This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Nonlocal Filtration Equations with Rough Kernels in the Heisenberg Group

Rong Tang
Abstract

Motivated by the extensive investigations of integro-differential equations on n\mathbb{R}^{n}, we consider nonlocal filtration type equations with rough kernels on the Heisenberg group n\mathbb{H}^{n}. We prove the existence and uniqueness of weak solutions corresponding to suitable initial data. Furthermore, we obtain the large time behavior of solutions and the uniform Hölder regularity of sign-changing solutions for the porous medium type equations (m1m\geq 1). Notice that both conformal fractional operators α/2\mathscr{L}_{\alpha/2} and pure power fractional operators α/2\mathscr{L}^{\alpha/2} on the Heisenberg group n\mathbb{H}^{n} have their integral representations with suitable kernels. Therefore, all the results in this paper will hold for these equations with operators α/2\mathscr{L}_{\alpha/2} or α/2\mathscr{L}^{\alpha/2}.

1 Introduction

The classic filtration equations are the nonlinear parabolic equations

tu=Δn(um),m>0.\displaystyle\partial_{t}u=\Delta_{\mathbb{R}^{n}}(u^{m}),\leavevmode\nobreak\ \leavevmode\nobreak\ m>0.

Here, um=|u|m1uu^{m}=|u|^{m-1}u. The above equations with m>1m>1 are usually called the porous medium equations. Otherwise, they are called the fast diffusion equations. There are extensive investigations of the above equations, especially in the applications of physics, see [28].

In 2012, Pablo, Quirós, Rodríguez and Vázquez\autocitesafp_dpaagf_dpa developed a theory of the following fractional version of filtration type equation, which can be considered as a model in statistical mechanics.

{tu+(Δ)nα/2(|u|m1u)=0,xn,t>0u(x,0)=u0(x),xn.\left\{\begin{array}[]{ll}\partial_{t}u+(-\Delta)_{\mathbb{R}^{n}}^{\alpha/2}(|u|^{m-1}u)=0,&x\in\mathbb{R}^{n},t>0\\ u(x,0)=u_{0}(x),&x\in\mathbb{R}^{n}.\end{array}\right.

Here 0<α<20<\alpha<2. The authors developed a theory of existence, uniqueness and qualitative properties of the weak solutions through the Caffarelli-Silvestre extension technique [6]. Moreover, this nonlocal equation forces the infinite propagation speed. This result is different comparing to the finite propagation speed with free boundary in the local filtration equations. This difference results from the nonlocal virtue of the fractional Laplacian operator (Δ)nα/2(-\Delta)_{\mathbb{R}^{n}}^{\alpha/2}.

Considering the Laplace-Beltrami operator (ΔM)(-\Delta_{M}) on the general Riemannian manifolds MM, the porous medium type equations have been studied for instance in [18]. Furthermore, the pure power fractional Laplacian (ΔM)α/2(-\Delta_{M})^{\alpha/2} can be defined through a semigroup approach. The corresponding fractional porous medium equations are investigated, see [3].

However, the pure power fractional operator (ΔM)α/2(-\Delta_{M})^{\alpha/2} is not conformally covariant. Let (X,h+)(X,h^{+}) be a Poincaré-Einstein manifold with conformal infinity (M,[g])(M,[g]). The conformal fractional Laplacian Pα/2gP_{\alpha/2}^{g} on (M,g)(M,g) is constructed as the Dirichlet-to-Neumann operator for a generalized eigenvalue problem on (X,h+)(X,h^{+}) (see the details in \autocitesfli_cs).

Pα/2gP_{\alpha/2}^{g} is a self-adjoint pseudo-differential operator on (M,g)(M,g) with principal symbol the same as (ΔM)α/2(-\Delta_{M})^{\alpha/2}, and satisfies the conformal transformation relation,

Pα/2ρ4nαg(ϕ)=ρn+αnαPα/2g(ρϕ),ρ,ϕC(M)withρ>0.\displaystyle P_{\alpha/2}^{\rho^{\frac{4}{n-\alpha}}g}(\phi)=\rho^{-\frac{n+\alpha}{n-\alpha}}P_{\alpha/2}^{g}(\rho\phi),\leavevmode\nobreak\ \leavevmode\nobreak\ \forall\leavevmode\nobreak\ \rho,\phi\in C^{\infty}(M)\leavevmode\nobreak\ \text{with}\leavevmode\nobreak\ \rho>0.

In the case that M=nM=\mathbb{R}^{n} with the Euclidean metric |dx|2|dx|^{2}, Chang and González [7] illustrated that the conformal fractional Laplacian Pα/2|dx|n2P_{\alpha/2}^{|dx|^{2}_{\mathbb{R}^{n}}} coincides with the pure power fractional Laplacian (Δn)α/2(-\Delta_{\mathbb{R}^{n}})^{\alpha/2} for 0<α<20<\alpha<2.

Let (𝕊n,g𝕊n)(\mathbb{S}^{n},g_{\mathbb{S}^{n}}) with n2n\geq 2 be the unit sphere equipped with the standard round metric g𝕊ng_{\mathbb{S}^{n}}, the conformal fractional Laplacian Pα/2g𝕊nP_{\alpha/2}^{g_{\mathbb{S}^{n}}} is the pull back of the fractional Laplacian (Δn)α/2(-\Delta_{\mathbb{R}^{n}})^{\alpha/2} via the stereographic projection through

(Pα/2g𝕊n(ϕ))Ψ=(detdΨ)n+α2n(Δn)α/2((detdΨ)nα2nϕΨ),forϕC2(𝕊n).\displaystyle(P_{\alpha/2}^{g_{\mathbb{S}^{n}}}(\phi))\circ\Psi=(\det\mathop{}\!\mathrm{d}\Psi)^{-\frac{n+\alpha}{2n}}(-\Delta_{\mathbb{R}^{n}})^{\alpha/2}((\det\mathop{}\!\mathrm{d}\Psi)^{\frac{n-\alpha}{2n}}\phi\circ\Psi),\leavevmode\nobreak\ \leavevmode\nobreak\ \text{for}\leavevmode\nobreak\ \phi\in C^{2}(\mathbb{S}^{n}).

where Ψ\Psi is the inverse of the stereographic projection and Ψ(g𝕊n)=(detdΨ)2n|dx|2\Psi^{*}(g_{\mathbb{S}^{n}})=(\det\mathop{}\!\mathrm{d}\Psi)^{\frac{2}{n}}|\mathop{}\!\mathrm{d}x|^{2}.

Define Rα/2g=Pα/2g(1)R_{\alpha/2}^{g}=P_{\alpha/2}^{g}(1), which is called the fractional order curvature on 𝕊n\mathbb{S}^{n}. Considering the following fractional Nirenberg problem using a geometric flow. Given a positive smooth function ff on 𝕊n\mathbb{S}^{n} and g0[g𝕊n]g_{0}\in[g_{\mathbb{S}^{n}}], we define the flow g=g(t)g=g(t):

gt=(γfRα/2g)g,g|t=0=g0,\displaystyle\frac{\partial g}{\partial t}=(\gamma f-R_{\alpha/2}^{g})g,\leavevmode\nobreak\ g|_{t=0}=g_{0},

where γ(t)=𝕊nRα/2gdVg𝕊nfdVg\gamma(t)=\frac{\int_{\mathbb{S}^{n}}R_{\alpha/2}^{g}\mathop{}\!\mathrm{d}V_{g}}{\int_{\mathbb{S}^{n}}f\mathop{}\!\mathrm{d}V_{g}} and 0<α<20<\alpha<2. The parabolic equation of this flow can be rewritten as the fractional filtration equation on n\mathbb{R}^{n} by rescalings through the stereographic projection. This approach has been well used in the investigations of fractional Nirenberg problem and fractional Yamabe problem on the unit sphere 𝕊n\mathbb{S}^{n} in \autocitesafc_cxfys_hpafy_jt.

In a similar way, we will consider the fractional geometric flow of the Cauchy Riemannian (CR) structure instead of the Riemannian structure. The investigations of CR geometry and sub-Riemannian geometry originated from the failure of the Riemann mapping theorem in complex analysis with several variables. The flat space in the CR geometry or sub-Riemannian geometry is the Heisenberg group n\mathbb{H}^{n}.

Let -\mathscr{L} be the sub-Laplacian on the Heisenberg group. The fractional power of the operator \mathscr{L} can be defined through semigroup approach, denoted by α/2\mathscr{L}^{\alpha/2}. An extension technique involving α/2\mathscr{L}^{\alpha/2} akin to that of Caffarelli-Silvestre has been developed in a general setting by Stinga and Torrea in [26]. This naturally leads to a theory of existence, uniqueness and quantitative properties of the fractional filtration equations on n\mathbb{H}^{n} via the Caffarelli-Silvestre type extension in a similar way.

However, the pure power fractional operator α/2\mathscr{L}^{\alpha/2} is not conformally covariant with respect to conformal class of pseudo-Hermitian structures, which seems to be very counterintuitive since (Δn)α/2(-\Delta_{\mathbb{R}^{n}})^{\alpha/2} are conformally covariant. There is another pseudodifferential operator α/2\mathscr{L}_{\alpha/2} satisfying a conformally covariant formula. This is where a new story begins.

The conformally covariant operator α/2\mathscr{L}_{\alpha/2} was introduced by Branson, Fontana and Morpurgo in [4] via the spectral formula,

α/2=2α2|s|α2Γ(12|s|1+1+α22)Γ(12|s|1+1α22), 0<α<2.\displaystyle\mathscr{L}_{\alpha/2}=2^{\frac{\alpha}{2}}|\partial_{s}|^{\frac{\alpha}{2}}\frac{\Gamma(-\frac{1}{2}\mathscr{L}|\partial_{s}|^{-1}+\frac{1+\frac{\alpha}{2}}{2})}{\Gamma(-\frac{1}{2}\mathscr{L}|\partial_{s}|^{-1}+\frac{1-\frac{\alpha}{2}}{2})},\leavevmode\nobreak\ \leavevmode\nobreak\ 0<\alpha<2.

In [17], Frank, González, Monticelli and Tan formulated the following extension problem,

{ttU+(1α)t1tU+t2ssU+U=0,inn×+,U(x,0)=f(x),x=(z,s)n.\left\{\begin{array}[]{ll}\partial_{tt}U+(1-\alpha)t^{-1}\partial_{t}U+t^{2}\partial_{ss}U+\mathscr{L}U=0,&\text{in}\leavevmode\nobreak\ \mathbb{H}^{n}\times\mathbb{R}^{+},\\ U(x,0)=f(x),&x=(z,s)\in\mathbb{H}^{n}.\end{array}\right.

The authors recovered the operator α/2f\mathscr{L}_{\alpha/2}f as the Neumann data of the above extension function UU. The additional term t2sst^{2}\partial_{ss}, which is a fourth-order term with respect to the dilations δλ\delta_{\lambda} on n\mathbb{H}^{n}, makes the extension problem harder than the Caffarelli-Silvestre type extension.

Let (Xn+1,h+)(X^{n+1},h^{+}) be a Kähler-Einstein manifold with strictly pseudoconvex boundary (M,[θ])(M,[\theta]). The conformal fractional operator Pα/2θP_{\alpha/2}^{\theta} on (M,θ)(M,\theta) is constructed as the Dirichlet-to-Neumann operator for a generalized eigenvalue problem (see [17]). The fractional order curvature associated to Pα/2θP_{\alpha/2}^{\theta} is defined as Rα/2θ=Pα/2θ(1)R_{\alpha/2}^{\theta}=P^{\theta}_{\alpha/2}(1). Here, 0<α<20<\alpha<2.

Let (𝕊2n+1,θ^0)(\mathbb{S}^{2n+1},\hat{\theta}_{0}) with n1n\geq 1 be the unit sphere equipped with the standard pseudo-Hermitian structure θ^0\hat{\theta}_{0}. The correspondence between the CR sphere and the Heisenberg group arises from Cayley transform, see [11]. As pointed out in [4], Pα/2θ^0P_{\alpha/2}^{\hat{\theta}_{0}} is the pull back of the conformally covariant operator α/2\mathscr{L}_{\alpha/2} via the Cayley transformation through

(Pα/2θ^0(ϕ))ΨC=(detdΨC)2(n+1)+α4(n+1)α/2((detdΨC)2(n+1)α4(n+1)ϕΨC),forϕC2(𝕊2n+1).\displaystyle(P_{\alpha/2}^{\hat{\theta}_{0}}(\phi))\circ\Psi_{C}=(\det\mathop{}\!\mathrm{d}\Psi_{C})^{-\frac{2(n+1)+\alpha}{4(n+1)}}\mathscr{L}_{\alpha/2}((\det\mathop{}\!\mathrm{d}\Psi_{C})^{\frac{2(n+1)-\alpha}{4(n+1)}}\phi\circ\Psi_{C}),\leavevmode\nobreak\ \leavevmode\nobreak\ \text{for}\leavevmode\nobreak\ \phi\in C^{2}(\mathbb{S}^{2n+1}).

where ΨC\Psi_{C} is the Cayley transformation and ΨC(θ^0)=(|JC|)1n+1θ0\Psi_{C}^{*}(\hat{\theta}_{0})=(|J_{C}|)^{\frac{1}{n+1}}\theta_{0}.

Considering the CR fractional curvature flow on 𝕊2n+1\mathbb{S}^{2n+1}, it is defined as the evolution of the contact form θ^(t)\hat{\theta}(t):

θ^t=(Qα/2θ^γf)θ^,θ^(0)=v042(n+1)αθ^0.\displaystyle\frac{\partial\hat{\theta}}{\partial t}=-(Q_{\alpha/2}^{\hat{\theta}}-\gamma f)\hat{\theta},\leavevmode\nobreak\ \hat{\theta}(0)=v_{0}^{\frac{4}{2(n+1)-\alpha}}\hat{\theta}_{0}.

where γ(t)=𝕊2n+1Qα/2θ^dVθ^𝕊2n+1fdVθ^\gamma(t)=\frac{\int_{\mathbb{S}^{2n+1}}{Q_{\alpha/2}^{\hat{\theta}}\mathop{}\!\mathrm{d}V_{\hat{\theta}}}}{\int_{\mathbb{S}^{2n+1}}f\mathop{}\!\mathrm{d}V_{\hat{\theta}}} and 0<α<20<\alpha<2.

The Cayley transform is a natural way to transform the evolution equation of this geometric flow on the CR sphere 𝕊2n+1\mathbb{S}^{2n+1} to the nonlocal equation of filtration type involving α/2\mathscr{L}_{\alpha/2} on the Heisenberg group n\mathbb{H}^{n}. However, the extension map corresponding to α/2\mathscr{L}_{\alpha/2} is not good enough as the Caffarelli-Silvestre type extension, we can not obtain analogous result of the nonlocal filtration type equation on n\mathbb{H}^{n} through the extension technique as [24].

On the other hand, Roncal and Thangavelu [25] established the essential pointwise integral representation of the operator α/2\mathscr{L}_{\alpha/2} using tools from noncommutative harmonic analysis. To be precise, for Schwartz function ff,

α/2f(x)=Cn,αP.V.nf(x)f(y)|y1x|nQ+αdμ(y), 0<α<2.\displaystyle\mathscr{L}_{\alpha/2}f(x)=C_{n,\alpha}\text{P.V.}\int_{\mathbb{H}^{n}}\frac{f(x)-f(y)}{|y^{-1}\cdot x|_{\mathbb{H}^{n}}^{Q+\alpha}}\mathop{}\!\mathrm{d}\mu({y}),\leavevmode\nobreak\ \leavevmode\nobreak\ 0<\alpha<2.

Furthermore, as pointed out in [13], the operator α/2\mathscr{L}^{\alpha/2} also has a pointwise integro-representation. To be precise, for the Schwartz function ff,

α/2f(x)=P.V.n(f(x)f(y))R~α(y1x)dμ(y), 0<α<2.\displaystyle\mathscr{L}^{\alpha/2}f(x)=\text{P.V.}\int_{\mathbb{H}^{n}}{(f(x)-f(y))}\tilde{R}_{-\alpha}(y^{-1}\cdot x)\mathop{}\!\mathrm{d}\mu(y),\leavevmode\nobreak\ \leavevmode\nobreak\ 0<\alpha<2.

where R~α(x)\tilde{R}_{-\alpha}(x) is a positive smooth function on n\{0}\mathbb{H}^{n}\backslash\{0\} and comparable with |x|nαQ|x|_{\mathbb{H}^{n}}^{-\alpha-Q}.

Therefore, the integro-representations of α/2\mathscr{L}^{\alpha/2} and α/2\mathscr{L}_{\alpha/2} motivate the investigations of filtration type equations involving the integro-kernels with rough estimates on n\mathbb{H}^{n}.

Let us recall the past investigations of the integro-differential equations on n\mathbb{R}^{n}. Integro-differential equations naturally arise from models in physics, engineering, and finance that involve long range interactions, see for instance [27]. They are also a natural generalization of fractional differential equations, since the fractional Laplacian (Δn)α/2(-\Delta_{\mathbb{R}^{n}})^{\alpha/2} is a classic example of nonlocal operators with a specific integro-kernel,

(Δn)α/2f(x)=Cn,αP.V.nf(x)f(z)|xz|n+αdz\displaystyle(-\Delta_{\mathbb{R}^{n}})^{\alpha/2}f(x)=C_{n,\alpha}\text{P.V.}\int_{\mathbb{R}^{n}}\frac{f(x)-f(z)}{|x-z|^{n+\alpha}}\mathop{}\!\mathrm{d}z

where Cn,α=2α1αΓ((n+α)/2)πn/2Γ(1α/2)C_{n,\alpha}=\frac{2^{\alpha-1}\alpha\Gamma((n+\alpha)/2)}{\pi^{n/2}\Gamma(1-\alpha/2)}.

The existence, uniqueness and quantitative properties of the solutions to the following nonlocal filtration type equations have been investigated in \autocitesttp_airtfp_clnfe_dpartf_dpa,

{tu+um=0,xn,t>0u(x,0)=u0(x),xn.\left\{\begin{array}[]{ll}\partial_{t}u+\mathcal{L}u^{m}=0,&x\in\mathbb{R}^{n},t>0\\ u(x,0)=u_{0}(x),&x\in\mathbb{R}^{n}.\end{array}\right.

with suitable initial data u0u_{0}. There is no sign restriction of uu. The nonlocal operator \mathcal{L} is defined formally as

f(x)=P.V.n(f(x)f(y))J(x,y)dy\displaystyle\mathcal{L}f(x)=\text{P.V.}\int_{\mathbb{R}^{n}}(f(x)-f(y))J(x,y)\mathop{}\!\mathrm{d}y

with a measurable kernel JJ satisfying

{J(x,y)=J(y,x)0x,yn,xy,1Λ|xy|n+αJ(x,y)Λ|xy|Q+α,0<α<2,Λ>0.\left\{\begin{array}[]{ll}J(x,y)=J(y,x)\geq 0\nobreak\leavevmode\nobreak\leavevmode&x,y\in\mathbb{R}^{n},x\neq y,\\ \frac{1}{\Lambda|x-y|^{n+\alpha}}\leq J(x,y)\leq\frac{\Lambda}{|x-y|^{Q+\alpha}},\nobreak\leavevmode\nobreak\leavevmode&0<\alpha<2,\leavevmode\nobreak\ \leavevmode\nobreak\ \Lambda>0.\end{array}\right.

The integro-operator satisfying the required conditions has been greatly studied especially in the probability theory, for instance the Markov jump process and the martingale problem. For the corresponding filtration equations with these integro-operators, the critical value m:=nαnm^{*}\mathrel{\mathop{\mathchar 58\relax}}=\frac{n-\alpha}{n} gives a natural way to develop non-identical theories for m>mm>m^{*} and m<mm<m^{*}, see more details in [24].

In this paper, we establish the existence, uniqueness and quantitative properties of solutions to the nonlocal filtration type equations with integro-operators on the Heisenberg group n\mathbb{H}^{n}.

{tu+(|u|m1u)=0,xn,t>0u(x,0)=u0(x),xn.\left\{\begin{array}[]{ll}\partial_{t}u+\mathcal{L}(|u|^{m-1}u)=0,&x\in\mathbb{H}^{n},t>0\\ u(x,0)=u_{0}(x),&x\in\mathbb{H}^{n}.\end{array}\right. (1.1)

with suitable initial data u0u_{0}. There is no sign restriction of uu. Similarly, the nonlocal operator \mathcal{L} is defined formally as

f(x)=P.V.n(f(x)f(y))J(x,y)dμ(y)\displaystyle\mathcal{L}f(x)=\text{P.V.}\int_{\mathbb{H}^{n}}(f(x)-f(y))J(x,y)\mathop{}\!\mathrm{d}\mu(y) (1.2)

with a measurable kernel JJ satisfying

{J(x,xy)=J(x,xy1)0x,yn,xy,1Λ|x1y|Q+αJ(x,y)Λ|x1y|Q+α,0<α<2,Λ>0.\left\{\begin{array}[]{ll}J(x,x\cdot y)=J(x,x\cdot y^{-1})\geq 0\nobreak\leavevmode\nobreak\leavevmode&x,y\in\mathbb{H}^{n},x\neq y,\\ \frac{1}{\Lambda|x^{-1}\cdot y|^{Q+\alpha}}\leq J(x,y)\leq\frac{\Lambda}{|x^{-1}\cdot y|^{Q+\alpha}},\nobreak\leavevmode\nobreak\leavevmode&0<\alpha<2,\leavevmode\nobreak\ \leavevmode\nobreak\ \Lambda>0.\end{array}\right. (1.3)

Here |||\cdot| is the standard homogeneous quasi-norm on n\mathbb{H}^{n} and Q=2n+2Q=2n+2 is the corresponding homogeneous degree. dμ(y)\mathop{}\!\mathrm{d}\mu(y) is the volume form with respect to standard Haar measure on n\mathbb{H}^{n}. Under the symmetric assumption J(x,xy)=J(x,xy1)J(x,x\cdot y)=J(x,x\cdot y^{-1}), the operator has a pointwise expression,

f(x)=12n(2f(x)f(xy)f(xy1))J(x1y)dμ(y).\displaystyle\mathcal{L}f(x)=\frac{1}{2}\int_{\mathbb{H}^{n}}({2f(x)-f(x\cdot y)-f(x\cdot y^{-1})})J(x^{-1}\cdot y)\mathop{}\!\mathrm{d}\mu(y). (1.4)

for regular enough ff. Without misunderstandings, J(x,y)J(x,y) is denoted by J(x1y)J(x^{-1}\cdot y). Comparably, the critical value m:=QαQm^{*}\mathrel{\mathop{\mathchar 58\relax}}=\frac{Q-\alpha}{Q} gives a natural way to develop non-identical theories for m>mm>m^{*} and m<mm<m^{*}.

In a forthcoming paper we shall apply the results in this paper to prove the long time existence of the fractional order curvature flow on the CR sphere through Cayley transform.

1.1 Main Results

Theorem 1.

Let m>mm>m^{*}, 0<α<20<\alpha<2. JJ satisfies assumption (1.3). Then for every u0L1(n)u_{0}\in L^{1}(\mathbb{H}^{n}), there exists a unique weak solution of Cauchy problem (1.1). Moreover,

  1. (i)

    tuL((τ,):L1(n))\partial_{t}u\in L^{\infty}((\tau,\infty)\mathrel{\mathop{\mathchar 58\relax}}L^{1}({\mathbb{H}^{n}})) for every τ>0\tau>0.

  2. (ii)

    Conservation of mass: nu(x,t)dμ=nu0(x,t)dμ\int_{\mathbb{H}^{n}}u(x,t)\mathop{}\!\mathrm{d}\mu=\int_{\mathbb{H}^{n}}u_{0}(x,t)\mathop{}\!\mathrm{d}\mu.

  3. (iii)

    LpL^{p} norm of uu is non-increasing in time, for each 1p1\leq p\leq\infty.

  4. (iv)

    L1L^{1} contraction property holds:

    n(u(,t)v(,t))+n(u(,0)v(,0))+,fort0.\int_{\mathbb{H}^{n}}(u(\cdot,t)-v(\cdot,t))_{+}\leq\int_{\mathbb{H}^{n}}(u(\cdot,0)-v(\cdot,0))_{+},\leavevmode\nobreak\ \leavevmode\nobreak\ \text{for}\leavevmode\nobreak\ t\geq 0.
  5. (v)

    LpLL^{p}-L^{\infty} smoothing effects hold:

    supxn|u(x,t)|C(m,p,α,Q)tγpfpδp.\sup\limits_{x\in\mathbb{H}^{n}}|u(x,t)|\leq C(m,p,\alpha,Q)t^{-\gamma_{p}}\mathinner{\!\left\lVert f\right\rVert}_{p}^{\delta_{p}}.

    with γp=(m1+αp/Q)1\gamma_{p}=(m-1+\alpha p/Q)^{-1} and δp=αpγp/Q\delta_{p}=\alpha p\gamma_{p}/Q for each 1p1\leq p\leq\infty.

Remark 1.

Weak solutions will be defined in Section 2. Furthermore, if solution uu satisfies condition (i), it can be considered as a strong solution in some sense.

If mmm\leq m^{*}, we establish the existence of weak solutions corresponding to more restrictive initial data compared with m>mm>m^{*}.

Theorem 2.

Let 0<mm0<m\leq m^{*}, 0<α<20<\alpha<2. JJ satisfies assumption (1.3). Then for every u0L1(n)Lp(n)u_{0}\in L^{1}(\mathbb{H}^{n})\cap L^{p}(\mathbb{H}^{n}) with p>p(m)=(1m)Q/αp>p^{*}(m)=(1-m)Q/\alpha, there exists a unique weak solution of the Cauchy problem (1.1). Furthermore,

  1. (i)

    tuL((τ,):L1(n))\partial_{t}u\in L^{\infty}((\tau,\infty)\mathrel{\mathop{\mathchar 58\relax}}L^{1}({\mathbb{H}^{n}})) for every τ>0\tau>0.

  2. (ii)

    Conservation of mass: nu(x,t)dμ=nu0(x,t)dμ\int_{\mathbb{H}^{n}}u(x,t)\mathop{}\!\mathrm{d}\mu=\int_{\mathbb{H}^{n}}u_{0}(x,t)\mathop{}\!\mathrm{d}\mu holds for m=mm=m^{*}. Otherwise, there exists a finite time T>0T>0 such that u(x,T)0u(x,T)\equiv 0 for 0<m<m0<m<m^{*}.

  3. (iii)

    L1L^{1} contraction property holds; LpL^{p} norm of uu is non-increasing in time, 1p1\leq p\leq\infty.

  4. (iv)

    LpLL^{p}-L^{\infty} smoothing effects hold:

    supxn|u(x,t)|C(m,p,α,Q)tγpfpδp.\sup\limits_{x\in\mathbb{H}^{n}}|u(x,t)|\leq C(m,p,\alpha,Q)t^{-\gamma_{p}}\mathinner{\!\left\lVert f\right\rVert}_{p}^{\delta_{p}}.

    with γp=(m1+αp/Q)1\gamma_{p}=(m-1+\alpha p/Q)^{-1} and δp=αpγp/Q\delta_{p}=\alpha p\gamma_{p}/Q. Here p>p(m)p>p^{*}(m).

Remark 2.

Conservation of mass for m>nαnm>\frac{n-\alpha}{n} on n\mathbb{R}^{n} has been formulated in [21]. This property also holds for m=nαnm=\frac{n-\alpha}{n} on n\mathbb{R}^{n} by a same proof as the case of Heisenberg group n\mathbb{H}^{n}.

Furthermore, we obtain some important regularity results of solutions to the equations (1.1). The first result is continuity at the points away from t=0t=0.

Theorem 3.

Let m>0m>0, 0<α<20<\alpha<2. J satisfies assumption (1.3). Let uu be the weak solution of the Cauchy problem (1.1). Then uC((0,+)×n)u\in C((0,+\infty)\times\mathbb{H}^{n}).

We also prove the Hölder regularity holds for the positive solution with uniform positive lower bound. There will be no degenerary (m>1m>1) or singularity (m<1m<1) since uδ>0u\geq\delta>0.

Theorem 4.

Let m>0m>0, 0<α<20<\alpha<2. J satisfies assumption (1.3). Let uu be the weak solution of the Cauchy problem (1.1). If uu is nonnegative with uniform positive lower bound, for each ϵ>0\epsilon>0, there exists 0<α<10<\alpha<1 such that uCα((ϵ,+)×n)u\in C^{\alpha}((\epsilon,+\infty)\times\mathbb{H}^{n}).

