθ|2dx+R{[Δj,Hθ]Δjθx}Δjθdx. $ (15)

    By the Bernstein inequality, we have

    $ HθxLCθ˙B322,1. $ (16)

    We then apply Lemma 2.1 to the second term in the right-hand side of 15 to obtain

    $ R[Δj,Hθ]ΔjθxΔjθdxCcj232jθ2˙B322,1ΔjθL2. $ (17)

    By 15, 16, and 17, we have

    $ \frac{d}{dt}\|\theta\|^{2}_{\dot{B}^{\frac{3}{2}}_{2, 1}}\leq C\|\theta\|^{3}_{\dot{B}^{\frac{3}{2}}_{2, 1}}, $

    from which we deduce

    This completes the proof.

    We next consider 14a-14b with rough initial data. More precisely, we assume that $ \theta_{0} $ satisfies the following conditions

    $ θ00,θ0L1L. $ (18)

    Since $ \theta $ satisfies the transport equation, we have

    $ θ(t,x)0,θL(R)for all time. $ (19)

    If we follow the usual weak formulation of 14a-14b, for all $ \psi\in C^{\infty}_{c}([0, T)\times \mathbb{R}) $

    $ T0R[θψt+(Hθ)θψx+(Λθ)θψ]dxdt=Rθ0(x)ψ(x,0)dx. $ (20)

    For $ \theta_{0}\ge 0 $, there is gain of a half derivative from the structure of the nonlinearity, that is

    $ θ(t)L1+t0Λ12θ(s)2L2ds=θ0L1. $ (21)

    So, we can rewrite the left-hand side of 20 as

    $ \int^{T}_{0}\int_{\mathbb{R}} \left[-\theta \psi_{t} + \left(\mathcal{H}\theta\right)\theta \psi_{x}+\Lambda^{\frac{1}{2}}\theta \left[\Lambda^{\frac{1}{2}}, \psi\right]\theta +\left|\Lambda^{\frac{1}{2}}\theta\right|^{2}\psi\right] dxdt = \int_{\mathbb{R}}\theta_{0}(x)\psi(x, 0)dx. $

    However, the $ \dot{H}^{\frac{1}{2}} $ regularity derived from 21 is not enough to pass to the limit in

    $ \int^{T}_{0}\int_{\mathbb{R}} \left|\Lambda^{\frac{1}{2}}\theta^{\epsilon}\right|^{2}\psi dxdt $

    from the $ \epsilon $-regularized equations described below. So, we introduce a new notion of solution. Let

    $ \mathcal{A}_{T} = L^{\infty}_{T}\left(L^{1}\cap L^{\infty}\right)\cap L^{2}_{T}H^{\frac{1}{2}}. $

    Definition 3.2. We say $ \theta $ is a weak super-solution of 14a-14b on the time interval $ [0, T] $ if $ \theta(t, x)\ge 0 $ for all $ t\in [0, T] $, $ \theta\in \mathcal{A}_{T} $, and for each nonnegative $ \psi\in C^{\infty}_{c}([0, T)\times\mathbb{R}) $,

    $ T0R[θψt+(Hθ)θψx+Λ12θ[Λ12,ψ]θ+|Λ12θ|2ψ]dxdtRθ0(x)ψ(x,0)dx. $ (22)

    To deal with the third term in 22, we use the following Lemma.

    Lemma 3.3. [3] For $ f\in L^{\frac{3}{2}} $, $ g\in L^{\frac{3}{2}} $ and $ \psi \in W^{1, \infty} $, we have

    $ \left\| \left[\Lambda^{\frac{1}{2}}, \psi \right]f- \left[\Lambda^{\frac{1}{2}}, \psi \right]g\right\|_{L^{6}}\leq C\|\psi\|_{W^{1, \infty}} \left\|f-g\right\|_{L^{\frac{3}{2}}}. $

    The second result in our paper is the following theorem.

    Theorem 3.4. For any $ \theta_{0} $ satisfying 18, there exists a weak super-solution of 14a-14b in $ \mathcal{A}_{T} $.

    Proof. We first regularize initial data as $ \theta^{\epsilon}_{0} = \rho_{\epsilon}\ast \theta_{0} $ where $ \rho_{\epsilon} $ is a standard mollifier that preserve the positivity of the regularized initial data. We then regularize the equation by introducing the Laplacian term with a coefficient $ \epsilon>0 $, namely

    $ θϵtHθϵθϵx=ϵθϵxx. $ (23)

    For the proof of the existence of a global-in-time smooth solution we refer to [21]. Moreover, $ \theta^{\epsilon} $ satisfies that $ \theta^{\epsilon}\ge 0 $ and

    $ θϵ(t)L1+θϵ(t)L+t0Λ12θϵ(s)2L2dsθ0L1+θ0L. $ (24)

    Therefore, $ (\theta^{\epsilon}) $ is bounded in $ \mathcal{A}_{T} $ uniformly in $ \epsilon>0 $.

