1 Introduction

In the studies of fluid flows, interfacial dynamics is broad and mathematically challenging. Some interfacial problems in fluid dynamics that have been rigorously studied include the water wave problem, the compressible free-boundary Euler equations, the Hele-Shaw problem, the Muskat problem and the Stefan problem. The dynamics of the interface between the fluids strongly depends on properties of the fluids and of the media through which they flow. However, a common feature in all the above problems is that the interface is driven by gravity and surface tension. Gravity is incorporated in the momentum equations as an external force. On the other hand, surface tension balances the pressure jump across the interface (Young-Laplace equation):

$$\begin{aligned} \llbracket p\rrbracket ={\mathfrak {s}}H, \end{aligned}$$
(1.1)

where \(\llbracket p\rrbracket \) is the pressure jump, H is the mean-curvature of the interface, and \({\mathfrak {s}}>0\) is the surface tension coefficient. Well-posedness in Sobolev spaces always holds when surface tension is taken into account but only holds under the Rayleigh-Taylor stability condition on the initial data when surface tension is neglected. It is a natural problem to justify the models without surface tension as the limit of the corresponding full models as surface tension vanishes. This question was addressed in [4, 5, 8, 9, 29, 36, 37] for the problems listed above. The common theory is the following: if the initial data is stable, i.e. it satisfies the Rayleigh-Taylor stability condition, and is sufficiently smooth, then solutions to the problem with surface tension converge to the unique solution of the problem without surface tension locally in time. The general strategy of proof consists of two points.

  1. (i)

    To leading order, the surface tension term \({\mathfrak {s}}H\) provides a regularizing effect. For sufficiently smooth solutions, the difference between \({\mathfrak {s}}H\) and its leading contribution can be controlled by the energy of the problem without surface tension. This yields a uniform time of existence \(T_*\) for the problem with surface tension \({\mathfrak {s}}\rightarrow 0\).

  2. (ii)

    For sufficiently smooth solutions and for some \(\theta \in [0, 1)\), the weighted mean curvature \({\mathfrak {s}}^\theta H\) is uniformly in \({\mathfrak {s}}\) bounded (in some appropriate Sobolev norm) by the energy of the problem without surface tension. It follows that \({\mathfrak {s}}H\), the difference between the two problems, vanishes as \({\mathfrak {s}}^{1-\theta }\) as \({\mathfrak {s}}\rightarrow 0\), establishing the convergence on the time interval \([0, T_*]\). Note that the optimal rate corresponds to \(\theta =0\).

Therefore, the vanishing surface tension limit becomes subtle if the initial data is sufficiently rough so that it can accommodate curvature singularities. As a matter of fact, in the aforementioned works, the initial curvature is at least bounded. In this paper, we prove that for the Muskat problem, the zero surface tension limit can be established for rough initial interfaces whose curvatures are not bounded or even not locally \(L^2\). Regarding quantitive properties of the zero surface tension limit, the convergence rates in the aforementioned works are either unspecified or suboptimal. In this paper, we obtain the optimal convergence rate for the Muskat problem. The next subsections are devoted to a description of the Muskat problem and the statement of our main result.

1.1 The Muskat problem

The Muskat problem [43] concerns the interface evolution between two fluids of densities \(\rho ^\pm \) and viscosities \(\mu ^\pm \) governed by Darcy’s law for flows through porous media. Specifically, the fluids occupy two domains \(\Omega ^\pm = \Omega ^\pm _t \subset {\mathbb {R}}^{d+1}\) separated by an interface \(\Sigma = \Sigma _t\), with \(\Omega ^+\) confined below a rigid boundary \(\Gamma ^+\), and \(\Omega ^-\) likewise above \(\Gamma ^-\). We consider the case when the surfaces \(\Gamma ^\pm \) and \(\Sigma \) are given by the graphs of functions, that is, we designate \(b^\pm : {\mathbb {R}}^{d}_x \rightarrow {\mathbb {R}}\) and \(\eta : {\mathbb {R}}_t\times {\mathbb {R}}^{d}_x \rightarrow {\mathbb {R}}\) for which

$$\begin{aligned} \Sigma&= \{(x,\eta (t,x)) \ : \ x \in {\mathbb {R}}^d\}, \end{aligned}$$
(1.2)
$$\begin{aligned} \Gamma ^\pm&= \{(x,b^\pm (x)) \ : \ x \in {\mathbb {R}}^{d} \}, \end{aligned}$$
(1.3)
$$\begin{aligned} \Omega ^-&= \{(x,y) \in {\mathbb {R}}^{d}\times {\mathbb {R}} \ : \ b^-(x)< y < \eta (t,x) \}, \end{aligned}$$
(1.4)
$$\begin{aligned} \Omega ^+&= \{(x,y) \in {\mathbb {R}}^{d}\times {\mathbb {R}} \ : \eta (t,x)< y < b^+ (x) \}, \end{aligned}$$
(1.5)
$$\begin{aligned} \Omega&= \Omega ^+ \cup \Omega ^-. \end{aligned}$$
(1.6)

We also consider the case where one or both of \(\Gamma ^\pm = \emptyset \). In each domain \(\Omega ^\pm \), the fluid velocity \(u^\pm \) and pressure \(p^\pm \) obey Darcy’s law:

$$\begin{aligned} \mu ^\pm u^\pm + \nabla _{x,y} p^\pm = -\rho ^\pm g \mathbf {e}_{d+1},\quad \mathrm {div}_{x,y} u^\pm =0 \quad \text { in } \Omega ^{\pm }, \end{aligned}$$
(1.7)

where g denotes the gravitational acceleration, and \(\mathbf {e}_{d+1}\) is the upward unit vector in the vertical direction. For any two objects \(A^+\) and \(A^-\) associated with the domains \(\Omega ^+\) and \(\Omega ^-\) respectively, we denote the jump

$$\begin{aligned}{}[\![A ]\!]= A^-- A^+ \end{aligned}$$

whenever this difference is well-defined. In particular, set

$$\begin{aligned} {\mathfrak {g}} =g [\![\rho ]\!]. \end{aligned}$$

At the interface, there are three boundary conditions. First, the normal component of the fluid velocity is continuous across the interface

$$\begin{aligned}{}[\![u \cdot n]\!]=0\quad \text {on}\quad \Sigma , \end{aligned}$$
(1.8)

where we fix n to be the upward normal of the interface, specifically \(n =\langle \nabla \eta \rangle ^{-1} (-\nabla \eta , 1)\) with

$$\begin{aligned} \langle \cdot \rangle = \sqrt{1+|\cdot |^2}. \end{aligned}$$

Second, the interface is transported by the normal fluid velocity, leading to the kinematic boundary condition

$$\begin{aligned} \eta _t = \langle \nabla \eta \rangle u^{-} \cdot n |_{\Sigma _t}. \end{aligned}$$
(1.9)

Third, according to the Young-Laplace equation, the pressure jump is proportional to the mean curvature through surface tension:

$$\begin{aligned}{}[\![p ]\!]= {\mathfrak {s}} H(\eta ) \equiv -{\mathfrak {s}} \text {div}( \langle \nabla \eta \rangle ^{-1}\nabla \eta ) \quad \text {on } \Sigma _t \end{aligned}$$
(1.10)

where \({\mathfrak {s}}\ge 0\) is the surface tension coefficient and

$$\begin{aligned} H(\eta )=-\text {div}( \langle \nabla \eta \rangle ^{-1}\nabla \eta ) \end{aligned}$$
(1.11)

is twice the mean curvature of \(\Sigma \).

Finally, there is no transportation of fluid through the rigid boundaries:

$$\begin{aligned} u^\pm \cdot \nu ^\pm = 0\quad \text {on}\quad \Gamma ^\pm , \end{aligned}$$
(1.12)

where \(\nu ^\pm = \pm \langle \nabla b\rangle ^{-1}(-\nabla b^\pm , 1)\) is the outward normal of \(\Gamma ^\pm \). If \(\Gamma ^\pm = \emptyset \), this condition is replaced by the decay condition

$$\begin{aligned} \lim _{y \rightarrow \pm \infty } u^\pm (x,y)=0. \end{aligned}$$
(1.13)

For the two-phase problem, we have \(\rho ^\pm \) and \(\mu ^\pm \) both as positive quantities. We will also consider the one-phase problem where the top fluid is treated as a vacuum by setting \(\rho ^+ = \mu ^+ = 0\) and \(\Gamma ^+ = \emptyset \).

In the absence of the boundaries \(\Gamma ^\pm \), both the Muskat problems with and without surface tension to leading order admit \(\dot{H}^{1+\frac{d}{2}}({\mathbb {R}}^d)\) as the scaling invariant Sobolev space in view of the scaling

$$\begin{aligned} \eta (x, t)\mapsto \lambda ^{-1}\eta (\lambda x, \lambda ^3 t)\quad \text {and}\quad \eta (x, t)\mapsto \lambda ^{-1}\eta (\lambda x, \lambda t). \end{aligned}$$

In either case, the problem is quasilinear. The literature on well-posedness for the Muskat problem is vast. Early results can be found in [6, 7, 17, 23, 31, 49, 51]. For more recent developments, we refer to [22, 24,25,26, 30, 34] for well-posedness, to [12, 20,21,22, 27, 28, 32, 34] for global existence, and to [13, 14] for singularity formation. Directly related to the problem addressed in the current paper is local well-posedness for low regularity large data. Consider first the problem without surface tension. In [18], the authors obtained well-posedness for \(H^2({\mathbb {T}})\) data for the one-phase problem, allowing for unbounded curvature. For the 2D Muskat problem without viscosity jump, i.e. \(\mu ^+=\mu ^-\), [22] proves well-posedness for data in all subcritical Sobolev spaces \(W^{2, 1+}({\mathbb {R}})\). In \(L^2\)-based Sobolev spaces, [39] obtains well-posedness for data in all subcritical spaces \(H^{\frac{3}{2}+}({\mathbb {R}})\). We also refer to [2] for a generalization of this result to homogeneous Sobolev spaces \(\dot{H}^1({\mathbb {R}})\cap \dot{H}^s({\mathbb {R}})\), \(s\in (\frac{3}{2}, 2)\), allowing non-\(L^2\) data. In [45], local well-posedness for the Muskat problem in the general setting as described above was obtained for initial data in all subcritical Sobolev spaces \(H^{1+\frac{d}{2}+}({\mathbb {R}}^d)\), \(d\ge 1\). The case of one fluid with infinite depth was independently obtained by [3]. Regarding the problem with surface tension, [39, 40] consider initial data in \(H^{2+}({\mathbb {R}})\). In the recent work [46], well-posedness for data in all subcritical Sobolev spaces \( H^{1+\frac{d}{2}+}({\mathbb {R}}^d)\), \(d\ge 1\), was established.

1.2 Main result

In order to state the Rayleigh-Taylor stability condition solely in terms of the interface, we define the operator

$$\begin{aligned} \mathrm {RT}(\eta )=1-[\![{\mathfrak {B}}(\eta ) J (\eta )]\!]\eta \equiv 1-\Big ({\mathfrak {B}}^-(\eta ) J^- (\eta )\eta -{\mathfrak {B}}^+(\eta ) J^+ (\eta )\eta \Big ), \end{aligned}$$
(1.14)

where \(J^\pm (\eta )\) and \(\mathfrak {B}^\pm (\eta )\) are respectively defined by and (2.14) and (2.21) below. Our main result is the following.

Theorem 1.1

Consider either the one-phase Muskat problem or the two-phase Muskat problem in the stable regime \(\rho ^->\rho ^+\). The boundaries \(\Gamma ^\pm \) can be empty or graphs of Lipschitz functions \(b^\pm \in \dot{W}^{1, \infty }({\mathbb {R}}^d)\). For any \(d\ge 1\), let \(s > 1+\frac{d}{2}\) be an arbitrary subcritical Sobolev index. Consider an initial datum \(\eta _0\in H^s({\mathbb {R}}^d)\) satisfying

$$\begin{aligned}&\inf _{x\in {\mathbb {R}}^d} \mathrm {RT}(\eta _0)\ge 2{\mathfrak {a}}>0, \quad {{\,\mathrm{dist}\,}}(\eta _0, \Gamma ^\pm )\ge 2h>0. \end{aligned}$$
(1.15)

Let \({\mathfrak {s}}_n\) be a sequence of surface tension coefficients converging to 0. Then, there exists \(T_*>0\) depending only on \(\Vert \eta _0\Vert _{H^s}\) and \(({\mathfrak {a}}, h, s, \mu ^\pm , {\mathfrak {g}})\) such that the following holds.

(i) The Muskat problems without surface tension and with surface tension \({\mathfrak {s}}_n\) have a unique solution on \([0, T_*]\), denoted respectively by \(\eta \) and \(\eta _n\), that satisfy

$$\begin{aligned}&\eta _n\in C([0, T_*]; H^s({\mathbb {R}}^d))\cap L^2([0, T_*]; H^{s+\frac{3}{2}}({\mathbb {R}}^d)), \end{aligned}$$
(1.16)
$$\begin{aligned}&\eta \in L^\infty ([0, T_*]; H^s({\mathbb {R}}^d))\cap L^2([0, T_*]; H^{s+\frac{1}{2}}({\mathbb {R}}^d))\cap C([0, T_*]; H^{s'}({\mathbb {R}}^d))\quad \forall s'<s, \end{aligned}$$
(1.17)
$$\begin{aligned}&\Vert (\eta _n, \eta )\Vert _{L^\infty ([0,T_*];H^s)}+ \Vert (\eta _n, \eta )\Vert _{L^2([0,T_*];H^{s+\frac{1}{2}})} + \sqrt{{\mathfrak {s}}} \Vert \eta _n\Vert _{L^2([0,T_*];H^{s+\frac{3}{2}})} \nonumber \\&\quad \le {\mathcal {F}}(\Vert \eta _0\Vert _{H^s}, {\mathfrak {a}}^{-1}), \end{aligned}$$
(1.18)
$$\begin{aligned}&\inf _{t\in [0, T_*]}\inf _{x\in {\mathbb {R}}^d} \mathrm {RT}(\eta _n(t))>{\mathfrak {a}}, \quad \inf _{t\in [0, T_*]}{{\,\mathrm{dist}\,}}(\eta _n(t), \Gamma ^\pm )>h,\end{aligned}$$
(1.19)
$$\begin{aligned}&\inf _{t\in [0, T_*]}\inf _{x\in {\mathbb {R}}^d} \mathrm {RT}(\eta (t))>{\mathfrak {a}}, \quad \inf _{t\in [0, T_*]}{{\,\mathrm{dist}\,}}(\eta (t), \Gamma ^\pm )> h, \end{aligned}$$
(1.20)

where \({\mathcal {F}}:{\mathbb {R}}^+\times {\mathbb {R}}^+\rightarrow {\mathbb {R}}^+\) is nondecreasing and depends only on \((h, s, \mu ^\pm , {\mathfrak {g}})\).

(ii) As \(n\rightarrow \infty \), \(\eta _n\) converges to \(\eta \) on \([0, T_*]\) with the rate \(\sqrt{{\mathfrak {s}}_n}\):

$$\begin{aligned} \Vert \eta _n-\eta \Vert _{L^\infty ([0,T_*];H^{s-1})}+\Vert \eta _n- \eta \Vert _{L^2([0,T_*];H^{s-\frac{1}{2}})} \le \sqrt{{\mathfrak {s}}_n} {\mathcal {F}}(\Vert \eta _0\Vert _{H^s}, {\mathfrak {a}}^{-1}). \end{aligned}$$
(1.21)

If in addition \(s\ge 2\), then we have the convergence with optimal rate \({\mathfrak {s}}_n\):

$$\begin{aligned} \Vert \eta _n-\eta \Vert _{L^\infty ([0,T_*];H^{s-2})}+\Vert \eta _n -\eta \Vert _{L^2([0,T_*];H^{s-\frac{3}{2}})} \le {{\mathfrak {s}}_n} {\mathcal {F}}(\Vert \eta _0\Vert _{H^s}, {\mathfrak {a}}^{-1}). \end{aligned}$$
(1.22)

The convergence (1.21) holds for initial data in any subcritical Sobolev spaces \(H^{1+\frac{d}{2}+}({\mathbb {R}}^d)\). In particular, this allows for initial interfaces whose curvatures are unbounded in all dimensions and not locally square integrable in one dimension. The former is because \(H(\eta _0)\in H^{-1+\frac{d}{2}+}({\mathbb {R}}^d)\not \subset L^\infty ({\mathbb {R}}^d)\) and latter is due to the fact that in one dimension we have \(H(\eta _0)\in H^{-\frac{1}{2}+\varepsilon }({\mathbb {R}})\not \subset L^2_{loc}({\mathbb {R}})\). This appears to be the first result on vanishing surface tension that can accommodate curvature singularities of the initial interface. On the other hand, the convergence (1.22) has optimal rate \({\mathfrak {s}}_n\) and holds under the additional condition that \(s\ge 2\). This is only a condition in one dimension since \(s>1+\frac{d}{2}\ge 2\) for \(d\ge 2\). Note also that \(s \ge 2\) is the minimal regularity to ensure that the initial curvature is square integrable, yet it still allows for unbounded curvature. See the technical Remark 1.3.

The proof of Theorem 1.1 exploits the Dirichlet-Neumann reformulation [45, 46] for the Muskat problem in a general setting. See also [3] for the one-fluid case. Part (i) of Theorem 1.1 is a uniform local well-posedness with repsect to surface tension. The key tool in proving this is paralinearization results for the Dirichlet-Neumann operator taken from [1, 45]. The convergences (1.21) and (1.22) rely on contraction estimates for the Dirichlet-Neumann operator proved in [45] for a large Sobolev regularity range of the Dirichlet data. Together with [45] and [46], Theorem 1.1 provides a rather complete local regularity theory for (large) subcritical data.

Remark 1.2

In [8] and [9], the first results on the zero surface tension limit for Muskat were obtained respectively in 2D and 3D for smooth initial data, i.e. \(\Sigma _0\in H^{s_0}\) for some sufficiently large \(s_0\). The interface is not necessarily a graph but if it is then the convergence estimates therein translate into

$$\begin{aligned} \Vert \eta _n-\eta \Vert _{L^\infty ([0, T_*]; H^1({\mathbb {T}}^d))}\le C\sqrt{{\mathfrak {s}}_n},\quad d=1, 2, \end{aligned}$$
(1.23)

which has the same rate as (1.21).

Remark 1.3

The condition \(s\ge 2\) in (1.22) is due to the control of low frequencies in the paralinearization and contraction estimates for the Dirichlet-Neumann operator \(G(\eta )f\) (see (2.8) for its definition). Precisely, the best currently available results (see Sects. 2.4 and 2.5 below) require \(f\in H^\sigma ({\mathbb {R}}^d)\) with \(\sigma \ge \frac{1}{2}\). The proof of the \(L^\infty _t H^{s-2}_x\) convergence in (1.22) appeals to these results with \(\sigma =s-\frac{3}{2}\).

Remark 1.4

It was proved in [45] that the Rayleigh-Taylor condition holds unconditionally in the following configurations:

  • the one-phase problem without bottom or with Lipschitz bottoms;

  • the two-phase problem with constant viscosity (\(\mu ^+=\mu ^-\)).

When the Rayleigh-Taylor condition is violated, analytic solutions to the problem without surface tension exist [49]. The works [47, 48] and [15, 16] strongly indicate that these solutions are not limits of solutions to the problem with surface tension. We also refer to [35] for the instability of the trivial solution (with surface tension) and to [33] for the stability of bubbles (without surface tension).