Moreover, if m1m\geq 1, the Hölder regularity will also hold for the sign-changing solutions since we can overcome the difficulty arising from the degeneracy at the vanishing points in a quantitative way. Here, the vanishing points (x0,t0)(x_{0},t_{0}) are the points satisfying u(x0,t0)=0u(x_{0},t_{0})=0.

Theorem 5.

Let m1m\geq 1, 0<α<20<\alpha<2. J satisfies assumption (1.3). Let uu be the bounded weak solution of the Cauchy problem (1.1). Then for each ϵ>0\epsilon>0, there exists 0<α<10<\alpha<1 such that uCα((ϵ,+)×n)u\in C^{\alpha}((\epsilon,+\infty)\times\mathbb{H}^{n}).

Remark 3.

The modulus of continuity follows the method introduced by De Giorgi in his classical proof to elliptic equations, see [12]. The linear nonlocal operator has been considered by Caffarelli, Chan and Vasseur in [5] using the De Giorgi’s method. Furthermore, the Hölder modulus of continuity for positive solution with uniform positive lower bound to the analogous nonlocal filtration type equation on n\mathbb{R}^{n} has been well studied in \autocitesttp_ainfe_dpartf_dpa, by using the non-quadratic energies and the De Giorgi’s method.

Remark 4.

The new result is, for the porous medium type equations (m1m\geq 1), the uniform Hölder continuity holds for all the points including the vanishing points. We constructed the iterated sequence of solutions with iterated nonlinearity functions in decreasing space-time cylinders. As the function value u(,t)u(\cdot,t) approaches zero, the derivatives of the iterated nonlinearity functions are not uniformly bounded, which results in the degeneracy of the oscillation reduction. To overcome this difficulty, we work on cylinders suitably scaled to reflect in a precise quantitative way the degeneracy at the vanishing points using the idea from \autociteshef_detpm_vj.

2 Preliminaries

2.1 Heisenberg group n\mathbb{H}^{n}

We recall some definitions and a few well-known facts concerning with the Heisenberg group n\mathbb{H}^{n}. For further details we refer to the book by Fischer and Ruzhansky, [14].

We identify the Heisenberg group n\mathbb{H}^{n} with 2n+1\mathbb{R}^{2n+1}. An element in the Heisenberg group n\mathbb{H}^{n} is denoted by

x:=(z,s)=(ξ1,,ξn,η1,,ηn,s).x\mathrel{\mathop{\mathchar 58\relax}}=(z,s)=(\xi_{1},\cdots,\xi_{n},\eta_{1},\cdots,\eta_{n},s).

For any x,x2n+1x,x^{\prime}\in\mathbb{R}^{2n+1}, the group multiplication is given by

xx:\displaystyle x\cdot x^{\prime}\mathrel{\mathop{\mathchar 58\relax}} =(ξ+ξ,η+η,s+s+2η,ξ2ξ,η)\displaystyle=(\xi+\xi^{\prime},\eta+\eta^{\prime},s+s^{\prime}+2\langle\eta,\xi^{\prime}\rangle-2\langle\xi,\eta^{\prime}\rangle)
=(ξ1+ξ1,,ξn+ξn,η1+η1,,ηn+ηn,s+s+2i=1n(ηiξiξiηi)).\displaystyle=(\xi_{1}+\xi^{\prime}_{1},\cdots,\xi_{n}+\xi^{\prime}_{n},\eta_{1}+\eta^{\prime}_{1},\cdots,\eta_{n}+\eta^{\prime}_{n},s+s^{\prime}+2\sum_{i=1}^{n}(\eta_{i}\xi^{\prime}_{i}-\xi_{i}\eta^{\prime}_{i})).

where ,,\langle\cdot,\cdot,\rangle is the standard inner product in n\mathbb{R}^{n}.

The neutral element of n\mathbb{H}^{n} is 0n=(0n,0n,0)0_{\mathbb{H}^{n}}=(0_{\mathbb{R}^{n}},0_{\mathbb{R}^{n}},0) and the inverse (z,s)1(z,s)^{-1} of the element (z,s)(z,s) is (z,s)(-z,-s). Define the dilations δλ\delta_{\lambda} on n\mathbb{H}^{n},

δλ(x)=(λξ,λη,λ2s),λ>0.\displaystyle\delta_{\lambda}(x)=(\lambda\xi,\lambda\eta,\lambda^{2}s),\leavevmode\nobreak\ \leavevmode\nobreak\ \lambda>0.

The Heisenberg group n\mathbb{H}^{n} can be considered as a sub-Riemannian manifold. The orthonormal basis of the Heisenberg group n\mathbb{H}^{n} is given by

Xj:=ξj+2ηjs,Yj:=ηj2ξjs, 1jn,T=s.\displaystyle X_{j}\mathrel{\mathop{\mathchar 58\relax}}=\frac{\partial}{\partial\xi_{j}}+2\eta_{j}\frac{\partial}{\partial s},\leavevmode\nobreak\ \leavevmode\nobreak\ Y_{j}\mathrel{\mathop{\mathchar 58\relax}}=\frac{\partial}{\partial\eta_{j}}-2\xi_{j}\frac{\partial}{\partial s},\leavevmode\nobreak\ \leavevmode\nobreak\ 1\leq j\leq n,\leavevmode\nobreak\ \leavevmode\nobreak\ T=\frac{\partial}{\partial s}.

which is the Jacobian basis of the Heisenberg Lie algebra 𝐡n\mathbf{h}^{n} of n\mathbb{H}^{n}. Here,

rank(Lie{X1,,Xn,Y1,,Yn,T}(0n))=2n+1.\displaystyle\text{rank}(\text{Lie}\{X_{1},\cdots,X_{n},Y_{1},\cdots,Y_{n},T\}(0_{\mathbb{H}^{n}}))=2n+1.

This shows that n\mathbb{H}^{n} is a Carnot group with the following stratification

𝐡n=span{X1,,Xn,Y1,,Yn}span{T}.\displaystyle\mathbf{h}^{n}=\text{span}\{X_{1},\cdots,X_{n},Y_{1},\cdots,Y_{n}\}\oplus\text{span}\{T\}.

Let (|z|4+t2)1/4=((|ξ|2+|η|2)2+t2)1/4(|z|^{4}+t^{2})^{1/4}=((|\xi|^{2}+|\eta|^{2})^{2}+t^{2})^{1/4} be the homogeneous quasi-norm on n\mathbb{H}^{n}, which is denote by ||n|{\cdot}|_{\mathbb{H}^{n}}. The Korányi-Cygan metric dn:n×nd_{\mathbb{H}^{n}}\mathrel{\mathop{\mathchar 58\relax}}\mathbb{H}^{n}\times\mathbb{H}^{n}\rightarrow\mathbb{R} is defined as

dn(x,x)=|x1x|n.\displaystyle d_{\mathbb{H}^{n}}(x,x^{\prime})=|x^{-1}\cdot x^{\prime}|_{\mathbb{H}^{n}}.

For the sake of readability, the subscript will be often omitted without causing misunderstandings. The Lebesgue measure dμ()\mathop{}\!\mathrm{d}\mu(\cdot) on 2n+1\mathbb{R}^{2n+1} will be the Haar measure on n\mathbb{H}^{n} which is uniquely defined up to some positive constant. Let Q=2n+2Q=2n+2 be the homogeneous degree corresponding to the automorphisms (δλ)λ>0(\delta_{\lambda})_{\lambda>0}.

For any fixed x0nx_{0}\in\mathbb{H}^{n} and R>0R>0, denote with BR(x0)B_{R}(x_{0}) the Korányi ball with center x0x_{0} and radius RR defined as

BR(x0):={xn:|x01x|R}.\displaystyle B_{R}(x_{0})\mathrel{\mathop{\mathchar 58\relax}}=\{x\in\mathbb{H}^{n}\mathrel{\mathop{\mathchar 58\relax}}|x_{0}^{-1}\cdot x|\leq R\}.

Let Ωn\Omega\subset\mathbb{H}^{n} be a domain, for each fC2(Ω;)f\in C^{2}(\Omega;\mathbb{R}), the negative sublaplacian \mathscr{L} is defined by,

f:=i=1nXi2fi=1nYi2f.\displaystyle\mathscr{L}f\mathrel{\mathop{\mathchar 58\relax}}=-\sum_{i=1}^{n}X_{i}^{2}f-\sum_{i=1}^{n}Y_{i}^{2}f.

The pure fractional powers of α/2\mathscr{L}^{\alpha/2} has been well defined and studied in \autocitessea_fg,hso_fg,hif_ff. For α>0\alpha>0, the operator α/2\mathscr{L}^{\alpha/2} can be written as

α/2=0+λα/2dE(λ)\displaystyle\mathscr{L}^{\alpha/2}=\int_{0}^{+\infty}\lambda^{\alpha/2}\mathop{}\!\mathrm{d}E(\lambda)

with domain

Wα/2,2(n):={uL2(n):α/4uL2(n)<}.W^{\alpha/2,2}(\mathbb{H}^{n})\mathrel{\mathop{\mathchar 58\relax}}=\{u\in L^{2}(\mathbb{H}^{n})\mathrel{\mathop{\mathchar 58\relax}}\mathinner{\!\left\lVert\mathscr{L}^{\alpha/4}u\right\rVert}_{L^{2}(\mathbb{H}^{n})}<\infty\}.

Here dE(λ)\mathop{}\!\mathrm{d}E(\lambda) is the spectral resolution of \mathscr{L} in L2(n)L^{2}(\mathbb{H}^{n}).

The fractional homogeneous Sobolev space H˙α/2(n)\dot{H}^{\alpha/{2}}(\mathbb{H}^{n}) is defined as the completion of C0(n)C_{0}^{\infty}(\mathbb{H}^{n}) with the norm

uH˙α/2(n)=α/4uL2(n).\mathinner{\!\left\lVert u\right\rVert}_{\dot{H}^{\alpha/{2}}(\mathbb{H}^{n})}=\mathinner{\!\left\lVert\mathscr{L}^{\alpha/{4}}u\right\rVert}_{L^{2}(\mathbb{H}^{n})}.

2.2 Weak Solutions

For simplicity, we denote Lp(n)\mathinner{\!\left\lVert\cdot\right\rVert}_{L^{p}(\mathbb{H}^{n})} by p\mathinner{\!\left\lVert\cdot\right\rVert}_{p} for 1p1\leq p\leq\infty and use the simplified notation umu^{m} instead of |u|m1u|u|^{m-1}u for sign changing solutions. In order to define weak solution, we consider the bilinear Dirichlet form,

J(u,v)=12nn(u(x)u(y))(v(x)v(y))J(x,y)dμ(x)dμ(y).\mathcal{E}_{J}(u,v)=\frac{1}{2}\int_{\mathbb{H}^{n}}\int_{\mathbb{H}^{n}}(u(x)-u(y))(v(x)-v(y))J(x,y)\mathop{}\!\mathrm{d}\mu(x)\mathop{}\!\mathrm{d}\mu(y).

For kernels satisfying the symmetry condition J(x,xy)=J(x,xy1)J(x,x\cdot y)=J(x,x\cdot y^{-1}) and functions f,gC02(n)f,g\in C_{0}^{2}(\mathbb{H}^{n}), we have

<f,g>=J(u,v).<\mathcal{L}f,g>=\mathcal{E}_{J}(u,v).

Denote the space ˙(n)\dot{\mathcal{H}}_{\mathcal{L}}(\mathbb{H}^{n}) as the closure of space C0(n)C_{0}^{\infty}(\mathbb{H}^{n}) under the seminorm J(u,u)12\mathcal{E}_{J}(u,u)^{\frac{1}{2}}. We also define (n)={fL2(n):J(f,f)<}\mathcal{H}_{\mathcal{L}}(\mathbb{H}^{n})=\{f\in L^{2}(\mathbb{H}^{n})\mathrel{\mathop{\mathchar 58\relax}}\mathcal{E}_{J}(f,f)<\infty\}. The condition (1.3) implies

c1(n,α,Λ)J(u,u)α/4u22c2(n,α,Λ)J(u,u).\displaystyle c_{1}(n,\alpha,\Lambda)\mathcal{E}_{J}(u,u)\leq\mathinner{\!\left\lVert\mathscr{L}^{\alpha/{4}}u\right\rVert}_{2}^{2}\leq c_{2}(n,\alpha,\Lambda)\mathcal{E}_{J}(u,u). (2.1)

Now we are ready to define the weak solutions of Cauchy problem (1.1).

Definition 1.

A function uu is a weak solution to equation (1.1) if:

  • uC([0,):L1(n))u\in C([0,\infty)\mathrel{\mathop{\mathchar 58\relax}}L^{1}(\mathbb{H}^{n})) and umLloc2((0,):˙(n))u^{m}\in L^{2}_{\text{loc}}((0,\infty)\mathrel{\mathop{\mathchar 58\relax}}\dot{\mathcal{H}}_{\mathcal{L}}(\mathbb{H}^{n}));

  • 0nutζ0J(um,ζ)=0\int_{0}^{\infty}\int_{\mathbb{H}^{n}}u\partial_{t}\zeta-\int_{0}^{\infty}\mathcal{E}_{J}(u^{m},\zeta)=0 for each ζCc2(n×(0,))\zeta\in C^{2}_{c}(\mathbb{H}^{n}\times(0,\infty));

  • u(x,0)=u0(x)u(x,0)=u_{0}(x) almost everywhere.

If the equation (1.1) is rewritten equivalently as

{tw1/m+w=0,xn,t>0w(x,0)=w0(x)=u0m(x),xn.\left\{\begin{array}[]{ll}\partial_{t}w^{1/m}+\mathcal{L}w=0,&x\in\mathbb{H}^{n},t>0\\ w(x,0)=w_{0}(x)=u_{0}^{m}(x),&x\in\mathbb{H}^{n}.\end{array}\right. (2.2)

with w01/mL1(n)w_{0}^{1/m}\in L^{1}(\mathbb{H}^{n}) and w=umw=u^{m}. Equivalently, we define the weak solutions to the Cauchy problem (2.2).

Definition 2.

A function ww is a weak solution to equation (2.2) if:

  • w1/mC([0,):L1(n))w^{1/m}\in C([0,\infty)\mathrel{\mathop{\mathchar 58\relax}}L^{1}(\mathbb{H}^{n})) and wLloc2((0,):˙(n))w\in L^{2}_{\text{loc}}((0,\infty)\mathrel{\mathop{\mathchar 58\relax}}\dot{\mathcal{H}}_{\mathcal{L}}(\mathbb{H}^{n}));

  • 0nw1/mtζ0J(w,ζ)=0\int_{0}^{\infty}\int_{\mathbb{H}^{n}}w^{1/m}\partial_{t}\zeta-\int_{0}^{\infty}\mathcal{E}_{J}(w,\zeta)=0 for each ζCc2(n×(0,))\zeta\in C^{2}_{c}(\mathbb{H}^{n}\times(0,\infty));

  • w(x,0)=w0(x)w(x,0)=w_{0}(x) almost everywhere.

2.3 Some Inequalities

We quote or prove some important inequalities which are useful in the following sections. The Stroock Varopoulos inequality will be proved by the extension map which can be used to reconstruct the operator α/2(u)\mathscr{L}^{\alpha/2}(u), see \autociteshif_ff,epa_sp.

Lemma 1.

Let 0<α<20<\alpha<2. If ψ1(v),v˙α/2(n)\psi_{1}(v),v\in\dot{\mathcal{H}}^{\alpha/2}(\mathbb{H}^{n}) and (Ψ)2ψ1(v)(\Psi^{\prime})^{2}\leq\psi_{1}^{\prime}(v), then

n(α/2ψ1(v))(v)n(α/2Ψ(v))Ψ(v).\displaystyle\int_{\mathbb{H}^{n}}(\mathscr{L}^{\alpha/2}\psi_{1}(v))(v)\geq\int_{\mathbb{H}^{n}}(\mathscr{L}^{\alpha/2}\Psi(v))\Psi(v).
Proof.

Recall that there exists a Caffarelli-Silvestre type extension for the pure power fractional operator α/2\mathscr{L}^{\alpha/2}, see \autociteshif_ff,epa_sp. Define the space X˙α/2(n×+)\dot{X}^{\alpha/2}(\mathbb{H}^{n}\times\mathbb{R}^{+}) as the completion of C(n×[0,))C^{\infty}(\mathbb{H}^{n}\times[0,\infty)) under the seminorm

wX˙α/2(n×+)=0ny1α|Hw(x)|2dμ(x)dy.\mathinner{\!\left\lVert w\right\rVert}_{\dot{X}^{\alpha/2}(\mathbb{H}^{n}\times\mathbb{R}^{+})}=\int_{0}^{\infty}\int_{\mathbb{H}^{n}}y^{1-\alpha}|\nabla_{H}w(x)|^{2}\mathop{}\!\mathrm{d}\mu(x)\mathop{}\!\mathrm{d}y.

Here H\nabla_{H} is the horizontal gradient on n×(0,)\mathbb{H}^{n}\times(0,\infty) defined by,

Hf:=i=1n(Xif)Xi+i=1n(Yif)Yi+(yf)y.\displaystyle\nabla_{H}f\mathrel{\mathop{\mathchar 58\relax}}=\sum_{i=1}^{n}(X_{i}f)X_{i}+\sum_{i=1}^{n}(Y_{i}f)Y_{i}+(\partial_{y}f)\partial_{y}.

and

|Hf|2:=i=1n|Xif|2+i=1n|Yif|2+|yf|2.\displaystyle|\nabla_{H}f|^{2}\mathrel{\mathop{\mathchar 58\relax}}=\sum_{i=1}^{n}|X_{i}f|^{2}+\sum_{i=1}^{n}|Y_{i}f|^{2}+|\partial_{y}f|^{2}.

For every uH˙α/2(n)u\in\dot{H}^{\alpha/2}(\mathbb{H}^{n}), there exists a norm-preserving extension map E:H˙α/2(n)X˙α/2(n×+)E\mathrel{\mathop{\mathchar 58\relax}}\dot{H}^{\alpha/2}(\mathbb{H}^{n})\rightarrow\dot{X}^{\alpha/2}(\mathbb{H}^{n}\times\mathbb{R}^{+}) satisfying,

{(y1α+y1αyy+(1α)yαy)E(u)(x,y)=0,(x,y)n×+E(u)(x,0)=u(x),xn.\left\{\begin{array}[]{ll}(y^{1-\alpha}\mathscr{L}+y^{1-\alpha}\partial_{yy}+(1-\alpha)y^{-\alpha}\partial_{y})E(u)(x,y)=0,&(x,y)\in\mathbb{H}^{n}\times\mathbb{R}^{+}\\ E(u)(x,0)=u(x),&x\in\mathbb{H}^{n}.\end{array}\right.

Moreover, the pure power opertaor α/2\mathscr{L}^{\alpha/2} can be reconstructed as the Neumann data of the extension map,

α/2u(x)=μαlimy0+y1αE(u)y,μα=2α1Γ(α/2)Γ(1α/2).\mathscr{L}^{\alpha/2}u(x)=-\mu_{\alpha}\lim\limits_{y\rightarrow 0^{+}}y^{1-\alpha}\frac{\partial{E(u)}}{\partial y},\leavevmode\nobreak\ \leavevmode\nobreak\ \mu_{\alpha}=2^{\alpha-1}\Gamma({\alpha/2})\Gamma(1-\alpha/2).

Then,

n(α/2ψ1(v))(v)\displaystyle\int_{\mathbb{H}^{n}}(\mathscr{L}^{\alpha/2}\psi_{1}(v))(v) =μα0ny1αHE(ψ1(v)),HE(v)dμ(x)dy\displaystyle=\mu_{\alpha}\int_{0}^{\infty}\int_{\mathbb{H}^{n}}y^{1-\alpha}\langle\nabla_{H}E(\psi_{1}(v)),\nabla_{H}E(v)\rangle\mathop{}\!\mathrm{d}\mu(x)\mathop{}\!\mathrm{d}y
=μα0ny1αHψ1(E(v)),HE(v)dμ(x)dy\displaystyle=\mu_{\alpha}\int_{0}^{\infty}\int_{\mathbb{H}^{n}}y^{1-\alpha}\langle\nabla_{H}\psi_{1}(E(v)),\nabla_{H}E(v)\rangle\mathop{}\!\mathrm{d}\mu(x)\mathop{}\!\mathrm{d}y
μα0ny1αHE(Ψ(v)),HE(Ψ(v))dμ(x)dy\displaystyle\geq\mu_{\alpha}\int_{0}^{\infty}\int_{\mathbb{H}^{n}}y^{1-\alpha}\langle\nabla_{H}E(\Psi(v)),\nabla_{H}E(\Psi(v))\rangle\mathop{}\!\mathrm{d}\mu(x)\mathop{}\!\mathrm{d}y
=n(α/2Ψ(v))Ψ(v).\displaystyle=\int_{\mathbb{H}^{n}}(\mathscr{L}^{\alpha/2}\Psi(v))\Psi(v).

From the equivalence of the norm: c1(n,α,Λ)J(u,u)α/4u22c2(n,α,Λ)J(u,u)c_{1}(n,\alpha,\Lambda)\mathcal{E}_{J}(u,u)\leq\mathinner{\!\left\lVert\mathscr{L}^{\alpha/{4}}u\right\rVert}_{2}^{2}\leq c_{2}(n,\alpha,\Lambda)\mathcal{E}_{J}(u,u), we obtain

Lemma 2.

Let 0<α<20<\alpha<2, then

J(ψ1(v),v)c(n,α,Λ)J(Ψ(v),Ψ(v)),\displaystyle\mathcal{E}_{J}(\psi_{1}(v),v)\geq c(n,\alpha,\Lambda)\mathcal{E}_{J}(\Psi(v),\Psi(v)), (2.3)

whenever ψ1(v),v˙(n)=˙α/2(n)\psi_{1}(v),v\in\dot{\mathcal{H}}_{\mathcal{L}}(\mathbb{H}^{n})=\dot{\mathcal{H}}^{\alpha/2}(\mathbb{H}^{n}) and (Ψ)2ψ1(v).(\Psi^{\prime})^{2}\leq\psi_{1}^{\prime}(v).

The Hardy-Littlewood-Sobolev’s inequality also holds for the pure power fractional operator α/2\mathscr{L}^{\alpha/2} on n\mathbb{H}^{n}, which had been proved in [15],

Lemma 3.

For each vv such that α/2vLr(n)\mathscr{L}^{\alpha/2}v\in L^{r}(\mathbb{H}^{n}), 1<r<Q/α1<r<Q/\alpha, 0<α<20<\alpha<2, the following inequality holds:

vr1C(Q,r,α)α/2vr.\displaystyle\mathinner{\!\left\lVert v\right\rVert}_{{r_{1}}}\leq C(Q,r,\alpha)\mathinner{\!\left\lVert\mathscr{L}^{\alpha/2}v\right\rVert}_{r}. (2.4)

Here r1=QrQαrr_{1}=\frac{Qr}{Q-\alpha r}.

By the Hardy-Littlewood-Sobolev’s inequality and Hölder’s inequality, we obtain,

Lemma 4.

Let p>1p>1, r>1r>1, 0<α<20<\alpha<2. For each vLp(n)v\in L^{p}(\mathbb{H}^{n}) with α/2vLr(n)\mathscr{L}^{\alpha/2}v\in L^{r}(\mathbb{H}^{n}), we have

vr2γ+1C(p,r,α,Q)α/2vrvpγ.\displaystyle\mathinner{\!\left\lVert v\right\rVert}_{{r_{2}}}^{\gamma+1}\leq C(p,r,\alpha,Q)\mathinner{\!\left\lVert\mathscr{L}^{\alpha/2}v\right\rVert}_{r}\mathinner{\!\left\lVert v\right\rVert}_{p}^{\gamma}. (2.5)

where r2=Q(rp+rp)r(Qα)r_{2}=\frac{Q(rp+r-p)}{r(Q-\alpha)}, γ=p(r1)r\gamma=\frac{p(r-1)}{r}.

3 Existence and Uniqueness of Weak Solutions

3.1 Existence of Weak Solutions for Bounded Initial Data

Crandall and Liggett [9] introduced a useful method to construct mild solutions, which is called Implicit Time Discretization. In order to apply the Crandall-Liggett’s Theorem, we prove the existence of weak solution to the following equation,

v1/m+v=ginn,gL1(n)L(n).\displaystyle v^{1/m}+\mathcal{L}v=g\leavevmode\nobreak\ \leavevmode\nobreak\ \text{in}\leavevmode\nobreak\ \leavevmode\nobreak\ \mathbb{H}^{n},\leavevmode\nobreak\ \leavevmode\nobreak\ g\in L^{1}(\mathbb{H}^{n})\cap L^{\infty}(\mathbb{H}^{n}). (3.1)

and the contractivity of solutions in L1(n)L^{1}(\mathbb{H}^{n}).

Proposition 1.

For each gL1(n)L(n)g\in L^{1}(\mathbb{H}^{n})\cap L^{\infty}(\mathbb{H}^{n}), there exists a unique weak solution v˙(n)v\in\dot{\mathcal{H}}_{\mathcal{L}}(\mathbb{H}^{n}) to the equation (3.1), such that the following weak formulation holds,

J(v,ζ)+nv1/mζngζ=0,foreveryζC0(n).\displaystyle\mathcal{E}_{J}(v,\zeta)+\int_{\mathbb{H}^{n}}v^{1/m}\zeta-\int_{\mathbb{H}^{n}}g\zeta=0,\leavevmode\nobreak\ \leavevmode\nobreak\ \text{for}\leavevmode\nobreak\ \text{every}\leavevmode\nobreak\ \zeta\in C_{0}^{\infty}(\mathbb{H}^{n}). (3.2)

Furthermore, v1/mg\mathinner{\!\left\lVert v^{1/m}\right\rVert}_{\infty}\leq\mathinner{\!\left\lVert g\right\rVert}_{\infty} and v1/m1g1\mathinner{\!\left\lVert v^{1/m}\right\rVert}_{1}\leq\mathinner{\!\left\lVert g\right\rVert}_{1}. If vv and v~\tilde{v} are solutions corresponding to gg and g~\tilde{g} respectively, the following T-contraction inequality holds,

n[v1/m(x)v~1/m(x)]+dμ(x)n[g(x)g~(x)]+dμ(x).\displaystyle\int_{\mathbb{H}^{n}}[v^{1/m}(x)-\tilde{v}^{1/m}(x)]_{+}\mathop{}\!\mathrm{d}\mu(x)\leq\int_{\mathbb{H}^{n}}[g(x)-\tilde{g}(x)]_{+}\mathop{}\!\mathrm{d}\mu(x). (3.3)
Proof.

The weak solution is the minimizer of the following convex functional

J(v)=12J(v,v)+nmm+1|v|1m+1nvg.J(v)=\frac{1}{2}\mathcal{E}_{J}(v,v)+\int_{\mathbb{H}^{n}}\frac{m}{m+1}|v|^{\frac{1}{m}+1}-\int_{\mathbb{H}^{n}}vg.

for each v˙(n)v\in\dot{\mathcal{H}}_{\mathcal{L}}(\mathbb{H}^{n}). Since

|nvg|\displaystyle|\int_{\mathbb{H}^{n}}vg| ϵv2QQα2+C(ϵ)g2QQ+α2\displaystyle\leq\epsilon\mathinner{\!\left\lVert v\right\rVert}_{\frac{2Q}{Q-\alpha}}^{2}+C(\epsilon)\mathinner{\!\left\lVert g\right\rVert}_{\frac{2Q}{Q+\alpha}}^{2}
C(Q,α)ϵvH˙0α/22+C(ϵ)g2QQ+α2\displaystyle\leq C(Q,\alpha)\epsilon\mathinner{\!\left\lVert v\right\rVert}_{\dot{H}^{\alpha/2}_{0}}^{2}+C(\epsilon)\mathinner{\!\left\lVert g\right\rVert}_{\frac{2Q}{Q+\alpha}}^{2}
C(Q,α,Λ)ϵJ(v,v)+C(ϵ)g2QQ+α2.\displaystyle\leq C(Q,\alpha,\Lambda)\epsilon\mathcal{E}_{J}(v,v)+C({\epsilon})\mathinner{\!\left\lVert g\right\rVert}_{\frac{2Q}{Q+\alpha}}^{2}.