    The first two terms on the left-hand side of 24 imply

    $ \left\|\theta^{\epsilon}\right\|_{L^{p}_{T}L^{q}}\leq \left\|\theta_{0}\right\|_{L^{1}} + \left\|\theta_{0}\right\|_{L^{\infty}} $

    for any $ p, q\in[1, \infty] $. In particular,

    $ \mathcal{H}\theta^{\epsilon}\in L^{4}_{T}L^{2}, \quad \theta^{\epsilon}\in L^{2}_{T}L^{\infty}. $

    These two bounds imply

    $ \left(\left(\mathcal{H}\theta^{\epsilon}\right)\theta^{\epsilon}\right)_x\in L^{\frac{4}{3}}_{T}\dot{H}^{-1}\subset L^{\frac{4}{3}}_{T}H^{-2} $

    by the embedding 6. From $ \theta^{\epsilon}\in L^{2}_{T}\dot{H}^{\frac{1}{2}} $, we also have

    $ \epsilon \theta^{\epsilon}_{xx}\in L^{2}_{T}\dot{H}^{-\frac{3}{2}}\subset L^{2}_{T}H^{-2} $

    by the embedding 6. Moreover, for any $ \phi \in H^{2} $,

    $ \int_{\mathbb{R}}\left|\theta^{\epsilon} \Lambda\theta^{\epsilon} \phi \right| dx \leq \left\|\Lambda^{\frac{1}{2}}\theta^{\epsilon}\right\|^{2}_{L^{2}} \left\|\phi\right\|_{L^{\infty}} + \left\|\Lambda^{\frac{1}{2}}\theta^{\epsilon}\right\|_{L^{2}} \left\|\theta^{\epsilon}\right\|_{L^{\infty}} \left\|\Lambda^{\frac{1}{2}}\phi\right\|_{L^{\infty}} $

    which implies that

    $ \theta^{\epsilon} \Lambda\theta^{\epsilon} \in L^{1}_{T}H^{-2}. $

    Combining all together, we obtain

    $ \theta^{\epsilon}_{t} = \mathcal{H}\theta^{\epsilon}\theta^{\epsilon}_x +\epsilon \theta^{\epsilon}_{xx} = \left(\mathcal{H}\theta^{\epsilon}\theta^{\epsilon}\right)_x -\theta^{\epsilon} \Lambda\theta^{\epsilon}+\epsilon \theta^{\epsilon}_{xx}\in L^{1}_{T}H^{-2}. $

    To pass to the limit into the weak super-solution formulation, we extract a subsequence of $ \left(\theta^{\epsilon}\right) $, using the same index $ \epsilon $ for simplicity, and a function $ \theta \in \mathcal{A}_T $ such that

    $ θϵθinLpTLqfor all p,q(1,),θϵθinL2TH12,θϵθin L2TLploc for all 1<p<, $ (25)

    where we use Lemma 2.2 for the strong convergence with

    $ X_0 = L^2_{T}H^{\frac{1}{2}}, \quad X_{1} = L^2_{T}L^p_{\text{loc}}, \quad X_2 = L^1_{T}H^{-2}. $

    We now multiply 23 by a nonnegative test function $ \psi\in \mathcal{C}^{\infty}_{c}\left([0, T)\times\mathbb{R}\right) $ and integrate over $ \mathbb{R} $. Then,

    $ T0R[θϵψt+(Hθϵ)θϵψxI+ϵθϵψxx]dxdtRθϵ0(x)ψ(0,x)dx=T0RΛ12θϵ[Λ12,ψ]θϵIIdxdtT0R|Λ12θϵ|2ψIIIdxdt. $ (26)

    We note that we are able to rearrange terms in the usual weak formulation into 26 since $ \theta^{\epsilon} $ is smooth. By the strong convergence in 25, we can pass to the limit to $ \text{I} $. Moreover, since

    $ \left[\Lambda^{\frac{1}{2}}, \psi \right]\theta^{\epsilon} \rightarrow \left[\Lambda^{\frac{1}{2}}, \psi \right]\theta $

    strongly in $ L^{2}_{T}L^{6} $ by Lemma 3.3 and the weak convergence in 25, we can pass to the limit to $ \text{II} $. Lastly, define

    $ g^\epsilon = \Lambda^{\frac{1}{2}}\theta^{\epsilon}\sqrt{\psi} \quad \text{and} \quad g = \Lambda^{\frac{1}{2}}\theta\sqrt{\psi}. $

    We then have that $ g^\epsilon\rightharpoonup g $ in $ L^2([0, T]\times\mathbb{R}) $. Since the $ L^2 $ norm is weakly lower semicontinuous, we find that

    $ \liminf\limits_{\epsilon\rightarrow 0}\|g^\epsilon\|_{L^2([0, T]\times\mathbb{R})}\geq\|g\|_{L^2([0, T]\times\mathbb{R})}, $

    or, equivalently,

    $ \liminf\limits_{\epsilon\rightarrow 0}\int^{T}_{0}\int_{\mathbb{R}} \left|\Lambda^{\frac{1}{2}}\theta^{\epsilon}\right|^{2}\psi dxdt \ge \int^{T}_{0}\int_{\mathbb{R}} \left|\Lambda^{\frac{1}{2}}\theta\right|^{2}\psi dxdt. $

    Combining all the limits together, we obtain that

    $ T0R[θψt+(Hθ)θψx+Λ12θ[Λ12,ψ]θ+|Λ12θ|2ψ]dxdtRθ0(x)ψ(x,0)dx. $ (27)

    This completes the proof.