Remark 1.5

In Theorem 1.1, the initial data is fixed for all surface tension coefficients \({\mathfrak {s}}_n\). In general, one can consider \(\eta _n\vert _{t=0}=\eta _{n, 0}\) uniformly bounded bounded in \(H^s({\mathbb {R}}^d)\) such that the conditions in (1.15) hold uniformly in n. Then, for any \(s>1+\frac{d}{2}\), we have

$$\begin{aligned}&\Vert \eta _n-\eta \Vert _{L^\infty ([0,T_*];H^{s-1})}+\Vert \eta _n -\eta \Vert _{L^2([0,T_*];H^{s-\frac{1}{2}})}\nonumber \\&\quad \le \big (\sqrt{{\mathfrak {s}}_n} +\Vert \eta _{n, 0}-\eta _0\Vert _{H^{s-1}}\big ){\mathcal {F}}(\Vert \eta _0\Vert _{H^s}, {\mathfrak {a}}^{-1}). \end{aligned}$$
(1.24)

On the other hand, if \(s>1+\frac{d}{2}\) and \(s\ge 2\) then

$$\begin{aligned} \Vert \eta _n-\eta \Vert _{L^\infty ([0,T_*];H^{s-2})}+\Vert \eta _n -\eta \Vert _{L^2([0,T_*];H^{s-\frac{3}{2}})} \le \big ({{\mathfrak {s}}_n}+\Vert \eta _{n, 0}-\eta _0\Vert _{H^{s-2}}\big ) {\mathcal {F}}(\Vert \eta _0\Vert _{H^s}, {\mathfrak {a}}^{-1}). \end{aligned}$$
(1.25)

Remark 1.6

By interpolating the convergence estimate (1.21) and the uniform bounds in (1.18), we obtain the vanishing surface tension limit in \(H^{s'}\) for all \(s'\in [s-1, s)\):

$$\begin{aligned} \Vert \eta _n-\eta \Vert _{L^\infty ([0,T_*];H^{s'})}+ \Vert \eta _n-\eta \Vert _{L^2([0,T_*];H^{s'+\frac{1}{2}})} \le {\mathfrak {s}}_n^{\frac{s-s'}{2}} {\mathcal {F}}(\Vert \eta _0\Vert _{H^s}, {\mathfrak {a}}^{-1}). \end{aligned}$$
(1.26)

Convergence in the highest regularity \(L^\infty _tH^s_x\) is more subtle and can possibly be established using the Bona-Smith type argument [10]. This would imply in particular that \(\eta \) is continuous in time with values in \(H^s_x\), \(\eta \in C_t H^s_x\). In the context of vanishing viscosity limit, this question was addressed in [41], while convergence in lower Sobolev spaces (compared to initial data) was proved in [19]. For gravity water waves, the Bona-Smith type argument was applied in [44] to establish the continuity of the flow map in the highest regularity.

Remark 1.7

The proof of the local well-posedness in all subcritical Sobolev spaces in [45] uses a parabolic regularization. Theorem 1.1 provides an alternative proof via regularization by vanishing surface tension. We stress that the assertions about \(\eta \) in Theorem 1.1 do not make use of the local well-posedness results in [45].

The rest of this paper is organized as follows. In Sect. 2, we recall the reformulation of the Muskat problem in terms of the Dirichlet-Neumann operator together with results on the Dirichlet-Neumann operator established in [1, 45]. Section 3 is devoted to the proof of uniform-in-\({\mathfrak {s}}\) a priori estimates. In Sect. 4, we prove contraction estimates for the operators \(J^\pm \) which arise in the reformulation of the two-phase problem. The proof of Theorem 1.1 is given in Sect. 5. Finally, in “Appendix 5”, we recall the symbolic paradifferential calculus and the Gåarding inequality for paradifferential operators.

2 Preliminaries

2.1 Big-O notation

If X and Y are Banach spaces and \(T: X\rightarrow Y\) is a bounded linear operator, \(T\in {\mathcal {L}}(X, Y)\), with operator norm bounded by A, we write

$$\begin{aligned} T=O_{X\rightarrow Y}(A). \end{aligned}$$

We define the space of operators of order \(m \in {\mathbb {R}}\) in the scale of Sobolev spaces \(H^s({\mathbb {R}}^d)\):

$$\begin{aligned} Op^m\equiv Op^m({\mathbb {R}}^d) = \bigcap _{s \in {\mathbb {R}}} {\mathcal {L}}(H^s({\mathbb {R}}^d),H^{s-m}({\mathbb {R}}^d)). \end{aligned}$$

We shall write \(T=O_{Op^m}(A)\) when \(T\in Op^m\) and for all \(s\in {\mathbb {R}}\), there exists \(C=C(s)\) such that \(T=O_{H^{s}\rightarrow H^{s-m}}(CA)\).

2.2 Function spaces

In the general setting, to reformulate the dynamics of the Muskat problem solely in terms of the interface, we require some function spaces.

We shall always assume that \(\eta \in W^{1, \infty }({\mathbb {R}}^d)\) and either \(\Gamma ^\pm =\emptyset \) or \(b^\pm \in \dot{W}^{1, \infty }({\mathbb {R}}^d)\) with \({{\,\mathrm{dist}\,}}(\eta , b^\pm )>0\). Recall that the fluid domains \(\Omega ^\pm \) are given in (1.2). Define

$$\begin{aligned} \dot{H}^1(\Omega ^\pm )=\{v\in L^1_{loc}(\Omega ^\pm ): \nabla _{x, y}v\in L^2(\Omega ^\pm )\}/{\mathbb {R}},\quad \Vert v\Vert _{\dot{H}^1(\Omega ^\pm )}=\Vert \nabla _{x, y}v\Vert _{L^2(\Omega ^\pm )}. \end{aligned}$$
(2.1)

For any \(\sigma \in {\mathbb {R}}\), we define the ‘slightly homogeneous’ Sobolev space

$$\begin{aligned} H^{1,\sigma }({\mathbb {R}}^d) =\left\{ f\in L^1_{loc}({\mathbb {R}}^d): \nabla f \in H^{\sigma -1}({\mathbb {R}}^d)\right\} /{\mathbb {R}},\quad \Vert f\Vert _{H^{1,\sigma }({\mathbb {R}}^d)}=\Vert \nabla f\Vert _{H^{\sigma -1}({\mathbb {R}}^d)}. \end{aligned}$$
(2.2)

When \(b^\pm \in \dot{W}^{1, \infty }({\mathbb {R}}^d)\), we fix an arbitrary number \(a\in (0, 1)\) and define the ‘screened’ fractional Sobolev spaces

(2.3)

According to Proposition 3.2 [45], the spaces are independent of \(\eta \) in \(W^{1, \infty }({\mathbb {R}}^d)\) satisfying \({{\,\mathrm{dist}\,}}(\eta , b^\pm )>h>0\). Thus, we can set

(2.4)

It was proved in [38, 50] that there exist unique continuous trace operators

$$\begin{aligned} \mathrm {Tr}_{\Omega ^\pm \rightarrow \Sigma }: \dot{H}^1(\Omega ^\pm )\rightarrow \widetilde{H}^\frac{1}{2}_\pm (\Sigma )\equiv \widetilde{H}^\frac{1}{2}_\pm ({\mathbb {R}}^d) \end{aligned}$$
(2.5)

with norm depending only on \(\Vert \eta \Vert _{\dot{W}^{1, \infty }({\mathbb {R}}^d)}\) and \(\Vert b^\pm \Vert _{\dot{W}^{1, \infty }({\mathbb {R}}^d)}\).

Finally, for \(\sigma > \frac{1}{2}\), we define

$$\begin{aligned} \widetilde{H}^\sigma _\pm ({\mathbb {R}}^d)=\widetilde{H}^\frac{1}{2}_\pm ({\mathbb {R}}^d)\cap H^{1, \sigma }({\mathbb {R}}^d) \end{aligned}$$
(2.6)

and equip it with the norm \(\Vert \cdot \Vert _{\widetilde{H}_\pm ^\sigma }=\Vert \cdot \Vert _{\widetilde{H}^\frac{1}{2}_\pm }+\Vert \cdot \Vert _{H^{1, \sigma }}\).

2.3 Dirichlet–Neumann formulation

Given a function \(f \in {\mathbb {R}}^d\), let \(\phi \) solve the Laplace equation

$$\begin{aligned} {\left\{ \begin{array}{ll} \Delta _{x,y} \phi = 0 \quad \text {in } \Omega ^-, \\ \phi = f \quad \text {on } \Sigma , \\ \frac{\partial \phi }{\partial \nu ^-} =0 \quad \text {on } \Gamma ^-, \end{array}\right. } \end{aligned}$$
(2.7)

with the final condition replaced by decay at infinity of \(\phi \) if \(\Gamma ^- = \emptyset \). Then, we define the (rescaled) Dirichlet-Neumann operator \(G^- \equiv G^-(\eta )\) by

$$\begin{aligned} G^- f = \langle \nabla \eta \rangle \frac{\partial \phi ^-}{\partial n}. \end{aligned}$$
(2.8)

The operator \(G^+(\eta )\) for the top fluid domain \(\Omega ^+\) is defined similarly. The solvability of (2.7) is given in the next proposition.

Proposition 2.1

([45] Propositions 3.4 and 3.6). Assume that either \(\Gamma ^-=\emptyset \) or \(b^-\in \dot{W}^{1, \infty }({\mathbb {R}}^d)\). If \(\eta \in W^{1, \infty }({\mathbb {R}}^d)\) and \({{\,\mathrm{dist}\,}}(\eta , \Gamma ^-)>h>0\), then for every \(f\in \widetilde{H}^\frac{1}{2}_-({\mathbb {R}}^d)\) there exists a unique variational solution \(\phi ^-\in \dot{H}^1(\Omega ^-)\) to (2.7). Precisely, \(\phi \) satisfies \(\mathrm {Tr}_{\Omega ^-\rightarrow \Sigma }\phi =f\),

$$\begin{aligned} \int _{\Omega ^-}\nabla _{x, y}\phi \cdot \nabla _{x, y}\varphi dxdy=0\quad \forall \varphi \in \big \{v\in \dot{H}^1(\Omega ^-):~\mathrm {Tr}_{\Omega ^-\rightarrow \Sigma }v=0\big \}, \end{aligned}$$
(2.9)

together with the estimate

$$\begin{aligned} \Vert \nabla _{x, y}\phi \Vert _{L^2(\Omega ^-)}\le {\mathcal {F}}(\Vert \nabla \eta \Vert _{L^\infty })\Vert f\Vert _{\widetilde{H}^\frac{1}{2}_-({\mathbb {R}}^d)} \end{aligned}$$
(2.10)

for some \({\mathcal {F}}: {\mathbb {R}}^+\rightarrow {\mathbb {R}}^+\) depending only on h and \(\Vert \nabla _xb^-\Vert _{L^\infty ({\mathbb {R}}^d)}\).

As the functions \(b^\pm \) are fixed in \(\dot{W}^{1, \infty }({\mathbb {R}}^d)\), we shall omit the dependence on \(\Vert \nabla _xb^-\Vert _{L^\infty ({\mathbb {R}}^d)}\) in various estimates below.

The Muskat problem can be reformulated in terms of \(G^\pm \) as follows.

Proposition 2.2

([46] Proposition 1.1). (i) If \((u, p, \eta )\) solve the one-phase Muskat problem then \(\eta :{\mathbb {R}}^d\rightarrow {\mathbb {R}}\) obeys the equation

$$\begin{aligned} \partial _t\eta =-\frac{1}{\mu ^-} G^{-}(\eta )\big (g\rho ^-\eta +{\mathfrak {s}}H(\eta )\big ). \end{aligned}$$
(2.11)

Conversely, if \(\eta \) is a solution of (2.11) then the one-phase Muskat problem has a solution which admits \(\eta \) as the free surface.

(ii) If \((u^\pm , p^\pm , \eta )\) is a solution of the two-phase Muskat problem then

$$\begin{aligned} \partial _t\eta =-\frac{1}{\mu ^-}G^{-}(\eta )f^-, \end{aligned}$$
(2.12)

where \(f^\pm :=p^\pm \vert _{\Sigma }+\rho ^\pm g \eta \) satisfy

$$\begin{aligned} {\left\{ \begin{array}{ll} f^--f^+={\mathfrak {g}}\eta +{\mathfrak {s}}H(\eta ),\\ \frac{1}{\mu ^-}G^-(\eta )f^--\frac{1}{\mu ^+}G^+(\eta )f^+=0. \end{array}\right. } \end{aligned}$$
(2.13)

Conversely, if \(\eta \) is a solution of (2.12) where \(f^\pm \) solve (2.13) then the two-phase Muskat problem has a solution which admits \(\eta \) as the free interface.

To make use of the results on the Dirichlet-Neumann operator established in [1, 45, 46], it is convenient to introduce the linear operators

$$\begin{aligned} J ^\pm = J ^\pm (\eta ):\quad v \mapsto f^\pm \end{aligned}$$
(2.14)

where \(f^\pm :{\mathbb {R}}^d \rightarrow {\mathbb {R}}\) are solutions to the system

$$\begin{aligned} \begin{pmatrix} \text {Id} &{} -\mathrm {Id}\\ \mu ^+ G^-(\eta ) &{} -\mu ^- G^+(\eta ) \end{pmatrix} \begin{pmatrix} f^- \\ f^+ \end{pmatrix} = \begin{pmatrix} v \\ 0 \end{pmatrix}. \end{aligned}$$
(2.15)

Introduce

$$\begin{aligned} L = L(\eta ) = \frac{\mu ^+ + \mu ^-}{\mu ^-}G^- J ^-. \end{aligned}$$
(2.16)

For the two-phase case, L coincides with \(\frac{\mu ^+ + \mu ^-}{\mu ^+}G^+ J ^+\) in view of (2.13). Thus, writing \(L = (\frac{\mu ^-}{\mu ^+ + \mu ^-} +\frac{\mu ^+}{\mu ^+ + \mu ^-})L\) yields the symmetric formula

$$\begin{aligned} L = G^- J ^- + G^+ J ^+. \end{aligned}$$
(2.17)

This formula holds for the one phase problem (2.11) as well. Indeed, when \(\mu ^+ = 0\) we have \( J ^+ = 0\) and \( J ^- = \text {Id}\), and hence \(L = G^-\). In view of (2.16), Proposition 2.2 implies the following.

Lemma 2.3

The Muskat problem (both one-phase and two-phase) is equivalent to

$$\begin{aligned} \partial _t\eta +\frac{1}{\mu ^+ + \mu ^-}L(\eta ) ({\mathfrak {g}} \eta + {\mathfrak {s}} H(\eta ))= 0. \end{aligned}$$
(2.18)

The next proposition gathers results on the existence and boundedness of the operators \(J ^\pm \), \(G^\pm \), and L in Sobolev spaces.

Proposition 2.4

([1] Theorem 3.12, and [45] Propositions 3.8, 4.8, 4.10 and Remark 4.9). Let \(\mu ^+ \ge 0\) and \(\mu ^- > 0\). Assume \({{\,\mathrm{dist}\,}}(\eta , \Gamma ^\pm )>h>0\).

(i) If \(\eta \in W^{1, \infty }({\mathbb {R}}^d)\) then

$$\begin{aligned} \Vert J ^\pm \Vert _{{\mathcal {L}}(H^{\frac{1}{2}},\widetilde{H}^{\frac{1}{2}}_\pm )} + \Vert G^\pm \Vert _{{\mathcal {L}}(\widetilde{H}^{\frac{1}{2}}_\pm , H^{-\frac{1}{2}})} + \Vert L\Vert _{{\mathcal {L}}(H^{\frac{1}{2}},H^{-\frac{1}{2}})}\le {\mathcal {F}}(\Vert \eta \Vert _{W^{1, \infty }}), \end{aligned}$$
(2.19)

where \({\mathcal {F}}\) is nondecreasing and depends only on \((h, \mu ^\pm )\).

(ii) If \(\eta \in H^s({\mathbb {R}}^d)\) with \(s > 1+\frac{d}{2}\) then for any \(\sigma \in [\frac{1}{2}, s]\), we have

$$\begin{aligned} \Vert J ^\pm \Vert _{{\mathcal {L}}(H^{\sigma },\widetilde{H}^{\sigma }_\pm )}+ \Vert G^\pm \Vert _{{\mathcal {L}}(\widetilde{H}^{\sigma }_\pm , H^{\sigma -1})} +\Vert L\Vert _{{\mathcal {L}}(H^{\sigma },H^{\sigma -1})} \le {\mathcal {F}}(\Vert \eta \Vert _{H^s}), \end{aligned}$$
(2.20)

where \({\mathcal {F}}\) is nondecreasing and depends only on \((h, \mu ^\pm , s, \sigma )\).

2.4 Paralinearization

Given a function f, define the operators

$$\begin{aligned} {{\mathfrak {B}}}^\pm f\equiv {{\mathfrak {B}}}^\pm (\eta )f&= \langle \nabla \eta \rangle ^{-2}(\nabla \eta \cdot \nabla + G^\pm (\eta ))f,\end{aligned}$$
(2.21)
$$\begin{aligned} {{\mathfrak {V}}}^\pm f \equiv {{\mathfrak {V}}}^\pm (\eta )f&=(\nabla - \nabla \eta {{\mathfrak {B}}}^\pm )f. \end{aligned}$$
(2.22)

We note here that \({{\mathfrak {B}}}^\pm f = \partial _y \phi ^\pm |_\Sigma \) and \({{\mathfrak {V}}}^\pm f= \nabla _x \phi ^\pm |_{\Sigma }\), where \(\phi \) solves (2.7). Moreover, as a consequence of (2.20) and product rules, we have

$$\begin{aligned} \Vert {{\mathfrak {B}}}^\pm (\eta )\Vert _{\widetilde{H}^\sigma _\pm \rightarrow H^{\sigma -1}}+\Vert {{\mathfrak {V}}}^\pm (\eta )\Vert _{\widetilde{H}^\sigma _\pm \rightarrow H^{\sigma -1}}\le {\mathcal {F}}(\Vert \eta \Vert _{H^s}) \end{aligned}$$
(2.23)

for all \(s>1+\frac{d}{2}\) and \(\sigma \in [\frac{1}{2}, s]\).

The principal symbol of \(G^-(\eta )\) is

$$\begin{aligned} \lambda (x, \xi ) = \sqrt{\langle \nabla \eta \rangle ^2 |\xi |^2 - (\nabla \eta \cdot \xi )^2}, \end{aligned}$$
(2.24)

while that of \(G^+(\eta )\) is \(-\lambda (x, \xi )\). Note that \(\lambda (x, \xi ) \ge |\xi |\) with equality when \(d=1\). Next we record results on paralinearization of \(G^\pm (\eta )\).

Theorem 2.5

([1] Propostion 3.13, [45] Theorem 3.18). Let \(s>1+\frac{d}{2}\) and let \(\delta \in (0,\frac{1}{2}]\) satisfy \(\delta <s-1-\frac{d}{2}\). Assume that \(\eta \in H^s\) and \({{\,\mathrm{dist}\,}}(\eta , \Gamma ^\pm )>h>0\).

(i) For any \(\sigma \in [\frac{1}{2}, s-\delta ]\), there exists a nondecreasing function \({\mathcal {F}}\) depending only on \((h, s,\sigma ,\delta )\) such that

(2.25)
$$\begin{aligned} L(\eta ) = T_\lambda +O_{H^\sigma \rightarrow H^{\sigma - 1+\delta }}( {\mathcal {F}}(\Vert \eta \Vert _{H^s})). \end{aligned}$$
(2.26)

(ii) For any \(\sigma \in [\frac{1}{2}, s]\), there exists a nondecreasing function \({\mathcal {F}}\) depending only on \((h, s,\sigma ,\delta )\) such that for all \(f \in {\tilde{H}}^\sigma _\pm ({\mathbb {R}}^d)\), we have

(2.27)
$$\begin{aligned} L(\eta )f = T_\lambda (f - T_{[\![{{\mathfrak {B}}} J ]\!]f}\eta ) - T_{[\![{{\mathfrak {V}}} J ]\!]f}\cdot \nabla \eta + O_{H^\sigma \rightarrow H^{\sigma - \frac{1}{2}}}({\mathcal {F}}(\Vert \eta \Vert _{H^s})(1 + \Vert \eta \Vert _{H^{s + \frac{1}{2}-\delta }}))f. \end{aligned}$$
(2.28)

Proof

(2.25) was proved in Proposition 3.13 in [1] for \(\sigma \in [\frac{1}{2}, s-\frac{1}{2}]\) but its proof allows for \(\sigma \in [\frac{1}{2}, s-\delta ]\). On the other hand, (2.27) was proved in Theorem 3.17 in [45]. Let us prove (2.26) and (2.28). We recall from (2.17) that \(L= G^- J ^-+G^+J^+\). Applying (2.25) we have

By virtue of (2.20) we have \(\Vert J ^\pm \Vert _{{\mathcal {L}}(H^{\sigma },\widetilde{H}^{\sigma }_\pm )}\le {\mathcal {F}}(\Vert \eta \Vert _{H^s})\), and thus

It follows that

$$\begin{aligned} L(\eta )f&=T_\lambda [\![J]\!]f + O_{ H^\sigma \rightarrow H^{\sigma - 1 + \delta }}({\mathcal {F}}(\Vert \eta \Vert _{H^s}))f\\&=T_\lambda f + O_{ H^\sigma \rightarrow H^{\sigma - 1 + \delta }}({\mathcal {F}}(\Vert \eta \Vert _{H^s}))f, \end{aligned}$$

where in the second equality we have used the fact that \( [\![J]\!]=\mathrm {Id}\). This completes the proof of (2.26). Finally, (2.28) can be proved similarly upon using the paralinearizaion (2.27). \(\quad \square \)

Finally, the mean curvature operator \(H(\cdot )\), defined by (1.11), can be paralinearized as follows.