Therefore, J(v)J(v) is coercive, convex and lower semi-continuous and there exists a unique weak solution vv to equation (3.1). Let vv and v~\tilde{v} be two weak solutions of equation (3.1) corresponding to the inhomogeneous terms gg and g~\tilde{g}. We use pn(vv~)p_{n}(v-\tilde{v}) as test functions in the weak formulation. Here pnp_{n} is the smooth approximation of sign function with 0pn()10\leq p_{n}(\cdot)\leq 1 and pn()0p_{n}^{\prime}(\cdot)\geq 0. We obtain

n(v1/mv~1/m)pn(vv~)+J(vv~,pn(vv~))=n(gg~)pn(vv~).\displaystyle\int_{\mathbb{H}^{n}}(v^{1/m}-\tilde{v}^{1/m})p_{n}(v-\tilde{v})+\mathcal{E}_{J}(v-\tilde{v},p_{n}(v-\tilde{v}))=\int_{\mathbb{H}^{n}}(g-\tilde{g})p_{n}(v-\tilde{v}).

Passing to the limit,

n(v1/mv~1/m)+n(gg~)sign(vv~)n(gg~)+.\displaystyle\int_{\mathbb{H}^{n}}(v^{1/m}-\tilde{v}^{1/m})_{+}\leq\int_{\mathbb{H}^{n}}(g-\tilde{g})\text{sign}(v-\tilde{v})\leq\int_{\mathbb{H}^{n}}(g-\tilde{g})_{+}.

Here we apply the Stroock-Varopoulos inequality to obtain

J(vv~,pn(vv~))0.\displaystyle\mathcal{E}_{J}(v-\tilde{v},p_{n}(v-\tilde{v}))\geq 0.

since pn()0p_{n}^{\prime}(\cdot)\geq 0. Furthermore, v=gmv=\mathinner{\!\left\lVert g\right\rVert}_{\infty}^{m} is a supersolution to equation (3.1), we get

v1/mg.\displaystyle\mathinner{\!\left\lVert v^{1/m}\right\rVert}_{\infty}\leq\mathinner{\!\left\lVert g\right\rVert}_{\infty}.

Choosing vv as the test function, we obtain

J(v,v)ngvC.\displaystyle\mathcal{E}_{J}(v,v)\leq\int_{\mathbb{H}^{n}}gv\leq C. (3.4)

Equivalently, define v=umv=u^{m}, we obtain

Proposition 2.

For every gL1(n)L(n)g\in L^{1}(\mathbb{H}^{n})\cap L^{\infty}(\mathbb{H}^{n}), there exists a unique weak solution um˙(n)u^{m}\in\dot{\mathcal{H}}_{\mathcal{L}}(\mathbb{H}^{n}) to the equation u+um=gu+\mathcal{L}u^{m}=g in n\mathbb{H}^{n}, such that the following weak formulation holds,

J(um,ζ)+nuζngζ=0,foreveryζC0(n).\displaystyle\mathcal{E}_{J}(u^{m},\zeta)+\int_{\mathbb{H}^{n}}u\zeta-\int_{\mathbb{H}^{n}}g\zeta=0,\leavevmode\nobreak\ \leavevmode\nobreak\ \text{for}\leavevmode\nobreak\ \text{every}\leavevmode\nobreak\ \zeta\in C_{0}^{\infty}(\mathbb{H}^{n}). (3.5)

Furthermore, ug\mathinner{\!\left\lVert u\right\rVert}_{\infty}\leq\mathinner{\!\left\lVert g\right\rVert}_{\infty} and u1g1\mathinner{\!\left\lVert u\right\rVert}_{1}\leq\mathinner{\!\left\lVert g\right\rVert}_{1}. If vv and v~\tilde{v} are two solutions corresponding to gg and g~\tilde{g} respectively, the following T-contraction inequality holds,

n[u(x)u~(x)]+dμ(x)n[g(x)g~(x)]+dμ(x).\displaystyle\int_{\mathbb{H}^{n}}[u(x)-\tilde{u}(x)]_{+}\mathop{}\!\mathrm{d}\mu(x)\leq\int_{\mathbb{H}^{n}}[g(x)-\tilde{g}(x)]_{+}\mathop{}\!\mathrm{d}\mu(x). (3.6)

From the above Proposition, the operator \mathcal{L} is accretive and satisfies the rank condition required in the Crandall-Liggett’s theorem. We apply this theorem to equation (1.1) to obtain the so-called mild solution. Moreover, it is a weak solution.

Proposition 3.

For each u0L1(n)L(n)u_{0}\in L^{1}(\mathbb{H}^{n})\cap L^{\infty}(\mathbb{H}^{n}), there exists a weak solution uC([0,):L1(n))u\in C([0,\infty)\mathrel{\mathop{\mathchar 58\relax}}L^{1}(\mathbb{H}^{n})) to equation (1.1). Moreover, uu0\mathinner{\!\left\lVert u\right\rVert}_{\infty}\leq\mathinner{\!\left\lVert u_{0}\right\rVert}_{\infty} and u1u01\mathinner{\!\left\lVert u\right\rVert}_{1}\leq\mathinner{\!\left\lVert u_{0}\right\rVert}_{1}. In addition, if uu and u~\tilde{u} are two such solutions corresponding to u0u_{0} and u0~\tilde{u_{0}}, the following T-contraction inequality holds,

n[v1/m(x)v~1/m(x)]+dμ(x)n[u0(x)u~0(x)]+dμ(x).\displaystyle\int_{\mathbb{H}^{n}}[v^{1/m}(x)-\tilde{v}^{1/m}(x)]_{+}\mathop{}\!\mathrm{d}\mu(x)\leq\int_{\mathbb{H}^{n}}[u_{0}(x)-\tilde{u}_{0}(x)]_{+}\mathop{}\!\mathrm{d}\mu(x). (3.7)
Proof.

For each T>0T>0, we divide interval [0,T][0,T] into NN subintervals. Let ϵ=T/N\epsilon=T/N, we construct the piecewise constant function uϵ,ku_{\epsilon,k} in each interval (tk1,tk](t_{k-1},t_{k}], where tk=kϵt_{k}=k\epsilon (k=1,2,,Nk=1,2,\cdots,N) as the solutions to the following elliptic equation,

uϵ,km+1ϵ(uϵ,kuϵ,k1)=0.\displaystyle\mathcal{L}u_{\epsilon,k}^{m}+\frac{1}{\epsilon}(u_{\epsilon,k}-u_{\epsilon,k-1})=0. (3.8)

with uϵ,0=u0u_{\epsilon,0}=u_{0}. By Crandall-Liggett theorem, uϵu_{\epsilon} converges in L1(n)L^{1}(\mathbb{H}^{n}) to some function uC([0,):L1(n))u\in C([0,\infty)\mathrel{\mathop{\mathchar 58\relax}}L^{1}(\mathbb{H}^{n})). Furthermore, uϵu0\mathinner{\!\left\lVert u_{\epsilon}\right\rVert}_{\infty}\leq\mathinner{\!\left\lVert u_{0}\right\rVert}_{\infty} inherited from the elliptic equations. And uϵu_{\epsilon} converges in weak topology to uL(n×[0,))u\in L^{\infty}(\mathbb{H}^{n}\times[0,\infty)). Here uu is the mild solution to Cauchy problem (1.1). Multiplying the equation (3.8) by uϵ,kmu_{\epsilon,k}^{m} and applying Young’s inequality, we get

0TJ(uϵm,uϵm)dt1m+1n|u0|m+1dμ(x).\displaystyle\int_{0}^{T}\mathcal{E}_{J}(u_{\epsilon}^{m},u_{\epsilon}^{m})\mathop{}\!\mathrm{d}t\leq\frac{1}{m+1}\int_{\mathbb{H}^{n}}|u_{0}|^{m+1}\mathop{}\!\mathrm{d}\mu(x).

Passing to the limit, we obtain

0TJ(um,um)dt1m+1n|u0|m+1dμ(x).\displaystyle\int_{0}^{T}\mathcal{E}_{J}(u^{m},u^{m})\mathop{}\!\mathrm{d}t\leq\frac{1}{m+1}\int_{\mathbb{H}^{n}}|u_{0}|^{m+1}\mathop{}\!\mathrm{d}\mu(x). (3.9)

Passing to the limit, we obtain the parabolic weak formulation, see the analogues in [24]. ∎

3.2 Uniqueness of Solutions

In this section, we prove the uniqueness of the constructed solutions. If m>mm>m^{*}, the weak solution defined in section 2 is always unique. However, if mmm\leq m^{*}, we need a stronger assumption of solutions to guarantee the uniqueness.

Proposition 4.

Let u0L1(n)u_{0}\in L^{1}(\mathbb{H}^{n}) and m>mm>m^{*}, Problem (1.1) has at most one weak solution.

Proof.

Following the analogues in [24], we claim that

uLm+1(n×(0,T))ifm>m.\displaystyle u\in L^{m+1}(\mathbb{H}^{n}\times(0,T))\leavevmode\nobreak\ \text{if}\leavevmode\nobreak\ m>m^{*}.

By inequalites (2,3), (2.4) and Hölder’s inequality, we get

0Tn|u|m+1dμ(x)dt\displaystyle\int_{0}^{T}\int_{\mathbb{H}^{n}}|u|^{m+1}\mathop{}\!\mathrm{d}\mu(x)\mathop{}\!\mathrm{d}t 0T(n|u|dμ(x))β((n|u|2QmQαdμ(x))1β)dt\displaystyle\leq\int_{0}^{T}(\int_{\mathbb{H}^{n}}|u|\mathop{}\!\mathrm{d}\mu(x))^{\beta}((\int_{\mathbb{H}^{n}}|u|^{\frac{2Qm}{Q-\alpha}}\mathop{}\!\mathrm{d}\mu(x))^{1-\beta})\mathop{}\!\mathrm{d}t
C(T)maxt[0,T]u(,t)1β[0Tum2QQα2dt]1γ\displaystyle\leq C(T)\max_{t\in[0,T]}\mathinner{\!\left\lVert u(\cdot,t)\right\rVert}_{1}^{\beta}[\int_{0}^{T}\mathinner{\!\left\lVert u^{m}\right\rVert}^{2}_{\frac{2Q}{Q-\alpha}}\mathop{}\!\mathrm{d}t]^{1-\gamma}
C(T)[0TJ(um,um)dt]1γ\displaystyle\leq C(T)[\int_{0}^{T}\mathcal{E}_{J}(u^{m},u^{m})\mathop{}\!\mathrm{d}t]^{1-\gamma}

where β=Q(m1)+α(m+1)Q(2m1)+α\beta=\frac{Q(m-1)+\alpha(m+1)}{Q(2m-1)+\alpha} and γ=Q(m1)+αQ(2m1)+α\gamma=\frac{Q(m-1)+\alpha}{Q(2m-1)+\alpha}.

Define ψ(x,t)=tT(umu~m)(x,s)ds\psi(x,t)=\int_{t}^{T}(u^{m}-\tilde{u}^{m})(x,s)\mathop{}\!\mathrm{d}s for 0tT0\leq t\leq T and ψ=0\psi=0 for tTt\geq T. Let uu and u~\tilde{u} be two weak solutions to equation (1.1) with u(,0)=u~(,0)=u0()u(\cdot,0)=\tilde{u}(\cdot,0)=u_{0}(\cdot). From the parabolic weak formulation, we have

0Tn\displaystyle\int_{0}^{T}\int_{\mathbb{H}^{n}} (uu~)(x,t)(umu~m)(x,t)dμ(x)dt\displaystyle(u-\tilde{u})(x,t)(u^{m}-\tilde{u}^{m})(x,t)\mathop{}\!\mathrm{d}\mu(x)\mathop{}\!\mathrm{d}t
+12J(0T(umu~m)(x,s)ds,0T(umu~m)(x,s)ds)=0.\displaystyle+\frac{1}{2}\mathcal{E}_{J}(\int_{0}^{T}(u^{m}-\tilde{u}^{m})(x,s)\mathop{}\!\mathrm{d}s,\int_{0}^{T}(u^{m}-\tilde{u}^{m})(x,s)\mathop{}\!\mathrm{d}s)=0.

Therefore, we prove u=u~u=\tilde{u} if they have the same initial data u0u_{0}. ∎

Considering the case of mmm\leq m^{*}, we prove the uniqueness of the weak solutions which is strong in the sense;

tuL((τ,):L1(n)),foreveryτ>0.\displaystyle\partial_{t}u\in L^{\infty}((\tau,\infty)\mathrel{\mathop{\mathchar 58\relax}}L^{1}(\mathbb{H}^{n})),\rm{\leavevmode\nobreak\ \leavevmode\nobreak\ for\leavevmode\nobreak\ every\leavevmode\nobreak\ }\tau>0. (3.10)

We prove the following comparison theorem as [23] to obtain the uniqueness of the weak solutions satisfying (3.10), which are called strong solutions throughout this paper.

Proposition 5.

Let m>0m>0, if u1u_{1} and u2u_{2} are strong solutions to Problem (1.1) with initial data u0,1,u0,2L1(n)u_{0,1},\leavevmode\nobreak\ u_{0,2}\in L^{1}(\mathbb{H}^{n}), then for 0t1t20\leq t_{1}\leq t_{2}, we have

n(u1u2)+(x,t2)dμ(x)n(u0,1u0,2)+(x,t1)dμ(x)\displaystyle\int_{\mathbb{H}^{n}}(u_{1}-u_{2})_{+}(x,t_{2})\mathop{}\!\mathrm{d}\mu(x)\leq\int_{\mathbb{H}^{n}}(u_{0,1}-u_{0,2})_{+}(x,t_{1})\mathop{}\!\mathrm{d}\mu(x) (3.11)
Proof.

Let pnp_{n} be the smooth approximation of sign function with 0pn()10\leq p_{n}(\cdot)\leq 1 and pn()0p_{n}^{\prime}(\cdot)\geq 0. Using pn(u1u2)p_{n}(u_{1}-u_{2}) as the test function in the parabolic weak formulation, for 0<t1<t20<t_{1}<t_{2}, we have

t1t2nt(u1u2)pn(u1u2)dμ(x)dt=t1t2J(u1u2,pn(u1u2))dt.\displaystyle\int_{t_{1}}^{t_{2}}\int_{\mathbb{H}^{n}}\partial_{t}(u_{1}-u_{2})p_{n}(u_{1}-u_{2})\mathop{}\!\mathrm{d}\mu(x)\mathop{}\!\mathrm{d}t=\int_{t_{1}}^{t_{2}}\mathcal{E}_{J}(u_{1}-u_{2},p_{n}(u_{1}-u_{2}))\mathop{}\!\mathrm{d}t.

Using Stroock-Varopoulos inequality, we have

t1t2J(u1u2,pn(u1u2))dt0.\displaystyle\int_{t_{1}}^{t_{2}}\mathcal{E}_{J}(u_{1}-u_{2},p_{n}(u_{1}-u_{2}))\mathop{}\!\mathrm{d}t\leq 0.

Hence, we obtain

t1t2nt(u1u2)pn(u1u2)dμ(x)dt0\displaystyle\int_{t_{1}}^{t_{2}}\int_{\mathbb{H}^{n}}\partial_{t}(u_{1}-u_{2})p_{n}(u_{1}-u_{2})\mathop{}\!\mathrm{d}\mu(x)\mathop{}\!\mathrm{d}t\leq 0

Passing to the limit, we have

n(u1u2)+(x,t2)dμ(x)n(u1u2)+(x,t1)dμ(x)\displaystyle\int_{\mathbb{H}^{n}}(u_{1}-u_{2})_{+}(x,t_{2})\mathop{}\!\mathrm{d}\mu(x)\leq\int_{\mathbb{H}^{n}}(u_{1}-u_{2})_{+}(x,t_{1})\mathop{}\!\mathrm{d}\mu(x)

Let t1t_{1} approach zero, we prove this inequality for t1=0t_{1}=0. ∎

4 Strong solutions

In this section, we prove the constructed weak solutions are indeed strong solutions. First, we prove that tu\partial_{t}u is a bounded radon measure.

Lemma 5.

Assume m1m\neq 1, let uu be a weak solution to equation (1.1), then tu\partial_{t}u is a bounded radon measure.

Proof.

If uu is a solution with initial data u0u_{0} and λ\lambda is a positive constant, then

u~(x,t)=λu(x,λm1t).\tilde{u}(x,t)=\lambda u(x,\lambda^{m-1}t).

is the weak solution with data u~0(x)=λu0(x)\tilde{u}_{0}(x)=\lambda u_{0}(x). Now fix t>0t>0 and h>0h>0 and put λm1t=t+h\lambda^{m-1}t=t+h so that λ>1\lambda>1. By L1L^{1} contractivity estimate, we get

u(x,t+h)u(x,t)12u01(m1)t(h+o(h)).\displaystyle\mathinner{\!\left\lVert u(x,t+h)-u(x,t)\right\rVert}_{1}\leq 2\frac{\mathinner{\!\left\lVert u_{0}\right\rVert}_{1}}{(m-1)t}(h+o(h)). (4.1)

Therefore, tu\partial_{t}u is a finite radon measure. ∎

First, we prove tum+12Lloc2(n×(0,))\partial_{t}u^{\frac{m+1}{2}}\in L_{\text{loc}}^{2}(\mathbb{H}^{n}\times(0,\infty)) by the method of Steklov averages.

Lemma 6.

The function z=um+12z=u^{\frac{m+1}{2}} has the property that tzLloc2(n×(0,))\partial_{t}z\in L_{\text{loc}}^{2}(\mathbb{H}^{n}\times(0,\infty)).

Proof.

Since tum\partial_{t}u^{m} may not be a function, it’s not a reasonable test function. The Steklov averages can be used to overcome the difficulties. For any gL1(n×(0,+))g\in L^{1}(\mathbb{H}^{n}\times(0,+\infty)), we define the Steklov average

gh(x,t)=1htt+hg(x,s)ds.\displaystyle g^{h}(x,t)=\frac{1}{h}\int_{t}^{t+h}g(x,s)\mathop{}\!\mathrm{d}s.

We have

δhg(x,t)=tgh(x,t):=g(x,t+h)g(x,t)halmosteverywhere.\displaystyle\delta^{h}g(x,t)=\partial_{t}g^{h}(x,t)\mathrel{\mathop{\mathchar 58\relax}}=\frac{g(x,t+h)-g(x,t)}{h}\leavevmode\nobreak\ \leavevmode\nobreak\ \text{almost}\leavevmode\nobreak\ \text{everywhere}.

Since tuhL1(n×(0,+))\partial_{t}u^{h}\in L^{1}(\mathbb{H}^{n}\times(0,+\infty)), we rewrite the weak formulation of equation (1.1) as

0ntuhζdμ(x)dt+0J((um)h,ζ)dt=0foreachζCc2(n×(0,)).\displaystyle\int_{0}^{\infty}\int_{\mathbb{H}^{n}}\partial_{t}u^{h}\zeta\mathop{}\!\mathrm{d}\mu(x)\mathop{}\!\mathrm{d}t+\int_{0}^{\infty}\mathcal{E}_{J}((u^{m})^{h},\zeta)\mathop{}\!\mathrm{d}t=0\leavevmode\nobreak\ \leavevmode\nobreak\ \text{for}\leavevmode\nobreak\ \text{each}\leavevmode\nobreak\ \zeta\in C^{2}_{c}(\mathbb{H}^{n}\times(0,\infty)). (4.2)

Take ζ=η(t)t(um)h\zeta=\eta(t)\partial_{t}(u^{m})^{h} as the admissible test function in the weak formulation (4.2), where η(t)C0((0,+))\eta(t)\in C_{0}^{\infty}((0,+\infty)) is a cut-off function satisfying 0η(t)10\leq\eta(t)\leq 1, η(t)=1\eta(t)=1 for t[t1,t2]t\in[t_{1},t_{2}]. We have

0nηtuht(um)hdμ(x)dt\displaystyle\int_{0}^{\infty}\int_{\mathbb{H}^{n}}\eta\partial_{t}u^{h}\partial_{t}(u^{m})^{h}\mathop{}\!\mathrm{d}\mu(x)\mathop{}\!\mathrm{d}t =120η(t)tJ((um)h,(um)h)dt\displaystyle=-\frac{1}{2}\int_{0}^{\infty}\eta(t)\partial_{t}\mathcal{E}_{J}((u^{m})^{h},(u^{m})^{h})\mathop{}\!\mathrm{d}t
=120tη(t)J((um)h,(um)h)dt.\displaystyle=\frac{1}{2}\int_{0}^{\infty}\partial_{t}\eta(t)\mathcal{E}_{J}((u^{m})^{h},(u^{m})^{h})\mathop{}\!\mathrm{d}t.

Notice that (δhum)(δhu)c(δhu(m+1)/2)2(\delta^{h}u^{m})(\delta^{h}u)\geq c(\delta^{h}u^{(m+1)/2})^{2}, see [23], we obtain

t1t2n(δhu(m+1)/2)2dμ(x)dtC.\displaystyle\int_{t_{1}}^{t_{2}}\int_{\mathbb{H}^{n}}(\delta^{h}u^{(m+1)/2})^{2}\mathop{}\!\mathrm{d}\mu(x)\mathop{}\!\mathrm{d}t\leq C.

Therefore, tu(m+1)/2Lloc2(n×(0,))\partial_{t}u^{(m+1)/2}\in L^{2}_{\text{loc}}(\mathbb{H}^{n}\times(0,\infty)). ∎

Proposition 6.

The weak solutions to equation (1.1) are indeed strong solutions. Moreover, tu(,t)12|m1|tu01\mathinner{\!\left\lVert\partial_{t}u(\cdot,t)\right\rVert}_{1}\leq\frac{2}{|m-1|t}\mathinner{\!\left\lVert u_{0}\right\rVert}_{1}, for m1m\neq 1 for all τ>0\tau>0.

Proof.

We begin with the case m1m\neq 1. From Lemma 5 and Lemma 6, tu\partial_{t}u is a bounded radon measure and tu(m+1)/2Lloc2(n×(0,))\partial_{t}u^{(m+1)/2}\in L^{2}_{\text{loc}}(\mathbb{H}^{n}\times(0,\infty)). Following Theorem 1.1 in [2], we obtain that tuLloc1(n×(0,))\partial_{t}u\in L^{1}_{\text{loc}}(\mathbb{H}^{n}\times(0,\infty)). Furthermore, we get tuL((τ,):L1(n))\partial_{t}u\in L^{\infty}((\tau,\infty)\mathrel{\mathop{\mathchar 58\relax}}L^{1}(\mathbb{H}^{n})) by estimate (4.1).

If m=1m=1, we obtain tuLloc1(n×(0,))\partial_{t}u\in L^{1}_{\text{loc}}(\mathbb{H}^{n}\times(0,\infty)) from Lemma 6. By L1L^{1} contractivity estimate, for every tτ>0t\geq\tau>0, we get

u(x,t+h)u(x,t)1hu(x,τ+h)u(x,τ)1hc(τ).\displaystyle\frac{\mathinner{\!\left\lVert u(x,t+h)-u(x,t)\right\rVert}_{1}}{h}\leq\frac{\mathinner{\!\left\lVert u(x,\tau+h)-u(x,\tau)\right\rVert}_{1}}{h}\leq c(\tau).

Therefore, we prove tuL((τ,):L1(n))\partial_{t}u\in L^{\infty}((\tau,\infty)\mathrel{\mathop{\mathchar 58\relax}}L^{1}(\mathbb{H}^{n})) in the linear case m=1m=1. ∎

Now it’s reasonable to use umu^{m} as a test function in the weak formulation of equation (1.1),

0tJ(um,um)ds+1m+1n|u|m+1(x,t)dμ(x)=1m+1n|u0|m+1dμ(x).\displaystyle\int_{0}^{t}\mathcal{E}_{J}(u^{m},u^{m})\mathop{}\!\mathrm{d}s+\frac{1}{m+1}\int_{\mathbb{H}^{n}}|u|^{m+1}(x,t)\mathop{}\!\mathrm{d}\mu(x)=\frac{1}{m+1}\int_{\mathbb{H}^{n}}|u_{0}|^{m+1}\mathop{}\!\mathrm{d}\mu(x). (4.3)

Then we have the following estimate:

Proposition 7.

If the initial data u0L1(n)L(n)u_{0}\in L^{1}(\mathbb{H}^{n})\cap L^{\infty}(\mathbb{H}^{n}), then the strong solution to equation (1.1) satisfies

0J(um,um)dt1m+1u0m+1m+1.\displaystyle\int_{0}^{\infty}\mathcal{E}_{J}(u^{m},u^{m})\mathop{}\!\mathrm{d}t\leq\frac{1}{m+1}\mathinner{\!\left\lVert u_{0}\right\rVert}_{m+1}^{m+1}. (4.4)

In fact, all the LpL^{p} norms are non-increasing as time goes by.

Proposition 8.

Any LpL^{p} norm, 1p1\leq p\leq\infty, of the strong solution to equation (1.1) with u0L1(n)L(n)u_{0}\in L^{1}(\mathbb{H}^{n})\cap L^{\infty}(\mathbb{H}^{n}) is non-increasing in time.

Proof.

The cases of p=1p=1 and p=p=\infty have been obtained from the contraction property in Proposition 3.

When 1<p<1<p<\infty, we multiply the equation (1.1) by up1u^{p-1}. By Stroock-Varopoulos inequality, we have

ddtn|u|p(x,t)dμ(x)\displaystyle\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}\int_{\mathbb{H}^{n}}|u|^{p}(x,t)\mathop{}\!\mathrm{d}\mu(x) pJ(um,up1)\displaystyle\leq-p\mathcal{E}_{J}(u^{m},u^{p-1}) (4.5)
CJ(um+p12,um+p12)0.\displaystyle\leq-C\mathcal{E}_{J}(u^{\frac{m+p-1}{2}},u^{\frac{m+p-1}{2}})\leq 0.

5 Existence and Uniqueness with General Initial Data

In order to prove the existence of solutions corresponding to less restrictive initial data, we first prove the smoothing effects to obtain that the solutions are uniformly bounded away from t=0t=0. These estimates will be used in the approximation process to obtain solutions with the general initial data. The smoothing effects are analogues of the results in [24].

Proposition 9.

Let 0<α<20<\alpha<2, m>0m>0, and take p>p(m)=max{1,(1m)Q/α}p>p^{*}(m)=\max\{1,(1-m)Q/\alpha\}. Then for every u0L1(n)L(n)u_{0}\in L^{1}(\mathbb{H}^{n})\cap L^{\infty}(\mathbb{H}^{n}), the solution to the Problem (1.1) satisfies

supxn|u(x,t)|C(m,p,α,Q,Λ)tγpu0pδp\displaystyle\sup\limits_{x\in\mathbb{H}^{n}}|u(x,t)|\leq C(m,p,\alpha,Q,\Lambda)t^{-\gamma_{p}}\mathinner{\!\left\lVert u_{0}\right\rVert}_{p}^{\delta_{p}} (5.1)

with γp=(m1+αp/Q)1\gamma_{p}=(m-1+\alpha p/Q)^{-1} and δp=αpγp/Q\delta_{p}=\alpha p\gamma_{p}/Q.

Proof.

The parabolic Moser iterative technique will be used to prove the estimates. Fix tt, consider the following sequence tk=(12k)tt_{k}=(1-2^{-k})t with 0k<+0\leq k<+\infty. Multiplying the equation (1.1) by upk1u^{p_{k}-1}, then integrate in n×[tk,tk+1]\mathbb{H}^{n}\times[t_{k},t_{k+1}]. Here pkp0>1p_{k}\geq p_{0}>1 which will be determined afterwards. Using the Stroock-Varopoulos inequality and the decay of LpL^{p} norms, we have

u(,tk)pkpk\displaystyle\mathinner{\!\left\lVert u(\cdot,t_{k})\right\rVert}_{p_{k}}^{p_{k}} C(Q,α,Λ)4mpk(pk1)(pk+m1)2tktk+1J(|u|pk+m12(,τ),|u|pk+m12(,τ))dτ\displaystyle\geq C(Q,\alpha,\Lambda)\frac{4mp_{k}(p_{k}-1)}{(p_{k}+m-1)^{2}}\int_{t_{k}}^{t_{k+1}}\mathcal{E}_{J}(|u|^{\frac{p_{k}+m-1}{2}}(\cdot,\tau),|u|^{\frac{p_{k}+m-1}{2}}(\cdot,\tau))\mathop{}\!\mathrm{d}\tau
1dku(,tk)pkpktktk+1u(,τ)pkpkJ(|u|pk+m12(,τ),|u|pk+m12(,τ))dτ.\displaystyle\geq\frac{1}{d_{k}\mathinner{\!\left\lVert u(\cdot,t_{k})\right\rVert}_{p_{k}}^{p_{k}}}\int_{t_{k}}^{t_{k+1}}\mathinner{\!\left\lVert u(\cdot,\tau)\right\rVert}_{p_{k}}^{p_{k}}\mathcal{E}_{J}(|u|^{\frac{p_{k}+m-1}{2}}(\cdot,\tau),|u|^{\frac{p_{k}+m-1}{2}}(\cdot,\tau))\mathop{}\!\mathrm{d}\tau.