    We now consider the following equation:

    $ θt(H(xx)αθ)θx+Λγθ=0, $ (28a)
    $ θ(0,x)=θ0(x), $ (28b)

    where $ \alpha, \gamma>0 $. In this case, we focus on the existence of weak solutions under some conditions of $ (\alpha, \gamma) $. As before, we assume that $ \theta_{0} $ satisfies the following conditions

    $ θ00,θ0L1L. $ (29)

    Let

    $ \mathcal{B}_{T} = L^{\infty}_{T}\left(L^{1}\cap L^{\infty}\right)\cap L^{2}_{T}H^{\frac{\gamma}{2}}. $

    Definition 4.1. We say $ \theta $ is a weak solution of 28a-28b on the time interval $ [0, T] $ if $ \theta(t, x)\ge 0 $ for all $ t\in [0, T] $, $ \theta\in \mathcal{B}_{T} $, and for each $ \psi\in C^{\infty}_{c}([0, T)\times\mathbb{R}) $,

    $ T0R[θψt(H(xx)αθ)θψxΛ1γ2(xx)αθΛγ2(θψ)θΛγψ]dxdt=Rθ0(x)ψ(x,0)dx. $

    The third result in the paper is the following.

    Theorem 4.2. Suppose that two positive numbers $ \alpha $ and $ \gamma $ satisfy

    $ 0<γ<1,12γ2<α<12. $ (30)

    Then, for any $ \theta_{0} $ satisfying 29, there exists a weak solution of 28a-28b in $ \mathcal{B}_{T} $ for all $ T>0 $.

    Proof. As in the proof of Theorem 3.4, we regularize $ \theta_{0} $ and the equation as

    $ θϵ0=ρϵθ0,θϵt(H(xx)αθϵ)θϵx+Λγθϵ=ϵθϵxx. $ (31)

    Then, the corresponding $ \theta^{\epsilon} $ satisfies

    $ θϵ(t,x)0,θϵ(t)Lθ0Lfor all time  $ (32)

    and

    $ θϵ(t)L1+t0Λ12(xx)α2θϵ(s)2L2dsθ0L1. $ (33)

    We next multiply 31 by $ \theta^{\epsilon} $ and integrate over $ \mathbb{R} $. Then,

    $ 12ddtθϵ(t)2L2+Λγ2θϵ(t)2L2+ϵθϵx2L2=12R{Λ(xx)αθϵ(t)}(θϵ(t))2dx=12R{(1Δ)γ4Λ(xx)αθϵ(t)}(1Δ)γ4(θϵ(t))2dxC(1Δ)γ4Λ(xx)αθϵ(t)L2(1Δ)γ4θϵ(t)L2θϵ(t)L, $

    where we use the fractional product rule 8 to obtain

    $ \left\|(1-\Delta)^{\frac{\gamma}{4}}\left(\theta^{\epsilon}(t)\right)^{2}\right\|_{L^{2}}\leq C \|\theta^{\epsilon}(t)\|_{L^{\infty}}\left\|(1-\Delta)^{\frac{\gamma}{4}}\theta^{\epsilon}(t)\right\|_{L^{2}}. $

    By this bound and 32, we have

    $ (1Δ)γ4(θϵ(t))2L2Cθ0L(θϵ(t)L2+Λγ2θϵ(t)L2). $ (34)

    We now consider $ \left\|(1-\Delta)^{-\frac{\gamma}{4}}\Lambda (\partial_{xx})^{-\alpha}\theta^{\epsilon}(t)\right\|_{L^{2}} $. For $ |\xi|\leq 1 $,

    For $ |\xi|\ge 1 $,

    So,

    $ (1Δ)γ4Λ(xx)αθϵ(t)L2C(θϵ(t)L2+Λ12(xx)α2θϵ(t)L2). $ (35)

    By 34 and 35, we obtain

    $ ddtθϵ(t)2L2+Λγ2θϵ(t)2L2+ϵθϵx2L2Cθ0L(θϵ(t)L2+Λγ2θϵ(t)L2)(θϵ(t)L2+Λ12(xx)α2θϵ(t)L2)C(θ0L+θ02L)θϵ(t)2L2+C(1+θ02L)Λ12(xx)α2θϵ(t)2L2+12Λγ2θϵ(t)2L2 $ (36)

    and so

    $ ddtθϵ(t)2L2+Λγ2θϵ(t)2L2+ϵθϵx2L2C(θ0L+θ02L)θϵ(t)2L2+C(1+θ02L)Λ12(xx)α2θϵ(t)2L2. $