Proposition 2.6

([46] Proposition 3.1). Let \(s > 1 + \frac{d}{2}\) and let \(\delta \in (0,\frac{1}{2}]\) satisfying \(\delta <s-1-\frac{d}{2}\). Then there exists a nondecreasing function \({\mathcal {F}}\) depending only on s such that

$$\begin{aligned} H(\eta ) = T_{l}\eta + O_{H^{s+\frac{3}{2}} \rightarrow H^{s-\frac{1}{2}+ \delta }}({\mathcal {F}}(\Vert \eta \Vert _{H^s}) )\eta \end{aligned}$$
(2.29)

where

$$\begin{aligned} l = \langle \nabla \eta \rangle ^{-3}\lambda ^2. \end{aligned}$$
(2.30)

In addition, if \(\sigma \ge -1\) then

$$\begin{aligned} \Vert H(\eta )\Vert _{H^\sigma } \le {\mathcal {F}}(\Vert \eta \Vert _{H^s})\Vert \eta \Vert _{H^{\sigma +2}}. \end{aligned}$$
(2.31)

2.5 Contraction estimates

Let \(s>1+\frac{d}{2}\) and consider \(\eta _j\in H^s({\mathbb {R}}^d)\) satisfying \({{\,\mathrm{dist}\,}}(\eta _j, \Gamma ^\pm )>h>0\), \(j=1, 2\). We have the following contraction estimates for \(G^\pm (\eta _1)-G^\pm (\eta _2)\).

Theorem 2.7

([45] Corollary 3.25 and Proposition 3.31). For any \(\sigma \in [\frac{1}{2}, s]\), there exists a nondecreasing function \({\mathcal {F}}:{\mathbb {R}}^+\rightarrow {\mathbb {R}}^+\) depending only on \((h, s, \sigma )\) such that

$$\begin{aligned} \Vert G^\pm (\eta _1)-G^\pm (\eta _2) \Vert _{\widetilde{H}^s_\pm \rightarrow H^{\sigma -1}}\le {\mathcal {F}}\big (\Vert (\eta _1, \eta _2)\Vert _{H^s}\big )\Vert \eta _1-\eta _2\Vert _{H^\sigma } \end{aligned}$$
(2.32)

and

$$\begin{aligned} \Vert G^\pm (\eta _1)-G^\pm (\eta _2) \Vert _{\widetilde{H}^\sigma _\pm \rightarrow H^{\sigma -1}}\le {\mathcal {F}}\big (\Vert (\eta _1, \eta _2) \Vert _{H^s}\big )\Vert \eta _1-\eta _2\Vert _{H^s}. \end{aligned}$$
(2.33)

Theorem 2.8

([45] Theorem 3.24). Let \(\delta \in (0, \frac{1}{2}]\) satisfy \(\delta <s-1-\frac{d}{2}\). Let \(\sigma \in [\frac{1}{2}+\delta , s]\). For any \(f \in \widetilde{H}^s_\pm \), there exists a nondecreasing function \({\mathcal {F}}\) depending only on \((h, s, \sigma )\) such that

(2.34)

where \(\lambda _1\) is defined by (2.24) with \(\eta =\eta _1\).

3 Uniform A Priori Estimates

Conclusion (i) in Theorem 1.1 concerns the uniform local well-posedness of the Muskat problem with surface tension. Key to that is the following a priori estimates that are uniform in the vanishing surface tension limit \({\mathfrak {s}}\rightarrow 0\).

Proposition 3.1

Let \(s > 1+\frac{d}{2}\), \(\mu ^- > 0\), \(\mu ^+ \ge 0\), \({\mathfrak {s}} > 0\), and \(h>0\). Suppose

$$\begin{aligned} \eta \in C([0, T]; H^s({\mathbb {R}}^d))\cap L^2([0, T]; H^{s+\frac{3}{2}}({\mathbb {R}}^d)) \end{aligned}$$
(3.1)

is a solution to (2.18) with initial data \(\eta _0\in H^s({\mathbb {R}}^d)\) such that

$$\begin{aligned}&\inf _{t\in [0, T]}\inf _{x\in {\mathbb {R}}^d} \mathrm {RT}(\eta (t))> {\mathfrak {a}}>0, \end{aligned}$$
(3.2)
$$\begin{aligned}&\inf _{t\in [0, T]}{{\,\mathrm{dist}\,}}(\eta (t), \Gamma ^\pm )> h. \end{aligned}$$
(3.3)

Then, there exists a nondecreasing function \({\mathcal {F}}:{\mathbb {R}}^+\times {\mathbb {R}}^+\rightarrow {\mathbb {R}}^+\) depending only on \((h, s, \mu ^\pm )\) such that

$$\begin{aligned} \begin{aligned}&\Vert \eta \Vert _{L^\infty ([0,T];H^s)} \le \Vert \eta _0\Vert _{H^s} \exp \Big (({\mathfrak {s}} + {\mathfrak {g}})T{\mathcal {F}}\big (\Vert \eta \Vert _{L^\infty ([0,T];H^s)}, {\mathfrak {a}}^{-1}\big )\Big ) \end{aligned} \end{aligned}$$
(3.4)

and

$$\begin{aligned}&{\mathfrak {s}} \Vert \eta \Vert _{L^2([0,T];H^{s+\frac{3}{2}})}^2 +{\mathfrak {g}} \Vert \eta \Vert _{L^2([0,T];H^{s+\frac{1}{2}})}^2\nonumber \\&\quad \le {\mathcal {F}}_1\left( \Vert \eta _0\Vert _{H^s} \exp \Big (({\mathfrak {s}} + {\mathfrak {g}})T{\mathcal {F}} \big (\Vert \eta \Vert _{L^\infty ([0,T];H^s)},{\mathfrak {a}}^{-1}\big ) \Big ),{\mathfrak {a}}^{-1}\right) \end{aligned}$$
(3.5)

where \({\mathcal {F}}_1(m, n)=m^2{\mathcal {F}}(m, n)\).

Proof

Set \( B= [\![{\mathfrak {B}}(\eta ) J (\eta )]\!]\eta \) and \(V= [\![{\mathfrak {V}}(\eta ) J (\eta )]\!]\eta \). We shall write \({\mathcal {Q}}={\mathcal {Q}}(t)={\mathcal {F}}(\Vert \eta (t)\Vert _{H^s}, {\mathfrak {a}}^{-1})\) when \({\mathcal {F}}(\cdot , \cdot )\) is nondecreasing and depends only on \((h, s, \mu ^\pm )\). Note that \({\mathcal {F}}\) may change from line to line. From (2.18) we have

$$\begin{aligned} (\mu ^+ + \mu ^-)\partial _t\eta +{\mathfrak {g}}L(\eta )\eta + {\mathfrak {s}}L(\eta )H(\eta )= 0 \end{aligned}$$
(3.6)

for both the one-phase and two-phase problems. Fix \(\delta \in \big (0, \min (\frac{1}{2}, s-1-\frac{d}{2})\big )\). By virtue of the paralinearization (2.28) (with \(\sigma =s\)) we have

$$\begin{aligned} {\mathfrak {g}}L(\eta )\eta ={\mathfrak {g}} (T_\lambda (\eta - T_{B}\eta ) -T_V\cdot \nabla \eta )+O_{H^s \rightarrow H^{s-\frac{1}{2}}}\big ({\mathfrak {g}} {\mathcal {Q}}(1+\Vert \eta \Vert _{H^{s+\frac{1}{2}-\delta }})\big )\eta . \end{aligned}$$

On the other hand, (2.26) (with \(\sigma =s-\frac{1}{2}\)) together with (2.31) gives

$$\begin{aligned} {\mathfrak {s}}L(\eta )H(\eta )={\mathfrak {s}}T_\lambda H(\eta )+O_{H^{s+\frac{3}{2}}\rightarrow H^{s-\frac{3}{2}+\delta }}({\mathfrak {s}}{\mathcal {F}}(\Vert \eta \Vert _{H^s}))\eta \end{aligned}$$

Combining this with the linearization (2.29) for \(H(\eta )\) yields

$$\begin{aligned} {\mathfrak {s}}L(\eta )H(\eta )={\mathfrak {s}}T_\lambda T_l\eta +O_{H^{s+\frac{3}{2}}\rightarrow H^{s-\frac{3}{2}+\delta }}({\mathfrak {s}}{\mathcal {F}}(\Vert \eta \Vert _{H^s}))\eta , \end{aligned}$$

where we have applied Theorem A.2 to have \(T_\lambda =O_{Op^1}({\mathcal {F}}(\Vert \eta \Vert _{H^s}))\).

Then in view of (3.6) we obtain

$$\begin{aligned} \begin{aligned}&(\mu ^+ + \mu ^-)\partial _t \eta + {\mathfrak {s}} T_\lambda T_{l} \eta + {\mathfrak {g}} (T_\lambda (\eta - T_{B}\eta ) -T_V\cdot \nabla \eta ) \\&\ \ \ \ \ \ \ \ \ \ = O_{H^s \rightarrow H^{s-\frac{1}{2}}}\Big ({\mathfrak {g}} {\mathcal {F}}(\Vert \eta \Vert _{H^s}) (1+\Vert \eta _s\Vert _{H^{\frac{1}{2}-\delta }})\Big )\eta + O_{H^{s+\frac{3}{2}} \rightarrow H^{s-\frac{3}{2}+\delta }}({\mathfrak {s}} {\mathcal {F}}(\Vert \eta \Vert _{H^s}))\eta . \end{aligned}\nonumber \\ \end{aligned}$$
(3.7)

Note that \(\lambda \in \Gamma ^1_\delta \), \(l\in \Gamma ^2_\delta \) and \((B, V)\in W^{1+\delta , \infty }\subset \Gamma ^0_\delta \), with seminorms bounded by \({\mathcal {F}}(\Vert \eta \Vert _{H^s})\). The symbolic calculus in Theorem A.3 then gives

$$\begin{aligned}&T_\lambda T_l \eta = T_{\lambda l}\eta + O_{Op^{3-\delta }}({\mathcal {F}}(\Vert \eta \Vert _{H^s})),\\&T_{\lambda }(\text {Id} - T_{ B}) = T_{\lambda (1-B)} + O_{Op^{1-\delta }}({\mathcal {F}}(\Vert \eta \Vert _{H^s})),\\&T_{V}\cdot \nabla =iT_{\xi \cdot V}= i\mathrm {Re}(T_{\xi \cdot V}) +O_{Op^{1-\delta }}({\mathcal {F}}(\Vert \eta \Vert _{H^s})), \end{aligned}$$

where \(\mathrm {Re}(T_{\xi \cdot V}) = \frac{1}{2}(T_{\xi \cdot V} + T_{\xi \cdot V}^*)\). It then follows from (3.7) that

$$\begin{aligned} \begin{aligned}&(\mu ^+ + \mu ^-)\partial _t \eta + {\mathfrak {s}} T_{\lambda l} \eta + {\mathfrak {g}} \Big (T_{\lambda (1-B)}\eta -i\text {Re}(T_{\xi \cdot V}) \eta \Big ) \\&\ \ \ \ \ \ \ \ \ \ = O_{H^s \rightarrow H^{s-\frac{1}{2}}}({\mathfrak {g}} {\mathcal {F}}(\Vert \eta \Vert _{H^s}) (1+\Vert \eta _s\Vert _{H^{\frac{1}{2}-\delta }}))\eta + O_{H^{s+\frac{3}{2}} \rightarrow H^{s-\frac{3}{2}+\delta }}({\mathfrak {s}} {\mathcal {F}}(\Vert \eta \Vert _{H^s}))\eta . \end{aligned}\nonumber \\ \end{aligned}$$
(3.8)

We set \(\eta _s = \langle D\rangle ^s \eta \). Appealing to Theorem A.3 again we have

$$\begin{aligned}&[\langle D\rangle ^s,T_{\lambda l}] = O_{Op^{s+3-\delta }}({\mathcal {F}}(\Vert \eta \Vert _{H^s})),\\&[\langle D\rangle ^s,T_{\lambda (1-B)}] = O_{Op^{s+1-\delta }}({\mathcal {F}}(\Vert \eta \Vert _{H^s}, {\mathfrak {a}}^{-1})), \\&[\langle D\rangle ^s,\mathrm {Re}(T_{\xi \cdot V})]= O_{Op^{s+1-\delta }}({\mathcal {F}}(\Vert \eta \Vert _{H^s})), \end{aligned}$$

where \([A, B]=AB-BA\) and in the second line we have used the lower bound (3.2) for \((1-B)\) together with the fact that \(\lambda \ge |\xi |\). Note that we have adopted the convention that \({\mathcal {F}}(\Vert \eta \Vert _{H^s})\equiv {\mathcal {F}}(\Vert \eta \Vert _{H^s}, 0)\). This implies

$$\begin{aligned} \begin{aligned}&(\mu ^+ + \mu ^-)\partial _t \eta _s + {\mathfrak {s}} T_{\lambda l} \eta _s + {\mathfrak {g}} \Big (T_{\lambda (1-B)}\eta _s -\mathrm {Re}(T_{\xi \cdot V})\eta _s\Big ) \\&\ \ \ \ \ \ \ \ \ \ = O_{L^2 \rightarrow H^{-\frac{1}{2}}}({\mathfrak {g}} {\mathcal {Q}}(1+\Vert \eta _s\Vert _{H^{\frac{1}{2}-\delta }})) \eta _s+ O_{Op^{1-\delta }}({\mathfrak {g}} {\mathcal {Q}})\eta _s+ O_{H^{\frac{3}{2}} \rightarrow H^{-\frac{3}{2}+\delta }}({\mathfrak {s}} {\mathcal {Q}})\eta _s . \end{aligned}\nonumber \\ \end{aligned}$$
(3.9)

Since \(i\text {Re}(T_{\xi \cdot V})\) is skew-adjoint, by testing (3.9) against \(\eta _s\), we obtain

$$\begin{aligned} \begin{aligned}&\frac{(\mu ^+ + \mu ^-)}{2}\frac{d}{dt}\Vert \eta _s\Vert _{L^2} ^2+ {\mathfrak {s}}(T_{{\lambda l}}\eta _s,\eta _s)_{L^2}+ {\mathfrak {g}}(T_{{\lambda (1-B)}}\eta _s,\eta _s)_{L^2} \\&\ \ \ \ \ \ \ \ \ \ \le {\mathcal {Q}}\Big \{{\mathfrak {g}}\Big [(1+\Vert \eta _s\Vert _{H^{\frac{1}{2}-\delta }}) \Vert \eta _s\Vert _{L^2}\Vert \eta _s\Vert _{H^{\frac{1}{2}}} + \Vert \eta _s\Vert _{H^{\frac{1}{2}-\delta }}\Vert \eta _s\Vert _{H^{\frac{1}{2}}}\Big ]+ {\mathfrak {s}} \Vert \eta _s\Vert _{H^{\frac{3}{2}-\delta }}\Vert \eta _s\Vert _{H^{\frac{3}{2}}}\Big \}, \end{aligned}\nonumber \\ \end{aligned}$$
(3.10)

where \((\cdot , \cdot )_{L^2}\) denotes the \(L^2\) pairing. The term involving \(1+ \Vert \eta _s\Vert _{H^{\frac{1}{2}-\delta }}\) is treated as follows:

$$\begin{aligned} \begin{aligned} (1+ \Vert \eta _s\Vert _{H^{\frac{1}{2}-\delta }}) \Vert \eta _s\Vert _{L^2}\Vert \eta _s\Vert _{H^{\frac{1}{2}}}&\le \Vert \eta _s\Vert _{L^2}\Vert \eta _s\Vert _{H^{\frac{1}{2}}}+ \Vert \eta _s\Vert _{L^2}\Vert \eta _s\Vert _{H^{\frac{1}{2}-\delta }}\Vert \eta _s\Vert _{H^{\frac{1}{2}}} \\&\le {\mathcal {Q}}\Vert \eta _s\Vert _{H^{\frac{1}{2}-\delta }}\Vert \eta _s\Vert _{H^{\frac{1}{2}}}. \end{aligned}\nonumber \\ \end{aligned}$$
(3.11)

In view of (3.2) and the fact that \(\lambda (x, \xi )\ge |\xi |\), we have the lower bounds

$$\begin{aligned} \lambda (1-B)\ge {\mathfrak {a}}|\xi |,\quad l\lambda =\langle \nabla \eta \rangle ^{-3}\lambda ^3\ge \frac{1}{\langle \Vert \eta \Vert _{W^{1, \infty }}\rangle ^3}|\xi |^3. \end{aligned}$$

Moreover, \(\lambda (1-B)\in \Gamma ^1_\delta \) and \(l\lambda \in \Gamma ^3_\delta \) with seminorms bounded by \({\mathcal {F}}(\Vert \eta \Vert _{H^s})\). Then applying the Gåarding’s inequality (A.9) gives

$$\begin{aligned} \Vert \Psi (D)\eta _s\Vert _{H^{\frac{3}{2}}}^2&\le {\mathcal {Q}}\Big (\Vert \eta _s\Vert _{H^\frac{3}{2}}\Vert \eta _s\Vert _{H^{\frac{3}{2}-\delta }}+ (T_{l\lambda }\eta _s,\eta _s)_{L^2}\Big ), \end{aligned}$$
(3.12)
$$\begin{aligned} \Vert \Psi (D)\eta _s\Vert _{H^{\frac{1}{2}}}^2&\le {\mathcal {Q}} \Big (\Vert \eta _s\Vert _{H^\frac{1}{2}}\Vert \eta _s\Vert _{H^{\frac{1}{2}-\delta }}^2 + (T_{\lambda (1-B)}\eta _s,\eta _s)_{L^2}\Big ), \end{aligned}$$
(3.13)

where \(\Psi (D)\) denotes the Fourier multiplier with symbol \(\Psi \) defined by (A.3). In addition, we have

$$\begin{aligned} \Vert u\Vert _{H^r}\le C(\Vert \Psi (D)u\Vert _{H^r}+\Vert u\Vert _{L^2})\quad \forall r\in {\mathbb {R}}. \end{aligned}$$

Thus, (3.10) amounts to

$$\begin{aligned} \begin{aligned}&\frac{(\mu ^+ + \mu ^-)}{2}\frac{d}{dt}\Vert \eta _s\Vert _{L^2} ^2+ \frac{1}{{\mathcal {Q}}}\Big ({{\mathfrak {s}}}\Vert \eta _s\Vert _{H^{\frac{3}{2}}}^2 + {{\mathfrak {g}}}\Vert \eta _s\Vert _{H^{\frac{1}{2}}}^2\Big ) \\&\ \ \ \ \ \ \ \ \ \ \le {\mathcal {Q}}({\mathfrak {g}}\Vert \eta _s\Vert _{H^{\frac{1}{2}-\delta }}\Vert \eta _s\Vert _{H^{\frac{1}{2}}}+ {\mathfrak {s}} \Vert \eta _s\Vert _{H^{\frac{3}{2}-\delta }}\Vert \eta _s\Vert _{H^{\frac{3}{2}}}). \end{aligned} \end{aligned}$$
(3.14)

We use Young’s inequality and interpolation as follows:

$$\begin{aligned} \Vert \eta _s\Vert _{H^{\frac{1}{2}-\delta }}\Vert \eta _s\Vert _{H^{\frac{1}{2}}}&\le \Vert \eta _s\Vert _{L^2}^{2\delta }\Vert \eta _s\Vert _{H^{\frac{1}{2}}}^{2(1-\delta )} \le (10{\mathcal {Q}})^{\frac{2(1-\delta )}{\delta }}\Vert \eta _s\Vert _{L^2}^2 + \frac{1}{100{\mathcal {Q}}^2}\Vert \eta _s\Vert _{H^{\frac{1}{2}}}^2 \end{aligned}$$
(3.15)

and similarly,

$$\begin{aligned} \Vert \eta _s\Vert _{H^{\frac{3}{2}-\delta }}\Vert \eta _s\Vert _{H^{\frac{3}{2}}}\le \Vert \eta _s\Vert _{L^2}^{2\frac{\delta }{3}}\Vert \eta _s\Vert _{H^{\frac{3}{2}}}^{2(1-\frac{\delta }{3})}\le (10{\mathcal {Q}})^{\frac{2(3-\delta )}{\delta }}\Vert \eta _s\Vert _{L^2}^2 + \frac{1}{100{\mathcal {Q}}^2}\Vert \eta _s\Vert _{H^{\frac{3}{2}}}^2. \end{aligned}$$
(3.16)

Applying these inequalities to (3.14), and then subtracting terms involving \(\Vert \eta _s\Vert _{H^{\frac{1}{2}}}\), we obtain for a larger \({\mathcal {Q}}\) if needed that

$$\begin{aligned} \frac{\mu ^+ + \mu ^-}{2}\frac{d}{dt} \Vert \eta _s\Vert _{L^2}^2 + \frac{1}{{\mathcal {Q}}}({\mathfrak {s}} \Vert \eta _s\Vert _{H^{\frac{3}{2}}}^2 + {\mathfrak {g}} \Vert \eta _s\Vert _{H^{\frac{1}{2}}}^2) \le ({\mathfrak {g}} + {\mathfrak {s}}) {\mathcal {Q}}\Vert \eta _s\Vert _{L^2}^2. \end{aligned}$$
(3.17)

A Grönwall’s argument then leads to

$$\begin{aligned}&\Vert \eta \Vert _{L^\infty ([0,T];H^s)}^2 + \frac{1}{{\mathcal {Q}}_T}\Big ({\mathfrak {s}} \Vert \eta \Vert _{L^2([0,T];H^{s+\frac{3}{2}})}^2 +{\mathfrak {g}} \Vert \eta \Vert _{L^2([0,T];H^{s+\frac{1}{2}})}^2\Big )\\&\quad \le \Vert \eta _0\Vert _{H^s}^2 \exp \Big (({\mathfrak {s}} + {\mathfrak {g}})T{\mathcal {Q}}_T\Big ), \end{aligned}$$

where \({\mathcal {Q}}_T={\mathcal {F}}(\Vert \eta \Vert _{L^\infty ([0,T];H^s)}, {\mathfrak {a}}^{-1})\) with \({\mathcal {F}}\) depending only on \((h, s, \mu ^\pm )\). In particular, we have the \(H^s\) estimate (3.4). As for the dissipation estimate, we have

$$\begin{aligned}&{\mathfrak {s}} \Vert \eta \Vert _{L^2([0,T];H^{s+\frac{3}{2}})}^2 +{\mathfrak {g}} \Vert \eta \Vert _{L^2([0,T];H^{s+\frac{1}{2}})}^2\\&\quad \le \Vert \eta _0\Vert _{H^s}^2 \exp \Big (({\mathfrak {s}} + {\mathfrak {g}})T{\mathcal {F}}\big (\Vert \eta \Vert _{L^\infty ([0,T];H^s)}, {\mathfrak {a}}^{-1}\big )\Big ){\mathcal {Q}}_T. \end{aligned}$$

On the other hand, plugging (3.4) into \({\mathcal {Q}}_T\) gives

$$\begin{aligned} {\mathcal {Q}}_T&={\mathcal {F}}(\Vert \eta \Vert _{L^\infty ([0,T];H^s)},{\mathfrak {a}}^{-1})\\&\le {\mathcal {F}}\left( \Vert \eta _0\Vert _{H^s} \exp \Big (({\mathfrak {s}} + {\mathfrak {g}})T{\mathcal {F}}\big (\Vert \eta \Vert _{L^\infty ([0,T];H^s)},{\mathfrak {a}}^{-1}\big )\Big ),{\mathfrak {a}}^{-1}\right) \end{aligned}$$

Therefore, upon setting \({\mathcal {F}}_1(m, n)=m^2{\mathcal {F}}(m, n)\) we obtain

$$\begin{aligned}&{\mathfrak {s}} \Vert \eta \Vert _{L^2([0,T];H^{s+\frac{3}{2}})}^2 +{\mathfrak {g}} \Vert \eta \Vert _{L^2([0,T];H^{s+\frac{1}{2}})}^2\\&\quad \le {\mathcal {F}}_1\left( \Vert \eta _0\Vert _{H^s} \exp \Big (({\mathfrak {s}} + {\mathfrak {g}})T{\mathcal {F}}\big (\Vert \eta \Vert _{L^\infty ([0,T];H^s)}, {\mathfrak {a}}^{-1}\big )\Big ),{\mathfrak {a}}^{-1}\right) \end{aligned}$$

which finishes the proof of (3.5). \(\quad \square \)

4 Contraction Estimates for \(J^\pm \)

Our goal in this section is to prove contraction estimates for \(J^\pm (\eta )\) at two different surfaces \(\eta _1\) and \(\eta _2\). This is only a question for the two-phase problem since for the one-phase problem we have \(J^-=\mathrm {Id}\) and \(J^+\equiv 0\). Given an object X depending on \(\eta \), we shall denote \(X_j = X|_{\eta = \eta _j}\) and the difference

$$\begin{aligned} X_\delta = X_1 - X_2. \end{aligned}$$

Proposition 4.1

Let \(s > 1 + \frac{d}{2}\) and consider \(\eta _j\in H^s({\mathbb {R}}^d)\) satisfying \({{\,\mathrm{dist}\,}}(\eta _j, \Gamma ^\pm )> h>0\), \(j=1, 2\). For any \(\sigma \in [\frac{1}{2}, s]\), there exists \({\mathcal {F}}:{\mathbb {R}}^+\rightarrow {\mathbb {R}}^+\) depending only on \((h, s, \sigma , \mu ^\pm )\) such that

$$\begin{aligned}&\Vert J ^\pm _\delta \Vert _{ H^s\rightarrow \widetilde{H}^\sigma _\pm } \le {\mathcal {F}}(N_s)\Vert \eta _\delta \Vert _{H^\sigma }, \end{aligned}$$
(4.1)
$$\begin{aligned}&\Vert J ^\pm _\delta \Vert _{ H^\sigma \rightarrow \widetilde{H}^\sigma _\pm } \le {\mathcal {F}}(N_s)\Vert \eta _\delta \Vert _{H^s}, \end{aligned}$$
(4.2)

where we denoted

$$\begin{aligned} N_s = \Vert (\eta _1, \eta _2)\Vert _{H^s}. \end{aligned}$$
(4.3)

We shall prove Proposition 4.1 for the most general case of two fluids and with bottoms, i.e. \(\mu ^+>0\) and \(\Gamma ^\pm \ne \emptyset \). Adaption to the other cases is straightforward.

4.1 Flattening the domain

There exist \(\eta _*^\pm \in C_b^{s+100}({\mathbb {R}}^d)\) such that

$$\begin{aligned} b^-(x)+\frac{h}{2}\le \eta ^-_*(x)\le \eta _j(x)-\frac{h}{2},\quad \eta _j(x)\le \eta _*^+(x)-\frac{h}{2}\le b^+(x)-h\quad \forall x\in {\mathbb {R}}^d \end{aligned}$$
(4.4)

and for some \(C=C(h, s, d)\),

$$\begin{aligned} \Vert \eta _*^\pm \Vert _{C_b^{s+100}({\mathbb {R}}^d)}\le C(1+\Vert \eta _1\Vert _{L^\infty }+\Vert \eta _2\Vert _{L^\infty }). \end{aligned}$$
(4.5)

For \(j=1, 2\) we set \(\Omega _*=\Omega ^+_{j, *}\cup \Omega ^-_{j, *}\) where

$$\begin{aligned} \Omega _{*, j}^\pm = \{(x,y):x\in {\mathbb {R}}^d, \pm \eta (x)\le \pm y\le \pm \eta ^\pm _*(x)\}. \end{aligned}$$
(4.6)

Note that \(\Omega _*=\{(x,y):x\in {\mathbb {R}}^d, \eta ^-_*(x)\le y\le \eta ^+_*(x)\}\) is independent of \(j\in \{1, 2\}\). For small \(\tau >0\) to be chosen, define \(\rho _j(x, z): {\mathbb {R}}^d\times [-1, 1]\) by

$$\begin{aligned} \varrho _j(x,z)= (1-z^2)e^{-\tau |z|\langle D_x \rangle }\eta _j(x) -\frac{1}{2}z(1-z) \eta ^-_*(x)+\frac{1}{2}z(1+z)\eta ^+_*(x). \end{aligned}$$
(4.7)

Lemma 4.2

There exists \(K>0\) depending only on (sd) such that if \(\tau K\Vert \eta _j\Vert _{H^s}\le \frac{h}{12}\) then

$$\begin{aligned} \partial _z\varrho _j(x, z)\ge \frac{h}{12}\quad a.e.~(x, z)\in {\mathbb {R}}^d\times (-1, 1),~j=1, 2. \end{aligned}$$
(4.8)

For \(j =1,2\), the mapping

$$\begin{aligned} \Phi _j: {\mathbb {R}}^d\times [-1, 1] \ni (x, z) \mapsto (x, \varrho _j(x, z))\in \Omega _* \end{aligned}$$

is a Lipschitz diffeomorphism and respectively maps \({\mathbb {R}}^d\times [0, 1]\) and \({\mathbb {R}}^d\times [-1, 0]\) onto \(\Omega _{*, j}^+\) and \(\Omega _{*, j}^-\). Moreover, there exists \(C=C(h, s, d)\) such that

$$\begin{aligned}&\Vert \nabla \Phi _j\Vert _{L^\infty ({\mathbb {R}}^d\times (-1, 1))}\le C(1+N_s), \end{aligned}$$
(4.9)
$$\begin{aligned}&\Vert \nabla (\Phi _j^{-1})\Vert _{L^\infty (\Omega _*)}\le C(1+N_s). \end{aligned}$$
(4.10)

Proof

We first note that \(\Phi _j(x, 0)=\Sigma _j=\{(x, \eta _j(x)): x\in {\mathbb {R}}^d\}\), \(\Phi _j(x, 1)=\{(x, \eta ^+_*(x)): x\in {\mathbb {R}}^d\}\) and \(\Phi _j(x, -1)=\{(x, \eta ^-_*(x)): x\in {\mathbb {R}}^d\}\). Thus, in order to prove that \(\Phi _j\) is one-to-one and onto, it suffices to prove that \(\partial _z\varrho _j(x, z)\ge c>0\) for a.e. \((x, z)\in {\mathbb {R}}^d\times (-1, 1)\). For \(z\in (-1, 1){\setminus }\{0\}\) we have

$$\begin{aligned} \partial _z\varrho _j(x, z)&=\frac{1}{2}(1-2z)(\eta _j(x)-\eta ^{-}_{*}(x))+\frac{1}{2}(1+2z)(\eta ^{+}_{*}(x)-\eta _j(x))\\&\quad -2z(e^{-\tau |z|\langle D_x \rangle }-1)\eta _j(x)-\text {sign}(z)\tau (1-z^2)e^{-\tau |z|\langle D_x \rangle }\langle D_x \rangle \eta _j(x). \end{aligned}$$

For \(z\in [\frac{1}{3}, 1]\), \(1-2z\in [-1, \frac{1}{3}]\) and \(1+2z\in [\frac{5}{3}, 3]\). In addition, by (4.4) we have . Consequently,

$$\begin{aligned} \frac{1}{2}(1-2z)(\eta _j(x)-\eta ^-_*(x))+\frac{1}{2}(1+2z)(\eta ^+_*(x) -\eta _j(x))\ge \frac{h}{6}. \end{aligned}$$
(4.11)

Similarly we obtain (4.11) for \(z\in (-1, -\frac{1}{3})\) and \(z\in [-\frac{1}{3}, \frac{1}{3}]{\setminus }\{0\}\). Next writing

$$\begin{aligned} (e^{-\tau |z|\langle D_x \rangle }-1)\eta _j(x)=-\tau \int _0^{|z|}e^{-\tau z'\langle D_x \rangle }\langle D_x \rangle \eta _j(x)dz' \end{aligned}$$

we obtain that

$$\begin{aligned} \partial _z\varrho _j(x, z)\ge \frac{h}{6}-\tau K\Vert \eta _j\Vert _{H^s}\quad \forall (x, z)\in {\mathbb {R}}^d\times (-1, 1) \end{aligned}$$

for some constant \(K=K(s, d)\). Note that the condition \(s>1+\frac{d}{2}\) has been used. Choosing \(\tau >0\) such that \(\tau K\Vert \eta _j\Vert _{H^s}\le h/12\) gives \(\partial _z\varrho _j(x, z)\ge h/12\) for a.e. \((x, z)\in {\mathbb {R}}^d\times (-1, 1)\).

Since

$$\begin{aligned} \partial _z\varrho _j(x, z)&=-2ze^{-\tau |z|\langle D_x \rangle }\eta _j(x)-\text {sign}(z)\tau (1-z^2)e^{-\tau |z|\langle D_x \rangle }\langle D_x \rangle \eta _j(x)\\&-\frac{1}{2}(1-2z)\eta ^-_*(x)+\frac{1}{2}(1+2z)\eta ^+_*(x), \end{aligned}$$

there exists \(K'=K'(s, d)\) such that

$$\begin{aligned} \Vert \partial _z\varrho _j(x, z)\Vert _{W^{1, \infty }({\mathbb {R}}^d\times (-1, 1))}\le K'\Vert \eta _j\Vert _{H^{s-1}}+K'\tau \Vert \eta _j\Vert _{H^s}+K'(1+N_{s-1}), \end{aligned}$$

where we have used (). Then in view of the fact that \(\tau \Vert \eta _j\Vert _{H^s}\le hK^{-1}\), we obtain

$$\begin{aligned} \Vert \varrho _j\Vert _{W^{1, \infty }({\mathbb {R}}^d\times (-1, 1))}\le C(1+N_s), \quad C=C(h, s, d), \end{aligned}$$
(4.12)

whence (4.9) follows. On the other hand, we have \(\Phi ^{-1}_j(x, y)=(x, \kappa _j(x, y))\) where

$$\begin{aligned} y= \varrho _j(x, z) \iff z= \kappa _j(x, y)\quad a.e.~ (x, z)\in {\mathbb {R}}^d\times (-1, 1). \end{aligned}$$
(4.13)

Then the relation \(\kappa _j(x, \varrho _j(x, z))=z\) yields

$$\begin{aligned} \partial _y\kappa _j(x, \varrho _j(x, z))= \frac{1}{\partial _z\varrho _j(x, z))},\quad \partial _x\kappa _j(x, \varrho _j(x, z))=-\frac{\partial _x\varrho _j(x, z)}{\partial _z\varrho _j(x, z))}. \end{aligned}$$
(4.14)

Thus, in view of (4.8) and (4.12), we obtain (4.10). \(\quad \square \)

Lemma 4.3

Set

$$\begin{aligned} \Upsilon (x, y)= {\left\{ \begin{array}{ll} \Phi _1\circ \Phi _2^{-1},\quad (x, y)\in \Omega _*,\\ (x, y),\quad (x, y)\in \Omega {\setminus }\Omega _* \end{array}\right. } \end{aligned}$$
(4.15)

and

$$\begin{aligned} M=\frac{\nabla \Upsilon \nabla \Upsilon ^t}{|\det \nabla \Upsilon |}. \end{aligned}$$
(4.16)

Then, \(\Upsilon \) is a Lipschitz diffeomorphism on \(\Omega \) and

$$\begin{aligned}&\frac{1}{C(1+N_s)}\le \det \nabla \Upsilon (x, y) \le C(1+N_s)\quad a.e.~ (x, y)\in \Omega , \end{aligned}$$
(4.17)
$$\begin{aligned}&\Vert \nabla \Upsilon \Vert _{L^\infty (\Omega )}+ \Vert \nabla (\Upsilon ^{-1}) \Vert _{L^\infty (\Omega )}\le {\mathcal {F}}(N_s). \end{aligned}$$
(4.18)

Moreover, M satisfies

$$\begin{aligned}&\Vert M-\mathrm {Id}\Vert _{L^\infty (\Omega )}\le {\mathcal {F}}(N_s)\Vert \eta _\delta \Vert _{H^s}, \end{aligned}$$
(4.19)
$$\begin{aligned}&\Vert M-\mathrm {Id}\Vert _{L^2(\Omega )}\le {\mathcal {F}}(N_s)\Vert \eta _\delta \Vert _{H^\frac{1}{2}}. \end{aligned}$$
(4.20)

Proof

According to Lemma 4.2, \(\Upsilon \) is a Lipschitz diffeomorphism on \(\Omega _*\). For \((x, y)\in \Omega {\setminus }\Omega _*\) we have \(\Upsilon =\mathrm {Id}\), and hence \(\Upsilon \) is a Lipschitz diffeomorphism on \(\Omega \) and \(M-\mathrm {Id}=0\). It thus suffices to consider \((x, y)\in \Omega _*\). On \(\Omega _*\) we have \(\Upsilon (x, y)=(x, \varrho _1(x, \kappa _2(x, y))\), and so

$$\begin{aligned} \nabla \Upsilon (x, y)=\begin{pmatrix} 1 &{}\quad 0\\ a(x, y) &{}\quad b(x, y) \end{pmatrix} \end{aligned}$$

where

$$\begin{aligned} a(x, y)= & {} \partial _x\varrho _1(x, \kappa _2(x, y))+\partial _z\varrho _1(x, \kappa _2(x, y))\partial _x\kappa _2(x, y),\\&\quad b(x, y)=\partial _z\varrho _1(x, \kappa _2(x, y))\partial _y\kappa _2(x, y). \end{aligned}$$

Using (4.14) (with \(j=2\)) gives

$$\begin{aligned}&a(x, y)=\frac{\partial _x\varrho _1(x, \kappa _2(x, y))\partial _z\varrho _2(x, \kappa _2(x, y))-\partial _x\varrho _2(x, \kappa _2(x, y))\partial _z\varrho _1(x, \kappa _2(x, y))}{\partial _z\varrho _2(x, \kappa _2(x, y))},\\&b(x, y)=\frac{\partial _z\varrho _1(x, \kappa _2(x, y))}{\partial _z\varrho _2(x, \kappa _2(x, y))}. \end{aligned}$$

In view of (4.8) and (4.12) we obtain

$$\begin{aligned} \frac{1}{C(1+N_s)}\le \det \nabla \Upsilon (x, y)=b(x, y) \le C(1+N_s)\quad a.e.~ (x, y)\in \Omega _*. \end{aligned}$$
(4.21)

Next we compute

$$\begin{aligned} M-\mathrm {Id}=\frac{1}{b}\begin{pmatrix} 1-b&{} a\\ a&{} a^2+b(b-1) \end{pmatrix}. \end{aligned}$$

Using the above formulas for a and b together with (4.7) and (4.8) we deduce that

$$\begin{aligned}&\Vert (a, b-1)\Vert _{L^\infty (\Omega _*)}\le {\mathcal {F}}(N_s)\Vert \eta _\delta \Vert _{H^s},\\&\Vert (a, b-1)\Vert _{L^2(\Omega _*)}\le {\mathcal {F}}(N_s)\Vert \eta _\delta \Vert _{H^\frac{1}{2}}. \end{aligned}$$

This combined with (4.21) leads to (4.18), (4.19) and (4.20). \(\quad \square \)

4.2 Proof of Proposition 4.1

The proof proceeds in three steps.

Step 1. We first recall from (2.15) that \(J^\pm v=f^\pm \) where \(f^\pm \) solve

$$\begin{aligned} \begin{pmatrix} \text {Id} &{} -\mathrm {Id}\\ \mu ^+ G^-(\eta ) &{} -\mu ^- G^+(\eta ) \end{pmatrix} \begin{pmatrix} f^- \\ f^+ \end{pmatrix} = \begin{pmatrix} v \\ 0 \end{pmatrix}. \end{aligned}$$
(4.22)

From the definition of the Dirichlet-Neumann operator we see that \(f^\pm =q^\pm \vert _\Sigma \) where \(q^\pm \) solve the two-phase elliptic problem

$$\begin{aligned} {\left\{ \begin{array}{ll} \Delta q^\pm =0\quad \text {in}~\Omega ^\pm ,\\ q^--q^+=v \quad \text {on}~\Sigma ,\\ \frac{\partial _nq^-}{\mu ^-}-\frac{\partial _nq^+}{\mu ^+}=0\quad \text {on}~\Sigma ,\\ \partial _{\nu ^\pm }q^\pm =0\quad \text {on}~\Gamma ^\pm . \end{array}\right. } \end{aligned}$$
(4.23)

To remove the jump of q at \(\Sigma \) we take a function \(\theta :\Omega \rightarrow {\mathbb {R}}\) satisfying

$$\begin{aligned}&\theta (x, \eta (x))=-\frac{1}{2}v(x),\quad \theta \equiv 0\quad \text {near}~ \Gamma ^\pm , \end{aligned}$$
(4.24)
$$\begin{aligned}&\Vert \theta \Vert _{\dot{H}^1(\Omega )}\le C(1+\Vert \eta \Vert _{W^{1, \infty }})\Vert v\Vert _{H^\frac{1}{2}({\mathbb {R}}^d)},\quad C=C(d). \end{aligned}$$
(4.25)

Then, the solution of (4.23) can be taken to be \(q^\pm :=(r\pm \theta )\vert _{\Omega ^\pm }\) where \(r\in \dot{H}^1(\Omega )\) solves

$$\begin{aligned} {\left\{ \begin{array}{ll} -\Delta r=\pm \Delta \theta \quad \text {in}~\Omega ^\pm ,\\ \frac{\partial _n r}{\mu ^+}-\frac{\partial _n r}{\mu ^-}=-\partial _n\theta (\frac{1}{\mu ^+}+\frac{1}{\mu ^-})\quad \text {on}~\Sigma ,\\ \partial _{\nu ^\pm }r=0\quad \text {on}~\Gamma ^\pm . \end{array}\right. } \end{aligned}$$
(4.26)

A function \(\theta \) satisfying (4.24) and (4.25) can be constructed as follows. Let \(\varsigma (z):{\mathbb {R}}\rightarrow {\mathbb {R}}^+\) be a cutoff function that is identically 1 for \(|z|\le \frac{1}{2}\) and vanishes for \(|z|\ge 1\). Set

$$\begin{aligned} {{\underline{\theta }}}(x, z)=-\frac{1}{2}\varsigma (z)e^{-|z|\langle D_x\rangle }v(x),\quad \theta (x, y) =\theta \big (x, \frac{y-\eta (x)}{h}\big ). \end{aligned}$$
(4.27)

Then, \(\theta (x, y)=0\) for \(|y-\eta (x)|\ge h\), and hence \(\theta _1\equiv 0\) near \(\Gamma ^\pm \) in view of the condition \({{\,\mathrm{dist}\,}}(\eta , \Gamma ^\pm )>h\). Moreover, (4.30) is satisfied.

Integration by parts leads to the following variational form of (4.26):

$$\begin{aligned} \int _\Omega \Big (\frac{1_{\Omega ^-}}{\mu ^-}+ \frac{1_{\Omega ^+}}{\mu ^+}\Big )\nabla r \cdot \nabla \phi \ dxdy= \int _\Omega \Big (\frac{1_{\Omega ^-}}{\mu ^-}-\frac{1_{\Omega ^+}}{\mu ^+}\Big )\nabla \theta \cdot \nabla \phi \ dxdy,\quad \forall \phi \in \dot{H}^1(\Omega ).\nonumber \\ \end{aligned}$$
(4.28)

For example, for \(\varsigma (z):{\mathbb {R}}\rightarrow {\mathbb {R}}^+\) a cutoff function that is identically 1 for \(|z|\le \frac{1}{2}\) and vanishes for \(|z|\ge 1\), we set

$$\begin{aligned} {{\underline{\theta }}}(x, z)=-\frac{1}{2}\varsigma (z)e^{-|z|\langle D_x\rangle }v(x),\quad \theta (x, y) =\theta \big (x, \frac{y-\eta (x)}{h}\big ). \end{aligned}$$
(4.29)

Then, \(\theta (x, y)=0\) for \(|y-\eta (x)|\ge h\), and hence \(\theta _1\equiv 0\) near \(\Gamma ^\pm \) in view of the condition \({{\,\mathrm{dist}\,}}(\eta , \Gamma ^\pm )>h\). Moreover, we have

$$\begin{aligned} \Vert \theta \Vert _{\dot{H}^1(\Omega )}\le C(1+\Vert \eta \Vert _{W^{1, \infty }})\Vert v\Vert _{H^\frac{1}{2}({\mathbb {R}}^d)},\quad C=C(d). \end{aligned}$$
(4.30)

By virtue of the Lax-Milgram theorem, there exists a unique solution \(r\in \dot{H}^1(\Omega )\) to (4.28) which obeys the bound

$$\begin{aligned} \Vert r\Vert _{\dot{H}^1(\Omega )}\le C(\mu ^\pm )\Vert \theta \Vert _{\dot{H}^1(\Omega )}\le C(1+\Vert \eta \Vert _{W^{1, \infty }})\Vert v\Vert _{H^\frac{1}{2}({\mathbb {R}}^d)}. \end{aligned}$$
(4.31)

Consequently,

$$\begin{aligned} J^\pm v=f^\pm =\mathrm {Tr}_{\Omega ^\pm \rightarrow \Sigma }(r\pm \theta ), \end{aligned}$$
(4.32)

and hence by the trace operation (2.5),

$$\begin{aligned} \Vert J^\pm v\Vert _{\widetilde{H}^\frac{1}{2}_\pm ({\mathbb {R}}^d)}\le {\mathcal {F}}(N_s)\Vert v\Vert _{H^\frac{1}{2}({\mathbb {R}}^d)}. \end{aligned}$$
(4.33)

We note that \(\theta \) defined by (4.27) depends on \(\eta \), and so does r.

Step 2. In this step we prove contraction estimates for \(J^\pm _\delta = J^\pm _1-J^\pm _2\) in \(\widetilde{H}^\frac{1}{2}_\pm \). Recall that \(\Upsilon \) defined by (4.15) is a Lipschitz diffeomorphism on \(\Omega \). Let \(\theta _1\) be defined as in (4.29) with \(\eta =\eta _1\), and let \(\theta _2=\theta _1\circ \Upsilon \). Let us check that \(\theta _2\) obeys (4.24) and (4.25) for \(\eta =\eta _2\). Indeed, using the fact that \(\Upsilon :\Sigma _2\rightarrow \Sigma _1\) we have

$$\begin{aligned} \theta _2(x, \eta _2(x))=\theta _1(x, \Upsilon (x, \eta _2(x)))= \theta _1(x, \eta _1(x))=-\frac{1}{2}v(x). \end{aligned}$$

On the other hand, since \(\Upsilon \equiv \mathrm {Id}\) near \(\Gamma ^\pm \) and \(\theta _1\equiv 0\) near \(\Gamma ^\pm \), we deduce that \(\theta _1\equiv 0\) near \(\Gamma ^\pm \). Finally, the bound (4.30) for \(\theta _2\) follows from (4.25) for \(\theta _1\) and the Lipschitz bound (4.18) for \(\Upsilon \).

According to Step 1, we have

$$\begin{aligned} J^\pm _j v=\mathrm {Tr}_{\Omega _j^\pm \rightarrow \Sigma _j}(r_j\pm \theta _j) \end{aligned}$$

where \(r_j\in \dot{H}^1(\Omega )\) satisfies

$$\begin{aligned} \int _\Omega \Big (\frac{1_{\Omega ^-}}{\mu ^-}+ \frac{1_{\Omega ^+}}{\mu ^+}\Big )\nabla r_j \cdot \nabla \phi \ dxdy= \int _\Omega \Big (\frac{1_{\Omega ^-}}{\mu ^-}-\frac{1_{\Omega ^+}}{\mu ^+}\Big )\nabla \theta _j \cdot \nabla \phi \ dxdy\quad \forall \phi \in \dot{H}^1(\Omega ).\nonumber \\ \end{aligned}$$
(4.34)

Set \(\widetilde{r}_2 = r_2 \circ \Upsilon ^{-1}\) and recall that \(\theta _1=\theta _2\circ \Upsilon ^{-1}\). Combining (4.31) and (4.17) gives

$$\begin{aligned} \Vert \widetilde{r}_2\Vert _{\dot{H}^1(\Omega )}\le {\mathcal {F}}(N_s). \end{aligned}$$

Since map \(\phi \mapsto \phi \circ \Upsilon ^{-1}\) is an isomorphism on \(\dot{H}^1 (\Omega )\), with \(M = \nabla \Upsilon \nabla \Upsilon ^t/|\det \nabla \Upsilon |\) we have for all \(\phi \in \dot{H}^1(\Omega )\) that

$$\begin{aligned} \begin{aligned}&\int _\Omega \Big (\frac{1_{\Omega _1^-}}{\mu ^-}+ \frac{1_{\Omega _1^+}}{\mu ^+}\Big ) \nabla {\tilde{r}}_2 M \nabla \phi ^t \ dxdy = \int _\Omega \Big (\frac{1_{\Omega _1^-}}{\mu ^-}-\frac{1_{\Omega _1^+}}{\mu ^+}\Big )\nabla \theta _1 M\nabla \phi ^t \ dxdy, \end{aligned} \end{aligned}$$
(4.35)

where gradients of scalar functions are understood as row vectors, and the rows of the Jacobian matrix \(\nabla \Upsilon \) are the gradients of each component of \(\Upsilon \). Taking the difference between (4.35) with \(j=2\) and (4.34) with \(j=1\), we obtain

$$\begin{aligned} \begin{aligned} \int _\Omega \Big (\frac{1_{\Omega _1^-}}{\mu ^-}+ \frac{1_{\Omega _1^+}}{\mu ^+}\Big ) \nabla (\widetilde{r}_2 -r_1) \nabla \phi ^t \ dxdy&= -\int _\Omega \Big (\frac{1_{\Omega _1^-}}{\mu ^-}+\frac{1_{\Omega _1^+}}{\mu ^+}\Big )\nabla \widetilde{r}_2 (M-\mathrm {Id})\nabla \phi ^t \ dxdy\\&\quad +\int _\Omega \Big (\frac{1_{\Omega _1^-}}{\mu ^-}-\frac{1_{\Omega _1^+}}{\mu ^+}\Big )\nabla \theta _1 (M-\mathrm {Id})\nabla \phi ^t \ dxdy \end{aligned}\nonumber \\ \end{aligned}$$
(4.36)

for all \(\phi \in \dot{H}^1(\Omega )\). Setting \(\phi =\widetilde{r}_2 -r_1\) and using the estimate (4.19) for \(M-\mathrm {Id}\) in \(L^\infty (\Omega )\) we obtain

$$\begin{aligned} \begin{aligned} \Vert \widetilde{r}_2 -r_1\Vert _{\dot{H}^1(\Omega )}&\le C(\mu ^\pm )\Vert M-\mathrm {Id}\Vert _{L^\infty (\Omega )}(\Vert \widetilde{r}_2\Vert _{\dot{H}^1(\Omega )}+\Vert \theta _1\Vert _{\dot{H}^1(\Omega )}) \\&\le {\mathcal {F}}(N_s)\Vert \eta _\delta \Vert _{H^s}\Vert v\Vert _{H^\frac{1}{2}}. \end{aligned} \end{aligned}$$
(4.37)

On the other hand, using (4.20) for \(M-\mathrm {Id}\) in \(L^2(\Omega )\) instead gives

$$\begin{aligned} \begin{aligned} \Vert \widetilde{r}_2 -r_1\Vert _{\dot{H}^1(\Omega )}&\le C(\mu ^\pm )\Vert M-\mathrm {Id}\Vert _{L^2(\Omega )}(\Vert \nabla \widetilde{r}_2\Vert _{L^\infty (\Omega )}+\Vert \nabla \theta _1\Vert _{L^\infty (\Omega )}) \\&\le {\mathcal {F}}(N_s)\Vert \eta _\delta \Vert _{H^\frac{1}{2}}\Vert v\Vert _{H^s}, \end{aligned} \end{aligned}$$
(4.38)

where in the last inequality we have used the fact that

$$\begin{aligned} \Vert \widetilde{r}_2\Vert _{L^\infty (\Omega )}+\Vert \theta _1\Vert _{L^\infty (\Omega )}\le {\mathcal {F}}(N_s)\Vert v\Vert _{H^s}. \end{aligned}$$

Since \(r_2=\widetilde{r}_2\circ \Upsilon \), \(\theta _2= \theta \circ \Upsilon \), and \(\Upsilon :\Omega _2^\pm \rightarrow \Omega _1^\pm ,~\Sigma _2\rightarrow \Sigma _1\) we have

$$\begin{aligned} \mathrm {Tr}_{\Omega ^\pm _2\rightarrow \Sigma _2}(r_2\pm \theta _2)=\mathrm {Tr}_{\Omega _1^\pm \rightarrow \Sigma _1}(\widetilde{r}_2\pm \theta _1), \end{aligned}$$

and hence

$$\begin{aligned} J^\pm _\delta v&=\mathrm {Tr}_{\Omega _1^\pm \rightarrow \Sigma _1}(r_1\pm \theta _1)-\mathrm {Tr}_{\Omega _2^\pm \rightarrow \Sigma _2}(r_2\pm \theta _2)\\&=\mathrm {Tr}_{\Omega ^\pm _1\rightarrow \Sigma _1}(r_1\pm \theta _1)-\mathrm {Tr}_{\Omega _1^\pm \rightarrow \Sigma _1}(\widetilde{r}_2\pm \theta _1)\\&=\mathrm {Tr}_{\Omega _1^\pm \rightarrow \Sigma _1}(r_1-\widetilde{r}_2). \end{aligned}$$

In view of (4.37) and (4.38), the trace operation (2.5) yields

$$\begin{aligned}&\Vert J^\pm _\delta v\Vert _{H^\frac{1}{2}_\pm }\le {\mathcal {F}}(N_s)\Vert \eta _\delta \Vert _{H^s}\Vert v\Vert _{H^\frac{1}{2}}, \end{aligned}$$
(4.39)
$$\begin{aligned}&\Vert J^\pm _\delta v\Vert _{H^\frac{1}{2}_\pm }\le {\mathcal {F}}(N_s)\Vert \eta _\delta \Vert _{H^\frac{1}{2}}\Vert v\Vert _{H^s}. \end{aligned}$$
(4.40)

For the proof of (4.1), we shall only need (4.40).

Step 3. We have \(J^\pm _jv\equiv J^\pm (\eta _j)v=f^\pm _j\) where

$$\begin{aligned} {\left\{ \begin{array}{ll} f^-_j-f^+_j=v,\\ \frac{1}{\mu ^-}G^-_jf_j^--\frac{1}{\mu ^+}G^+_jf_j^+=0, \end{array}\right. } \quad j=1, 2. \end{aligned}$$
(4.41)

By taking differences we obtain \(f^-_\delta =f^+_\delta \) and

$$\begin{aligned} \frac{1}{\mu ^-} G^-_1 f^-_\delta -\frac{1}{\mu ^+} G^+_1 f^+_\delta =\frac{1}{\mu ^+}G^+_\delta f^+_2-\frac{1}{\mu ^-}G^-_\delta f^- _2, \end{aligned}$$
(4.42)

where we recall the notation \(G^\pm _\delta =G^\pm _1-G^\pm _2\). Combining the contraction estimate (2.32) with the continuity (2.20) for \(J^\pm \), we deduce that

$$\begin{aligned}&\Vert G^+_\delta f^+_2\Vert _{H^{\sigma -1}} +\Vert G^-_\delta f^-_2\Vert _{H^{\sigma -1}}\le {\mathcal {F}}(N_s)\Vert \eta _\delta \Vert _{H^\sigma }\Vert v\Vert _{H^s}\quad \forall \sigma \in [\frac{1}{2}, s]. \end{aligned}$$
(4.43)

We take \(\delta \in (0, \frac{1}{2}]\) satisfying \(\sigma <s-1-\frac{d}{2}\). In light of the paralinearization (2.25) for \(G^\pm _1\), we have

$$\begin{aligned} \begin{aligned}&\frac{1}{\mu ^-} G^-_1 f^-_\delta -\frac{1}{\mu ^+} G^+_1 f^+_\delta = T_{\lambda _1}\big (\frac{1}{\mu ^-} f^-_\delta + \frac{1}{\mu ^+} f^+_\delta \big ) + R^1=\frac{\mu ^++\mu ^-}{\mu ^-\mu ^+}T_{\lambda _1} f^-_\delta +R^1,\\&\Vert R^1\Vert _{H^{\nu -1+\delta }}\le {\mathcal {F}}(N_s)\Vert f^-_\delta \Vert _{\widetilde{H}^\nu _-}\quad \forall \nu \in [\frac{1}{2}, s-\delta ]. \end{aligned} \end{aligned}$$
(4.44)

It follows from (4.42),(4.43) and (4.44) that if

$$\begin{aligned} \nu \in [\frac{1}{2}, s-\delta ],\quad \nu +\delta \le \sigma ,\quad \sigma \in [\frac{1}{2}, s], \end{aligned}$$
(4.45)

then

$$\begin{aligned}&\Vert T_{\lambda _1} f^-_\delta \Vert _{H^{\nu -1+\delta }}\le {\mathcal {F}}(N_s)\Vert f^-_\delta \Vert _{\widetilde{H}^\nu _-}+{\mathcal {F}}(N_s)\Vert \eta _\delta \Vert _{H^\sigma }\Vert v\Vert _{H^s}. \end{aligned}$$
(4.46)

Applying Lemma A.6 we have

$$\begin{aligned} \Vert \Psi (D) f^-_\delta \Vert _{H^{\nu +\delta }}\le {\mathcal {F}}(\Vert \eta _1\Vert _{H^s})\Vert T_{\lambda _1} f^-_\delta \Vert _{H^{\nu +\delta -1}}+{\mathcal {F}}(\Vert \eta _1\Vert _{H^s})\Vert f^-_\delta \Vert _{H^{1, \nu }}. \end{aligned}$$

But for \(\tau >\frac{1}{2}\),

$$\begin{aligned} \Vert \cdot \Vert _{\widetilde{H}^{\tau }_\pm }\le C\Vert \Psi (D)\cdot \Vert _{H^\tau }+ C\Vert \cdot \Vert _{\widetilde{H}^\frac{1}{2}_\pm }, \end{aligned}$$

whence (4.46) implies that

$$\begin{aligned}&\Vert f^-_\delta \Vert _{\widetilde{H}^{\nu +\delta }_-}\le {\mathcal {F}}(N_s)\Vert f^-_\delta \Vert _{\widetilde{H}^\nu _-}+{\mathcal {F}}(N_s)\Vert \eta _\delta \Vert _{H^\sigma }\Vert v\Vert _{H^s} \end{aligned}$$
(4.47)

provided that \(\nu \) and \(\sigma \) satisfy (4.45).

We note that (4.40) implies that

$$\begin{aligned} \Vert f^-_\delta \Vert _{\widetilde{H}^\frac{1}{2}_-}\le {\mathcal {F}}(N_s)\Vert \eta _\delta \Vert _{H^\frac{1}{2}}\Vert v\Vert _{H^s}, \end{aligned}$$
(4.48)

and hence (4.1) holds for \(\sigma =\frac{1}{2}\). Now fix \(\sigma \in (\frac{1}{2}, s]\). We use (4.47) to bootstrap the base estimate (4.48) to

$$\begin{aligned} \Vert f^-_\delta \Vert _{\widetilde{H}^\sigma _-}\le {\mathcal {F}}(N_s)\Vert \eta _\delta \Vert _{H^\sigma }\Vert v\Vert _{H^s}. \end{aligned}$$
(4.49)

Indeed, we can always take \(\delta \) smaller if necessary so that \(\frac{1}{2}+\delta <\sigma \). Plugging (4.48) into (4.47) with \(\nu =\frac{1}{2}\) yields (4.49) with \(\frac{1}{2}+\delta \) in place of \(\sigma \). Continuing this n steps, n being the greatest integer such that \(\frac{1}{2}+n\delta \le \sigma \), we obtain (4.49) for \(\frac{1}{2}+n\delta \) in place of \(\sigma \). This is justified since \(\nu =\frac{1}{2}+(n-1)\delta \) satisfies (4.45). Thus, for possibly one more step to gain \(\sigma -(\frac{1}{2}+n\delta )\) derivative, we obtain (4.49). The proof of (4.1) is complete.

Finally, (4.2) can be proved similarly except that one uses the contraction estimate (2.33) to estimate \(G^\pm _\delta f^\pm _2\) in (4.43).

5 Proof of Theorem 1.1

Let \(s > 1+\frac{d}{2}\), \(\mu ^- > 0\), \(\mu ^+ \ge 0\), and \({\mathfrak {s}} \in (0, 1]\). Consider an initial datum \(\eta _0\in H^s({\mathbb {R}}^d)\) satisfying

$$\begin{aligned}&\inf _{x\in {\mathbb {R}}^d} \mathrm {RT}(\eta _0)\ge 2{\mathfrak {a}}>0, \end{aligned}$$
(5.1)
$$\begin{aligned}&{{\,\mathrm{dist}\,}}(\eta _0, \Gamma ^\pm )\ge 2h>0. \end{aligned}$$
(5.2)

Theorem 1.1 will be proved in Propositions 5.1, 5.2 and 5.5 below. Precisely, Proposition 5.1 establishes the uniform local well-posedness for the Muskat problem with surface tension belonging to any bounded set, say \({\mathfrak {s}}\in (0, 1]\). Then, in Propositions 5.2 and 5.5, we prove that in appropriate topologies, \(\eta ^{({\mathfrak {s}})}\) converges to \(\eta \) with the rate \(\sqrt{\mathfrak {s}}\) and \({\mathfrak {s}}\) respectively.

Proposition 5.1

There exists a time \(T_*>0\) depending only on \(\Vert \eta _0\Vert _{H^s}\) and \(({\mathfrak {a}}, h, s, \mu ^\pm , {\mathfrak {g}})\) such that the following holds. For each \({\mathfrak {s}}\in (0, 1]\), there exists a unique solution

$$\begin{aligned} \eta ^{({\mathfrak {s}})}\in C([0, T_*]; H^s({\mathbb {R}}^d))\cap L^2([0, T_*]; H^{s+\frac{3}{2}}({\mathbb {R}}^d)) \end{aligned}$$
(5.3)

to the Muskat problem with surface tension \({\mathfrak {s}}\), \(\eta ^{({\mathfrak {s}})}\vert _{t=0}=\eta _0\) and

$$\begin{aligned} \Vert \eta ^{({\mathfrak {s}})}\Vert _{L^\infty ([0,T_*];H^s)}^2+ \Vert \eta ^{({\mathfrak {s}})}\Vert _{L^2([0,T_*];H^{s+\frac{1}{2}})}^2 + {\mathfrak {s}} \Vert \eta ^{({\mathfrak {s}})}\Vert _{L^2([0,T_*];H^{s+\frac{3}{2}})}^2 \le {\mathcal {F}}(\Vert \eta _0\Vert _{H^s}, {\mathfrak {a}}^{-1}) \end{aligned}$$
(5.4)

for some nondecreasing function \({\mathcal {F}}:{\mathbb {R}}^+\times {\mathbb {R}}^+\rightarrow {\mathbb {R}}^+\) depending only on \((h, s, \mu ^\pm , {\mathfrak {g}})\). Furthermore, for all \({\mathfrak {s}}\in (0, 1]\) we have

$$\begin{aligned}&\inf _{t\in [0, T_*]}\inf _{x\in {\mathbb {R}}^d} \mathrm {RT}(\eta ^{({\mathfrak {s}})}(t)) > \frac{3}{2}{\mathfrak {a}}, \end{aligned}$$
(5.5)
$$\begin{aligned}&\inf _{t\in [0, T_*]}{{\,\mathrm{dist}\,}}(\eta ^{({\mathfrak {s}})}(t), \Gamma ^\pm )> \frac{3}{2}h. \end{aligned}$$
(5.6)

Proof

According to Theorems 1.2 and 1.3 in [46], for each initial datum \(\eta _0\in H^s\) satisfying \({{\,\mathrm{dist}\,}}(\eta _0, \Gamma ^\pm )\ge 2h>0\) and for each \({\mathfrak {s}}>0\), there exists \(T_{\mathfrak {s}}>0\) such that the Muskat problem has a unique solution

$$\begin{aligned} \eta ^{({\mathfrak {s}})}\in C([0, T_{\mathfrak {s}}]; H^s)\cap L^2([0, T_{\mathfrak {s}}]; H^{s+\frac{3}{2}}) \end{aligned}$$

satisfying \(\inf _{t\in T_{\mathfrak {s}}}{{\,\mathrm{dist}\,}}(\eta ^{(s)}(t),\Gamma ^\pm )> \frac{3}{2}h\). We stress the continuity in time of the \(H^s\) norm of \(\eta ^{({\mathfrak {s}})}\). Now we have in addition that \(\eta _0\) satisfies the Rayleigh-Taylor condition (5.1). Thus, we define

$$\begin{aligned} T^*_{\mathfrak {s}}=\sup \{T\in (0, T_{\mathfrak {s}}]: \inf _{t\in [0, T]}\inf _{x\in {\mathbb {R}}^d} \mathrm {RT}(\eta ^{({\mathfrak {s}})}(t))> \frac{3}{2}{\mathfrak {a}}\}. \end{aligned}$$
(5.7)

We shall prove that \(T^*_{\mathfrak {s}}>0\) for each \({\mathfrak {s}}\in (0, 1]\) and there exists \(T_*>0\) such that \(T^*_{\mathfrak {s}}\ge T_*\) for all \(s\in (0, 1]\).

Step 1. We claim that there exist \(\theta >0\) depending only on s, and \({\mathcal {F}}_0: {\mathbb {R}}^+\rightarrow {\mathbb {R}}^+\) depending only on \((h, s, \mu ^\pm , {\mathfrak {g}})\) such that

$$\begin{aligned} \Big \Vert [\![{\mathfrak {B}}(\eta ^{({\mathfrak {s}})}(t))J(\eta ^{({\mathfrak {s}})}(t))]\!]\eta ^{({\mathfrak {s}})}(t)- [\![{\mathfrak {B}}(\eta _0)J(\eta _0)]\!]\eta _0\Big \Vert _{L^\infty ({\mathbb {R}}^d)}\le (t^{\frac{\theta }{2}}+ t^{\theta }) {\mathcal {F}}_0(E_{\mathfrak {s}}(t)) \end{aligned}$$
(5.8)

for all \(t\le T_{\mathfrak {s}}\), where

$$\begin{aligned} E_{\mathfrak {s}}(t)=\Vert \eta ^{({\mathfrak {s}})}\Vert _{C([0, t]; H^s)}^2+{\mathfrak {s}}\Vert \eta ^{({\mathfrak {s}})}\Vert ^2_{L^2([0, t]; H^{s+\frac{3}{2}})}. \end{aligned}$$
(5.9)

Set

$$\begin{aligned} A(t)=[\![{\mathfrak {B}}(\eta ^{({\mathfrak {s}})}(t))J(\eta ^{({\mathfrak {s}})}(t))]\!]\eta ^{({\mathfrak {s}})}(t)- [\![{\mathfrak {B}}(\eta _0)J(\eta _0)]\!]\eta _0. \end{aligned}$$

The continuity properties (2.20) and (2.23) of \(J^\pm \) and \(B^\pm \) imply that

$$\begin{aligned} \begin{aligned} \Vert A(t)\Vert _{H^{s-1}}&\le \Vert [\![{\mathfrak {B}}(\eta ^{({\mathfrak {s}})}(t))J(\eta ^{({\mathfrak {s}})}(t))]\!]\eta ^{({\mathfrak {s}})}(t)\Vert _{H^{s-1}}+\Vert [\![{\mathfrak {B}}(\eta _0)J(\eta _0)]\!]\eta _0\Vert _{H^{s-1}}\\&\le {\mathcal {F}}\big (\Vert \eta ^{({\mathfrak {s}})}(t)\Vert _{H^s}\big )+{\mathcal {F}}\big (\Vert \eta ^{({\mathfrak {s}})}(0)\Vert _{H^s}\big )\\&\le {\mathcal {F}}(E_{\mathfrak {s}}(t)). \end{aligned} \end{aligned}$$
(5.10)

On the other hand, denoting \({\mathfrak {B}}^\pm (\eta ^{({\mathfrak {s}})}(t))=\mathfrak {B}^\pm _t\) and \(J^\pm (\eta ^{({\mathfrak {s}})}(t))=J^\pm _t\), we can write

$$\begin{aligned} A(t)&=\big ({\mathfrak {B}}^-_tJ^-_t-\mathfrak {B}^+_tJ^+_t\big )\eta ^{({\mathfrak {s}})}(t)-\big ({\mathfrak {B}}^-_0J^-_0-\mathfrak {B}^+_0J^+_0\big )\eta _0\\&=(\mathfrak {B}^-_t-\mathfrak {B}^-_0)J^-_t\eta ^{({\mathfrak {s}})}(t)+\mathfrak {B}^-_0(J^-_t-J^-_0)\eta ^{({\mathfrak {s}})}(t)+\mathfrak {B}_0^-J_0^-(\eta ^{({\mathfrak {s}})}(t)-\eta _0)\\&\qquad -(\mathfrak {B}^+_t-\mathfrak {B}^+_0)J^+_t\eta ^{({\mathfrak {s}})}(t)-\mathfrak {B}^+_0(J^+_t-J^+_0)\eta ^{({\mathfrak {s}})}(t)-\mathfrak {B}_0^+J_0^+(\eta ^{({\mathfrak {s}})}(t)-\eta _0). \end{aligned}$$

We treat the first two terms since the other terms are either similar or easier. The contraction estimate (2.32) with \(\sigma =s-\frac{1}{2}\) gives

$$\begin{aligned} \Vert G^\pm (\eta ^{({\mathfrak {s}})}(t))-G^\pm (\eta _0)\Vert _{\widetilde{H}^s_\pm \rightarrow H^{s-\frac{3}{2}}}\le {\mathcal {F}}(E_{\mathfrak {s}}(t))\Vert \eta ^{({\mathfrak {s}})}(t)-\eta _0\Vert _{H^{s-\frac{1}{2}}}. \end{aligned}$$

From this and the definition of \(\mathfrak {B}^\pm \) it is easy to prove that

$$\begin{aligned} \Vert \mathfrak {B}^+_t-\mathfrak {B}^+_0\Vert _{ \widetilde{H}^s_\pm \rightarrow H^{s-\frac{3}{2}}}\le {\mathcal {F}}(E_{\mathfrak {s}}(t))\Vert \eta ^{({\mathfrak {s}})}(t)-\eta _0\Vert _{H^{s-\frac{1}{2}}}. \end{aligned}$$

Then recalling the continuity of \(J^\pm \) from \(H^{s-\frac{1}{2}}\rightarrow \widetilde{H}^{s-\frac{1}{2}}_\pm \) we obtain

$$\begin{aligned} \Vert (\mathfrak {B}^-_t-\mathfrak {B}^-_0)J^-_t\eta ^{({\mathfrak {s}})}(t)\Vert _{H^{s-\frac{3}{2}}}\le {\mathcal {F}}(E_{\mathfrak {s}}(t))\Vert \eta ^{({\mathfrak {s}})}(t)-\eta _0\Vert _{H^{s-\frac{1}{2}}}. \end{aligned}$$

Next regarding \(\mathfrak {B}^-_0(J^-_t-J^-_0)\eta ^{({\mathfrak {s}})}(t)\) we use the contraction estimate (4.1) with \(\sigma =s-\frac{1}{2}\)

$$\begin{aligned} \Vert J ^-_t-J^-_0 \Vert _{ H^s\rightarrow \widetilde{H}^{s-\frac{1}{2}}_\pm } \le {\mathcal {F}}(E_{\mathfrak {s}}(t))\Vert \eta ^{({\mathfrak {s}})}(t)-\eta _0\Vert _{H^{s-\frac{1}{2}}} \end{aligned}$$

together with the continuity (2.23) to have

$$\begin{aligned} \Vert \mathfrak {B}^-_0(J^-_t-J^-_0)\eta ^{({\mathfrak {s}})}(t)\Vert _{H^{s-\frac{3}{2}}} \le {\mathcal {F}}(E_{\mathfrak {s}}(t))\Vert \eta ^{({\mathfrak {s}})}(t)-\eta _0\Vert _{H^{s-\frac{1}{2}}}. \end{aligned}$$

Therefore, we arrive at

$$\begin{aligned} \Vert A(t)\Vert _{H^{s-\frac{3}{2}}} \le {\mathcal {F}}(E_{\mathfrak {s}}(t))\Vert \eta ^{({\mathfrak {s}})}(t)-\eta _0\Vert _{H^{s-\frac{1}{2}}}. \end{aligned}$$
(5.11)

To bound \(\Vert \eta ^{({\mathfrak {s}})}(t)-\eta _0\Vert _{H^{s-\frac{1}{2}}}\) we first use the mean-value theorem and equation (2.18) to have

$$\begin{aligned} \Vert \eta ^{({\mathfrak {s}})}(t)-\eta _0\Vert _{H^{-\frac{1}{2}}}&\le \int _0^t \Vert \partial _t\eta ^{({\mathfrak {s}})}(\tau )\Vert _{H^{-\frac{1}{2}}}d\tau \\&\le {\mathcal {F}}(E_{\mathfrak {s}}(t))\int _0^t {\mathfrak {g}}\Vert \eta ^{({\mathfrak {s}})}(\tau )\Vert _{H^\frac{1}{2}}+{\mathfrak {s}}\Vert H\big (\eta ^{({\mathfrak {s}})}(\tau )\big )\Vert _{H^\frac{1}{2}}d\tau \\&\le {\mathcal {F}}(E_{\mathfrak {s}}(t)) \Big (t{\mathfrak {g}}\Vert \eta ^{({\mathfrak {s}})}\Vert _{C([0, t]; H^s)}+t^\frac{1}{2}{\mathfrak {s}}\Vert \eta ^{({\mathfrak {s}})}\Vert _{L^2([0, t]; H^{s+\frac{3}{2}})}\Big )\\&\le (t^\frac{1}{2}+t){\mathcal {F}}(E_{\mathfrak {s}}(t)) \Big ({\mathfrak {g}}\Vert \eta ^{({\mathfrak {s}})}\Vert _{C([0, t]; H^s)}+\sqrt{\mathfrak {s}}\Vert \eta ^{({\mathfrak {s}})}\Vert _{L^2([0, t]; H^{s+\frac{3}{2}})}\Big ) \end{aligned}$$

for all \({\mathfrak {s}}\in (0, 1]\) and \(t\le T_{\mathfrak {s}}\). Interpolating this with the obvious bound \(\Vert \eta ^{({\mathfrak {s}})}(t)-\eta _0\Vert _{H^s}\le {\mathcal {F}}(M_s(t))\) gives

$$\begin{aligned} \Vert \eta ^{({\mathfrak {s}})}(t)-\eta _0\Vert _{H^{s-\frac{1}{2}}}\le (t^{\frac{\theta _0}{2}}+ t^{\theta _0}){\mathcal {F}}(E_{\mathfrak {s}}(t)) \end{aligned}$$
(5.12)

for some \(\theta _0\in (0, 1)\) and \({\mathcal {F}}\) depending only on \((h, s, \mu ^\pm , {\mathfrak {g}})\). Then in view of (5.11), this implies

$$\begin{aligned} \Vert A(t)\Vert _{H^{s-\frac{3}{2}}}\le (t^{\frac{\theta _0}{2}}+ t^{\theta _0}){\mathcal {F}}(E_{\mathfrak {s}}(t)). \end{aligned}$$

Fixing \(s'\in (\max \{1+\frac{d}{2}, s-\frac{3}{2}\}, s)\) and interpolating this with the \(H^s\) bound (5.10) we obtain

$$\begin{aligned} \Vert A(t)\Vert _{H^{s'-1}}\le (t^{\frac{\theta }{2}}+ t^{\theta }){\mathcal {F}}(E_{\mathfrak {s}}(t)) \end{aligned}$$

for some \(\theta \in (0, 1)\). Using the embedding \(H^{s'-1}\subset L^\infty ({\mathbb {R}}^d)\) we conclude the proof of (5.8).

Step 2. We note that (5.8) implies the continuity of

$$\begin{aligned}{}[0, T_{\mathfrak {s}}]\ni t\mapsto \inf _{x\in {\mathbb {R}}^d} \mathrm {RT}(\eta ^{({\mathfrak {s}})}(t)). \end{aligned}$$

Thus, in view of the definition (5.7) and the initial condition (5.1), we have \(T^*_{\mathfrak {s}}>0\) for all \({\mathfrak {s}}\in (0, 1]\).

By the definition of \(T^*_{\mathfrak {s}}\), conditions (3.1), (3.2) and (3.3) in Proposition 3.1 are satisfied for all \(T\le T^*_{\mathfrak {s}}\). Thus, the estimates (3.4) and (3.5) imply the existence of a strictly increasing \({\mathcal {F}}_2:{\mathbb {R}}^+\times {\mathbb {R}}^+\rightarrow {\mathbb {R}}^+\) depending only on \((h, s, \mu ^\pm , {\mathfrak {g}})\) such that

$$\begin{aligned} {\mathcal {F}}_2(m, 0)>m^2\quad \forall m>0 \end{aligned}$$
(5.13)

and

$$\begin{aligned} M_{\mathfrak {s}}(T)\le {\mathcal {F}}_2\Big ( \Vert \eta _0\Vert _{H^s} +T{\mathcal {F}}_2\big (M_{\mathfrak {s}}(T),{\mathfrak {a}}^{-1}\big ), {\mathfrak {a}}^{-1}\Big ) \end{aligned}$$
(5.14)

for all \({\mathfrak {s}}\in (0, 1]\) and \(T\le T_{\mathfrak {s}}^*\), where

$$\begin{aligned} M_{\mathfrak {s}}(T) =\Vert \eta ^{({\mathfrak {s}})}\Vert _{C([0, T]; H^s)}^2+\Vert \eta ^{({\mathfrak {s}})} \Vert _{L^2([0, T]; H^{s+\frac{1}{2}})}^2+{\mathfrak {s}}\Vert \eta ^{({\mathfrak {s}})}\Vert ^2_{L^2([0, T]; H^{s+\frac{3}{2}})}. \end{aligned}$$

Set

$$\begin{aligned} T_2= \frac{\Vert \eta _0\Vert _{H^s}}{2{\mathcal {F}}_2\Big ({\mathcal {F}}_2\big ( 2\Vert \eta _0\Vert _{H^s},{\mathfrak {a}}^{-1}\big ),{\mathfrak {a}}^{-1}\Big )} \end{aligned}$$
(5.15)

independent of \({\mathfrak {s}}\). We claim that

$$\begin{aligned} M_{\mathfrak {s}}(T)\le K_0:={\mathcal {F}}_2\big (2\Vert \eta _0\Vert _{H^s}, {\mathfrak {a}}^{-1}\big )\quad \forall T\le \min \{T_2, T^*_{\mathfrak {s}}\},\quad \forall {\mathfrak {s}}\in (0, 1]. \end{aligned}$$
(5.16)

Assume not, then there exists \({\mathfrak {s}}_0\in (0, 1]\) and \(T_3\le \min \{T_2, T^*_{{\mathfrak {s}}_0}\}\) such that \(M_{{\mathfrak {s}}_0}(T_3)> K_0\). Since \(M_{{\mathfrak {s}}_0}(0)=\Vert \eta _0\Vert _{H^s}^2<K_0\), the continuity of \(T\mapsto E_{{\mathfrak {s}}_0}(T)\) then yields the existence of \(T_4\in (0, T_3)\) such that \(M_{{\mathfrak {s}}_0}(T_4)= K_0\). Consequently, at \(T=T_4\), (5.14) gives

$$\begin{aligned} {\mathcal {F}}_2\big ( 2\Vert \eta _0\Vert _{H^s},{\mathfrak {a}}^{-1}\big )&\le {\mathcal {F}}_2\Big ( \Vert \eta _0\Vert _{H^s} +T_4{\mathcal {F}}_2\Big ({\mathcal {F}}_2\big ( 2\Vert \eta _0\Vert _{H^s},{\mathfrak {a}}^{-1}\big ), {\mathfrak {a}}^{-1}\Big ),{\mathfrak {a}}^{-1}\Big )\\&\le {\mathcal {F}}_2\Big ( \Vert \eta _0\Vert _{H^s} +T_2{\mathcal {F}}_2\Big ({\mathcal {F}}_2\big ( 2\Vert \eta _0\Vert _{H^s},{\mathfrak {a}}^{-1}\big ),{\mathfrak {a}}^{-1}\Big ), {\mathfrak {a}}^{-1}\Big )\\&\le {\mathcal {F}}_2\Big ( \frac{3}{2}\Vert \eta _0\Vert _{H^s},{\mathfrak {a}}^{-1}\Big ), \end{aligned}$$

where we have used the definition of \(T_2\) in the last inequality. This contradicts the fact that \({\mathcal {F}}_2\) was chosen to be strictly increasing.

Now for all \(T\le \min \{T_2, T^*_{\mathfrak {s}}\}\), we use (5.8) and the fact that \(E_{\mathfrak {s}}(\cdot )\le M_{\mathfrak {s}}(\cdot )\) to have

$$\begin{aligned} \inf _{t\in [0, T]}\inf _{x\in {\mathbb {R}}^d} \mathrm {RT} (\eta ^{({\mathfrak {s}})}(t))&\ge \inf _{x\in {\mathbb {R}}^d} \mathrm {RT}(\eta _0)- (T^{\frac{\theta }{2}}+ T^{\theta }) {\mathcal {F}}_0(M_{\mathfrak {s}}(T))\\&\ge 2{\mathfrak {a}}- (T^{\frac{\theta }{2}}+ T^{\theta }) {\mathcal {F}}_0(K_0). \end{aligned}$$

Choosing \(T_*\le T_2\) sufficiently small so that

$$\begin{aligned} (T_*^{\frac{\theta }{2}}+ T_*^{\theta }) {\mathcal {F}}_0(K_0)<\frac{1}{2}{\mathfrak {a}}\end{aligned}$$
(5.17)

we obtain

$$\begin{aligned} \inf _{t\in [0, T]}\inf _{x\in {\mathbb {R}}^d} \mathrm {RT}(\eta ^{({\mathfrak {s}})}(t))>\frac{3}{2}{\mathfrak {a}}\quad \forall T\le \min \{T_*, T_{\mathfrak {s}}^*\}. \end{aligned}$$
(5.18)

Clearly, \(T_*\) is independent of \({\mathfrak {s}}\). Moreover, (5.18) and the definition of \(T_{\mathfrak {s}}^*\) show that \(T_*\le T^*_{\mathfrak {s}}\) for all \({\mathfrak {s}}\in (0, 1]\). Finally, since \(T_*\le \min \{T_2, T_{\mathfrak {s}}^*\}\), (5.16) and the definition of \(T^*_{\mathfrak {s}}\) guarantee that the estimates (5.4), (5.5) and (5.6) hold true. \(\quad \square \)

Proposition 5.2

There exists \({\mathcal {F}}:{\mathbb {R}}^+\times {\mathbb {R}}^+\rightarrow {\mathbb {R}}^+\) depending only on \((h, s, \mu ^\pm , {\mathfrak {g}})\) such that

$$\begin{aligned} \begin{aligned}&\Vert \eta ^{({\mathfrak {s}}_1)}-\eta ^{({\mathfrak {s}}_2)}\Vert _{L^\infty ([0,T_*];H^{s-1})}^2+ \Vert \eta ^{({\mathfrak {s}}_1)}-\eta ^{({\mathfrak {s}}_2)}\Vert _{L^2([0,T_*];H^{s-\frac{1}{2}})}^2 \\&\quad \le ({\mathfrak {s}}_1+{\mathfrak {s}}_2){\mathcal {F}}\big (\Vert \eta _0\Vert _{H^s}, {\mathfrak {a}}^{-1}\big ) \end{aligned} \end{aligned}$$
(5.19)

for all \({\mathfrak {s}}_1\) and \({\mathfrak {s}}_2\) in (0, 1].

Proof

Denote \(\eta _j=\eta ^{({\mathfrak {s}}_j)}\), \(j=1, 2\), and \(\eta _\delta =\eta _1-\eta _2\) which exists on \([0, T_*]\). We fix \(\delta \in \big (0, \min (\frac{1}{2}, s-1-\frac{d}{2})\big )\). From (2.18) we have that \(\eta _\delta \) evolves according to

$$\begin{aligned} (\mu ^+ + \mu ^-) \partial _t \eta _\delta = -{\mathfrak {g}} (L_\delta \eta _1+ L_2\eta _\delta )- {\mathfrak {s}}_1 L_1 H(\eta _1)+{\mathfrak {s}}_2 L_2 H(\eta _2). \end{aligned}$$
(5.20)

By (2.20) and (2.31),

$$\begin{aligned} \Vert L_j H(\eta _j)\Vert _{H^{s-\frac{3}{2}}}\le {\mathcal {F}}(\Vert \eta _j\Vert _{H^s})\Vert H(\eta _j)\Vert _{H^{s-\frac{1}{2}}}\le {\mathcal {F}}(\Vert \eta _j\Vert _{H^s})\Vert \eta _j\Vert _{H^{s+\frac{3}{2}}}. \end{aligned}$$
(5.21)

We now paralinearize \(L_2\) and \(L_\delta \). Applying (2.26) with \(\sigma =s-\frac{1}{2}-\delta \) gives

$$\begin{aligned} L_2 \eta _\delta = T_{\lambda _2}\eta _\delta + O_{H^{s-\frac{1}{2}-\delta } \rightarrow H^{s -\frac{3}{2}}}({\mathcal {F}}(N_s))\eta _\delta . \end{aligned}$$
(5.22)

As for \(L_\delta \) we write

$$\begin{aligned} L_\delta&= \sum _{\pm = +,-} G_\delta ^\pm J _1^\pm + G^\pm _2 J _\delta ^\pm . \end{aligned}$$
(5.23)

Using (2.25) at \(\sigma = s-\frac{1}{2}-\delta \), we have

$$\begin{aligned}&\sum _{\pm = +,-} G_2^\pm J _\delta ^\pm = T_{\lambda _2}(J _\delta ^--J^+_\delta ) + \sum _{\pm = +,-} O_{\widetilde{H}^{s-\frac{1}{2}-\delta }_\pm \rightarrow H^{s-\frac{3}{2}}}({\mathcal {F}}(N_s)) J _\delta ^\pm . \end{aligned}$$

Recall that \(N_s\) is given by (4.3). However,

$$\begin{aligned} J _\delta ^--J^+_\delta =(J_1^--J_1^+)-(J_2^--J_2^+)= \text {Id} - \text {Id} = 0 \end{aligned}$$

and by virtue of Proposition 4.1 (with \(\sigma =s-\frac{1}{2}-\delta )\),

$$\begin{aligned}&J^\pm _\delta =O_{H^s\rightarrow \widetilde{H}^{s-\frac{1}{2}-\delta }_\pm }(\Vert \eta _\delta \Vert _{H^{s-\frac{1}{2}-\delta }}). \end{aligned}$$

We thus obtain

$$\begin{aligned}&\sum _{\pm = +,-}G_2^\pm J _\delta ^\pm = O_{H^{s} \rightarrow H^{s -\frac{3}{2}}}\big ({\mathcal {F}}(N_s)\Vert \eta _\delta \Vert _{H^{s-\frac{1}{2}-\delta }}\big ). \end{aligned}$$
(5.24)

As for \(G_\delta ^\pm J _1^\pm \), we apply Theorem 2.8 with \(\sigma =s-\frac{1}{2}\) and (2.20) to have

and hence

$$\begin{aligned} \sum _{\pm = +,-} G_\delta ^\pm J _1^\pm f=-T_{ \lambda _1 [\![{\mathfrak {B}}_1 J _1]\!]f}\eta _\delta -T_{[\![{\mathfrak {V}}_1 J _1]\!]f} \cdot \nabla \eta _\delta + O_{H^s \rightarrow H^{s-\frac{3}{2}}}\big ({\mathcal {F}}(N_s) \Vert \eta _\delta \Vert _{H^{s-\frac{1}{2}-\delta }}\big )f. \end{aligned}$$

Combining this with (5.24) yields

$$\begin{aligned} L_\delta f =- T_{ \lambda _1 [\![{\mathfrak {B}}_1 J _1]\!]f}\eta _\delta -T_{[\![{\mathfrak {V}}_1 J _1]\!]f} \cdot \nabla \eta _\delta + O_{H^s \rightarrow H^{s-\frac{3}{2}}}\big ({\mathcal {F}}(N_s) \Vert \eta _\delta \Vert _{H^{s-\frac{1}{2}-\delta }}\big )f. \end{aligned}$$
(5.25)

From (5.22) and (5.25) we have

$$\begin{aligned} L_2\eta _\delta +L_\delta \eta _1&=T_{\lambda _2}\eta _\delta - T_{ \lambda _1 [\![{\mathfrak {B}}_1 J _1]\!]\eta _1}\eta _\delta -T_{[\![{\mathfrak {V}}_1 J _1]\!]\eta _1} \cdot \nabla \eta _\delta \\&\qquad + O_{H^{s-\frac{1}{2}-\delta } \rightarrow H^{s -\frac{3}{2}}}({\mathcal {F}}(N_s))\eta _\delta + O_{H^s \rightarrow H^{s-\frac{3}{2}}}\big ({\mathcal {F}}(N_s) \Vert \eta _\delta \Vert _{H^{s-\frac{1}{2}-\delta }}\big )\eta _1 \end{aligned}$$

Interchanging \(\eta _1\) and \(\eta _2\) gives

$$\begin{aligned} L_1\eta _\delta +L_\delta \eta _2&=T_{\lambda _1}\eta _\delta - T_{ \lambda _2 [\![{\mathfrak {B}}_2 J _2]\!]\eta _2}\eta _\delta -T_{[\![{\mathfrak {V}}_2 J _2]\!]\eta _2} \cdot \nabla \eta _\delta \\&\qquad + O_{H^{s-\frac{1}{2}-\delta } \rightarrow H^{s -\frac{3}{2}}}({\mathcal {F}}(N_s))\eta _\delta +O_{H^s \rightarrow H^{s-\frac{3}{2}}}\big ({\mathcal {F}}(N_s) \Vert \eta _\delta \Vert _{H^{s-\frac{1}{2}-\delta }}\big )\eta _2. \end{aligned}$$

But \(L_1\eta _\delta +L_\delta \eta _2=L_2\eta _\delta +L_\delta \eta _1\), thus taking the average of the above identities yields

$$\begin{aligned} L_2\eta _\delta +L_\delta \eta _1&= T_{ (\lambda (1-[\![{\mathfrak {B}} J ]\!]\eta ))_\alpha }\eta _\delta -T_{([\![{\mathfrak {V}} J]\!]\eta )_\alpha } \cdot \nabla \eta _\delta + O_{H^{s-\frac{1}{2}-\delta } \rightarrow H^{s -\frac{3}{2}}}({\mathcal {F}}(N_s))\eta _\delta \\&\quad + O_{H^s \rightarrow H^{s-\frac{3}{2}}}\big ({\mathcal {F}}(N_s) \Vert \eta _\delta \Vert _{H^{s-\frac{1}{2}-\delta }}\big )(\eta _1+\eta _2), \end{aligned}$$

where

$$\begin{aligned} \big (\lambda (1-[\![{\mathfrak {B}} J ]\!]\eta )\big )_\alpha =\frac{1}{2}\Big (\lambda _1 (1-[\![{\mathfrak {B}}_1 J_1 ]\!]\eta _1)+\lambda _2 (1-[\![{\mathfrak {B}}_2 J_2 ]\!]\eta _2)\Big ) \end{aligned}$$

and similarly for \(([\![{\mathfrak {V}} J]\!]\eta )_\alpha \). It then follows from (5.20) and (5.21) that

$$\begin{aligned} \begin{aligned}&(\mu ^+ + \mu ^-) \partial _t \eta _\delta =- {\mathfrak {g}}T_{ (\lambda (1-[\![{\mathfrak {B}} J ]\!]\eta ))_\alpha }\eta _\delta +{\mathfrak {g}}T_{([\![{\mathfrak {V}} J]\!]\eta )_\alpha } \cdot \nabla \eta _\delta +{\mathcal {R}}_1,\\&\Vert {\mathcal {R}}_1\Vert _{H^{s-\frac{3}{2}}}\le {\mathfrak {g}}{\mathcal {F}}(N_s)\Vert \eta _\delta \Vert _{H^{s-\frac{1}{2}-\delta }} +({\mathfrak {s}}_1+{\mathfrak {s}}_2){\mathcal {F}}(N_s)N_{s+\frac{3}{2}}. \end{aligned} \end{aligned}$$
(5.26)

Next we perform \(H^{s-1}\) energy estimate for (5.26). Introduce \(\eta _{\delta , s-1}=\langle D\rangle ^{s-1}\eta _\delta \). Upon commuting (5.26) with \(\langle D\rangle ^{s-1}\) and applying Theorem A.3 we arrive at

$$\begin{aligned} \begin{aligned}&(\mu ^+ + \mu ^-) \partial _t \eta _{\delta , s-1} =-{\mathfrak {g}}T_{ (\lambda (1-[\![{\mathfrak {B}} J ]\!]\eta ))_\alpha }\eta _{\delta , s-1} +{\mathfrak {g}}i\mathrm {Re}\big (T_{([\![{\mathfrak {V}} J]\!]\eta )_\alpha \cdot \xi }\big ) \eta _{\delta , s-1}+{\mathcal {R}},\\&\Vert {\mathcal {R}}\Vert _{H^{-\frac{1}{2}}}\le {\mathfrak {g}}{\mathcal {F}}(N_s)\Vert \eta _{\delta , s-1}\Vert _{H^{\frac{1}{2}-\delta }}+{\mathcal {F}}(N_s)\sum _{j=1}^2{\mathfrak {s}}_j\Vert \eta _j\Vert _{H^{s+\frac{3}{2}}}, \end{aligned} \end{aligned}$$
(5.27)

where \({\mathcal {F}}\) depends only on \((h, s, \mu ^\pm )\). Moreover, the uniform estimate () implies that

$$\begin{aligned} N_s\le {\mathcal {F}}(\Vert \eta _0\Vert _{H^s}, {\mathfrak {a}}^{-1}). \end{aligned}$$
(5.28)

Testing (5.27) against \(\eta _{\delta , s-1}\) yields

$$\begin{aligned} \frac{(\mu ^+ + \mu ^-)}{2}\frac{d}{dt}\Vert \eta _{\delta , s-1} \Vert _{L^2}^2 =&- {\mathfrak {g}} \big ( T_{ (\lambda (1-[\![{\mathfrak {B}} J ]\!]\eta ))_\alpha }\eta _{\delta , s-1},\eta _{\delta , s-1}\big )_{L^2} +({\mathcal {R}}, \eta _{\delta , s-1})_{L^2}, \end{aligned}$$

where we have used the fact that \(i\mathrm {Re}\big (T_{([\![{\mathfrak {V}} J]\!]\eta )_\alpha \cdot \xi }\big )\) is skew-adjoint.

From (5.5) we have that \( (\lambda (1-[\![{\mathfrak {B}} J ]\!]\eta ))_\alpha \) is an elliptic symbol in \(\Gamma ^1_\delta \):

$$\begin{aligned} (\lambda (1-[\![{\mathfrak {B}} J ]\!]\eta ))_\alpha \ge {\mathfrak {a}}|\xi |,\quad M^1_\delta \big ( (\lambda (1-[\![{\mathfrak {B}} J ]\!]\eta ))_\alpha \big )\le {\mathcal {F}}(N_s, {\mathfrak {a}}^{-1}), \end{aligned}$$

where \({\mathcal {F}}\) depends only on \((h, s, \mu ^\pm , {\mathfrak {g}})\). Applying the Gårding inequality (A.9), we have

$$\begin{aligned} \Vert \eta _{\delta , s-1} \Vert ^2_{H^\frac{1}{2}}\le {\mathcal {F}}(\Vert \eta _0\Vert _{H^s}, {\mathfrak {a}}^{-1})\Big (\big ( T_{ (\lambda (1-[\![{\mathfrak {B}} J ]\!]\eta ))_\alpha }\eta _{\delta , s-1},\eta _{\delta , s-1}\big )_{L^2}+\Vert \eta _{\delta , s-1}\Vert _{H^\frac{1}{2}}\Vert \eta _{\delta , s-1}\Vert _{H^{\frac{1}{2}-\delta }} \Big ). \end{aligned}$$

This combined with the estimate for \({\mathcal {R}}\) in (5.27) and (5.28) implies

$$\begin{aligned} \begin{aligned}&\frac{(\mu ^+ + \mu ^-)}{2}\frac{d}{dt}\Vert \eta _{\delta , s-1} \Vert _{L^2}^2 +\frac{{\mathfrak {g}}}{{\mathcal {Q}}_{T_*}}\Vert \eta _{\delta , s-1} \Vert ^2_{H^\frac{1}{2}}\\&\quad \le {\mathfrak {g}}{\mathcal {Q}}_{T_*}\Vert \eta _{\delta , s-1}\Vert _{H^{\frac{1}{2}-\delta }}\Vert \eta _{\delta , s-1}\Vert _{H^{\frac{1}{2}}}+{\mathcal {Q}}_{T_*}\Vert \eta _{\delta , s-1}\Vert _{H^{\frac{1}{2}}}\sum _{j=1}^2{\mathfrak {s}}_j\Vert \eta _j\Vert _{H^{s+\frac{3}{2}}}, \end{aligned} \end{aligned}$$
(5.29)

where

$$\begin{aligned} {\mathcal {Q}}_{T_*}={\mathcal {F}}\big (\Vert \eta _0\Vert _{H^s}, {\mathfrak {a}}^{-1}\big ), \end{aligned}$$
(5.30)

\({\mathcal {F}}\) depending only on \((h, s, \mu ^\pm , {\mathfrak {g}})\). Using interpolation and Young’s inequality, we have

$$\begin{aligned} \begin{aligned}&\Vert \eta _{\delta , s-1}\Vert _{H^{\frac{1}{2}-\delta }}\Vert \eta _s\Vert _{H^{\frac{1}{2}}} \le \Vert \eta _{\delta , s-1}\Vert _{L^2}^{2\delta }\Vert \eta _{\delta , s-1}\Vert _{H^{\frac{1}{2}}}^{2(1-\delta )} \\&\quad \le (10{\mathcal {Q}}_{T_*})^{\frac{2(1-\delta )}{\delta }}\Vert \eta _{\delta , s-1}\Vert _{L^2}^2 + \frac{1}{100{\mathcal {Q}}^2_{T_*}}\Vert \eta _{\delta , s-1}\Vert _{H^{\frac{1}{2}}}^2,\\&{\mathfrak {s}}_j\Vert \eta _j\Vert _{H^{s+\frac{3}{2}}}\Vert \eta _{\delta , s-1}\Vert _{H^{\frac{1}{2}}}\le {\mathfrak {s}}_j^2100{\mathcal {Q}}^2_{T_*}\Vert \eta _j\Vert _{H^{s+\frac{3}{2}}}^2+ \frac{1}{100{\mathcal {Q}}^2_{T_*}}\Vert \eta _{\delta , s-1}\Vert _{H^{\frac{1}{2}}}^2. \end{aligned} \end{aligned}$$
(5.31)

Thus, for possibly a larger \({\mathcal {F}}\) in \({\mathcal {Q}}_{T_*}\), we obtain

$$\begin{aligned} \begin{aligned}&\frac{(\mu ^+ + \mu ^-)}{2}\frac{d}{dt}\Vert \eta _{\delta , s-1} \Vert _{L^2}^2 +\frac{{\mathfrak {g}}}{{\mathcal {Q}}_{T_*}}\Vert \eta _{\delta , s-1}\Vert _{H^{\frac{1}{2}}}^2\le {\mathfrak {g}}{\mathcal {Q}}_{T_1}\Vert \eta _{\delta , s-1}\Vert ^2_{L^2}+{\mathcal {Q}}_{T_*} \sum _{j=1}^2{\mathfrak {s}}_j^2\Vert \eta _j\Vert _{H^{s+\frac{3}{2}}}^2. \end{aligned} \end{aligned}$$
(5.32)

Finally, since \(\eta _\delta \vert _{t=0}=0\) and by (5.4)

$$\begin{aligned} {\mathfrak {s}}_j \int _0^{T_*}\Vert \eta _j\Vert ^2_{H^{s+\frac{3}{2}}}dt \le {\mathcal {F}}(\Vert \eta _0\Vert _{H^s}, {\mathfrak {a}}^{-1}), \end{aligned}$$

an application of Grönwall’s lemma leads to the estimate (5.19). \(\quad \square \)

Now let \({\mathfrak {s}}_n\rightarrow 0\) and rename \(\eta _n=\eta ^{({\mathfrak {s}}_n)}\) solution to the Muskat problem with surface tension \({\mathfrak {s}}_n\) on \([0, T_*]\). The uniform estimates in (5.4) show that along a subsequence \(\eta _n\) converges weakly-* to

$$\begin{aligned} \eta \in L^\infty ([0, T_*]; H^s)\cap L^2([0, T_*]; H^{s+\frac{1}{2}}) \end{aligned}$$
(5.33)

together with the bounds

$$\begin{aligned}&\Vert \eta \Vert _{L^\infty ([0,T_*];H^s)}+ \Vert \eta \Vert _{L^2([0,T_*]; H^{s+\frac{1}{2}})} \le {\mathcal {F}}(\Vert \eta _0\Vert _{H^s}, {\mathfrak {a}}^{-1}). \end{aligned}$$
(5.34)

The estimate (5.19) implies that \((\eta _n)_n\) is a Cauchy sequence in \(C([0, T_*]; H^{s-1})\cap L^2([0, T_*]; H^{s-\frac{1}{2}})\). Therefore,

$$\begin{aligned} \eta _n\rightarrow \eta \quad \text {in}~C([0, T_*]; H^{s-1})\cap L^2([0, T_*]; H^{s-\frac{1}{2}}); \end{aligned}$$
(5.35)

in particular, \(\eta \vert _{t=0}=\eta _0\). Moreover, by interpolating between \(L^\infty _t H^s_x\) and \(C_t H^{s-1}\), we deduce that \(\eta \in C([0, T_*]; H^{s'})\) for all \(s'<s\). Since \(\eta _n\rightarrow \eta \) in \(C_t H^{s-1}_x\subset C_tL^\infty _x\), (5.6) gives

$$\begin{aligned} \inf _{t\in [0, T_*]}{{\,\mathrm{dist}\,}}(\eta (t), \Gamma ^\pm )\ge \frac{3}{2}h. \end{aligned}$$
(5.36)

Lemma 5.3

\(\eta \) is a solution on \([0, T_*]\) of the Muskat problem without surface tension with initial data \(\eta _0\).

Proof

For each n, we have from (2.18) that

$$\begin{aligned} \partial _t\eta _n +\frac{1}{\mu ^+ + \mu ^-}L(\eta _n) ({\mathfrak {g}} \eta _n + {\mathfrak {s}} H(\eta _n))= 0 \end{aligned}$$
(5.37)

For any compactly supported test function \(\varphi \in C^\infty ((0, T_*)\times {\mathbb {R}}^d)\), we have

$$\begin{aligned} \int _0^{T_*}\eta _n\partial _t\varphi dxdt=\frac{1}{\mu ^+ + \mu ^-}\int _0^{T_*}\int _{{\mathbb {R}}^d}\varphi L(\eta _n)( {\mathfrak {g}} \eta _n + {\mathfrak {s}} H(\eta _n))dxdt. \end{aligned}$$
(5.38)

Clearly, (5.35) implies that

$$\begin{aligned} \int _0^{T_*}\eta _n\partial _t\varphi dxdt\rightarrow \int _0^{T_*}\eta \partial _t\varphi dxdt. \end{aligned}$$

The continuity (2.20) of L combined with (2.31) and the uniform bound (5.4) yields

$$\begin{aligned} {\mathfrak {s}}_n \Vert L(\eta _n)H(\eta _n)\Vert _{L^2([0, T_*]; H^{s-\frac{3}{2}})}&\le {\mathcal {F}}(\Vert \eta _n\Vert _{L^\infty ([0, T_*]; H^s)}){\mathfrak {s}}_n\Vert \eta _n\Vert _{L^2([0, T_*]; H^{s+\frac{3}{2}})}\\&\lesssim \sqrt{{\mathfrak {s}}_n}{\mathcal {F}}(\Vert \eta _0\Vert _{H^s}, {\mathfrak {a}}^{-1}). \end{aligned}$$

Since \(s-\frac{3}{2}>0\), this implies

$$\begin{aligned} \left| \int _0^{T_*}\int _{{\mathbb {R}}^d}\varphi L(\eta _n) \big ({\mathfrak {s}}_nH(\eta _n)\big )dxdt\right|&\le \Vert \varphi \Vert _{L^2_{x, t}}\Vert L(\eta _n) \big ({\mathfrak {s}}_nH(\eta _n)\big )\Vert _{L^2_{x, t}}\\&\lesssim \sqrt{{\mathfrak {s}}_n}\Vert \varphi \Vert _{L^2_{x, t}}{\mathcal {F}}(\Vert \eta _0\Vert _{H^s}, {\mathfrak {a}}^{-1})\rightarrow 0. \end{aligned}$$

Next we write

$$\begin{aligned} L(\eta _n)\eta _n-L(\eta )\eta&=\big (L(\eta _n) -L(\eta )\big )\eta _n+L(\eta _n)(\eta _n-\eta )\\&=\frac{\mu ^++\mu ^-}{\mu ^-}\big (G^-(\eta _n)J^- (\eta _n)-G^-(\eta )J^-(\eta )\big )\eta _n+L(\eta _n)(\eta _n-\eta )\\&=\frac{\mu ^++\mu ^-}{\mu ^-}\Big \{\big (G^-(\eta _n) -G^-(\eta )\big )J^-(\eta _n)\eta _n-G^-(\eta ) \big (J^-(\eta _n)-J^-(\eta )\big )\eta _n\Big \}\\&\qquad +L(\eta _n)(\eta _n-\eta ). \end{aligned}$$

Combining (2.32) and (2.20) we obtain

$$\begin{aligned}&\Vert \big (G^-(\eta _n)-G^-(\eta )\big )J^-(\eta _n)\eta _n\Vert _{L^2([0, T_*]; H^{s-\frac{3}{2}})}\\&\quad \le {\mathcal {F}}(\Vert (\eta _n, \eta )\Vert _{L^\infty ([0, T_*]; H^s)})\Vert \eta _n-\eta \Vert _{L^2([0, T_*]; H^{s-\frac{1}{2}})}. \end{aligned}$$

On the other hand, (2.20) and (4.1) yield

$$\begin{aligned}&\Vert G^-(\eta )\big (J^-(\eta _n)-J^-(\eta )\big )\eta _n\Vert _{L^2([0, T_*]; H^{s-\frac{3}{2}})}\\&\quad \le {\mathcal {F}}(\Vert (\eta _n, \eta )\Vert _{L^\infty ([0, T_*]; H^s)})\Vert \eta _n-\eta \Vert _{L^2([0, T_*]; H^{s-\frac{1}{2}})}. \end{aligned}$$

Finally, by (2.20) we have

$$\begin{aligned} \Vert L(\eta _n)(\eta _n-\eta )\Vert _{L^2([0, T_*]; H^{s-\frac{3}{2}})}\le {\mathcal {F}}(\Vert \eta _n\Vert _{L^\infty ([0, T_*]; H^s)})\Vert \eta _n-\eta \Vert _{L^2([0, T_*]; H^{s-\frac{1}{2}})}. \end{aligned}$$

Putting together the above considerations, we obtain

$$\begin{aligned}&\left| \int _0^{T_*}\int _{{\mathbb {R}}^d}\varphi \big (L(\eta _n) \eta _n-L(\eta )\eta \big )\right| \\&\quad \le \Vert \varphi \Vert _{L^2_{x,t}}{\mathcal {F}}(\Vert (\eta _n, \eta )\Vert _{L^\infty ([0, T_*]; H^s)})\Vert \eta _n-\eta \Vert _{L^2([0, T_*]; H^{s-\frac{1}{2}})}\rightarrow 0 \end{aligned}$$

by virtue of the strong convergence (5.35) and the uniform \(H^s\) bound in (5.4). We have proved that

$$\begin{aligned} \int _0^{T_*}\eta \partial _t\varphi dxdt=\frac{1}{\mu ^+ + \mu ^-}\int _0^{T_*}\int _{{\mathbb {R}}^d}\varphi L(\eta )({\mathfrak {g}} \eta )dxdt \end{aligned}$$

for all compactly supported smooth test functions \(\varphi \). Therefore, \(\eta \) is a solution on \([0, T_*]\) of the Muskat problem without surface tension. \(\quad \square \)

Lemma 5.4

We have

$$\begin{aligned} \inf _{t\in [0, T_*]}\inf _{x\in {\mathbb {R}}^d} \mathrm {RT}(\eta (t)) \ge \frac{3}{2}{\mathfrak {a}}. \end{aligned}$$
(5.39)

Proof

Set

$$\begin{aligned} K=[\![{\mathfrak {B}}(\eta _n)J(\eta _n)]\!]\eta _n-[\![{\mathfrak {B}}(\eta )J(\eta )]\!]\eta . \end{aligned}$$

Arguing as in the proof of (5.11) we find that

$$\begin{aligned} \begin{aligned} \Vert K\Vert _{H^{s-2}}&\le {\mathcal {F}}(\Vert (\eta _n, \eta ) \Vert _{H^s})\Vert \eta _n-\eta \Vert _{H^{s-1}}\\&\lesssim {\mathcal {F}}(\Vert \eta _0\Vert _{H^s}, {\mathfrak {a}}^{-1})\Vert \eta _n-\eta \Vert _{H^{s-1}}. \end{aligned} \end{aligned}$$

On the other hand, by estimating each term in K we have

$$\begin{aligned} \Vert K\Vert _{H^{s-1}}\le {\mathcal {F}}(\Vert (\eta _n, \eta ) \Vert _{H^s}) \lesssim {\mathcal {F}}(\Vert \eta _0\Vert _{H^s}, {\mathfrak {a}}^{-1}). \end{aligned}$$

Choosing \(s'\in (\max \{\frac{d}{2}, s-2\}, s-1)\), then interpolating the above estimates gives

$$\begin{aligned} \Vert K\Vert _{L^\infty ([0, T_*]; L^\infty ({\mathbb {R}}^d)}\le {\mathcal {F}}(\Vert \eta _0\Vert _{H^s}, {\mathfrak {a}}^{-1})\Vert \eta _n-\eta \Vert _{L^\infty ([0, T]; H^{s-1})}^\theta \end{aligned}$$

for some \(\theta \in (0, 1)\). Then, (5.39) follows from this and (5.5). \(\quad \square \)

Now in view of the properties (5.34), (5.36) and (5.39) of \(\eta \), we see that in the proof of (5.19), if we replace \(\eta ^{({\mathfrak {s}}_1)}\) with \(\eta _n\), \(\eta ^{({\mathfrak {s}}_2)}\) with \(\eta \), and \(({\mathfrak {s}}_1, {\mathfrak {s}}_2)\) with \(({\mathfrak {s}}_n, 0)\), then we obtain the convergence estimate

$$\begin{aligned} \Vert \eta _n-\eta \Vert _{L^\infty ([0,T_*];H^{s-1})}+\Vert \eta _n- \eta \Vert _{L^2([0,T_*];H^{s-\frac{1}{2}})} \le \sqrt{{\mathfrak {s}}_n}{\mathcal {F}}(\Vert \eta _0\Vert _{H^s}, {\mathfrak {a}}^{-1}). \end{aligned}$$
(5.40)

Furthermore, assume that \(\eta _1\) and \(\eta _2\) are two solutions on \([0, T_*]\) of the Muskat problem without surface tension with the same initial data \(\eta _0\) and that both satisfy (5.34), (5.36) and (5.39). Then the proof of (5.19) with \({\mathfrak {s}}_1={\mathfrak {s}}_2=0\) yields that \(\eta _1\equiv \eta _2\) on \([0, T_*]\). This proves the uniqueness of \(\eta \). In other words, we have obtained an alternative proof for the local well-posedness of the Muskat problem without surface tension for any subcritical data satisfying (5.1) and (5.2).

The next proposition improves the rate in (5.40) to the optimal rate.

Proposition 5.5

If in addition \(s\ge 2\), then

$$\begin{aligned} \begin{aligned}&\Vert \eta _n-\eta \Vert _{L^\infty ([0,T_*];H^{s-2})}+ \Vert \eta _n-\eta \Vert _{L^2([0,T_*];H^{s-\frac{3}{2}})} \le {\mathfrak {s}}{\mathcal {F}}(\Vert \eta _0\Vert _{H^s}, {\mathfrak {a}}^{-1}), \end{aligned} \end{aligned}$$
(5.41)

where \({\mathcal {F}}:{\mathbb {R}}^+\times {\mathbb {R}}^+\rightarrow {\mathbb {R}}^+\) depends only on \((h, s, \mu ^\pm , {\mathfrak {g}})\).

Proof

We follow the notation in the proof of Proposition 5.2 but set \(\eta _1=\eta _n\) and \(\eta _2=\eta \). Then, \(\eta _\delta =\eta _n-\eta \) satisfies

$$\begin{aligned} (\mu ^+ + \mu ^-) \partial _t \eta _\delta = -{\mathfrak {g}} (L_\delta \eta _1+ L_2\eta _\delta )- {\mathfrak {s}}_n L_1 H(\eta _1). \end{aligned}$$
(5.42)

Applying (2.20) and (2.31) with \(\sigma =s-\frac{3}{2}\ge \frac{1}{2}\) yields

$$\begin{aligned} \Vert L_1 H(\eta _1)\Vert _{H^{s-\frac{5}{2}}}\le {\mathcal {F}}(\Vert \eta _1\Vert _{H^s})\Vert H(\eta _1)\Vert _{H^{s-\frac{3}{2}}}\le {\mathcal {F}}(\Vert \eta _1\Vert _{H^s})\Vert \eta _1\Vert _{H^{s+\frac{1}{2}}}. \end{aligned}$$
(5.43)

Next we paralinearize \(L_2\) and \(L_\delta \). For \(L_2\) we apply (2.26) with \(\sigma =s-\frac{3}{2}\ge \frac{1}{2}\)

$$\begin{aligned} L_2 \eta _\delta = T_{\lambda _2}\eta _\delta + O_{H^{s-\frac{3}{2}} \rightarrow H^{s -\frac{5}{2}+\delta }}({\mathcal {F}}(N_s))\eta _\delta . \end{aligned}$$
(5.44)

\(L_\delta \) can be written as in (5.23). Using (2.25) with \(\sigma = s-\frac{3}{2}\ge \frac{1}{2}\) together with the fact that \(J _\delta ^--J^+_\delta =0\), we obtain

$$\begin{aligned}&\sum _{\pm = +,-} G_2^\pm J _\delta ^\pm = \sum _{\pm = +,-} O_{\widetilde{H}^{s-\frac{3}{2}}_\pm \rightarrow H^{s-\frac{5}{2}+\delta }}({\mathcal {F}}(N_s)) J _\delta ^\pm . \end{aligned}$$

Applying Proposition 4.1 with \(\sigma =s-\frac{3}{2}\ge \frac{1}{2}\), we obain

$$\begin{aligned} J^\pm _\delta =O_{H^s\rightarrow \widetilde{H}^{s-\frac{3}{2}}_\pm }(\Vert \eta _\delta \Vert _{H^{s-\frac{3}{2}}}), \end{aligned}$$

and hence

$$\begin{aligned} \sum _{\pm = +,-} G_2^\pm J _\delta ^\pm = O_{ H^s\rightarrow H^{s-\frac{5}{2}+\delta }}({\mathcal {F}}(N_s)\Vert \eta _\delta \Vert _{H^{s-\frac{3}{2}}}). \end{aligned}$$

On the other hand, Theorem 2.8 can be applied with \(\sigma =s-\frac{3}{2}\ge \frac{1}{2}\), implying

$$\begin{aligned} \sum _{\pm = +,-} G_\delta ^\pm J _1^\pm f=-T_{ \lambda _1 [\![{\mathfrak {B}}_1 J _1]\!]f}\eta _\delta -T_{[\![{\mathfrak {V}}_1 J _1]\!]f} \cdot \nabla \eta _\delta + O_{H^s \rightarrow H^{s-\frac{5}{2}+\delta }}\big ({\mathcal {F}}(N_s) \Vert \eta _\delta \Vert _{H^{s-\frac{3}{2}}}\big )f. \end{aligned}$$

We thus obtain

$$\begin{aligned} \begin{aligned} L_\delta f&=- T_{ \lambda _1 [\![{\mathfrak {B}}_1 J _1]\!]f}\eta _\delta -T_{[\![{\mathfrak {V}}_1 J _1]\!]f} \cdot \nabla \eta _\delta +O_{ H^s\rightarrow H^{s-\frac{5}{2}+\delta }}({\mathcal {F}}(N_s)\Vert \eta _\delta \Vert _{H^{s-\frac{3}{2}}})f. \end{aligned} \end{aligned}$$
(5.45)

Applying this with \(f=\eta _1\), then combining with (5.44) and symmetrizing we arrive at

$$\begin{aligned} L_2\eta _\delta +L_\delta \eta _1&= T_{ (\lambda (1-[\![{\mathfrak {B}} J ]\!]\eta ))_\alpha }\eta _\delta -T_{([\![{\mathfrak {V}} J]\!]\eta )_\alpha } \cdot \nabla \eta _\delta \\&\qquad + O_{H^{s-\frac{3}{2}} \rightarrow H^{s -\frac{5}{2}+\delta }}({\mathcal {F}}(N_s))\eta _\delta +O_{ H^s\rightarrow H^{s-\frac{5}{2}+\delta }}({\mathcal {F}}(N_s)\Vert \eta _\delta \Vert _{H^{s-\frac{3}{2}}})\eta _1. \end{aligned}$$

Plugging this and (5.43) into (5.20) leads to

$$\begin{aligned} \begin{aligned}&(\mu ^+ + \mu ^-) \partial _t \eta _\delta =- {\mathfrak {g}}T_{ (\lambda (1-[\![{\mathfrak {B}} J ]\!]\eta ))_\alpha }\eta _\delta +{\mathfrak {g}}T_{([\![{\mathfrak {V}} J]\!]\eta )_\alpha } \cdot \nabla \eta _\delta +{\mathcal {R}}'_1+{\mathcal {R}}_2',\\&\Vert {\mathcal {R}}_1'\Vert _{H^{s-\frac{5}{2}+\delta }}\le {\mathfrak {g}}{\mathcal {F}}(N_s)\Vert \eta _\delta \Vert _{H^{s-\frac{3}{2}}},\\&\Vert {\mathcal {R}}_2'\Vert _{H^{s-\frac{5}{2}}}\le {\mathfrak {s}}_n{\mathcal {F}}(N_s) \Vert \eta _1\Vert _{H^{s+\frac{1}{2}}}. \end{aligned} \end{aligned}$$
(5.46)

Next we set \(\eta _{\delta , s-2}=\langle D\rangle ^{s-2}\eta _\delta \) and commute the first equation in (5.46) with \(\langle D\rangle ^{s-2}\) to obtain after applying Theorem A.3 that

$$\begin{aligned} \begin{aligned}&(\mu ^+ + \mu ^-) \partial _t \eta _{\delta , s-2} =- {\mathfrak {g}}T_{ (\lambda (1-[\![{\mathfrak {B}} J ]\!]\eta ))_\alpha }\eta _{\delta , s-2} +{\mathfrak {g}}i\mathrm {Re}\big (T_{([\![{\mathfrak {V}} J]\!]\eta )_\alpha \cdot \xi }\big ) \eta _{\delta , s-2}+{\mathcal {R}}_1+{\mathcal {R}}_2,\\&\Vert {\mathcal {R}}_1\Vert _{H^{-\frac{1}{2}+\delta }}\le {\mathfrak {g}}{\mathcal {F}}(N_s)\Vert \eta _{\delta , s-2}\Vert _{H^{\frac{1}{2}}},\\&\Vert {\mathcal {R}}_2\Vert _{H^{-\frac{1}{2}}}\le {\mathfrak {s}}_n{\mathcal {F}}(N_s) \Vert \eta _1\Vert _{H^{s+\frac{1}{2}}}, \end{aligned} \end{aligned}$$
(5.47)

where \({\mathcal {F}}\) depends only on \((h, s, \mu ^\pm )\). An \(L^2\) energy estimate as in (5.29) yields

$$\begin{aligned} \begin{aligned}&\frac{(\mu ^+ + \mu ^-)}{2}\frac{d}{dt}\Vert \eta _{\delta , s-2} \Vert _{L^2}^2 +\frac{{\mathfrak {g}}}{{\mathcal {Q}}_{T_*}}\Vert \eta _{\delta , s-2} \Vert ^2_{H^\frac{1}{2}}\\&\quad \le {\mathfrak {g}}{\mathcal {Q}}_{T_*}\Vert \eta _{\delta , s-2}\Vert _{H^{\frac{1}{2}-\delta }}\Vert \eta _{\delta , s-2}\Vert _{H^{\frac{1}{2}}}+{\mathfrak {s}}_n{\mathcal {Q}}_{T_*} \Vert \eta _1\Vert _{H^{s+\frac{1}{2}}}\Vert \eta _{\delta , s-2}\Vert _{H^{\frac{1}{2}}}, \end{aligned} \end{aligned}$$
(5.48)

where \({\mathcal {Q}}_{T_*}\) is given by (5.30). Interpolating as in (5.31) we obtain

$$\begin{aligned} \begin{aligned}&\frac{(\mu ^+ + \mu ^-)}{2}\frac{d}{dt}\Vert \eta _{\delta , s-2} \Vert _{L^2}^2 +\frac{{\mathfrak {g}}}{{\mathcal {Q}}_{T_*}}\Vert \eta _{\delta , s-2}\Vert _{H^{\frac{1}{2}}}^2\le {\mathfrak {g}}{\mathcal {Q}}_{T_1}\Vert \eta _{\delta , s-2}\Vert ^2_{L^2}+{\mathfrak {s}}_n^2{\mathcal {Q}}_{T_*} \Vert \eta _1\Vert _{H^{s+\frac{1}{2}}}^2. \end{aligned} \end{aligned}$$
(5.49)

From the uniform estimate (5.4) we have

$$\begin{aligned} \int _0^{T_*}\Vert \eta _1\Vert ^2_{H^{s+\frac{1}{2}}}dt \le {\mathcal {F}}(\Vert \eta _0\Vert _{H^s}, {\mathfrak {a}}^{-1}). \end{aligned}$$

Thus, applying Gönwall’s lemma to (5.49) we arrive at (5.41). \(\quad \square \)