Here dk=1C(Q,α,Λ)(pk+m1)24mpk(pk1)d_{k}=\frac{1}{C(Q,\alpha,\Lambda)}\cdot\frac{(p_{k}+m-1)^{2}}{4mp_{k}(p_{k}-1)}. By using the inequalities (2.4) and (2.5), we have

tktk+1u(,τ)pkpkJ(|u|pk+m12(,τ),|u|pk+m12(,τ))dτ\displaystyle\int_{t_{k}}^{t_{k+1}}\mathinner{\!\left\lVert u(\cdot,\tau)\right\rVert}_{p_{k}}^{p_{k}}\mathcal{E}_{J}(|u|^{\frac{p_{k}+m-1}{2}}(\cdot,\tau),|u|^{\frac{p_{k}+m-1}{2}}(\cdot,\tau))\mathop{}\!\mathrm{d}\tau Ctktk+1u(,τ)Q(2pk+m1)2Qα2pk+m1dτ\displaystyle\geq C\int_{t_{k}}^{t_{k+1}}\mathinner{\!\left\lVert u(\cdot,\tau)\right\rVert}_{\frac{Q(2p_{k}+m-1)}{2Q-\alpha}}^{2p_{k}+m-1}\mathop{}\!\mathrm{d}\tau
C2ktu(,tk+1)Q(2pk+m1)2Qα2pk+m1.\displaystyle\geq C2^{-k}t\mathinner{\!\left\lVert u(\cdot,t_{k+1})\right\rVert}_{\frac{Q(2p_{k}+m-1)}{2Q-\alpha}}^{2p_{k}+m-1}.

Therefore,

u(,tk+1)pk+1(C2kdkt1)s2pk+1u(,tk)pkspkpk+1,\displaystyle\mathinner{\!\left\lVert u(\cdot,t_{k+1})\right\rVert}_{p_{k+1}}\leq(C2^{k}d_{k}t^{-1})^{\frac{s}{2p_{k+1}}}\mathinner{\!\left\lVert u(\cdot,t_{k})\right\rVert}_{p_{k}}^{\frac{sp_{k}}{p_{k+1}}}, (5.2)

where pk+1=s(pk+m12)p_{k+1}=s(p_{k}+\frac{m-1}{2}) and s=2Q2Qα>1s=\frac{2Q}{2Q-\alpha}>1. Hereafter, {pk}\{p_{k}\} is an increasing sequence and limk+pk=+\lim\limits_{k\rightarrow+\infty}p_{k}=+\infty, we discover that

pk=A(sk1)+p0,A=p0(1m)Qα.\displaystyle p_{k}=A(s^{k}-1)+p_{0},\leavevmode\nobreak\ \leavevmode\nobreak\ A=p_{0}-\frac{(1-m)Q}{\alpha}.

To guarantee that A>0A>0, we impose the following condition on p0p_{0}:

p0(1m)Qα.\displaystyle p_{0}\geq\frac{(1-m)Q}{\alpha}.

Notice that if m1m\geq 1, we have

dk\displaystyle d_{k} =C(Q,α,Λ)(pk+m1)2mpk2pkpk1C(Q,α,Λ)4(m1)mpkpk1\displaystyle=C(Q,\alpha,\Lambda)\frac{(p_{k}+m-1)^{2}}{mp_{k}^{2}}\cdot\frac{p_{k}}{p_{k}-1}\leq C(Q,\alpha,\Lambda)\frac{4(m-1)}{m}\cdot\frac{p_{k}}{p_{k}-1}
C(Q,α,Λ)pkpk1C(Q,α,Λ)pkp01.\displaystyle\leq C(Q,\alpha,\Lambda)\cdot\frac{p_{k}}{p_{k}-1}\leq C(Q,\alpha,\Lambda)\frac{p_{k}}{p_{0}-1}.

If m<1m<1, we have

dk\displaystyle d_{k} =C(Q,α,Λ)(pk+m1)2mpk2pkpk1C(Q,α,Λ)αm(1m)Qαpk.\displaystyle=C(Q,\alpha,\Lambda)\frac{(p_{k}+m-1)^{2}}{mp_{k}^{2}}\cdot\frac{p_{k}}{p_{k}-1}\leq C(Q,\alpha,\Lambda)\frac{\alpha m}{(1-m)Q-\alpha}\cdot p_{k}.

Define Uk=u(,tk)pkU_{k}=\mathinner{\!\left\lVert u(\cdot,t_{k})\right\rVert}_{p_{k}}. From the iterative inequalities (5.2), we have

Uk+1Ckpk+1ts2pk+1Ukspkpk+1.\displaystyle U_{k+1}\leq C^{\frac{k}{p_{k+1}}}t^{-\frac{s}{2p_{k+1}}}U_{k}^{\frac{sp_{k}}{p_{k+1}}}.

Therefore, we obtain

UkCαktβkU0σk.\displaystyle U_{k}\leq C^{\alpha_{k}}t^{-\beta_{k}}U_{0}^{\sigma_{k}}.

Here

αk=1pkj=1k1(kj)sj,βk=12pkj=1ksj,σk=skp0pk.\displaystyle\alpha_{k}=\frac{1}{p_{k}}\sum_{j=1}^{k-1}(k-j)s^{j},\leavevmode\nobreak\ \leavevmode\nobreak\ \beta_{k}=\frac{1}{2p_{k}}\sum_{j=1}^{k}s^{j},\leavevmode\nobreak\ \leavevmode\nobreak\ \sigma_{k}=\frac{s^{k}p_{0}}{p_{k}}.

Let kk approach infinity, by the iteration process, we have

limk+αk=Q(Qα)α2A,limk+βk=QαA,limk+σk=p0A.\displaystyle\lim\limits_{k\rightarrow+\infty}\alpha_{k}=\frac{Q(Q-\alpha)}{\alpha^{2}A},\leavevmode\nobreak\ \leavevmode\nobreak\ \lim\limits_{k\rightarrow+\infty}\beta_{k}=\frac{Q}{\alpha A},\leavevmode\nobreak\ \leavevmode\nobreak\ \lim\limits_{k\rightarrow+\infty}\sigma_{k}=\frac{p_{0}}{A}.

Hereafter,

u(,t)\displaystyle\mathinner{\!\left\lVert u(\cdot,t)\right\rVert}_{\infty} =limk+u(,t12k)pk\displaystyle=\lim\limits_{k\rightarrow+\infty}\mathinner{\!\left\lVert u(\cdot,\frac{t}{1-2^{-k}})\right\rVert}_{p_{k}}
limk+C(Q,α,Λ)(t12k)QαAU0p0A\displaystyle\leq\lim\limits_{k\rightarrow+\infty}C(Q,\alpha,\Lambda)(\frac{t}{1-2^{-k}})^{-\frac{Q}{\alpha A}}U_{0}^{\frac{p_{0}}{A}}
C(Q,α,Λ)tQ(m1)Q+p0αfp0p0α(m1)Q+p0α.\displaystyle\leq C(Q,\alpha,\Lambda)t^{-\frac{Q}{(m-1)Q+p_{0}\alpha}}\mathinner{\!\left\lVert f\right\rVert}_{p_{0}}^{\frac{p_{0}\alpha}{(m-1)Q+p_{0}\alpha}}.

By the above Moser iteration technique, we obtain the desired smoothing effects estimates (5.1). ∎

Remark 5.

If mmm\leq m^{*}, the constant AA goes to zero when pp(m)p\rightarrow p*(m). Hereafter, the constants in the estimates (5.1) blow up as pp(m)p\rightarrow p*(m). If m>mm>m^{*} and p1p\rightarrow 1, the 1p1\frac{1}{p-1} of constant dkd_{k} in Proposition 9 blow up.

However, an iterative interpolation argument, see the analogues in [24], will give rise to the L1LL^{1}-L^{\infty} smoothing effect as m>mm>m_{*}.

Proposition 10.

Let 0<α<20<\alpha<2, m>mm>m_{*}. For every u0L1(n)L(n)u_{0}\in L^{1}(\mathbb{H}^{n})\cap L^{\infty}(\mathbb{H}^{n}), the solution to equation (1.1) satisfies

supxn|u(x,t)|C(m,α,Q)tγ1u01δ1\displaystyle\sup\limits_{x\in\mathbb{H}^{n}}|u(x,t)|\leq C(m,\alpha,Q)t^{-\gamma_{1}}\mathinner{\!\left\lVert u_{0}\right\rVert}_{1}^{\delta_{1}} (5.3)

with γ1=(m1+α/Q)1\gamma_{1}=(m-1+\alpha/Q)^{-1} and δ1=αγ1/Q\delta_{1}=\alpha\gamma_{1}/Q.

Proof.

For fixed tt, let τk=2kt\tau_{k}=2^{-k}t. Here 0k<+0\leq k<+\infty. Since p(m)<1p^{*}(m)<1 if m>mm>m^{*}, Applying the smoothing effects estimates (5.1) in Proposition 9 on the interval [τ1,τ0][\tau_{1},\tau_{0}] with p=2p=2, we have

u(,t)C(Q,α,Λ)(t2)γ2u(,τ1)1αγ2Qu(,τ1)αγ2Q.\displaystyle\mathinner{\!\left\lVert u(\cdot,t)\right\rVert}_{\infty}\leq C(Q,\alpha,\Lambda)(\frac{t}{2})^{-\gamma_{2}}\mathinner{\!\left\lVert u(\cdot,\tau_{1})\right\rVert}_{1}^{\frac{\alpha\gamma_{2}}{Q}}\mathinner{\!\left\lVert u(\cdot,\tau_{1})\right\rVert}_{\infty}^{\frac{\alpha\gamma_{2}}{Q}}.

Considering the iterative sequence τk\tau_{k}, applying the estimates (5.1) with p=2p=2 on the interval [τ2,τ1][\tau_{2},\tau_{1}], we have

u(,t)C(Q,α,Λ)(t2)γ2u(,τ1)1αγ2Q(C(Q,α,Λ)(t4)γ2u(,τ2)22αγ2Q)αγ2Q\displaystyle\mathinner{\!\left\lVert u(\cdot,t)\right\rVert}_{\infty}\leq C(Q,\alpha,\Lambda)(\frac{t}{2})^{-\gamma_{2}}\mathinner{\!\left\lVert u(\cdot,\tau_{1})\right\rVert}_{1}^{\frac{\alpha\gamma_{2}}{Q}}\cdot(C(Q,\alpha,\Lambda)(\frac{t}{4})^{-\gamma_{2}}\mathinner{\!\left\lVert u(\cdot,\tau_{2})\right\rVert}_{2}^{\frac{2\alpha\gamma_{2}}{Q}})^{\frac{\alpha\gamma_{2}}{Q}}

Following the iterative process, since u(,t)1u(,0)1\mathinner{\!\left\lVert u(\cdot,t)\right\rVert}_{1}\leq\mathinner{\!\left\lVert u(\cdot,0)\right\rVert}_{1} for each t0t\geq 0, we obtain

u(,t)C(Q,α,Λ)ak2bktcku01dku(,τk)2ek.\displaystyle\mathinner{\!\left\lVert u(\cdot,t)\right\rVert}_{\infty}\leq C(Q,\alpha,\Lambda)^{a_{k}}\cdot 2^{b_{k}}\cdot t^{-c_{k}}\cdot\mathinner{\!\left\lVert u_{0}\right\rVert}_{1}^{d_{k}}\cdot\mathinner{\!\left\lVert u(\cdot,\tau_{k})\right\rVert}_{2}^{e_{k}}.

Since m>mm>m^{*}, the constant γ2αQ<1\frac{\gamma_{2}\alpha}{Q}<1. Then,

limk+ak=limk+j=0k1(γ2αQ)j=(m1)Q+2α(m1)Q+α,\displaystyle\lim\limits_{k\rightarrow+\infty}a_{k}=\lim\limits_{k\rightarrow+\infty}\sum_{j=0}^{k-1}(\frac{\gamma_{2}\alpha}{Q})^{j}=\frac{(m-1)Q+2\alpha}{(m-1)Q+\alpha},
limk+bk=limk+j=0k1γ2(j+1)(γ2αQ)j=(m1)Q+2α((m1)Q+α)2,\displaystyle\lim\limits_{k\rightarrow+\infty}b_{k}=\lim\limits_{k\rightarrow+\infty}\sum_{j=0}^{k-1}\gamma_{2}(j+1)(\frac{\gamma_{2}\alpha}{Q})^{j}=\frac{(m-1)Q+2\alpha}{((m-1)Q+\alpha)^{2}},
limk+ck=limk+j=0k1γ2(γ2αQ)j=Q(m1)Q+α=γ1,\displaystyle\lim\limits_{k\rightarrow+\infty}c_{k}=\lim\limits_{k\rightarrow+\infty}\sum_{j=0}^{k-1}\gamma_{2}(\frac{\gamma_{2}\alpha}{Q})^{j}=\frac{Q}{(m-1)Q+\alpha}=\gamma_{1},
limk+dk=limk+j=1k1(γ2αQ)j=α(m1)Q+α=γ1αQ,\displaystyle\lim\limits_{k\rightarrow+\infty}d_{k}=\lim\limits_{k\rightarrow+\infty}\sum_{j=1}^{k-1}(\frac{\gamma_{2}\alpha}{Q})^{j}=\frac{\alpha}{(m-1)Q+\alpha}=\gamma_{1}\cdot\frac{\alpha}{Q},
limk+ek=limk+2(γ2αQ)k=0.\displaystyle\lim\limits_{k\rightarrow+\infty}e_{k}=\lim\limits_{k\rightarrow+\infty}2(\frac{\gamma_{2}\alpha}{Q})^{k}=0.

which are similar to the constants in [24]. Using the fact that u(,τk)2u02\mathinner{\!\left\lVert u(\cdot,\tau_{k})\right\rVert}_{2}\leq\mathinner{\!\left\lVert u_{0}\right\rVert}_{2}, we prove

supxn|u(x,t)|C(m,α,Q)tγ1u01γ1αQ\displaystyle\sup\limits_{x\in\mathbb{H}^{n}}|u(x,t)|\leq C(m,\alpha,Q)t^{-\gamma_{1}}\mathinner{\!\left\lVert u_{0}\right\rVert}_{1}^{\frac{\gamma_{1}\alpha}{Q}}

A generalization of the existence of the weak solutions to the less restrictive initial data, indeed which are strong with respect to time tt variable, will be proved at the end of this section.

If m>mm>m^{*}, constructing the approximating solutions corresponding to the approximating initial data {u0,k}L1(n)L(n)\{u_{0,k}\}\in L^{1}(\mathbb{H}^{n})\cap L^{\infty}(\mathbb{H}^{n}). Since we have already obtained the existence of the weak solutions to the bounded initial data, the point here is to employ the smoothing effects estimates (5.3) for m>mm>m^{*} to give a uniform L(n)L^{\infty}(\mathbb{H}^{n}) norm of the approximating solutions away t=0t=0. Hereafter, passing the weak formualtions to the limit, we prove the weak formulation holds for the general inital data u0L1(n)u_{0}\in L^{1}(\mathbb{H}^{n}).

If mmm\leq m^{*}, the smoothing effects estimates (5.2) only hold for p>max{1,(1m)Q/α}p>\max\{1,(1-m)Q/\alpha\}. Therefore, we obtain the existence of the weak solutions for each u0L1(n)Lp(n)u_{0}\in L^{1}(\mathbb{H}^{n})\cap L^{p}(\mathbb{H}^{n}) with suitable parameter pp.

Theorem 6.

Let 0<α<20<\alpha<2, m>0m>0. Assume u0L1(n)u_{0}\in L^{1}(\mathbb{H}^{n}) if m>mm>m^{*}; u0L1(n)Lp(n)u_{0}\in L^{1}(\mathbb{H}^{n})\cap L^{p}(\mathbb{H}^{n}) with p>max{1,(1m)Q/α}p>\max\{1,(1-m)Q/\alpha\} if mmm\leq m^{*}, there exists a strong solution to equation (1.1) with the initial data u0u_{0}.

Proof.

Let {u0,k}L1(n)L(n)\{u_{0,k}\}\subset L^{1}(\mathbb{H}^{n})\cap L^{\infty}(\mathbb{H}^{n}) be a sequence of functions converging to u0u_{0} in L1(n)Lp(n)L^{1}(\mathbb{H}^{n})\cap L^{p}(\mathbb{H}^{n}) satisfying

u0,kpu0p.\displaystyle\mathinner{\!\left\lVert u_{0,k}\right\rVert}_{p}\leq\mathinner{\!\left\lVert u_{0}\right\rVert}_{p}.

Denote by {uk}\{u_{k}\} the sequence of the solutions corresponding to the initial data u0,ku_{0,k}.

Applying the L1L^{1}-contraction property and Crandall-Ligget’s theorem, we obtain ukuu_{k}\rightarrow u in C([0,):L1(n))C([0,\infty)\mathrel{\mathop{\mathchar 58\relax}}L^{1}(\mathbb{H}^{n})). By the smoothing effects estimates (5.2) or (5.3), we have

supxn|uk(x,t)|C(m,p,α,Q,Λ)tγpu0pδpC(m,p,α,Q,Λ)tγp.\displaystyle\sup\limits_{x\in\mathbb{H}^{n}}|u_{k}(x,t)|\leq C(m,p,\alpha,Q,\Lambda)t^{-\gamma_{p}}\mathinner{\!\left\lVert u_{0}\right\rVert}_{p}^{\delta_{p}}\leq C(m,p,\alpha,Q,\Lambda)t^{-\gamma_{p}}.

Since {uk}\{u_{k}\} are strong solutions satisfying (3.10), the weak formulations implies

t1t2J(ukm,ukm)dt+1m+1n|uk|m+1(x,t2)dμ(x)=1m+1n|uk|m+1(x,t1)dμ(x).\displaystyle\int_{t_{1}}^{t_{2}}\mathcal{E}_{J}(u_{k}^{m},u_{k}^{m})\mathop{}\!\mathrm{d}t+\frac{1}{m+1}\int_{\mathbb{H}^{n}}|u_{k}|^{m+1}(x,t_{2})\mathop{}\!\mathrm{d}\mu(x)=\frac{1}{m+1}\int_{\mathbb{H}^{n}}|u_{k}|^{m+1}(x,t_{1})\mathop{}\!\mathrm{d}\mu(x).

Hereafter, the norm L2((τ,):˙(n)){L^{2}((\tau,\infty)\mathrel{\mathop{\mathchar 58\relax}}\dot{\mathcal{H}_{\mathcal{L}}}(\mathbb{H}^{n}))} of the solutions {uk}\{u_{k}\} are uniformly bounded which are independent of kk. By the Banach-Alaoglu theorem, {uk}\{u_{k}\} converge to uu in the weak- topology on Lloc2((0,):˙(n))L^{2}_{\rm{loc}}((0,\infty)\mathrel{\mathop{\mathchar 58\relax}}\dot{\mathcal{H}_{\mathcal{L}}}(\mathbb{H}^{n})). Notice that the following weak formulations of the solutions {uk}\{u_{k}\} hold,

0nuktζ0J(ukm,ζ)=0,foreachζCc2(n×(0,)).\displaystyle\int_{0}^{\infty}\int_{\mathbb{H}^{n}}u_{k}\partial_{t}\zeta-\int_{0}^{\infty}\mathcal{E}_{J}(u_{k}^{m},\zeta)=0,\leavevmode\nobreak\ \leavevmode\nobreak\ \rm{for\leavevmode\nobreak\ each}\leavevmode\nobreak\ \zeta\in C^{2}_{c}(\mathbb{H}^{n}\times(0,\infty)).

Passing to the limit, we obtain

0nutζ0J(um,ζ)=0,foreachζCc2(n×(0,)).\displaystyle\int_{0}^{\infty}\int_{\mathbb{H}^{n}}u\partial_{t}\zeta-\int_{0}^{\infty}\mathcal{E}_{J}(u^{m},\zeta)=0,\leavevmode\nobreak\ \leavevmode\nobreak\ \rm{for\leavevmode\nobreak\ each}\leavevmode\nobreak\ \zeta\in C^{2}_{c}(\mathbb{H}^{n}\times(0,\infty)).

Furthermore, we have

n|u(x,t)u0(x)|dμ(x)\displaystyle\int_{\mathbb{H}^{n}}|u(x,t)-u_{0}(x)|\mathop{}\!\mathrm{d}\mu(x) n|u(x,t)uk(x,t)|dμ(x)+n|uk(x,t)u0,k(x)|dμ(x)\displaystyle\leq\int_{\mathbb{H}^{n}}|u(x,t)-u_{k}(x,t)|\mathop{}\!\mathrm{d}\mu(x)+\int_{\mathbb{H}^{n}}|u_{k}(x,t)-u_{0,k}(x)|\mathop{}\!\mathrm{d}\mu(x)
+n|u0,k(x)u0(x)|dμ(x).\displaystyle+\int_{\mathbb{H}^{n}}|u_{0,k}(x)-u_{0}(x)|\mathop{}\!\mathrm{d}\mu(x).

Passing to the limit, we derive the initial condition.

u(x,0)=u0(x)almosteverywhere.\displaystyle u(x,0)=u_{0}(x)\leavevmode\nobreak\ \leavevmode\nobreak\ \rm{almost\leavevmode\nobreak\ everywhere}.

The above constructed weak solution uu is strong satisfying (3.10) by Proposition 6. ∎

6 Further Properties

6.1 Conservation of mass if m>mm>m^{*}

Proposition 11.

Let m>mm>m^{*}, 0<α<20<\alpha<2. JJ satisfies assumption (1.3). For each u0L1(n)u_{0}\in L^{1}(\mathbb{H}^{n}), the corresponding solution uu satisfies the property, conservation of mass: nu(x,t)dμ(x)=nu0(x,t)dμ(x)\int_{\mathbb{H}^{n}}u(x,t)\mathop{}\!\mathrm{d}\mu(x)=\int_{\mathbb{H}^{n}}u_{0}(x,t)\mathop{}\!\mathrm{d}\mu(x).

Proof.

First we consider u0L1(n)L(n)u_{0}\in L^{1}(\mathbb{H}^{n})\cap L^{\infty}(\mathbb{H}^{n}), then the corresponding solution uL1(n)L(n)u\in L^{1}(\mathbb{H}^{n})\cap L^{\infty}(\mathbb{H}^{n}). Choosing a non-negative and nonincreasing cutoff function ψ(s)\psi(s) such that ψ(s)=1\psi(s)=1 for 0s10\leq s\leq 1, ψ(s)=0\psi(s)=0 for s2s\geq 2. Define ψR(x)=ψ(|x|/R)\psi_{R}(x)=\psi(|x|/R). Employing ψR\psi_{R} as a test function in the weak formulation, we have

nu(x,t)ψR(x)dμ(x)nu(x,0)ψR(x)dμ(x)=0tnumψRdμ(x)dt.\int_{\mathbb{H}^{n}}u(x,t)\psi_{R}(x)\mathop{}\!\mathrm{d}\mu(x)-\int_{\mathbb{H}^{n}}u(x,0)\psi_{R}(x)\mathop{}\!\mathrm{d}\mu(x)=-\int_{0}^{t}\int_{\mathbb{H}^{n}}u^{m}\mathcal{L}\psi_{R}\mathop{}\!\mathrm{d}\mu(x)\mathop{}\!\mathrm{d}t.

By the scaling property of test functions, we have

ψRqCRα+Qq,forq1.\displaystyle\mathinner{\!\left\lVert\mathcal{L}\psi_{R}\right\rVert}_{q}\leq CR^{-\alpha+\frac{Q}{q}},\leavevmode\nobreak\ \leavevmode\nobreak\ \rm{for\leavevmode\nobreak\ }q\geq 1.

Applying Hölder’s inequality, we have

|nu(x,t)ψR(x)dμ(x)nu(x,0)ψR(x)dμ(x)|CtRα+Q(p1)p.|\int_{\mathbb{H}^{n}}u(x,t)\psi_{R}(x)\mathop{}\!\mathrm{d}\mu(x)-\int_{\mathbb{H}^{n}}u(x,0)\psi_{R}(x)\mathop{}\!\mathrm{d}\mu(x)|\leq CtR^{-\alpha+\frac{Q(p-1)}{p}}.

Here p=max{1,1m}p=\max\{1,\frac{1}{m}\}. Since m>mm>m^{*}, the exponent of RR is negative. Letting RR go to infinity, we derive the conservation of mass property.

If u0L1(n)u_{0}\in L^{1}(\mathbb{H}^{n}), let {u0,k}L1(n)L(n)\{u_{0,k}\}\subset L^{1}(\mathbb{H}^{n})\cap L^{\infty}(\mathbb{H}^{n}) be a sequence of functions converging to u0u_{0} in L1(n)L^{1}(\mathbb{H}^{n}). Denote by {uk}\{u_{k}\} the sequence of the solutions corresponding to the initial data {u0,k}\{u_{0,k}\}. The conservation of mass for bounded initial data gives us

nuk(x,t)dμ=nu0,k(x,t)dμ.\displaystyle\int_{\mathbb{H}^{n}}u_{k}(x,t)\mathop{}\!\mathrm{d}\mu=\int_{\mathbb{H}^{n}}u_{0,k}(x,t)\mathop{}\!\mathrm{d}\mu.

Passing to the limit, we get

nu(x,t)dμ=nu0(x,t)dμforeachu0L1(n).\displaystyle\int_{\mathbb{H}^{n}}u(x,t)\mathop{}\!\mathrm{d}\mu=\int_{\mathbb{H}^{n}}u_{0}(x,t)\mathop{}\!\mathrm{d}\mu\leavevmode\nobreak\ \leavevmode\nobreak\ \text{for}\leavevmode\nobreak\ \text{each}\leavevmode\nobreak\ u_{0}\in L^{1}(\mathbb{H}^{n}).

6.2 Conservation of mass if m=mm=m^{*}

If m=mm=m^{*}, the exponent α+Q(p1)p-\alpha+\frac{Q(p-1)}{p} in the proof of Proposition 11 is zero and we can not obtain the conservation of mass by the above proof. In order to overcome this difficulty, we split the integral ddtnu(x,t)ψR(x)dμ(x)\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}\int_{\mathbb{H}^{n}}u(x,t)\psi_{R}(x)\mathop{}\!\mathrm{d}\mu(x) into two parts and estimate each part separately.

Proposition 12.

Let m=mm=m^{*}, 0<α<20<\alpha<2. JJ satisfies assumption (1.3). For each u0L1(n)Lp(n)u_{0}\in L^{1}(\mathbb{H}^{n})\cap L^{p}(\mathbb{H}^{n}) with p>1p>1, the solution uu corresponding to the initial data u0u_{0} satisfies the property, conservation of mass: nu(x,t)dμ=nu0(x,t)dμ\int_{\mathbb{H}^{n}}u(x,t)\mathop{}\!\mathrm{d}\mu=\int_{\mathbb{H}^{n}}u_{0}(x,t)\mathop{}\!\mathrm{d}\mu.

Proof.

For each given δ>0\delta>0, choose R0R_{0} such that n|uχ{|x|R0}|δ\int_{\mathbb{H}^{n}}|u\cdot\chi_{\{|x|\geq R_{0}\}}|\leq\delta. Then uu can be rewritten as u=u1+u2u=u_{1}+u_{2} with

u1=uχ{|x|R0},u2=uχ{|x|R0}.\displaystyle u_{1}=u\cdot\chi_{\{|x|\leq R_{0}\}},\leavevmode\nobreak\ \leavevmode\nobreak\ u_{2}=u\cdot\chi_{\{|x|\geq R_{0}\}}.

Observe that um=u1m+u2mu^{m}=u_{1}^{m}+u_{2}^{m}. We begin with the bounded solution uu corresponding to the bounded initial data u0L1(n)L(n)u_{0}\in L^{1}(\mathbb{H}^{n})\cap L^{\infty}(\mathbb{H}^{n}).