    By Gronwall's inequality,

    $ \left\|\theta^{\epsilon}(t)\right\|^{2}_{L^{2}}\leq \left(\|\theta_{0}\|^{2}_{L^{2}} +C(1+\|\theta_{0}\|^{2}_{L^{\infty}})\|\theta_{0}\|_{L^{1}}\right)e^{C\left(\|\theta_{0}\|_{L^{\infty}}+\|\theta_{0}\|^{2}_{L^{\infty}}\right)t}, $

    where we use 33 to bound the time integral of $ \left\|\Lambda^{\frac{1}{2}}(\partial_{xx})^{-\frac{\alpha}{2}}\theta^{\epsilon}(t)\right\|^{2}_{L^{2}} $. Hence we finally derive the following

    $ θϵ(t)2L2+t0Λγ2θϵ(s)2L2ds+ϵt0θϵx(s)2L2dsθ02L2+(θ02L2+C(1+θ02L)θ0L1)eC(θ0L+θ02L)tC(θ0L+θ02L). $ (37)

    Therefore, $ (\theta^{\epsilon}) $ is bounded in $ \mathcal{B}_{T} $ uniformly in $ \epsilon>0 $.

    By 32 and 33,

    $ θϵLT(L1L). $ (38)

    We next consider $ \mathcal{H}(\partial_{xx})^{-\alpha}\theta^{\epsilon} $. We first choose $ \beta \in \left[0, \frac{1}{2}\right) $ also satisfying

    $ 2α12<β2α+γ2. $ (39)

    Then,

    $ R|ξ|2(β2α)|^θϵ(ξ)|2dξ=|ξ|1|ξ|2(β2α)|^θϵ(ξ)|2dξ+|ξ|1|ξ|2(β2α)|^θϵ(ξ)|2dξ^θϵ2L+Λγ2θϵ2L2θϵ2L1+Λγ2θϵ2L2 $

    and so

    $ \mathcal{H}(\partial_{xx})^{-\alpha}\theta^{\epsilon}\in L^{2}_{T}\dot{H}^{\beta}. $

    Moreover, by Sobolev embedding,

    $ H(xx)αθϵL2TLp,1p=12β $ (40)

    where $ \beta $ is defined in 39. By 37, we also have

    $ \Lambda^{\gamma}\theta^{\epsilon}+\epsilon \theta^{\epsilon}_{xx}\in L^{2}_{T}H^{-2}. $

    Combining all together, we derive that

    $ \theta^{\epsilon}_{t}\in L^{1}_{T}H^{-2}. $

    Finally, 7 and 38 imply that

    $ Λγ2(θϵψ)L2TL2. $ (41)

    To pass the limit to this formulation, we extract a subsequence of $ \left(\theta^{\epsilon}\right) $, using the same index $ \epsilon $ for simplicity, and a function $ \theta \in \mathcal{B}_T $ such that

    $ θϵθinL2THγ2, $ (42a)
    $ θϵθin L2TLploc for all 1<p<21γ, $ (42b)
    $ θϵθin L2TH1γ22α,. $ (42c)

    Here, we use Lemma 2.2 with

    $ X_0 = L^2_{T}H^{\frac{\gamma}{2}}, \quad X_{1} = L^2_{T}L^p_{\text{loc}}, \quad X_2 = L^1_{T}H^{-2} $

    to obtain 42b. Similarly, we use Lemma 2.2 with the condition 30 and

    $ X_0 = L^2_{T}H^{\frac{\gamma}{2}}, \quad X_{1} = L^{2}_{T}H^{1-\frac{\gamma}{2}-2\alpha}, \quad X_2 = L^1_{T}H^{-2} $

    to obtain 42c.

    We now multiply 31 by a test function $ \psi\in \mathcal{C}^{\infty}_{c}\left([0, T)\times\mathbb{R}\right) $ and integrate over $ \mathbb{R} $. Then,

    $ T0[θϵψt(H(xx)αθϵ)θϵψxI+Λγθϵψ+ϵθϵψxx]dxdtθϵ0(x)ψ(0,x)dx=T0Λ1γ2H(xx)αθϵΛγ2(θϵψ)IIdxdt. $ (43)

    By 40 and the strong convergence in 42b, we can pass to the limit to $ \text{I} $. By 41 and the strong convergence in 42c, we can also pass to the limit to $ \text{II} $. Therefore, we obtain

    $ T0R[θψt(H(xx)αθ)θψxΛ1γ2(xx)αθΛγ2(θψ)θΛγψ]dxdt=Rθ0(x)ψ(x,0)dx. $

    This completes the proof of Theorem 4.2.