Multiplying the same test function ψR()\psi_{R}(\cdot) as the proof of Proposition 11, we have

|ddtnu(x,t)ψR(x)dμ(x)|\displaystyle|\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}\int_{\mathbb{H}^{n}}u(x,t)\psi_{R}(x)\mathop{}\!\mathrm{d}\mu(x)| =|nu1mψRdμ(x)+nu2mψRdμ(x)|\displaystyle=|\int_{\mathbb{H}^{n}}u_{1}^{m}\mathcal{L}\psi_{R}\mathop{}\!\mathrm{d}\mu(x)+\int_{\mathbb{H}^{n}}u_{2}^{m}\mathcal{L}\psi_{R}\mathop{}\!\mathrm{d}\mu(x)|
|nu1mψRdμ(x)|+|nu2mψRdμ(x)|\displaystyle\leq|\int_{\mathbb{H}^{n}}u_{1}^{m}\mathcal{L}\psi_{R}\mathop{}\!\mathrm{d}\mu(x)|+|\int_{\mathbb{H}^{n}}u_{2}^{m}\mathcal{L}\psi_{R}\mathop{}\!\mathrm{d}\mu(x)|
:=I1+I2.\displaystyle\mathrel{\mathop{\mathchar 58\relax}}=I_{1}+I_{2}.

By Hölder’s inequality, we have

u1mru11m|{|x|R0}|1mrrCR0Q(1mr)rforevery 1r<1/m.\displaystyle\mathinner{\!\left\lVert u_{1}^{m}\right\rVert}_{r}\leq\mathinner{\!\left\lVert u_{1}\right\rVert}_{1}^{m}|\{|x|\leq R_{0}\}|^{\frac{1-mr}{r}}\leq CR_{0}^{\frac{Q(1-mr)}{r}}\leavevmode\nobreak\ \leavevmode\nobreak\ \text{for}\leavevmode\nobreak\ \text{every}\leavevmode\nobreak\ 1\leq r<1/m.

Hereafter,

I1u1mrψRrr1CR0Q(1mr)rRα+Q(r1)r.\displaystyle I_{1}\leq\mathinner{\!\left\lVert u_{1}^{m}\right\rVert}_{r}\mathinner{\!\left\lVert\mathcal{L}\psi_{R}\right\rVert}_{\frac{r}{r-1}}\leq CR_{0}^{\frac{Q(1-mr)}{r}}R^{-\alpha+\frac{Q(r-1)}{r}}.

Observe that Q(1mr)r=αQ(r1)r\frac{Q(1-mr)}{r}=\alpha-\frac{Q(r-1)}{r} when m=m=QαQm=m^{*}=\frac{Q-\alpha}{Q}, we obtain

I1C(R0/R)Q(1mr)r.\displaystyle I_{1}\leq C(R_{0}/R)^{\frac{Q(1-mr)}{r}}.

For fixed δ\delta and R0R_{0}, I1I_{1} goes to zero by letting RR go to infinity. Furthermore,

u2m1m=u21mδm.\displaystyle\mathinner{\!\left\lVert u_{2}^{m}\right\rVert}_{\frac{1}{m}}=\mathinner{\!\left\lVert u_{2}\right\rVert}_{1}^{m}\leq\delta^{m}.

Then we have,

I2u2m1mψR11mCδmR0=Cδm.\displaystyle I_{2}\leq\mathinner{\!\left\lVert u_{2}^{m}\right\rVert}_{\frac{1}{m}}\mathinner{\!\left\lVert\mathcal{L}\psi_{R}\right\rVert}_{\frac{1}{1-m}}\leq C\delta^{m}R^{0}=C\delta^{m}.

Letting RR\rightarrow\infty, we obtain I1+I2CδmI_{1}+I_{2}\leq C\delta^{m}. We conclude that

|ddtnu(x,t)ψR(x)dμ(x)|0.\displaystyle|\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}\int_{\mathbb{H}^{n}}u(x,t)\psi_{R}(x)\mathop{}\!\mathrm{d}\mu(x)|\rightarrow 0.

by letting δ0\delta\rightarrow 0. We obtain the conservation of mass with u0L1(n)L(n)u_{0}\in L^{1}(\mathbb{H}^{n})\cap L^{\infty}(\mathbb{H}^{n}) if m=mm=m^{*}.

If u0L1(n)Lp(n)u_{0}\in L^{1}(\mathbb{H}^{n})\cap L^{p}(\mathbb{H}^{n}) with arbitrary p>1p>1, let {u0,k}L1(n)L(n)\{u_{0,k}\}\subset L^{1}(\mathbb{H}^{n})\cap L^{\infty}(\mathbb{H}^{n}) be a sequence of functions converging to u0u_{0} in L1(n)Lp(n)L^{1}(\mathbb{H}^{n})\cap L^{p}(\mathbb{H}^{n}). Denote by {uk}\{u_{k}\} the sequence of the solutions corresponding to the initial data {u0,k}\{u_{0,k}\}. The conservation of mass for bounded initial data gives us

nuk(x,t)dμ=nu0,k(x,t)dμ.\displaystyle\int_{\mathbb{H}^{n}}u_{k}(x,t)\mathop{}\!\mathrm{d}\mu=\int_{\mathbb{H}^{n}}u_{0,k}(x,t)\mathop{}\!\mathrm{d}\mu.

Passing to the limit, we obtain

nu(x,t)dμ=nu0(x,t)dμforeachu0L1(n)Lp(n).\displaystyle\int_{\mathbb{H}^{n}}u(x,t)\mathop{}\!\mathrm{d}\mu=\int_{\mathbb{H}^{n}}u_{0}(x,t)\mathop{}\!\mathrm{d}\mu\leavevmode\nobreak\ \leavevmode\nobreak\ \text{for}\leavevmode\nobreak\ \text{each}\leavevmode\nobreak\ u_{0}\in L^{1}(\mathbb{H}^{n})\cap L^{p}(\mathbb{H}^{n}).

6.3 Extinction if m<mm<m^{*}

We have already proved that the mass of the solution does not change as tt\rightarrow\infty if mmm\geq m^{*}. On the other hand, for m<mm<m^{*}, there is a finite extinction time T>0T>0 such that u(,T)0u(\cdot,T)\equiv 0 almost everywhere in n\mathbb{H}^{n}, see an analogous result in [24].

Proposition 13.

Let m<mm<m^{*}, 0<α<20<\alpha<2, JJ satisfies assumption (1.3). For every u0L1(n)Lp(n)u_{0}\in L^{1}(\mathbb{H}^{n})\cap L^{p}(\mathbb{H}^{n}) with p>max{1,(1m)Q/α}p>\max\{1,(1-m)Q/\alpha\}, there is a finite time T>0T>0 of the corresponding solution uu such that u(,T)0u(\cdot,T)\equiv 0 almost everywhere in n\mathbb{H}^{n}.

Proof.

From inequality (2.3), (2.5) and (4.5), we have

ddtn|u|p(x,t)dμ(x)\displaystyle\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}\int_{\mathbb{H}^{n}}|u|^{p}(x,t)\mathop{}\!\mathrm{d}\mu(x) pJ(um,up1)\displaystyle\leq-p\mathcal{E}_{J}(u^{m},u^{p-1})
CJ(um+p12,um+p12)\displaystyle\leq-C\mathcal{E}_{J}(u^{\frac{m+p-1}{2}},u^{\frac{m+p-1}{2}})
Cum+p122QQα2=C(n|u|(p+m1)QQαdμ)QαQ.\displaystyle\leq-C\mathinner{\!\left\lVert u^{\frac{m+p-1}{2}}\right\rVert}_{\frac{2Q}{Q-\alpha}}^{2}=-C(\int_{\mathbb{H}^{n}}|u|^{\frac{(p+m-1)Q}{Q-\alpha}}\mathop{}\!\mathrm{d}\mu)^{\frac{Q-\alpha}{Q}}.

Observe that p=(p+m1)QQαp=\frac{(p+m-1)Q}{Q-\alpha} if p=(1m)Qαp=\frac{(1-m)Q}{\alpha}. Define

J(t)=u(,t)pp,\displaystyle J(t)=\mathinner{\!\left\lVert u(\cdot,t)\right\rVert}_{p}^{p},

Then,

J(t)+CJQαQ(t)0.\displaystyle J^{\prime}(t)+CJ^{\frac{Q-\alpha}{Q}}(t)\leq 0.

Hereafter, we derive the extinction of the solution in finite time since J(0)J(0) is finite. ∎

7 Cα regularity

In this section, we prove continuity results of the constructed solutions to the equation (1.1). Instead, the equation (1.1) can be rewritten as

tu1/m+u=0.\partial_{t}u^{1/m}+\mathcal{L}u=0.

In order to prove the continuity results, we will construct the iterated sequence of solutions with iterated nonlinearity functions in decreasing space-time cylinders. For instance, we will construct the following iterative functions:

uk(x,t)=u(Rkx,c0Rαkt)μkωk.\displaystyle u_{k}(x,t)=\frac{u(R^{-k}x,c_{0}R^{-\alpha k}t)-\mu_{k}}{\omega_{k}}.

Here the constants RR, μk\mu_{k}, and ωk\omega_{k} will be chosen in the subsequent proof. Moreover, if the sequence of functions {uk}\{u_{k}\} satisfies the specific consistent regularity, we can obtain the Hölder modulus continuity of the solution uu.

Observe that the function uku_{k} is a solution to the following equation with the nonlinearity function βk(s)\beta_{k}(s),

tβk(u)+u=0,\displaystyle\partial_{t}\beta_{k}(u)+\mathcal{L}u=0,
βk(s)=|ωks+μk|1m1(ωks+μk)ωkc0.\displaystyle\beta_{k}(s)=\frac{|\omega_{k}s+\mu_{k}|^{\frac{1}{m}-1}(\omega_{k}s+\mu_{k})}{\omega_{k}\cdot c_{0}}.

This naturally leads to the investigations of the following equations with the general nonlinearity functions β(s)\beta(s):

tβ(u)+u=0,\displaystyle\partial_{t}\beta(u)+\mathcal{L}u=0, (7.1)

Here,

β(s)=a|bs+c|1m1(bs+c)forsomea,b>0,s.\beta(s)=a|bs+c|^{\frac{1}{m}-1}(bs+c)\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \text{for}\leavevmode\nobreak\ \text{some}\leavevmode\nobreak\ a,b>0,\leavevmode\nobreak\ s\in\mathbb{R}.

We begin with an energy inequality about the weak solution to the equation (7.1) by the weak formulation. First of all, we use the following equalities frequently on the Heisenberg group n\mathbb{H}^{n}. These equalities are mentioned in Corollary 1.6 [15] using the polar coordinates on the nilpotent groups.

a|x|b|x|αQdx={C0α1(bαaα)ifα0,C0log(b/a)ifα=0.\int_{a\leq|x|\leq b}|x|^{\alpha-Q}\mathop{}\!\mathrm{d}x=\left\{\begin{array}[]{ll}C_{0}\alpha^{-1}(b^{\alpha}-a^{\alpha})\nobreak\leavevmode\nobreak\leavevmode&\text{if}\leavevmode\nobreak\ \alpha\neq 0,\\ C_{0}\log(b/a)\nobreak\leavevmode\nobreak\leavevmode&\text{if}\leavevmode\nobreak\ \alpha=0.\\ \end{array}\right. (7.2)

Here aa\in\mathbb{C} and 0<a<b<0<a<b<\infty.

To obtain the energy inequality, we regard ζ=(uψ)+\zeta=(u-\psi)_{+} as a test function in the weak formulation. Here ψ\psi is a nonnegative Lipschitz barrier function satisfying,

|x1y|>1|ψ(x)ψ(y)|J(x,y)dy<C<foreveryxn.\displaystyle\int_{|x^{-1}\cdot y|>1}|\psi(x)-\psi(y)|J(x,y)\mathop{}\!\mathrm{d}y<C<\infty\leavevmode\nobreak\ \leavevmode\nobreak\ \text{for}\leavevmode\nobreak\ \text{every}\leavevmode\nobreak\ x\in\mathbb{H}^{n}. (7.3)

Recall that we have obtained tβ(u)L((τ,):L1(n))\partial_{t}\beta(u)\in L^{\infty}((\tau,\infty)\mathrel{\mathop{\mathchar 58\relax}}L^{1}(\mathbb{H}^{n})) in Section 4. Hereafter, the function ζ\zeta is an admissible test function. Define the functional,

ψ(u)=0(uψ)+β(s+ψ)sds.\displaystyle\mathcal{B}_{\psi}(u)=\int_{0}^{(u-\psi)_{+}}\beta^{\prime}(s+\psi)s\mathop{}\!\mathrm{d}s. (7.4)

Therefore,

nψ(u(x,t))dx|t1t2+t1t2J(u,(uψ)+)(t)dt=0.\displaystyle\int_{\mathbb{H}^{n}}\mathcal{B}_{\psi}(u(x,t))\mathop{}\!\mathrm{d}x|^{t_{2}}_{t_{1}}+\int_{t_{1}}^{t_{2}}\mathcal{E}_{J}(u,(u-\psi)_{+})(t)\mathop{}\!\mathrm{d}t=0. (7.5)

In order to obtain a useful inverse Sobolev energy inequality, we estabish the next Lemma first to give an estimate of the functional ψ(u)\mathcal{B}_{\psi}(u).

Lemma 7.

If l=infψ0l=\inf\psi\geq 0 and M=sup{uψ}u<M=\sup_{\{u\geq\psi\}}u<\infty, the following inequality holds

Λ1(uψ)+2ψ(u)Λ2(uψ)+.\displaystyle\Lambda_{1}(u-\psi)_{+}^{2}\leq\mathcal{B}_{\psi}(u)\leq\Lambda_{2}(u-\psi)_{+}. (7.6)

Here Λ1=12inflsMβ(s)\Lambda_{1}=\frac{1}{2}\inf\limits_{l\leq s\leq M}\beta^{\prime}(s) and Λ2=β(M)β(l).\Lambda_{2}=\beta(M)-\beta(l).

Proof.

First, we give the lower bound of ψ(u)\mathcal{B}_{\psi}(u),

ψ(u)2Λ10(uψ)+sds=Λ1(uψ)+2.\displaystyle\mathcal{B}_{\psi}(u)\geq 2\Lambda_{1}\int_{0}^{(u-\psi)_{+}}s\mathop{}\!\mathrm{d}s=\Lambda_{1}(u-\psi)_{+}^{2}.

Since β(s)0\beta^{\prime}(s)\geq 0 for all ss\in\mathbb{R}, we have

ψ(u)(uψ)+0(uψ)+β(s+ψ)dsΛ2(uψ)+.\displaystyle\mathcal{B}_{\psi}(u)\leq(u-\psi)_{+}\cdot\int_{0}^{(u-\psi)_{+}}\beta^{\prime}(s+\psi)\mathop{}\!\mathrm{d}s\leq\Lambda_{2}(u-\psi)_{+}.

Now are ready to prove the inverse Sobolev energy inequality, see the similar results on n\mathbb{R}^{n} in \autocitesrtfp_clnfe_dpa. Recall that J(f,f)12\mathcal{E}_{J}(f,f)^{\frac{1}{2}} is equivalent to the homogeneous fractional Sobolev norm α/4f2\mathinner{\!\left\lVert\mathscr{L}^{\alpha/4}f\right\rVert}_{2}, we obtain the following energy estimates,

Lemma 8.

For any nonnegative Lipschitz barrier function ψ\psi satisfying (7.3). If uu is a weak solution to (7.1) satisfying l=infψ0l=\inf\psi\geq 0 and M=sup{uψ}u<M=\sup_{\{u\geq\psi\}}u<\infty, then

Λ1(uψ)+(,t2)22+C1\displaystyle\Lambda_{1}\mathinner{\!\left\lVert(u-\psi)_{+}(\cdot,t_{2})\right\rVert}_{2}^{2}+C_{1} t1t2α/4((uψ)+)(,t)22dt\displaystyle\int_{t_{1}}^{t_{2}}\mathinner{\!\left\lVert\mathscr{L}^{\alpha/4}((u-\psi)_{+})(\cdot,t)\right\rVert}_{2}^{2}\mathop{}\!\mathrm{d}t (7.7)
Λ2(uψ)+(,t1)1+C2t1t2z(t)dt,\displaystyle\leq\Lambda_{2}\mathinner{\!\left\lVert(u-\psi)_{+}(\cdot,t_{1})\right\rVert}_{1}+C_{2}\int_{t_{1}}^{t_{2}}z(t)\mathop{}\!\mathrm{d}t,

Where z(t)=(uψ)+(,t)1+𝟙{u(,t)>ψ()}1z(t)=\mathinner{\!\left\lVert(u-\psi)_{+}(\cdot,t)\right\rVert}_{1}+\mathinner{\!\left\lVert\mathbbm{1}_{\{u(\cdot,t)>\psi(\cdot)\}}\right\rVert}_{1}. Notice that C1C_{1} and C2C_{2} are positive numbers which are independent of the choice of β\beta.

Proof.

From equation (7.5), we have

nψ(u(x,t2))dμ(x)+t1t2J(u,(uψ)+)(t)dt=nψ(u(x,t1))dx.\int_{\mathbb{H}^{n}}\mathcal{B}_{\psi}(u(x,t_{2}))\mathop{}\!\mathrm{d}\mu(x)+\int_{t_{1}}^{t_{2}}\mathcal{E}_{J}(u,(u-\psi)_{+})(t)\mathop{}\!\mathrm{d}t=\int_{\mathbb{H}^{n}}\mathcal{B}_{\psi}(u(x,t_{1}))\mathop{}\!\mathrm{d}x.

To estimate the term t1t2J(u,(uψ)+)(t)dt\int_{t_{1}}^{t_{2}}\mathcal{E}_{J}(u,(u-\psi)_{+})(t)\mathop{}\!\mathrm{d}t, we split it into three parts,

t1t2J(u,(uψ)+)(t)dt=\displaystyle\int_{t_{1}}^{t_{2}}\mathcal{E}_{J}(u,(u-\psi)_{+})(t)\mathop{}\!\mathrm{d}t= t1t2J((uψ)+,(uψ)+)(t)dt\displaystyle\int_{t_{1}}^{t_{2}}\mathcal{E}_{J}((u-\psi)_{+},(u-\psi)_{+})(t)\mathop{}\!\mathrm{d}t
+t1t2J((uψ),(uψ)+)(t)dt\displaystyle+\int_{t_{1}}^{t_{2}}\mathcal{E}_{J}((u-\psi)_{-},(u-\psi)_{+})(t)\mathop{}\!\mathrm{d}t
+t1t2J(ψ,(uψ)+)(t)dt:=I1+I2+I3\displaystyle+\int_{t_{1}}^{t_{2}}\mathcal{E}_{J}(\psi,(u-\psi)_{+})(t)\mathop{}\!\mathrm{d}t\mathrel{\mathop{\mathchar 58\relax}}=I_{1}+I_{2}+I_{3}

Here I20I_{2}\geq 0. Furthermore, we have the following estimate of J(ψ,(uψ)+)(t)\mathcal{E}_{J}(\psi,(u-\psi)_{+})(t),

J(ψ,(uψ)+)(t)\displaystyle\mathcal{E}_{J}(\psi,(u-\psi)_{+})(t) =n|x1y|1[ψ(x)ψ(y)][(uψ)+(x)(uψ)+(y)]J(x,y)\displaystyle=\int_{\mathbb{H}^{n}}\int_{|x^{-1}\cdot y|\geq 1}[\psi(x)-\psi(y)][(u-\psi)_{+}(x)-(u-\psi)_{+}(y)]J(x,y)
+n|x1y|1[ψ(x)ψ(y)][(uψ)+(x)(uψ)+(y)]J(x,y)\displaystyle\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ +\int_{\mathbb{H}^{n}}\int_{|x^{-1}\cdot y|\leq 1}[\psi(x)-\psi(y)][(u-\psi)_{+}(x)-(u-\psi)_{+}(y)]J(x,y)
:=J1+J2\displaystyle\mathrel{\mathop{\mathchar 58\relax}}=J_{1}+J_{2}

Since the function ψ\psi satisfies the condition (7.3), we have

|J1|Cn(uψ)+(y)dμ(y)|J_{1}|\leq C\int_{\mathbb{H}^{n}}(u-\psi)_{+}(y)\mathop{}\!\mathrm{d}\mu(y)

Notice that |(uψ)+(x)(uψ)+(y)||(uψ)+(x)(uψ)+(y)|(𝟙{u(x,t)>ψ(x)}+𝟙{u(y,t)>ψ(y)})|(u-\psi)_{+}(x)-(u-\psi)_{+}(y)|\leq|(u-\psi)_{+}(x)-(u-\psi)_{+}(y)|\cdot(\mathbbm{1}_{\{u(x,t)>\psi(x)\}}+\mathbbm{1}_{\{u(y,t)>\psi(y)\}}), by Hölder inequality,

|J2|\displaystyle|J_{2}|\leq C(ϵ)n|x1y|1[ψ(x)ψ(y)]2J(x,y)𝟙{u(x,t)>ψ(x)}dμ(y)dμ(x)\displaystyle C(\epsilon)\int_{\mathbb{H}^{n}}\int_{|x^{-1}\cdot y|\leq 1}[\psi(x)-\psi(y)]^{2}J(x,y)\mathbbm{1}_{\{u(x,t)>\psi(x)\}}\mathop{}\!\mathrm{d}\mu(y)\mathop{}\!\mathrm{d}\mu(x)
+ϵJ((uψ)+,(uψ)+)\displaystyle+\epsilon\mathcal{E}_{J}((u-\psi)_{+},(u-\psi)_{+})

Since ψ\psi is a Lipschitz function and the equivalence (2.1), we obtain the desired inequality. ∎

Next, we prove the first De Giorgi type oscillation lemma: if u is mostly negative in time-space cylinder, the supreme goes down in the half cylinder. Denote the cylinder {|x|R,at0}\{|x|\leq R,-a\leq t\leq 0\} by ΓR,a\Gamma_{R,a}.

Lemma 9.

There is a constant δ(β)(0,1)\delta(\beta)\in(0,1) depending on the choice of nonlinearity β\beta such that, for any 0<a10<a\leq 1, if u:n×[2,0]u\mathrel{\mathop{\mathchar 58\relax}}\mathbb{H}^{n}\times[-2,0]\rightarrow\mathbb{R} is a weak solution to equation (7.1) satisfying

u(x,t)1+(|x|α/41)+inn×[2,0],\displaystyle u(x,t)\leq 1+(|x|^{\alpha/4}-1)_{+}\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \text{in}\leavevmode\nobreak\ \mathbb{H}^{n}\times[-2,0], (7.8)

and

|{u>0}Γ2,2a|Cδ(β)1+Qαa(1+Qα),\displaystyle|\{u>0\}\cap\Gamma_{2,2a}|\leq C\delta(\beta)^{1+\frac{Q}{\alpha}}a^{(1+\frac{Q}{\alpha})}, (7.9)

then

u(x,t)1/2if(x,t)Γ1,a.\displaystyle u(x,t)\leq 1/2\leavevmode\nobreak\ \leavevmode\nobreak\ \text{if}\leavevmode\nobreak\ (x,t)\in\Gamma_{1,a}. (7.10)

In the proof, we have an explicit bound of δ(β)\delta(\beta),

δ(β)=12inf1/4s2β(s)1+β(2)β(1/4).\displaystyle\delta(\beta)=\frac{1}{2}\frac{\inf\limits_{1/4\leq s\leq 2}\beta^{\prime}(s)}{1+\beta(2)-\beta(1/4)}. (7.11)
Proof.

We begin with establishing a nonlinear recurrence relation to the following energy quantity

Uk=supt[Tk,0]n(uψLk)+2(t,x)dμ(x)+Tk0α/4((uψLk)+)(,t)22dt\displaystyle U_{k}=\sup_{t\in[-T_{k},0]}\int_{\mathbb{H}^{n}}(u-\psi_{L_{k}})_{+}^{2}(t,x)\mathop{}\!\mathrm{d}\mu(x)+\int_{T_{k}}^{0}\mathinner{\!\left\lVert\mathscr{L}^{\alpha/4}((u-\psi_{L_{k}})_{+})(\cdot,t)\right\rVert}_{2}^{2}\mathop{}\!\mathrm{d}t (7.12)

where Lk=12(112k)L_{k}=\frac{1}{2}(1-\frac{1}{2^{k}}), ψLk=Lk+(|x|α/21)+\psi_{L_{k}}=L_{k}+(|x|^{\alpha/2}-1)_{+} and Tk=(112k)aT_{k}=(-1-\frac{1}{2^{k}})a. Using uk(t,x)u_{k}(t,x) to denote (uψLk)+(t,x)(u-\psi_{L_{k}})_{+}(t,x). Since uψLku\geq\psi_{L_{k}}, we take M=2M=2 in Lemma 7. Furthermore, due to the fact that Lk1/4L_{k}\geq 1/4, we take l=1/4l=1/4 in Lemma 7. For Tk1σTkt0T_{k-1}\leq\sigma\leq T_{k}\leq t\leq 0, by the energy inequality (7.7) in Lemma 8, we have

Λ1(uψk)+(,t)22+C1\displaystyle\Lambda_{1}\mathinner{\!\left\lVert(u-\psi_{k})_{+}(\cdot,t)\right\rVert}_{2}^{2}+C_{1} σtα/4((uψk)+)(,s)22ds\displaystyle\int_{\sigma}^{t}\mathinner{\!\left\lVert\mathscr{L}^{\alpha/4}((u-\psi_{k})_{+})(\cdot,s)\right\rVert}_{2}^{2}\mathop{}\!\mathrm{d}s
Λ2(uψk)+(,σ)1+C2σtz(s)ds.\displaystyle\leq\Lambda_{2}\mathinner{\!\left\lVert(u-\psi_{k})_{+}(\cdot,\sigma)\right\rVert}_{1}+C_{2}\int_{\sigma}^{t}z(s)\mathop{}\!\mathrm{d}s.

By taking the average of σ[Tk1,Tk]\sigma\in[T_{k-1},T_{k}], and taking the supreme over t[Tk,0]t\in[T_{k},0], we obtain

UkC2k1+Λ2Λ1aTk10z(s)ds.\displaystyle U_{k}\leq C2^{k}\frac{1+\Lambda_{2}}{\Lambda_{1}\cdot a}\int_{T_{k-1}}^{0}z(s)\mathop{}\!\mathrm{d}s.

Using the inequality (2.4) and Interpolation inequality, we get

ukL2(1+αQ)(n×[Tk,0])CUk12.\displaystyle\mathinner{\!\left\lVert u_{k}\right\rVert}_{L^{2(1+\frac{\alpha}{Q})}(\mathbb{H}^{n}\times[T_{k},0])}\leq CU_{k}^{\frac{1}{2}}.

Now we deal with the controlling term Tk10z(s)ds\int_{T_{k-1}}^{0}z(s)\mathop{}\!\mathrm{d}s. By Tchebychev inequality, we have

Tk10z(s)ds\displaystyle\int_{T_{k-1}}^{0}z(s)\mathop{}\!\mathrm{d}s n×[Tk1,0]uk1χ{uk1>12k+1}+n×[Tk1,0]χ{uk1>12k+1}\displaystyle\leq\int\int_{\mathbb{H}^{n}\times[T_{k-1},0]}u_{k-1}\chi_{\{u_{k-1}>\frac{1}{2^{k+1}}\}}+\int\int_{\mathbb{H}^{n}\times[T_{k-1},0]}\chi_{\{u_{k-1}>\frac{1}{2^{k+1}}\}}
C(22+2αQ)kUk11+αQ.\displaystyle\leq C(2^{2+\frac{2\alpha}{Q}})^{k}U_{k-1}^{1+\frac{\alpha}{Q}}.

Therefore, we obtain

Uk1+Λ2Λ1a(CQ,Λ,α)kUk11+αQ.\displaystyle U_{k}\leq\frac{1+\Lambda_{2}}{\Lambda_{1}\cdot a}(C_{Q,\Lambda,\alpha})^{k}U_{k-1}^{1+\frac{\alpha}{Q}}.

If (1+Λ2Λ1a)Q/αU1ϵ0(Q,Λ,α)(\frac{1+\Lambda_{2}}{\Lambda_{1}\cdot a})^{Q/\alpha}U_{1}\leq\epsilon_{0}(Q,\Lambda,\alpha), then limkUk=0\lim\limits_{k\rightarrow\infty}U_{k}=0 which gives u(x,t)1/2if(x,t)Γ1,au(x,t)\leq 1/2\leavevmode\nobreak\ \leavevmode\nobreak\ \text{if}\leavevmode\nobreak\ (x,t)\in\Gamma_{1,a}. By Tchebychev inequality, we get

U1CQ,Λ,α1+Λ2Λ1an×[2a,0](u(|x|α/21)+)+.\displaystyle U_{1}\leq C_{Q,\Lambda,\alpha}\frac{1+\Lambda_{2}}{\Lambda_{1}\cdot a}\int\int_{\mathbb{H}^{n}\times[-2a,0]}(u-(|x|^{\alpha/2}-1)_{+})_{+}.