    Remark 1. Theorem 4.2 improves Theorem 1.4 in [3], where $ (\alpha, \gamma) $ is assumed to satisfy $ \alpha\ge \frac{1}{2}-\frac{\gamma}{4} $. The main idea of taking weaker regularization in 28a-28b is that the Hilbert transform in front of $ (1-\partial_{xx})^{-\alpha} $ gives 33 which makes to obtain 37. We choose $ \alpha> \frac{1}{2}-\frac{\gamma}{2} $ instead of $ \alpha\ge \frac{1}{2}-\frac{\gamma}{2} $ to apply compactness argument when we pass to the limit to $ \epsilon $-regularized equations.

    In this section, we consider the following equation

    $ θt(H(xx)βθ)θx+Λγθ=0, $ (44a)
    $ θ(0,x)=θ0(x) $ (44b)

    where $ \beta, \gamma>0 $. Depending on the range of $ \beta $ and $ \gamma $, we will have four different results.

    We begin with the local well-posedness result.

    Theorem 5.1. Let $ 0<\gamma<2 $ and $ 0<\beta\leq \frac{\gamma}{4} $. For $ \theta_0 \in H^2 (\Bbb R) $ there exists $ T = T(\|\theta_0\|_{H^2}) $ such that a unique solution of 44a-44b exists in $ C_{T}H^2 $. Moreover, we have the following blow-up criterion:

    $ limsuptTθ(t)H2=  if and only if  T0(ux(s)L+θx(s)2L)ds=, $ (45)

    where $u = -\mathcal{H}(\partial_{xx})^{\beta}\theta$.

    Proof. Operating $\partial^{l}$ on 44a, taking its $L^2$ inner product with $\partial^l \theta$, and summing over $l = 0, 1, 2$,

    $ 12ddtθ(t)2H2+Λγ2θ2H2=2l=0l(uθx)lθdx=2l=0(l(uθx)ulθx)lθdx2l=0ulθxlθdx=I1+I2. $ (46)

    Using the commutator estimate 9, we have

    $ I12l=0l(uθx)ulθxL2θH2C(uxLθH2+uH2θxL)θH2Cκ(uxL+θx2L)θ2H2+κu2H2. $ (47)

    And by integration by parts,

    $ I2=122l=0ux|lθ|2dx=122l=0ux|lθ|2dxCuxLθ2H2. $ (48)

    Since $ \beta\leq \frac{\gamma}{4} $, for a sufficiently small $ \kappa>0 $

    $ \kappa\|u\|^{2}_{H^2}\leq \frac{1}{2}\left\|\Lambda^{\frac{\gamma}{2}}\theta\right\|_{H^2}^2. $

    By 47 and 48, we obtain

    $ ddtθ2H2+Λγ2θ2H2C(uxL+θx2L)θ2H2Cθ3H2+Cθ4H2, βγ4 $ (49)

    from which we deduce that there is $ T = T(\|\theta_{0}\|_{H^{2}}) $ such that

    $ \|\theta(t)\|_{H^2}\leq 2 \|\theta_0\|_{H^2} \quad \text{for}\ \ \text{all}~ \ t < T. $

    49 also implies 45.

    To show the uniqueness, let $ \theta_{1} $ and $ \theta_{2} $ be two solutions of 44a-44b, and let $ \theta = \theta_{1}-\theta_{2} $ and $ u = u_{1}-u_{2} = -\mathcal{H}(\partial_{xx})^{\beta}\theta_{1} + -\mathcal{H}(\partial_{xx})^{\beta}\theta_{2} $. Then, $ (\theta, u) $ satisfies the following equations

    $ \theta_t+u_{1}\theta_x-u\theta_{2x} = -\Lambda^{\gamma}\theta, \quad u = -\mathcal{H}(\partial_{xx} )^{\beta} \theta, \quad \theta(0, x) = 0. $

    By taking the $ L^{2} $ product of the equation with $ \theta $,

    $ ddtθ2L2+2Λγ2θ2L2C(u1xL+θ2x2L)θ2L2C(Λγ2θ1H2+θ22H2)θ2L2. $

    So, $ \theta = 0 $ in $ L^{2} $ and thus a solution is unique. This completes the proof of Theorem 5.1.

    Theorem 5.1 provides a local existence result for $ \beta \nearrow \frac{1}{2} $ as $ \gamma\nearrow 2 $. But, we can increase the range of $ \beta $ when we deal with 44a-44b directly with $ \gamma = 2 $ because we can do the integration by parts.

    Theorem 5.2. Let $ \gamma = 2 $ and $ 0<\beta<1 $. For $ \theta_0 \in H^2 (\Bbb R) $ there exists $ T = T(\|\theta_0\|_{H^2}) $ such that a unique solution of 44a-44b exists in $ C_{T}H^2 $.

    Proof. We begin the $ L^{2} $ bound:

    $ \frac{1}{2}\frac{d}{dt}\left\|\theta\right\|^{2}_{L^{2}}+ \left\|\theta_{x}\right\|^{2}_{L^{2}}\leq \|\theta\|_{L^{\infty}} \left\|\mathcal{H}(\partial_{xx})^{\beta}\theta\right\|_{L^{2}} \left\|\theta_{x}\right\|_{L^{2}}\leq C\|\theta\|^{3}_{H^{2}}. $

    We next estimate $ \theta_{xx} $. Indeed, after several integration parts, we have

    $ 12ddtθ2˙H2+θ2˙H3={H(xx)βθx}θxθxxxdx+12{H(xx)βθx}θxxθxxdx=I1+I2. $

    When $ 0<\beta<1 $,

    $ |I1|θxLH(xx)βθxL2θxxxL2=θxLΛ2β+1θL2θxxxL2CθH2θx1βL2θxxx1+βL2Cθ4H2+Cθ42β1βH2+14θxxx2L2. $

    And

    $ |I2|H(xx)βθxL2θxx2L4CH(xx)βθxL2θxx32L2θxxx12L2Cθ4H2+14θxxx2L2. $

    Therefore, we obtain

    $ ddtθ2H2+θx2H2Cθ4H2+Cθ42β1βH2. $ (50)

    This implies that there exists $ T = T(\|\theta_0\|_{H^2}) $ such that there exists a unique solution of 44a-44b in $ C_{T}H^2 $.

    We may lower the regularity of the initial data to prove a local existence result of a weak solution by considering initial data in $ \dot H^{\frac{1}{2}} $. The main tools to achieve this will be the use of the Hardy-BMO duality together with interpolation arguments. However, in order to simplify the computation, we consider an equivalent equation by changing the sign of the nonlinearity:

    $ θt+(H(xx)βθ)θx+Λγθ=0, $ (51a)
    $ θ(0,x)=θ0(x). $ (51b)

    This can be obtained from 51a-51b via $ \theta \mapsto -\theta $. For this equation, we do $ \dot H^{\frac{1}{2}} $ estimates and prove that there exists a local existence of a unique solution in that special case.

    Theorem 5.3. Let $ \gamma = 2 $ and $ 0<\beta<\frac{1}{2} $. For any $ \theta_0 \in \dot H^{\frac{1}{2}} (\Bbb R) $, there exists $ T = T(\Vert \theta_{0} \Vert_{\dot H^{\frac{1}{2}}}) $ such that there exists a unique local-in-time solution in $ C_{T}\dot H^{\frac{1}{2}} \cap L^2_{T}\dot{H}^{\frac{3}{2}} $.

    Proof. By recalling that $ \Lambda^{2\beta} = (-\partial_{xx})^{\beta} $ we get

    $ 12ddtθ2˙H12+Λ1+γ2θ2L2=Λ12θΛ12{(H(xx)βθ)θx}dx=θxΛθ H(xx)βθdx=θxHθxH(xx)βθdx. $

    We now use the $ \mathcal{H}^1 $-BMO duality to estimate the right hand side of the last equality. By using the estimate 5 and $ \dot H^{\frac{1}{2}} \hookrightarrow BMO $, we obtain

    $ \Vert \theta_{x} \mathcal{H} \theta_{x} \Vert_{\mathcal{H}^1} \leq \Vert \theta \Vert^{2}_{\dot H^{1}}, \quad \left\|\mathcal{H}(-\partial_{xx})^{\beta}\theta\right\|_{{BMO}}\leq C\Vert \theta \Vert_{\dot H^{2\beta +\frac{1}{2}}} $

    and thus we have

    $ \frac{1}{2} \frac{d}{dt} \Vert \theta \Vert^{2}_{\dot H^{\frac{1}{2}}} +\left\Vert \Lambda^{\frac{1+\gamma}{2}} \theta \right\Vert^{2}_{L^{2}} \leq C\Vert \theta \Vert^{2}_{\dot H^{1}} \Vert \theta \Vert_{\dot H^{2\beta +\frac{1}{2}}}. $

    By fixing $ \gamma = 2 $ and by using the interpolation inequalities

    $ \Vert \theta \Vert^{2}_{\dot H^1} \leq \Vert \theta \Vert_{\dot H^{\frac{3}{2}}}\Vert \theta \Vert_{\dot H^{\frac{1}{2}}}, \quad \Vert \theta \Vert_{\dot H^{2\beta +\frac{1}{2}}} \leq \Vert \theta \Vert^{2\beta}_{\dot H^{\frac{3}{2}}} \Vert \theta \Vert^{1-2\beta}_{\dot H^{\frac{1}{2}}}, $

    where we use $ \frac{1}{2}\leq 2\beta+\frac{1}{2} \leq \frac{3}{2} $ for $ \beta \in \left(0, \frac{1}{2}\right) $ to get the second inequality. Hence, we obtain

    $ 12ddtθ2˙H12+Λ32θ2L2θ2˙H1θ˙H2β+12θ1+2β˙H32θ22β˙H1212θ2˙H32+2θ41β12β˙H12, $

    where we use the condition $ \beta \in \left(0, \frac{1}{2}\right) $ again to derive the inequality. This implies local existence of a unique solution up to some time $ T = T(\Vert \theta_{0} \Vert_{\dot H^{\frac{1}{2}}}) $.