To guarantee that (1+Λ2Λ1a)Q/αU1ϵ0(Q,Λ,α)(\frac{1+\Lambda_{2}}{\Lambda_{1}\cdot a})^{Q/\alpha}U_{1}\leq\epsilon_{0}(Q,\Lambda,\alpha), we only need to impose the following condition,

n×[2a,0](u(|x|α/21)+)+ϵ0(Q,Λ,α)(Λ1a1+Λ2)1+Qα.\displaystyle\int\int_{\mathbb{H}^{n}\times[-2a,0]}(u-(|x|^{\alpha/2}-1)_{+})_{+}\leq\epsilon_{0}(Q,\Lambda,\alpha)(\frac{\Lambda_{1}\cdot a}{1+\Lambda_{2}})^{1+\frac{Q}{\alpha}}. (7.13)

The condition (7.13) can be connected to the condition (7.9) in the Lemma 9 through a scaling argument. For each point (x0,t0)(x_{0},t_{0}) in Γ1,a\Gamma_{1,a}, considering the following function uR(x,t)u_{R}(x,t),

uR(x,t)=u(x0R1x,t0+Rαt),t[2,0]byrequiringRα>2.\displaystyle u_{R}(x,t)=u(x_{0}\cdot R^{-1}x,t_{0}+R^{-\alpha}t),\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ t\in[-2,0]\leavevmode\nobreak\ \text{by}\leavevmode\nobreak\ \text{requiring}\leavevmode\nobreak\ R^{\alpha}>2.

which is a solution to equation (7.1) corresponding to the rescaled kernel

JR(x,y)=R(Q+α)J(x0R1x,x01R1x)\displaystyle J_{R}(x,y)=R^{-(Q+\alpha)}J(x_{0}\cdot R^{-1}x,x_{0}^{-1}\cdot R^{-1}x)

We claim that the rescaled solution uRu_{R} satisfies (7.13). Notice that there exists large enough RR such that 1+((|R1x|+1)α/41)+(|x|α/21)+1+((|R^{-1}x|+1)^{\alpha/4}-1)_{+}\leq(|x|^{\alpha/2}-1)_{+} for |x|R|x|\geq R. Therefore,

n×[2a,0](uR(|x|α/21)+)+\displaystyle\int\int_{\mathbb{H}^{n}\times[-2a,0]}(u_{R}-(|x|^{\alpha/2}-1)_{+})_{+} =2a0BR(0)(uR)+\displaystyle=\int_{-2a}^{0}\int_{B_{R}(0)}{(u_{R})}_{+}
=RQ+αt02aRαt0B1(x0)u(x,t)+dμ(x)dt.\displaystyle=R^{Q+\alpha}\int_{t_{0}-2aR^{-\alpha}}^{t_{0}}\int_{B_{1}(x_{0})}u(x,t)_{+}\mathop{}\!\mathrm{d}\mu(x)\mathop{}\!\mathrm{d}t.

Observe that B1(x0)×[t02aRα,t0]Γ2,2aB_{1}(x_{0})\times[t_{0}-2aR^{-\alpha},t_{0}]\subset\Gamma_{2,2a} if Rα>2R^{\alpha}>2 and (x0,t0)Γ1,a(x_{0},t_{0})\in\Gamma_{1,a}. Furthermore, u(x,t)+2α/4u(x,t)_{+}\leq 2^{\alpha/4} where (x,t)Γ2,a(x,t)\in\Gamma_{2,a}. Hereafter,

n×[2a,0](uR(|x|α/21)+)+\displaystyle\int\int_{\mathbb{H}^{n}\times[-2a,0]}(u_{R}-(|x|^{\alpha/2}-1)_{+})_{+} C2α/4RQ+α|{u>0}Γ2,2a|\displaystyle\leq C2^{\alpha/4}R^{Q+\alpha}|\{u>0\}\cap\Gamma_{2,2a}|
ϵ0(Q,Λ,α)(Λ1a1+Λ2)1+Qα.\displaystyle\leq\epsilon_{0}(Q,\Lambda,\alpha)(\frac{\Lambda_{1}\cdot a}{1+\Lambda_{2}})^{1+\frac{Q}{\alpha}}.

by requiring

|{u>0}Γ2,2a|ϵ1(Q,Λ,α)(Λ1a1+Λ2)1+Qα=c(Q,Λ,α)δ(β)1+Qαa1+Qα\displaystyle|\{u>0\}\cap\Gamma_{2,2a}|\leq\epsilon_{1}(Q,\Lambda,\alpha)(\frac{\Lambda_{1}\cdot a}{1+\Lambda_{2}})^{1+\frac{Q}{\alpha}}=c(Q,\Lambda,\alpha)\delta(\beta)^{1+\frac{Q}{\alpha}}\cdot a^{1+\frac{Q}{\alpha}}

where δ(β)=Λ11+Λ2=12inf1/4s2β(s)1+β(2)β(1/4)\delta(\beta)=\frac{\Lambda_{1}}{1+\Lambda_{2}}=\frac{\frac{1}{2}\inf\limits_{1/4\leq s\leq 2}\beta^{\prime}(s)}{1+\beta(2)-\beta(1/4)} is given in the proof. Therefore, we obtain u(x0,t0)1u(x_{0},t_{0})\leq 1 for each (x0,t0)Γ1,a(x_{0},t_{0})\in\Gamma_{1,a}. ∎

Remark 6.

If u is mostly positive in the cylinder Γ2,2α\Gamma_{2,2^{\alpha}}, applying Lemma 9 to u-u with β~(s)=β(s)\tilde{\beta}(s)=-\beta(-s), we obtain that the infimum has a upper bound in Γ1,1\Gamma_{1,1} with the explicit bound δ(β~)\delta(\tilde{\beta}).

Remark 7.

We observe that the constant δ(β)\delta(\beta) depends on the choice of β\beta. For instance, if inf1/4s2β(s)0\inf\limits_{1/4\leq s\leq 2}\beta^{\prime}(s)\rightarrow 0, then δ(β)0\delta(\beta)\rightarrow 0. Hereafter, the condition (7.9) becomes harder to achieve.

The next step is to prove second De Giorgi Lemma which says some mass is lost between two level sets. Define

ψλ(x)=((|x|λ4/α)+α/41)+,λ(0,13),xn.\displaystyle\psi_{\lambda}(x)=((|x|-\lambda^{-4/\alpha})_{+}^{\alpha/4}-1)_{+},\leavevmode\nobreak\ \leavevmode\nobreak\ \lambda\in(0,\frac{1}{3}),\leavevmode\nobreak\ x\in\mathbb{H}^{n}. (7.14)

which can be used to control the growth at infinity.

F(x)=sup(1,inf(0,|x|29)),xn.\displaystyle F(x)=\sup(-1,\inf(0,|x|^{2}-9)),\leavevmode\nobreak\ \leavevmode\nobreak\ x\in\mathbb{H}^{n}. (7.15)

which can be used to localize the problem in the ball B3B_{3}. We follow the ideas in [5] to obtain the following lemma,

Lemma 10.

Assume C1β(s)C2C_{1}\leq\beta^{\prime}(s)\leq C_{2} for every 1/2s2{1}/{2}\leq s\leq 2. Assume 0<a<10<a<1. For each very small μ,ν>0\mu,\nu>0, there exists λ¯(0,13)\bar{\lambda}\in(0,\frac{1}{3}) depending on μ,ν,α,a,Q,C1,C2\mu,\nu,\alpha,a,Q,C_{1},C_{2} such that for any λ(0,λ¯)\lambda\in(0,\bar{\lambda}) and solution u:n×[2,0]u\mathrel{\mathop{\mathchar 58\relax}}\mathbb{H}^{n}\times[-2,0]\rightarrow\mathbb{R} to equation (7.1) satisfying

u(x,t)1+ψλ(x)inn×[2,0]and|{u<0}(B2×(2,2a))|>μ,\displaystyle u(x,t)\leq 1+\psi_{\lambda}(x)\leavevmode\nobreak\ \leavevmode\nobreak\ \text{in}\leavevmode\nobreak\ \mathbb{H}^{n}\times[-2,0]\leavevmode\nobreak\ \leavevmode\nobreak\ \text{and}\leavevmode\nobreak\ \leavevmode\nobreak\ |\{u<0\}\cap(B_{2}\times(-2,-2a))|>\mu, (7.16)

We have the following implication: If

|{u>1+λ2F}(B3×(2a,0))|ν,\displaystyle|\{u>1+\lambda^{2}F\}\cap(B_{3}\times(-2a,0))|\geq\nu, (7.17)

then

|{1+F<u<1+λ2F}(B3×(2,0))|γ(μ,ν,α,a,Q,C1,C2).\displaystyle|\{1+F<u<1+\lambda^{2}F\}\cap(B_{3}\times(-2,0))|\geq\gamma(\mu,\nu,\alpha,a,Q,C_{1},C_{2}). (7.18)
Proof.

Define

ϕ0(x)=1+ψλ(x)+F(x),\displaystyle\phi_{0}(x)=1+\psi_{\lambda}(x)+F(x),
ϕ1(x)=1+ψλ(x)+λF(x),\displaystyle\phi_{1}(x)=1+\psi_{\lambda}(x)+\lambda F(x),
ϕ2(x)=1+ψλ(x)+λ2F(x).\displaystyle\phi_{2}(x)=1+\psi_{\lambda}(x)+\lambda^{2}F(x).

Notice that the barrier function ϕ1\phi_{1} is an intermediate state between ϕ0\phi_{0} and ϕ2\phi_{2}, we expect these barrier functions to provide a quantitative picture of the loss of mass between the level sets.

Energy inequality: Recall the equation (7.5) and regard the barrier function ϕ1\phi_{1} as an admissible test function, for 2t1t20-2\leq t_{1}\leq t_{2}\leq 0, we have

nϕ1(u(x,t2))dμ(x)|t1t2+t1t2J(u,(uϕ1)+)(t)dt=0.\displaystyle\int_{\mathbb{H}^{n}}\mathcal{B}_{\phi_{1}}(u(x,t_{2}))\mathop{}\!\mathrm{d}\mu(x)\Big{|}_{t_{1}}^{t_{2}}+\int_{t_{1}}^{t_{2}}\mathcal{E}_{J}(u,(u-\phi_{1})_{+})(t)\mathop{}\!\mathrm{d}t=0.

Hereafter,

nϕ1(u(x,t2))\displaystyle\int_{\mathbb{H}^{n}}\mathcal{B}_{\phi_{1}}(u(x,t_{2})) dμ(x)|t1t2+t1t2J((uϕ1)+,(uϕ1)+)(t)dt\displaystyle\mathop{}\!\mathrm{d}\mu(x)\Big{|}_{t_{1}}^{t_{2}}+\int_{t_{1}}^{t_{2}}\mathcal{E}_{J}((u-\phi_{1})_{+},(u-\phi_{1})_{+})(t)\mathop{}\!\mathrm{d}t (7.19)
+t1t2J((uϕ1),(uϕ1)+)(t)dt=t1t2J(ϕ1,(uϕ1)+)(t)dt.\displaystyle+\int_{t_{1}}^{t_{2}}\mathcal{E}_{J}((u-\phi_{1})_{-},(u-\phi_{1})_{+})(t)\mathop{}\!\mathrm{d}t=-\int_{t_{1}}^{t_{2}}\mathcal{E}_{J}(\phi_{1},(u-\phi_{1})_{+})(t)\mathop{}\!\mathrm{d}t.

In the proof of the inverse Sobolev energy inequality in Lemma 8, we neglect the nonnegative term t1t2J((uϕ1),(uϕ1)+)(t)dt\int_{t_{1}}^{t_{2}}\mathcal{E}_{J}((u-\phi_{1})_{-},(u-\phi_{1})_{+})(t)\mathop{}\!\mathrm{d}t. However, controlling this term is important in the following proof. Hence, we keep this term in the above equation.

Now we estimate the right hand side of the above equation. Observe that u(x)>ϕ1u(x)>\phi_{1} implies xB3x\in B_{3}. we have

|t1t2J(ϕ1,(uϕ1)+)(t)dt|\displaystyle|\int_{t_{1}}^{t_{2}}\mathcal{E}_{J}(\phi_{1},(u-\phi_{1})_{+})(t)\mathop{}\!\mathrm{d}t|\leq 12t1t2J((uϕ1)+,(uϕ1)+)(t)dt\displaystyle\frac{1}{2}\int_{t_{1}}^{t_{2}}\mathcal{E}_{J}((u-\phi_{1})_{+},(u-\phi_{1})_{+})(t)\mathop{}\!\mathrm{d}t
+2t1t2B3n[ϕ1(x)ϕ1(y)]2J(x,y)dμ(y)dμ(x)dt\displaystyle+2\int_{t_{1}}^{t_{2}}\int_{B_{3}}\int_{\mathbb{H}^{n}}[\phi_{1}(x)-\phi_{1}(y)]^{2}J(x,y)\mathop{}\!\mathrm{d}\mu(y)\mathop{}\!\mathrm{d}\mu(x)\mathop{}\!\mathrm{d}t
:=J1+J2.\displaystyle\mathrel{\mathop{\mathchar 58\relax}}=J_{1}+J_{2}.

The first term J1J_{1} can be absorbed into the left hand side of the equation (7.19). Due to the definition of function ϕ1\phi_{1}, we control J2J_{2} in terms of ψλ(x)\psi_{\lambda}(x) and F(x)F(x) respectively.

J2\displaystyle J_{2} 4λ2t1t2nn[F(x)F(y)]2J(x,y)dμ(x)dμ(y)dt\displaystyle\leq 4\lambda^{2}\int_{t_{1}}^{t_{2}}\int_{\mathbb{H}^{n}}\int_{\mathbb{H}^{n}}[F(x)-F(y)]^{2}J(x,y)\mathop{}\!\mathrm{d}\mu(x)\mathop{}\!\mathrm{d}\mu(y)\mathop{}\!\mathrm{d}t
+4t1t2B3n[ψλ(x)ψλ(y)]2J(x,y)dμ(x)dμ(y)dt:=L1+L2.\displaystyle+4\int_{t_{1}}^{t_{2}}\int_{B_{3}}\int_{\mathbb{H}^{n}}[\psi_{\lambda}(x)-\psi_{\lambda}(y)]^{2}J(x,y)\mathop{}\!\mathrm{d}\mu(x)\mathop{}\!\mathrm{d}\mu(y)\mathop{}\!\mathrm{d}t\mathrel{\mathop{\mathchar 58\relax}}=L_{1}+L_{2}.

Since FF is a Lipschitz function with compact support, we have L1Cλ2L_{1}\leq C\lambda^{2}. Furthermore, notice that ψλ(x)=0\psi_{\lambda}(x)=0 for xB3x\in B_{3} by requiring 0<λ<1/30<\lambda<1/3, we get

L2\displaystyle L_{2} =4t1t2B3nψλ(y)2J(x,y)dμ(x)dμ(y)dt\displaystyle=4\int_{t_{1}}^{t_{2}}\int_{B_{3}}\int_{\mathbb{H}^{n}}\psi_{\lambda}(y)^{2}J(x,y)\mathop{}\!\mathrm{d}\mu(x)\mathop{}\!\mathrm{d}\mu(y)\mathop{}\!\mathrm{d}t
4Λ|B3|t1t2{|y|>λ4/α}((|y|λ4/α)+α/41)+2(|y|3)Q+αdμ(y)dt\displaystyle\leq 4\Lambda|B_{3}|\int_{t_{1}}^{t_{2}}\int_{\{|y|>\lambda^{-4/\alpha}\}}\frac{((|y|-\lambda^{-4/\alpha})_{+}^{\alpha/4}-1)_{+}^{2}}{(|y|-3)^{Q+\alpha}}\mathop{}\!\mathrm{d}\mu(y)\mathop{}\!\mathrm{d}t
4Λ|B3|λ2t1t2{|z|>1}((|z|1)+α/4λ)+2(|z|3λ4/α)Q+αdμ(z)dt\displaystyle\leq 4\Lambda|B_{3}|\lambda^{2}\int_{t_{1}}^{t_{2}}\int_{\{|z|>1\}}\frac{((|z|-1)_{+}^{\alpha/4}-\lambda)_{+}^{2}}{(|z|-3\lambda^{4/\alpha})^{Q+\alpha}}\mathop{}\!\mathrm{d}\mu(z)\mathop{}\!\mathrm{d}t
Cλ2.\displaystyle\leq C\lambda^{2}.

Therefore, we obtained

nϕ1(u(x,t2))dμ(x)|t1t2\displaystyle\int_{\mathbb{H}^{n}}\mathcal{B}_{\phi_{1}}(u(x,t_{2}))\mathop{}\!\mathrm{d}\mu(x)\Big{|}_{t_{1}}^{t_{2}} +12t1t2J((uϕ1)+,(uϕ1)+)(t)dt\displaystyle+\frac{1}{2}\int_{t_{1}}^{t_{2}}\mathcal{E}_{J}((u-\phi_{1})_{+},(u-\phi_{1})_{+})(t)\mathop{}\!\mathrm{d}t
+t1t2J((uϕ1),(uϕ1)+)(t)dtCλ2.\displaystyle+\int_{t_{1}}^{t_{2}}\mathcal{E}_{J}((u-\phi_{1})_{-},(u-\phi_{1})_{+})(t)\mathop{}\!\mathrm{d}t\leq C\lambda^{2}.

Define H(t):=nϕ1(u(x,t))dμ(x)H(t)\mathrel{\mathop{\mathchar 58\relax}}=\int_{\mathbb{H}^{n}}\mathcal{B}_{\phi_{1}}(u(x,t))\mathop{}\!\mathrm{d}\mu(x), we have

H(t)Cλ2.\displaystyle H^{\prime}(t)\leq C\lambda^{2}.

Meanwhile, observe that ϕ11λ1/2\phi_{1}\geq 1-\lambda\geq 1/2 and we choose l=1/2l=1/2 in Lemma 7. Also, if u(x)>ϕ1u(x)>\phi_{1}, we get xB3x\in B_{3} and u(x)2u(x)\leq 2. We choose M=2M=2 in Lemma 7. Furthermore, we have (uϕ1)+λ𝟙B3(u-\phi_{1})_{+}\leq\lambda\mathbbm{1}_{B_{3}}. Hereafter, since C1β(s)C2C_{1}\leq\beta^{\prime}(s)\leq C_{2} for every 12s2\frac{1}{2}\leq s\leq 2, we obtain

H(t)C2n(uϕ1)+2dμ(x)CC2λ2.\displaystyle H(t)\leq C_{2}\int_{\mathbb{H}^{n}}(u-\phi_{1})_{+}^{2}\mathop{}\!\mathrm{d}\mu(x)\leq CC_{2}\lambda^{2}.

Hence, we obtain

t1t2J((uϕ1)+,(uϕ1)+)(t)dtCC2λ2,\displaystyle\int_{t_{1}}^{t_{2}}\mathcal{E}_{J}((u-\phi_{1})_{+},(u-\phi_{1})_{+})(t)\mathop{}\!\mathrm{d}t\leq CC_{2}\lambda^{2},

and

t1t2J((uϕ1),(uϕ1)+)(t)dtCC2λ2.\displaystyle\int_{t_{1}}^{t_{2}}\mathcal{E}_{J}((u-\phi_{1})_{-},(u-\phi_{1})_{+})(t)\mathop{}\!\mathrm{d}t\leq CC_{2}\lambda^{2}.

An estimate of n(uϕ𝟏)+𝟐(t,x)𝐝μ(x)\int_{\mathbb{H}^{n}}(u-\phi_{1})_{+}^{2}(t,x)\mathop{}\!\mathrm{d}\mu(x):

Notice that ϕ0(x)=0\phi_{0}(x)=0 for xB2x\in B_{2}. The condition (7.17) can be rewritten as |{u<ϕ0}(B2×(2,2a))|>μ|\{u<\phi_{0}\}\cap(B_{2}\times(-2,-2a))|>\mu. Define

Σ:={t[2,2a]:|{u(,t)<ϕ0}B2|μ/4}.\displaystyle\Sigma\mathrel{\mathop{\mathchar 58\relax}}=\{t\in[-2,-2a]\mathrel{\mathop{\mathchar 58\relax}}|\{u(\cdot,t)<\phi_{0}\}\cap B_{2}|\geq\mu/4\}.

Then,

μ<|{u<ϕ0}(B2×(2,2a))|μ/4(22a)+|B2||Σ|,\displaystyle\mu<|\{u<\phi_{0}\}\cap(B_{2}\times(-2,-2a))|\leq\mu/4\cdot(2-2a)+|B_{2}|\cdot|\Sigma|,

and

|Σ|μ2|B2|.\displaystyle|\Sigma|\geq\frac{\mu}{2|B_{2}|}.

Next we prove that the integral n(uϕ1)+2(t,x)dμ(x)\int_{\mathbb{H}^{n}}(u-\phi_{1})_{+}^{2}(t,x)\mathop{}\!\mathrm{d}\mu(x) is very small for most of the time tt in Σ\Sigma. Since (uϕ1)+λ(u-\phi_{1})_{+}\leq\lambda, we have

CC2λ2\displaystyle CC_{2}\lambda^{2} 22aJ((uϕ1)+,(uϕ1))dt\displaystyle\geq\int_{-2}^{-2a}\mathcal{E}_{J}((u-\phi_{1})_{+},(u-\phi_{1})_{-})\mathop{}\!\mathrm{d}t
CΣB3{|{u(,t)<ϕ0}B2|μ/4}(uϕ1)+(x)(1λ)J(x,y)dμ(x)dμ(y)dt\displaystyle\geq C\int_{\Sigma}\int_{B_{3}}\int_{\{|\{u(\cdot,t)<\phi_{0}\}\cap B_{2}|\geq\mu/4\}}(u-\phi_{1})_{+}(x)\cdot(1-\lambda)\cdot J(x,y)\mathop{}\!\mathrm{d}\mu(x)\mathop{}\!\mathrm{d}\mu(y)\mathop{}\!\mathrm{d}t
CΛ1μ8Σn(uϕ1)+(x)dμ(x)dt\displaystyle\geq C\Lambda^{-1}\frac{\mu}{8}\int_{\Sigma}\int_{\mathbb{H}^{n}}(u-\phi_{1})_{+}(x)\mathop{}\!\mathrm{d}\mu(x)\mathop{}\!\mathrm{d}t
Cμ8λΣn(uϕ1)+2(x)dμ(x)dt.\displaystyle\geq C\frac{\mu}{8\lambda}\int_{\Sigma}\int_{\mathbb{H}^{n}}(u-\phi_{1})_{+}^{2}(x)\mathop{}\!\mathrm{d}\mu(x)\mathop{}\!\mathrm{d}t.

Hereafter,

Σn(uϕ1)+2(x)dμ(x)dtCC2λ3μλ318,\displaystyle\int_{\Sigma}\int_{\mathbb{H}^{n}}(u-\phi_{1})_{+}^{2}(x)\mathop{}\!\mathrm{d}\mu(x)\mathop{}\!\mathrm{d}t\leq CC_{2}\frac{\lambda^{3}}{\mu}\leq\lambda^{3-\frac{1}{8}},

by requiring

λ(μCC2)8.\displaystyle\lambda\leq(\frac{\mu}{CC_{2}})^{8}.

By Tchebychev inequality, we have

n(uϕ1)+2(x,t)dμ(x)λ314\displaystyle\int_{\mathbb{H}^{n}}(u-\phi_{1})_{+}^{2}(x,t)\mathop{}\!\mathrm{d}\mu(x)\leq\lambda^{3-\frac{1}{4}} (7.20)

except for a tiny set AΣA\subset\Sigma with |A|λ18|A|\leq\lambda^{\frac{1}{8}}. Furthermore,

|Σ|μ2|B2|2λ18ifλ(μ4|B2|)8.\displaystyle|\Sigma|\geq\frac{\mu}{2|B_{2}|}\geq 2\lambda^{\frac{1}{8}}\leavevmode\nobreak\ \leavevmode\nobreak\ \rm{if}\leavevmode\nobreak\ \lambda\leq(\frac{\mu}{4|B_{2}|})^{8}.

Therefore, (7.20) holds for most of the time t[2,2a]t\in[-2,-2a] with a measure greater than μ/4|B2|\mu/4|B_{2}|.

Searching an intermediate set between two level sets in the setting:

Define Σ1:={t[2a,0]:|{u(,t)>ϕ2}|>ν/4a}\Sigma_{1}\mathrel{\mathop{\mathchar 58\relax}}=\{t\in[-2a,0]\mathrel{\mathop{\mathchar 58\relax}}|\{u(\cdot,t)>\phi_{2}\}|>\nu/4a\} The condition (7.17) implies

ν|B3||Σ1|+ν4a2a.\displaystyle\nu\leq|B_{3}|\cdot|\Sigma_{1}|+\frac{\nu}{4a}\cdot 2a.

Hence |Σ1|>0|\Sigma_{1}|>0 and there exists T0[2a,0]T_{0}\in[-2a,0] such that

|{(uϕ2)+(T0,)>0}|>ν4aν4.\displaystyle|\{(u-\phi_{2})_{+}(T_{0},\cdot)>0\}|>\frac{\nu}{4a}\geq\frac{\nu}{4}.

Let time T0T_{0} go backwards, following the inequality (7.20) from the first step, we can find another T1[2,2a]T_{1}\in[-2,-2a] satisfying

n(uϕ1)+2(T1,x)dμ(x)λ314.\displaystyle\int_{\mathbb{H}^{n}}(u-\phi_{1})_{+}^{2}(T_{1},x)\mathop{}\!\mathrm{d}\mu(x)\leq\lambda^{3-\frac{1}{4}}.

Then,

H(T1)C2λ314.\displaystyle H(T_{1})\leq C_{2}\lambda^{3-\frac{1}{4}}.

Moreover, we will derive the estimate of the integral n(uϕ1)+2(T0,x)dμ(x)\int_{\mathbb{H}^{n}}(u-\phi_{1})_{+}^{2}(T_{0},x)\mathop{}\!\mathrm{d}\mu(x) at time T0T_{0}.

n(uϕ1)+2(T0,x)dμ(x)\displaystyle\int_{\mathbb{H}^{n}}(u-\phi_{1})_{+}^{2}(T_{0},x)\mathop{}\!\mathrm{d}\mu(x) n(ϕ1ϕ2)2𝟙{(uϕ2)+>0}\displaystyle\geq\int_{\mathbb{H}^{n}}(\phi_{1}-\phi_{2})^{2}\mathbbm{1}_{\{(u-\phi_{2})_{+}>0\}}
n(λλ2)2F2(x)𝟙{(uϕ2)+>0}CFλ2ν34.\displaystyle\geq\int_{\mathbb{H}^{n}}(\lambda-\lambda^{2})^{2}F^{2}(x)\mathbbm{1}_{\{(u-\phi_{2})_{+}>0\}}\geq C_{F}\frac{\lambda^{2}\nu^{3}}{4}.

Then,

H(T0)C1CFλ2ν34.\displaystyle H(T_{0})\geq C_{1}C_{F}\frac{\lambda^{2}\nu^{3}}{4}.

By requiring λ114C1CFC2ν364\lambda^{1-\frac{1}{4}}\leq\frac{C_{1}C_{F}}{C_{2}}\cdot\frac{\nu^{3}}{64}, then we have H(T1)C1CFλ2ν364H(T_{1})\leq C_{1}C_{F}\frac{\lambda^{2}\nu^{3}}{64}. Define

Σ2:={t[T1,T0]:C1CFλ2ν364<H(t)<C1CFλ2ν34}.\displaystyle\Sigma_{2}\mathrel{\mathop{\mathchar 58\relax}}=\{t\in[T_{1},T_{0}]\mathrel{\mathop{\mathchar 58\relax}}C_{1}C_{F}\frac{\lambda^{2}\nu^{3}}{64}<H(t)<C_{1}C_{F}\frac{\lambda^{2}\nu^{3}}{4}\}.

Since H(t)λ2H^{\prime}(t)\leq\lambda^{2}, we get |Σ2|C1CFν3/8|\Sigma_{2}|\geq C_{1}C_{F}\nu^{3}/8. Now we are ready to search an intermediate level set between the level sets in (7.16) and (7.17). For any time τΣ2\tau\in\Sigma_{2}, we have |{(uϕ2)+(τ,)0}|<ν4.|\{(u-\phi_{2})_{+}(\tau,\cdot)\geq 0\}|<\frac{\nu}{4}. Otherwise, we can obtain H(τ)C1CFλ2ν34H(\tau)\geq C_{1}C_{F}\frac{\lambda^{2}\nu^{3}}{4} and τΣ2\tau\notin\Sigma_{2}, which is a contradiction.

To make sure that we can find an intermediate level set {ϕ0<u<ϕ2}\{\phi_{0}<u<\phi_{2}\} whose measure is positive, we have to give a small upper bound of |{(uϕ0)+(,t)0}B3||\{(u-{\phi_{0}})_{+}(\cdot,t)\leq 0\}\cap B_{3}| for most of the time tt in Σ2\Sigma_{2}. Define Σ3:={t:|{(uϕ0)+(,t)0}B3|μ}\Sigma_{3}\mathrel{\mathop{\mathchar 58\relax}}=\{t\mathrel{\mathop{\mathchar 58\relax}}|\{(u-{\phi_{0}})_{+}(\cdot,t)\leq 0\}\cap B_{3}|\geq\mu\}. Then

CC2λ2\displaystyle CC_{2}\lambda^{2} 20n(uϕ1)+(uϕ1)J(x,y)\displaystyle\geq\int_{-2}^{0}\int_{\mathbb{H}^{n}}(u-\phi_{1})_{+}(u-\phi_{1})_{-}J(x,y)
CμΣ3B3(uϕ1)+dμ(x)dt\displaystyle\geq C\mu\int_{\Sigma_{3}}\int_{B_{3}}(u-\phi_{1})_{+}\mathop{}\!\mathrm{d}\mu(x)\mathop{}\!\mathrm{d}t
Cμλ|Σ3|(C1CFλ2ν364).\displaystyle\geq\frac{C\mu}{\lambda}|\Sigma_{3}|\cdot(\frac{C_{1}C_{F}\lambda^{2}\nu^{3}}{64}).

Hereafter,

|Σ3|CC2λC1μν3.\displaystyle|\Sigma_{3}|\leq\frac{CC_{2}\lambda}{C_{1}\mu\nu^{3}}.

By requiring λ<CC12C2μν6\lambda<C\frac{C_{1}^{2}}{C_{2}}\mu\nu^{6}, we have

|Σ3|C1CFν3/16|Σ2|/2.\displaystyle|\Sigma_{3}|\leq C_{1}C_{F}\nu^{3}/16\leq|\Sigma_{2}|/2.

Therefore, for those time tΣ2\Σ3t\in\Sigma_{2}\backslash\Sigma_{3}, we have

|{ϕ0<u(,t)ϕ2}B3||B3|μν/41/2.\displaystyle|\{\phi_{0}<u(\cdot,t)\leq\phi_{2}\}\cap B_{3}|\geq|B_{3}|-\mu-\nu/4\geq 1/2.

And

|{1+F<u<1+λ2F}(B3×(2,0))|\displaystyle|\{1+F<u<1+\lambda^{2}F\}\cap(B_{3}\times(-2,0))| Σ2\Σ31/2dt\displaystyle\geq\int_{\Sigma_{2}\backslash\Sigma_{3}}1/2\mathop{}\!\mathrm{d}t (7.21)
|Σ2|/4C1CFν3/8.\displaystyle\geq|\Sigma_{2}|/4\geq C_{1}C_{F}\nu^{3}/8.

Remark 8.

In the above proof, we use the inequality {|z|>1}((|z|1)+α/4λ)+2(|z|3λ4/α)Q+αdμ(z)<\int_{\{|z|>1\}}\frac{((|z|-1)_{+}^{\alpha/4}-\lambda)_{+}^{2}}{(|z|-3\lambda^{4/\alpha})^{Q+\alpha}}\mathop{}\!\mathrm{d}\mu(z)<\infty, which is similar to the inequality {{|x|>1}n}((|x|1)+α/4λ)+2(|x|3λ4/α)n+αdx<\int_{\{\{|x|>1\}\cap\mathbb{R}^{n}\}}\frac{((|x|-1)_{+}^{\alpha/4}-\lambda)_{+}^{2}}{(|x|-3\lambda^{4/\alpha})^{n+\alpha}}\mathop{}\!\mathrm{d}x<\infty on n\mathbb{R}^{n}. The difference is we use the homogeneous dimension QQ instead on the Heisenberg group n\mathbb{H}^{n}.

We are ready to show the oscillation Lemma by combining the first and second De Giorgi Lemma. For any λ\lambda in Lemma 10, we define a new barrier function with more restrictive control:

Hλ,ϵ(x)=((|x|λ4/α)+ϵ1)+,xn.\displaystyle H_{\lambda,\epsilon}(x)=((|x|-\lambda^{-4/\alpha})_{+}^{\epsilon}-1)_{+},\leavevmode\nobreak\ \leavevmode\nobreak\ x\in\mathbb{H}^{n}.
Lemma 11.

Assume 0<C1β(s)C20<C_{1}\leq\beta^{\prime}(s)\leq C_{2} for every 1/4s21/4\leq s\leq 2, or 2s1/4-2\leq s\leq-1/4. If u:n×[2,0]u\mathrel{\mathop{\mathchar 58\relax}}\mathbb{H}^{n}\times[-2,0]\rightarrow\mathbb{R} is a weak solution to equation (7.1) satisfying

|u(x,t)|1+Hλ,ϵ(x)inn×[2,0],\displaystyle|u(x,t)|\leq 1+H_{\lambda,\epsilon}(x)\leavevmode\nobreak\ \leavevmode\nobreak\ \text{in}\leavevmode\nobreak\ \mathbb{H}^{n}\times[-2,0], (7.22)

then there exists 0<a<10<a<1 and 0<θ<120<\theta<\frac{1}{2}, depending on the nonlinearity function β\beta and the dimensional constants, such that

supΓ1,ainfΓ1,a2θ.\displaystyle\sup\limits_{\Gamma_{1,a}}-\inf\limits_{\Gamma_{1,a}}\leq 2-\theta. (7.23)
Proof.

Choosing a=1a=1 in Lemma 9, if

|{u>0}Γ2,2|Cδ(β)1+Qα\displaystyle|\{u>0\}\cap\Gamma_{2,2}|\leq C\delta(\beta)^{1+\frac{Q}{\alpha}}

holds, then we obtain the oscillation by Lemma 9. Otherwise, if

|{u>0}Γ2,2|Cδ(β)1+Qα\displaystyle|\{u>0\}\cap\Gamma_{2,2}|\geq C\delta(\beta)^{1+\frac{Q}{\alpha}}

then

|{u>0}(B2×(2,2a))|>C2δ(β)1+Qα\displaystyle|\{u>0\}\cap(B_{2}\times(-2,-2a))|>\frac{C}{2}\delta(\beta)^{1+\frac{Q}{\alpha}}

Here we take 0<a<10<a<1 such that |Γ2,2a|<C2δ(β)1+Qα|\Gamma_{2,2a}|<\frac{C}{2}\delta(\beta)^{1+\frac{Q}{\alpha}}. Notice that aa only depends on the nonlinearity function β\beta and the dimensional constants. Therefore, we work on the solution u-u with the nonlinearity function:

β~(s)=β(s)\displaystyle\tilde{\beta}(s)=-\beta(-s)

Furthermore, u-u satisfies (7.17) in Lemma 10 by choosing μ=C2δ(β)1+Qα\mu=\frac{C}{2}\delta(\beta)^{1+\frac{Q}{\alpha}}. To be clear without misunderstandings, we assume that uu satisfies the condition

|{u<0}(B2×(2,2a))|>C2δ(β)1+Qα|\{u<0\}\cap(B_{2}\times(-2,-2a))|>\frac{C}{2}\delta(\beta)^{1+\frac{Q}{\alpha}}

and

0<C1β(s)C2forevery 1/4s2.0<C_{1}\leq\beta^{\prime}(s)\leq C_{2}\leavevmode\nobreak\ \leavevmode\nobreak\ \text{for}\leavevmode\nobreak\ \text{every}\leavevmode\nobreak\ 1/4\leq s\leq 2.

Considering the sequence of rescaled functions:

uk+1=uk(1λ2)λ2,u0=u.\displaystyle u_{k+1}=\frac{u_{k}-(1-\lambda^{2})}{\lambda^{2}},\leavevmode\nobreak\ \leavevmode\nobreak\ u_{0}=u.

Then uku_{k} is a weak solution to

tβk(uk)+Luk=0.\displaystyle\partial_{t}\beta_{k}(u_{k})+Lu_{k}=0.

with a nonlinearity recurrence relation

βk+1(s)=βk(λ2s+1λ2)λ2,β0=β.\displaystyle\beta_{k+1}(s)=\frac{\beta_{k}(\lambda^{2}s+1-\lambda^{2})}{\lambda^{2}},\leavevmode\nobreak\ \leavevmode\nobreak\ \beta_{0}=\beta.

and

βk(s)=β(λ2ks+1λ2k)λ2k.\displaystyle\beta_{k}(s)=\frac{\beta(\lambda^{2k}s+1-\lambda^{2k})}{\lambda^{2k}}.

Here, βk(s)=β(λ2ks+1λ2k)\beta_{k}^{\prime}(s)=\beta^{\prime}(\lambda^{2k}s+1-\lambda^{2k}). When 1/4s21/4\leq s\leq 2, we have

2λ2ks+1λ2k1/4,forallk0.2\geq\lambda^{2k}s+1-\lambda^{2k}\geq 1/4,\leavevmode\nobreak\ \leavevmode\nobreak\ \text{for}\leavevmode\nobreak\ \text{all}\leavevmode\nobreak\ k\geq 0.

Hence, for all k0k\geq 0, we have the uniform bound of βk(s)\beta_{k}^{\prime}(s),

C1β(λ2ks+1λ2k)=βk(s)C2,forall 1/4s2.\displaystyle C_{1}\leq\beta^{\prime}(\lambda^{2k}s+1-\lambda^{2k})=\beta_{k}^{\prime}(s)\leq C_{2},\leavevmode\nobreak\ \leavevmode\nobreak\ \text{for}\leavevmode\nobreak\ \text{all}\leavevmode\nobreak\ 1/4\leq s\leq 2.

Moreover, the constants δ(βk)\delta(\beta_{k}) for k0k\geq 0 in Lemma 9 must have uniform positive lower bounds δ0\delta_{0}. We choose ν\nu in Lemma 10 as the constant constructed in Lemma 9:

ν=Cδ01+Qαa(1+Qα).\nu=C\delta_{0}^{1+\frac{Q}{\alpha}}a^{(1+\frac{Q}{\alpha})}.

We choose k0k_{0} as the smallest integer greater than |B3×(2,0)|/γ|B_{3}\times(-2,0)|/\gamma and ϵ\epsilon small enough such that

(|x|ϵ1)+λ2(k0+1)(|x|α/41)+,forallxn.\displaystyle\frac{(|x|^{\epsilon}-1)_{+}}{\lambda^{2(k_{0}+1)}}\leq(|x|^{\alpha/4}-1)_{+},\leavevmode\nobreak\ \leavevmode\nobreak\ \text{for}\leavevmode\nobreak\ \text{all}\leavevmode\nobreak\ x\in\mathbb{H}^{n}.

Here γ\gamma is the constant in (7.19) corresponding to the previous determined constants μ,ν,C1,C2\mu,\nu,C_{1},C_{2}. On the other hand, since uk+1(x,t)=1+uk(x,t)1λ2u_{k+1}(x,t)=1+\frac{u_{k}(x,t)-1}{\lambda^{2}}, we have

uk(x,t)=1+u(x,t)1λ2k.\displaystyle u_{k}(x,t)=1+\frac{u(x,t)-1}{\lambda^{2k}}.

Therefore, for each 0kk0+10\leq k\leq{k_{0}+1},

uk(x,t)1+((|x|λ4/α)+ϵ1)+λ2(k0+1)1+((|x|λ4/α)+α/41)+=1+ψλ(x).\displaystyle u_{k}(x,t)\leq 1+\frac{((|x|-\lambda^{-4/\alpha})_{+}^{\epsilon}-1)_{+}}{\lambda^{2(k_{0}+1)}}\leq 1+((|x|-\lambda^{-4/\alpha})_{+}^{\alpha/4}-1)_{+}=1+\psi_{\lambda}(x).

As long as |{uk>1+λ2F}(B3×(2a,0))|ν|\{u_{k}>1+\lambda^{2}F\}\cap(B_{3}\times(-2a,0))|\geq\nu for all 1kk01\leq k\leq k_{0}, by Lemma 10, we have

|{uk1+λ2F}(B3×(2a,0))|\displaystyle|\{u_{k}\geq 1+\lambda^{2}F\}\cap(B_{3}\times(-2a,0))| |{uk1+F}(B3×(2a,0))|γ\displaystyle\leq|\{u_{k}\geq 1+F\}\cap(B_{3}\times(-2a,0))|-\gamma
|{uk11+λ2F}(B3×(2a,0))|γ\displaystyle\leq|\{u_{k-1}\geq 1+\lambda^{2}F\}\cap(B_{3}\times(-2a,0))|-\gamma
|B3×[2,0]|kγ.\displaystyle\leq|B_{3}\times[-2,0]|-k\cdot\gamma.

This can not true up to k0k_{0}. Hence, there exists 1kk01\leq k^{\prime}\leq k_{0} such that |{uk>1+λ2F}(B3×(2a,0))|ν|\{u_{k^{\prime}}>1+\lambda^{2}F\}\cap(B_{3}\times(-2a,0))|\leq\nu. And we have,

|{uk+1>0}(B2×(2a,0))||{uk>1+λ2F}(B3×(2a,0))|ν.\displaystyle|\{u_{k^{\prime}+1}>0\}\cap(B_{2}\times(-2a,0))|\leq|\{u_{k^{\prime}}>1+\lambda^{2}F\}\cap(B_{3}\times(-2a,0))|\leq\nu.

By Lemma 9, we have uk+112u_{k^{\prime}+1}\leq\frac{1}{2} in Γ1,a\Gamma_{1,a}. This gives the result with 0<θ=λ2(k0+2)2<120<\theta=\frac{\lambda^{2(k_{0}+2)}}{2}<\frac{1}{2}. ∎

Now we are ready to prove the continuity of the weak solutions to equation (1.1) by contradiction. Based on the oscillation Lemma 11, we will construct the iterative sequence in which are the solutions to the rescaled equation (7.1). And the iterative sequence exhibits the quantitative behavior of the solution around some given point since they are constructed over a decreasing parabolic cylinder.

Proof of Theorem 3:

Proof.

We first use translations and rescaling arguments to move each point (x0,t0)(x_{0},t_{0}) to the origin. Let τ0=inf{1,t0/3}\tau_{0}=\inf\{1,t_{0}/3\} and A=supt0/3tt0u(,t)A=\sup\limits_{t_{0}/3\leq t\leq t_{0}}\mathinner{\!\left\lVert u(\cdot,t)\right\rVert}_{\infty}. The existence of AA is guaranteed by the smoothing effects. Define the following rescaled function:

u0(x,t)=1Au(x0τ01/αx,t0+τ0t),t[2,0]\displaystyle u_{0}(x,t)=\frac{1}{A}u(x_{0}\cdot\tau_{0}^{1/\alpha}x,t_{0}+\tau_{0}t),\leavevmode\nobreak\ \leavevmode\nobreak\ t\in[-2,0]

which is a solution to equation (7.1) with the nonlinearity function β0(s)=1Aβ(As)\beta_{0}(s)=\frac{1}{A}\beta(As) and the measurable kernel

J0(x,y)=τ0Q+ααJ(x0τ01/αx,x0τ01/αy).J_{0}(x,y)=\tau_{0}^{\frac{Q+\alpha}{\alpha}}J(x_{0}\cdot\tau_{0}^{1/\alpha}x,x_{0}\cdot\tau_{0}^{1/\alpha}y).

From the setting, we have |u0(x,t)|1|u_{0}(x,t)|\leq 1 in n×[2,0]\mathbb{H}^{n}\times[-2,0] and still use uu to denote u0u_{0}.
Let Qk=ΓRk,RkαQ_{k}=\Gamma_{R^{-k},R^{-k\alpha}} for every k0k\geq 0 and some large enough R>1R>1 to be determined later. Define the semi-oscillation of uu in Qk1Q_{k-1},

ωk=supQk1uinfQk1u2.\displaystyle{\omega}_{k}=\frac{\sup_{Q_{k-1}}u-\inf_{Q_{k-1}}u}{2}.

Our goal is to prove ωk0\omega_{k}\rightarrow 0 as kk\rightarrow\infty. We prove it by contradiction and assume ωkξ>0\omega_{k}\geq\xi>0. Given k1k\geq 1, we define

uk(x,t)=u(Rkx,Rαkt)μkωk,μk=supQk1u+infQk1u2.\displaystyle u_{k}(x,t)=\frac{u(R^{-k}x,R^{-\alpha k}t)-\mu_{k}}{\omega_{k}},\leavevmode\nobreak\ \leavevmode\nobreak\ \mu_{k}=\frac{\sup_{Q_{k-1}}u+\inf_{Q_{k-1}}u}{2}.

The functions uku_{k} satisfy the equation (7.1) with

βk(s)=β(ωks+μk)ωk,Jk(x,y)=R(Q+α)kJ0(Rkx,Rky).\displaystyle\beta_{k}(s)=\frac{\beta(\omega_{k}s+\mu_{k})}{\omega_{k}},\leavevmode\nobreak\ \leavevmode\nobreak\ J_{k}(x,y)=R^{-(Q+\alpha)k}J_{0}(R^{-k}x,R^{-k}y).

Here, βk(s)=β(ωks+μk)\beta_{k}^{\prime}(s)=\beta^{\prime}(\omega_{k}s+\mu_{k}). When 2s2-2\leq s\leq 2, we have

2\displaystyle-2 supQk1u+3infQk1u2ωks+μk\displaystyle\leq\frac{-\sup_{Q_{k-1}}u+3\inf_{Q_{k-1}}u}{2}\leq\omega_{k}s+\mu_{k}
2supQk1u2infQk1u+supQk1u+infQk1u2\displaystyle\leq\frac{2\sup_{Q_{k-1}}u-2\inf_{Q_{k-1}}u+\sup_{Q_{k-1}}u+\inf_{Q_{k-1}}u}{2}
2.\displaystyle\leq 2.

Hence, {βk(s)}k=0\{\beta_{k}^{\prime}(s)\}_{k=0}^{\infty} have uniform positive lower bound and upper bound for 1/4s21/4\leq s\leq 2 if μk0\mu_{k}\geq 0 or 2s1/4-2\leq s\leq-1/4 if μk0\mu_{k}\leq 0.

We claim that {uk}k=0\{u_{k}\}_{k=0}^{\infty} satisfy condition (7.21) in Lemma 11. If |x|R|x|\leq R, we have |uk|1|u_{k}|\leq 1 by requiring Rα>2R^{\alpha}>2. If |x|R|x|\geq R, choosing RR large enough such that Hλ,ϵ(R)2ξξH_{\lambda,\epsilon}(R)\geq\frac{2-\xi}{\xi}, we have

|uk(x,t)|2ωk2ξ1+Hλ,ϵ(R)1+Hλ,ϵ(x).\displaystyle|u_{k}(x,t)|\leq\frac{2}{\omega_{k}}\leq\frac{2}{\xi}\leq 1+H_{\lambda,\epsilon}(R)\leq 1+H_{\lambda,\epsilon}(x).

Apply Lemma 10, we have

supΓ1,aukinfΓ1,auk2θ.\displaystyle\frac{\sup_{\Gamma_{1,a}}u_{k}-\inf_{\Gamma_{1,a}}u_{k}}{2}\leq\theta.

Take RR large enough such that Rα>1aR^{\alpha}>\frac{1}{a}, we obtain

ωk=supQk1uinfQk1u2supΓ1,auk2infΓ1,auk22ωk2θωk2.\displaystyle{\omega}_{k}=\frac{\sup_{Q_{k-1}}u-\inf_{Q_{k-1}}u}{2}\leq\frac{\sup_{\Gamma_{1,a}}u_{k-2}-\inf_{\Gamma_{1,a}}u_{k-2}}{2}\omega_{k-2}\leq\theta\omega_{k-2}.

which leads to a contradiction. Therefore, we proved the continuity of weak solutions to equation (1.1) for general m>0m>0. ∎

We have proved the continuity of the weak solution by the oscillation Lemma 11 and the constructed iterative sequence. Based on this construction, it is natural to think about proving the Hölder continuity at some reasonable point. The critical point is to provide a uniform oscillation for each step of iteration.

Reviewing the conditions in Lemma 11, it requires C1β(s)C2C_{1}\leq\beta^{\prime}(s)\leq C_{2}. Let us recall that βk(s)=β(ωks+μk)\beta^{\prime}_{k}(s)=\beta^{\prime}(\omega_{k}s+\mu_{k}). Since ωk0\omega_{k}\rightarrow 0 following from the above continuity result, the easiest way is to give a good bound of μk\mu_{k}. Without loss of generality, if u(x0,t0)>0u(x_{0},t_{0})>0, then μkδ\mu_{k}\geq\delta for some δ>0\delta>0 when kk is very large due to the continuity. Hereafter, Lemma 11 provided a uniform oscillation for the iterative sequence.

Proof of Theorem 4:

Proof.

By rescaling arguments, we assume (x0,t0)=(0,0)(x_{0},t_{0})=(0,0) and u(0,0)>0u(0,0)>0. By the continuity of uu just proved above, there exists a large enough R0R_{0} such that

infΓR1,Rαuu(0,0)/2foreachRR0.\inf_{\Gamma_{R^{-1},R^{-\alpha}}}u\geq u(0,0)/2\leavevmode\nobreak\ \text{for}\leavevmode\nobreak\ \text{each}\leavevmode\nobreak\ R\geq R_{0}.

Notice that R0R_{0} varies for different points even with same function value. Define:

u0=u(R01x,R0αt),(x,t)n×[2,0].u_{0}=u(R_{0}^{-1}x,R_{0}^{-\alpha}t),\leavevmode\nobreak\ (x,t)\in\mathbb{H}^{n}\times[-2,0].

Let Qk=ΓRk,RkαQ_{k}=\Gamma_{R^{-k},R^{-k\alpha}} for every k0k\geq 0 and some large enough R>1R>1 to be determined later. Define the semi-oscillation of u0u_{0} in Qk1Q_{k-1},

ωk=supQk1u0infQk1u02.\displaystyle{\omega}_{k}=\frac{\sup_{Q_{k-1}}u_{0}-\inf_{Q_{k-1}}u_{0}}{2}.
μk=supQk1u0+infQk1u02\displaystyle\mu_{k}=\frac{\sup_{Q_{k-1}}u_{0}+\inf_{Q_{k-1}}u_{0}}{2}

Then, by the construction of u0u_{0},

μku(0,0)/2fork1.\displaystyle\mu_{k}\geq u(0,0)/2\leavevmode\nobreak\ \text{for}\leavevmode\nobreak\ k\geq 1.

We still use uu to denote u0u_{0} and define the following iterative sequence:

uk(x,t)=u(Rkx,Rαkt)μkνk,u0=u.\displaystyle u_{k}(x,t)=\frac{u(R^{-k}x,R^{-\alpha k}t)-\mu_{k}}{\nu_{k}},\leavevmode\nobreak\ \leavevmode\nobreak\ u_{0}=u.

Here, νk0\nu_{k}\geq 0 will be determined later. Observe that uku_{k} satisfies equation (7.1) with nonlinearity βk(s)\beta_{k}(s) and measurable kernel Jk(x,y)J_{k}(x,y),

βk(s)=β(νks+μk)νk,Jk(x,y)=R(Q+α)kJ(Rkx,Rky).\displaystyle\beta_{k}(s)={\frac{\beta(\nu_{k}s+\mu_{k})}{\nu_{k}}},\leavevmode\nobreak\ \leavevmode\nobreak\ J_{k}(x,y)=R^{-(Q+\alpha)k}J(R^{-k}x,R^{-k}y).

Since 2νks+μkmin(u(0,0)/2,1/4)2\geq\nu_{k}s+\mu_{k}\geq\min(u(0,0)/2,1/4) if 1/4s21/4\leq s\leq 2, the nonlinearity function {βk(s)}k=0\{\beta_{k}^{\prime}(s)\}_{k=0}^{\infty} share uniform upper and lower bound depending on the value u(0,0)u(0,0). Therefore, from Lemma 11, there exists 0<a,θ<10<a,\theta<1 corresponding to the above uniform constant of βk(s)\beta_{k}^{\prime}(s). To overcome the difficulties from a<1a<1, we define Qka=ΓRk,aRkα.Q_{k}^{a}=\Gamma_{R^{-k},aR^{-k\alpha}}. Let

μk=supQk1au+infQk1au2,ωk=supQk1auinfQk1au2\displaystyle\mu_{k}^{\prime}=\frac{\sup_{Q_{k-1}^{a}}u+\inf_{Q_{k-1}^{a}}u}{2},\leavevmode\nobreak\ \leavevmode\nobreak\ \omega_{k}^{\prime}=\frac{\sup_{Q_{k-1}^{a}}u-\inf_{Q_{k-1}^{a}}u}{2}

Notice that QkaQkQ_{k}^{a}\subset Q_{k} and μku(0,0)/2fork1\mu_{k}^{\prime}\geq u(0,0)/2\leavevmode\nobreak\ \leavevmode\nobreak\ \text{for}\leavevmode\nobreak\ k\geq 1. Hence, the constants aa and θ\theta in Lemma 11 don’t have to be changed. To be clear, we use μk,ωk\mu_{k},\omega_{k} to denote μk,ωk\mu_{k}^{\prime},\omega_{k}^{\prime}. Choose νk=θk\nu_{k}=\theta^{k} for k0k\geq 0 and μ0=0\mu_{0}=0. We claim that uk(x,t)1+Hλ,ϵ(x)u_{k}(x,t)\leq 1+H_{\lambda,\epsilon}(x) and prove it by induction. Choose RR large enough such that

2+Hλ,ϵ(x/R)θ1+Hλ,ϵ(x)for|x|R.\displaystyle\frac{2+H_{\lambda,\epsilon}(x/R)}{\theta}\leq 1+H_{\lambda,\epsilon}(x)\leavevmode\nobreak\ \leavevmode\nobreak\ \text{for}\leavevmode\nobreak\ |x|\geq R.

Assume uk(x,t)1+Hλ,ϵ(x)u_{k}(x,t)\leq 1+H_{\lambda,\epsilon}(x) holds for all 0kk00\leq k\leq k_{0}, our goal is to prove uk0+1(x,t)1+Hλ,ϵ(x)u_{{k_{0}}+1}(x,t)\leq 1+H_{\lambda,\epsilon}(x). By Lemma 11, take RR large enough such that R2/aR\geq 2/a, we obtain

ωk=supQk1auinfQk1au2supΓ1,auk1infΓ1,auk12νk1θνk1,for 1kk0+1.\displaystyle{\omega}_{k}=\frac{\sup_{Q_{k-1}^{a}}u-\inf_{Q_{k-1}^{a}}u}{2}\leq\frac{\sup_{\Gamma_{1,a}}u_{k-1}-\inf_{\Gamma_{1,a}}u_{k-1}}{2}\nu_{k-1}\leq\theta\nu_{k-1},\leavevmode\nobreak\ \leavevmode\nobreak\ \text{for}\leavevmode\nobreak\ \leavevmode\nobreak\ 1\leq k\leq k_{0}+1.

Therefore, we obtain

ωkνk,for 1kk0+1.\displaystyle\omega_{k}\leq\nu_{k},\leavevmode\nobreak\ \leavevmode\nobreak\ \text{for}\leavevmode\nobreak\ 1\leq k\leq k_{0}+1.

If |x|R|x|\leq R and t[2,0]t\in[-2,0], we have

(R(k0+1)x,Rα(k0+1)t)Qk0a\displaystyle(R^{-(k_{0}+1)}x,R^{-\alpha(k_{0}+1)}t)\in Q_{k_{0}}^{a}

Since ωk0+1νk0+1\omega_{k_{0}+1}\leq\nu_{k_{0}+1}, we obtain

|uk0+1(x,t)|1for(x,t)BR×[2,0].\displaystyle|u_{k_{0}+1}(x,t)|\leq 1\leavevmode\nobreak\ \leavevmode\nobreak\ \text{for}\leavevmode\nobreak\ (x,t)\in B_{R}\times[-2,0].