    Remark 2. In the case $ \beta = 1/2 $, the equation reduces to the following Hamilton-Jacobi equation (or primitive Burgers equation)

    $ \theta_{t} - \theta_{x}^2 +\Lambda^{\gamma}\theta = 0. $

    For this equation, it seems that a naive approach based in energy methods cannot work. Indeed, if we multiply by $ \Lambda\theta $ and integrate by parts the nonlinearity takes a commutator structure

    $ \int \theta_x^2\Lambda \theta dx = \int \theta_x^2\mathcal{H} \theta_xdx = -\frac{1}{2}\int \theta_x [\mathcal{H}, \theta_x]\theta_x \ dx . $

    However, it seems that, at this level of regularity, this commutator is comparable to an energy estimate:

    $ \int \theta_x^2\Lambda \theta dx\leq c\|\theta_x\|_{L^3}^3\leq c\|\theta\|_{H^{1+\frac{1}{6}}}^3\leq c\|\theta\|_{H^{1}}^2\|\theta\|_{H^{\frac{3}{2}}} $

    which is also equivalent to the use of Hardy-BMO duality:

    $ \int \theta_x^2\mathcal{H} \theta_xdx \leq \Vert \theta_{x} \Vert_{BMO} \Vert \theta_x \mathcal{H} \theta_x \Vert_{\mathcal{H}^{1}} \leq \|\theta\|_{H^{1}}^2\|\theta\|_{H^{\frac{3}{2}}}. $

    Also, the best estimate that one has for the commutator $ [\mathcal{H}, \theta_{x}] \theta_{x} $ in $ L^{2} $ is that it is controlled by $ \Vert\theta_{x}\Vert_{BMO}\Vert\theta\Vert_{\dot H^{1}} $ (see e.g. [22]) which is, once again, similar to the use of the Hardy-BMO duality. So the commutator structure is not that useful in this special case.

    Remark 3. It is also unclear whether the local solution starting from an arbitrary initial data becomes smooth. However, for smooth initial data satisfying size restriction in appropriate spaces, one can prove the desired smoothing effect.

    We finally deal with 51a-51b with $ \gamma = 2 $.

    Theorem 5.4. Let $ \gamma = 2 $ and $ \beta<\frac{1}{4} $. For any $ \theta_0 \in H^2 (\Bbb R) $, there exists a unique global-in-time solution in $ C_{T}H^2 $.

    Proof. By Theorem 5.1, we only need to control the quantities in 45. Let $ u = -\mathcal{H}(\partial_{xx})^{\beta}\theta $. We first note that 51a-51b satisfies the maximum principle and so

    $ \left\|\theta(t)\right\|_{L^{\infty}}\leq \left\|\theta_{0}\right\|_{L^{\infty}}\leq C \|\theta_{0}\|_{H^{2}}. $

    We take the $ L^2 $ inner product of 51a with $ \theta $. Then,

    $ 12ddtθ2L2+θx2L2=uθxθdxθ0LuL2θxL2. $ (52)

    Since

    $ \|u\|_{L^{2}} \leq C\|\theta\|^{1-2\beta}_{L^{2}} \|\theta_{x}\|^{2\beta}_{L^{2}} \quad \text{for $\beta < \frac{1}{2}$}, $

    we have

    $ θ(t)2L2+t0θx(s)2L2dsC(t,θ0H2). $ (53)

    We next take $ \partial_x $ to 51a, take its $ L^2 $ inner product with $ \theta_x $, and integrate by parts to obtain

    $ 12ddtθx2L2+θxx2L2=uθxθxxdx2u2Lθx2L2+12θxx2L2. $

    Since

    $ \|u\|^{2}_{L^{\infty}} \leq C\|\theta\|^{2}_{L^{2}}+C\|\theta_{x}\|^{2}_{L^{2}} \quad \text{when $\beta < \frac{1}{4}$}, $

    we obtain

    $ θx(t)2L2+t0θxx(s)2L2dsC(t,θ0L1,θ0H2)when β<14. $ (54)

    We also obtain

    $ θx2LC(θx2L2+θxx2L2),uxLC(θxL2+θxxL2)when β<14 $ (55)

    By 53, 54 and 55, we finally obtain

    $ t0(θx(s)2L+ux(s)L)dsCt0(θx(s)2L2+θxx(s)2L2+θx(s)L2+θxx(s)L2)dsC(t,θ0L1,θ0H2) $

    and so we complete the proof of Theorem 5.4.

    The authors acknowledge the referees for their valuable comments and suggestions that highly improved the manuscript.

    H.B. was supported by NRF-2018R1D1A1B07049015.

    R.G.B. was supported by the LABEX MILYON (ANR-10-LABX-0070) of Université de Lyon, within the program "Investissements d'Avenir" (ANR-11-IDEX-0007) operated by the French National Research Agency (ANR), and by the Universidad de Cantabria.