Furthermore, since

μk0μk0+1νk0\displaystyle\frac{\mu_{k_{0}}-\mu_{k_{0}+1}}{\nu_{k_{0}}} μk0μk0+1ωk0\displaystyle\leq\frac{\mu_{k_{0}}-\mu_{k_{0}+1}}{\omega_{k_{0}}}
supQk01au+infQk01ausupQk0auinfQk0ausupQk01auinfQk01au\displaystyle\leq\frac{\sup_{Q_{k_{0}-1}^{a}}u+\inf_{Q_{k_{0}-1}^{a}}u-\sup_{Q_{k_{0}}^{a}}u-\inf_{Q_{k_{0}}^{a}}u}{\sup_{Q_{k_{0}-1}^{a}}u-\inf_{Q_{k_{0}-1}^{a}}u}
1,\displaystyle\leq 1,

we have

uk0+1(x,t)\displaystyle u_{k_{0}+1}(x,t) =u(R(k0+1)x,Rα(k0+1)t)μk0+1νk0+1\displaystyle=\frac{u(R^{-(k_{0}+1)}x,R^{-\alpha(k_{0}+1)}t)-\mu_{k_{0}+1}}{\nu_{k_{0}+1}}
μk0μk0+1+νk0(1+Hλ,ϵ(x/R))νk0+1\displaystyle\leq\frac{\mu_{k_{0}}-\mu_{k_{0}+1}+\nu_{k_{0}}(1+H_{\lambda,\epsilon}(x/R))}{\nu_{k_{0}+1}}
2+Hλ,ϵ(x/R)θ\displaystyle\leq\frac{2+H_{\lambda,\epsilon}(x/R)}{\theta}
1+Hλ,ϵ(x).\displaystyle\leq 1+H_{\lambda,\epsilon}(x).

Remark 9.

If the solution uu has a uniform positive lower bound, then we can choose R0=1R_{0}=1 and obtain the uniform Hölder regularity.

Remark 10.

As u(x0,t0)u(x_{0},t_{0}) approaches zero, the uniform bounds of βk(s)\beta_{k}^{\prime}(s) approach zero. This will lead to the degeneracy of the oscillation in Lemma 11, which means we can not obtain uniform Hölder regularity from the above constructions near the degenerate or singular points (u(x0,t0)=0u(x_{0},t_{0})=0).

As mentioned in the above remark, we can not use the oscillation Lemma directly near the vanishing points (u(x0,t0)=0u(x_{0},t_{0})=0). However, if m1m\geq 1, we can prove the Hölder regularity for all the sign-changing solutions by constructing another iterative sequence with different nonlinearity functions. The bounds of βk(s)\beta_{k}^{\prime}(s) will be well controlled by considering the two different situations near the vanishing point and we can impose the previous oscillation Lemma in a quantitative way.

Proof of Theorem 5:

Proof.

Denote

𝒬R0ϵ((x0,t0))=BR01(x0)×(t0R0α+ϵ,t0),\mathcal{Q}_{R_{0}}^{\epsilon}((x_{0},t_{0}))=B_{R_{0}^{-1}}(x_{0})\times(t_{0}-R_{0}^{-{\alpha+\epsilon}},t_{0}),
𝒬R0((x0,t0),ω)=BR01(x0)×(t0R0αωσ,t0),\mathcal{Q}_{R_{0}}((x_{0},t_{0}),\omega)=B_{R_{0}^{-1}}(x_{0})\times(t_{0}-R_{0}^{-\alpha}\omega^{-\sigma},t_{0}),
𝒬R0ρ((x0,t0),ω)=BR01(x0)×(t012ρR0αωσ,t0).\mathcal{Q}^{\rho}_{R_{0}}((x_{0},t_{0}),\omega)=B_{R_{0}^{-1}}(x_{0})\times(t_{0}-\frac{1}{2}\rho R_{0}^{-\alpha}\omega^{-\sigma},t_{0}).

Here σ=11m\sigma=1-\frac{1}{m}. There is no lack of generality to assume (x0,t0)=(0,0)(x_{0},t_{0})=(0,0) by scalings and translations. First, we consider the nonnegative solutions, the case of sign changing solution will be considered subsequently. For abrevity, denote 𝒬R0ϵ((0,0)):=𝒬R0ϵ\mathcal{Q}_{R_{0}}^{\epsilon}((0,0))\mathrel{\mathop{\mathchar 58\relax}}=\mathcal{Q}_{R_{0}}^{\epsilon}, 𝒬R((0,0),ω):=𝒬R(ω)\mathcal{Q}_{R}((0,0),\omega)\mathrel{\mathop{\mathchar 58\relax}}=\mathcal{Q}_{R}(\omega) and 𝒬Rρ((0,0),ω):=𝒬Rρ(ω)\mathcal{Q}^{\rho}_{R}((0,0),\omega)\mathrel{\mathop{\mathchar 58\relax}}=\mathcal{Q}^{\rho}_{R}(\omega). Taking a fixed R0>0R_{0}>0 such that 𝒬R0ϵ\mathcal{Q}_{R_{0}}^{\epsilon} is contained in n×[3,0]\mathbb{H}^{n}\times[-3,0]. Let

ω0=sup𝒬R0ϵuinf𝒬R0ϵu2:=osc(u;𝒬R0ϵ)2.\omega_{0}=\frac{\sup_{\mathcal{Q}_{R_{0}}^{\epsilon}}u-\inf_{\mathcal{Q}_{R_{0}}^{\epsilon}}u}{2}\mathrel{\mathop{\mathchar 58\relax}}=\frac{\text{osc}(u;\mathcal{Q}_{R_{0}}^{\epsilon})}{2}.

There exists RR0R\geq R_{0} such that

ω0σRϵ.\omega_{0}^{\sigma}\geq R^{-\epsilon}.

Hence,

osc(u;𝒬R(ω0))/2ω0.\displaystyle\text{osc}(u;\mathcal{Q}_{R}(\omega_{0}))/2\leq\omega_{0}.

(i) Assuming ω0σR0ϵ\omega_{0}^{\sigma}\geq R_{0}^{-\epsilon}, we have osc(u;𝒬R0(ω0))/2ω0\text{osc}(u;\mathcal{Q}_{R_{0}}(\omega_{0}))/2\leq\omega_{0}. Define

μ0=sup𝒬R0(ω0)u+inf𝒬R0(ω0)u2.\displaystyle\mu_{0}=\frac{\sup_{\mathcal{Q}_{R_{0}}(\omega_{0})}u+\inf_{\mathcal{Q}_{R_{0}}(\omega_{0})}u}{2}.

We also define

u1(x,t)=u(xC0R0,tC0αR0αω0σ)μ0ω0.u_{1}(x,t)=\frac{u(\frac{x}{C_{0}R_{0}},\frac{t}{C_{0}^{\alpha}R_{0}^{\alpha}{\omega_{0}}^{\sigma}})-\mu_{0}}{\omega_{0}}.

Choose C0C_{0} large enough such that C0α2C_{0}^{\alpha}\geq 2, for |x|C0|x|\leq C_{0} and 2t0-2\leq t\leq 0, we have

(xC0R0,tC0αR0αω0σ)𝒬R0(ω0).\displaystyle(\frac{x}{C_{0}R_{0}},\frac{t}{C_{0}^{\alpha}R_{0}^{\alpha}{\omega_{0}}^{\sigma}})\in\mathcal{Q}_{R_{0}}(\omega_{0}).

Then,

|u(x,t)|1,for(x,t)BC0×[2,0].\displaystyle|u(x,t)|\leq 1,\leavevmode\nobreak\ \leavevmode\nobreak\ \text{for}\leavevmode\nobreak\ (x,t)\in B_{C_{0}}\times[-2,0].

Considering equation (7.1) with the following nonlinearity function,

β(s)=|as+b|1/ma1σ,with 0ba32.\displaystyle\beta(s)=\frac{|as+b|^{1/m}}{a^{1-\sigma}},\leavevmode\nobreak\ \leavevmode\nobreak\ \text{with}\leavevmode\nobreak\ 0\leq\frac{b}{a}\leq\frac{3}{2}. (7.24)

Then

β(s)=|s+ba|σ.\beta^{\prime}(s)=|s+\frac{b}{a}|^{-\sigma}.

Notice that β(s)\beta^{\prime}(s) share the uniform upper and lower bound for 14s2\frac{1}{4}\leq s\leq 2. Let θ\theta and aa be the corresponding constants in Lemma 11. Denote 2θ2\frac{2-\theta}{2} by η\eta, which will be important in the following constructions. Furthermore, choose C0C_{0} large enough such that

2R0ϵ/σ1+Hλ,ϵ(x),for|x|C0,\displaystyle 2R_{0}^{{\epsilon}/{\sigma}}\leq 1+H_{\lambda,\epsilon}(x),\leavevmode\nobreak\ \leavevmode\nobreak\ \text{for}\leavevmode\nobreak\ |x|\geq C_{0},

and

2+Hλ,ϵ(xC0)η1+Hλ,ϵ(x),for|x|C0.\displaystyle\frac{2+H_{\lambda,\epsilon}(\frac{x}{C_{0}})}{\eta}\leq 1+H_{\lambda,\epsilon}(x),\leavevmode\nobreak\ \leavevmode\nobreak\ \text{for}\leavevmode\nobreak\ |x|\geq C_{0}. (7.25)

Then,

|u1(x,t)|2ω01+Hλ,ϵ(x)for|x|C0.\displaystyle|u_{1}(x,t)|\leq\frac{2}{\omega_{0}}\leq 1+H_{\lambda,\epsilon}(x)\leavevmode\nobreak\ \leavevmode\nobreak\ \text{for}\leavevmode\nobreak\ |x|\geq C_{0}.

Here we use

ω0R0ϵσsinceσ0\displaystyle\omega_{0}\geq R_{0}^{-\frac{\epsilon}{\sigma}}\leavevmode\nobreak\ \leavevmode\nobreak\ \text{since}\leavevmode\nobreak\ \sigma\geq 0 (7.26)

By Lemma 11, we obtain

supΓ1,au1infΓ1,au12η.\displaystyle\frac{\sup_{\Gamma_{1,a}}u_{1}-\inf_{\Gamma_{1,a}}u_{1}}{2}\leq\eta.

Hence, define the folllowing iterative sequence,

Rk=K0kR0,ωk=ηkω0,fork0.\displaystyle R_{k}=K_{0}^{k}\cdot R_{0},\leavevmode\nobreak\ \leavevmode\nobreak\ \omega_{k}=\eta^{k}\cdot\omega_{0},\leavevmode\nobreak\ \leavevmode\nobreak\ \text{for}\leavevmode\nobreak\ k\geq 0.

Here, K0K_{0} will be determined later. Define the corresponding solution uku_{k}:

uk(x,t)=u(xC0Rk1,tC0αRk1αωk1σ)μk1ωk1,μk1=sup𝒬Rk1(ωk1)u+inf𝒬Rk1(ωk1)u2.\displaystyle u_{k}(x,t)=\frac{u(\frac{x}{C_{0}R_{k-1}},\frac{t}{C_{0}^{\alpha}R_{k-1}^{\alpha}{\omega_{k-1}^{\sigma}}})-\mu_{k-1}}{\omega_{k-1}},\leavevmode\nobreak\ \leavevmode\nobreak\ \mu_{k-1}=\frac{\sup\limits_{\mathcal{Q}_{R_{k-1}}(\omega_{k-1})}u\leavevmode\nobreak\ +\leavevmode\nobreak\ \inf\limits_{\mathcal{Q}_{R_{k-1}}(\omega_{k-1})}u}{2}.

The corresponding nonlinearity functions will be:

βk(s)=|ωk1s+μk1|1/mωk11σ,\displaystyle\beta_{k}(s)=\frac{|\omega_{k-1}s+\mu_{k-1}|^{1/m}}{\omega_{k-1}^{1-\sigma}},

and

βk(s)=|s+μk1ωk1|σ.\displaystyle\beta_{k}^{\prime}(s)=|s+\frac{\mu_{k-1}}{\omega_{k-1}}|^{-\sigma}.

If we assume

inf𝒬Rk(ωk)uωk12for 0kk0,\frac{\inf\limits_{\mathcal{Q}_{R_{k}}(\omega_{k})}u}{\omega_{k}}\leq\frac{1}{2}\leavevmode\nobreak\ \leavevmode\nobreak\ \text{for}\leavevmode\nobreak\ 0\leq k\leq k_{0},

we have

μk1ωk132for 1kk0+1.\displaystyle\frac{\mu_{k-1}}{\omega_{k-1}}\leq\frac{3}{2}\leavevmode\nobreak\ \leavevmode\nobreak\ \text{for}\leavevmode\nobreak\ 1\leq k\leq k_{0}+1.

We claim that

|uk(x,t)|1+Hλ,ϵ(x),for 1kk0+3,\displaystyle|u_{k}(x,t)|\leq 1+H_{\lambda,\epsilon}(x),\leavevmode\nobreak\ \leavevmode\nobreak\ \text{for}\leavevmode\nobreak\ 1\leq k\leq k_{0}+3,

and

osc(u;𝒬Rk1(ωk1))/2ωk1,for 1kk0+2.\displaystyle\text{osc}(u;\mathcal{Q}_{R_{k-1}}({\omega_{k-1}}))/2\leq\omega_{k-1},\leavevmode\nobreak\ \leavevmode\nobreak\ \text{for}\leavevmode\nobreak\ 1\leq k\leq k_{0}+2.

We prove them by induction. Assume that |uk(x,t)|1+Hλ,ϵ(x)|u_{k}(x,t)|\leq 1+H_{\lambda,\epsilon}(x) and

osc(u;𝒬Rk1(ωk1))/2ωk1for 1kk1with 1k1k0+1.\displaystyle\text{osc}(u;\mathcal{Q}_{R_{k-1}}({\omega_{k-1}}))/2\leq\omega_{k-1}\leavevmode\nobreak\ \rm{for}\leavevmode\nobreak\ 1\leq k\leq k_{1}\leavevmode\nobreak\ \rm{with}\leavevmode\nobreak\ 1\leq k_{1}\leq k_{0}+1.

Therefore, by Lemma 11, we obtain

supΓ1,aukinfΓ1,auk2η,for 1kk1.\displaystyle\frac{\sup_{\Gamma_{1,a}}u_{k}-\inf_{\Gamma_{1,a}}u_{k}}{2}\leq\eta,\leavevmode\nobreak\ \leavevmode\nobreak\ \text{for}\leavevmode\nobreak\ 1\leq k\leq k_{1}.

Choose K0K_{0} large enough such that K0C0ησαa1αK_{0}\geq C_{0}\cdot\eta^{-\frac{\sigma}{\alpha}}\cdot a^{-\frac{1}{\alpha}}, then

𝒬Rk(ωk)B1C0Rk1×[aC0αRk1αωk1σ,0],for 1kk1.\displaystyle\mathcal{Q}_{R_{k}}(\omega_{k})\subset B_{\frac{1}{C_{0}R_{k-1}}}\times[\frac{-a}{C_{0}^{\alpha}R_{k-1}^{\alpha}\omega_{k-1}^{\sigma}},0],\leavevmode\nobreak\ \leavevmode\nobreak\ \text{for}\leavevmode\nobreak\ 1\leq k\leq k_{1}.

Hence,

osc(u;𝒬Rk(ωk))/2ωk,for 0kk1.\displaystyle\text{osc}(u;\mathcal{Q}_{R_{k}}({\omega_{k}}))/2\leq\omega_{k},\leavevmode\nobreak\ \leavevmode\nobreak\ \text{for}\leavevmode\nobreak\ 0\leq k\leq k_{1}.

Furthermore, if |x|C0|x|\leq C_{0} and 2t0-2\leq t\leq 0, we have

(xC0Rk1,tC0αRk1αωk1σ)𝒬Rk1(ωk1).\displaystyle(\frac{x}{C_{0}R_{k_{1}}},\frac{t}{C_{0}^{\alpha}R_{k_{1}}^{\alpha}{\omega_{k_{1}}^{\sigma}}})\subset\mathcal{Q}_{R_{k_{1}}}(\omega_{k_{1}}).

Then, we obtain

|uk1+1(x,t)|1,for(x,t)BC0×[2,0].\displaystyle|u_{k_{1}+1}(x,t)|\leq 1,\leavevmode\nobreak\ \leavevmode\nobreak\ \text{for}\leavevmode\nobreak\ (x,t)\in B_{C_{0}}\times[-2,0].

Since osc(u;𝒬Rk11(ωk11))/2ωk11\text{osc}(u;\mathcal{Q}_{R_{k_{1}-1}}({\omega_{k_{1}-1}}))/2\leq\omega_{k_{1}-1}, we have

μk11μk1ωk111,\displaystyle\frac{\mu_{k_{1}-1}-\mu_{k_{1}}}{\omega_{k_{1}-1}}\leq 1,

Therefore, by (7.24), we obtain

uk1+1(x,t)\displaystyle u_{k_{1}+1}(x,t) =ωk11uk1(xK0,tK0αησ)+μk11μk1ωk1\displaystyle=\frac{\omega_{k_{1}-1}u_{k_{1}}(\frac{x}{K_{0}},\frac{t}{K_{0}^{\alpha}\eta^{\sigma}})+\mu_{k_{1}-1}-\mu_{k_{1}}}{\omega_{k_{1}}}
2+Hλ,ϵ(xK0)η\displaystyle\leq\frac{2+H_{\lambda,\epsilon}(\frac{x}{K_{0}})}{\eta}
2+Hλ,ϵ(xC0)η1+Hλ,ϵ(x),for|x|C0.\displaystyle\leq\frac{2+H_{\lambda,\epsilon}(\frac{x}{C_{0}})}{\eta}\leq 1+H_{\lambda,\epsilon}(x),\leavevmode\nobreak\ \leavevmode\nobreak\ \text{for}\leavevmode\nobreak\ |x|\geq C_{0}.

Since osc(u;𝒬Rk1(ωk1))/2ωk1\text{osc}(u;\mathcal{Q}_{R_{k_{1}}}({\omega_{k_{1}}}))/2\leq\omega_{k_{1}}, we also obtain

|uk1+2(x,t)|1+Hλ,ϵ(x,t).\displaystyle|u_{k_{1}+2}(x,t)|\leq 1+H_{\lambda,\epsilon}(x,t).

(ii) If there exists some kk such that the average μk\mu_{k} is not comparatively small with the corresponding oscillation. Assume

inf𝒬Rk(ωk)uωk12for 0kk0,\frac{\inf\limits_{\mathcal{Q}_{R_{k}}(\omega_{k})}u}{\omega_{k}}\leq\frac{1}{2}\leavevmode\nobreak\ \leavevmode\nobreak\ \text{for}\leavevmode\nobreak\ 0\leq k\leq k_{0},

and

inf𝒬Rk0+1(ωk0+1)uωk0+112.\displaystyle\frac{\inf\limits_{\mathcal{Q}_{R_{k_{0}+1}}(\omega_{k_{0}+1})}u}{\omega_{k_{0}+1}}\geq\frac{1}{2}.

Consider

uk0+2(x,t)=u(xC0Rk0+1,tC0αRk0+1αωk0+1σ)μk0+1ωk0+1,\displaystyle u_{k_{0}+2}(x,t)=\frac{u(\frac{x}{C_{0}R_{k_{0}+1}},\frac{t}{C_{0}^{\alpha}R_{k_{0}+1}^{\alpha}{\omega_{k_{0}+1}^{\sigma}}})-\mu_{k_{0}+1}}{\omega_{k_{0}+1}},

Notice that |uk0+2(x,t)|1+Hλ,ϵ(x,t)|u_{k_{0}+2}(x,t)|\leq 1+H_{\lambda,\epsilon}(x,t) and infΓ1,1uk0+212\inf_{\Gamma_{1,1}}u_{k_{0}+2}\geq\frac{1}{2}. Combining with the previous proof, we obtain the uniform Hölder regularity including the vanishing points.

(iii) If ω0σR0ϵ\omega_{0}^{\sigma}\geq R_{0}^{-\epsilon} does not hold. In order to apply previous arguments, we try to see if it holds for 2R02R_{0} instead of R0R_{0}. If eventually we can find an integer kk such that

osc(u;𝒬2kR0ϵ)(2kR0)ϵσ.\displaystyle\text{osc}(u;\mathcal{Q}_{2^{k}R_{0}}^{\epsilon})\geq(2^{k}R_{0})^{-\frac{\epsilon}{\sigma}}.

Then we can apply step (i) and step (ii) again. otherwise,

osc(u;𝒬2kR0ϵ)(2kR0)ϵσ.\displaystyle\text{osc}(u;\mathcal{Q}_{2^{k}R_{0}}^{\epsilon})\leq(2^{k}R_{0})^{-\frac{\epsilon}{\sigma}}.

for all k0k\geq 0. In this case, we also get an uniform Hölder regularity.

(iv) Considering the sign-changing solutions with the initial data u(x,0)u(x,0). Let u(x,0):=u+(x,0)u(x,0)u(x,0)\mathrel{\mathop{\mathchar 58\relax}}=u_{+}(x,0)-u_{-}(x,0). Here, u+(x,0):=max(0,u(x,0))u_{+}(x,0)\mathrel{\mathop{\mathchar 58\relax}}=\max(0,u(x,0)) and u(x,0)=min(0,u(x,0))u_{-}(x,0)=-\min(0,u(x,0)). By the L1L^{1}-contraction property, we obtain

u(x,t)=u1(x,t)u2(x,t).\displaystyle u(x,t)=u_{1}(x,t)-u_{2}(x,t).

Here u1u_{1} and u2u_{2} are nonnegative solutions corresponding to the initial data u+(x,0)u_{+}(x,0) and u(x,0)u_{-}(x,0). By the previous parts (i)(ii)(iii), u1u_{1} and u2u_{2} are uniformly Hölder continuous. Therefore, the solution uu is also uniformly Hölder continuous. ∎

Acknowledgements

The author would like to thank her advisor Yannick Sire for bringing the investigations of fractional calculus and CR manifolds into her sight, suggesting studying them together, and also helpful discussions and valuable criticism on this paper.

References

  • [1] Ioannis Athanasopoulos, Luis Caffarelli and Emmanouil Milakis “The Two-Phase Stefan Problem with Anomalous Diffusion” arXiv, 2021 DOI: 10.48550/ARXIV.2111.15600
  • [2] Philippe Benilan and Ronald Gariepy “Strong Solutions in L1 of Degenerate Parabolic Equations” In Journal of Differential Equations 119, 1995 DOI: 10.1006/jdeq.1995.1099
  • [3] Elvise Berchio, Matteo Bonforte, Gabriele Grillo and Matteo Muratori “The Fractional Porous Medium Equation on noncompact Riemannian manifolds” arXiv, 2021 DOI: 10.48550/ARXIV.2109.10732
  • [4] Thomas P. Branson, Luigi Fontana and Carlo Morpurgo “Moser-Trudinger and Beckner-Onofri’s inequalities on the CR sphere” In Annals of Mathematics 177, 2013 DOI: 10.4007/annals.2013.177.1.1
  • [5] Luis Caffarelli, Chi Hin Chan and Alexis Vasseur “Regularity theory for parabolic nonlinear integral operators” In Journal of the American Mathematical Society 24, 2011 DOI: 10.1090/s0894-0347-2011-00698-x
  • [6] Luis Caffarelli and Luis Silvestre “An extension problem related to the fractional laplacian” In Communications in Partial Differential Equations 32, 2007 DOI: 10.1080/03605300600987306
  • [7] Sun Yung Alice Chang and María Mar González “Fractional Laplacian in conformal geometry” In Advances in Mathematics 226, 2011 DOI: 10.1016/j.aim.2010.07.016
  • [8] Xuezhang Chen and Pak Tung Ho “A fractional conformal curvature flow on the unit sphere” arXiv, 2019 DOI: 10.48550/ARXIV.1906.08434
  • [9] M.. Crandall and T.. Liggett “Generation of Semi-Groups of Nonlinear Transformations on General Banach Spaces” In American Journal of Mathematics 93, 1971 DOI: 10.2307/2373376
  • [10] Emmanuele DiBenedetto and Avner Friedman “Hölder estimates for nonlinear degenerate parabolic systems” In Journal für die reine und angewandte Mathematik (Crelles Journal) 1985, 1985 DOI: 10.1515/crll.1985.357.1
  • [11] Sorin Dragomir and Giuseppe Tomassini “Differential Geometry and Analysis on CR Manifolds” Birkhäuser, 2006 DOI: 10.1007/0-8176-4483-0
  • [12] Memoria Ennio De Giorgi “Sulla differenziabilità e l’analiticità delle estremali degli integrali multipli regolari” In Mem. Accad. Sci. Torino. Cl. Sci. Fis. Mat. Nat., 1957
  • [13] Fausto Ferrari and Bruno Franchi “Harnack inequality for fractional sub-Laplacians in Carnot groups” In Mathematische Zeitschrift 279, 2015 DOI: 10.1007/s00209-014-1376-5
  • [14] V Fischer and M Ruzhansky “Quantization on Nilpotent Lie Groups” In Monograph to appear, 2016
  • [15] G.. Folland “Subelliptic estimates and function spaces on nilpotent Lie groups” In Arkiv för Matematik 13, 1975 DOI: 10.1007/BF02386204
  • [16] G.. Folland and E.. Stein “Hardy Spaces on Homogeneous Groups. (MN-28), Volume 28” In Hardy Spaces on Homogeneous Groups. (MN-28), Volume 28, 2020 DOI: 10.2307/j.ctv17db3q0
  • [17] Rupert L. Frank, María Mar González, Dario D. Monticelli and Jinggang Tan “An extension problem for the CR fractional Laplacian” In Advances in Mathematics 270, 2015 DOI: 10.1016/j.aim.2014.09.026
  • [18] Gabriele Grillo, Matteo Muratori and Juan Luis Vázquez “The porous medium equation on Riemannian manifolds with negative curvature. The large-time behaviour” In Advances in Mathematics 314, 2017 DOI: 10.1016/j.aim.2017.04.023
  • [19] Pak Tung Ho and Rong Tang “Fractional Yamabe solitons and fractional Nirenberg problem” In Communications on Pure and Applied Analysis 20, 2021 DOI: 10.3934/cpaa.2021103
  • [20] Tianling Jin and Jingang Xiong “A fractional Yamabe flow and some applications” In Journal für die reine und angewandte Mathematik (Crelles Journal), 2014 DOI: 10.1515/crelle-2012-0110
  • [21] Arturo De Pablo, Fernando Quirós and Ana Rodríguez “Nonlocal filtration equations with rough kernels” In Nonlinear Analysis, Theory, Methods and Applications 137, 2016 DOI: 10.1016/j.na.2016.01.026
  • [22] Arturo De Pablo, Fernando Quirós and Ana Rodríguez “Regularity theory for singular nonlocal diffusion equations” In Calculus of Variations and Partial Differential Equations 57, 2018 DOI: 10.1007/s00526-018-1410-2
  • [23] Arturo De Pablo, Fernando Quirós, Ana Rodríguez and Juan Luis Vázquez “A fractional porous medium equation” In Advances in Mathematics 226, 2011 DOI: 10.1016/j.aim.2010.07.017
  • [24] Arturo De Pablo, Fernando Quirós, Ana Rodríguez and Juan Luis Vázquez “A General Fractional Porous Medium Equation” In Communications on Pure and Applied Mathematics 65, 2012 DOI: 10.1002/cpa.21408
  • [25] Luz Roncal and Sundaram Thangavelu “Hardy’s inequality for fractional powers of the sublaplacian on the Heisenberg group” In Advances in Mathematics 302, 2016 DOI: 10.1016/j.aim.2016.07.010
  • [26] Pablo Raúl Stinga and José Luis Torrea “Extension problem and Harnack’s inequality for some fractional operators” In Communications in Partial Differential Equations 35, 2010 DOI: 10.1080/03605301003735680
  • [27] Peter Tankov “Financial Modelling with Jump Processes” In Financial Modelling with Jump Processes, 2003 DOI: 10.1201/9780203485217
  • [28] Juan Luis Vázquez “The Porous Medium Equation : Mathematical Theory” In The Porous Medium Equation : Mathematical Theory, 2007 DOI: 10.1093/acprof:oso/9780198569039.001.0001

Department of Mathematics, Johns Hopkins University, 3400 N. North Charles, Baltimore, MD 21218

E-mail address: [email protected]