    O.L. was partially supported by the Marie-Curie Grant, acronym: TRANSIC, from the FP7-IEF program and by the ERC through the Starting Grant project H2020-EU.1.1.-63922.

    Both O. L. and R.G.B. were partially supported by the Grant MTM2014-59488-P from the former Ministerio de Economía y Competitividad (MINECO, Spain).



    [1] On the well-posedness of various one-dimensional model equations for fluid motion. Nonlinear Anal. (2017) 160: 25-43.
    [2] Global existence for some transport equations with nonlocal velocity. Adv. Math. (2015) 269: 197-219.
    [3]

    H. Bae, R. Granero-Belinchón and O. Lazar, Global existence of weak solutions to dissipative transport equations with nonlocal velocity, Nonlinearity, 31 (2018) 1484–1515.

    [4]

    H. Bahouri, J-Y. Chemin and R. Danchin, Fourier Analysis and Nonlinear Partial Differential Equations, Grundlehren der Mathematischen Wissenschaften, 343, Springer, 2011.

    [5] Analytic structure of 1D transport equations with nonlocal fluxes. Physica D. (1996) 91: 349-375.
    [6] A mass-transportation approach to a one dimensional fluid mechanics model with nonlocal velocity. Adv. Math. (2012) 231: 306-327.
    [7] Global existence, singularities and ill-posedness for a nonlocal flux. Adv. Math. (2008) 219: 1916-1936.
    [8] Self-similar solutions for a transport equation with non-local flux. Chinese Annals of Mathematics, Series B (2009) 30: 505-512.
    [9] Finite time singularities in a 1D model of the quasi-geostrophic equation. Adv. Math. (2005) 194: 203-223.
    [10] A maximum principle applied to quasi-geostrophic equations. Comm. Math. Phys. (2004) 249: 511-528.
    [11] Formation of singularities for a transport equation with nonlocal velocity. Ann. of Math. (2005) 162: 1-13.
    [12] A unified theory of Hilbert transforms and ergodic theorems. Rev. Mat. Cuyana (1955) 1: 105-167.
    [13] On a one-dimensional model for the 3D vorticity equation. J. Statist. Phys. (1990) 59: 1251-1263.
    [14]

    H. Dong, Well-posedness for a transport equation with nonlocal velocity, J. Funct. Anal., 255, (2008), 3070–3097.

    [15]

    H. Dong, On a multi-dimensional transport equation with nonlocal velocity, Adv. Math., 264 (2014), 747–761.

    [16]

    H. Dong and D. Li, On a one-dimensional $\alpha$-patch model with nonlocal drift and fractional dissipation, Trans. Amer. Math. Soc., 366 (2014), 2041–2061.

    [17]

    J. Duoandikoetxea, Fourier Analysis, Translated and revised from the 1995 Spanish original by David Cruz-Uribe, Graduate Studies in Mathematics, 29, American Mathematical Society, 2000.

    [18] Commutator estimates and the Euler and Navier-Stokes equations. Comm. Pure Appl. Math. (1988) 41: 891-907.
    [19] Regularity and blow up for active scalars. Math. Model. Math. Phenom. (2010) 5: 225-255.
    [20] On a 1D nonlocal transport equation with nonlocal velocity and subcritical or supercritical diffusion. Journal of Diff. Eq. (2016) 261: 4974-4996.
    [21]

    O. Lazar and P.-G. Lemarié-Rieusset, Infinite energy solutions for a 1D transport equation with nonlocal velocity, Dynamics of PDEs, 13 (2016), 107-131.

    [22]

    D. Li, On Kato-Ponce and fractional Leibniz, arXiv: 1609.01780.

    [23] Blow-up of solutions for a 1D transport equation with nonlocal velocity and supercritical dissipation. Adv. Math. (2008) 217: 2563-2568.
    [24] On a one-dimensional nonlocal flux with fractional dissipation. SIAM J. Math. Anal. (2011) 43: 507-526.
    [25] Further properties of a continuum of model equations with globally defined flux. J. Math. Anal. Appl. (1998) 221: 132-160.
    [26] Compact sets in the space $L^{p}(0, T; B)$. Ann. Mat. Pura Appl. (1987) 146: 65-96.
  • This article has been cited by:

    1. Lucas C.F. Ferreira, Valter V.C. Moitinho, Global well-posedness, regularity and blow-up for the β-CCF model, 2023, 344, 00220396, 230, 10.1016/j.jde.2022.10.032
    2. Lucas Ferreira, Valter Moitinho, Global smoothness for a 1D supercritical transport model with nonlocal velocity, 2020, 148, 0002-9939, 2981, 10.1090/proc/14984
    3. Martina Magliocca, On a fourth order equation describing single-component film models, 2024, 79, 14681218, 104137, 10.1016/j.nonrwa.2024.104137
  • Reader Comments
  • © 2019 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(4008) PDF downloads(280) Cited by(3)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog