1 Introduction

We consider the irrotational incompressible capillary-gravity water waves problem in a two-dimensional domain \(\Omega _t\), where \(\Omega _t\) is a bounded domain with an upper free surface \(\Gamma _t\) and a fixed bottom \(\Gamma _b\). This moving domain contains two moving contact points, \(p_{l}, p_{r}\) (left and right), with the contact angles \(\omega _l, \omega _r\in (0, \pi /2)\), which are the intersection points of \(\Gamma _t, \Gamma _b\):

$$\begin{aligned} \Gamma _t\cap \Gamma _b=\{p_l, p_r\}. \end{aligned}$$

Moreover, the fixed bottom \(\Gamma _b\) is assumed to be smooth enough, and it becomes straight near the contact points \(p_i\) (\(i=l, r\)) for the sake of simplicity.

figure a

The water waves problem has been widely studied in centuries; see, for example, [49, 63]. This problem focuses on the motion of an ideal fluid and describes the evolution of the free surface \(\Gamma _t\) as well as the velocity field v. Mathematically, it is described by Euler’s equation with boundary conditions and initial conditions, and in our case, we also need some boundary conditions at contact points.

We express the water waves problem on the corner domain \(\Omega _t\) as the following system \(\text{(WW) }\) of velocity v and pressure P:

figure b

Here, (1.1) from \(\text{(WW) }\) is Euler’s equation where, \(\textbf{g}=-g \mathbf{e_z}\) is the vertical gravity vector; (1.2) describes the incompressibility and irrotationality; (1.3) is the condition of the pressure on the free surface in the case with surface tension, where \(\sigma \) is the coefficient of surface tension and \(\kappa \) is the mean curvature of \(\Gamma _t\) (see Section 1.2); (1.4) is the classical kinematic condition on the free surface \(\Gamma _t\) with \(D_t\) the material derivative, Meanwhile, (1.5) describes that the velocity along the fixed bottom \(\Gamma _b\) is always tangential, where \(n_b\) is the unit outward normal vector of \(\Gamma _b\). These equations and conditions are standard in water waves, see [45, 46].

In particular, (1.6) gives the conditions at contact points, which come from [58]. We denote by \(v_i\) the upward tangential components of the velocity at the corner points along \(\Gamma _b\):

$$\begin{aligned} v_l=-v\cdot \tau _b\quad \hbox {at}\ p_{l},\quad \hbox {and}\quad v_r=v\cdot \tau _b\quad \hbox {at}\ p_{r}. \end{aligned}$$

Here \(\omega _s\) is the stationary contact angle decided by the materials of the bottom and the fluid (see [76]), and \(\beta _c\) denotes the effective friction coefficient. This condition shows that the slip velocity is dominated by the unbalanced Young stress, and it is indeed an effective variation of Young’s law (1805) for stationary contact angles [76]. In fact, this kind of conditions are commonly seen, see [11, 15, 27, 62].

Before presenting our results, we recall briefly earlier works on the well-posedness of classical water waves problem, where one has smooth surfaces \(\Gamma _t\) satisfying \(\Gamma _t\cap \Gamma _b=\emptyset \).

We recall results on the local well-posedness. When the fluid is irrotational, some early works like Nalimov [55], Yosihara [74, 75] and Craig [22] established the two-dimensional local well-posedness with small initial data in Sobolev spaces. In the late 1990s, Wu [68, 69] proved for the first time the local well-posedness with general initial data in Sobolev spaces and showed that the Taylor sign condition

$$\begin{aligned} -\nabla _{n_t} P\geqq c_0>0 \end{aligned}$$
(1.7)

held on \(\Gamma _t\) as long as \(\Gamma _t\) was not self-intersecting. Iguchi, Tanaka and Tani [38] and Iguchi [36] proved the local well-posedness in two-dimensional case respectively. Later on, Lannes [45] derived the local well-posedness of the gravity water waves under Zakharov formulation, which is convenient to link with approximate models. Alazard, Burq and Zuily in [2,3,4,5] used paradifferential operators and Strichartz estimates to study the problem in a low-regularity space. On the other hand, when the fluid is rotational, Christodoulou and Lindblad [18] proved a priori estimates based on the geometry of the moving domain. Lindblad [50] obtained the existence of solutions using Nash-Moser iteration. In 2007, Coutand and Shkoller [20] used Lagrangian coordinates to show the local well-posedness. Shatah and Zeng [60, 61] adopted a geometric point of view to reformulate the problem and prove the local well-posedness, while Beyer and Günther [12, 13] used a similar geometric approach to study the irrotational flow. Zhang and Zhang [77] proved the local well-posedness for rotational flow using a framework of Clifford analysis introduced by Wu [69]. For more references see [7,8,9, 46, 54, 56, 57, 59, 67, 77] etc..

For the global well-posedness of small data, Wu [72] and Germain, Masmoudi and Shatah [28] proved the global three-dimensional existence of gravity water waves respectively using different approaches. One can check [6, 25, 33,34,35, 39, 66, 71, 73] e.t.c. and their references for more results on gravity or capillary-gravity water waves,. Meanwhile, there are also some works concerning geometric singularities on the free surfaces. The authors in [16] proved the existence of a wave which is given initially as the graph of a function and then can overturn at a later time. Later on, the authors in [17] showed the existence for some “splash" or “splat” singularities. This result was extended to three-dimensional case and some other models in [21].

Compared to the rich literature on the well-posedness of classical water waves, the research on the well-posedness of water waves problem with non-smooth boundaries (we call it “non-smooth water waves”) just started several years ago and there are a lot of open questions. In general, there are two kinds of non-smooth water waves problems: The first kind of problem has contact points (or contact lines) between the free surface and the bottom, i.e. \(\Gamma _t\cap \Gamma _b\ne \emptyset \); The other kind contains crests or cusps on the free surface, i.e. the surface is Lipschitz. We would like to mention that in the case with large crest angles, the famous Stokes waves can be dated back to papers by G. Stokes [63, 64] which obtained traveling-wave solutions with limit crest angle \(2\pi /3\). Obviously, the main difference here compared to classical water waves lies in the corners on boundaries. As a result, the analysis involving the corners (i.e. domain singularities) becomes the key point in the non-smooth water waves.

Now, we are in a position to state an informal version of the main result. To begin with, we introduce the following compatibility conditions at \(t=0\):

$$\begin{aligned} \beta _c D_t^k v_i(0)=\sigma [D_t^k(\cos {\omega _s}-\cos {\omega _{i}})](0)\quad \hbox {at}\quad p_i \,(i=l, r), \quad k=0,1,2.\nonumber \\ \end{aligned}$$
(1.8)

When we reduce system \(\text{( }WW)\) into some initial-boundary value problem essentially related to the mean curvature and the pressure (see (3.15), (5.30) and (6.6)), these compatibility conditions are needed in the linearized system (Section 6.3) in a natural way as long as we try to solve this problem.

Theorem 1.1

Let the initial data belong to a suitable space and the initial contact angles \(\omega _{i0}\in (0, \pi /2)\) for \(i=l, r\). If the compatibility conditions (1.8) are satisfied, there exists a time interval depending on the initial data such that system \(\text{(WW) }\) is locally well-posed in a suitable space.

Remark 1.1

The suitable space is defined by \(\Sigma _h\) in Section 6.3 for some good unknowns built upon the mean curvature, the pressure and the position information of contact points. In fact, the proof of Theorem 1.1 is divided into three parts. In Section 4 and Section 5, we obtain a priori estimates (see Theorem 4.1 and Theorem 5.1). In the last section, we construct approximate solutions to prove the well-posedness based on the a priori estimates. The precise statement of local well-posedness for a geometric form of \(\text{( }WW)\) is given in Theorem 6.1 (Section 6.3).

Remark 1.2

When there is no surface tension, the authors of [19, 40] studied the case with contact angles. They assume that the wall \(\Gamma _b\) is vertical and the contact angle \(\omega \in (0, \pi /4)\). Then, by a symmetric extension, they reduced the problem to the crest case with the crest angle less than \(\pi /2\). As a result, [19, 40] do not contain the case of limit Stokes waves with the “ \(2\pi /3\)” crest. In that case, one needs a contact angle of \(\pi /3\) even with the help of a symmetric extension. In this paper, we only require that the contact angles \(\omega _i\in (0, \pi /2)\), which brings us some useful experiences to deal with domain singularities in a more general case for water waves.

Remark 1.3

Noticing that

$$\begin{aligned} \cos \omega _i=-\tau _t\cdot \tau _b |_{p_i}, \quad i=l, r \end{aligned}$$

we can simply have

$$\begin{aligned} (\sin \omega _i) D_t\omega _i=F(\partial v, \partial \Gamma _t)|_{p_i} \end{aligned}$$

with some polynomial F. Therefore, the time evolution of contact angles \(\omega _i=\omega _i(t)\) depends on the derivatives of v and the free surface \(\Gamma _t\). A continuity argument in time shows that as long as the initial contact angles \(\omega _{i0}\in (0, \pi /2)\), we have \(\omega _i(t)\in (0, \pi /2)\) in a short time interval.

Remark 1.4

The compatibility conditions (1.8) seem to be complicated, but they can be satisfied according to the proof of our main theorem. In fact, when one considers the initial data \((v_0, \Gamma _{t0})\) for system \(\text{( }WW)\), it suffices to find the corresponding initial data for essentially the mean curvature \(\kappa \) (and its time derivative), the pressure P and the positions of contact points (see the initial data in Section 6.3). According to the linear system (6.7) in the iteration, compatibility conditions are required there which are linear, localized and derived from (1.8). Therefore, one can find corresponding initial values for this linear system satisfying these compatibility conditions. As a result, based on the proof of the existence of the solution to system (6.6) in Theorem 6.1, we can show that conditions (1.8) are satisfied thanks to the linear system (6.7).

We recall some works concerning the local well-posedness of non-smooth water waves. In the case where there are crests or cusps on the free surface (the “crest” case), Alazard, Burq and Zuily [5] study a special case (without surface tension) when the contact angle is equal to \(\pi /2\) (the right angle), where they used symmetric and periodic extension to turn this problem into a classical smooth periodic case. A breakthrough in this subject is made by Kinsey and Wu, see [40, 70]. They focus on gravity water waves where the crest angle is less than \(\pi /2\). The main difficulty is that the free surface is a non-\(C^1\) interface with angled crest and the Taylor sign \(-\nabla _{n_t}P\) degenerates at the crest point. To be more precise, they start with reducing the water waves problem into the following equation in [40, 70]:

$$\begin{aligned} \partial _t^2 u+{\mathfrak {a}}\nabla _{n} u= f(u, \partial _t u). \end{aligned}$$

there \({\mathfrak {a}}=-\nabla _{n_t}P\) is the Taylor sign. When \({\mathfrak {a}}\) degenerates at the crest point, the above system will loss its hyperbolicity, and classical analysis does not apply any more. To solve the problem, Kinsey and Wu flatten the domain with a Riemann mapping, and some singular weights appear naturally in their equation. As a result, they introduce some weighted Sobolev spaces accordingly for energy estimates to deal with these singular weights. Based on these works, Agrawal [1] show that these singularities are “rigid", which means that the angle of these crests can not change in time. Very recently, Córdoba, Enciso and Grubic [19] study a similar case with cusps and crests without gravity, where the angles of these crests are less than \(\pi /2\) and change in time.

For the other case where there are contact angles (that is \(\Gamma _t\cap \Gamma _b\ne \emptyset \)), things become different from the “crest” case. First, the corners appear due to intersections of the free surface and the bottom (or wall). Second, there are different boundary conditions in corresponding elliptic systems compared to the crest case. In fact, various boundary conditions may have no big difference if we only focus on the elliptic theory, but there will be a series of consequences in water waves when boundary conditions change. For example, the evolution of the free surface is different from the crest case. Moreover, when there is surface tension, boundary conditions at the contact points as (1.6) are needed in order to close the system, and dissipations appear at the contact points too (See Section 4 and [52]).

In the case with contact angles, de Poyferré [23] prove a priori estimates in bounded n-dimensional corner domains without surface tension. The contact angle is assumed to be small to ensure sufficient Sobolev regularity near the corner. Meanwhile, under a similar assumption of small contact angles, we obtain the local well-posedness in a two-dimensional corner domain (beach type) with surface tension, see [52, 53]. Meanwhile, we notice that [23] and [52, 53] use similar geometry formulations introduced in [60].

To explain why small contact angles are needed in [23, 52, 53], we look at a typical mixed-boundary system in water waves:

$$\begin{aligned} {\left\{ \begin{array}{ll}{} \Delta u=0\qquad in \quad \Omega _t,\\ u|_{\Gamma _t}=f,\quad \nabla _{n_b}u|_{\Gamma _b}=0. \end{array}\right. } \end{aligned}$$

In fact, the elliptic theory on corner domains is well known already, see for example [30, 43, 44]. Generally, one still has variational solution \(u\in H^1(\Omega _t)\) if the right-side functions lie in proper Sobolev spaces. If one seeks for \(H^2\) and above regularities, singularity decompositions are needed naturally for the solution, which decompose the solution into a singular part \(u_s\) (i.e. not good enough) near the corners and a regular part \(u_r\):

$$\begin{aligned} u=u_s+u_r\qquad \hbox {according to the required regularity.} \end{aligned}$$

For the mixed-boundary system above, the most singular part in \(u_s\) is like \(r^{\pi /2\omega }\), where r is the radius with respect to the corner point and \(\omega \) is the contact angle. Consequently, when \(\omega \) is small enough, the singular part \(u_s\) will be good enough so that we will have enough regularities from elliptic systems as in classical water waves to close the energy estimates. One can also find a singularity decomposition for v in Proposition 2.3 (Section 2).

In contrast, when the contact angle is larger or more general, the idea of taking small angles to improve regularities in [23, 52, 53] does not work any more. Meanwhile, there is no obvious weighted space to use due to the structure of the water waves problem.

We want to show here the main ingredients of this paper. Similarly as in our previous works [52, 53], we still adopt the geometric formulation from Shatah and Zeng [60, 61]. In fact, we rewrite \((\hbox {WW})\) into an equation for \({\mathfrak {J}}=\nabla {\mathfrak {K}}_{{\mathcal {H}}}\) with boundary conditions, where \({\mathfrak {K}}\) is the modified mean curvature on \(\Gamma _t\) and \({{\mathcal {H}}}\) means the harmonic extension in \(\Omega _t\) (see Section 3, and we only present a simpler form for this equation here),

$$\begin{aligned} D^2_t{\mathfrak {J}}+\sigma {{\mathcal {A}}}{\mathfrak {J}}=R\quad \hbox {with}\quad {{\mathcal {A}}}{\mathfrak {J}}=\nabla {{\mathcal {H}}}(-\Delta _{\Gamma _t}{\mathfrak {J}}^\perp ), \end{aligned}$$
(1.9)

where R is some remainder part.

The trouble here is that singularities from the domain \(\Omega _t\) affect directly the regularity of the solution to \(\text{(WW) }\), which means singularities always appear in related elliptic systems even if the boundary conditions are good enough in Sobolev spaces (see for example [30]). More precisely, the natural norm \(\Vert {{\mathcal {A}}}^{k}f \Vert _{L^2(\Omega _t)}+\Vert f\Vert _{L^2(\Omega _t)}\) with \(k\geqq 1\) arising from this equation above is not equivalent with \(\Vert f \Vert _{H^{3k}(\Omega _t)}\) due to larger contact angles, and apparently “singular parts" are contained in this norm.

Due to this kind of singularities from elliptic systems, we only have limited regularities for some quantities (for example, the velocity v) in Sobolev spaces. Compared to [52, 53], we must make full use of the maximal regularity for each quantity (especially for v) while the contact angles \(\omega _i\in (0, \pi /2)\). To do this, some delicate estimates together with singular parts from singularity decompositions (see [30, 51]) are carefully used. Meanwhile, it is also very important to gain more information from the structure of \(\text{(WW) }\).

The main part of the (lower-order) energy functional is defined as

$$\begin{aligned} E(t)=\Vert D_t {\mathfrak {J}}\Vert ^2_{L^2(\Omega _t)}+\Vert {\mathfrak {J}}\cdot n_t\Vert ^2_{H^1(\Gamma _t)}, \end{aligned}$$

which gives us the estimate (see Section 4.1)

$$\begin{aligned}{} & {} \Vert v\Vert _{H^2(\Omega _t)}+\Vert D_t v\Vert _{H^{3/2}(\Omega _t)}+\Vert D_t^2 v\Vert _{L^2(\Omega _t)}+\Vert \nabla v\Vert _{L^\infty (\Omega _t)}+\Vert \kappa \Vert _{H^2(\Gamma _t)}\\{} & {} \quad \leqq P(E(t)), \end{aligned}$$

with P(E(t)) the positive-coefficient polynomial of E(t). This means that the free surface \(\Gamma _t\) still has enough regularity and v is Lipschitz.

In our previous work [52, 53], one main part from the remainder term R in (1.9) is about the higher-order derivative terms of \(P_{v,v}\), where \(P_{v, v}\) is defined by an elliptic system with mixed-boundary conditions. According to the elliptic theory on corner domains (see for example [30]), the system of \(P_{v,v}\) only gives \(P_{v,v}\) limited regularity around \(H^2(\Omega _t)\), when the contact angles are less than \(\pi /2\). To improve its regularity, we modify the definition of \(P_{v, v}\) in this paper to have a Neumann-boundary system, see (3.1). In fact, thanks to the elliptic theory, solutions of Neumann-boundary system (or Dirichlet system) may become more regular than solutions of mixed-boundary problem in Sobolev spaces, while the right-side data have the same regularity. As a result, we obtain a bit more regularity from the elliptic system of \(P_{v,v}\) (we have \(P_{v,v}\in H^3(\Omega _t)\) indeed), which is important in the energy estimates.

Moreover, another main part of R lies in the higher-order derivative terms of the velocity v. Here, we point out that compared to [52, 53], v loses some regularity due to the existence of corners. Fortunately, \(D_t v\) and \(D_t^2 v\) have the same regularities as before. As a result, when we deal with v, sometimes we need to use material derivatives \(D_t\) instead of spatial derivatives. Meanwhile, we also need to apply singularity decompositions to v and its potential \(\phi \) in the estimates, see for example Lemma 4.5.

For the higher-order energy estimates, we use the material derivative \(D_t\) instead of \({{\mathcal {A}}}^{1/2}\) from [60]. The higher-order energy is defined as

$$\begin{aligned} E_1(t)=\Vert \nabla _{\tau _t} D_t (\mathfrak {J}\cdot n_t) \Vert ^2_{L^2(\Gamma _t)} +\Vert D^2_t \mathfrak {J}\Vert ^2_{L^2(\Omega _t)}. \end{aligned}$$

In fact, one will see that \(D_t\) is convenient to use when there are contact points. For example, taking \(D_t\) on elliptic systems does not violate boundary conditions, while it will change boundary conditions if one takes spatial derivatives. Moreover, one will find in the higher-order energy estimate that taking \(D_t\) leads to better regularities for some quantities than their own regularities (like \(D_t v\) for the velocity v). However, due to the singularities of the boundary (contact points), it is not as convenient as before to turn these regularities into spatial regularities, which explains the reason why we need to choose very carefully for “a suitable space” in Theorem 1.1. As long as the energy estimates for \(E(t), E_1(t)\) are finished, we use the equation of \({\mathfrak {J}}\) to gain more spatial regularities.

we now mention some other related works. Lannes and Métivier [48] studied the Green-Naghdi system in a beach-type domain, which is a shallow-water model of the water waves problem. Lannes [47] studied the floating-body problem and proposed a new formulation that can be easily generalized in order to take into account the presence of a floating body. Lannes and Iguchi [37] proved some sharp results for initial boundary value problem with a free boundary arising in wave-structure interaction, and it contains the floating problem in the shallow water regime. In addition, Guo and Tice [31] showed a priori estimates for the contact line problem in the case of the stokes equations. Later on, Tice and Zheng proved the local well-posedness of the contact line problem in two-dimensional Stokes flow, see [65]. In 2020, Guo and Tice [32] proved a priori estimates for the contact line problem for two-dimensional Navire-Stokes flow. For Darcy’s flow, one can see [41, 42].

1.1 Organization of the Paper

In Section 2, we present various useful lemmas including singularity decompositions and estimates for elliptic systems. In Section 3, we derive the equation for the good unknown \({\mathfrak {J}}\) from \((\hbox {WW})\) with modified curvature \({\mathfrak {K}}\) and modified pressure \(P_{v,v}\). The lower-order energy is constructed and the energy estimate is proved in Section 4, where estimates for various quantities like \(\Gamma _t\), v, \(P_{v,v}\) are proved. Moreover, we consider the higher-order energy estimate using \(D_t\) in Section 5. In Section 6, we present the precise main theorem in our paper and show the local well-posedness.

1.2 Notations

- X stands for a point in . \(p_l, p_r\) are the left and right contact points. \(n_j (j=t, b)\) are the unit outward normal vectors on \(\Gamma _j\), and \(\tau _j\) are the corresponding unit tangential vectors obeying the right-hand rule with \(n_j\).

- \(\sigma \) is the surface tension coefficient. \(\beta _c\) is the effective friction coefficient determined by interfacial widths, interactions between the fluid and the bottom, and the normal stress contributions.

- \(\chi _{\omega }\) is a characteristic function of contact angles:

$$\begin{aligned} \chi _{\omega }(\theta )= {\left\{ \begin{array}{ll} 1,\quad \theta \in (\pi /3, \pi /2),\\ 0,\quad \theta \in (0,\pi /3]. \end{array}\right. } \end{aligned}$$

- \(\chi _i\) (\(i=l, r\)) are cut-off functions near the corner points \(p_i\):

$$\begin{aligned} \chi _{i}(X)= {\left\{ \begin{array}{ll} 1,\quad \hbox {when}\ |X-p_i|\leqq r_0,\ X\in \Omega _t;\\ 0,\quad \ \hbox {otherwise} \end{array}\right. } \end{aligned}$$

with some small \(r_0>0\).

  • \(f|_c=\chi _l (f|_{p_l})+\chi _r (f|_{p_r})\) stands for taking values of f at the corner points.

  • \({{\mathcal {S}}}_{t,i}\) are straightened sector of \(\Omega _t\) with radius \(r_0>0\) near the corner points \(p_i\).

  • \(D_t=\partial _t+\nabla _v\) is the material derivative.

  • \(M^*\) denotes the transport of a matrix M.

  • \(w^\perp \) on \(\Gamma _t\): \(w\cdot n_t\) for a vector \(w\in T_X\Gamma _t\).

  • \(w^\top \) on \(\Gamma _t\): \((w\cdot \tau _t)\,\tau _t\). Sometimes we also use \(w^\top \) on \(\Gamma _b\) with a similar definition.

  • \(\Pi \) is the second fundamental form on \(\Gamma _t\), where \(\Pi (w)=\nabla _w n_t\in T_X\Gamma _t\) for \(w\in T_X\Gamma _t\).

  • \(\kappa =tr \Pi =\nabla _{\tau _t}n_t\cdot \tau _t\) is the mean curvature of the surface \(\Gamma _t\).

  • We define on \(\Gamma _t\) that \({{\mathcal {D}}}w={{\mathcal {D}}}_{\tau _t}w=(\nabla _{\tau _t}w)^\top =(\nabla _{\tau _t}w\cdot \tau _t\big )\tau _t\) for a vector \(w\in T_X\Gamma _t\).

  • \({{\mathcal {D}}}^2f(\tau _1, \tau _2)=D^2f(\tau _1,\tau _2)-\big (\Pi (\tau _1)\cdot \tau _2\big )\nabla _{n_t}f\) for any two vector \(\tau _1, \tau _2\in T_X\Gamma _t\).

  • \(\Delta _{\Gamma _t}\) is the Beltrami-Laplace operator on \(\Gamma _t\):

    $$\begin{aligned} \Delta _{\Gamma _t}f={{\mathcal {D}}}^2f(\tau _t, \tau _t)=\nabla _{\tau _t}\nabla _{\tau _t}f-\nabla _{(\nabla _{\tau _t}\tau _t)^\top }f. \end{aligned}$$

    - \({{\mathcal {H}}}(f)\) or \(f_{{\mathcal {H}}}\) is the harmonic extension for some function f on \(\Gamma _t\), which is defined by the elliptic system

    $$\begin{aligned} \left\{ \begin{array}{ll} \Delta {{\mathcal {H}}}(f)=0\qquad \hbox {in}\quad \Omega _t,\\ {{\mathcal {H}}}(f)|_{\Gamma _t}=f,\quad \nabla _{n_b}{{\mathcal {H}}}(f)|_{\Gamma _b}=0. \end{array}\right. \end{aligned}$$

    - \({{\mathcal {N}}}=\nabla _{n_t}{{\mathcal {H}}}\) is the Dirichlet-Neumann operator on \(\Gamma _t\). - \(\Delta ^{-1}(h,g)\) stands for the solution u to the system

    $$\begin{aligned} \left\{ \begin{array}{ll} \Delta u=h\qquad \hbox {in}\quad \Omega _t\\ u|_{\Gamma _t}=0,\qquad \nabla _{n_b}u|_{\Gamma _b}=g. \end{array} \right. \end{aligned}$$

    - The Sobolev norm \(H^s\) for the boundary \(\Gamma _t\) or \(\Gamma _b\) is defined by local coordinates and local graphs. - \({\tilde{H}}^{1/2}(\Gamma _j)\) (\(j=t,b\)) (see [30]) is a subspace of \(H^{1/2}(\Gamma _j)\) related to corner domains

    $$\begin{aligned} {\tilde{H}}^{1/2}(\Gamma _j)=\Big \{u\in \dot{H}^{1/2}(\Gamma _j)\Big | \,\rho _i^{-1/2}u\in L^2(\Gamma _j),\ i=l, r\Big \}\end{aligned}$$

    where \(\dot{H}^{1/2}(\Gamma _j)\) is the closure of \(\mathscr {D}(\Gamma _j)\) in \(H^{1/2}(\Gamma _j)\), and \(\rho _i=\rho _i(X)\) (\(i=l, r\)) is the distance (arc length) between the point \(X\in \Gamma _j\) and the end \(p_i\). We define the norm

    $$\begin{aligned} \Vert u\Vert ^2_{{\tilde{H}}^{1/2}(\Gamma _j)}=\Vert u\Vert ^2_{H^{1/2}(\Gamma _j)}+\int _{\Gamma _j} \rho _l^{-1}|u|^2{\text {d}}X+\int _{\Gamma _j} \rho _r^{-1}|u|^2{\text {d}}X. \end{aligned}$$

    - \({\tilde{H}}^{-1/2}(\Gamma _j)\) stands for the dual space of \({\tilde{H}}^{1/2}(\Gamma _j)\). For more details, see [30]. - We define \(P_{w,v}\) (with \(w\ne v\)) by the following Neumann system:

    $$\begin{aligned} \left\{ \begin{array}{ll} \Delta P_{w,v}=-tr (\nabla w \nabla v),\qquad \hbox {in}\quad \Omega _t\\ \nabla _{n_t} P_{w,v}|_{\Gamma _t}=C_{w,v}(t)-(w\cdot \tau _t) \nabla _{\tau _t} v\cdot n_t, \qquad \nabla _{n_b} P_{w,v}|_{\Gamma _b}=w\cdot \nabla _v n_b. \end{array}\right. \nonumber \\ \end{aligned}$$
    (1.10)

    Here \(C_{w,v}\) is a function of t satisfying the compatibility condition

    $$\begin{aligned} |\Gamma _t|C_{w, v}(t)= -\int _{\Omega _t}tr (\nabla w \nabla v){\text {d}}X+\int _{\Gamma _t}(w\cdot \tau _t) \nabla _{\tau _t} v\cdot n_t {\text {d}}s-\int _{\Gamma _b}w\cdot \nabla _v n_b {\text {d}}s. \end{aligned}$$

    - \(P_{v,v}\) is defined in (3.1) from Section 3. - C(a) stands for a positive constant C depending on a quantity a. P(E(t)) stands for a polynomial of the energy E(t) with positive constant coefficients.

2 Preliminaries

To get started, we recall some useful estimates for Sobolev space.

Lemma 2.1

(Sobolev Embeddings) One has the inclusion

$$\begin{aligned} W^{s, p}(\Omega )\subseteq L^q( \Omega ) \end{aligned}$$

for \(1/q=1/p-s/n\) and \(\Omega \) any bounded open subset of with a Lipschitz boundary. Moreover, one also has

$$\begin{aligned} W^{s, p}(\Omega )\subseteq W^{t,q}(\bar{\Omega }) \end{aligned}$$

for \(t\leqq s, q\geqq p\) such that \(s-n/p=t-n/q\).

Lemma 2.2

(Product estimates) (1) For functions \(f\in H^{1/2}(\Omega _t)\) and \(g\in H^1(\Omega _t)\cap L^\infty (\Omega _t)\), one has the product estimate

$$\begin{aligned} \Vert f\,g\Vert _{H^{1/2}(\Omega _t)}\leqq C\Vert f\Vert _{H^{1/2}(\Omega _t)}\big (\Vert g\Vert _{H^1(\Omega _t)}+\Vert g\Vert _{L^\infty (\Omega _t)}\big ), \end{aligned}$$

with a constant C independent of fg;

(2) For functions \(f, g\in L^\infty (\Gamma _t)\cap H^{1/2}(\Gamma _t)\), one has

$$\begin{aligned} \Vert f\,g\Vert _{H^{1/2}(\Gamma _t)}\leqq C \big (\Vert f\Vert _{L^\infty (\Gamma _t)}\Vert g\Vert _{H^{1/2}(\Gamma _t)}+\Vert f\Vert _{H^{1/2}(\Gamma _t)}\Vert g\Vert _{L^\infty (\Gamma _t)}\big ) \end{aligned}$$

with a constant C independent of fg.

Proof

(1) In fact, one firstly extends fg to be defined on the full plane with a control of their corresponding norms. Secondly, one can apply standard para-product analysis to prove the estimate on . The details are omitted here. (2) The proof is similar to the proof of (1). \(\quad \square \)

We quote some Hardy inequalities here.

Lemma 2.3

(Hardy inequalities) (1)([14, Corollary 2.3]) Let \(f\in H^1(0, d)\cap C^0(0, d)\) with \(d>0\) and \(f(0)=0\). Then there exists a positive constant \(C=C(\epsilon , d)\) such that

$$\begin{aligned} \int ^d_0 r^{-2\varepsilon }|f(r)|^2dr\leqq C\int ^d_0r^{-2\varepsilon +2}|f'(r)|^2dr \end{aligned}$$

for \(\epsilon \in (1/2, 1)\);

(2) (Fractional-order version, see [26, 29]) For a number \(\varepsilon \in (0, 1)\) and any function with \(f(0)=0\), there exists a positive constant \(C=C(\varepsilon )\) such that

Lemma 2.4

(Traces on \(\Gamma _t\) or \(\Gamma _b\), [52, Theorem 5.3]) The maps

$$\begin{aligned} u\mapsto \{u, \nabla _{n_j} u\}|_{\Gamma _j},\quad \hbox {for}\quad j=t, b \end{aligned}$$

have unique continuous extensions as operators from \(H^{s}(\Omega _t)\) onto \(\Pi ^1_{i=1} H^{s-i-1/2}(\Gamma _j)\) for \(s>3/2\).

Moreover, one has the estimate:

$$\begin{aligned} \Vert u\Vert _{H^{s-1/2}(\Gamma _j)}+\Vert \nabla _{n_j} u\Vert _{H^{s-3/2 }(\Gamma _j )}\leqq C\big (\Vert \Gamma _t\Vert _{H^{s-1/2} }\big )\Vert u\Vert _{H^{s}(\Omega _t)}. \end{aligned}$$

Next, we present some some special trace theorems on corner domains involving \({\tilde{H}}^{1/2}(\Gamma _j)\) and \({\tilde{H}}^{-1/2}(\Gamma _j)\) (\(j=t, b\)).

Lemma 2.5

Assume that \(u|_{\Gamma _t}=f\) with \(f|_{p_i}=0\) (\(i=l,r\)) for a function \(u\in H^1(\Omega _t)\). Then one has \(f\in {\tilde{H}}^{1/2}(\Gamma _t)\) and

$$\begin{aligned} \Vert f\Vert _{{\tilde{H}}^{1/2}(\Gamma _t)}\leqq C(\Vert \Gamma _t\Vert _{H^{5/2} }) \Vert u\Vert _{H^1(\Omega _t)}. \end{aligned}$$

The case on \(\Gamma _b\) holds similarly.

Proof

First, one has \(f\in H^{1/2}(\Gamma _t)\) immediately by Lemma 2.4. Second, noticing \(f|_{p_i}=0\) and applying Lemma 2.3(2) with some straightening localizations near \(p_i\), one can see that \(f\in {\tilde{H}}^{1/2}(\Gamma _t)\) with the desired estimate. Moreover, a lemma similar to Lemma 5.5 [52] can be follow \(\quad \square \)

Lemma 2.6

[52, Lemma 5.6] Let \(u\in H^{1/2}(\Gamma _j)\) (\(j=t, b\)), then \(\nabla _{\tau _j} u\) belongs to \({\tilde{H}}^{-1/2}(\Gamma _j)\) and satisfies the estimate

$$\begin{aligned} \Vert \nabla _{\tau _b}u\Vert _{{\tilde{H}}^{-1/2}(\Gamma _j)}\leqq C(\Vert \Gamma _j\Vert _{H^{5/2} } ) \Vert u\Vert _{H^{1/2}(\Gamma _j)}. \end{aligned}$$

We now recall some elliptic systems and estimates including singular decompositions in corner domains. First, for the mixed-boundary system

$$\begin{aligned} \text{(MBVP) }\quad \left\{ \begin{array}{ll} \Delta u=h,\qquad \hbox {in}\quad \Omega _t\\ u\,|_{\Gamma _t}=f,\qquad \nabla _{n_b}u \,|_{\Gamma _b}=g, \end{array}\right. \end{aligned}$$

we quote directly the following variational result:

Lemma 2.7

[52, Lemma 5.9] For a given function \(f\in H^{1/2}(\Gamma _t)\), the system

$$\begin{aligned} \left\{ \begin{array}{ll} \Delta {{\mathcal {H}}}(f)=0\qquad \hbox {in}\quad \Omega _t,\\ {{\mathcal {H}}}(f)|_{\Gamma _t}=f,\quad \nabla _{n_b}{{\mathcal {H}}}(f)|_{\Gamma _b}=0 \end{array}\right. \end{aligned}$$
(2.1)

admits a unique solution \(f_{{\mathcal {H}}}={{\mathcal {H}}}(f)\in H^1(\Omega _t)\), and there holds

$$\begin{aligned} \Vert {{\mathcal {H}}}(f)\Vert _{H^1(\Omega _t)}\leqq C(\Vert \Gamma _t\Vert _{H^{5/2} }) \Vert f\Vert _{H^{1/2}(\Gamma _t)}. \end{aligned}$$

The next proposition shows the existence and estimate for solutions in \(H^2(\Omega _t)\). Notice that when the contact angles are below \(\pi /2\), no singularity appears.

Proposition 2.1

Let \(h\in L^2(\Omega _t)\), \(f\in H^{3/2}(\Gamma _t)\), \(g\in H^{1/2}(\Gamma _b)\) and \(\Gamma _t \in H^{5/2}\) be given in \(\text{(MBVP) }\). The contact angles \(\omega _i\in (0, \pi /2)\). Then there exists a unique solution \(u\in H^2(\Omega _t)\) to \(\text{(MBVP) }\). Moreover, one has

$$\begin{aligned} \Vert u\Vert _{H^2(\Omega _t)}\leqq C\big (\Vert h\Vert _{L^2(\Omega )}+\Vert f\Vert _{H^{3/2}(\Gamma _t)}+\Vert g\Vert _{H^{1/2}(\Gamma _b)}\big ) \end{aligned}$$

with the constant \(C=C(\Vert \Gamma _t\Vert _{H^{5/2}})\).

Proof

This proposition is a direct conclusion from Proposition 5.1, Lemma 5.2 and Theorem 5.3 in [51]. \(\quad \square \)

Second, we consider the Neumann-boundary system

$$\begin{aligned} \text{(NBVP) }\quad \left\{ \begin{array}{l} \Delta u=h\quad \text {in}\quad \Omega _t, \\ \nabla _{n_t}u|_{\Gamma _t}=f,\quad \quad \nabla _{n_b} u |_{\Gamma _b}=g \end{array} \right. , \end{aligned}$$
(2.2)

satisfying the compatibility condition

$$\begin{aligned} \int _{\Omega _t} h{\text {d}}X=\int _{\Gamma _t} f{\text {d}}s+\int _{\Gamma _b}g{\text {d}}s. \end{aligned}$$

The following results for \(H^3\)-and-below case are needed in our paper:

Proposition 2.2

(1) Let \(h\in L^2(\Omega _t)\), \(f\in L^2(\Gamma _t)\), \(g\in L^2(\Gamma _b)\) in \(\text{(NBVP) }\). Then there exists a unique (up to an additive constant) variational solution \(u\in H^1(\Omega _t)\) to \(\text{(NBVP) }\) satisfying

$$\begin{aligned} \Vert u\Vert _{H^1(\Omega _t)}\leqq C(\Vert \Gamma _t\Vert _{H^{5/2}})\big (\Vert h\Vert _{L^2(\Omega _t)}+\Vert f\Vert _{L^{2}(\Gamma _t)}+\Vert g\Vert _{L^{2}(\Gamma _b)}+\Vert u\Vert _{L^2(\Omega _t)}\big ); \end{aligned}$$

(2) Let \(h\in H^s(\Omega _t)\), \(f\in H^{1/2+s}(\Gamma _t)\), \(g\in H^{1/2+s}(\Gamma _b)\) in \(\text{(NBVP) }\). The contact angles \(\omega _i\in (0, \pi /2)\). Then there exists a unique (up to an additive constant) solution \(u\in H^3(\Omega _t)\) to \(\text{(NBVP) }\) satisfying

$$\begin{aligned} \Vert u\Vert _{H^{2+s}(\Omega _t)}\leqq & {} C(\Vert \Gamma _t\Vert _{H^{3}})\big (\Vert h\Vert _{H^s(\Omega _t)}+\Vert f\Vert _{H^{1/2+s}(\Gamma _t)}\\{} & {} +\Vert g\Vert _{H^{1/2+s}(\Gamma _b)}+\Vert u\Vert _{L^2(\Omega _t)}\big ) \end{aligned}$$

for a constant \(s\in [0,1]\).

Proof

We only write a sketch for the proof here, since one can find similar details in our previous papers [51,52,53]. First, the existence of the variational solution \(u\in H^1(\Omega _t)\) can be proved directly by a standard variation procedure based on Lemma 4.4.3.1 [30]. Note that here one doesn’t require that the contact angles are below \(\pi /2\). Second, one needs to notice that, there is no singular part since the most singular part for \(\text{(NBVP) }\) behaves like \(r^{\pi /\omega }\) near the corners (see for example [30]). Therefore, one can have directly \(H^2\) estimate from Theorem 3.2.5 [10] and Proposition 5.1 [51], Theorem 4.3.1.4 [30]. Similarly to the proof for Proposition 5.13 [51], one has \(H^3\) estimate without singular part. As a result, the desired estimate in (2) can be proved by an interpolation. \(\quad \square \)

Remark 2.1

For the first estimate in Proposition 2.2, when one has additionally

$$\begin{aligned} \int _{\Omega _t}u{\text {d}}X=0, \end{aligned}$$

the norm \(\Vert u\Vert _{L^2(\Omega _t)}\) can be deleted from the right side during the proof.

Moreover, we present the \(H^4\) singularity decomposition and estimate.

Proposition 2.3

Let \(h\in H^2(\Omega _t)\), \(f\in H^{5/2}(\Gamma _t)\), \(g\in H^{5/2}(\Gamma _b)\) and \(\Gamma _t \in H^{4}\) in \(\text{(NBVP) }\). The contact angles \(\omega _i\in (0, \pi /2)\). Then there exists a unique (up to an additive constant) solution \(u\in H^3(\Omega _t)\) to \(\text{(NBVP) }\) such that

$$\begin{aligned} u=u_r+ u_s\quad \hbox {with the singular part}\ u_s=\chi _{\omega }(\omega _l)\chi _{l}\,c_l S_l\circ T_l+\chi _{\omega }(\omega _r)\chi _{r} \, c_r S_r\circ T_r, \end{aligned}$$

and the regular part \(u_r\in H^4(\Omega _t)\). Here the cut-off functions \(\chi _\omega , \chi _i\) are defined in the notation part. \(T_i\in H^{4}(\Omega _t)\) are boundary-straightening diffeomorphisms from \(\Omega _t\) onto the sectors \({{\mathcal {S}}}_{t,i}\) near the corner points \(p_i\) (from [51]), and \(S_i= r^{\pi /\omega _i}s(\theta )\) with r the radius with respect to \(p_i\) in \({{\mathcal {S}}}_{t, i}\) and \(s_i(\theta )\) some fixed sine or cosine functions.

Moreover, one has estimates for the singular coefficients \(c_i\) (\(i=l,r\)) and the regular part \(u_r\):

$$\begin{aligned} |c_l|+\Vert u_r\Vert _{H^4(\Omega _t)}\leqq C\big (\Vert h\Vert _{H^2(\Omega _t)}+\Vert f\Vert _{H^{5/2}(\Gamma _t)}+\Vert g\Vert _{H^{5/2}(\Gamma _b)}+\Vert u\Vert _{L^2(\Omega _t)}\big ) \end{aligned}$$

with the constant \(C=C(\Vert \Gamma _t\Vert _{H^{4}})\).

Proof

Similarly to the proofs of Proposition 5.17, Proposition 5.18 [51] for the mixed-boundary problem, the proof can be checked and follows line by line thanks to [30], so we omit the details here. \(\quad \square \)

Remark 2.2

A direct conclusion from this proposition is that when the contact angles \(\omega _i\in (0,\pi /3]\), the solution u to \(\text{(NBVP) }\) with the same right side will be in \(H^4(\Omega _t)\) with the corresponding estimate.

Remark 2.3

Based on the proposition above and estimate (9.15) in [24], when the contact angles \(\omega _i\in (\pi /3, \pi /2)\), one can have a more delicate and also natural estimate for \(\text{(NBVP) }\) with \(h\in H^{1+\epsilon }(\Omega _t)\), \(f\in H^{3/2+\epsilon }(\Gamma _t)\), \(g\in H^{3/2+\epsilon }(\Gamma _b)\) and \(\Gamma _t \in H^{4}\):

$$\begin{aligned} \Vert u\Vert _{H^{3+\epsilon }(\Omega _t)}\leqq & {} C(\Vert \Gamma _t\Vert _{H^{4}})\big (\Vert h\Vert _{H^{1+\epsilon }(\Omega _t)}+\Vert f\Vert _{H^{3/2+\epsilon }(\Gamma _t)}\\{} & {} +\Vert g\Vert _{H^{3/2+\epsilon }(\Gamma _b)}+\Vert u\Vert _{L^2(\Omega _t)}\big ), \end{aligned}$$

where \(\epsilon \in (0, \pi /\omega -2)\subset (0,1)\).

In the end, we recall some useful expressions and commutators from [52, 53, 60].

- \(D_tn_t\) and \(D_t\tau _t\). One has

$$\begin{aligned} D_t n_t=-\big ((\nabla v)^*n_t\big )^\top ,\quad D_t\tau _t=(\nabla _{\tau _t}v\cdot n_t\big )n_t. \end{aligned}$$
(2.3)

- \([D_t,{{\mathcal {H}}}]\). One has for a smooth function f on \(\Gamma _t\) that

$$\begin{aligned}{}[D_t,{{\mathcal {H}}}]f=\Delta ^{-1}\big (2\nabla v\cdot \nabla ^2f_{{{\mathcal {H}}}}+\Delta v\cdot \nabla f_{{{\mathcal {H}}}},\, (\nabla _{N_b}v-\nabla _v N_b)\cdot \nabla f_{{{\mathcal {H}}}}\big ) \quad \hbox {in}\quad \Omega _t,\nonumber \\ \end{aligned}$$
(2.4)

where \(\Delta ^{-1}(h,g)\) and \({{\mathcal {H}}}\) are defined in the notation part.

- \([D_t, {{\mathcal {N}}}]\). One has

$$\begin{aligned} \begin{aligned}{}[D_t,\,{{\mathcal {N}}}]f=&\nabla _{n_t}\Delta ^{-1}\Big (2\nabla v\cdot \nabla ^2f_{{{\mathcal {H}}}}+\Delta v\cdot \nabla f_{{{\mathcal {H}}}},\,(\nabla _{n_b}v-\nabla _{v}n_b)\cdot \nabla f_{{{\mathcal {H}}}}\Big )\\&\,-\nabla _{n_t}v\cdot \nabla f_{{{\mathcal {H}}}}-\nabla _{(\nabla f_{{{\mathcal {H}}}})^\top } v\cdot n_t\qquad \qquad \hbox {on}\quad \Gamma _t. \end{aligned} \end{aligned}$$
(2.5)

- \([D_t,\Delta _{\Gamma _t}]\). For a smooth function f on \(\Gamma _t\), there holds

$$\begin{aligned}{}[D_t,\,\Delta _{\Gamma _t}]f=2{{\mathcal {D}}}^2f\big (\tau _t,(\nabla _{\tau _t}v)^\top \big )-(\nabla f)^\top \cdot \Delta _{\Gamma _t}v+\kappa \nabla _{(\nabla f)^\top } v\cdot n_t \qquad \hbox {on}\quad \Gamma _t.\nonumber \\ \end{aligned}$$
(2.6)

- \([D_t,\,\Delta ^{-1}]\). We have

$$\begin{aligned} D_t\Delta ^{-1}(h,g)=\Delta ^{-1}(D_th,\,D_tg)+\Delta ^{-1}(h_1, g_1) \end{aligned}$$
(2.7)

with

$$\begin{aligned} h_1= & {} 2\nabla v\cdot \nabla ^2\Delta ^{-1}(h,g)+\Delta v\cdot \nabla \Delta ^{-1}(h,g),\\{} & {} g_1=(\nabla _{N_b}v-\nabla _{v}N_b)\cdot \nabla \Delta ^{-1}(h,g). \end{aligned}$$

- \([D_t, \nabla _{\tau _t}]\). Direct computations lead to

$$\begin{aligned}{}[D_t, \nabla _{\tau _t}]=(D_t\tau _t-\nabla _{\tau _t}v)\cdot \nabla =(n_t \nabla _{\tau _t}v\cdot n_t-\nabla _{\tau _t}v)\cdot \nabla =-(\nabla _{\tau _t}v\cdot \tau _t)\nabla _{\tau _t}.\nonumber \\ \end{aligned}$$
(2.8)

3 Reformulation of the Problem

In this section, we derive the equation for a good unknown \({\mathfrak {J}}\), which is slightly different from the quantity \(J=\nabla \kappa _{{{\mathcal {H}}}}\) introduced in [60] and used in our previous papers [52, 53].

To begin with, we define \(P_{v, v}\) by the Neumann-boundary system

$$\begin{aligned} \left\{ \begin{array}{ll} \Delta P_{v, v}=-tr (\nabla v \nabla v)\qquad \hbox {in}\quad \Omega _t\\ \nabla _{n_t}P_{v, v}|_{\Gamma _t}=C_{v,v}(t),\qquad \nabla _{n_b}P_{v, v}|_{\Gamma _b}=v\cdot \nabla _v n_b. \end{array} \right. ,\end{aligned}$$
(3.1)

where \(C_{v,v}(t)\) satisfies

$$\begin{aligned} |\Gamma _t| C_{v,v}(t) = -\int _{\Omega _t}tr (\nabla v \nabla v){\text {d}}X-\int _{\Gamma _b}v\cdot \nabla _v n_b {\text {d}}s. \end{aligned}$$

Moreover, due to the non-uniqueness of the variational solution to this Neumann problem, we assume that \(P_{v, v}\) satisfies

$$\begin{aligned} \int _{\Omega _t} P_{v, v}{\text {d}}X=0. \end{aligned}$$
(3.2)

In this way, we will have a unique solution to (3.1) in Sobolev space.

Let the pressure P be decomposed into

$$\begin{aligned} P={\mathfrak {K}}_{{{\mathcal {H}}}}+ P_{v,v}, \end{aligned}$$
(3.3)

where \({\mathfrak {K}}\) is the modified curvature on \(\Gamma _t\) defined by

$$\begin{aligned} {\mathfrak {K}}=\sigma \kappa -P_{v,v}\qquad \text{ on }\ {\Gamma _t}. \end{aligned}$$
(3.4)

Now, we can define the new quantity

$$\begin{aligned} {\mathfrak {J}}=\nabla \mathfrak {K}_{{{\mathcal {H}}}} \end{aligned}$$

and we are ready to derive its equation from \(\hbox {(WW)}\). In fact, since the derivation for the equation of \({\mathfrak {J}}\) is much similar to the derivation for J in [52], many details are omitted here.

First, recall the equation for the curvature \(\kappa \) from [52],

$$\begin{aligned} D_t\kappa =-\Delta _{\Gamma _t}v^\perp -v^\perp |\nabla _{\tau _t}n_t|^2+(\nabla _{\tau _t}\nabla _{v^\top }n_t)\cdot \tau _t-(\nabla _{\tau _t}v^\top )^\top \cdot \nabla n_t\cdot \tau _t\nonumber \\ \end{aligned}$$
(3.5)

or equivalently,

$$\begin{aligned} D_t\kappa =-\Delta _{\Gamma _t}v\cdot n_t-2\Pi (\tau _t)\cdot \nabla _{\tau _t}v \qquad \hbox {on}\quad \Gamma _t. \end{aligned}$$
(3.6)

Applying \(D_t\) to (3.6) and using (3.4), we have

$$\begin{aligned} D^2_t{\mathfrak {K}}=-\sigma \Delta _{\Gamma _t}D_tv \cdot n_t +2\sigma \Pi (\tau _t)\cdot \nabla _{\tau _t} J+\sigma R_1- D_t^2 P_{v,v}\qquad \hbox {on}\quad \Gamma _t\nonumber \\ \end{aligned}$$
(3.7)

with

$$\begin{aligned} \begin{aligned} R_1&= - [D_t, \Delta _{\Gamma _t}]v\cdot n_t-\Delta _{\Gamma _t}v\cdot D_tn_t+2 \Pi (\tau _t)\cdot \nabla _{\tau _t}\nabla P_{v,v}\\&\quad -2 \Pi (\tau _t)\cdot [D_t, \nabla _{\tau _t}]v-2 D_t\big (\Pi (\tau _t)\big )\cdot \nabla _{\tau _t}v. \end{aligned} \end{aligned}$$

Moreover, direct computations almost the same as in [52] lead to

$$\begin{aligned} D^2_t{\mathfrak {J}}=\nabla {{\mathcal {H}}}(D^2_t{\mathfrak {K}})+A_1+A_2+A_3, \end{aligned}$$
(3.8)

where the remainder terms are

$$\begin{aligned} \begin{aligned} A_1=\nabla [D_t,{{\mathcal {H}}}]D_t{\mathfrak {K}}=&\nabla \Delta ^{-1}\big (2\nabla v\cdot \nabla +\Delta v\cdot ,(\nabla _{n_b}v-\nabla _vn_b)\cdot \big )\\&\quad \Big (D_t{\mathfrak {J}}-\nabla \Delta ^{-1}\big (2\nabla v\cdot \nabla {\mathfrak {J}}\\&\quad +\Delta v\cdot {\mathfrak {J}},\,(\nabla _{n_b}v-\nabla _v n_b)\cdot {\mathfrak {J}}\big )+(\nabla v)^*{\mathfrak {J}}\Big ), \end{aligned} \end{aligned}$$
(3.9)
$$\begin{aligned} \begin{aligned} A_2&=\nabla D_t\Delta ^{-1}\big (2\nabla v\cdot \nabla {\mathfrak {J}}+\Delta v\cdot {\mathfrak {J}},(\nabla _{n_b}v-\nabla _vn_b)\cdot {\mathfrak {J}}\big )\\&=\nabla \Delta ^{-1}\big (2\nabla v\cdot \nabla ^2w_{A2}+\Delta v\cdot \nabla w_{A2},(\nabla _{n_b}v-\nabla _vn_b)\cdot \nabla w_{A2}\big )\\&\quad +\nabla \Delta ^{-1}(h_{A2},\, g_{A2}) \end{aligned} \end{aligned}$$
(3.10)

with

$$\begin{aligned} \begin{aligned} w_{A2}=&\Delta ^{-1}\big (2\nabla v\cdot \nabla {\mathfrak {J}}+\Delta v\cdot {\mathfrak {J}},\,(\nabla _{n_b}v-\nabla _v n_b)\cdot {\mathfrak {J}}\big ),\\ h_{A2}=&2\nabla v\cdot (\nabla D_t{\mathfrak {J}}-(\nabla v)^*{\mathfrak {J}})+2(\nabla D_tv-(\nabla v)^*\nabla v)\cdot \nabla {\mathfrak {J}}+D_t{\mathfrak {J}}\cdot \Delta v\\&+{\mathfrak {J}}\cdot (\Delta D_tv-\Delta v\cdot \nabla v-2\nabla v\cdot \nabla ^2 v),\\ g_{A2}=&(\nabla _vn_b-\nabla _{n_b}v)\cdot \nabla v\cdot {\mathfrak {J}}+\nabla _{n_b}D_t v\cdot {\mathfrak {J}}-(D_t v-\nabla _v v)\cdot \nabla n_b\cdot {\mathfrak {J}}\\&+\nabla _v\big ((\nabla v)^*n_b\big )^\top \cdot {\mathfrak {J}}+(\nabla _{n_b}v-\nabla _vn_b)\cdot D_t{\mathfrak {J}}, \end{aligned} \end{aligned}$$

and

$$\begin{aligned} A_3=-2(\nabla v)^*D_t{\mathfrak {J}}-(\nabla D_t v)^* {\mathfrak {J}}-\big ((\nabla v)^2\big )^*{\mathfrak {J}}+\big ((\nabla v)^*\nabla v\big ){\mathfrak {J}}. \end{aligned}$$
(3.11)

Here we note that the leading-order terms in \(A_1, A_2\) are like \({\mathfrak {J}},D_t{\mathfrak {J}},\nabla v,\nabla D_tv\).

Substituting (3.7) into (3.8) and applying Euler’s equation, we can have after some more computations that

$$\begin{aligned} D^2_t{\mathfrak {J}}=\sigma \nabla {{\mathcal {H}}}(\Delta _{\Gamma _t}{\mathfrak {J}}^\perp )+R_0+D_t\nabla {{\mathcal {H}}}( D_t P_{v,v} ). \end{aligned}$$
(3.12)

with

$$\begin{aligned} R_0= & {} -\sigma \nabla {{\mathcal {H}}}({\mathfrak {J}}\cdot \Delta _{\Gamma _t}n_t)+\sigma \nabla {{\mathcal {H}}}(n_t\cdot \Delta _{\Gamma _t}\nabla P_{v,v})\\{} & {} +\sigma \nabla {{\mathcal {H}}}(R_1 ) +A_1+A_2+A_3+[D_t, \nabla {{\mathcal {H}}}]( D_t P_{v,v}). \end{aligned}$$

Here, one observation is that since \(\nabla _{n_t} P_{v,v}\big |_{\Gamma _t} =C_{v,v}(t)\) from (3.1), we find immediately in \(R_0\) that

$$\begin{aligned} n_t\cdot \Delta _{\Gamma _t}\nabla P_{v,v} =[n_t,\Delta _{\Gamma _t} ] \cdot \nabla P_{v,v}. \end{aligned}$$

Moreover, we use the following Hodge decomposition from [52, 60]:

$$\begin{aligned} {{\mathcal {D}}}_t {\mathfrak {J}}=D_t {\mathfrak {J}}+\nabla {{\mathcal {P}}}_{{\mathfrak {J}},v} \end{aligned}$$
(3.13)

such that \({{\mathcal {D}}}_t {\mathfrak {J}}\) satisfies

$$\begin{aligned} \nabla \cdot {{\mathcal {D}}}_t{\mathfrak {J}}=0\quad \text{ on } \ \Omega _t,\quad \hbox {and}\quad n_b\cdot {{\mathcal {D}}}_t{\mathfrak {J}}\big |_{\Gamma _b}=0, \end{aligned}$$

and \({{\mathcal {P}}}_{{\mathfrak {J}},v}\) satisfies the Neumann-boundary system

$$\begin{aligned} \left\{ \begin{array}{ll} \Delta {{\mathcal {P}}}_{{\mathfrak {J}},v}=-tr (\nabla {\mathfrak {J}}\nabla v)\qquad \hbox {in}\quad \Omega _t\\ \nabla _{n_t} {{\mathcal {P}}}_{{\mathfrak {J}},v}|_{\Gamma _t}=C_{{\mathfrak {J}},v}(t)-\nabla _{\tau _t} {\mathfrak {K}}\nabla _{\tau _t} v\cdot n_t,\quad \nabla _{n_b} {{\mathcal {P}}}_{{\mathfrak {J}},v}|_{\Gamma _b}= {\mathfrak {J}}\cdot \nabla _v n_b \end{array}\right. \end{aligned}$$
(3.14)

with

$$\begin{aligned} |\Gamma _t| C_{{\mathfrak {J}},v}(t) = -\int _{\Omega _t}tr (\nabla {\mathfrak {J}}\nabla v){\text {d}}X+\int _{\Gamma _t}\nabla _{\tau _t} {\mathfrak {K}}\nabla _{\tau _t} v\cdot n_t {\text {d}}s-\int _{\Gamma _b}{\mathfrak {J}}\cdot \nabla _v n_b {\text {d}}s. \end{aligned}$$

Here the assumption

$$\begin{aligned} \int _{\Omega _t} {{\mathcal {P}}}_{{\mathfrak {J}}, v}{\text {d}}X=0 \end{aligned}$$

holds similarly to guarantee the uniqueness as for \(P_{v,v}\) system.

This implies immediately

$$\begin{aligned} \begin{aligned} D_t{{\mathcal {D}}}_t{\mathfrak {J}}= D_t(D_t{\mathfrak {J}}+\nabla {{\mathcal {P}}}_{{\mathfrak {J}},v})=D^2_t{\mathfrak {J}}+D_t\nabla {{\mathcal {P}}}_{{\mathfrak {J}},v}. \end{aligned} \end{aligned}$$

As a result, we finally conclude the equation for \({\mathfrak {J}}\) in the form based on (3.12),

$$\begin{aligned} D_t\big [{{\mathcal {D}}}_t {\mathfrak {J}}-\nabla {{\mathcal {H}}}\big ( D_t P_{v,v} -v\cdot (\nabla P_{v,v}|_c)\big )\big ]+\sigma {{\mathcal {A}}}{\mathfrak {J}}={{\mathcal {R}}}\end{aligned}$$
(3.15)

where the third-order elliptic operator \({{\mathcal {A}}}\) is defined in the same way as in [52, 60],

$$\begin{aligned} {{\mathcal {A}}}(w)=\nabla {{\mathcal {H}}}\big (-\Delta _{\Gamma _t}(w|_{\Gamma _t})^\perp \big ) \end{aligned}$$

for any smooth-enough function w defined on \(\Omega _t\), and the remainder term \({{\mathcal {R}}}\) is

$$\begin{aligned} {{\mathcal {R}}}=R_0 +D_t\nabla {{\mathcal {P}}}_{{\mathfrak {J}},v}+D_t\nabla {{\mathcal {H}}}\big ( v\cdot (\nabla P_{v,v}|_c)\big ). \end{aligned}$$

Moreover, one can find the definition of \(\nabla P_{v,v}|_c\) in the notation part.

Remark 3.1

Instead of (3.12), we write the equation of \({\mathfrak {J}}\) in a more complicated form (3.15). This happens because of technical reasons. In fact, \({{\mathcal {D}}}_t {\mathfrak {J}}\) is more handy to use in the energy estimates, while as a price we have the remainder \({{\mathcal {P}}}_{{\mathfrak {J}},v}\) part to deal with. Meanwhile, the part \(v\cdot (\nabla P_{v,v}|_c)\) is added to derive better estimate for \(D_tP_{v,v}\), see Lemma 4.7.

4 Lower-Order Energy Estimates

The lower-order energy E(t) is defined as

$$\begin{aligned} E(t)=\Vert \nabla _{\tau _t}{\mathfrak {J}}^\bot \Vert ^2_{L^2(\Gamma _t)} +\Vert {{\mathcal {D}}}_t {\mathfrak {J}}\Vert ^2_{L^2(\Omega _t)}+\Vert \Gamma _t\Vert ^2_{H^{5/2}}+\Vert v\Vert ^2_{H^{3/2}(\Omega _t)}, \end{aligned}$$

where we write

$$\begin{aligned} E_l(t)= \Vert \Gamma _t\Vert ^2_{H^{5/2}}+\Vert v\Vert ^2_{H^{3/2}(\Omega _t)}. \end{aligned}$$

Moreover, recalling from [52, 53] that our \(\text{(WW) }\) system has some dissipation at corner points, we have the following dissipation term at corner points

$$\begin{aligned} F(t)=\sum _{i=l,r}\big |(\sin \omega _i)\nabla _{\mathbf{\tau }_t}{\mathfrak {J}}^\perp |_{p_i}\big |^2. \end{aligned}$$

Theorem 4.1

Let the contact angles \(\omega _i\in (0,\pi /2)\). Assume that \(E(t), \int _0^T F(t){\text {d}}t\) are both bounded above in [0, T] for some \(T>0\). Then the following a priori estimate holds for system \(\text{(WW) }\):

$$\begin{aligned} \sup _{0\leqq t\leqq T} E(t)+\int _0^T F(t){\text {d}}t\leqq P(E(0))+\int _0^T P(E(t)){\text {d}}t. \end{aligned}$$

there \(P(\cdot )\) is some polynomial with positive constant coefficients depending on \(\sigma , \beta _c\).

One can see immediately that, our energy is defined mainly in forms of \({\mathfrak {J}}\). As a result, we need to related (vP), \(\Gamma _t\) and other quantities to \({\mathfrak {J}}\) firstly before we prove this energy estimate.

4.1 Dependence on E(t)

We show in this part that all the quantities related to our problem can be controlled by E(t). To start with, the following proposition focuses on the estimates for \(P_{v,v}, n_t, {\mathfrak {K}}_{{\mathcal {H}}}, {\mathfrak {J}}\):

Lemma 4.1

Assuming that \(E(t)\in L^\infty [0,T]\) for some \(T>0\), one has the following estimates:

$$\begin{aligned} \Vert P_{v, v}\Vert _{H^{2}(\Omega _t)} \leqq P(E_l(t)) \end{aligned}$$

and

$$\begin{aligned} \Vert n_t\Vert _{H^2(\Gamma _t)}+\Vert {\mathfrak {J}}\Vert _{H^{3/2}(\Omega _t)}+\Vert {\mathfrak {K}}_{{\mathcal {H}}}\Vert _{H^{5/2}(\Omega _t)}\leqq P(E(t)). \end{aligned}$$

Proof

(1) The first inequality. Applying Proposition 2.2 with an interpolation to \(P_{v,v}\) system (3.1), we get

$$\begin{aligned}{} & {} \Vert P_{v, v}\Vert _{H^{2}(\Omega _t)} \leqq C(\Vert \Gamma _t\Vert _{H^{5/2}})\big (\Vert tr(\nabla v \nabla v)\Vert _{L^2(\Omega _t)}\\{} & {} \quad +|\Gamma _t|^{1/2}|C_{v,v}(t)|+\Vert v\cdot \nabla _v n_b\Vert _{H^{1/2}(\Gamma _b)}+\Vert P_{v,v}\Vert _{L^2(\Omega _t)}\big ). \end{aligned}$$

For the right-hand side, we have firstly by Lemma 2.1 that

$$\begin{aligned} \Vert tr(\nabla v \nabla v)\Vert _{L^2(\Omega _t)}\leqq C(\Vert \Gamma _t\Vert _{H^{5/2}})\Vert v\Vert ^2_{H^{3/2}(\Omega _t)}. \end{aligned}$$

Second, it’s straightforward to find

$$\begin{aligned} |C_{v,v}(t)|\leqq C\Vert v\Vert ^2_{H^{3/2}(\Omega _t)} \end{aligned}$$

and

$$\begin{aligned} \begin{aligned} \Vert P_{v,v}\Vert _{L^2(\Omega _t)}&=\big \Vert P_{v,v}-\int _{\Omega _t}P_{v,v}{\text {d}}X\big \Vert _{L^2(\Omega _t)}\leqq C\Vert \nabla P_{v,v}\Vert _{L^2(\Omega _t)}\\&\leqq C(\Vert \Gamma _t\Vert _{H^{5/2}})\big ( \Vert tr(\nabla v \nabla v)\Vert _{L^2(\Omega _t)}\\&\quad +|\Gamma _t|^{1/2}|C_{v,v}(t)|+\Vert v\cdot \nabla _v n_b\Vert _{L^2(\Gamma _b)}\big ), \end{aligned} \end{aligned}$$

where (3.2) is used and the variational estimate above is proved similarly as Proposition 2.2(1).

Consequently, summing up all these inequalities above, we conclude that

$$\begin{aligned} \Vert P_{v, v}\Vert _{H^{2}(\Omega _t)} \leqq C(\Vert \Gamma _t\Vert _{H^{5/2}})\Vert v\Vert ^2_{H^{3/2}(\Omega _t)}\leqq P(E_l(t)). \end{aligned}$$

(2) The second inequality. In fact, the key point of the proof lies in the estimates for mean curvature \(\kappa \) and \({\mathfrak {K}}=\sigma \kappa -P_{v,v}\).

To begin with, one can see immediately that \({\mathfrak {K}}_{{\mathcal {H}}}\) satisfies the system

$$\begin{aligned} \left\{ \begin{array}{ll} \Delta {\mathfrak {K}}_{{\mathcal {H}}}=0\qquad \hbox {in}\quad \Omega _t\\ \nabla _{n_t}{\mathfrak {K}}_{{\mathcal {H}}}|_{\Gamma _t}={\mathfrak {J}}^\perp |_{\Gamma _t}\in H^1(\Gamma _t),\qquad \nabla _{n_b} {\mathfrak {K}}_{{\mathcal {H}}}|_{\Gamma _b}=0. \end{array}\right. \end{aligned}$$
(4.1)

Applying Proposition 2.2(2), we find

$$\begin{aligned} \begin{aligned} \Vert {\mathfrak {K}}_{{{\mathcal {H}}}}\Vert _{H^{5/2}(\Omega _t)}&\leqq C(\Vert \Gamma _t\Vert _{H^{5/2}})(\Vert {\mathfrak {J}}^\perp \Vert _{H^1(\Gamma _t)}+\Vert {\mathfrak {K}}_{{\mathcal {H}}}\Vert _{L^2(\Omega _t)})\\&\leqq C(\Vert \Gamma _t\Vert _{H^{5/2}})( \Vert \nabla _{\tau _t} {\mathfrak {J}}^\bot \Vert _{L^2(\Gamma _t)} +\Vert {\mathfrak {K}}_{{\mathcal {H}}}\Vert _{L^2(\Omega _t)}), \end{aligned} \end{aligned}$$
(4.2)

where interpolation for \({\mathfrak {K}}_{{\mathcal {H}}}\) is applied to \(\Vert {\mathfrak {J}}^\perp \Vert _{L^2}\).

Moreover, from the definition of \({\mathfrak {K}}\) and Lemma 2.7 we have

$$\begin{aligned} \Vert {\mathfrak {K}}_{{\mathcal {H}}}\Vert _{L^2(\Omega _t)}\leqq C(\Vert \Gamma _t\Vert _{H^{5/2}})\big (\Vert \kappa \Vert _{H^{1/2}(\Gamma _t)}+\Vert P_{v,v}\Vert _{H^{1/2}(\Gamma _t)}\big ), \end{aligned}$$

which together with the previous inequality leads to

$$\begin{aligned} \Vert {\mathfrak {K}}_{{{\mathcal {H}}}}\Vert _{H^{5/2}(\Omega _t)}\leqq C(\Vert \Gamma _t\Vert _{H^{5/2}}) (1+ \Vert \nabla _{\tau _t} {\mathfrak {J}}^\bot \Vert _{L^2(\Gamma _t)}+\Vert P_{v, v}\Vert _{H^{1/2}(\Omega _t)} ). \end{aligned}$$
(4.3)

This implies immediately the desired \(H^{3/2}\) estimate for \({\mathfrak {J}}\).

For the estimate for \(n_t\), we find by Lemma 2.4, (4.3) and part (1) that

$$\begin{aligned} \begin{aligned} \Vert \kappa \Vert _{H^{1}(\Gamma _t)}&\leqq \sigma ^{-1}(\Vert {\mathfrak {K}}\Vert _{H^{1}(\Gamma _t)}+\Vert P_{v,v}\Vert _{H^{1}(\Gamma _t)})\\&\leqq C(\Vert \Gamma _t\Vert _{H^{\frac{5}{2}}}) (1+ \Vert \nabla _{\tau _t} {\mathfrak {J}}^\bot \Vert _{L^2(\Gamma _t)}+\Vert P_{v, v}\Vert _{H^{3/2}(\Omega _t)}) \leqq P(E(t)). \end{aligned} \end{aligned}$$
(4.4)

As a result, the proof is finished thanks to the fact that

$$\begin{aligned} \kappa =\nabla _{\tau _t}n_t\cdot \tau _t\quad \hbox {where}\ \nabla _{\tau _t}n_t\Vert \tau _t\quad \hbox {on}\ \Gamma _t. \end{aligned}$$

\(\square \)

Lemma 4.2

Assuming that \(E(t)\in L^\infty [0,T]\) for some \(T>0\), one has the following estimates for v:

$$\begin{aligned} \Vert v\Vert _{H^2(\Omega _t)}+\Vert v^\perp \Vert _{H^{5/2}(\Gamma _t)}\leqq P(E(t)). \end{aligned}$$

Meanwhile, one also has

$$\begin{aligned} \Vert D_tP_{v, v}\Vert _{H^{1}(\Omega _t)} +\Vert {{\mathcal {P}}}_{{\mathfrak {J}}, v}\Vert _{H^1(\Omega _t)}\leqq P(E(t)) \end{aligned}$$

Proof

- \(H^1\) estimate for \({{\mathcal {P}}}_{{\mathfrak {J}}, v}\). From system (3.14) of \({{\mathcal {P}}}_{{\mathfrak {J}}, v}\), we can have the following variational estimate by Proposition 2.2(1) and Remark 2.1:

$$\begin{aligned} \begin{aligned} \Vert {{\mathcal {P}}}_{{\mathfrak {J}}, v}\Vert _{H^{1}(\Omega _t)} \leqq&C(\Vert \Gamma _t\Vert _{H^{5/2}})\big (\Vert tr(\nabla {\mathfrak {J}}\nabla v) \Vert _{L^2(\Omega _t)}\\&\quad +|\Gamma _t|^{1/2}|C_{{\mathfrak {J}},v}(t)|+\Vert \nabla _{\tau _t} {\mathfrak {K}}\cdot \nabla _{\tau _t} v\cdot n_t\Vert _{L^2(\Gamma _t)}\\&\quad +\Vert {\mathfrak {J}}\cdot \nabla _v n_b\Vert _{L^2(\Gamma _b)}\big ). \end{aligned} \end{aligned}$$

which Together with Lemma 4.1 thus leads to the desired estimate for \({{\mathcal {P}}}_{{\mathfrak {J}}, v}\).

- \(v^\perp \) estimate. In fact, it is straightforward to get by Lemma 2.4 that

$$\begin{aligned}{} & {} \Vert D_t \kappa \Vert _{H^{1/2}(\Gamma _t)} \leqq C(\Vert \Gamma _t\Vert _{H^{5/2}})\Vert D_t\kappa _{{\mathcal {H}}}\Vert _{H^1(\Omega _t)}\\{} & {} \quad \leqq C(\Vert \Gamma _t\Vert _{H^{5/2}})(\Vert \nabla D_t \kappa _{{\mathcal {H}}}\Vert _{L^2(\Omega _t)} +\Vert D_t \kappa \Vert _{L^2(\Gamma _t)} ). \end{aligned}$$

Rewriting \(\nabla \kappa _{{\mathcal {H}}}\) in terms of \({\mathfrak {J}}, \nabla P_{v,v}\) leads to

$$\begin{aligned} \begin{aligned} \Vert D_t \kappa \Vert _{H^{1/2}(\Gamma _t)}&\leqq C(\Vert \Gamma _t\Vert _{H^{5/2}})\big (\Vert D_t {\mathfrak {J}}\Vert _{L^2(\Omega _t)}+\Vert D_t\nabla P_{v,v}\Vert _{L^2(\Omega _t)}\\&\quad +\Vert \nabla v \cdot {\mathfrak {J}}\Vert _{L^2(\Omega _t)}+\Vert \nabla v\cdot \nabla P_{v,v}\Vert _{L^2(\Omega _t)}\\&\qquad +\Vert D_t \kappa \Vert _{L^2(\Gamma _t)} \big ) \\&\leqq P(E(t)) \big ( 1+\Vert \nabla {{\mathcal {P}}}_{{\mathfrak {J}}, v}\Vert _{L^2(\Omega _t)}+\Vert D_tP_{v,v}\Vert _{H^1(\Omega _t)}\\&\quad +\Vert P_{v,v}\Vert _{H^{3/2}(\Omega _t)}+\Vert {\mathfrak {J}}\Vert _{H^{3/2}(\Omega _t)} \\&\qquad +\Vert D_t \kappa \Vert _{L^2(\Gamma _t)} \big ). \end{aligned} \end{aligned}$$

Applying Lemma 4.1, we have

$$\begin{aligned} \Vert D_t \kappa \Vert _{H^{1/2}(\Gamma _t)}\leqq P(E(t)) \big ( 1+\Vert \nabla {{\mathcal {P}}}_{{\mathfrak {J}}, v}\Vert _{L^2(\Omega _t)}+\Vert D_tP_{v,v}\Vert _{H^1(\Omega _t)} +\Vert D_t \kappa \Vert _{L^2(\Gamma _t)} \big ). \end{aligned}$$

On the other hand, we rewrite (3.5) into

$$\begin{aligned} \Delta _{\Gamma _t}v^\perp =-D_t\kappa -v^\perp |\nabla _{\tau _t}n_t|^2+(\nabla _{\tau _t}\nabla _{v^\top }n_t)\cdot \tau _t-(\nabla _{\tau _t}v^\top )^\top \cdot \nabla n_t\cdot \tau _t, \end{aligned}$$

which together with the inequality above for \(\Vert D_t \kappa \Vert _{H^{1/2}(\Gamma _t)}\) and the estimate for \({{\mathcal {P}}}_{{\mathfrak {J}}, v}\) imply

$$\begin{aligned} \begin{aligned} \Vert \Delta _{\Gamma _t}v^\perp \Vert _{H^{1/2}(\Gamma _t)}&\leqq C(\Vert \Gamma _t\Vert _{H^{5/2}},\Vert n_t\Vert _{H^{5/2}(\Gamma _t)})\big (\Vert D_t\kappa \Vert _{H^{1/2}(\Gamma _t)} \\&\quad + \Vert v\Vert _{H^2(\Omega _t)}\big )\\&\leqq P(E(t))\big (1+\Vert \nabla {{\mathcal {P}}}_{{\mathfrak {J}}, v}\Vert _{L^2(\Omega _t)}+\Vert D_tP_{v,v}\Vert _{H^1(\Omega _t)}\\&\quad +\Vert D_t \kappa \Vert _{L^2(\Gamma _t)} + \Vert v\Vert _{H^2(\Omega _t)}\big )\\&\leqq P(E(t))\big (1 +\Vert D_tP_{v, v}\Vert _{H^1(\Omega _t)} + \Vert v\Vert _{H^2(\Omega _t)}\big ). \end{aligned} \end{aligned}$$
(4.5)

Here we use (3.6) to obtain

$$\begin{aligned} \Vert D_t \kappa \Vert _{L^2(\Gamma _t)}\leqq C(\Vert \Gamma _t\Vert _{H^{5/2}}) \Vert v\Vert _{H^2(\Omega _t)}. \end{aligned}$$

As a result, the estimate for \(v^\perp \) depends on the estimates for \(\Vert D_tP_{v, v}\Vert _{H^1(\Omega _t)}, \Vert v\Vert _{H^2(\Omega _t)}\).

- \(H^2\) estimate for v. Recalling system (4.10) for velocity potential \(\phi \) with \(v=\nabla \phi \), we have by Proposition 2.2 that

$$\begin{aligned} \begin{aligned} \Vert v\Vert _{H^2(\Omega _t)}&\leqq \Vert \phi \Vert _{H^3(\Omega _t)}\leqq C(\Vert \Gamma _t\Vert _{H^{5/2}})\big (\Vert v^\perp \Vert _{H^{3/2}(\Gamma _t)}+\Vert \phi \Vert _{L^2(\Omega _t)}\big )\\&\leqq C(\Vert \Gamma _t\Vert _{H^{5/2}})\big (\Vert v^\perp \Vert _{H^{3/2}(\Gamma _t)}+\Vert v\Vert _{L^2(\Omega _t)}\big )\\&\leqq P(E(t))\big (1 +\Vert v^\perp \Vert _{H^{3/2}(\Gamma _t)}\big ). \end{aligned} \end{aligned}$$

Bringing the above estimate into (4.5) and applying interpolations to \(\Vert v^\perp \Vert _{H^{3/2}(\Gamma _t)}\), we arrive at

$$\begin{aligned} \Vert \Delta _{\Gamma _t}v^\perp \Vert _{H^{1/2}(\Gamma _t)}+\Vert v\Vert _{H^2(\Omega _t)}\leqq P(E(t))\big (1+\Vert D_tP_{v, v}\Vert _{H^1(\Omega _t)} \big ). \end{aligned}$$
(4.6)

Therefore, it remains to deal with \(\Vert D_tP_{v, v}\Vert _{H^1(\Omega _t)}\).

- \(D_tP_{v,v}\) estimates. From the definition of \(P_{v, v}\), we derive the system for \(D_tP_{v,v}\):

$$\begin{aligned} \left\{ \begin{array}{ll} \Delta D_t P_{v, v}=-D_t tr(\nabla v \nabla v)+ 2tr(\nabla v \nabla ^2 P_{v, v})\qquad \hbox {in}\quad \Omega _t\\ \nabla _{n_t} D_t P_{v, v}|_{\Gamma _t}=C'_{v,v}(t)+\nabla _{n_t} v\cdot \nabla P_{v, v}|_{\Gamma _t},\quad \nabla _{n_b} D_tP_{v, v}|_{\Gamma _b}\\ \qquad =D_t(v\cdot \nabla _v n_b)+\nabla _{n_b} v\cdot \nabla P_{v, v}|_{\Gamma _b}. \end{array} \right. \end{aligned}$$
(4.7)

with

$$\begin{aligned} \int _{\Omega _t}D_tP_{v,v}{\text {d}}X=0. \end{aligned}$$

It is easy to see from Euler’s equation that

$$\begin{aligned} D_t tr(\nabla v \nabla v)=-2tr (\nabla v\cdot \nabla v\, \nabla v)-2tr \big (\nabla (\nabla P_{v,v}+\nabla {\mathfrak {K}}_{{\mathcal {H}}})\nabla v\big ), \end{aligned}$$
(4.8)

and similar expression can be derived for \(D_t(v\cdot \nabla _v n_b)\) on \(\Gamma _b\).

Moreover, we have

$$\begin{aligned} C'_{v,v} =|\Gamma _t|^{-1}\int _{\Omega _t} 2tr(D_t \nabla v\nabla v){\text {d}}X -|\Gamma _t|^{-2}\partial _t|\Gamma _t|\int _{\Omega _t}tr(\nabla v\nabla v){\text {d}}X, \end{aligned}$$

where \(\frac{\text {d}}{{\text {d}}t}|\Gamma _t|=\int _{\Gamma _t}\big (v^\perp \kappa +\nabla _{\tau _t}(v\cdot \tau _t)\big ){\text {d}}s\). So we obtain by Lemma 4.1, (4.3) and (4.4) that

$$\begin{aligned} |C'_{v,v}|\leqq & {} C(\Vert \Gamma _t\Vert _{H^{5/2}}, \Vert v\Vert _{H^{3/2}(\Omega _t)}, \Vert P_{v,v}\Vert _{H^2(\Omega _t)}, \Vert {\mathfrak {K}}_{{\mathcal {H}}}\Vert _{H^2(\Omega _t)}, \Vert \kappa \Vert _{L^2(\Gamma _t)}\big )\nonumber \\\leqq & {} P(E(t)). \end{aligned}$$
(4.9)

To prove the \(H^1\) estimate for \(D_tP_{v,v}\), we use the following form of variation equation:

$$\begin{aligned} \begin{aligned} \int _{\Omega _t}\big |\nabla D_tP_{v,v}\big |^2{\text {d}}X=&-\int _{\Omega _t}(\Delta D_tP_{v,v})D_tP_{v,v}{\text {d}}X+\int _{\Gamma _t}(\nabla _{n_t}D_tP_{v,v})D_tP_{v,v}{\text {d}}s\\&+\int _{\Gamma _b}(\nabla _{n_b}D_tP_{v,v})D_tP_{v,v}{\text {d}}s. \end{aligned} \end{aligned}$$

Substituting the expressions in (4.7) into the equality above, we can deal with the integrals one by one, where Lemma 2.1 is applied. For example, for the highest-order terms of v on the boundary, we use Green’s Formula and Lemma 2.1 to find

$$\begin{aligned} \begin{aligned}&\int _{\Gamma _t}\nabla _{n_t}v\cdot P_{v,v} \,D_tP_{v,v}{\text {d}}s+\int _{\Gamma _b}\nabla _{n_b}v\cdot P_{v,v} \,D_tP_{v,v}{\text {d}}s\\&\quad =\int _{\Omega _t}tr(\nabla v \nabla ^2 P_{v,v}) D_tP_{v,v}{\text {d}}X+\int _{\Omega _t}\nabla v\cdot \nabla P_{v,v}\cdot \nabla D_tP_{v,v}{\text {d}}X\\&\quad \leqq \Vert \nabla v\Vert _{L^4(\Omega _t)}\Vert \nabla ^2P_{v,v}\Vert _{L^2(\Omega _t)}\Vert D_tP_{v,v}\Vert _{L^4(\Omega _t)}\\&\qquad +|\nabla v\Vert _{L^4(\Omega _t)}\Vert \nabla P_{v,v}\Vert _{L^4(\Omega _t)}\Vert \nabla D_tP_{v,v}\Vert _{L^2(\Omega _t)}\\&\quad \leqq C(\Vert \Gamma _t\Vert _{H^{5/2}})\Vert v\Vert _{H^{3/2}(\Omega _t)}\Vert P_{v,v}\Vert _{H^2(\Omega _t)}\Vert D_tP_{v,v}\Vert _{H^1(\Omega _t)}. \end{aligned} \end{aligned}$$

The other integrals above can be handled similarly and the details are omitted.

As a result, we obtain by Lemma 4.1 the following desired estimate:

$$\begin{aligned} \Vert D_t P_{v, v}\Vert _{H^1(\Omega _t)} \leqq P(E(t))\big (\Vert P_{v,v}\Vert _{H^2(\Omega _t)}+\Vert {\mathfrak {K}}_{{\mathcal {H}}}\Vert _{H^{5/2}(\Omega _t)}+|C'_{v,v} |\big )\leqq P(E(t)). \end{aligned}$$

Consequently, applying (4.6) leads to the estimates for \(v^\perp \) and v. \(\quad \square \)

Remark 4.1

We know immediately from the proof of Lemma 4.2 the following estimate:

$$\begin{aligned} \Vert D_t\kappa \Vert _{H^{1/2}(\Gamma _t)}\leqq P(E(t)). \end{aligned}$$

Moreover, we also have

$$\begin{aligned} \Vert D_t{\mathfrak {K}}_{{\mathcal {H}}}\Vert _{H^1(\Omega _t)}\leqq \sigma \Vert D_t\kappa _{{\mathcal {H}}}\Vert _{H^1(\Omega _t)}+\Vert D_tP_{v,v}\Vert _{H^1(\Omega _t)}\leqq P(E(t)). \end{aligned}$$

Based on these lemmas above, we are ready to show some higher-order estimates.

Lemma 4.3

Let \(E(t)\in L^\infty [0,T]\) for some \(T>0\), then one has

$$\begin{aligned} \Vert P_{v, v}\Vert _{H^{5/2}(\Omega _t)}+\Vert n_t\Vert _{H^3(\Gamma _t)}\leqq P(E(t)). \end{aligned}$$

Proof

For \(P_{v,v}\), similar arguments as in the proof of Lemma 4.1 lead to

$$\begin{aligned} \begin{aligned} \Vert P_{v, v}\Vert _{H^{5/2}(\Omega _t)}&\leqq C(\Vert \Gamma _t\Vert _{H^{5/2}})\big (\Vert tr(\nabla v \nabla v)\Vert _{H^{1/2}(\Omega _t)}\\&\quad +|\Gamma _t|^{1/2}|C_{v,v}(t)|+\Vert v\cdot \nabla _v n_b\Vert _{H^{1}(\Gamma _b)}\big )\\&\leqq P(E(t))(1+\Vert v\Vert _{H^2(\Omega _t)} ). \end{aligned} \end{aligned}$$

Applying Lemma 4.2, we have the desired estimate.

Again, similar arguments as (4.4) in the proof of Lemma 4.1 lead directly to the estimate

$$\begin{aligned}{} & {} \Vert \kappa \Vert _{H^{2}(\Gamma _t)} \leqq \sigma ^{-1}\big (\Vert {\mathfrak {K}}\Vert _{H^{2}(\Gamma _t)}+\Vert P_{v,v}\Vert _{H^{2}(\Gamma _t)}\big )\\{} & {} \quad \leqq C(\Vert \Gamma _t\Vert _{H^{5/2}}) (1+ \Vert \nabla _{\tau _t} {\mathfrak {J}}^\bot \Vert _{L^2(\Gamma _t)}+\Vert P_{v, v}\Vert _{H^{5/2}(\Omega _t)}), \end{aligned}$$

which implies

$$\begin{aligned} \Vert n_t\Vert _{H^3(\Gamma _t)}\leqq P(E(t))( 1+\Vert P_{v, v}\Vert _{H^{5/2}(\Omega _t)}). \end{aligned}$$

Therefore, combining this with the estimate for \(P_{v,v}\), the proof is finished. \(\quad \square \)

We also present more estimates for v.

Lemma 4.4

Assuming that \(E(t)\in L^\infty [0,T]\) for some \(T>0\), one has

$$\begin{aligned} \Vert D_tv\Vert _{H^{3/2}(\Omega _t)} +\Vert D^2_tv\Vert _{L^2(\Omega _t)} \leqq P(E(t)) \end{aligned}$$

Proof

First, one recalls Euler’s equation to get that

$$\begin{aligned} \Vert D_tv\Vert _{H^{3/2}(\Omega _t)} \leqq C\big (1+\Vert {\mathfrak {J}}\Vert _{H^{3/2}(\Omega _t)}+ \Vert \nabla P_{v, v}\Vert _{H^{3/2}(\Omega _t)}\big ), \end{aligned}$$

and applying Lemma 4.3 leads to the estimate for \(D_tv\).

Second, taking \(D_t\) on Euler’s equation leads to the following estimate:

$$\begin{aligned} \Vert D^2_tv\Vert _{L^2(\Omega _t)} \leqq C(\Vert {{\mathcal {D}}}_t{\mathfrak {J}}\Vert _{L^2(\Omega _t)} +\Vert \nabla {{\mathcal {P}}}_{{\mathfrak {J}}, v}\Vert _{L^2(\Omega _t)} + \Vert D_t\nabla P_{v, v}\Vert _{L^2(\Omega _t)} ). \end{aligned}$$

Using Lemma 4.2 again, we can finish the estimate for \(D^2_t v\). \(\quad \square \)

We now consider the following Neumann system for velocity potential \(\phi \):

$$\begin{aligned} \left\{ \begin{array}{ll} \Delta \phi =0,\qquad \hbox {on}\quad \Omega _t\\ \nabla _{n_t}\phi |_{\Gamma _t}=v^\perp ,\quad \nabla _{n_b}\phi |_{\Gamma _b}=0, \end{array}\right. \end{aligned}$$
(4.10)

where in order to obtain the uniqueness we chose \(\phi \) to satisfy

$$\begin{aligned} \int _{\Omega _t}\phi {\text {d}}X=0 \end{aligned}$$

without loss of generality.

We show by singularity decompositions that \(\nabla v\) lies in \(L^\infty (\Omega _t)\), which is a key ingredient in the following estimates:

Lemma 4.5

Assume that \(E(t)\in L^\infty [0,T]\) for some \(T>0\), then there exists a unique \(\phi \in H^3(\Omega _t)\) to system (4.10) and one finds the singular decomposition

$$\begin{aligned} \phi =\phi _r+\phi _s \end{aligned}$$

where the regular part \(\phi _r\in H^4(\Omega _t)\), and the singular part is expressed in the same way as \(u_s\) in Proposition 2.3. Moreover, one has

$$\begin{aligned} \Vert \nabla v\Vert _{L^{\infty }(\Omega _t)}\leqq P(E(t)). \end{aligned}$$

Proof

In fact, applying Proposition 2.3, we know immediately about the existence of \(\phi \in H^3(\Omega _t)\) and the singular decomposition. Therefore, we have

$$\begin{aligned} v=\nabla \phi _r+\nabla \phi _s=v_r+v_s,\qquad \hbox {with}\quad v_r\in H^3(\Omega _t). \end{aligned}$$
(4.11)

It only remains to show the estimate for \(\nabla v\). Since we have

$$\begin{aligned} \phi _s=\chi _{\omega }(\omega _l)\chi _{l}\,c_l r^{\pi /\omega _l}\circ T_l+\chi _{\omega }(\omega _r)\chi _{r} \, c_r r^{\pi /\omega _r}\circ T_r \end{aligned}$$

where the singular part exists when \(\omega _i\in (\pi /3, \pi /2)\). In this case, we find by Proposition 2.3 that

$$\begin{aligned} \begin{aligned} \Vert \nabla ^2\phi _s\Vert _{L^\infty (\Omega _t)}&\leqq C(\Vert \Gamma _t\Vert _{H^4})\big (\Vert v^\perp \Vert _{H^{5/2}(\Gamma _t)}+\Vert \phi \Vert _{L^2(\Omega _t)}\big ) \big (\Vert \chi _{l}r^{\pi /\omega _l-2}\circ T_l\Vert _{L^\infty (\Omega _t)}\\&\quad +\Vert \chi _{r}r^{\pi /\omega _r-2}\circ T_r\Vert _{L^\infty (\Omega _t)}\big ) \end{aligned} \end{aligned}$$

with \(\pi /\omega _i-2\in (0,1)\). Consequently, we obtain

$$\begin{aligned} \Vert \nabla ^2\phi _s\Vert _{L^\infty (\Omega _t)}\leqq & {} C(\Vert \Gamma _t\Vert _{H^4})\big (\Vert v^\perp \Vert _{H^{5/2}(\Gamma _t)}+\Vert \phi \Vert _{L^2(\Omega _t)}\big )\\\leqq & {} C(\Vert \Gamma _t\Vert _{H^4})\Vert v^\perp \Vert _{H^{5/2}(\Gamma _t)} \end{aligned}$$

thanks to a direct variational estimate

$$\begin{aligned} \Vert \phi \Vert _{L^2(\Omega _t)}\leqq C(\Vert \Gamma _t\Vert _{H^{5/2}}) \Vert \nabla \phi \Vert _{L^2(\Omega _t)}\leqq C(\Vert \Gamma _t\Vert _{H^{5/2}})\Vert v^\perp \Vert _{L^2(\Gamma _t)}. \end{aligned}$$

Moreover, we also have by Proposition 2.3 the following estimate:

$$\begin{aligned} \Vert \phi _r\Vert _{H^4(\Omega _t)}\leqq C(\Vert \Gamma _t\Vert _{H^4})\big (\Vert v^\perp \Vert _{H^{5/2}(\Gamma _t)}+\Vert \phi \Vert _{L^2(\Omega _t)}\big ) \leqq C(\Vert \Gamma _t\Vert _{H^4})\Vert v^\perp \Vert _{H^{5/2}(\Gamma _t)}. \end{aligned}$$

As a result, we find that

$$\begin{aligned} \Vert \nabla v\Vert _{L^\infty (\Omega _t)}\leqq \Vert \nabla ^2 \phi _r\Vert _{L^\infty (\Omega _t)} +\Vert \nabla ^2 \phi _s\Vert _{L^\infty (\Omega _t)} \leqq C(\Vert \Gamma _t\Vert _{H^4})\Vert v^\perp \Vert _{H^{5/2}(\Gamma _t)}, \end{aligned}$$

which together with Lemma 4.2 lead to the desired estimate. \(\quad \square \)

As long as we have \(\Vert \nabla v\Vert _{L^\infty (\Omega _t)}\) estimate, we are able to deal with more higher-order estimates.

Lemma 4.6

Let \(E(t)\in L^\infty [0,T]\) for some \(T>0\), one has the following estimate:

$$\begin{aligned} \Vert P_{v, v}\Vert _{H^{3}(\Omega _t)}+ \Vert P_{{\mathfrak {J}}, v}\Vert _{H^{5/2}(\Omega _t)}+\Vert D_tP_{v,v}\Vert _{H^{3/2}(\Omega _t)}\leqq P(E(t)) \end{aligned}$$

Proof

- \(P_{v,v}\) estimate. In fact, similar arguments as in the proof of Lemma 4.1 show that

$$\begin{aligned} \Vert P_{v, v}\Vert _{H^{3}(\Omega _t)} \leqq P(E(t))(1+\Vert v\Vert _{H^2(\Omega _t)})(1+\Vert \nabla v\Vert _{L^\infty (\Omega _t)}). \end{aligned}$$

Consequently, applying Lemmas 4.2 and 4.5 leads to the desired estimate.

- Estimate for \(P_{{\mathfrak {J}}, v}\). Apply Proposition 2.2, we have

$$\begin{aligned} \begin{aligned} \Vert P_{{\mathfrak {J}}, v}\Vert _{H^{5/2}(\Omega _t)}&\leqq C(\Vert \Gamma _t\Vert _{H^{5/2}})\big (\Vert tr(\nabla {\mathfrak {J}}\nabla v) \Vert _{H^{1/2}(\Omega _t)}\\&\quad +|C_{{\mathfrak {J}},v}(t)|+\Vert \nabla _{\tau _t} {\mathfrak {K}}\nabla _{\tau _t} v\cdot n_t\Vert _{H^{1}(\Gamma _t)}\\&\quad +\Vert {\mathfrak {J}}\cdot \nabla _v n_b\Vert _{H^{1}(\Gamma _b)}+\Vert P_{{\mathfrak {J}}, v}\Vert _{L^{2}(\Omega _t)}\big ). \end{aligned} \end{aligned}$$

Using Lemmas 2.1, 2.2, 4.1, 4.2 and 4.5, we can finish the esitmate.

- \(H^{3/2}\) estimate for \(D_tP_{v,v}\). Following the proof of Lemma 4.2 and using Lemma 4.2 again to improve estimates for v, one can easily see that we have the desired estimate. \(\quad \square \)

We now give a high-order estimate for \(D_t P_{v, v}\).

Lemma 4.7

Assuming \(E(t)\in L^\infty [0,T]\) for some \(T>0\), one has

$$\begin{aligned} \big \Vert D_t P_{v, v}-v\cdot (\nabla P_{v,v}|_c)\big \Vert _{H^{5/2}(\Omega _t)} \leqq P(E(t)). \end{aligned}$$

Proof

We denote by

$$\begin{aligned} w=D_t P_{v, v}-v\cdot (\nabla P_{v,v}|_c)\quad \hbox {with}\quad \nabla P_{v,v}|_c=\chi _l(\nabla P_{v,v}|_{p_l})+\chi _r(\nabla P_{v,v}|_{p_r}). \end{aligned}$$

A direct computation using (4.7) leads to

$$\begin{aligned} \left\{ \begin{array}{ll} \Delta w=-trD_t(\nabla v \nabla v) +[\Delta ,v]\cdot (\nabla P_{v, v}- \nabla P_{v,v}|_c\big )-v\cdot \Delta (\nabla P_{v,v}|_c) \quad \hbox {in}\quad \Omega _t\\ \nabla _{n_t}w|_{\Gamma _t}=C_{v,v}'(t)+\nabla _{n_t} v\cdot (\nabla P_{v, v}-\nabla P_{v,v}|_{c})-v\cdot \nabla _{n_t}(\nabla P_{v,v}|_c)\big |_{\Gamma _t},\\ \nabla _{n_b}w|_{\Gamma _b}=D_t(v\cdot \nabla _v n_b)+\nabla _{n_b} v\cdot (\nabla P_{v, v}-\nabla P_{v,v}|_{c})-v\cdot \nabla _{n_b}(\nabla P_{v,v}|_c)\big |_{\Gamma _b}. \end{array} \right. \end{aligned}$$
(4.12)

Applying Proposition 2.2 to system (4.12), we have

$$\begin{aligned} \Vert w\Vert _{H^{5/2}(\Omega _t)}\leqq & {} C(\Vert \Gamma _t\Vert _{H^{5/2}})\big (\Vert \Delta w\Vert _{H^{1/2}(\Omega _t)}+\Vert \nabla _{n_t}w\Vert _{H^1(\Gamma _t)}\\{} & {} +\Vert \nabla _{n_b}w\Vert _{H^1(\Gamma _b)}+\Vert w\Vert _{L^2(\Omega _t)}\big ). \end{aligned}$$

We need to deal with the terms on the right side above. First, thanks to (4.8), Lemmas 4.2, 4.3 and 4.5, it is straightforward to check that

$$\begin{aligned}{} & {} \Vert D_t tr(\nabla v \nabla v)\Vert _{H^{1/2}(\Omega _t)}+\big \Vert [\Delta ,v]\cdot (\nabla P_{v, v}- \nabla P_{v,v}|_{c}) \big \Vert _{H^{1/2}(\Omega _t)} \\{} & {} \quad +\Vert v\cdot \Delta (\nabla P_{v,v}|_c)\Vert _{H^{1/2}(\Omega _t)} \leqq P(E(t)). \end{aligned}$$

Notice that here we need to use Lemma 2.2 for the product estimate of \(\Vert \nabla ^2{\mathfrak {K}}_{{\mathcal {H}}}\nabla v\Vert _{H^{1/2}(\Omega _t)}\) from \(\Vert D_t tr(\nabla v \nabla v)\Vert _{H^{1/2}(\Omega _t)}\):

$$\begin{aligned} \Vert \nabla ^2{\mathfrak {K}}_{{\mathcal {H}}}\nabla v\Vert _{H^{1/2}(\Omega _t)}\leqq & {} C\Vert \nabla ^2{\mathfrak {K}}_{{\mathcal {H}}}\Vert _{H^{1/2}(\Omega _t)}\big (\Vert \nabla v\Vert _{H^1(\Omega _t)}+ \Vert \nabla v\Vert _{L^\infty (\Omega _t)}\big )\\\leqq & {} P(E(t)). \end{aligned}$$

The estimate for \(\Vert \nabla ^2 P_{v,v}\nabla v\Vert _{H^{1/2}(\Omega _t)}\) also follows in a similar way.

Second, for the boundary terms, we only need to take care of the following estimate from \(\Vert \nabla _{n_t} v\cdot (\nabla P_{v, v}-\nabla P_{v,v}|_{c})\Vert _{H^1(\Gamma _t)}\):

$$\begin{aligned} \begin{aligned}&\big \Vert \nabla _{\tau _t}\nabla _{n_t} v\cdot (\nabla P_{v, v}-\nabla P_{v,v}|_{c})\big \Vert _{L^2(\Gamma _t)} \\&\quad \leqq \Vert \nabla _{\tau _t}\nabla _{n_t} v\cdot \chi _i(\nabla P_{v, v}-\nabla P_{v,v}|_{p_i})\Vert _{L^2(\Gamma _t)}\\&\qquad +\Vert \nabla _{\tau _t}\nabla _{n_t} v\cdot (1-\chi _l-\chi _r)\nabla P_{v, v}\Vert _{L^2(\Gamma _t)}\\&\quad \leqq \Vert \nabla _{\tau _t}\nabla _{n_t} v\cdot \chi _i(\nabla P_{v, v}-\nabla P_{v,v}|_{p_i})\Vert _{L^2(\Gamma _t)} +P(E(t)). \end{aligned} \end{aligned}$$
(4.13)

Here the estimate for the second term in the equality above holds thanks to (4.11) and Proposition 2.3.

To handle the first term in (4.13), using a straightening differeomorphism as \(T_i\) from Proposition 2.3 and (4.11), we know immediately that

$$\begin{aligned} \begin{aligned}&\big \Vert \nabla _{\tau _t}\nabla _{n_t} v\cdot \chi _i(\nabla P_{v, v}-\nabla P_{v,v}|_{p_i})\big \Vert _{L^2(\Gamma _t)}\\&\quad \leqq C(\Vert \Gamma _t\Vert _{H^{5/2}})\big \Vert (\nabla _{\tau _t}\nabla _{n_t} v)\circ T^{-1}_i\cdot (\nabla P_{v,v}-\nabla P_{v,v}|_{p_i})\circ T^{-1}_i\big \Vert _{L^2(0, r_0)}\\&\quad \leqq C(\Vert \Gamma _t\Vert _{H^{3}})\big (\big \Vert c_ir^{\pi /\omega -3} (\nabla P_{v,v}-\nabla P_{v,v}|_{p_i})\circ T^{-1}_i\big \Vert _{L^2(0, r_0)}\\&\qquad +\big \Vert (\partial ^2\nabla \phi _r)\circ T^{-1}_i\cdot (\nabla P_{v,v}-\nabla P_{v,v}|_{p_i})\circ T^{-1}_i\big \Vert _{L^2(0, r_0)}\big )\\&\quad \leqq P(E(t))\big (\Vert r^{\pi /\omega -3+3/4}\Vert _{L^2(0,r_0)}\big \Vert r^{-3/4}(\nabla P_{v,v}-\nabla P_{v,v}|_{p_i})\circ T^{-1}_i\big \Vert _{L^\infty (0, r_0)}+1\big ). \end{aligned} \end{aligned}$$
(4.14)

Moreover, applying Lemma 2.3 leads to

$$\begin{aligned} \begin{aligned}&\Vert r^{-3/4}(\nabla P_{v,v}-\nabla P_{v,v}|_{p_i})\circ T_i\Vert _{L^2(0, r_0)}\\&\quad \leqq C\Vert (\nabla P_{v,v}-\nabla P_{v,v}|_{p_i})\circ T_i\Vert _{H^{3/4}(0, r_0)} \quad \hbox {and}\\&\Vert r^{-3/4}(\nabla P_{v,v}-\nabla P_{v,v}|_{p_i})\circ T_i\Vert _{H^1(0, r_0)}\\&\quad \leqq C\Vert (\nabla P_{v,v}-\nabla P_{v,v}|_{p_i})\circ T_i\Vert _{H^{1+3/4}(0, r_0)}, \end{aligned} \end{aligned}$$

so we obtain by an interpolation the inequality

$$\begin{aligned} \begin{aligned}&\big \Vert r^{-3/4}(\nabla P_{v,v}-\nabla P_{v,v}|_{p_i})\circ T_i\big \Vert _{L^\infty (0, r_0)}\leqq C(\Vert \Gamma _t\Vert _{H^{5/2}})\big \Vert r^{-3/4}(\nabla P_{v,v}\\&\qquad -\nabla P_{v,v}|_{p_i})\circ T_i\big \Vert _{H^{3/4}(0, r_0)}\\&\quad \leqq C(\Vert \Gamma _t\Vert _{H^{5/2}})\Vert (\nabla P_{v,v}-\nabla P_{v,v}|_{p_i})\circ T_i\Vert _{H^{3/2}(0, r_0)}. \end{aligned} \end{aligned}$$

As a result, combining this with (4.14), we conclude that

$$\begin{aligned} \big \Vert \nabla _{\tau _t}\nabla _{n_t} v\cdot \chi _i(\nabla P_{v, v}-\nabla P_{v,v}|_{p_i})\big \Vert _{L^2(\Gamma _t)}\leqq & {} P(E(t))(\Vert P_{v,v}\Vert _{H^3(\Omega _t)}+1)\\\leqq & {} P(E(t)). \end{aligned}$$

Substituting this estimate into (4.13), we finally obtain

$$\begin{aligned} \big \Vert \nabla _{n_t} v\cdot (\nabla P_{v, v}-\nabla P_{v,v}|_{c})\big \Vert _{H^1(\Gamma _t)}\leqq P(E(t)). \end{aligned}$$

In the end, the estimate for \(\Vert \nabla _{n_b} v\cdot (\nabla P_{v, v}-\nabla P_{v,v}|_{c})\Vert _{H^1(\Gamma _b)}\) follows in a similar way, and the proof can be finished. \(\quad \square \)

In the end, we give the estimate of \(\int _{\Omega _t} \nabla D_t{{\mathcal {P}}}_{\mathfrak {J}, v}\cdot D_t \mathfrak {J}{\text {d}}X\).

Lemma 4.8

Let \(E(t)\in L^\infty [0,T]\) for some \(T>0\). Then, we have the following estimate:

$$\begin{aligned} \Big |\int _{\Omega _t} \nabla D_t{{\mathcal {P}}}_{\mathfrak {J}, v} \cdot D_t \mathfrak {J}{\text {d}}X-\frac{\text {d}}{{\text {d}}t}\int _{\Omega _t} \nabla {{\mathcal {P}}}_{\mathfrak {J}, v}\cdot \nabla v\cdot \nabla \mathfrak {K}_{{{\mathcal {H}}}}{\text {d}}X \Big |\leqq P(E(t)). \end{aligned}$$

Proof

Recalling system (3.14), we get

$$\begin{aligned} \left\{ \begin{array}{ll} \Delta D_t{{\mathcal {P}}}_{ \mathfrak {J},v}=- D_ttr(\nabla \mathfrak {J}\nabla v)+2tr(\nabla v \nabla ^2 {{\mathcal {P}}}_{ \mathfrak {J},v})\qquad \hbox {in}\quad \Omega _t\\ \nabla _{n_t} D_t{{\mathcal {P}}}_{{\mathfrak {J}},v}|_{\Gamma _t}= C'_{ \mathfrak {J}, v}-D_t(\nabla _{\tau _t}\mathfrak {K}_{{{\mathcal {H}}}}\cdot \nabla _{\tau _t} v \cdot n_t)+\nabla _{n_t}v\cdot \nabla {{\mathcal {P}}}_{ \mathfrak {J},v},\\ \nabla _{n_b} D_tP_{ \mathfrak {J},v}|_{\Gamma _b}=D_t( {\mathfrak {J}}\cdot \nabla _v n_b)+\nabla _{n_b}v\cdot \nabla P_{ \mathfrak {J},v}. \end{array}\right. \end{aligned}$$
(4.15)

Therefore, we have, directly that

$$\begin{aligned} \begin{aligned}&\int _{\Omega _t} \nabla D_t {{\mathcal {P}}}_{\mathfrak {J}, v} \cdot D_t \mathfrak {J}{\text {d}}X= \int _{\Omega _t} \nabla D_t {{\mathcal {P}}}_{\mathfrak {J}, v}\cdot D_t \nabla \mathfrak {K}_{{{\mathcal {H}}}}{\text {d}}X\\&\quad = \int _{\Omega _t} \nabla D_t {{\mathcal {P}}}_{\mathfrak {J}, v}\cdot \nabla D_t \mathfrak {K}_{{{\mathcal {H}}}}{\text {d}}X+\int _{\Omega _t} \nabla D_t {{\mathcal {P}}}_{\mathfrak {J}, v}\cdot \nabla v\cdot \nabla \mathfrak {K}_{{{\mathcal {H}}}}{\text {d}}X\\&\quad = \int _{\Omega _t} \nabla D_t {{\mathcal {P}}}_{\mathfrak {J}, v}\cdot \nabla D_t \mathfrak {K}_{{{\mathcal {H}}}}{\text {d}}X \\&\qquad +\frac{\text {d}}{{\text {d}}t}\int _{\Omega _t} \nabla {{\mathcal {P}}}_{\mathfrak {J}, v}\cdot \nabla v\cdot \nabla \mathfrak {K}_{{{\mathcal {H}}}}{\text {d}}X-\int _{\Omega _t} \nabla {{\mathcal {P}}}_{\mathfrak {J}, v}\cdot D_t(\nabla v\cdot \nabla \mathfrak {K}_{{{\mathcal {H}}}}){\text {d}}X\\&\qquad +\int _{\Omega _t}\nabla v\cdot \nabla {{\mathcal {P}}}_{{\mathfrak {J}},v}\cdot \nabla v\cdot \nabla {\mathfrak {K}}_{{\mathcal {H}}}{\text {d}}X, \end{aligned} \end{aligned}$$

which leads to the following equality:

$$\begin{aligned} \begin{aligned}&\int _{\Omega _t} \nabla D_t {{\mathcal {P}}}_{\mathfrak {J}, v} \cdot D_t \mathfrak {J}{\text {d}}X-\frac{\text {d}}{{\text {d}}t}\int _{\Omega _t} \nabla {{\mathcal {P}}}_{\mathfrak {J}, v}\cdot \nabla v\cdot \nabla \mathfrak {K}_{{{\mathcal {H}}}}{\text {d}}X\\&\quad =\int _{\Omega _t} \nabla D_t {{\mathcal {P}}}_{\mathfrak {J}, v}\cdot \nabla D_t \mathfrak {K}_{{{\mathcal {H}}}}{\text {d}}X-\int _{\Omega _t} \nabla {{\mathcal {P}}}_{\mathfrak {J}, v}\cdot D_t(\nabla v\cdot \nabla \mathfrak {K}_{{{\mathcal {H}}}}){\text {d}}X\\&\qquad +\int _{\Omega _t}\nabla v\cdot \nabla {{\mathcal {P}}}_{{\mathfrak {J}},v}\cdot \nabla v\cdot \nabla {\mathfrak {K}}_{{\mathcal {H}}}{\text {d}}X. \end{aligned} \end{aligned}$$

To finish the proof, the key lies in the analysis for the first integral. The remainder part is controlled by P(E(t)) thanks to Lemma 4.1, Lemma 4.2, Remark 4.1 and Lemma 4.6.

For the first integral on the right side of the equality above, we have by Green’s Formula that

$$\begin{aligned} \begin{aligned}&\int _{\Omega _t} \nabla D_t {{\mathcal {P}}}_{\mathfrak {J}, v}\cdot \nabla D_t \mathfrak {K}_{{{\mathcal {H}}}}{\text {d}}X\\&=\int _{\Gamma _t}(\nabla _{n_t} D_t {{\mathcal {P}}}_{\mathfrak {J}, v}) D_t \mathfrak {K}_{{{\mathcal {H}}}}{\text {d}}s+\int _{\Gamma _b}(\nabla _{n_b} D_t{{\mathcal {P}}}_{\mathfrak {J}, v}) D_t \mathfrak {K}_{{{\mathcal {H}}}}{\text {d}}s-\int _{\Omega _t}\Delta D_t{{\mathcal {P}}}_{{\mathfrak {J}},v} D_t \mathfrak {K}_{{{\mathcal {H}}}}{\text {d}}X\\&=\int _{\Gamma _t}\big [C'_{{\mathfrak {J}}, v}-D_t(\nabla _{\tau _t}\mathfrak {K}_{{{\mathcal {H}}}}\cdot \nabla _{\tau _t} v \cdot n_t)+\nabla _{n_t}v\cdot \nabla {{\mathcal {P}}}_{ \mathfrak {J},v}\big ] D_t \mathfrak {K}_{{{\mathcal {H}}}}{\text {d}}s \\&\quad +\int _{\Gamma _b}D_t( {\mathfrak {J}}\cdot \nabla _v n_b)D_t \mathfrak {K}_{{{\mathcal {H}}}}{\text {d}}s\\&\quad +\int _{\Gamma _b}\nabla _{n_b}v\cdot \nabla P_{ \mathfrak {J},v}\,D_t \mathfrak {K}_{{{\mathcal {H}}}}{\text {d}}s +\int _{\Omega _t} D_ttr(\nabla \mathfrak {J}\nabla v)\,D_t \mathfrak {K}_{{{\mathcal {H}}}}{\text {d}}X\\&\quad -\int _{\Omega _t}2tr(\nabla v \nabla ^2 {{\mathcal {P}}}_{ \mathfrak {J},v})\,D_t \mathfrak {K}_{{{\mathcal {H}}}}{\text {d}}X\\&=C'_{{\mathfrak {J}}, v}\int _{\Gamma _t}D_t \mathfrak {K}{\text {d}}s-\int _{\Gamma _t}\big (\nabla _{\tau _t}D_t\mathfrak {K}\cdot \nabla _{\tau _t} v \cdot n_t\big ) D_t \mathfrak {K}{\text {d}}s+\int _{\Gamma _b}(D_t{\mathfrak {J}}\cdot \nabla _v n_b) D_t \mathfrak {K}_{{{\mathcal {H}}}}{\text {d}}s\\&\quad +\int _{\Omega _t}tr(\nabla D_t{\mathfrak {J}}\nabla v)D_t{\mathfrak {K}}_{{\mathcal {H}}}{\text {d}}X +\hbox {l.o.t.}, \end{aligned} \end{aligned}$$
(4.16)

where the remainder lower-order terms can be controlled by P(E(t)) in a similar way as before.

Now we deal with the terms in (4.16) one by one. To begin with, we write from the definition of \(C_{{\mathfrak {J}},v}\) in system (3.14) that

$$\begin{aligned} \begin{aligned} \frac{\text {d}}{{\text {d}}t}\big (|\Gamma _t|C_{{\mathfrak {J}},v}\big )&=-\int _{\Omega _t}D_t tr(\nabla {\mathfrak {J}}\nabla v){\text {d}}X+\int _{\Gamma _t}D_t(\nabla _{\tau _t}{\mathfrak {K}}\nabla _{\tau _t}v\cdot n_t {\text {d}}s)\\&\quad -\int _{\Gamma _b}D_t({\mathfrak {J}}\cdot \nabla _v n_b {\text {d}}s)\\&=-\int _{\Omega _t}tr(\nabla D_t{\mathfrak {J}}\nabla v){\text {d}}X+\int _{\Gamma _t}\nabla _{\tau _t}D_t{\mathfrak {K}}\nabla _{\tau _t}v\cdot n_t{\text {d}}s\\&\quad -\int _{\Gamma _b}D_tJ\cdot \nabla _v n_b{\text {d}}s+\hbox {l.o.t.}, \end{aligned} \end{aligned}$$

where the lower-order terms can be controlled again and hence the details are omitted. Applying Green’s Formula again and using the decompositions

$$\begin{aligned} \begin{aligned}&\nabla D_t{\mathfrak {K}}_{{\mathcal {H}}}=(\nabla _{\tau _t}D_t{\mathfrak {K}})\tau _t+(\nabla _{n_t}D_t{\mathfrak {K}}_{{\mathcal {H}}})n_t\qquad \hbox {on}\quad \Gamma _t,\\&\nabla D_t{\mathfrak {K}}_{{\mathcal {H}}}=(\nabla _{\tau _b}D_t{\mathfrak {K}}_{{\mathcal {H}}})\tau _b+(\nabla _{n_b}D_t{\mathfrak {K}}_{{\mathcal {H}}})n_b\qquad \hbox {on}\quad \Gamma _b, \end{aligned} \end{aligned}$$

we find

$$\begin{aligned} \begin{aligned}&\frac{\text {d}}{{\text {d}}t}\big (|\Gamma _t|C_{{\mathfrak {J}},v}\big )\\&\quad =-\int _{\Gamma _t}D_t{\mathfrak {J}}\cdot \nabla v\cdot n_t{\text {d}}s-\int _{\Gamma _b}D_t{\mathfrak {J}}\cdot \nabla v\cdot n_b{\text {d}}s\\&\qquad +\int _{\Gamma _t}\nabla _{\tau _t}D_t{\mathfrak {K}}\nabla _{\tau _t}v\cdot n_t{\text {d}}s-\int _{\Gamma _b}D_tJ\cdot \nabla _v n_b{\text {d}}s+\hbox {l.o.t.}\\&\quad =-\int _{\Gamma _t}\nabla D_t{\mathfrak {K}}_{{\mathcal {H}}}\cdot \nabla v\cdot n_t{\text {d}}s+\int _{\Gamma _t}\nabla _{\tau _t}D_t{\mathfrak {K}}\nabla _{\tau _t}v\cdot n_t{\text {d}}s\\&\qquad -\int _{\Gamma _b}\nabla D_t{\mathfrak {K}}_{{\mathcal {H}}}\cdot \nabla v\cdot n_b{\text {d}}s\\&\qquad -\int _{\Gamma _b}\nabla D_t{\mathfrak {K}}_{{\mathcal {H}}}\cdot \nabla _v n_b{\text {d}}s+\hbox {l.o.t.}\\&\quad =-\int _{\Gamma _t}\nabla _{n_t}D_t{\mathfrak {K}}_{{\mathcal {H}}}\nabla _{n_t}v\cdot n_t{\text {d}}s-\int _{\Gamma _b}\nabla _{n_b}D_t{\mathfrak {K}}_{{\mathcal {H}}}\nabla _{n_b}v\cdot n_b{\text {d}}s\\&\quad -\int _{\Gamma _b}\nabla _{\tau _b}D_t{\mathfrak {K}}\cdot \nabla _{\tau _b} v\cdot n_b{\text {d}}s \\&\qquad -\int _{\Gamma _b}\nabla D_t{\mathfrak {K}}_{{\mathcal {H}}}\cdot \nabla _v n_b{\text {d}}s+\hbox {l.o.t.}. \end{aligned} \end{aligned}$$
(4.17)

Here, for the last three terms on \(\Gamma _b\), thanks to the assumption that \(n_b\) is constant near the contact points \(p_i\) and \(v\cdot n_b|_{\Gamma _b}=0\), we know that \(\nabla _{\tau _b} v\cdot n_b=-v\cdot \nabla _{\tau _b}n_b\) and \(\nabla _v n_b=(\nabla _v n_b\cdot \tau _b)\tau _b\) vanish near \(p_i\). Moreover, we also know from the definition of \({\mathfrak {K}}_{{\mathcal {H}}}\) that

$$\begin{aligned} \nabla _{n_b}D_t{\mathfrak {K}}_{{\mathcal {H}}}|_{\Gamma _b}=[\nabla _{n_b}, D_t]{\mathfrak {K}}_{{\mathcal {H}}}=\nabla _{n_b}v\cdot \nabla {\mathfrak {K}}_{{\mathcal {H}}}. \end{aligned}$$

Consequently, applying Lemmas 2.5 and 2.6, we have

$$\begin{aligned} \begin{aligned}&\int _{\Gamma _b}\nabla _{\tau _b}D_t{\mathfrak {K}}_{{\mathcal {H}}}\cdot \nabla _{\tau _b} v\cdot n_b{\text {d}}s+\int _{\Gamma _b}\nabla D_t{\mathfrak {K}}_{{\mathcal {H}}}\cdot \nabla _v n_b{\text {d}}s\\&\quad \leqq \Vert \nabla _{\tau _b}D_t{\mathfrak {K}}_{{\mathcal {H}}}\Vert _{{\tilde{H}}^{-1/2}(\Gamma _b)}\big (\Vert v\cdot \nabla _{\tau _b}n_b\Vert _{{\tilde{H}}^{1/2}(\Gamma _b)}+\Vert \nabla _v n_b\cdot \tau _b\Vert _{{\tilde{H}}^{1/2}(\Gamma _b)}\big )\\&\quad \leqq C(\Vert \Gamma _t\Vert _{H^{5/2}})\Vert D_t{\mathfrak {K}}_{{\mathcal {H}}}\Vert _{H^1(\Omega _t)}\big (\Vert v\cdot \nabla _{{{\mathcal {H}}}(\tau _b)}{{\mathcal {H}}}(n_b)\Vert _{H^1(\Omega _t)}\\&\qquad +\Vert \nabla _v {{\mathcal {H}}}(n_b)\cdot {{\mathcal {H}}}(\tau _b)\Vert _{H^1(\Omega _t)}\big )\leqq P(E(t)) \end{aligned} \end{aligned}$$

and

$$\begin{aligned} \int _{\Gamma _b}\nabla _{n_b}D_t{\mathfrak {K}}_{{\mathcal {H}}}\nabla _{n_b}v\cdot n_b{\text {d}}s\leqq P(E(t)). \end{aligned}$$

For the first integral in (4.17), applying Green’s Formula again leads to

$$\begin{aligned}{} & {} \int _{\Gamma _t}\nabla _{n_t}D_t{\mathfrak {K}}_{{\mathcal {H}}}\nabla _{n_t}v\cdot n_t{\text {d}}s=\int _{\Omega _t}[\Delta , D_t]{\mathfrak {K}}_{{\mathcal {H}}}\nabla _{{{\mathcal {H}}}(n_t)}v\cdot {{\mathcal {H}}}(n_t){\text {d}}s\\{} & {} \quad -\int _{\Gamma _b}[\nabla _{n_b},D_t]{\mathfrak {K}}_{{\mathcal {H}}}\nabla _{{{\mathcal {H}}}(n_t)}v\cdot {{\mathcal {H}}}(n_t){\text {d}}s, \end{aligned}$$

so this term is controlled by P(E(t)) again.

Therefore, summing up all these estimates above and going back to (4.17), we arrive at

$$\begin{aligned} \frac{\text {d}}{{\text {d}}t}\big (|\Gamma _t|C_{{\mathfrak {J}},v}\big )= \hbox {l.o.t.} \end{aligned}$$

with the lower-order terms controlled by P(E(t)). This leads to the following estimate

$$\begin{aligned} C'_{{\mathfrak {J}},v}\leqq P(E(t)). \end{aligned}$$

For the moment, we can go back and deal with the other terms in (4.16). In fact, using similar arguments as above, we also conclude that

$$\begin{aligned} \begin{aligned}&\int _{\Gamma _t}\big (\nabla _{\tau _t}D_t\mathfrak {K}_{{{\mathcal {H}}}}\cdot \nabla _{\tau _t} v \cdot n_t\big ) D_t \mathfrak {K}_{{{\mathcal {H}}}}{\text {d}}s\\&\quad +\int _{\Gamma _b}(D_t{\mathfrak {J}}\cdot \nabla _v n_b) D_t \mathfrak {K}_{{{\mathcal {H}}}}{\text {d}}s-\int _{\Omega _t}tr(\nabla D_t{\mathfrak {J}}\nabla v)D_t{\mathfrak {K}}_{{\mathcal {H}}}{\text {d}}X\leqq P(E(t)) \end{aligned} \end{aligned}$$

Therefore, the proof is finished. \(\quad \square \)

4.2 Boundary Terms at the Contact Points

Lemma 4.9

We have the following equation on \(\Gamma _b\):

$$\begin{aligned} D_t{\mathfrak {J}}= -\frac{\sigma ^2 }{\beta _c}(n_t\cdot \tau _b) (\nabla _{\tau _t} {\mathfrak {J}})^\bot \tau _b+R_{c1}. \end{aligned}$$
(4.18)

then there holds at the contact points \(p_i\) (\(i=l, r\)) that

$$\begin{aligned} (D_t{\mathfrak {J}})^\perp \,( \nabla _{\mathbf{\tau }_t}{\mathfrak {J}})^\perp \big |_{p_l}= & {} -\frac{\sigma ^2}{\beta _c}F_l(t) +R_{c2,l},\quad (D_t{\mathfrak {J}})^\perp \,( \nabla _{\mathbf{\tau }_t}{\mathfrak {J}})^\perp \big |_{p_r}\\= & {} \frac{\sigma ^2}{\beta _c}F_r(t) + R_{c2,r} \end{aligned}$$

where

$$\begin{aligned} \begin{aligned}&R_{c1}=r_c\,\tau _b- ( {\mathfrak {J}}\cdot D_t n_b)\,n_b,\\&\quad R_{c2,i}= \big (-r_c +\cot \omega ({\mathfrak {J}}\cdot D_t n_b)\big )\, (\sin \omega ) \nabla _{\mathbf{\tau }_t}{\mathfrak {J}}^\perp \big |_{p_i},\qquad \hbox {and}\\&r_c=-\sigma \sin \omega (\nabla _{\tau _t}v\cdot D_t n_t)+\sigma \tau _b\cdot D_t n_t\,(\nabla _{\tau _t}v\cdot n_t)\\&\qquad -\sigma \sin \omega \big (\nabla P_{v,v}\cdot \nabla _{\tau _t}n_t +[D_t, \nabla _{\tau _t} ] v\cdot n_t\big )\\&\qquad - \beta _c D_t \nabla P_{v,v}\cdot \tau _b \end{aligned} \end{aligned}$$

satisfy

$$\begin{aligned} |R_{c1}|\leqq P(E(t)),\quad |R_{c2,i}|\leqq P(E(t))F(t)^{1/2}. \end{aligned}$$

Proof

The proof follows the proof of Lemma 7.1 in [52]. For the estimates, we apply Lemmas 2.1, 4.3, 4.5 and 4.7. Here the differences compared to the proof of Lemma 7.1 in [52] lie in that we have different \({\mathfrak {J}}\) and

$$\begin{aligned} \nabla _{\tau _t}\nabla P_{v,v}\cdot n_t=-\nabla P_{v,v}\cdot \nabla _{\tau _t}n_t \end{aligned}$$

in \(r_c\) due to the definition of \(P_{v,v}\). \(\quad \square \)

4.3 Proof of Theorem 4.1

We are ready to show the a priori energy estimate in Theorem 4.1. In fact, taking \(L^2(\Omega _t)\) inner product with \({{\mathcal {D}}}_t J-\nabla {{\mathcal {H}}}( D_t P_{v,v} - v\cdot \big (\nabla P_{v,v}|_{c})\big )\) on both sides of (3.15), we have

$$\begin{aligned} \begin{aligned}&\int _{\Omega _t} D_t \big [ {{\mathcal {D}}}_t {\mathfrak {J}}-\nabla {{\mathcal {H}}}\big ( D_t P_{v,v} - v\cdot (\nabla P_{v,v}|_{c})\big )\big ]\\&\qquad \cdot \big [ {{\mathcal {D}}}_t {\mathfrak {J}}-\nabla {{\mathcal {H}}}\big ( D_t P_{v,v} - v\cdot (\nabla P_{v,v}|_{c})\big )\big ]{\text {d}}X \\&\qquad +\sigma \int _{\Omega _t} {{\mathcal {A}}}{\mathfrak {J}}\cdot \big [ {{\mathcal {D}}}_t {\mathfrak {J}}-\nabla {{\mathcal {H}}}\big ( D_t P_{v,v} - v\cdot (\nabla P_{v,v}|_{c})\big )\big ]{\text {d}}X \\&\quad =\int _{\Omega _t}{{\mathcal {R}}}\cdot \big [ {{\mathcal {D}}}_t {\mathfrak {J}}-\nabla {{\mathcal {H}}}\big ( D_t P_{v,v} - v\cdot (\nabla P_{v,v}|_{c})\big )\big ]{\text {d}}X. \end{aligned} \end{aligned}$$
(4.19)

We deal with these integrals above one by one. To get started, for the first term on the left side, we rewrite it as

$$\begin{aligned} \begin{aligned}&\int _{\Omega _t} D_t \big [ {{\mathcal {D}}}_t {\mathfrak {J}}-\nabla {{\mathcal {H}}}\big ( D_t P_{v,v} - v\cdot (\nabla P_{v,v}|_{c})\big )\big ]\\&\qquad \cdot \big [ {{\mathcal {D}}}_t {\mathfrak {J}}-\nabla {{\mathcal {H}}}\big ( D_t P_{v,v} - v\cdot (\nabla P_{v,v}|_{c})\big )\big ]{\text {d}}X\\&\quad =\frac{1}{2} \frac{\text {d}}{{\text {d}}t}\int _{\Omega _t} \big | {{\mathcal {D}}}_t {\mathfrak {J}}-\nabla {{\mathcal {H}}}\big ( D_t P_{v,v} - v\cdot (\nabla P_{v,v}|_{c})\big )\big |^2{\text {d}}X. \end{aligned} \end{aligned}$$

For the second term on the left side of (4.19), we have by Green’s Formula that

$$\begin{aligned} \begin{aligned}&\int _{\Omega _t} {{\mathcal {A}}}{\mathfrak {J}}\cdot \big [ {{\mathcal {D}}}_t {\mathfrak {J}}-\nabla {{\mathcal {H}}}\big ( D_t P_{v,v} - v\cdot (\nabla P_{v,v}|_{c})\big )\big ]{\text {d}}X\\&=-\int _{\Gamma _t} \Delta _{\Gamma _t} {\mathfrak {J}}^\bot \, {{\mathcal {D}}}_t {\mathfrak {J}}\cdot n_t{\text {d}}s +\int _{\Gamma _t} \Delta _{\Gamma _t} {\mathfrak {J}}^\bot \, \nabla {{\mathcal {H}}}\big ( D_t P_{v,v} - v\cdot (\nabla P_{v,v}|_{c})\big )\cdot n_t{\text {d}}s. \end{aligned} \end{aligned}$$
(4.20)

For the first term on the right side of (4.20), one deduces from Hodge decomposition (3.13) and integration by parts as in [53] that

$$\begin{aligned} \begin{aligned}&-\int _{\Gamma _t} \Delta _{\Gamma _t} {\mathfrak {J}}^\bot \, {{\mathcal {D}}}_t {\mathfrak {J}}\cdot n_t{\text {d}}s\\&\quad =\int _{\Gamma _t} \nabla _{\tau _t} {\mathfrak {J}}^\bot \, \nabla _{\tau _t}(D_t {\mathfrak {J}}\cdot n_t ){\text {d}}s-(D_t {\mathfrak {J}})^\bot \nabla _{\tau _t} {\mathfrak {J}}^\bot \big |^{p_l}_{p_r} -\int _{\Gamma _t} \Delta _{\Gamma _t} {\mathfrak {J}}^\bot (\nabla {{\mathcal {P}}}_{{\mathfrak {J}},v}\cdot n_t){\text {d}}s\\&\quad =\frac{1}{2}\,\frac{\text {d}}{{\text {d}}t}\int _{\Gamma _t}| \nabla _{\tau _t} {\mathfrak {J}}^\bot |^2{\text {d}}s-\int _{\Gamma _t} \nabla _{\tau _t} {\mathfrak {J}}^\bot \, [ D_t, \nabla _{\tau _t}] {\mathfrak {J}}^\bot {\text {d}}s\\&\qquad - \int _{\Gamma _t} \nabla _{\tau _t} {\mathfrak {J}}^\bot \, \nabla _{\tau _t}({\mathfrak {J}}\cdot D_t n_t ){\text {d}}s +\frac{\sigma ^2}{\beta _c}F(t) \\&\qquad -R_{c2,l}+R_{c2,r}-\int _{\Gamma _t} \Delta _{\Gamma _t} {\mathfrak {J}}^\bot (\nabla {{\mathcal {P}}}_{{\mathfrak {J}},v}\cdot n_t){\text {d}}s, \end{aligned} \end{aligned}$$

where we apply Lemma 4.9 at corner points.

For the second term in (4.20), one has in a similar way as above that

$$\begin{aligned} \begin{aligned}&\int _{\Gamma _t} \Delta _{\Gamma _t} {\mathfrak {J}}^\bot \, \nabla {{\mathcal {H}}}\big ( D_t P_{v,v} - v\cdot (\nabla P_{v,v}|_{c})\big )\cdot n_t{\text {d}}s\\&\quad =-\int _{\Gamma _t}\nabla _{\tau _t} J^\bot \cdot \nabla _{\tau _t}\nabla _{n_t}{{\mathcal {H}}}\big ( D_t P_{v,v} - v\cdot (\nabla P_{v,v}|_{c})\big ){\text {d}}s \\&\qquad +\nabla _{\tau _t} J^\bot \cdot \nabla _{n_t}{{\mathcal {H}}}\big ( D_t P_{v,v} - v\cdot (\nabla P_{v,v}|_{c})\big )\big |^{p_r}_{p_l}\\&\quad \leqq P(E(t))\big (1+ F(t)^{1/2}\big ), \end{aligned} \end{aligned}$$

where Lemma 2.1 and Lemma 4.7 are used.

As a result, going back to (4.19), we summerise that

$$\begin{aligned} \begin{aligned}&\frac{1}{2} \frac{\text {d}}{{\text {d}}t}\int _{\Omega _t} \big | {{\mathcal {D}}}_t {\mathfrak {J}}-\nabla {{\mathcal {H}}}\big ( D_t P_{v,v}\\&\qquad - v\cdot (\nabla P_{v,v}|_{c})\big )\big |^2{\text {d}}X+\frac{\sigma }{2}\,\frac{\text {d}}{{\text {d}}t}\int _{\Gamma _t}| \nabla _{\tau _t} {\mathfrak {J}}^\bot |^2{\text {d}}s + \frac{\sigma ^3}{2\beta _c} F(t)\\&\quad \leqq P(E(t))+\sigma \int _{\Gamma _t} \nabla _{\tau _t} {\mathfrak {J}}^\bot \, [ D_t, \nabla _{\tau _t}] {\mathfrak {J}}^\bot {\text {d}}s\\&\qquad +\sigma \int _{\Gamma _t} \nabla _{\tau _t} {\mathfrak {J}}^\bot \, \nabla _{\tau _t}({\mathfrak {J}}\cdot D_t n_t ){\text {d}}s +\sigma \int _{\Gamma _t} \Delta _{\Gamma _t} {\mathfrak {J}}^\bot (\nabla {{\mathcal {P}}}_{{\mathfrak {J}},v}\cdot n_t){\text {d}}s\\&\qquad +\int _{\Omega _t}{{\mathcal {R}}}\cdot \big [ {{\mathcal {D}}}_t {\mathfrak {J}}-\nabla {{\mathcal {H}}}\big ( D_t P_{v,v} - v\cdot (\nabla P_{v,v}|_{c})\big )\big ]{\text {d}}X. \end{aligned} \end{aligned}$$

Moreover, direct estimates similarly to those in [53] lead to the estimate

$$\begin{aligned} \begin{aligned}&\sigma \int _{\Gamma _t} \nabla _{\tau _t} {\mathfrak {J}}^\bot \, [ D_t, \nabla _{\tau _t}] {\mathfrak {J}}^\bot {\text {d}}s+\sigma \int _{\Gamma _t} \nabla _{\tau _t} {\mathfrak {J}}^\bot \, \nabla _{\tau _t}({\mathfrak {J}}\cdot D_t n_t ){\text {d}}s\\&\qquad +\sigma \int _{\Gamma _t} \Delta _{\Gamma _t} {\mathfrak {J}}^\bot (\nabla {{\mathcal {P}}}_{{\mathfrak {J}},v}\cdot n_t){\text {d}}s\\&\quad \leqq P(E(t))+\frac{\sigma ^3}{4\beta _c} F(t), \end{aligned} \end{aligned}$$

so we obtain

$$\begin{aligned} \begin{aligned}&\frac{1}{2} \frac{\text {d}}{{\text {d}}t}\Big (\int _{\Omega _t} \big | {{\mathcal {D}}}_t {\mathfrak {J}}-\nabla {{\mathcal {H}}}\big ( D_t P_{v,v} - v\cdot (\nabla P_{v,v}|_{c})\big )\big |^2{\text {d}}X +\sigma \int _{\Gamma _t}| \nabla _{\tau _t} {\mathfrak {J}}^\bot |^2{\text {d}}s\Big )\\&\qquad + \frac{\sigma ^3}{4\beta _c} F(t)\\&\quad \leqq P(E(t)) +\int _{\Omega _t}{{\mathcal {R}}}\cdot \big [ {{\mathcal {D}}}_t {\mathfrak {J}}-\nabla {{\mathcal {H}}}\big ( D_t P_{v,v} - v\cdot (\nabla P_{v,v}|_{c})\big )\big ]{\text {d}}X. \end{aligned} \end{aligned}$$
(4.21)

In order to finish the energy estimate, we still need to deal with the integral on the right side involving the remainder term \({{\mathcal {R}}}\) in (3.15).

- Estimates for the part \(R_0-\sigma \nabla {{\mathcal {H}}}(R_1)\) in \({{\mathcal {R}}}\).

Lemma 4.10

For the remainder term \(R_0\) defined in (3.12), we have the estimate

$$\begin{aligned} \Vert R_0-\sigma \nabla {{\mathcal {H}}}(R_1) \Vert _{L^2(\Omega _t)}\leqq P(E(t)). \end{aligned}$$

Proof

Recalling from (3.12), we know that

$$\begin{aligned} R_0-\sigma \nabla {{\mathcal {H}}}(R_1)= & {} -\sigma \nabla {{\mathcal {H}}}(J\cdot \Delta _{\Gamma _t}n_t)+\sigma \nabla {{\mathcal {H}}}(n_t\cdot \Delta _{\Gamma _t}\nabla P_{v,v}) \\{} & {} +[D_t, \nabla {{\mathcal {H}}}]( D_t P_{v,v} ) +A_1+A_2+A_3 \end{aligned}$$

where \(R_1\) and \(A_1,A_2,A_3\) are defined in (3.7) and (3.9), (3.10), (3.8) respectively.

For the term \(\nabla {{\mathcal {H}}}(n_t\cdot \Delta _{\Gamma _t}\nabla P_{v,v})\), we have

$$\begin{aligned} \Vert \nabla {{\mathcal {H}}}(n_t\cdot \Delta _{\Gamma _t}\nabla P_{v,v})\Vert _{L^2(\Omega _t)}=\Vert \nabla {{\mathcal {H}}}([n_t, \Delta _{\Gamma _t}]\nabla P_{v,v})\Vert _{L^2(\Omega _t)}\leqq P(E(t)), \end{aligned}$$

where the boundary condition \(\nabla _{n_t}P_{v,v}|_{\Gamma _t}=C_{v,v}(t)\) and Lemma 4.6 are applied.

For the term \([D_t, \nabla {{\mathcal {H}}}]( D_t P_{v,v} )\), direct computations using (2.4) and similar arguments as in the proof of Lemma 4.2 lead to

$$\begin{aligned}{} & {} \Vert [D_t, \nabla {{\mathcal {H}}}]( D_t P_{v,v} )\Vert _{L^2(\Omega _t)}\leqq \Vert \nabla v\cdot \nabla {{\mathcal {H}}}( D_t P_{v,v} )\Vert _{L^2(\Omega _t)}\\{} & {} \quad +\Vert [D_t, {{\mathcal {H}}}]( D_t P_{v,v} )\Vert _{H^1(\Omega _t)}\leqq P(E(t)). \end{aligned}$$

The estimates for the other terms follow from lemmas in the previous section and can be done similarly as [53], so we omit the details here. \(\quad \square \)

- The part \(\sigma \nabla {{\mathcal {H}}}(R_1 )\) in \({{\mathcal {R}}}\). In fact, the integral from the right side of (4.21) is rewritten as

$$\begin{aligned} \begin{aligned}&\int _{\Omega _t}\nabla {{\mathcal {H}}}(R_1 )\cdot \big [ {{\mathcal {D}}}_t {\mathfrak {J}}-\nabla {{\mathcal {H}}}\big ( D_t P_{v,v} - v\cdot (\nabla P_{v,v}|_{c})\big )\big ]{\text {d}}X\\&\quad = \int _{\Gamma _t}R_1\, {{\mathcal {D}}}_t {\mathfrak {J}}\cdot n_t {\text {d}}s-\int _{\Gamma _t}R_1 \nabla _{n_t} {{\mathcal {H}}}\big ( D_t P_{v,v} - v\cdot (\nabla P_{v,v}|_{c})\big ) {\text {d}}s\\&\quad =\int _{\Gamma _t}R_1\, D_t {\mathfrak {J}}\cdot n_t {\text {d}}s+\int _{\Gamma _t}R_1 \,\nabla {{\mathcal {P}}}_{{\mathfrak {J}},v}\cdot n_t {\text {d}}s\\&\qquad -\int _{\Gamma _t}R_1 \nabla _{n_t} {{\mathcal {H}}}\big ( D_t P_{v,v} - v\cdot (\nabla P_{v,v}|_{c})\big ) {\text {d}}s \end{aligned} \end{aligned}$$

On the other hand, noticing from (3.7) that \(R_1\) contains terms like \(\nabla ^2 v,\nabla n_t, \kappa \) and \(\nabla ^2 P_{v, v}\) and using (4.11) and lemmas in the previous subsections, we have

$$\begin{aligned}{} & {} \int _{\Gamma _t}R_1 \nabla _{n_t} {{\mathcal {H}}}\big ( D_t P_{v,v} - v\cdot (\nabla P_{v,v}|_{c})\big ) {\text {d}}s\\{} & {} \quad \leqq \int _{\Gamma _t}|R_1 |{\text {d}}s\, \Vert \nabla _{n_t} {{\mathcal {H}}}\big ( D_t P_{v,v} - v\cdot (\nabla P_{v,v}|_{c})\big )\Vert _{L^\infty (\Gamma _t)}\leqq P(E(t)). \end{aligned}$$

The details in the estimate above are omitted and we only note that for \(\partial ^2 v\) terms in \(R_1\), we can have the following estimate thanks to (4.11):

$$\begin{aligned} \int _{\Gamma _t}|\nabla ^2 v|{\text {d}}s\leqq \int _{\Gamma _t}|\nabla ^3\phi _r|{\text {d}}s+\int _{\Gamma _t}|\nabla ^3\phi _s|{\text {d}}s\leqq P(E(t)). \end{aligned}$$

Similarly but more easily, we also obtain

$$\begin{aligned} \int _{\Gamma _t}R_1 \,\nabla {{\mathcal {P}}}_{{\mathfrak {J}},v}\cdot n_t {\text {d}}s\leqq P(E(t)). \end{aligned}$$

In addition, for the part \(\int _{\Gamma _t}R_1 D_t {\mathfrak {J}}\cdot n_t {\text {d}}s\), we put \(D_t\) out of the integral:

$$\begin{aligned} \begin{aligned}&\int _{\Gamma _t}R_1 D_t {\mathfrak {J}}\cdot n_t {\text {d}}s=\int _{\Gamma _t}R_1 D_t({\mathfrak {J}}-{\mathfrak {J}}|_c){\text {d}}s+\int _{\Gamma _t}R_1 D_t({\mathfrak {J}}|_c){\text {d}}s\\&\quad =\frac{\text {d}}{{\text {d}}t}\int _{\Gamma _t} R_1\,({\mathfrak {J}}-{\mathfrak {J}}|_c){\text {d}}s-\int _{\Gamma _t}D_t R_1\,({\mathfrak {J}}-{\mathfrak {J}}|_c){\text {d}}s\\&\qquad -\int _{\Gamma _t}R_1\,({\mathfrak {J}}-{\mathfrak {J}}|_c)D_t{\text {d}}s+\int _{\Gamma _t}R_1 D_t({\mathfrak {J}}|_c){\text {d}}s. \end{aligned} \end{aligned}$$

Here we need to take care of the terms in \(\int _{\Gamma _t}D_t R_1\,({\mathfrak {J}}-{\mathfrak {J}}|_c){\text {d}}s\) and \(\int _{\Gamma _t}R_1\,({\mathfrak {J}}-{\mathfrak {J}}|_c)D_t{\text {d}}s\). In fact, similarly to the analysis in (4.14), the key terms like \(\Vert \partial ^2 v ({\mathfrak {J}}-{\mathfrak {J}}|_c)\Vert _{L^2(\Gamma _t)}\) can be handled as follows:

$$\begin{aligned}{} & {} \Vert \partial ^2 v ({\mathfrak {J}}-{\mathfrak {J}}|_c)\Vert _{L^2(\Gamma _t)}\nonumber \\{} & {} \quad \leqq C(\Vert \Gamma _t\Vert _{H^{5/2}})\Vert r^{\delta } (\partial ^2 v)\circ T^{-1}_i \Vert _{L^\infty (\Gamma _t)}\Vert r^{-\delta }({\mathfrak {J}}-{\mathfrak {J}}|_{p_i})\circ T^{-1}_i\Vert _{L^2(\Gamma _t)}\nonumber \\{} & {} \quad \leqq P(E(t)). \end{aligned}$$
(4.22)

there the constant \(\delta \in (0,1)\) is chosen to satisfy

$$\begin{aligned} \delta +\pi /\omega _i-3>0,\quad \hbox { when}\quad \omega _i\in (\pi /3, \pi /2). \end{aligned}$$

Consequently, we derive

$$\begin{aligned} \int _{\Gamma _t}R_1 D_t {\mathfrak {J}}\cdot n_t {\text {d}}s=\frac{\text {d}}{{\text {d}}t}\int _{\Gamma _t} R_1 ({\mathfrak {J}}- {\mathfrak {J}}|_c)\cdot n_t {\text {d}}s+l.o.t., \end{aligned}$$

where all the lower-order terms are controlled by P(E(t)).

As a result, we conclude that

$$\begin{aligned}{} & {} \int _{\Omega _t}\nabla {{\mathcal {H}}}(R_1 )\cdot \big [ {{\mathcal {D}}}_t {\mathfrak {J}}-\nabla {{\mathcal {H}}}\big ( D_t P_{v,v} - v\cdot (\nabla P_{v,v}|_{c})\big )\big ]{\text {d}}X\\{} & {} \quad =\frac{\text {d}}{{\text {d}}t}\int _{\Gamma _t} R_1 ({\mathfrak {J}}- {\mathfrak {J}}|_c)\cdot n_t {\text {d}}s+l.o.t. \end{aligned}$$

with the lower-order terms controlled by P(E(t)).

- The terms \(D_t\nabla {{\mathcal {P}}}_{{\mathfrak {J}},v}+D_t\nabla {{\mathcal {H}}}\big (v\cdot (\nabla P_{v,v}|_c)\big )\) in \({{\mathcal {R}}}\). First, one has directly from Lemma 4.8 that

$$\begin{aligned} \int _{\Omega _t} \nabla D_t{{\mathcal {P}}}_{\mathfrak {J}, v} \cdot D_t \mathfrak {J}{\text {d}}X=\frac{\text {d}}{{\text {d}}t}\int _{\Omega _t} \nabla {{\mathcal {P}}}_{\mathfrak {J}, v}\cdot \nabla v\cdot \nabla \mathfrak {K}_{{{\mathcal {H}}}}{\text {d}}X +l.o.t. \end{aligned}$$

where the lower-order terms are all controlled by P(E(t)).

Second, for the integral

$$\begin{aligned} \int _{\Omega _t} \nabla D_t{{\mathcal {P}}}_{\mathfrak {J}, v} \cdot \nabla {{\mathcal {H}}}\big ( D_t P_{v,v} - v\cdot (\nabla P_{v,v}|_{c})\big ){\text {d}}X, \end{aligned}$$

applying Green’s formula and similar calculations as in (4.16) show directly that it can be controlled by P(E(t)), and we omit the details.

Therefore, we conclude that

$$\begin{aligned}{} & {} \int _{\Omega _t} \nabla D_t{{\mathcal {P}}}_{\mathfrak {J}, v} \cdot \big [ {{\mathcal {D}}}_t {\mathfrak {J}}-\nabla {{\mathcal {H}}}\big ( D_t P_{v,v} - v\cdot (\nabla P_{v,v}|_{c})\big )\big ]{\text {d}}X\\{} & {} \quad =\frac{\text {d}}{{\text {d}}t}\int _{\Omega _t} \nabla {{\mathcal {P}}}_{\mathfrak {J}, v}\cdot \nabla v\cdot \nabla \mathfrak {K}_{{{\mathcal {H}}}}{\text {d}}X+l.o.t., \end{aligned}$$

where all the lower-order terms are controlled by P(E(t)).

Summing up all these estimates related to \({{\mathcal {R}}}\) above and going back to (4.21), we obtain the following estimate:

$$\begin{aligned} \begin{aligned}&\frac{1}{2} \frac{\text {d}}{{\text {d}}t}\Big (\int _{\Omega _t} \big | {{\mathcal {D}}}_t {\mathfrak {J}}-\nabla {{\mathcal {H}}}\big ( D_t P_{v,v} - v\cdot (\nabla P_{v,v}|_{c})\big )\big |^2{\text {d}}X +\sigma \int _{\Gamma _t}| \nabla _{\tau _t} {\mathfrak {J}}^\bot |^2{\text {d}}s\Big )\\&\qquad + \frac{\sigma ^3}{4\beta _c} F(t)\\&\quad \leqq P(E(t))+\frac{\text {d}}{{\text {d}}t}\int _{\Gamma _t} R_1 ({\mathfrak {J}}- {\mathfrak {J}}|_c)\cdot n_t {\text {d}}s +\frac{\text {d}}{{\text {d}}t}\int _{\Omega _t} \nabla {{\mathcal {P}}}_{\mathfrak {J}, v}\cdot \nabla v\cdot \nabla \mathfrak {K}_{{{\mathcal {H}}}}{\text {d}}X. \end{aligned} \end{aligned}$$

Integrating on both sides with respect to time on [0, t], we have

$$\begin{aligned} \begin{aligned}&\frac{1}{2}\int _{\Omega _t} \big | {{\mathcal {D}}}_t {\mathfrak {J}}-\nabla {{\mathcal {H}}}\big ( D_t P_{v,v} - v\cdot (\nabla P_{v,v}|_{c})\big )\big |^2{\text {d}}X\\&\quad +\frac{\sigma }{2}\int _{\Gamma _t}| \nabla _{\tau _t} {\mathfrak {J}}^\bot |^2{\text {d}}s + \frac{\sigma ^3}{4\beta _c} \int ^t_0 F(t')dt'\\&\quad \leqq \frac{1}{2}\int _{\Omega _0} \big | {{\mathcal {D}}}_t {\mathfrak {J}}-\nabla {{\mathcal {H}}}\big ( D_t P_{v,v} - v\cdot (\nabla P_{v,v}|_{c})\big )\big |_{t'=0}\big |^2{\text {d}}X\\&\qquad +\frac{\sigma }{2}\int _{\Gamma _{t0}}| \nabla _{\tau _t} {\mathfrak {J}}^\bot |^2\big |_{t'=0}{\text {d}}s\\&\qquad +\int ^t_0 P(E(t'))dt'+\int _{\Gamma _{t'}} R_1 ({\mathfrak {J}}- {\mathfrak {J}}|_c)\cdot n_t{\text {d}}s\Big |^t_0\\&\qquad +\int _{\Omega _{t'}} \nabla {{\mathcal {P}}}_{\mathfrak {J}, v}\cdot \nabla v\cdot \nabla \mathfrak {K}_{{{\mathcal {H}}}}{\text {d}}X\Big |^t_0\\&\quad \leqq P(E(0))+\int ^t_0 P(E(t'))dt'+\int _{\Gamma _{t'}} R_1 ({\mathfrak {J}}- {\mathfrak {J}}|_c)\cdot n_t{\text {d}}s\Big |^t_0\\&\qquad +\int _{\Omega _{t'}} \nabla {{\mathcal {P}}}_{\mathfrak {J}, v}\cdot \nabla v\cdot \nabla \mathfrak {K}_{{{\mathcal {H}}}}{\text {d}}X\Big |^t_0. \end{aligned} \end{aligned}$$

Replacing the first integral with \(\Vert {{\mathcal {D}}}_t {\mathfrak {J}}\Vert ^2_{L^2(\Omega _t)}\), we arrive at the inequality

$$\begin{aligned} \begin{aligned} E(t)+ \int ^t_0 F(t')dt'&\leqq P(E(0))+\int ^t_0 P(E(t'))dt'+C\int _{\Gamma _{t'}} R_1 ({\mathfrak {J}}- {\mathfrak {J}}|_c)\cdot n_t{\text {d}}s\Big |^t_0\\&\quad +C\int _{\Omega _{t'}} \nabla {{\mathcal {P}}}_{\mathfrak {J}, v}\cdot \nabla v\cdot \nabla \mathfrak {K}_{{{\mathcal {H}}}}{\text {d}}X\Big |^t_0 \\&\quad +C\int _{\Omega _{t}} \big | \nabla {{\mathcal {H}}}\big ( D_t P_{v,v} - v\cdot (\nabla P_{v,v}|_{c})\big )\big |^2{\text {d}}X, \end{aligned} \end{aligned}$$

where the constant C depends on \(\sigma , \beta _c\).

We now deal with the last three integrals on the right side above one by one. First, for the terms in \(R_1\), similarly as in (4.22), we can have by careful estimates with interpolations that

$$\begin{aligned} \Vert r^{\delta }( \partial ^2 v)\circ T^{-1}_i \Vert _{L^\infty (0, r_0)}\leqq P(E_l(t))E(t)^{1/2} \end{aligned}$$

and

$$\begin{aligned} \Vert r^{-\delta }({\mathfrak {J}}-{\mathfrak {J}}|_{p_i})\circ T^{-1}_i\Vert _{L^2(0, r_0)}\leqq & {} C(\Vert \Gamma _t\Vert _{H^{5/2}})\big \Vert {\mathfrak {J}}-{\mathfrak {J}}|_c\big \Vert _{H^\delta (\Gamma _t)}\\\leqq & {} C(\Vert \Gamma _t\Vert _{H^{5/2}})\Vert {\mathfrak {J}}\Vert _{H^{\delta +1/2}(\Omega _t)}, \end{aligned}$$

where \(1>\delta >3-\pi /\omega \) when \(\omega \in (\pi /3, \pi /2)\).

Therefore, we can prove by interpolations that there exist \(\delta _0\in (1/2, 1)\) and \(\varepsilon \) small enough such that

$$\begin{aligned} \int _{\Gamma _t} R_1 ({\mathfrak {J}}-{\mathfrak {J}}|_c)\cdot n_t{\text {d}}s\leqq P(E_l(t))E(t)^{\delta _0}\leqq \varepsilon E(t)+C_{\varepsilon , \delta _0} P(E_l(t)). \end{aligned}$$

Moreover, the following estimates can also be proved in a similar but easier way:

$$\begin{aligned}{} & {} \int _{\Omega _t} \nabla {{\mathcal {P}}}_{\mathfrak {J}, v}\cdot \nabla v\cdot \nabla \mathfrak {K}_{{{\mathcal {H}}}}{\text {d}}X +\int _{\Omega _t} \big | \nabla {{\mathcal {H}}}\big ( D_t P_{v,v} - v\cdot (\nabla P_{v,v}|_{c})\big )\big |^2{\text {d}}X \\{} & {} \quad \leqq \varepsilon E(t)+C_{\varepsilon , \delta _0} P(E_l(t)). \end{aligned}$$

As a result, summing up the estimates above, we are able to conclude that

$$\begin{aligned} E(t)+ \int ^t_0 F(t')dt'\leqq P(E(0))+\int ^t_0 P(E(t'))dt' +P(E_l(t)). \end{aligned}$$

In the end, to close the energy estimates, we deal with \(E_l(t)\). First, we show by direct calculations using Euler’s equation and lemmas from the previous section that

$$\begin{aligned} \Vert v\Vert ^2_{H^k(\Omega _t)}\leqq C(T) \Vert v(0)\Vert ^2_{H^k(\Omega _0)}+C(T)\Vert {\mathfrak {J}}\Vert ^2_{L^2([0,T],H^k(\Omega _t))}+\int ^t_0P(E(t'))dt' \end{aligned}$$

for \(k=1,2\). Then we take the square root on both sides of the inequalities above and apply an interpolation between \(H^1(\Omega _t), H^2(\Omega _t)\) as well as Lemma 4.1 to obtain

$$\begin{aligned} \Vert v\Vert _{H^{3/2}(\Omega _t)}\leqq C(T) \Vert v(0)\Vert _{H^{3/2}(\Omega _0)}+\int ^t_0P(E(t'))dt'. \end{aligned}$$

Second, the estimate for \(\Vert \Gamma _t\Vert _{H^{5/2}}\) can be derived in a similar way as above. Consequently, we have

$$\begin{aligned} E_l(t)\leqq P(E(0))+\int ^t_0P(E(t'))dt'. \end{aligned}$$

Combining this estimate with the estimate above for E(t), we finish the lower-order energy estimates.

5 Higher-Order Time-Derivative Energy Estimates

In this section, we prove higher-order energy estimates with respect to \(D_t\), which is needed in the local well-posedness part. To start with, we define the energy functional

$$\begin{aligned} E_1(t)=\Vert \nabla _{\tau _t} D_t \mathfrak {J}^\bot \Vert ^2_{L^2(\Gamma _t)} +\Vert D^2_t \mathfrak {J}\Vert ^2_{L^2(\Omega _t)}, \end{aligned}$$

and the dissipation

$$\begin{aligned} F_1(t)=\sum _{i=l,r}\big |(\sin \omega _i)\nabla _{\mathbf{\tau }_t}D_t \mathfrak {J}^\perp |_{p_i}\big |^2. \end{aligned}$$

The main result of this section is as follows:

Theorem 5.1

Let the contact angles \(\omega _i\in (0,\pi /2)\) and \(E(t), \int _0^T F(t)dt, E_1(t), \int _0^T F_1(t)dt\) be bounded above in [0, T] for some \(T>0\). Then the following higher-order a priori estimate holds

$$\begin{aligned} \sup _{0\leqq t\leqq T} E_1(t)+\int _0^T F_1(t)dt\leqq P(E_1(0))+\int _0^T P(E_1(t))dt. \end{aligned}$$

To prove this theorem, we begin with a higher-order equation for \({\mathfrak {J}}\) and more delicate estimates involving \(D_t\) based on Section 4.1. With these preparations, we are able to finish the energy estimate in the last subsection of this part.

5.1 The Higher-Order Equation for \({\mathfrak {J}}\)

Firstly, we recall system (3.12) and rewrite it as follows:

$$\begin{aligned} D^2_t \mathfrak {J}=\sigma \nabla {{\mathcal {H}}}(\Delta _{\Gamma _t} \mathfrak {J}^\perp -h_v)+\tilde{R}_0, \end{aligned}$$
(5.1)

where

$$\begin{aligned} \tilde{R}_0= & {} -\sigma \nabla {{\mathcal {H}}}(\mathfrak {J} \cdot \Delta _{\Gamma _t}n_t)+\sigma \nabla {{\mathcal {H}}}([n_t, \Delta _{\Gamma _t}]\cdot \nabla P_{v,v})\\{} & {} +\sigma \nabla {{\mathcal {H}}}(R_1+h_v ) +A_1+A_2+A_3+ \nabla {{\mathcal {H}}}( D^2_t P_{v,v} ), \end{aligned}$$

and \(h_v\) contains all the second-order terms of v in \(R_1\) which come from the commutator \([D_t, n_t\Delta _{\Gamma _t}]\cdot v\). More precisely, we have

$$\begin{aligned} h_v= \Delta _{\Gamma _t} v\cdot D_t n_t +2D^2v\big (\tau _t, (\nabla _{\tau _t}v)^\top \big )\cdot n_t. \end{aligned}$$
(5.2)

As a result, \(R_1+h_v\) only contains lower-order derivatives like \(\partial v, \partial n_t, \kappa \).

Acting \(D_t\) on both sides of (3.12), we obtain

$$\begin{aligned} D_t ({{\mathcal {D}}}^2_t \mathfrak {J}) +\nabla {{\mathcal {H}}}\big (\Delta _{\Gamma _t} D_t \mathfrak {J}^\bot -D_th_v\big )\nonumber \\ = \tilde{{\mathcal {R}}}_2, \end{aligned}$$
(5.3)

with the right side

$$\begin{aligned} \begin{aligned} \tilde{{\mathcal {R}}}_2&=D_t \tilde{R}_0 +[\nabla , D_t]{{\mathcal {H}}}\big (\Delta _{\Gamma _t} \mathfrak {J}^\bot -h_v)\big ) +\nabla [{{\mathcal {H}}}, D_t]\big (\Delta _{\Gamma _t} \mathfrak {J}^\bot -h_v\big )\\&\quad +\nabla {{\mathcal {H}}}[\Delta _{\Gamma _t}, D_t]\mathfrak {J}^\bot \\&\quad +D_t\big (D_t\nabla P_{\mathfrak {J}, v}+\nabla {{\mathcal {P}}}_{\nabla P_{\mathfrak {J}, v}, v}+\nabla P_{{{\mathcal {D}}}_t \mathfrak {J}, v}\big ). \end{aligned} \end{aligned}$$

Here we use

$$\begin{aligned} -[{{\mathcal {A}}}, D_t]=[\nabla , D_t]{{\mathcal {H}}}\Delta _{\Gamma _t}+\nabla [{{\mathcal {H}}}, D_t]\Delta _{\Gamma _t}+\nabla {{\mathcal {H}}}[\Delta _{\Gamma _t}, D_t] \end{aligned}$$

with all these commutators from the end of Section 2, and

$$\begin{aligned} {{\mathcal {D}}}^2_t \mathfrak {J}= D_t^2 \mathfrak {J}+D_t\nabla P_{\mathfrak {J}, v}+\nabla P_{\nabla P_{\mathfrak {J}, v}, v}+\nabla P_{{{\mathcal {D}}}_t \mathfrak {J}, v}, \end{aligned}$$
(5.4)

where \(P_{w, v}\) is defined in (1.10).

5.2 More Preliminary Estimates.

Based on Section 4.1, we are going to prove higher-order estimates for different quantities using our higher-order energy.

- Higher-order estimates for \(\mathfrak {K}_{{{\mathcal {H}}}}\) and \(D_t{\mathfrak {J}}\). To get started, we recall system (4.1) to find

$$\begin{aligned} \left\{ \begin{array}{ll} \Delta D_t\mathfrak {K}_{{\mathcal {H}}}=2tr(\nabla v \nabla ^2 \mathfrak {K}_{{\mathcal {H}}})\qquad \hbox {in}\quad \Omega _t\\ \nabla _{n_t} D_t \mathfrak {K}_{{\mathcal {H}}}|_{\Gamma _t} = D_t {\mathfrak {J}}^\perp +\nabla _{n_t} v \cdot \nabla \mathfrak {K}_{{\mathcal {H}}}|_{\Gamma _t},\qquad \nabla _{n_b}D_t\mathfrak {K}_{{\mathcal {H}}}|_{\Gamma _b}=\nabla _{n_b}v\cdot \nabla {\mathfrak {K}}_{{\mathcal {H}}}|_{\Gamma _b}. \end{array}\right. \end{aligned}$$

Applying Proposition 2.2 and lemmas in Section 4.1, we have

$$\begin{aligned} \begin{aligned} \Vert D_t \mathfrak {K}_{{\mathcal {H}}}\Vert _{H^2(\Omega _t)}&\leqq C(\Vert \Gamma _t\Vert _{H^{5/2}})\big (\Vert 2tr(\nabla v \nabla ^2 \mathfrak {K}_{{\mathcal {H}}})\Vert _{L^2(\Omega _t)}+\Vert D_t {\mathfrak {J}}^\perp \\&\quad +\nabla _{n_t} v \cdot \nabla \mathfrak {K}_{{\mathcal {H}}}\Vert _{H^{1/2}(\Gamma _t)}\\&\quad +\Vert \nabla _{n_b}v\cdot \nabla {\mathfrak {K}}_{{\mathcal {H}}}\Vert _{H^{1/2}(\Gamma _b)}+\Vert D_t{\mathfrak {K}}_{{\mathcal {H}}}\Vert _{L^2(\Omega _t)}\big )\\&\leqq P(E(t))(1+ \Vert D_t \mathfrak {J}^\bot \Vert _{H^{1/2}(\Gamma _t)}). \end{aligned} \end{aligned}$$
(5.5)

Here we notice that the regularity of \(D_t\mathfrak {K}_{{\mathcal {H}}}\) is constrained due to the regularity of v. To get a higher-order estimate, we need to get rid of the worst part.

Similarly as in Lemma 4.7, we consider a good quantity \(D_t\mathfrak {K}_{{\mathcal {H}}}-v\cdot (\nabla \mathfrak {K}_{{\mathcal {H}}}|_{c})\) which satisfies

$$\begin{aligned} \left\{ \begin{array}{ll} \Delta \big (D_t\mathfrak {K}_{{\mathcal {H}}}-v\cdot (\nabla \mathfrak {K}_{{\mathcal {H}}}|_{c})\big )=2tr(\nabla v \nabla ^2 \mathfrak {K}_{{\mathcal {H}}})-\Delta \big (v\cdot (\nabla \mathfrak {K}_{{\mathcal {H}}}|_{c})\big ) \qquad \hbox {in}\quad \Omega _t\\ \nabla _{n_t}(D_t\mathfrak {K}_{{\mathcal {H}}}-v\cdot (\nabla \mathfrak {K}_{{\mathcal {H}}}|_{c}))\big |_{\Gamma _t}=D_t {\mathfrak {J}}^\perp +\nabla _{n_t} v \cdot \big (\nabla \mathfrak {K}_{{\mathcal {H}}}-\nabla \mathfrak {K}_{{\mathcal {H}}}|_{c}\big )-v\cdot \nabla _{n_t}(\nabla \mathfrak {K}_{{\mathcal {H}}}|_{c})\big |_{\Gamma _t},\\ \nabla _{n_b}(D_t\mathfrak {K}_{{\mathcal {H}}}-v\cdot (\nabla \mathfrak {K}_{{\mathcal {H}}}|_{c}))\big |_{\Gamma _b}=\nabla _{n_b}v\cdot \big (\nabla {\mathfrak {K}}_{{\mathcal {H}}}-\nabla \mathfrak {K}_{{\mathcal {H}}}|_{c}\big ) -v\cdot \nabla _{n_b}(\nabla \mathfrak {K}_{{\mathcal {H}}}|_{c})\big |_{\Gamma _b}. \end{array}\right. \end{aligned}$$

As a result, similar arguments as in (4.22) applied for the part \(\nabla v\cdot \big (\nabla {\mathfrak {K}}_{{\mathcal {H}}}-\nabla \mathfrak {K}_{{\mathcal {H}}}|_{c}\big )\), we can have

$$\begin{aligned} \Vert D_t\mathfrak {K}_{{\mathcal {H}}}-v\cdot (\nabla \mathfrak {K}_{{\mathcal {H}}}|_{c})\Vert _{H^{5/2}(\Omega _t)} \leqq P(E(t))(1+E_1(t)^{1/2}). \end{aligned}$$
(5.6)

where \(\Vert D_t{\mathfrak {J}}^\perp \Vert _{L^2(\Gamma _t)}\) is handled by (5.5).

On the other hand, we deal with \(D_t{\mathfrak {J}}\). First, we know directly from (5.5) that

$$\begin{aligned} \Vert D_t{\mathfrak {J}}\Vert _{H^1(\Omega _t)}\leqq P(E(t))(1+ \Vert D_t \mathfrak {J}^\bot \Vert _{H^{1/2}(\Gamma _t)}). \end{aligned}$$
(5.7)

Second, we notice that

$$\begin{aligned} D_t{\mathfrak {J}}=D_t\nabla \mathfrak {K}_{{\mathcal {H}}}=\nabla \big (D_t\mathfrak {K}_{{\mathcal {H}}}-v\cdot ({\mathfrak {J}}|_{c})\big )-\nabla v\cdot \big ({\mathfrak {J}}-{\mathfrak {J}}|_{c}\big ). \end{aligned}$$

Similarly as in (4.14), we apply Lemma 2.2(1) and (4.11) to obtain the estimate

$$\begin{aligned} \begin{aligned}&\Vert \nabla v\cdot ({\mathfrak {J}}-{\mathfrak {J}}|_{c})\Vert _{H^{3/2}(\Omega _t)}\\&\quad \leqq P(E(t))\big (1+\Vert (\partial \nabla v)\cdot ({\mathfrak {J}}-{\mathfrak {J}}|_{c})\Vert _{H^{1/2}(\Omega _t)}\big )\\&\quad \leqq P(E(t)\Big (1+\sum _{i=l,r}\big (\Vert r^{\pi /\omega _i-2}\Vert _{L^\infty (0,r_0)}\\&\qquad +\Vert r^{\pi /\omega _i-2}\Vert _{H^1(0,r_0)}\big ) \Vert r^{-1}({\mathfrak {J}}-{\mathfrak {J}}|_{p_i})\circ T^{-1}_i\Vert _{H^{1/2}({{\mathcal {S}}}_{t,i})}\Big )\\&\quad \leqq P(E(t)), \end{aligned} \end{aligned}$$

while recall that \({{\mathcal {S}}}_{t,i}\) are straightened sector of \(\Omega _t\) with radius \(r_0\) near the corners.

Therefore, combining this estimate with (5.6), we have

$$\begin{aligned} \Vert D_t{\mathfrak {J}}\Vert _{H^{3/2}(\Omega _t)} \leqq P(E(t))(1+E_1(t)^{1/2}) \end{aligned}$$
(5.8)

and also

$$\begin{aligned} \Vert D_t{\mathfrak {K}}_{{\mathcal {H}}}\Vert _{H^2(\Omega _t)}\leqq P(E(t))(1+E_1(t)^{1/2}). \end{aligned}$$

- Estimates for \(D_tP_{v,v}\). Checking system (4.7) carefully and applying lemmas from Section 4.1, one has immediately

$$\begin{aligned} \Vert D_tP_{v,v}\Vert _{H^{5/2}(\Omega _t)}\leqq P(E(t)), \end{aligned}$$
(5.9)

where we use arguments similar to (4.13) on boundary terms.

- Estimates for v. To begin with, applying Remark 2.3, we obtain for a small number \(\epsilon \in (0, \pi /\omega _i-2)\) (when \(\omega _i\in (\pi /3, \pi /2)\)) that

$$\begin{aligned} \begin{aligned} \Vert v\Vert _{H^{2+\epsilon }(\Omega _t)}\leqq P(E(t)). \end{aligned} \end{aligned}$$
(5.10)

Moreover, we improve the regularity of \(v^\perp \). In fact, similarly as in (4.5), we have thanks to (5.5),(5.9) and (5.10) the estimate

$$\begin{aligned} \begin{aligned} \Vert \Delta _{\Gamma _t}v^\perp \Vert _{H^{1/2+\epsilon }(\Gamma _t)}&\leqq C(\Vert \Gamma _t\Vert _{H^{4}})\big (\Vert D_t\kappa \Vert _{H^{1/2+\epsilon }(\Gamma _t)} + \Vert v\Vert _{H^{1/2+\epsilon }(\Omega _t)}\big )\\&\leqq P(E(t))(1+E_1(t)^{1/2}). \end{aligned} \end{aligned}$$
(5.11)

- Estimates for \(D^2_tv, D^3_tv\). First, acting \(D_t\) on both sides of Euler’s equation leads to

$$\begin{aligned} D_t^2 v =-D_t \mathfrak {J}-D_t\nabla P_{v,v}. \end{aligned}$$

Applying Lemma 4.2, Lemma 4.6, (5.7), and (5.8), we find

$$\begin{aligned} \Vert D_t^2v\Vert _{H^{3/2}(\Omega _t)}\leqq & {} \Vert D_t \mathfrak {J}\Vert _{H^{3/2}(\Omega _t)}+\Vert D_t\nabla P_{v,v}\Vert _{H^{3/2}(\Omega _t)}\nonumber \\\leqq & {} P(E(t))(1+E_1(t)^{1/2}). \end{aligned}$$
(5.12)

Next, we give the estimates of \(D_t^3 v\). Before that, we first prove the following lemma:

Lemma 5.1

Let \(P_{\mathfrak {J}, v}, P_{\nabla P_{\mathfrak {J}, v}, v}\) and \(P_{{{\mathcal {D}}}_t \mathfrak {J}, v}\) be defined by (1.10). Then there hold

$$\begin{aligned} \Vert D_t\nabla P_{\mathfrak {J}, v}+\nabla P_{\nabla P_{\mathfrak {J}, v}, v}+\nabla P_{{{\mathcal {D}}}_t \mathfrak {J}, v}\Vert _{H^1(\Omega _t)} \leqq P(E(t))\big (1+ E_1(t)^{1/2}\big ) \end{aligned}$$

and

$$\begin{aligned} \Vert D^2_t\nabla P_{v,v}\Vert _{L^2(\Omega _t)} \leqq P(E(t))\big (1+ E_1(t)^{1/2}\big ). \end{aligned}$$

Proof

First, we deal with \(D_t P_{\mathfrak {J}, v}\). Recalling system (4.15), we have by Proposition 2.2, (4.17), (5.5), (5.8) and lemmas from Section 4.1 that

$$\begin{aligned} \begin{aligned}&\Vert D_t P_{\mathfrak {J}, v}\Vert _{H^2(\Omega _t)}\\&\quad \leqq C(\Vert \Gamma _t\Vert _{H^{5/2}})\Big (\Vert - D_ttr(\nabla \mathfrak {J}\nabla v)+2tr(\nabla v \nabla ^2 {{\mathcal {P}}}_{ \mathfrak {J},v})\Vert _{L^2(\Omega _t)} + \Vert D_t( {\mathfrak {J}}\cdot \nabla _v n_b)\\&\qquad +\nabla _{n_b}v\cdot \nabla P_{ \mathfrak {J},v}\Vert _{H^{1/2}(\Gamma _b)}\\&\qquad +\Vert C'_{ \mathfrak {J}, v}-D_t(\nabla _{\tau _t}\mathfrak {K}_{{{\mathcal {H}}}}\cdot \nabla _{\tau _t} v \cdot n_t)+\nabla _{n_t}v\cdot \nabla P_{ \mathfrak {J},v}\Vert _{H^{1/2}(\Gamma _t)} +\Vert D_t P_{\mathfrak {J}, v}\Vert _{L^2(\Omega _t)}\Big )\\&\quad \leqq P(E(t))(1+E_1(t)^{1/2}). \end{aligned} \end{aligned}$$

Second, checking the definition (1.10) and applying Lemma 2.2 (1) and Lemma 4.6, we can have for \( P_{\nabla P_{\mathfrak {J}, v}, v}\) the following estimate immediately:

$$\begin{aligned} \Vert P_{\nabla P_{\mathfrak {J}, v}, v}\Vert _{H^{5/2}(\Omega _t)} \leqq P(E(t)). \end{aligned}$$
(5.13)

Moreover, a similar argument also leads to

$$\begin{aligned} \Vert P_{{{\mathcal {D}}}_t \mathfrak {J}, v}\Vert _{H^2(\Omega _t)} \leqq P(E(t))\big (1+ E_1(t)^{1/2}\big ). \end{aligned}$$

In the end, for the part \(D_t^2 P_{v,v}\), we deduce from (4.7) the following system:

$$\begin{aligned} \left\{ \begin{array}{ll} \Delta (D^2_t P_{v, v} )=-tr D^2_t (\nabla v \nabla v) +2trD_t (\nabla v \nabla ^2 P_{v, v})\qquad \hbox {in}\quad \Omega _t\\ \nabla _{n_t}( D^2_t P_{v, v} )|_{\Gamma _t}=C''_{v,v}(t)+D_t(\nabla _{n_t} v\cdot \nabla P_{v, v} )+\nabla _{n_t} v\cdot \nabla D_t P_{v, v}\big |_{\Gamma _t},\\ \nabla _{n_b} ( D^2_t P_{v, v}) |_{\Gamma _b}=D^2_t(v\cdot \nabla _v n_b)+D_t(\nabla _{n_b} v\cdot \nabla P_{v, v} )+\nabla _{n_b} v\cdot \nabla D_t P_{v, v} \big |_{\Gamma _b} \end{array} \right. \end{aligned}$$
(5.14)

with

$$\begin{aligned} \int _{\Omega _t}D^2_tP_{v,v}{\text {d}}X=0. \end{aligned}$$

Applying (5.16), (5.12) and checking term by term, we obtain the variational estimate

$$\begin{aligned} \Vert D^2_t P_{v, v} \Vert _{H^1(\Omega _t)} \leqq P(E(t))(1+E_1(t)^{1/2}), \end{aligned}$$
(5.15)

and the proof is finished. \(\quad \square \)

Now, we are in a position to give the estimate for \(D_t^3 v\). In fact, acting \(D_t^2\) on both sides of Euler’s equation, we derive by the previous lemma and lemmas from Section 4.1 that

$$\begin{aligned} \Vert D_t^3v\Vert _{L^2(\Omega _t)} \leqq \Vert D^2_t \mathfrak {J}\Vert _{L^2(\Omega _t)} +\Vert D^2_t\nabla P_{v,v}\Vert _{L^2(\Omega _t)} \leqq P(E(t))(1+E_1(t)^{1/2}). \nonumber \\\end{aligned}$$
(5.16)

- Estimate for \(v^\perp \). We obtain by the definition of \(\mathfrak {J}\) that

$$\begin{aligned} \nabla (D_t^2 \mathfrak {K}_{{\mathcal {H}}}) =D_t^2 \mathfrak {J}+ \nabla v\cdot \nabla D_t \mathfrak {K}_{{\mathcal {H}}}+ D_t(\nabla v\cdot \nabla \mathfrak {K}_{{\mathcal {H}}}), \end{aligned}$$

and applying (3.5) and (5.5) implies

$$\begin{aligned} \Vert D_t^2 \mathfrak {K}_{{\mathcal {H}}}\Vert _{H^1(\Omega _t)} \leqq P(E(t))(1+E_1(t)^{1/2}). \end{aligned}$$
(5.17)

Recalling (3.5) again, we know

$$\begin{aligned} \sigma D_t\kappa =D_t (\mathfrak {K}+P_{v,v})=-\Delta _{\Gamma _t}v^\perp -v^\perp |\nabla _{\tau _t}n_t|^2+\nabla _{\tau _t}\nabla _{v^\top }n_t\cdot \tau _t \qquad \hbox {on}\quad \Gamma _t, \end{aligned}$$

so we obtain by (5.15) and (5.17) the estimate

$$\begin{aligned} \Vert D_t \Delta _{\Gamma _t}v^\perp \Vert _{L^2(\Gamma _t)} \leqq P(E(t))(1+E_1(t)^{1/2}). \end{aligned}$$
(5.18)

- Some more higher-order estimates for \({\mathfrak {K}}_{{\mathcal {H}}}\), v and \({\mathfrak {J}}\).

Lemma 5.2

Assuming that \(E(t), E_1(t)\in L^\infty [0,T]\) for some \(T>0\), we have the following estimate:

$$\begin{aligned}{} & {} \Vert \mathfrak {K}_{{\mathcal {H}}}\Vert _{H^3(\Omega _t)}+\Vert D_t v\Vert _{H^2(\Omega _t)} +\Vert \nabla ^2 \mathfrak {K}_{{\mathcal {H}}}\Vert _{L^\infty (\Omega _t)} +\Vert D_t \nabla v\Vert _{L^\infty (\Omega _t)}\\{} & {} \quad \leqq P(E(t))(1+E_1(t)^{1/2}). \end{aligned}$$

Proof

(1) Higher regularity for \({\mathfrak {K}}_{{\mathcal {H}}}\). To begin with, we use Euler’s equation to rewrite (3.7) as follows:

$$\begin{aligned} D^2_t{\mathfrak {K}}= & {} \big (\Delta _{\Gamma _t}J^\perp -h_v\big )+[n_t,\Delta _{\Gamma _t}](J+\nabla P_{v,v}+\textbf{g})+2\sigma \Pi (\tau _t)\cdot \nabla _{\tau _t} J\\{} & {} +(R_1+h_v)- D_t^2 P_{v,v}\quad \hbox {on}\ \Gamma _t \end{aligned}$$

where we notice that \(\Delta _{\Gamma _t}(\nabla _{n_t} P_{v,v})|_{\Gamma _t}=0\) thanks to the definition of \(P_{v,v}\) and \((R_1+h_v)\) contains only \(\partial v\) terms instead of \(\partial ^2 v\) terms.

Meanwhile, we know from (5.15) and (5.17) that

$$\begin{aligned} \Vert D^2_t{\mathfrak {K}}\Vert _{H^{1/2}(\Gamma _t)}+\Vert D_t^2 P_{v,v}\Vert _{H^{1/2}(\Gamma _t)}\leqq P(E(t))(1+E_1(t)^{1/2}), \end{aligned}$$

so checking term by term in the equation above, we have immediately

$$\begin{aligned} \Vert \Delta _{\Gamma _t}J^\perp -h_v\Vert _{H^{1/2}(\Gamma _t)}\leqq P(E(t))(1+E_1(t)^{1/2}). \end{aligned}$$
(5.19)

On the other hand, we know from (4.11) that (when \(\omega _i\in (\pi /3, \pi /2)\))

$$\begin{aligned} h_v= & {} h_{v,r}+h_{v,s}\quad \hbox {with}\quad h_{v,r}\in H^{1/2}(\Gamma _t),\ h_s=\sum _i\big [\chi _{\omega } a_{i,1}r^{2\alpha _i-3}\\{} & {} +\chi _{\omega }a_{i,2}(\partial v_r\circ T^{-1}_i)r^{\alpha _i-2}\big ]\circ T_i, \end{aligned}$$

where we note \(\alpha _i=\pi /\omega _i-1\in (1,2)\), \(a_{i,k}\) (\(i=l, r\)) contain \(n_t, \tau _t, \partial n_t, \partial \tau _t\) and the singular coefficient from (4.11). Moreover, \(a_{i,3}(\partial v_r\circ T^{-1}_i)\) is linear with respect to \(\partial v_r\circ T^{-1}_i\), where we recall that \(v_r=\nabla \phi _r\in H^3(\Omega _t)\) and \(\Vert v_r\Vert _{H^3(\Omega _t)}\) is controlled by P(E(t)).

As a result, we find

$$\begin{aligned} 2\alpha _i-3\in (-1,1),\quad \alpha _i-2\in (-1,0), \end{aligned}$$

which implies immediately that

$$\begin{aligned} \Vert h_v\Vert _{L^p(\Gamma _t)}\leqq P(E(t)),\qquad \hbox {for}\quad 1<p<\min \{(2-\alpha _i)^{-1}, |3-2\alpha _i|^{-1}\}. \end{aligned}$$

Summing up the estimates above for \(\Vert \Delta _{\Gamma _t}J^\perp -h_v\Vert _{H^{1/2}(\Gamma _t)}\) and \(\Vert h_v\Vert _{L^p(\Gamma _t)}\), we obtain

$$\begin{aligned} \begin{aligned} \Vert \Delta _{\Gamma _t}J^\perp \Vert _{L^p(\Gamma _t)}&\leqq \Vert \Delta _{\Gamma _t}J^\perp -h_v\Vert _{L^p(\Gamma _t)}+\Vert h_v\Vert _{L^p(\Gamma _t)}\\&\leqq C(\Vert \Gamma _t\Vert _{H^{5/2}})\Vert \Delta _{\Gamma _t}J^\perp -h_v\Vert _{H^{1/2}(\Gamma _t)}+\Vert h_v\Vert _{L^p(\Gamma _t)}\\&\leqq P(E(t))(1+E_1(t)^{1/2}), \end{aligned} \end{aligned}$$

and this leads to

$$\begin{aligned} \Vert \nabla _{\tau _t}J^\perp \Vert _{W^{1,p}(\Gamma _t)}\leqq P(E(t))(1+E_1(t)^{1/2}). \end{aligned}$$

Applying Lemma 2.1, we finally show that

$$\begin{aligned} \Vert \nabla _{\tau _t}J^\perp \Vert _{L^\infty (\Gamma _t)}+\Vert J^\perp \Vert _{H^{3/2+\varepsilon }(\Gamma _t)}\leqq P(E(t))(1+E_1(t)^{1/2}) \end{aligned}$$
(5.20)

with \(\varepsilon =1-1/p\).

Consequently, applying Proposition 2.2, we have the desired estimate for \(\Vert {\mathfrak {K}}_{{\mathcal {H}}}\Vert _{H^{3}(\Omega _t)}\). Notice that we have in fact the estimate by Remark 2.3:

$$\begin{aligned} \Vert {\mathfrak {K}}_{{\mathcal {H}}}\Vert _{H^{3+\varepsilon }(\Omega _t)}\leqq P(E(t))(1+E_1(t)^{1/2}). \end{aligned}$$
(5.21)

Moreover, we apply Lemma 2.4 to have \({\mathfrak {K}}=\sigma \kappa -P_{v,v}\in H^{5/2+\varepsilon }(\Gamma _t)\) with the estimate

$$\begin{aligned} \Vert {\mathfrak {K}}\Vert _{H^{5/2+\varepsilon }(\Gamma _t)}\leqq P(E(t))(1+E_1(t)^{1/2}). \end{aligned}$$

(2) \(H^2\) estimate for \(D_tv\). Recalling the Euler’s equations and \(P_{v,v}\) estimate from Lemma 4.6, we get

$$\begin{aligned} \Vert D_t v\Vert _{H^2(\Omega _t)}\leqq \Vert {\mathfrak {J}}\Vert _{H^2(\Omega _t)}+\Vert \nabla P_{v,v}\Vert _{H^2(\Omega _t)}+\Vert \textbf{g}\Vert _{H^2(\Omega _t)} \leqq P(E(t))(1+E_1(t)^{1/2}). \end{aligned}$$

(3) \(L^\infty \) estimates. Applying Remark 2.3 and (5.21) with the same \(\varepsilon =1-1/p\) as above lead immediately to

$$\begin{aligned} \begin{aligned} \Vert \nabla ^2 \mathfrak {K}_{{\mathcal {H}}}\Vert _{L^\infty }&\leqq C(\Vert \Gamma _t\Vert _{H^{5/2}})\Vert \nabla ^2{\mathfrak {K}}_{{\mathcal {H}}}\Vert _{H^{1+\varepsilon }(\Omega _t)} \leqq P(E(t))(1+E_1(t)^{1/2}). \end{aligned} \end{aligned}$$

Applying Remark 2.3 to \(P_{v,v}\), we obtain

$$\begin{aligned} \begin{aligned}&\Vert \nabla ^2 P_{v,v}\Vert _{L^\infty (\Omega _t)}\leqq C(\Vert \Gamma _t\Vert _{H^{5/2}})\Vert \nabla ^2P_{v,v}\Vert _{H^{1+\epsilon }(\Omega _t)}\\&\quad \leqq C(\Vert \Gamma _t\Vert _{H^4})\big (\Vert tr(\nabla v\nabla v)\Vert _{H^{1+\varepsilon }(\Omega _t)}+\Vert C_{v,v}(t)\Vert _{H^{3/2+\epsilon }(\Gamma _t)}\\&\qquad +\Vert v\cdot \nabla _vn_b\Vert _{H^{1/2+\varepsilon }(\Gamma _b)} +\Vert P_{v,v}\Vert _{L^2(\Omega _t)}\big )\\&\quad \leqq P(E(t))\big (1+\Vert \nabla v\Vert ^2_{H^{1+\varepsilon }(\Omega _t)}\big ). \end{aligned} \end{aligned}$$

Using (4.11) and a similar argument as in (1), we have

$$\begin{aligned} \Vert \nabla v\Vert _{H^{1+\varepsilon }(\Omega _t)}\leqq \Vert \nabla ^2\phi _s\Vert _{H^{1+\varepsilon }(\Omega _t)}+\Vert \nabla ^2\phi _r\Vert _{H^2(\Omega _t)} \leqq P(E(t)), \end{aligned}$$

which implies

$$\begin{aligned} \Vert \nabla ^2 P_{v,v}\Vert _{L^\infty (\Omega _t)}\leqq P(E(t)). \end{aligned}$$

In the end, apply Euler’s equation again leads to

$$\begin{aligned} \Vert \nabla D_t v\Vert _{L^\infty (\Omega _t)}\leqq \Vert \nabla ^2{\mathfrak {K}}_{{\mathcal {H}}}\Vert _{L^\infty (\Omega _t)}+\Vert \nabla ^2 P_{v,v}\Vert _{L^\infty (\Omega _t)}\leqq P(E(t))(1+E_1(t)^{1/2}), \end{aligned}$$

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

- Estimate for \(D^2_tP_{v,v}\). Based the above estimates, we firstly improve the estimate for \(D_tP_{v,v}\). Using system (4.12) and Lemma 5.2, we improve the estimate in Lemma 4.7 and get

$$\begin{aligned} \begin{aligned}&\Vert D_t P_{v, v}-v\cdot (\nabla P_{v,v}|_{c}) \Vert _{H^3(\Omega _t)} \\&\quad \leqq C(\Vert \Gamma _t\Vert _{H^3})\Big (\Vert -tr D_t (\nabla v \nabla v) +[\Delta ,v]\cdot (\nabla P_{v, v}\\&\qquad - \nabla P_{v,v}|_{c})+v\cdot \Delta (\nabla P_{v,v}|_c) \Vert _{H^1(\Omega _t)}+| C'_{v,v}(t)|\\&\qquad +\Vert \nabla _{n_t} v\cdot (\nabla P_{v, v}-\nabla P_{v,v}|_{c})-v\cdot \nabla _{n_t}(\nabla P_{v,v}|_c)\Vert _{H^{3/2}(\Gamma _t)}\\&\qquad +\Vert D_t(v\cdot \nabla _v n_b)\Vert _{H^{3/2}(\Gamma _b)}\\&\qquad +\Vert \nabla _{n_b} v\cdot (\nabla P_{v, v}-\nabla P_{v,v}|_{c})-v\cdot \nabla _{n_b}(\nabla P_{v,v}|_c)\Vert _{H^{3/2}(\Gamma _b)}\Big )\\&\quad \leqq P(E(t))(1+E_1(t)^{1/2}). \end{aligned} \end{aligned}$$
(5.22)

Here we use Lemma 2.2 (2) for the boundary terms like \(\Vert \nabla _{n_t} v\cdot (\nabla P_{v, v}-\nabla P_{v,v}|_{c})\Vert _{H^{3/2}(\Gamma _t)}\), which are handled similarly as in the proof for (5.8).

Next, we derive the equation of \(D_t(D_t P_{v, v}-v\cdot (\nabla P_{v,v}|_{c}))\). To simplify the notation, we define

$$\begin{aligned} {\mathcal {P}}_{t,1}=D_t P_{v, v}-v\cdot (\nabla P_{v,v}|_{c}), \end{aligned}$$

and we rewrite (5.22) as

$$\begin{aligned} \Vert {\mathcal {P}}_{t,1} \Vert _{H^3(\Omega _t)}\leqq P(E(t))(1+E_1(t)^{1/2}). \end{aligned}$$
(5.23)

Direct computations lead to the following system for \(D_t{\mathcal {P}}_{t,1}\):

$$\begin{aligned} \left\{ \begin{array}{ll} &{}\Delta (D_t{\mathcal {P}}_{t,1})=-tr D^2_t (\nabla v \nabla v) +2trD_t\big (\nabla v\nabla (\nabla P_{v, v}- \nabla P_{v,v}|_{c}) \big )\\ &{}\quad -D_t\big (v\cdot \Delta (\nabla P_{v,v}|_c)\big )+[D_t, \Delta ]{\mathcal {P}}_{t,1}\qquad \hbox {in}\quad \Omega _t\\ &{}\nabla _{n_t}( D_t {\mathcal {P}}_{t,1})\big |_{\Gamma _t}=C''_{v,v}(t)+D_t(\nabla _{n_t} v\cdot \big (\nabla P_{v, v}-\nabla P_{v,v}|_{c})\big )\\ &{}\quad -D_t\big (v\cdot \nabla _{n_t}(\nabla P_{v,v}|_c)\big )+\nabla _{n_t} v\cdot \nabla {\mathcal {P}}_{t,1}\big |_{\Gamma _t},\\ &{}\nabla _{n_b}( D_t {\mathcal {P}}_{t,1})\big |_{\Gamma _b}= D^2_t(v\cdot \nabla _v n_b)+D_t\big (\nabla _{n_b} v\cdot (\nabla P_{v, v}-\nabla P_{v,v}|_{c})\big )\\ &{}\quad -D_t\big (v\cdot \nabla _{n_b}(\nabla P_{v,v}|_c)\big )+\nabla _{n_b} v\cdot \nabla {\mathcal {P}}_{t,1}\big |_{\Gamma _b}.\\ \end{array} \right. \end{aligned}$$
(5.24)

Moreover, we define

$$\begin{aligned} {\mathcal {P}}_{t,2}=D_t {\mathcal {P}}_{t,1}-v\cdot (\nabla {\mathcal {P}}_{t,1}|_{c})= D^2_tP_{v,v}-D_t\big (v\cdot (\nabla P_{v,v}|_c)\big )-v\cdot (\nabla {\mathcal {P}}_{t,1}|_{c}) \nonumber \\\end{aligned}$$
(5.25)

and we modify this system above as in Lemma 4.7 into a new system for \({\mathcal {P}}_{t,2}\) as follows:

$$\begin{aligned} \left\{ \begin{array}{ll} \Delta {\mathcal {P}}_{t,2}=-tr D^2_t (\nabla v \nabla v) +2trD_t\big (\nabla v\nabla (\nabla P_{v, v}- \nabla P_{v,v}|_{c}) \big )-D_t\big (v\cdot \Delta (\nabla P_{v,v}|_c)\big )\\ \qquad \qquad +[D_t, \Delta ]{\mathcal {P}}_{t,1}-v\cdot \Delta (\nabla {\mathcal {P}}_{t,1}|_c)\qquad \hbox {in}\quad \Omega _t\\ \nabla _{n_t}{\mathcal {P}}_{t,2}\big |_{\Gamma _t}=C''_{v,v}(t)+D_t(\nabla _{n_t} v\cdot \big (\nabla P_{v, v}-\nabla P_{v,v}|_{c})\big )-D_t\big (v\cdot \nabla _{n_t}(\nabla P_{v,v}|_c)\big )\\ \qquad \qquad \qquad +\nabla _{n_t} v\cdot (\nabla {\mathcal {P}}_{t,1}-\nabla {\mathcal {P}}_{t,1}|_c)-v\cdot \nabla _{n_t}(\nabla {\mathcal {P}}_{t,1}|_c)\big |_{\Gamma _t},\\ \nabla _{n_b}{\mathcal {P}}_{t,2}\big |_{\Gamma _b}= D^2_t(v\cdot \nabla _v n_b)+D_t\big (\nabla _{n_b} v\cdot (\nabla P_{v, v}-\nabla P_{v,v}|_{c})\big )-D_t\big (v\cdot \nabla _{n_b}(\nabla P_{v,v}|_c)\big )\\ \qquad \qquad \qquad +\nabla _{n_b} v\cdot (\nabla {\mathcal {P}}_{t,1}-\nabla {\mathcal {P}}_{t,1}|_c)-v\cdot \nabla _{n_b}(\nabla {\mathcal {P}}_{t,1}|_c)\big |_{\Gamma _b}.\\ \end{array} \right. \end{aligned}$$
(5.26)

Thanks to Lemma 4.7, (5.12) and (5.23), it is straightforward to show that

$$\begin{aligned} \Vert {\mathcal {P}}_{t,2}\Vert _{H^{5/2}(\Omega _t)}\leqq P(E(t))(1+E_1(t)^{1/2}). \end{aligned}$$
(5.27)

- The boundary condition for \(D_t^2 \mathfrak {J}\) at corner points.

Lemma 5.3

We have at the contact points \(p_i (i=l,r)\) the equations

$$\begin{aligned} D^2_t \mathfrak {J} =\frac{\sigma ^2 }{\beta _c}(\sin \omega _i) D_t(\nabla _{\tau _t} \mathfrak {J}^\bot ) \tau _b+R_{c, h, i}, \end{aligned}$$
(5.28)

where the remainder terms

$$\begin{aligned} R_{c, h, i}=D_tR_{c1} + \frac{\sigma ^2 }{\beta _c}D_t(\tau _b\sin \omega _i ) (\nabla _{\tau _t} \mathfrak {J})^\bot - \frac{\sigma ^2 }{\beta _c}(\sin \omega _i ) D_t(\nabla _{\tau _t} n_t\cdot \mathfrak {J})\tau _b\big |_{p_i} \end{aligned}$$

with \(R_{c1}\) defined in Lemma 4.9. Moreover, if holds for \(i=l, r\) that

$$\begin{aligned} | R_{c, h,i}|\leqq P(E(t))\big (1+E_1(t)^{1/2}\big ). \end{aligned}$$

Proof

First, using \(D_t\) directly on both sides of the equations in Lemma 4.9, one has the desired equations immediately. Next, apply lemmas from Section 4.1, Lemma 5.2 (5.21) and (5.27), the estimates follow. Moreover, we notice that the higher-order term \(\nabla ^2_{\tau _t}v^\perp (\tau _t\cdot {\mathfrak {J}})\) from \( D_t(\nabla _{\tau _t} n_t\cdot \mathfrak {J})\tau _b\) in \(R_{c,h,i}\) can be controlled thanks to (5.11), where we use

$$\begin{aligned} D_t n_t=-\big ((\nabla v)^*n_t\big )^ \top =-\big (n_t\cdot \nabla _{\tau _t}v\big )\tau _t=-(\nabla _{\tau _t}v^\perp )\tau _t+(\nabla _{\tau _t}n_t\cdot v)\tau _t. \nonumber \\\end{aligned}$$
(5.29)

\(\square \)

5.3 Proof of Theorem 5.1

At this moment, we are finally ready to prove the higher-order energy estimate in Theorem 5.1. To begin with, we rewrite equation (5.3) of \({{\mathcal {D}}}_t^2 \mathfrak {J}\) into

$$\begin{aligned}{} & {} D_t \big ({{\mathcal {D}}}^2_t \mathfrak {J} -\nabla {{\mathcal {H}}}({{\mathcal {P}}}_{t,2}) \big )+\nabla {{\mathcal {H}}}\big (\Delta _{\Gamma _t} D_t \mathfrak {J}^\bot -D_t h_v\big )\nonumber \\{} & {} \quad = {{\mathcal {R}}}_2 +D_t {{\mathcal {D}}}_t\nabla P_{\mathfrak {J}, v}+D_t \nabla P_{{{\mathcal {D}}}_t \mathfrak {J}, v}, \end{aligned}$$
(5.30)

where

$$\begin{aligned} \begin{aligned} {{\mathcal {R}}}_2&= D_t \big (\tilde{R}_0-\nabla {{\mathcal {H}}}(D_t^2P_{v,v}) \big )\\&\quad +[\nabla , D_t]{{\mathcal {H}}}(\Delta _{\Gamma _t} \mathfrak {J}^\bot -h_v)+\nabla [{{\mathcal {H}}}, D_t](\Delta _{\Gamma _t} \mathfrak {J}^\bot -h_v)+\nabla {{\mathcal {H}}}[\Delta _{\Gamma _t}, D_t]\mathfrak {J}^\bot \\&\quad +D_t( \nabla P_{\nabla P_{\mathfrak {J}, v}, v} ) -D_t \nabla {{\mathcal {H}}}\big [D_t\big (v\cdot (\nabla P_{v,v}|_{c})\big )+ v\cdot (\nabla {\mathcal {P}}_{t,1}|_{c})\big ]. \end{aligned} \end{aligned}$$

In fact, this complicated form is used due to similar technical reasons as (3.15), see Remark 3.1, where the terms \(\nabla {{\mathcal {H}}}({{\mathcal {P}}}_{t,2})\), \(D_t h_v\) are added on the left side to improve the estimates.

Next, we apply \(L^2(\Omega _t)\) inner product with \({{\mathcal {D}}}^2_t \mathfrak {J} -\nabla {{\mathcal {H}}}({{\mathcal {P}}}_{t,2})\) on both sides of (5.30) to get

$$\begin{aligned} \begin{aligned}&\frac{1}{2}\frac{\text {d}}{{\text {d}}t} \Vert {{\mathcal {D}}}^2_t \mathfrak {J} -\nabla {{\mathcal {H}}}({{\mathcal {P}}}_{t,2})\Vert ^2_{L^2(\Omega _t)} \\&\qquad -\int _{\Omega _t} \nabla {{\mathcal {H}}}\big (\Delta _{\Gamma _t} D_t \mathfrak {J}^\bot -D_t h_v\big )\cdot \big ({{\mathcal {D}}}^2_t \mathfrak {J} -\nabla {{\mathcal {H}}}({{\mathcal {P}}}_{t,2})\big ) {\text {d}}X \\&\quad =\int _{\Omega _t} {{\mathcal {R}}}_2 \cdot \big ({{\mathcal {D}}}^2_t \mathfrak {J} -\nabla {{\mathcal {H}}}({{\mathcal {P}}}_{t,2})\big ){\text {d}}X +\int _{\Omega _t}\big ( D_t {{\mathcal {D}}}_t\nabla P_{{\mathfrak {J}}, v}+D_t \nabla P_{{{\mathcal {D}}}_t {\mathfrak {J}}, v} \big )\\&\qquad \cdot \big ({{\mathcal {D}}}^2_t \mathfrak {J} -\nabla {{\mathcal {H}}}({{\mathcal {P}}}_{t,2})\big ){\text {d}}X. \end{aligned} \end{aligned}$$

Consequently, for any \(t'\in [0,T]\) if holds that

$$\begin{aligned} \begin{aligned}&\frac{1}{2}\big \Vert {{\mathcal {D}}}^2_t \mathfrak {J}(t') -\nabla {{\mathcal {H}}}({{\mathcal {P}}}_{t,2})(t')\big \Vert ^2_{L^2(\Omega _{t'})}\\&\qquad -\int _0^{t'}\int _{\Omega _t} \nabla {{\mathcal {H}}}\big (\Delta _{\Gamma _t} D_t \mathfrak {J}^\bot -D_th_v\big )\cdot \big ({{\mathcal {D}}}^2_t \mathfrak {J} -\nabla {{\mathcal {H}}}({{\mathcal {P}}}_{t,2})\big ) {\text {d}}X{\text {d}}t\\&\quad =\frac{1}{2} \big \Vert {{\mathcal {D}}}^2_t \mathfrak {J}(0) -\nabla {{\mathcal {H}}}({{\mathcal {P}}}_{t,2})(0)\big \Vert ^2_{L^2(\Omega _0)} +\int _0^{t'}\int _{\Omega _t} {{\mathcal {R}}}_2 \cdot \big ({{\mathcal {D}}}^2_t \mathfrak {J} -\nabla {{\mathcal {H}}}({{\mathcal {P}}}_{t,2})\big ){\text {d}}X{\text {d}}t\\&\qquad +\int _0^{t'}\int _{\Omega _t}\big ( D_t {{\mathcal {D}}}_t\nabla P_{J, v}+D_t \nabla P_{{{\mathcal {D}}}_t J, v} \big )\cdot \big ({{\mathcal {D}}}^2_t \mathfrak {J} -\nabla {{\mathcal {H}}}({{\mathcal {P}}}_{t,2})\big ){\text {d}}X{\text {d}}t. \end{aligned} \end{aligned}$$
(5.31)

5.3.1 Left Side of (5.31)

In this subsection, we prove estimates for the second term on the left side of (5.31).

Proposition 5.4

One has for some \(\delta _0\in (0,1)\) the following estimate:

$$\begin{aligned} \begin{aligned}&-\int _0^T\int _{\Omega _t} \nabla {{\mathcal {H}}}\big (\Delta _{\Gamma _t} D_t \mathfrak {J}^\bot -D_t h_v\big )\cdot \big ({{\mathcal {D}}}^2_t \mathfrak {J} -\nabla {{\mathcal {H}}}({{\mathcal {P}}}_{t,2})\big ){\text {d}}X{\text {d}}t \\&\quad \geqq \frac{1}{2} \sup _{t\in [0, T]}\int _{\Gamma _t}|\nabla _{\tau _t}D_t \mathfrak {J}^\bot |^2{\text {d}}s + \frac{1}{4}\int _0^T F_1(t)dt -\int _0^TP(E(t))\big (1+E_1(t)^{3/2}\big )dt\\&\qquad -\sup _{t\in [0, T]}P(E(t))\big (1+E_1(t)^{\delta _0}\big )-P(E(0))\big (1+E_1(0)^{\delta _0}\big ). \end{aligned} \end{aligned}$$

Proof

Applying Green’s Formula, we have

$$\begin{aligned}{} & {} - \int _{\Omega _t} \nabla {{\mathcal {H}}}\big (\Delta _{\Gamma _t} D_t \mathfrak {J}^\bot - D_t h_v\big )\cdot \big ({{\mathcal {D}}}^2_t \mathfrak {J} -\nabla {{\mathcal {H}}}({{\mathcal {P}}}_{t,2})\big ){\text {d}}X\nonumber \\{} & {} \quad =-\int _{\Gamma _t} \big (\Delta _{\Gamma _t}{{\mathcal {D}}}_t \mathfrak {J}^\bot -D_t h_v\big ) \big ({{\mathcal {D}}}^2_t \mathfrak {J} -\nabla {{\mathcal {H}}}({{\mathcal {P}}}_{t,2})\big )\cdot n_t{\text {d}}s\nonumber \\{} & {} \quad = -\int _{\Gamma _t} \big (\Delta _{\Gamma _t} D_t \mathfrak {J}^\bot - D_t h_v\big ) \big (D_t^2 \mathfrak {J}+D_t\nabla P_{\mathfrak {J}, v}+\nabla P_{\nabla P_{\mathfrak {J}, v}, v}+\nabla P_{{{\mathcal {D}}}_t \mathfrak {J}, v}\big )\cdot n_t{\text {d}}s\nonumber \\{} & {} \qquad +\int _{\Gamma _t} \big (\Delta _{\Gamma _t} D_t \mathfrak {J}^\bot - D_t h_v\big ) \nabla _{n_t} {{\mathcal {H}}}({{\mathcal {P}}}_{t,2}) {\text {d}}s\nonumber \\{} & {} \quad =-\int _{\Gamma _t} \Delta _{\Gamma _t} D_t \mathfrak {J}^\bot \, D_t^2 \mathfrak {J}\cdot n_t{\text {d}}s+\int _{\Gamma _t} D_t h_v \ D_t^2 \mathfrak {J} \cdot n_t{\text {d}}s +I_R, \end{aligned}$$
(5.32)

where

$$\begin{aligned} \begin{aligned} I_R&=-\int _{\Gamma _t} \big (\Delta _{\Gamma _t} D_t \mathfrak {J}^\bot -D_t h_v\big ) \big ( D_t\nabla P_{\mathfrak {J}, v}+\nabla P_{\nabla P_{\mathfrak {J}, v}, v}+\nabla P_{{{\mathcal {D}}}_t \mathfrak {J}, v}\big )\cdot n_t{\text {d}}s\\&\quad +\int _{\Gamma _t} \big (\Delta _{\Gamma _t} D_t \mathfrak {J}^\bot - D_t h_v\big )\cdot \nabla _{n_t} {{\mathcal {H}}}({{\mathcal {P}}}_{t,2}) {\text {d}}s. \end{aligned} \end{aligned}$$

We deal with these integrals above one by one. First, integrating by parts and taking one \(D_t\) out of the first integral in (5.32), we get

$$\begin{aligned}{} & {} -\int _{\Gamma _t} \Delta _{\Gamma _t} D_t \mathfrak {J}^\bot \, D_t^2 \mathfrak {J}\cdot n_t {\text {d}}s\nonumber \\{} & {} \quad = \frac{1}{2}\frac{\text {d}}{{\text {d}}t}\int _{\Gamma _t}|\nabla _{\tau _t}D_t \mathfrak {J}^\bot |^2{\text {d}}s-\frac{1}{2} \int _{\Gamma _t}|\nabla _{\tau _t}D_t \mathfrak {J}^\bot |^2\,D_t{\text {d}}s+ \nabla _{\tau _t} D_t \mathfrak {J}^\bot \,(D_t^2 \mathfrak {J}\cdot n_t)\big |^{p_r}_{p_l}\nonumber \\{} & {} \qquad +\int _{\Gamma _t} \nabla _{\tau _t} D_t \mathfrak {J}^\bot \, [\nabla _{\tau _t},D_t] D_t\mathfrak {J}^\bot {\text {d}}s +\int _{\Gamma _t} \nabla _{\tau _t} D_t \mathfrak {J}^\bot \,\nabla _{\tau _t} \big ([n_t, D_t^2] \cdot \mathfrak {J}\big ){\text {d}}s. \end{aligned}$$
(5.33)

Here, Lemma 5.3 is applied at the corner terms,

$$\begin{aligned} \begin{aligned}&\nabla _{\tau _t} D_t \mathfrak {J}^\bot \,(D_t^2 \mathfrak {J}\cdot n_t)\big |^{p_r}_{p_l}\\&\quad = \nabla _{\tau _t} D_t \mathfrak {J}^\bot \ \frac{\sigma ^2 }{\beta _c}(\sin \omega _l) D_t\nabla _{\tau _t} \mathfrak {J}^\bot (\tau _b\cdot n_t)\big |_{p_r}\\&\qquad -\nabla _{\tau _t} D_t \mathfrak {J}^\bot \ \frac{\sigma ^2 }{\beta _c}(\sin \omega _r) D_t\nabla _{\tau _t} \mathfrak {J}^\bot (\tau _b\cdot n_t)\big |_{p_l}\\&\qquad +\nabla _{\tau _t} D_t \mathfrak {J}^\bot (R_{c,h,l}\cdot n_t)\big |_{p_r}-\nabla _{\tau _t} D_t \mathfrak {J}^\bot (R_{c,h,r}\cdot n_t)\big |_{p_l}\\&\quad =F_1(t)+\nabla _{\tau _t} D_t \mathfrak {J}^\bot \ \frac{\sigma ^2 }{\beta _c}(\sin \omega _l) [D_t,\nabla _{\tau _t}] \mathfrak {J}^\bot (\tau _b\cdot n_t)\big |_{p_r}\\&\qquad -\nabla _{\tau _t} D_t \mathfrak {J}^\bot \ \frac{\sigma ^2 }{\beta _c}(\sin \omega _r) [D_t,\nabla _{\tau _t} ] \mathfrak {J}^\bot (\tau _b\cdot n_t)\big |_{p_l}\\&\qquad +\nabla _{\tau _t} D_t \mathfrak {J}^\bot (R_{c,h,l}\cdot n_t)\big |_{p_r}-\nabla _{\tau _t} D_t \mathfrak {J}^\bot (R_{c,h,r}\cdot n_t)\big |_{p_l}, \end{aligned} \end{aligned}$$

so one has, by (5.21) and lemmas from Section 4.1, that

$$\begin{aligned} \nabla _{\tau _t} D_t \mathfrak {J}^\bot \,(D_t^2 \mathfrak {J}\cdot n_t)\big |^{p_r}_{p_l}\geqq \frac{1}{2} F_1(t)-P(E(t))\big (1+E_1(t)\big ). \end{aligned}$$

On the other hand, since \(D_tds=(v^\perp \kappa +\nabla _{\tau _t}(v\cdot \tau _t)\big ){\text {d}}s\), we directly obtain the estimate:

$$\begin{aligned} \frac{1}{2}\int _{\Gamma _t}|\nabla _{\tau _t}D_t \mathfrak {J}^\bot |^2\,D_t{\text {d}}s\leqq P(E(t))E_1(t). \end{aligned}$$

Moreover, a direct computation using (5.29) shows that

$$\begin{aligned} \begin{aligned}{}[n_t, D_t^2] \cdot \mathfrak {J}&=- D_t(D_tn_t)\cdot {\mathfrak {J}}-2(D_tn_t)\cdot D_t{\mathfrak {J}}\\&=-D_t\big (-(\nabla _{\tau _t}v^\perp )\tau _t+(\nabla _{\tau _t}n_t\cdot v)\tau _t\big )\cdot {\mathfrak {J}}\\&\quad -2\big (-(\nabla _{\tau _t}v^\perp )\tau _t+(\nabla _{\tau _t}n_t\cdot v)\tau _t\big )\cdot D_t{\mathfrak {J}}. \end{aligned} \end{aligned}$$

Therefore, thanks to (5.8), Lemma 5.2 and checking term by term, we have the following estimate for the last two integrals in (5.33):

$$\begin{aligned}&\int _{\Gamma _t} \nabla _{\tau _t} D_t \mathfrak {J}^\bot \, [\nabla _{\tau _t},D_t] D_t\mathfrak {J}^\bot {\text {d}}s +\int _{\Gamma _t} \nabla _{\tau _t} D_t \mathfrak {J}^\bot \,\nabla _{\tau _t} \big ([n_t, D_t^2] \cdot \mathfrak {J}\big ){\text {d}}s\\&\quad \leqq P(E(t))\big (1+E_1(t)\big ). \end{aligned}$$

As a result, the proof is finished as long as we have Lemma 5.5 and Lemma 5.6. \(\quad \square \)

We deal with the remainder integrals in (5.32) in the next two lemmas.

Lemma 5.5

One has for the second term of the right hand of (5.32) that

$$\begin{aligned} \begin{aligned} \Big |\int _0^T\int _{\Gamma _t} D_t h_v \ D_t^2 \mathfrak {J} \cdot n_t{\text {d}}s \Big |&\leqq P(E(0))\big (1+E_1(0)^{\delta _0}\big )\\&\quad +\sup _{t\in [0, T]}P(E(t))\big (1+E_1(t)^{\delta _0}\big ) \\&\quad +\int _0^T P(E(t))\big (1+E_1(t)^{3/2}\big )dt+ \frac{1}{8} \int _0^T F_1(t). \end{aligned} \end{aligned}$$

Proof

To begin with, one recalls the definition of \(h_v\) in (5.2) to obtain

$$\begin{aligned} \int _{\Gamma _t} D_t h_v \ D_t^2 \mathfrak {J} \cdot n_t{\text {d}}s= & {} \int _{\Gamma _t} D_t\big (\Delta _{\Gamma _t} v\cdot D_t n_t \big ) D_t^2 \mathfrak {J} \cdot n_t{\text {d}}s\\{} & {} +\int _{\Gamma _t} D_t\big [2D^2v\big (\tau _t, (\nabla _{\tau _t}v)^\top \big )\cdot n_t\big ] D_t^2 \mathfrak {J} \cdot n_t{\text {d}}s. \end{aligned}$$

We only deal with the first integral in the integral above, since the second one can be handled in a similar way.

In fact, it is straightforward to see by (5.29) that

$$\begin{aligned} \begin{aligned}&\int _{\Gamma _t} D_t\big (\Delta _{\Gamma _t} v\cdot D_t n_t \big ) D_t^2 \mathfrak {J} \cdot n_t{\text {d}}s\\&\quad =\int _{\Gamma _t} D_t\big [\Delta _{\Gamma _t} v\cdot \tau _t(-\nabla _{\tau _t}v^\perp +\nabla _{\tau _t}n_t\cdot v )\big ] D_t^2 \mathfrak {J} \cdot n_t{\text {d}}s\\&\quad = \int _{\Gamma _t} D_t\Delta _{\Gamma _t} v\cdot \tau _t(-\nabla _{\tau _t}v^\perp +\nabla _{\tau _t}n_t\cdot v) D_t^2 \mathfrak {J} \cdot n_t{\text {d}}s\\&\quad + \int _{\Gamma _t}\Delta _{\Gamma _t} v\cdot D_t\big [\tau _t(-\nabla _{\tau _t}v^\perp +\nabla _{\tau _t}n_t\cdot v)\big ] D_t^2 \mathfrak {J} \cdot n_t{\text {d}}s \\&\quad \triangleq I_1+I_2. \end{aligned} \end{aligned}$$

The estimates for \(I_1, I_2\) are proved in the following lines.

- Estimates of \(I_1\). First, direct computations lead to

$$\begin{aligned} I_1= & {} \int _{\Gamma _t} D_t\Delta _{\Gamma _t} v\cdot \tau _t(-\nabla _{\tau _t}v^\perp +\nabla _{\tau _t}n_t\cdot v) D_t^2 (\mathfrak {J}-{\mathfrak {J}}|_c) \cdot n_t{\text {d}}s \\{} & {} +\int _{\Gamma _t} D_t\Delta _{\Gamma _t} v\cdot \tau _t(-\nabla _{\tau _t}v^\perp +\nabla _{\tau _t}n_t\cdot v) D_t^2 (\mathfrak {J}|_c) \cdot n_t{\text {d}}s \\= & {} \frac{\text {d}}{{\text {d}}t} \int _{\Gamma _t} D_t\Delta _{\Gamma _t} v\cdot \tau _t(-\nabla _{\tau _t}v^\perp +\nabla _{\tau _t}n_t\cdot v) D_t (\mathfrak {J}-{\mathfrak {J}}|_c) \cdot n_t{\text {d}}s\\{} & {} -\int _{\Gamma _t} D^2_t\Delta _{\Gamma _t} v\cdot \tau _t(-\nabla _{\tau _t}v^\perp +\nabla _{\tau _t}n_t\cdot v) D_t(\mathfrak {J}-{\mathfrak {J}}|_c) \cdot n_t{\text {d}}s\\{} & {} -\int _{\Gamma _t} D_t\Delta _{\Gamma _t} v\cdot D_t\big [\tau _t(-\nabla _{\tau _t}v^\perp +\nabla _{\tau _t}n_t\cdot v)\big ] D_t (\mathfrak {J}-{\mathfrak {J}}|_c) \cdot n_t{\text {d}}s\\{} & {} -\int _{\Gamma _t} D_t\Delta _{\Gamma _t} v\cdot \tau _t(-\nabla _{\tau _t}v^\perp +\nabla _{\tau _t}n_t\cdot v) \,D_t (\mathfrak {J}-{\mathfrak {J}}|_c) \cdot D_t(n_t{\text {d}}s)\\{} & {} +\int _{\Gamma _t} D_t\Delta _{\Gamma _t} v\cdot \tau _t(-\nabla _{\tau _t}v^\perp +\nabla _{\tau _t}n_t\cdot v) D_t^2 (\mathfrak {J}|_c) \cdot n_t{\text {d}}s \\{} & {} \triangleq \frac{\text {d}}{{\text {d}}t} \int _{\Gamma _t} D_t\Delta _{\Gamma _t} v\cdot \tau _t(-\nabla _{\tau _t}v^\perp +\nabla _{\tau _t}n_t\cdot v) D_t (\mathfrak {J}-{\mathfrak {J}}|_c) \cdot n_t{\text {d}}s+\sum _{i=1}^4I_{1i}. \end{aligned}$$

We deal with \(I_{1i}\) one by one. In fact, applying integration by parts, we have

$$\begin{aligned} \begin{aligned} I_{14}&=D_t\nabla _{\tau _t} v\cdot \tau _t(-\nabla _{\tau _t}v^\perp +\nabla _{\tau _t}n_t\cdot v) \,D_t^2 (\mathfrak {J}|_c) \cdot n_t\big |^{p_l}_{p_r}\\&\quad -\int _{\Gamma _t} D_t\nabla _{\tau _t} v\cdot \nabla _{\tau _t}\big [ \tau _t(-\nabla _{\tau _t}v^\perp +\nabla _{\tau _t}n_t\cdot v) D_t^2 (\mathfrak {J}|_c) \cdot n_t\big ]{\text {d}}s\\&\quad + \int _{\Gamma _t} [D_t,\nabla _{\tau _t}]\nabla _{\tau _t} v\cdot \tau _t(-\nabla _{\tau _t}v^\perp +\nabla _{\tau _t}n_t\cdot v) D_t^2 (\mathfrak {J}|_c) \cdot n_t{\text {d}}s\\&\quad - \int _{\Gamma _t} D_t\big ((\nabla _{\tau _t}\tau _t)^\top \cdot \nabla v\big )\cdot \tau _t(-\nabla _{\tau _t}v^\perp +\nabla _{\tau _t}n_t\cdot v) D_t^2 (\mathfrak {J}|_c) \cdot n_t{\text {d}}s. \end{aligned} \end{aligned}$$

Using (5.10) and similar arguments as to those in (4.22) for \(\partial ^2v\) and \(v^\perp \) terms and applying Lemma 5.3 to \(D^2{\mathfrak {J}}|_{p_i}\), we derive

$$\begin{aligned} |I_{14}|\leqq \frac{1}{16}F_1(t)+P(E(t))\big (1+E(t)\big ). \end{aligned}$$

For \(I_{11}\), since

$$\begin{aligned} \begin{aligned}&D^2_t\Delta _{\Gamma _t} v=D^2_t\big (\nabla _{\tau _t}\nabla _{\tau _t}v-(\nabla _{\tau _t}\tau _t\cdot \tau _t)\nabla _{\tau _t} v\big )\\&\quad =\nabla _{\tau _t}D^2_t\nabla _{\tau _t}v+[D^2_t, \nabla _{\tau _t}]\nabla _{\tau _t}v-D^2_t \big ((\nabla _{\tau _t}\tau _t\cdot \tau _t)\nabla _{\tau _t} v\big )\\&\quad =\nabla _{\tau _t}D^2_t\nabla _{\tau _t}v+D_t\big ((D_t\tau _t-\nabla _{\tau _t}v)\cdot \nabla \nabla _{\tau _t}v\big )\\&\qquad +(D_t\tau _t-\nabla _{\tau _t}v)\cdot \nabla D_t\nabla _{\tau _t}v-D^2_t \big ((\nabla _{\tau _t}\tau _t\cdot \tau _t)\nabla _{\tau _t} v\big )\\&\quad \triangleq h_{v,2}+l.o.t. \end{aligned} \end{aligned}$$

where

$$\begin{aligned} h_{v,2}= & {} \nabla _{\tau _t}D^2_t\nabla _{\tau _t}v+ 2\nabla _{\tau _t}\big ((D_t\tau _t-\nabla _{\tau _t}v)\cdot \nabla D_t v\big ) \\{} & {} -\nabla ^2_{\tau _t}D_tv\cdot n_t(n_t\cdot \tau _t)\nabla _{\tau _t}v -(\nabla _{\tau _t}\tau _t\cdot \tau _t)\nabla _{\tau _t}D^2_t v \end{aligned}$$

contains higher-order terms of v, and the remainder part contains products like \(\partial D_tv\,\partial ^2v\) and other lower-order terms and can be controlled by (5.8), (5.12), Lemma 5.2 and similar arguments as in (4.22).

Therefore, we arrive at

$$\begin{aligned} I_{11}=\int _{\Gamma _t} h_{v,2}\cdot \tau _t(-\nabla _{\tau _t}v^\perp +\nabla _{\tau _t}n_t\cdot v) D_t(\mathfrak {J}-{\mathfrak {J}}|_c) \cdot n_t{\text {d}}s+l.o.t.. \end{aligned}$$

To finish the estimate for \(I_{11}\), we firstly take care of the first term \(\nabla _{\tau _t}D^2_t\nabla _{\tau _t}v\) in \(h_{v,2}\). In fact, we use integration by parts on \(\Gamma _t\) to obtain

$$\begin{aligned} \begin{aligned}&\int _{\Gamma _t} \nabla _{\tau _t}D^2_t\nabla _{\tau _t}v\cdot \tau _t(-\nabla _{\tau _t}v^\perp +\nabla _{\tau _t}n_t\cdot v) D_t(\mathfrak {J}-{\mathfrak {J}}|_c) \cdot n_t{\text {d}}s\\&\quad =-\int _{\Gamma _t}D^2_t\nabla _{\tau _t}v\cdot \nabla _{\tau _t}\big (\tau _t(-\nabla _{\tau _t}v^\perp +\nabla _{\tau _t}n_t\cdot v)\big ) D_t(\mathfrak {J}-{\mathfrak {J}}|_c) \cdot n_t{\text {d}}s\\&\qquad -\int _{\Gamma _t} D^2_t\nabla _{\tau _t}v\cdot \tau _t(-\nabla _{\tau _t}v^\perp +\nabla _{\tau _t}n_t\cdot v) \nabla _{\tau _t}\big (D_t(\mathfrak {J}-{\mathfrak {J}}|_c) \cdot n_t\big ){\text {d}}s. \end{aligned} \end{aligned}$$

Similar arguments as to those in(4.22) and applying (5.8), (5.12), we derive

$$\begin{aligned} \int _{\Gamma _t} \nabla _{\tau _t}D^2_t\nabla _{\tau _t}v\cdot \tau _t(-\nabla _{\tau _t}v^\perp +\nabla _{\tau _t}n_t\cdot v) D_t(\mathfrak {J}-{\mathfrak {J}}|_c) \cdot n_t{\text {d}}s\leqq P(E(t))\big (1+E_1(t)\big ). \end{aligned}$$

Meanwhile, the other terms in \(I_{11}\) can also be handled similarly, so we conclude

$$\begin{aligned} |I_{11}|\leqq P(E(t))\big (1+E_1(t)\big ). \end{aligned}$$

For \(I_{12}\), we have

$$\begin{aligned} \begin{aligned} I_{12}&=-\int _{\Gamma _t} \big ([D_t,\nabla _{\tau _t}]\nabla _{\tau _t}v-D_t(\nabla _{\tau _t}\tau _t\cdot \tau _t\nabla _{\tau _t} v)\big )\\&\quad \cdot D_t\big [\tau _t(-\nabla _{\tau _t}v^\perp +\nabla _{\tau _t}n_t\cdot v)\big ] D_t (\mathfrak {J}-{\mathfrak {J}}|_c) \cdot n_t{\text {d}}s \\&\quad -\int _{\Gamma _t} \nabla _{\tau _t} D_t\nabla _{\tau _t}v\cdot D_t\big [\tau _t(-\nabla _{\tau _t}v^\perp +\nabla _{\tau _t}n_t\cdot v)\big ] D_t (\mathfrak {J}-{\mathfrak {J}}|_c) \cdot n_t{\text {d}}s, \end{aligned} \end{aligned}$$

where the first integral can be handled as above. For the second integral, we know from Lemma 2.5 that

$$\begin{aligned} D_t (\mathfrak {J}-{\mathfrak {J}}|_c)\in {\tilde{H}}^{1/2}(\Gamma _t), \end{aligned}$$

so we apply Lemmas 2.5, 2.6, (5.8) and 5.2 to find

$$\begin{aligned} \begin{aligned}&-\int _{\Gamma _t} \nabla _{\tau _t} D_t\nabla _{\tau _t}v\cdot D_t\big [\tau _t(-\nabla _{\tau _t}v^\perp +\nabla _{\tau _t}n_t\cdot v)\big ] D_t (\mathfrak {J}-{\mathfrak {J}}|_c) \cdot n_t{\text {d}}s\\&\quad \leqq \big \Vert D_t\big [\tau _t(-\nabla _{\tau _t}v^\perp +\nabla _{\tau _t}n_t\cdot v)\big ]\big \Vert _{L^\infty (\Gamma _t)}\Vert \nabla _{\tau _t} D_t\nabla _{\tau _t}v\Vert _{{\tilde{H}}^{-1/2}(\Gamma _t)}\Vert \Vert D_t (\mathfrak {J}-{\mathfrak {J}}|_c) \\&\qquad \cdot n_t\Vert _{{\tilde{H}}^{1/2}(\Gamma _t)}\\&\quad \leqq P(E(t))\Vert D_t\nabla _{\tau _t}v\Vert _{H^{1/2}(\Gamma _t)}\Vert D_t{\mathfrak {J}}\Vert _{H^1(\Omega _t)}\\&\quad \leqq P(E(t))\big (1+E_1(t)\big ). \end{aligned} \end{aligned}$$
(5.34)

As a result, we summarize that

$$\begin{aligned} |I_{12}|\leqq P(E(t))\big (1+E_1(t)^{3/2}\big ). \end{aligned}$$

Moreover, similar arguments as for \(I_{12}\), we also have

$$\begin{aligned} | I_{13}| \leqq P(E(t))\big (1+E_1(t)\big ). \end{aligned}$$

Together with all these estimates above for \(I_{11}\) to \(I_{14}\), we can go back to \(I_1\) expression and integrate on both sides with respect to time t on [0, T] to find

$$\begin{aligned} \begin{aligned} \int ^T_0 I_1 dt&\leqq \int _{\Gamma _t} D_t\Delta _{\Gamma _t} v\cdot \tau _t(-\nabla _{\tau _t}v^\perp +\nabla _{\tau _t}n_t\cdot v) D_t (\mathfrak {J}-{\mathfrak {J}}|_c) \cdot n_t{\text {d}}s\big |^T_0\\&\quad +\int ^T_0P(E(t))\big (1+E_1(t)^{3/2}\big )dt+\frac{1}{16} \int _0^T F_1(t). \end{aligned}\end{aligned}$$

Moreover, we have for the first integral on the right side above the following estimate

$$\begin{aligned} \begin{aligned}&\int _{\Gamma _t} D_t\Delta _{\Gamma _t} v\cdot \tau _t(-\nabla _{\tau _t}v^\perp +\nabla _{\tau _t}n_t\cdot v) D_t (\mathfrak {J}-{\mathfrak {J}}|_c) \cdot n_t{\text {d}}s\\&\quad = \int _{\Gamma _t} \big ([D_t,\nabla _{\tau _t}]\nabla _{\tau _t}v-D_t(\nabla _{\tau _t}\tau _t\cdot \tau _t\nabla _{\tau _t} v)\big )\cdot \tau _t(-\nabla _{\tau _t}v^\perp \\&\qquad +\nabla _{\tau _t}n_t\cdot v) D_t (\mathfrak {J}-{\mathfrak {J}}|_c) \cdot n_t{\text {d}}s\\&\qquad +\int _{\Gamma _t} \nabla _{\tau _t}D_t\nabla _{\tau _t} v\cdot \tau _t(-\nabla _{\tau _t}v^\perp +\nabla _{\tau _t}n_t\cdot v) D_t (\mathfrak {J}-{\mathfrak {J}}|_c) \cdot n_t{\text {d}}s. \end{aligned} \end{aligned}$$

Similar arguments as in (5.34) and checking carefully on the highest-order terms, we derive

$$\begin{aligned} \int _{\Gamma _t} D_t\Delta _{\Gamma _t} v\cdot \tau _t(-\nabla _{\tau _t}v^\perp +\nabla _{\tau _t}n_t\cdot v) D_t (\mathfrak {J}-{\mathfrak {J}}|_c) \cdot n_t{\text {d}}s \leqq P(E(t))\big (1+E_1(t)^{\delta _0}\big ) \end{aligned}$$

with the number \(\delta _0\in (0,1)\) as above.

Consequently, we have the following estimate for \(I_1\):

$$\begin{aligned} \begin{aligned} \Big |\int _0^T I_1 dt \Big |&\quad \leqq P(E(0))\big (1+E_1(0)^{\delta _0}\big )+\sup _{t\in [0, T]}P(E(t))\big (1+E_1(t)^{\delta _0}\big ) \\&\qquad +\int _0^T P(E(t))\big (1+E_1(t)^{3/2}\big )dt+ \frac{1}{16} \int _0^T F_1(t). \end{aligned} \end{aligned}$$
(5.35)

- Estimates of \(I_2\). The integral \(I_2\) can be handled in the same way as \(I_1\). We simply conclude that

$$\begin{aligned} \begin{aligned} \Big |\int _0^T I_2 dt \Big |&\leqq P(E(0))\big (1+E_1(0)^{\delta _0}\big )+\sup _{t\in [0, T]}P(E(t)\big (1+E_1(t)^{\delta _0}\big )\\&\quad +\int _0^T P(E(t))\big (1+E_1(t)^{3/2}\big )dt+ \frac{1}{16} \int _0^T F_1(t). \end{aligned}\end{aligned}$$
(5.36)

In the end, combing (5.35) with (5.36), the proof is finished. \(\quad \square \)

Next, we deal with \(I_R\).

Lemma 5.6

One has

$$\begin{aligned} |I_R|\leqq & {} \frac{1}{8} F_1(t)+P(E_1(t)). \end{aligned}$$
(5.37)

Proof

First, to simplify the notations in the first integral of \(I_R\), we denote by

$$\begin{aligned} C_P =(D_t\nabla {{\mathcal {P}}}_{\mathfrak {J}, v}+\nabla {{\mathcal {P}}}_{\nabla {{\mathcal {P}}}_{\mathfrak {J}, v}, v}+\nabla {{\mathcal {P}}}_{{{\mathcal {D}}}_t \mathfrak {J}, v})\cdot n_t. \end{aligned}$$

Thanks to the definition of \(P_{w, v}\) (see (1.10)), we have

$$\begin{aligned} C_P= & {} C'_{\mathfrak {J}, v}+{{\mathcal {C}}}_{\nabla P_{\mathfrak {J}, v}, v}+C_{{{\mathcal {D}}}_t \mathfrak {J}, v} -D_t n_t\cdot \nabla P_{\mathfrak {J}, v}\\{} & {} - D_t \big (({\mathfrak {J}}\cdot {\tau _t})\nabla _{\tau _t} v\cdot n_t\big )-(\nabla _{\tau _t}P_{\mathfrak {J}, v}+{{\mathcal {D}}}_t\mathfrak {J}\cdot \tau _t) \nabla _{\tau _t} v\cdot n_t. \end{aligned}$$

By (5.5), (5.8), Lemma 5.2 and (5.27), we show immediately

$$\begin{aligned} \Vert C_P\Vert _{H^{1}(\Gamma _t)}+\Vert \nabla _{n_t}{{\mathcal {H}}}({{\mathcal {P}}}_{t,2})\Vert _{H^{1}(\Gamma _t)}\leqq P(E(t))\big (1+E_1(t)^{1/2}\big ). \end{aligned}$$
(5.38)

Now we are ready to have estimate of \(I_R\). For the first integral in \(I_R\), we integrate by parts to derive

$$\begin{aligned} \begin{aligned} \int _{\Gamma _t} \Delta _{\Gamma _t} D_t \mathfrak {J}^\bot \, C_P {\text {d}}s&\quad =\nabla _{\tau _t} D_t \mathfrak {J}^\bot C_P\big |^{p_l}_{p_r}-\int _{\Gamma _t} \nabla _{\tau _t} D_t \mathfrak {J}^\bot \cdot \nabla _{\tau _t}C_P{\text {d}}s \\&\quad \leqq \frac{1}{8} F_1(t)+P(E(t))\big (1+E_1(t)\big ) \end{aligned} \end{aligned}$$

where (5.8), (5.11), (5.29), Lemma 5.3 and (5.38) are used.

Moreover, we have from (5.2) and (5.38) that

$$\begin{aligned} \begin{aligned} \int _{\Gamma _t} D_t h_v \, C_P{\text {d}}s&\quad =\int _{\Gamma _t}\big [\Delta _{\Gamma _t} v\cdot D_t n_t +2D^2v\big (\tau _t, (\nabla _{\tau _t}v)^\top \big )\cdot n_t\big ]C_P{\text {d}}s\\&\quad \leqq P(E(t))\big (1+E_1(t)\big ), \end{aligned} \end{aligned}$$

where similar analysis as in the proof of the previous lemma is applied.

On the other hand, the second integral in \(I_R\) can also be handled by a similar argument as above. As a result, the proof is finished. \(\quad \square \)

5.3.2 Right Side of (5.31)

We firstly deal with the integral involving \({{\mathcal {R}}}_2\).

Lemma 5.7

One has for some \(\delta _0\in (0,1)\) the following estimate:

$$\begin{aligned} \begin{aligned}&\Big |\int _0^T\int _{\Omega _t}{{\mathcal {R}}}_2 \cdot \big ({{\mathcal {D}}}_t^2 \mathfrak {J}-\nabla {{\mathcal {H}}}({{\mathcal {P}}}_{t,2})\big ){\text {d}}X{\text {d}}t \Big |\\&\quad \leqq P(E(0))\big (1+E_1(0)^{\delta _0}\big )\\&\qquad +\sup _{t\in [0, T]}P(E(t))\big (1+E_1(t)^{\delta _0}\big ) \\&\qquad +\int _0^T P(E(t))\big (1+E_1(t)^{3/2}\big )dt+ \frac{1}{16} \int _0^T F_1(t). \end{aligned} \end{aligned}$$

Proof

Recalling from (5.30), we deal with the terms in \({{\mathcal {R}}}_2\) one by one.

- Estimates of \(D_t \big (\tilde{R}_0-\nabla {{\mathcal {H}}}( D^2_t P_{v,v} ) \big ) \). In fact, one knows directly from (5.1) that

$$\begin{aligned} \tilde{R}_0- \nabla {{\mathcal {H}}}( D^2_t P_{v,v} )= & {} -\sigma \nabla {{\mathcal {H}}}(J\cdot \Delta _{\Gamma _t}n_t)\\{} & {} +\sigma \nabla {{\mathcal {H}}}\big ([n_t,\Delta _{\Gamma _t}]\cdot \nabla P_{v,v}\big )+\sigma \nabla {{\mathcal {H}}}(R_1+h_v )\\{} & {} +A_1+A_2+A_3, \end{aligned}$$

where recall that \(R_1+h_v\) only contains lower-order derivatives like \(\partial v, \partial n_t, \kappa \) and \(A_1\) to \(A_3\) are defined in (3.9)–(3.11).

As a result, checking term by term on \(\tilde{R}_0- \nabla {{\mathcal {H}}}( D^2_t P_{v,v} ) \), one finds that it contains \(\partial D_t v\), \(D_tJ\), \(\partial v\) and other lower-order terms, so acting \(D_t\) on \(\tilde{R}_0- \nabla {{\mathcal {H}}}( D^2_t P_{v,v} ) \) and applying lemmas in Sections 4.1 and 5.2 lead to the following estimate:

$$\begin{aligned} \big \Vert D_t \big (\tilde{R}_0-\nabla {{\mathcal {H}}}( D^2_t P_{v,v} ) \big ) \big \Vert _{L^2(\Omega _t)}\leqq P(E(t))\big (1+E_1(t)^{1/2}\big ). \end{aligned}$$

- Estimates of \([\nabla , D_t]{{\mathcal {H}}}(\Delta _{\Gamma _t}\mathfrak {J}^\bot -h_v) \). It is straightforward to show by (5.19) and Lemma 2.7 that

$$\begin{aligned}{} & {} \big \Vert [\nabla , D_t]{{\mathcal {H}}}(\Delta _{\Gamma _t} \mathfrak {J}^\bot -h_v) \big \Vert _{L^2(\Omega _t)} \\{} & {} \quad \leqq P(E(t))\Vert \Delta _{\Gamma _t}\mathfrak {J}^\bot -h_v \Vert _{H^{1/2}(\Gamma _t)}\leqq P(E(t))\big (1+E_1(t)^{1/2}\big ). \end{aligned}$$

- Estimates of \(\nabla [{{\mathcal {H}}}, D_t](\Delta _{\Gamma _t}\mathfrak {J}^\bot -h_v) \). Recalling (2.4) and using Lemma 2.7, (5.19) and applying variational estimates as in page 33 [52] imply that

$$\begin{aligned} \big \Vert \nabla [{{\mathcal {H}}}, D_t](\Delta _{\Gamma _t}\mathfrak {J}^\bot -h_v) \big \Vert _{L^2(\Omega _t)}\leqq & {} P(E(t))\Vert {{\mathcal {H}}}(\Delta _{\Gamma _t}\mathfrak {J}^\bot -h_v)\Vert _{H^1(\Omega _t)}\\\leqq & {} P(E(t))\big (1+E_1(t)^{1/2}\big ). \end{aligned}$$

- Estimates of \(\nabla {{\mathcal {H}}}[\Delta _{\Gamma _t}, D_t]\mathfrak {J}^\bot \). To estimate this term, we use similar arguments as in the proof of Lemma 5.5.

First, using Green’s Formula, we obtain

$$\begin{aligned} \begin{aligned}&\int _{\Omega _t} \nabla {{\mathcal {H}}}[\Delta _{\Gamma _t}, D_t]\mathfrak {J}^\bot \cdot {{\mathcal {D}}}_t^2 \mathfrak {J}{\text {d}}X= \int _{\Gamma _t}[\Delta _{\Gamma _t}, D_t]\mathfrak {J}^\bot \, {{\mathcal {D}}}_t^2 \mathfrak {J}\cdot n_t{\text {d}}s\\&\quad =\int _{\Gamma _t} [\big (\nabla _{\tau _t}\nabla _{\tau _t}-(\nabla _{\tau _t}\tau _t\cdot \tau _t)\nabla _{\tau _t}\big ), D_t]{\mathfrak {J}}^\perp \ {{\mathcal {D}}}_t^2{\mathfrak {J}}\cdot n_t{\text {d}}s\\&\quad = \int _{\Gamma _t} 2( \nabla _{\tau _t}v\cdot \tau _t)\nabla ^2_{\tau _t}{\mathfrak {J}}^\perp \ {{\mathcal {D}}}_t^2{\mathfrak {J}}\cdot n_t{\text {d}}s\\&\qquad + \int _{\Gamma _t}\big [\nabla _{\tau _t}( \nabla _{\tau _t}v\cdot \tau _t)+D_t(\nabla _{\tau _t}\tau _t\cdot \tau _t)-( \nabla _{\tau _t}\tau _t\cdot \tau _t)( \nabla _{\tau _t}v\cdot \tau _t) \big ] \nabla _{\tau _t}{\mathfrak {J}}^\perp \ {{\mathcal {D}}}_t^2 \mathfrak {J}\cdot n_t{\text {d}}s. \end{aligned} \end{aligned}$$

These two integrals can be handled in a similar way as before, and here we only give the details for the first part in the second integral above. Moreover, the analysis for \(\int _{\Omega _t} \nabla {{\mathcal {H}}}[\Delta _{\Gamma _t}, D_t]\mathfrak {J}^\bot \cdot \nabla {{\mathcal {H}}}({{\mathcal {P}}}_{t,2}){\text {d}}X\) can be done in a similar and easier way thanks to (5.27), and hence estimates for all these remainder parts are omitted.

Recalling (5.4) and the definition of \(C_P\), we rewrite the first part in the second integral above as follows:

$$\begin{aligned} \begin{aligned}&\int _{\Gamma _t}\nabla _{\tau _t}( \nabla _{\tau _t}v\cdot \tau _t) \nabla _{\tau _t} \mathfrak {J}^\bot \, {{\mathcal {D}}}_t^2 \mathfrak {J}\cdot n_t {\text {d}}s\\&\quad = \int _{\Gamma _t}\nabla _{\tau _t}( \nabla _{\tau _t}v\cdot \tau _t) \nabla _{\tau _t} \mathfrak {J}^\bot \, D_t^2 \mathfrak {J}\cdot n_t {\text {d}}s+ \int _{\Gamma _t}\nabla _{\tau _t}( \nabla _{\tau _t}v\cdot \tau _t) \nabla _{\tau _t} \mathfrak {J}^\bot \, C_P {\text {d}}s \triangleq I_3+I_4. \end{aligned} \end{aligned}$$

Similarly as the estimates for \(I_1\) in the proof of Lemma 5.5, one has for \(I_3\) that

$$\begin{aligned} \begin{aligned} I_3&=\int _{\Gamma _t}\nabla _{\tau _t}( \nabla _{\tau _t}v\cdot \tau _t) \nabla _{\tau _t} \mathfrak {J}^\bot \, D_t^2 (\mathfrak {J}-{\mathfrak {J}}|_c)\cdot n_t {\text {d}}s\\&\qquad +\int _{\Gamma _t}\nabla _{\tau _t}( \nabla _{\tau _t}v\cdot \tau _t) \nabla _{\tau _t} \mathfrak {J}^\bot \, D_t^2 (\mathfrak {J}|_c)\cdot n_t {\text {d}}s\\&\leqq \frac{1}{8} F_1(t)+P(E(t))\big (1+E_1(t)\big )\\&\qquad +\frac{\text {d}}{{\text {d}}t}\int _{\Gamma _t}\nabla _{\tau _t}( \nabla _{\tau _t}v\cdot \tau _t) \nabla _{\tau _t} \mathfrak {J}^\bot \, D_t (\mathfrak {J}-{\mathfrak {J}}|_c)\cdot n_t {\text {d}}s\\&\qquad -\int _{\Gamma _t}D_t\nabla _{\tau _t}( \nabla _{\tau _t}v\cdot \tau _t) \nabla _{\tau _t} \mathfrak {J}^\bot \, D_t (\mathfrak {J}-{\mathfrak {J}}|_c)\cdot n_t {\text {d}}s\\&\qquad -\int _{\Gamma _t}\nabla _{\tau _t}( \nabla _{\tau _t}v\cdot \tau _t) D_t\nabla _{\tau _t} \mathfrak {J}^\bot \, D_t (\mathfrak {J}-{\mathfrak {J}}|_c)\cdot n_t {\text {d}}s\\&\leqq \frac{1}{8} F_1(t)+P(E(t))\big (1+E_1(t)^{3/2}\big )\\&\qquad +\frac{\text {d}}{{\text {d}}t}\int _{\Gamma _t}\nabla _{\tau _t}( \nabla _{\tau _t}v\cdot \tau _t) \nabla _{\tau _t} \mathfrak {J}^\bot \, D_t (\mathfrak {J}-{\mathfrak {J}}|_c)\cdot n_t {\text {d}}s, \end{aligned} \end{aligned}$$

where we use (5.20), Lemma 5.2 and Lemma 5.3 to have

$$\begin{aligned} \begin{aligned}&\int _{\Gamma _t}\nabla _{\tau _t}( \nabla _{\tau _t}v\cdot \tau _t) D_t\nabla _{\tau _t} \mathfrak {J}^\bot \, D_t (\mathfrak {J}-{\mathfrak {J}}|_c)\cdot n_t {\text {d}}s\\&\quad \leqq P(E(t))\Big (1+\sum _i\big \Vert r^{\delta }(\nabla ^2_{\tau _t} v)\circ T^{-1}_i\big \Vert _{L^\infty (0, r_0)} \big \Vert \Vert D_t\nabla _{\tau _t}{\mathfrak {J}}^\perp \Vert _{L^2(\Gamma _t)}\big \Vert r^{-\delta }D_t (\mathfrak {J}-{\mathfrak {J}}|_c)\circ T^{-1}_i\big \Vert _{L^2(0, r_0)} \Big ) \end{aligned} \end{aligned}$$

with \(1>\delta >3-\pi /\omega _i\) for \(\omega _i\in (\pi /3,\pi /2)\) and

$$\begin{aligned} \begin{aligned} \Vert D_t\nabla _{\tau _t}{\mathfrak {J}}^\perp \Vert _{L^2(\Gamma _t)}&\leqq \big \Vert (\nabla _{\tau _t}v\cdot \tau _t)\nabla _{\tau _t}{\mathfrak {J}}^\perp \big \Vert _{L^2(\Gamma _t)}+\big \Vert \nabla _{\tau _t}D_t({\mathfrak {J}}\cdot n_t)\big \Vert _{L^2(\Gamma _t)}\\&\leqq P(E(t))\big (1+\Vert \nabla _{\tau _t}D_t{\mathfrak {J}}\Vert _{L^2(\Gamma _t)}+\Vert \nabla ^2_{\tau _t}v^\perp \Vert _{L^2(\Gamma _t)}\big )\\&\leqq P(E(t))\big (1+E_1(t)\big ). \end{aligned} \end{aligned}$$

As a result, we have

$$\begin{aligned} \begin{aligned} \Big |\int _0^T I_3 dt \Big |&\leqq P(E(0))\big (1+E_1(0)^{\delta _0}\big )+\sup _{t\in [0, T]}P(E(t))\big (1+E_1(t)^{\delta _0}\big ) \\&\quad +\int _0^T P(E(t))\big (1+E_1(t)^{3/2}\big )dt+ \frac{1}{64} \int _0^T F_1(t). \end{aligned} \end{aligned}$$

Since similar arguments can be applied to \(I_4\) and other integrals, we conclude directly

$$\begin{aligned} \begin{aligned} \Big |\int _0^T\int _{\Omega _t} \nabla {{\mathcal {H}}}[\Delta _{\Gamma _t}, D_t]\mathfrak {J}^\bot \cdot {{\mathcal {D}}}_t^2 \mathfrak {J}{\text {d}}X{\text {d}}t \Big |&\leqq P(E(0))\big (1+E_1(0)^{\delta _0}\big )\\&\quad +\sup _{t\in [0, T]}P(E(t))\big (1+E_1(t)^{\delta _0}\big ) \\&\quad +\int _0^T P(E(t))\big (1+E_1(t)^{3/2}\big )dt\\&\quad + \frac{1}{64} \int _0^T F_1(t). \end{aligned} \end{aligned}$$

- Estimates of \(D_t( \nabla P_{\nabla P_{\mathfrak {J}, v}, v} )\). Recalling the definition of \(P_{\mathfrak {J}, v}\) and \(P_{\nabla P_{\mathfrak {J}, v}, v} \) by (1.10), one has firstly the estimate (5.13). Moreover, one can take \(D_t\) on the system of \(P_{\nabla P_{\mathfrak {J}, v}, v}\) as in (4.7) to obtain the system for \(D_tP_{\nabla P_{\mathfrak {J}, v}, v}\).

Consequently, checking term by term and applying Lemma 4.6, Lemma 5.1, one derives

$$\begin{aligned} \Vert D_t( \nabla {{\mathcal {P}}}_{\nabla {{\mathcal {P}}}_{\mathfrak {J}, v}, v} )\Vert _{L^2(\Omega _t)} \leqq P(E(t))\big (1+E_1(t)^{1/2}\big ). \end{aligned}$$

- Estimates of the last term in \({{\mathcal {R}}}_2\). Thanks to (2.4), (5.12), (5.27) and Lemma 5.1, we have

$$\begin{aligned} \big \Vert D_t \nabla {{\mathcal {H}}}\big [D_t\big (v\cdot (\nabla P_{v,v}|_{c})\big )+ v\cdot (\nabla {\mathcal {P}}_{t,1}|_{c})\big ]\big \Vert _{L^2(\Omega _t)} \leqq P(E(t))\big (1+E_1(t)^{1/2}\big ). \end{aligned}$$

In the end, summing up all these estimates above, the proof is finished. \(\quad \square \)

At this moment, it remains to handle the last integral on the right side of (5.31). In fact, similar arguments as in the proof of Lemma 4.8, we can conclude the following estimate:

$$\begin{aligned} \begin{aligned}&\int _0^{t'}\int _{\Omega _t}\big ( D_t {{\mathcal {D}}}_t\nabla P_{J, v}+D_t \nabla P_{{{\mathcal {D}}}_t J, v} \big )\cdot \big ({{\mathcal {D}}}^2_t \mathfrak {J} -\nabla {{\mathcal {H}}}({{\mathcal {P}}}_{t,2})\big ){\text {d}}X{\text {d}}t\\&\quad \leqq P(E(0))\big (1+E_1(0)^{\delta _0}\big )+\sup _{t\in [0, T]}P(E(t))\big (1+E_1(t)^{\delta _0}\big )\\&\qquad +\int _0^T P(E(t))\big (1+E_1(t)^{3/2}\big )dt\\&\qquad + \frac{1}{16} \int _0^T F_1(t). \end{aligned} \end{aligned}$$

5.3.3 The End of the Higher-Order Energy Estimate

Summing up all the estimates in the previous two subsections, we finally arrive at:

$$\begin{aligned} \begin{aligned}&\sup _{t\in [0, T]}\Big (\big \Vert {{\mathcal {D}}}^2_t \mathfrak {J} -\nabla {{\mathcal {H}}}({{\mathcal {P}}}_{t,2})\big \Vert ^2_{L^2(\Omega _t)}+\int _{\Gamma _t}|\nabla _{\tau _t}D_t \mathfrak {J}^\bot |^2\Big )+\frac{1}{4}\int _0^T F_1(t)\\&\quad \leqq P(E(0))\big (1+E_1(0)^{\delta _0}\big )+\sup _{t\in [0, T]}P(E(t))\big (1+E_1(t)^{\delta _0}\big )\\&\qquad +\int _0^T P(E(t))\big (1+E_1(t)^{3/2}\big )dt. \end{aligned} \end{aligned}$$

Thanks to (5.27) and Lemma 5.1, we have for a number \(\delta _1>0\) small enough such that

$$\begin{aligned} \sup _{t\in [0, T]} E_1(t)+ \int _0^T F_1(t)\leqq & {} P(E(0))\big (1+E_1(0)^{\delta _0}\big )+\delta _1\sup _{t\in [0, T]} E_1(t)\\{} & {} +\int _0^TP(E(t))\big (1+E_1(t)^{3/2}\big )dt. \end{aligned}$$

Therefore, we’ve finished the proof of Theorem 5.1.

6 Well-Posedness of System \(\text{(WW) }\)

In this section, we use Picard iteration to prove the existence of solutions to \(\text{(WW) }\). The main idea is the same as [53], although some necessary modifications are needed and presented here. As a result, we only show the skecth of the proof. For more details, see [53].

6.1 Definitions of Surfaces and Domains

Following [53, 61], we introduce a map \(\Phi _{S_t}\) on the boundary \(S_t=\Gamma _t\cup \Gamma _b\) to fix the moving domain \(\Omega _t\). To start with, we choose a reference domain \(\Omega _*\) with upper surface \(\Gamma _{t*}\) and bottom \(\Gamma _{b*}\), which can be taken as the initial domain \(\Omega _0\) without loss of generality. The contact points of \(\Omega _*\) are denoted by \(p_{i*}\) (\(i=l, r\)) and the other notations follow similarly.

we define a unit upward vector field \(\mu \in H^s(\Gamma _{*}, {{\mathcal {S}}}^1)\) with some large s satisfying

$$\begin{aligned} \mu \cdot n_{t*}\geqq c_0\quad \hbox {on}\quad \Gamma _{t*},\qquad \hbox {and}\quad \mu |_{p_{l*}}=-\tau _{b*}|_{p_{l*}}, \ \mu |_{p_{r*}}=\tau _{b*}|_{p_{r*}} \end{aligned}$$

for some fixed constant \(c_0\in (0,1)\). Here we notice that the conditions above hold at \(p_{l*}, p_{r*}\).

Applying the Implicit Function Theorem, there exists a small constant \(d_0>0\) such that the map

is an \(H^s\) diffeomorphism from its domain to a neighborhood of \(\Gamma _*\).

Consequently, this map identifies each upper surface \(\Gamma _t\) near \(\Gamma _{t*}\) with a unique function

and we can define the following map

Meanwhile, we can use the function \(d_{\Gamma _t}(p)\) as the expression of the upper surface \(\Gamma _t\), and we have at the corner points that

$$\begin{aligned} d_{\Gamma _t}(p_{i*})=p_i,\quad i=l, r. \end{aligned}$$

Moreover, \(\Phi _{S_t}\) can be extended to the entire boundary \(S_*=\Gamma _{t*}\cup \Gamma _{b*}\). Consequently, we obtain the map on \(S_*\):

$$\begin{aligned} \Phi _{S_t}: \ \ S_* \rightarrow {S_t}. \end{aligned}$$

Using the harmonic extension, we define the following map on \(\Omega _*\):

$$\begin{aligned} {{\mathcal {T}}}_{S_t}:\ \Omega _*\rightarrow \Omega \quad \hbox {with}\quad {{\mathcal {T}}}_{S_t}={{\mathcal {H}}}_*\big (\Phi _{S_t}-Id_{S_*}\big )+Id. \end{aligned}$$

Here \({{\mathcal {H}}}_*\big (\Phi _{S_t}-Id_{S_*}\big )\) is the harmonic extension of \(\Phi _{S_t}-Id_{S_*}\) satisfying

$$\begin{aligned} \left\{ \begin{array}{ll} \Delta {{\mathcal {H}}}_*\big (\Phi _{S_t}-Id_{S_*}\big )=0\qquad \hbox {in}\quad \Omega _*,\\ {{\mathcal {H}}}_*\big (\Phi _{S_t}-Id_{S_*}\big )\big |_{\Gamma _{t*}}=d_{\Gamma _t}\mu , \quad {{\mathcal {H}}}_*\big (\Phi _{S_t}-Id_{S_*}\big )\big |_{\Gamma _{b*}}=\Phi _{S_t}|_{\Gamma _{b*}}-Id_{\Gamma _{b*}}. \end{array}\right. \end{aligned}$$

6.2 Recovery of the Velocity

When the domain \(\Omega _t\) is defined by \({{\mathcal {T}}}_S\), we can define the velocity v by the free surface function \(d_{\Gamma _t}\). To begin with, the kinematic condition on \(\Gamma _t\) in \(\text{(WW) }\) is rewritten into

$$\begin{aligned} \partial _t \Phi _{S_t}\cdot (n_t\circ \Phi _{S_t})=(v\cdot N_t)\circ \Phi _{S_t}\quad \hbox {with}\quad \partial _t\Phi _{S_t}=(\partial _t d_{\Gamma _t})\mu \quad \hbox {on}\quad \Gamma _*. \end{aligned}$$

So we obtain

$$\begin{aligned} \partial _t d_{\Gamma _t}=\frac{(v\cdot n_t)\circ \Phi _{S_t}}{\mu \cdot (n_t\circ \Phi _{S_t})}\quad \hbox {i.e.} \quad v\cdot n_t=(\partial _t d_{\Gamma _t}\mu )\circ \Phi ^{-1}_{S_t}\cdot n_t. \end{aligned}$$
(6.1)

Due to the assumption that the velocity v is irrotational, we define v by

$$\begin{aligned} v=\nabla \phi \end{aligned}$$
(6.2)

with \(\phi \) satisfying

$$\begin{aligned} \left\{ \begin{array}{ll} \Delta \phi =\xi \gamma \qquad \hbox {in}\quad \Omega _t,\\ \nabla _{n_t} \phi |_{\Gamma _t}= (\partial _t d_{\Gamma _t} \,\mu )\circ \Phi ^{-1}_{S_t}\cdot n_t, \quad \nabla _{n_b} \phi |_{\Gamma _b}= 0, \end{array}\right. \end{aligned}$$
(6.3)

where

$$\begin{aligned} \gamma =|\Omega _t|^{-1}\quad \hbox { and }\quad \xi =\int _{\Gamma _t} v\cdot n_t\,{\text {d}}s=\int _{\Gamma _t}(\partial _t d_{\Gamma _t}\mu )\circ \Phi ^{-1}_{S_t}\cdot n_t{\text {d}}s. \end{aligned}$$

Moreover, we define as in [53] that

$$\begin{aligned} D_{t^*}= \partial _t+\nabla _{v^*}, \end{aligned}$$

where \(v^{*}= D\Phi _{S_t}^{-1}(v^\top \circ \Phi _{S_t} -\partial _t d_{\Gamma _t} \mu ^{\top } )\). A direct computation shows that

$$\begin{aligned} (D_t f)\circ \Phi _{S_t}=D_{t^*}(f\circ \Phi _{S_t}), \end{aligned}$$

for a function f on \(\Gamma _t\).

6.3 The Modified Formulation and the Precise Form of the Main Theorem

Before we construct the approximate solutions, we derive a new equation modified from the previous sections, which turns out to be more convenient in this part.

First, we define \({\mathfrak {K}}_a\) based on the definition of \(\mathfrak {K}\):

$$\begin{aligned} \mathfrak {K}_a={\mathfrak {K}}+\sigma a \,d_{\Gamma _t}\circ \Phi _{S_t}^{-1}= \sigma (\kappa +a \,d_{\Gamma _t}\circ \Phi _{S_t}^{-1} )-P_{v, v} \quad \hbox {on} \quad \Gamma _t \end{aligned}$$
(6.4)

for some constant \(a>0\), where \(\kappa \) can be expressed by \(d_{\Gamma _t}\) and \(P_{v,v}\) is defined by (3.1). So we have

$$\begin{aligned} {\mathfrak {J}}_a= \nabla {{\mathcal {H}}}(\mathfrak {K}_a)={\mathfrak {J}}+\sigma a\nabla {{\mathcal {H}}}(d_{\Gamma _t}\circ \Phi _{S_t}^{-1}). \end{aligned}$$

Applying (3.7) and using \({\mathfrak {K}}_a\) instead of \({\mathfrak {K}}\), we obtain

$$\begin{aligned} D_t^2 \mathfrak {K}_a+\sigma (a-\Delta _{\Gamma _t}){{\mathcal {N}}}(\mathfrak {K}_a)=R_{a, 0}, \end{aligned}$$

where \(R_{a,0}\) is defined by

$$\begin{aligned} \begin{aligned} R_{a, 0}&=\sigma R_1 -D_t^2P_{v,v}+\sigma [n_t, \Delta _{\Gamma _t}]\cdot {\mathfrak {J}}+\sigma \Delta _{\Gamma _t}\nabla P_{v,v}\cdot n_t\\&\quad +2\sigma \nabla _{\tau _t}n_t\cdot \nabla _{\tau _t}{\mathfrak {J}}+\sigma a D^2_t(d_{\Gamma _t}\circ \Phi _{S_t}^{-1})\\&\quad -\sigma ^2 a \Delta _{\Gamma _t}{{\mathcal {N}}}(d_{\Gamma _t}\circ \Phi _{S_t}^{-1})+\sigma a {{\mathcal {N}}}{\mathfrak {K}}_a. \end{aligned} \end{aligned}$$

Acting \(\nabla {{\mathcal {H}}}\) on both sides of the above equation, we get the equation of \({\mathfrak {J}}_a\):

$$\begin{aligned} D^2_t{\mathfrak {J}}_a+ \sigma \nabla {{\mathcal {H}}}\big [(a-\Delta _{\Gamma _t}){\mathfrak {J}}_a^\perp +h_v\big ]=R_a, \end{aligned}$$

where

$$\begin{aligned} R_a= \nabla {{\mathcal {H}}}(R_{a, 0}+\sigma h_v)-[\nabla {{\mathcal {H}}}, D_t^2]\mathfrak {K}_a, \end{aligned}$$

and recall that \(h_v\) is defined in (5.2) and comes from \(R_1\).

Meanwhile, the condition (4.18) at the contact points are rewritten as

$$\begin{aligned} D_t{\mathfrak {J}}_a = \pm \frac{\sigma ^2 }{\beta _c}\sin \omega _i (\nabla _{\tau _t} {\mathfrak {J}}_a)^\bot \tau _b+R_{c1, a}\quad \hbox {at}\quad p_i (i=l, r) \end{aligned}$$

with

$$\begin{aligned} R_{c1, a}=R_{c1}+\sigma aD_t\nabla {{\mathcal {H}}}(d_{\Gamma _t}\circ \Phi _{S_t}^{-1})+\sigma a \frac{\sigma ^2 }{\beta _c}(n_t\cdot \tau _b) \big (\nabla _{\tau _t} \nabla {{\mathcal {H}}}(d_{\Gamma _t}\circ \Phi _{S_t}^{-1})\big )^\perp . \end{aligned}$$

Next, we consider how to recover the free surface and the domain from \(\mathfrak {K}_a\), which is slightly different from [53]. In [53], we use the equation of \({\mathfrak {N}}_a={{\mathcal {N}}}(\kappa +a \,d_{\Gamma _t}\circ \Phi _{S_t}^{-1})\) together with the boundary information \(d_i=d_{\Gamma _t}|_{p_{i*}}\) of \(d_{\Gamma _t}\) to recover the free surface \(d_{\Gamma _t}\), so the system of \((\mathfrak {K}_a, d_l, d_r)\) is needed In this paper, we use the quantity \({\mathfrak {K}}_a\) instead of \({\mathfrak {N}}_a\), where an extra \(P_{v,v}\) is added here in (6.4). Moreover, we will need to use the equation and norms of \({\mathfrak {J}}_a\) in the iteration scheme, where \({\mathfrak {K}}_a\) can be retrieved. In fact, we have \({\mathfrak {J}}_a^\perp ={\mathfrak {J}}_a\cdot n_t={{\mathcal {N}}}{\mathfrak {K}}_a\) on \(\Gamma _t\). Therefore, to identify \({\mathfrak {K}}_a\), we look at the Neumann-boundary elliptic system of \({{\mathcal {H}}}({\mathfrak {K}}_a)\) with the compatibility condition \(\int _{\Gamma _t}{\mathfrak {J}}_a^\perp {\text {d}}s=0\). We know immediately that there exists a unique solution \({{\mathcal {H}}}({\mathfrak {K}}_a)\) up to an additive constant to this system. As a result, as long as we have \({\mathfrak {J}}_a\), we obtain \(\mathfrak {K}_a\). (One can also check Lemma 2.5 in [53].)

Consequently, to recover \(d_{\Gamma _t}\) (which is the key to recover the water-waves system), we need the system of \(({\mathfrak {K}}_a, P_{v,v}, d_l, d_r)\). As long as we have proved the existence of the solution to this system, we obtain immediately the following quantity:

$$\begin{aligned} \kappa +a \,d_{\Gamma _t}\circ \Phi _{S_t}^{-1}=\sigma ^{-1}({\mathfrak {K}}_a+P_{v,v})\quad \hbox {with the boundary information} \ d_l, d_r. \end{aligned}$$

As a result, the the desired function \(d_{\Gamma _t}\) can be solved directly from these quantities above in a similar way as in Proposition 4.2 [53], and then we can finally recover our water-waves system \(\text{(WW) }\).

Based on the analysis above, we need to give the boundary condition of \(d_{\Gamma _t}\) which is deduce from (6.1) (for more details, see (4.27) in [53]). In fact, one has the evolution equations for \(d_i(t)=d_{\Gamma _t}(p_{i*})\) (\( i=l, r\)),

$$\begin{aligned} d_i''(t)=\mathfrak {B}_i,\quad i=l, r, \end{aligned}$$
(6.5)

where

$$\begin{aligned} \begin{aligned} \mathfrak {B}_i&=-\frac{1}{\mu \cdot (n_t \circ \Phi _{S_t})} \Big (\mu \cdot (n_t \circ \Phi _{S_t})\nabla _{v^*}\partial _td_{\Gamma _t}+\nabla _{v^*}\mu \cdot (n_t\circ \Phi _{S_t})\partial _t d_{\Gamma _t}\\&\quad +\sigma \mathfrak {K}_a\circ \Phi _{S_t}\\&\quad -\sigma ^2 a {{\mathcal {N}}}(d_{\Gamma _t}\circ \Phi ^{-1}_{S_t})\circ \Phi _{S_t}+(\nabla P_{v,v}+\textbf{g})\circ \Phi _{S_t}\cdot (n_t\circ \Phi _{S_t})\Big )\big |_{p_{i*}}. \end{aligned} \end{aligned}$$

We rewrite the equation for \(P_{v,v}\) by (4.7),

$$\begin{aligned} D_tP_{v,v}=\Delta ^{-1}_N(h_p, f_p, g_p), \end{aligned}$$

where \(\Delta ^{-1}_N\) means solving the Neumann-boundary system (4.7) with \(\int _{\Omega _t}D_tP_{v,v}{\text {d}}X=0\), and the right-side functions are

$$\begin{aligned} h_p= & {} 2tr (\nabla v\cdot \nabla v\, \nabla v)+2tr \big (\nabla (\nabla P_{v,v}+{\mathfrak {J}})\nabla v\big )+ 2tr(\nabla v \nabla ^2 P_{v, v}),\quad f_p\\= & {} C'_{v,v}(t)+\nabla _{n_t} v\cdot \nabla P_{v, v}\big |_{\Gamma _t}\end{aligned}$$

and

$$\begin{aligned} g_p=-(mJ+\nabla P_{v,v}+\textbf{g})\cdot \big (\nabla n_b+(\nabla n_b)^*\big )\cdot v+v\cdot D_t(\nabla n_b)\cdot v+\nabla _{n_b} v\cdot \nabla P_{v, v}. \end{aligned}$$

As a result, we sum up the system of \(({\mathfrak {K}}_a, P_{v,v}, d_l, d_r)\) as follows:

$$\begin{aligned} \left\{ \begin{array}{ll} D_{t}^2 \mathfrak {K}_a +\sigma (a-\Delta _{\Gamma _t}){{\mathcal {N}}}(\mathfrak {K}_a)=R_{a, 0},\\ D_{t}{\mathfrak {J}}_a = \pm \sigma ^2 \beta ^{-1}_c\sin \omega (\nabla _{\tau _t} {\mathfrak {J}}_a)^\bot \tau _b +R_{c1, a}, \quad \text {at} \quad p_{i} ( i=l, r),\\ D_t P_{v,v}=\Delta ^{-1}_N(h_p, f_p, g_p),\\ \frac{d^2}{dt^2}d_i(t)=\mathfrak {B}_i,\quad i=l, r. \end{array} \right. \end{aligned}$$
(6.6)

Based on these preparations above, we are finally ready to state our precise form of Theorem 1.1. We start with introducing the space \(\Sigma \) for given \(T, L>0\):

$$\begin{aligned} \Sigma =\big \{({\mathfrak {K}}_a, P_{v,v}, d_l, d_r)\big | \Vert ({\mathfrak {K}}_a, P_{v,v}, d_l, d_r)\Vert _\Sigma \leqq L\big \} \end{aligned}$$

where the norm

$$\begin{aligned} \begin{aligned} \Vert ({\mathfrak {K}}_a, P_{v,v}, d_l, d_r)\Vert _\Sigma&\triangleq \Vert \partial _t({\mathfrak {J}}_a\circ \Phi _{S_t}) \Vert _{C([0, T]; L^2(\Omega _*))}\\&\quad +\Vert ({\mathfrak {J}}_a\circ \Phi _{S_t})\cdot n_{t*}\Vert _{C([0, T]; H^1(\Gamma _*))} \\&\quad +\Vert P_{v,v}\circ \Phi _{S_t}\Vert _{C([0, T]; H^{5/2}(\Omega _*))}\\&\quad +\sum _{i=l, r}\Big (|d_i|_{C([0, T])}+| d(d_i)/ dt |_{C([0, T])}\Big ). \end{aligned} \end{aligned}$$

Meanwhile, according to the higher-order energy \(E_h(t)\), we also define the space (for given \(L_1>0\))

$$\begin{aligned} \Sigma _h=\big \{({\mathfrak {K}}_a, P_{v,v}, d_l, d_r)\big | \Vert ({\mathfrak {K}}_a, P_{v,v}, d_l, d_r)\Vert _{\Sigma _h}\leqq L_1\big \} \end{aligned}$$

by the norm

$$\begin{aligned} \begin{aligned} \Vert ({\mathfrak {K}}_a, P_{v,v}, d_l, d_r)\Vert _{\Sigma _h}&\triangleq \Vert ({\mathfrak {K}}_a, P_{v,v}, d_l, d_r)\Vert _\Sigma \\&\quad +\Vert \partial _t\big (({\mathfrak {J}}_a\circ \Phi _{S_t}) \cdot n_{t*}\big ) \Vert _{L^\infty ([0, T]; H^1(\Gamma _*))}\\&\quad +\Vert \partial ^2_t({\mathfrak {J}}_a\circ \Phi _{S_t}) \Vert _{L^\infty ([0, T]; L^2(\Omega _*))}. \end{aligned} \end{aligned}$$

The initial data is given by

$$\begin{aligned} \left\{ \begin{array}{ll} ({\mathfrak {J}}_a\circ \Phi _{S_t})\cdot n_{t*}\big |_{t=0}=\bar{\mathfrak {J}}_{a, 0}, \quad \partial _t({\mathfrak {J}}_a\circ \Phi _{S_t})\big |_{t=0} =\bar{\mathfrak {J}}_{a, 1},\\ P_{v,v}\circ \Phi _{S_t}\big |_{t=0}={\bar{P}}_{v, v, 0}, \quad d_i(0)=d_{i, 0},\quad \frac{\text {d}}{{\text {d}}t}d_i(0)=d_{i, 1}. \end{array} \right. \end{aligned}$$

Now we can present the local well-posedness theorem.

Theorem 6.1

Assume that the initial data and initial contact angles \(\omega _{i0}\in (0, \pi /2) \) for \(i=l, r\). When the compatibility conditions (1.8) at \(t=0\) are satisfied for \(k=0,1,2,3\), there exists a unique solution \(({\mathfrak {K}}_a, P_{v,v}, d_l, d_r)\in \Sigma _h\) to system (6.6). Moreover, system (6.6) is locally well-posed with \(({\mathfrak {K}}_a, P_{v,v}, d_l, d_r)\) depending continuously on the initial data in \(\Sigma \).

6.4 Iteration Scheme

In this subsection, we present the iteration scheme. First of all, we set the initial boundary \(S_0=S_*\) without loss of generality. To simplify the notations, we denote by

$$\begin{aligned} D_{t*}=\partial _t+v^k_*\cdot \nabla ,\quad D_t=\partial _t+v^k\cdot \nabla , \end{aligned}$$

when no confusion will be made.

When we have \(({\mathfrak {K}}^{k}_a, P^{k}_{v,v}, d^{k}_l, d^{k}_r)\), the linear system of \(({\mathfrak {K}}^{k+1}_a, P^{k+1}_{v,v}, d^{k+1}_l, d^{k+1}_r)\) for the iteration scheme is set to be

$$\begin{aligned} \left\{ \begin{array}{ll} D_{t*}^2 (\mathfrak {K}^{k+1}_a\circ \Phi ^k_{S_t})+\sigma (a-\Delta _{\Gamma _t}){{\mathcal {N}}}(\mathfrak {K}^{k+1}_a)\circ \Phi ^k_{S_t}=R^k_{a, 0}\circ \Phi ^k_{S_t},\\ D_{t*}({\mathfrak {J}}^{k+1}_a\circ \Phi ^k_{S_t}) = \pm \sigma ^2\beta ^{-1}_c\sin \omega ^k_i (\nabla _{\tau ^k_t} {\mathfrak {J}}_a^{k+1})^\bot \tau _b\circ \Phi ^k_{S_t}\\ \quad +R^k_{c1, a}\circ \Phi ^k_{S_t}\quad \text {at} \quad p_{i*},\quad i=l, r,\\ D_{t*}(P^{k+1}_{v,v}\circ \Phi ^k_{S_t})=\Delta ^{-1}_N(h^k_p, f^k_p, g^k_p)\circ \Phi ^k_{S_t}\\ \frac{d^2}{dt^2}d^{k+1}_i(t)=\mathfrak {B}^k_i,\qquad i=l, r. \end{array} \right. \end{aligned}$$
(6.7)

Here we use the superscript k on \(R_{a, 0}\) (for example) to denote that all the quantities there are obtained using \(({\mathfrak {K}}^{k}_a, P^{k}_{v,v}, d^{k}_l, d^{k}_r)\).

Moreover, we point out that the velocity v in the definition of \(R^k_{a, 0}\) is given by

$$\begin{aligned} D_{t}\tilde{v}^k =-J^k-\nabla P_{v^k, v^k}-\textbf{g}, \end{aligned}$$

while in the other quantities we use \(v^k\) defined by (6.1)–(6.3). This happens due to the difference of regularities using these two definitions, which can be seen already in the previous sections.

Besides, the initial data is given by

$$\begin{aligned} \left\{ \begin{array}{ll} ({\mathfrak {J}}^{k+1}_a\circ \Phi ^k_{S_t})\cdot n^k_{t*}\big |_{t=0}=\bar{\mathfrak {J}}_{a, 0}, \quad \partial _t({\mathfrak {J}}^{k+1}_a\circ \Phi ^k_{S_t})\big |_{t=0} =\bar{\mathfrak {J}}_{a, 1},\\ P^{k+1}_{v,v}\circ \Phi ^k_{S_t}\big |_{t=0}={\bar{P}}_{v, v, 0}, \quad d^{k+1}_i(0)=d_{i, 0},\quad \frac{\text {d}}{{\text {d}}t}d^{k+1}_i(0)=d_{i, 1}. \end{array} \right. \end{aligned}$$
(6.8)

As a result, a similar proof as the proof of Proposition 5.1 in [53], we show the existence of the solution \(({\mathfrak {K}}^{k+1}_a, P^{k+1}_{v,v}, d^{k+1}_l, d^{k+1}_r)\) to the linear system (6.7)–(6.8). The details for the proof are omitted.

Proposition 6.1

Let , \(\omega _{i0}\in (0, \pi /2)\) and \(({\mathfrak {K}}^{k}_a, P^{k}_{v,v}, d^{k}_l, d^{k}_r)\) be given correspondingly. Moreover we assume that the conditions for the corner points from (6.7) hold at \(t=0\). Then there exists a small \(T>0\) such that the system (6.7)–(6.8) has a unique solution on [0, T].

6.5 Uniform Estimates

Now we are ready to give the uniform estimates for the linear system (6.7)–(6.8). To begin with, we define the energy functional for as

$$\begin{aligned} E^{k+1}(t)= E^{k+1}_{low}(t)+ E^{k+1}_{high}(t), \end{aligned}$$

where \(E^{k+1}_{low}(t)\) and \(E^{k+1}_{high}(t)\) are defined by

$$\begin{aligned} \begin{aligned} E^{k+1}_{low}(t)&=a\Vert ({\mathfrak {J}}^{k+1})^\bot \Vert ^2_{L^2(\Gamma ^k_t)}+\Vert \nabla _{\tau ^k_t} ({\mathfrak {J}}^{k+1})^\bot \Vert ^2_{L^2(\Gamma ^k_t)} \\&\quad +\Vert D_t {\mathfrak {J}}^{k+1}\Vert ^2_{L^2(\Omega ^k_t)} +\Vert P^{k+1}_{v,v}\Vert ^2_{H^{5/2}(\Omega ^k_t)}\\&\quad +\sum _{i=l, r}\big (|d^{k+1}_i(t)|^2+\Big |\frac{\text {d}}{{\text {d}}t}d^{k+1}_i(t)\Big |^2\big ), \end{aligned} \end{aligned}$$

and

$$\begin{aligned} E_{high}(t)=a\Vert D_t({\mathfrak {J}}^k)^\bot \Vert ^2_{L^2(\Gamma ^k_t)}+\Vert \nabla _{\tau ^k_t}D_t({\mathfrak {J}}^{k+1})^\bot \Vert ^2_{L^2(\Gamma ^k_t)} +\Vert D_t^2 {\mathfrak {J}}^{k+1} \Vert ^2_{L^2(\Omega ^k_t)}. \end{aligned}$$

Moreover, the dissipation \(F^{k+1}(t)\) is

$$\begin{aligned} F^{k+1}(t)= F^{k+1}_{low}(t)+ F^{k+1}_{high}(t), \end{aligned}$$

where

$$\begin{aligned}{} & {} F^{k+1}_{low}(t)=\sum _{i=l,r}\big |(\sin \omega _i^k)\nabla _{ \tau ^k_t} (\mathfrak {J}^{k+1})^\perp |_{p_i}\big |^2,\\{} & {} \quad F^{k+1}_{high}(t)=\sum _{i=l,r}\big |(\sin \omega _i^k)\nabla _{ \tau ^k_t}D_t (\mathfrak {J}^{k+1})^\perp |_{p_i}\big |^2. \end{aligned}$$

Meanwhile, we define some more auxiliary functions. Recalling from (5.25), here \({{\mathcal {P}}}^k_{t,1}\) and \({{\mathcal {P}}}^k_{t,2}\) are defined by

$$\begin{aligned} {{\mathcal {P}}}^k_{t,1}=D_t P_{\tilde{v}^k, \tilde{v}^k}-v^k\cdot (\nabla P_{\tilde{v}^k, \tilde{v}^k}|_{c}),\quad {{\mathcal {P}}}^k_{t,2}=D_t{{\mathcal {P}}}^k_{t,1}-v^k\cdot ({{\mathcal {P}}}^k_{t,1}|_{c}). \end{aligned}$$

The following proposition is our main result on the uniform estimates:

Proposition 6.2

Let \(\big (\bar{\mathfrak {J}}_{a, 0}, \bar{\mathfrak {J}}_{a, 1}, {\bar{P}}_{v, v, 0}, d_{i,0}, d_{i, 1}\big )\) and \(({\mathfrak {K}}^{k}_a, P^{k}_{v,v}, d^{k}_l, d^{k}_r)\) be given as in Proposition 6.1. Then there exists constants \(T>0\) small enough and \(A>0\) large enough such that when \(a\geqq A\), the inequality below holds

$$\begin{aligned} \sup _{t\in [0, T]} E^{k+1}(t)+\int _0^T F^{k+1}(t) dt \leqq P(E(0)). \end{aligned}$$

Proof

Since the main steps of the proof follow Theorems 4.1 and 5.1, we only present the sketch of the proof here.

First, we consider the basic energy estimates \(E^{k+1}_l\), where we only focus on the estimates for \({\mathfrak {K}}_a\) or \({\mathfrak {J}}_a\) and the other estimates follow from Lemma 4.6 and [53].

To begin with, acting \((\Phi ^k_{S_t})^{-1}\) on both sides of (6.7), one has

$$\begin{aligned} D_{t}^2 \mathfrak {K}^{k+1}_a +\sigma (a-\Delta _{\Gamma _t}){{\mathcal {N}}}(\mathfrak {K}^{k+1}_a) =R^k_{a, 0}, \end{aligned}$$

which implies that

$$\begin{aligned}{} & {} D_t^2 {\mathfrak {J}}^{k+1}_a +\sigma \nabla {{\mathcal {H}}}\big [(a-\Delta _{\Gamma _t})({\mathfrak {J}}^{k+1}_a)^\perp +h_{\tilde{v}^k}\big ] \nonumber \\{} & {} \quad = \nabla {{\mathcal {H}}}(R^k_{a, 0}+\sigma h_{\tilde{v}^k})-[\nabla {{\mathcal {H}}}, D_t^2]\mathfrak {K}^{k+1}_a. \end{aligned}$$
(6.9)

Using the same arguments as in Section 4 and Section 5, we have

$$\begin{aligned} \big \Vert \nabla {{\mathcal {H}}}(R^k_{a, 0}+\sigma h_{\tilde{v}^k})-[\nabla {{\mathcal {H}}}, D_t^2]\mathfrak {K}^{k+1}_a\big \Vert _{L^2(\Omega _t^k)}\leqq P(E^{k}(t))(1+E^{k+1}_{low}(t)). \end{aligned}$$

Next, we define \(P_{J^{k+1}, v^k}\) on \(\Omega _t^k\) by system (3.14). Taking the \(L^2(\Omega ^k_t)\) inner product of (6.9) with \(D_t{\mathfrak {J}}^{k+1}_a +\nabla P_{J^{k+1}, v^k}\), one has

$$\begin{aligned} \begin{aligned}&\int _{\Omega _t^k} \Big ( D_t ^2{\mathfrak {J}}^{k+1}_a + \nabla {{\mathcal {H}}}\big [(a-\Delta _{\Gamma ^k_t})({\mathfrak {J}}^{k+1}_a)^\perp +h_{\tilde{v}^k}\big ]\Big )\cdot (D_t {\mathfrak {J}}^{k+1}_a +\nabla P_{J^{k+1}, v^k}){\text {d}}X\\&\quad =\int _{\Omega _t^k} \big (\nabla {{\mathcal {H}}}(R^k_{a, 0}+\sigma h_{\tilde{v}^k})-[\nabla {{\mathcal {H}}}, D_t^2]\mathfrak {K}^{k+1}_a\big ) \cdot (D_t {\mathfrak {J}}^{k+1}_a +\nabla P_{J^{k+1}, v^k}){\text {d}}X. \end{aligned} \end{aligned}$$

Following the energy estimates in Section 4, there exists a constant A large enough such that when \(a\geqq A\), one concludes

$$\begin{aligned} \sup _{t\in [0, T]} E^{k+1}_{low}(t)+\int _0^T F^{k+1}_{low}(t) \leqq P(E(0))+\int _0^TP(E^k(t)). \end{aligned}$$

On the other hand, for the higher-order energy \(E_h^{k+1}(t)\), we get similarly as (5.30) the equation

$$\begin{aligned}{} & {} D_t \big [D^2_t \mathfrak {J}_a^{k+1} -\nabla {{\mathcal {H}}}({{\mathcal {P}}}^k_{t,2}) +[\nabla {{\mathcal {H}}}, D_t^2]\mathfrak {K}^{k+1}_a\big ]\\{} & {} \quad +\sigma \nabla {{\mathcal {H}}}\big ((a-\Delta _{\Gamma ^k_t}) D_t( \mathfrak {J}_a^{k+1})^\bot +D_t h_{\tilde{v}^k} \big )= R^k_{a, 1}, \end{aligned}$$

where \(R^k_{a, 1}\) is given by

$$\begin{aligned} \begin{aligned} R^k_{a, 1}&=D_t \nabla {{\mathcal {H}}}(R^k_{a,0}+\sigma h_{{\tilde{v}}_k})-D_t\nabla {{\mathcal {H}}}({{\mathcal {P}}}^k_{t,2})\\&\quad +\sigma [\nabla {{\mathcal {H}}}, D_t]{{\mathcal {H}}}\big ((a-\Delta _{\Gamma _t})({\mathfrak {J}}^{k+1}_a)^\perp +\sigma h_{\tilde{v}^k}\big ) \\&\quad -\sigma \nabla {{\mathcal {H}}}[\Delta _{\Gamma _t}, D_t] ({\mathfrak {J}}^{k+1}_a)^\perp . \end{aligned} \end{aligned}$$

Compared to the energy estimate of Theorem 5.1, the main difference lies in the term \(D_t ([\nabla {{\mathcal {H}}}, D_t^2]\mathfrak {K}^{k+1}_a)\) which contains a term like \(D_t(\partial D_t v^k)\partial {{\mathcal {H}}}(\mathfrak {K}^{k+1}_a)\). In Theorem 5.1, since v satisfies Euler’s equation, we have the estimate for \(D_t^2\nabla v\). But here in the iteration scheme, \(D_t^2\partial v^k\) acts like \(\partial \partial _t^3 d_{\Gamma _t}\), which can not be controlled by the energy. Therefore, we put \(D_t ([\nabla {{\mathcal {H}}}, D_t^2]\mathfrak {K}^{k+1}_a)\) together with \(D^2_t \mathfrak {J}_a^{k+1}\) to go back to the form \(D_t^2 \mathfrak {K}^{k+1}_a\):

$$\begin{aligned} D^2_t \mathfrak {J}_a^{k+1}+[\nabla {{\mathcal {H}}}, D_t^2]\mathfrak {K}^{k+1}_a=\nabla {{\mathcal {H}}}(D_t^2 \mathfrak {K}^{k+1}_a). \end{aligned}$$

Taking the \(L^2(\Omega ^k_t)\) inner product of the above equation with \(\nabla {{\mathcal {H}}}(D_t^2 \mathfrak {K}^{k+1}_a) -\nabla {{\mathcal {H}}}({{\mathcal {P}}}^k_{t,2}) \), we derive

$$\begin{aligned} \begin{aligned}&\int _{\Omega _t^k}D_t\big (\nabla {{\mathcal {H}}}(D_t^2 \mathfrak {K}^{k+1}_a) -\nabla {{\mathcal {H}}}({{\mathcal {P}}}^k_{t,2})\big )\cdot \big (\nabla {{\mathcal {H}}}(D_t^2 \mathfrak {K}^{k+1}_a) -\nabla {{\mathcal {H}}}({{\mathcal {P}}}^k_{t,2})\big ){\text {d}}X\\&\quad +\sigma \int _{\Omega _t^k}\nabla {{\mathcal {H}}}\big ((a-\Delta _{\Gamma ^k_t}) D_t( \mathfrak {J}_a^{k+1})^\bot -D_t h_{\tilde{v}^k} \big )\cdot \big (\nabla {{\mathcal {H}}}(D_t^2 \mathfrak {K}^{k+1}_a) -\nabla {{\mathcal {H}}}({{\mathcal {P}}}^k_{t,2})\big ){\text {d}}X\\&\quad =\int _{\Omega _t^k}R^k_{a, 1} \cdot \big (\nabla {{\mathcal {H}}}(D_t^2 \mathfrak {K}^{k+1}_a) -\nabla {{\mathcal {H}}}({{\mathcal {P}}}^k_{t,2})\big ){\text {d}}X. \end{aligned} \end{aligned}$$

For the second integral on the left side of the above equation, we have by Green’s Formula that

$$\begin{aligned} \begin{aligned}&\int _{\Omega _t^k}\nabla {{\mathcal {H}}}\big ((a-\Delta _{\Gamma ^k_t}) D_t( \mathfrak {J}_a^{k+1})^\bot + D_t h_{\tilde{v}^k} \big )\cdot \big (\nabla {{\mathcal {H}}}(D_t^2 \mathfrak {K}^{k+1}_a) -\nabla {{\mathcal {H}}}({{\mathcal {P}}}^k_{t,2})\big ){\text {d}}X\\&=\int _{\Gamma ^k_t} \big ((a-\Delta _{\Gamma ^k_t}) D_t( \mathfrak {J}_a^{k+1})^\bot +D_t h_{\tilde{v}^k} \big )\cdot {{\mathcal {N}}}(D_t^2 \mathfrak {K}^{k+1}_a-{{\mathcal {P}}}^k_{t,2}){\text {d}}s\\&=\int _{\Gamma ^k_t} \big ((a-\Delta _{\Gamma ^k_t}) D_t( \mathfrak {J}_a^{k+1})^\bot + D_t h_{\tilde{v}^k} \big )\cdot D_t^2 (\mathfrak {J}_a^{k+1})^\bot {\text {d}}s\\&\quad +\int _{\Gamma ^k_t} \big ((a-\Delta _{\Gamma ^k_t}) D_t( \mathfrak {J}_a^{k+1})^\bot + D_t h_{\tilde{v}^k} \big )\cdot \big ([{{\mathcal {N}}},D_t^2] \mathfrak {K}^{k+1}_a-{{\mathcal {N}}}({{\mathcal {P}}}^k_{t,2})\big ){\text {d}}s, \end{aligned} \end{aligned}$$

and proceeding in a similar way as before, we obtain

$$\begin{aligned} \begin{aligned}&\int _{\Omega _t^k}\nabla {{\mathcal {H}}}\big ((a-\Delta _{\Gamma ^k_t}) D_t( \mathfrak {J}_a^{k+1})^\bot + D_t h_{\tilde{v}^k} \big )\cdot \big (\nabla {{\mathcal {H}}}(D_t^2 \mathfrak {K}^{k+1}_a) -\nabla {{\mathcal {H}}}({{\mathcal {P}}}^k_{t,2})\big ){\text {d}}X\\&\quad =\frac{\text {d}}{{\text {d}}t}\Big (a\Vert D_t({\mathfrak {J}}^k)^\bot \Vert ^2_{L^2(\Gamma ^k_t)}+\big \Vert \nabla _{\tau ^k_t}D_t({\mathfrak {J}}^{k+1})^\bot \big \Vert ^2_{L^2(\Gamma ^k_t)} \Big )\\&\qquad +\nabla _{\tau ^k_t}D_t({\mathfrak {J}}^{k+1}_a)^\perp \cdot D^2_t( \mathfrak {J}_a^{k+1})^\perp \big |^{p_r}_{p_l} \\&\qquad +\int _{\Gamma ^k_t} (a-\Delta _{\Gamma ^k_t}) D_t( \mathfrak {J}_a^{k+1})^\bot \cdot \big ([{{\mathcal {N}}},D_t^2] \mathfrak {K}^{k+1}_a-{{\mathcal {N}}}({{\mathcal {P}}}^k_{t,2})\big ){\text {d}}s\\&\qquad +\int _{\Gamma ^k_t} D_t h_{\tilde{v}^k} \cdot \big (D_t^2 (\mathfrak {J}_a^{k+1})^\bot +[{{\mathcal {N}}}, D_t^2] \mathfrak {K}^{k+1}_a-{{\mathcal {N}}}({{\mathcal {P}}}^k_{t,2})\big ){\text {d}}s. \end{aligned} \end{aligned}$$

On one hand, since the commutator \([D_t,\,{{\mathcal {N}}}]\) is already expressed in (2.5), one can conclude that

$$\begin{aligned} \Vert [{{\mathcal {N}}},D_t^2] \mathfrak {K}^{k+1}_a\Vert _{H^1(\Gamma _t^k)}\leqq P(E^{k+1}(t)). \end{aligned}$$

Consequently, we have

$$\begin{aligned}{} & {} \int _{\Gamma ^k_t} \big ((a-\Delta _{\Gamma ^k_t}) D_t( \mathfrak {J}_a^{k+1})^\bot \big )\cdot \big ([{{\mathcal {N}}},D_t^2] \mathfrak {K}^{k+1}_a-{{\mathcal {N}}}({{\mathcal {P}}}^k_{t,2})\big ){\text {d}}s\\{} & {} \quad \leqq \frac{1}{8} F_h^{k+1}(t)+ P(E^{k+1}(t)). \end{aligned}$$

On the other hand, using similar arguments as in Lemma 5.5, we get

$$\begin{aligned} \begin{aligned}&\int _0^T \int _{\Gamma ^k_t} D_t h_{\tilde{v}^k} \cdot \big (D_t^2 (\mathfrak {J}_a^{k+1})^\bot +[{{\mathcal {N}}}, D_t^2] \mathfrak {K}^{k+1}_a-{{\mathcal {N}}}({{\mathcal {P}}}^k_{t,2})\big ){\text {d}}sdt\\&\quad \leqq P(E(0))+\sup _{t\in [0, T]}\big (\frac{1}{8} E_{high}^{k+1}(t)+P(E_{low}^{k+1}(t))\big )\\&\quad +\int _0^T P(E^{k}(t))dt+ \frac{1}{8} \int _0^T F_{high}^{k+1}(t)dt. \end{aligned} \end{aligned}$$

In the end, using similar arguments as in Section 5.3, we can show that when T small enough, the following estimate holds

$$\begin{aligned} \sup _{t\in [0, T]} E^{k+1}_{high}(t)+\int _0^T F^{k+1}_{high}(t)dt\leqq & {} P(E(0))+\int _0^TP(E^k(t) )dt \\{} & {} +\sup _{t\in [0, T]} P(E_{low}^{k+1}(t)). \end{aligned}$$

Therefore, by a bootstrap argument, we can prove the desired result. \(\quad \square \)

6.6 Cauchy Sequence and Going Back to \(\text{(WW) }\)

In this part, we are finally in a position to prove that the sequence of \((\mathfrak {K}^{k}_a, P^k_{v,v}, d^{k}_l, d^{k}_r)\) is indeed a Cauchy sequence. In fact, we know from the previous subsection that

$$\begin{aligned} \sup _{t\in [0, T]} E^{k+1}(t) +\int _0^T F^{k+1}(t)dt \leqq C\quad \hbox {for all} \ k\in \mathbb N, \end{aligned}$$
(6.10)

where \(C>0\) is a constant depending on E(0).

To simplify the notation, we denote by

$$\begin{aligned} \overline{f}^{k+1}=f^{k+1}\circ \Phi ^k_{S_t},\qquad \delta _{\overline{f}^k} = \overline{f}^{k+1}- \overline{f}^k,\end{aligned}$$

and

$$\begin{aligned} (a-\Delta _{\Gamma _t})g^k=h_{\tilde{v}^k}\quad \hbox {with}\quad g^k|_{p_i}=0. \end{aligned}$$

Using (6.9) and rewriting it similarly as (3.15), we have the equation of \(\delta _{\overline{{\mathfrak {J}}}^{k}_a} \):

$$\begin{aligned} D_{t*}(D_{t*}\delta _{\overline{{\mathfrak {J}}}^{k}_a}-\delta _{\overline{{\nabla {{{\mathcal {H}}}}}({{\mathcal {P}}}^k_{t,1})}})+{\sigma {{{\mathcal {A}}}}}(d_{\Gamma ^k_t})(\delta _{\overline{{\mathfrak {J}}}^{k}_a}-\delta _{\overline{g}^k}) = D_{R^k}, \end{aligned}$$
(6.11)

where we note

$$\begin{aligned} {{\mathcal {A}}}(d_{\Gamma _t})f=\big ({{\mathcal {N}}}(a-\Delta _{\Gamma _t})(f\circ \Phi ^{-1}_{S_t})\big )\circ \Phi _{S_t}, \end{aligned}$$

and

$$\begin{aligned} \begin{aligned} D_{R^k}&=\big ((\partial _t+v^k_*\cdot \nabla )^2-(\partial _t+v^{k-1}_*\cdot \nabla )^2\big )\overline{{\mathfrak {J}}}^k_a\\&\quad +\big ({{\mathcal {A}}}(d_{\Gamma ^k_t})-{{\mathcal {A}}}(d_{\Gamma ^{k-1}_t})\big )(\overline{{\mathfrak {J}}}^{k}_a-\overline{g}^k) +\delta _{ \overline{\nabla {{\mathcal {H}}}(R^{k-1}_{a, 0}+\sigma h_{\tilde{v}^{k-1}})}}\\&\quad -\delta _{ \overline{[\nabla {{\mathcal {H}}}, D_t^2]\mathfrak {K}^{k}_a}}-D_{t*}\delta _{\overline{\nabla {{\mathcal {H}}}({{\mathcal {P}}}^k_{t,1})}}. \end{aligned}\end{aligned}$$

Besides, similar but simpler equations for \((\delta _{P^k_{v,v}}, \delta _{d^k_l}, \delta _{d^k_r})\) can be derived, and we omit the details here.

Moreover, we define the energy of the difference according to the definition of \(\Sigma \) as

$$\begin{aligned} \begin{aligned} E^k_\delta (t)&= \Vert \partial _t\delta _{\overline{{\mathfrak {J}}}^{k}_a} \Vert _{C([0, T]; L^2(\Omega _*))}+a\Vert \delta _{\overline{{\mathfrak {J}}}^{k}_a}\cdot n_{t*}\Vert _{C([0, T]; L^2(\Gamma _*))}\\&\quad +\Vert \delta _{\overline{{\mathfrak {J}}}^{k}_a}\cdot n_{t*}\Vert _{C([0, T]; H^1(\Gamma _*))} \\&\quad +\Vert \delta _{P^k_{v,v}}\Vert _{C([0, T]; H^{5/2}(\Omega _*))} +\sum _{i=l, r}\big (|\delta _{d^{k}_i}|_{C([0, T])}+\Big |\delta _{d (d^{k}_i)/dt}\Big |_{C([0, T])}\big ). \end{aligned} \end{aligned}$$

Before we consider the convergence, we need to deal with \(D_{R^k}\) first.

Lemma 6.1

The right side of (6.11) satisfies the following estimate:

$$\begin{aligned} \Vert D_{R^k}\Vert _{L^2(\Omega _*)}\leqq CE^k_\delta (t) \end{aligned}$$

with the positive constant C depending on E(0).

Proof

Here, we only give the outline of the proof, and one can see similar details in [53]. First, we consider the estimate for \(\big ((\partial _t+v^k_*\cdot \nabla )^2-(\partial _t+v^{k-1}_*\cdot \nabla )^2\big )\overline{{\mathfrak {J}}}^k_a\). Using similar arguments as Proposition 4.12 and Corollary 4.13 in [53] and applying (6.7) and (6.10), we can have

$$\begin{aligned} \begin{aligned}&\big \Vert \big ((\partial _t+v^k_*\cdot \nabla )^2-(\partial _t+v^{k-1}_*\cdot \nabla )^2\big )\overline{{\mathfrak {J}}}^k_a\big \Vert _{L^2(\Omega _*)}\\&\quad \leqq C\Vert \partial _t^2\delta _{d_{\Gamma ^{k-1}_t}}\Vert _{L^2(\Gamma _*)}+C\Vert \partial _t\delta _{d_{\Gamma ^{k-1}_t}}\Vert _{H^1(\Gamma _*)} \\&\quad \leqq C \big (a^{-1} \Vert \delta _{\overline{{\mathfrak {J}}}^{k}_a}\cdot n_*\Vert _{H^1(\Gamma _*)}+C \Vert \partial _t\delta _{\overline{{\mathfrak {J}}}^{k}_a} \Vert _{L^2(\Omega _*)}\big )\leqq C E^k_\delta (t). \end{aligned} \end{aligned}$$

Second, for the term \(({{\mathcal {A}}}(d_{\Gamma _t})-{{\mathcal {A}}}(d_{\Gamma ^{k-1}_t}))(\overline{{\mathfrak {J}}}^{k}_a-\overline{g}^k)\), we notice by (5.19) that

$$\begin{aligned} \big \Vert \overline{{\mathfrak {J}}}^{k}_a-\overline{g}^k\big \Vert _{H^{5/2}(\Gamma _*)}\leqq C. \end{aligned}$$

As a result, by similar arguments as in the proof of Proposition 6.3 [53], we have

$$\begin{aligned} \big \Vert \big ({{\mathcal {A}}}(d_{\Gamma _t})-{{\mathcal {A}}}(d_{\Gamma ^{k-1}_t})\big )(\overline{{\mathfrak {J}}}^{k}_a-\overline{g}^k)\big \Vert _{L^2(\Omega _*)}\leqq CE^k_\delta (t). \end{aligned}$$

In the end, similar arguments as in Lemma 4.6 [53], we can have

$$\begin{aligned} \big \Vert \delta _{ \overline{\nabla {{\mathcal {H}}}(R^{k-1}_{a, 0}+\sigma h_{\tilde{v}^{k-1}})}}-\delta _{ \overline{[\nabla {{\mathcal {H}}}, D_t^2]\mathfrak {K}^{k}_a}}-D_{t*}\delta _{\overline{\nabla {{\mathcal {H}}}({{\mathcal {P}}}^k_{t,1})}}\big \Vert _{L^2(\Omega _*)}\leqq CE^k_\delta (t). \end{aligned}$$

Combining all these estimates above, the proof is finished. \(\quad \square \)

Now, we are able to conclude about the convergence result.

Proposition 6.3

The sequence \((\mathfrak {K}^{k}_a, P^k_{v, v}, d^{k}_l, d^{k}_r)\) is a Cauchy sequence.

Proof

We follow the steps in Theorem 4.1 and Section 6.3 in [53] to conclude that there exists a constant T small enough and A large enough such that when \(a\geqq A\), we have

$$\begin{aligned} E^k_\delta (t)\leqq C\int _0^TE^k_\delta (t)dt. \end{aligned}$$

As a result, this implies immediately that \((\mathfrak {K}^{k}_a, P^k_{v, v}, d^{k}_l, d^{k}_r)\) is convergent. \(\quad \square \)

We are finally in a position to finish the proof for Theorem 6.1.

Proof of Theorem 6.1

We only present the sketch for the proof here. In fact, since we have proved in Proposition 6.3 that \((\mathfrak {K}^{k}_a, P^k_{v, v}, d^{k}_l, d^{k}_r)\in \Sigma \) is a Cauchy sequence, we know immediately that there exists \((\mathfrak {K}_a, P_{v, v}, d_l, d_r)\in \Sigma \) satisfying

$$\begin{aligned} (\mathfrak {K}^{k}_a, P^k_{v, v}, d^{k}_l, d^{k}_r)\rightarrow (\mathfrak {K}_a, P_{v, v}, d_l, d_r)\qquad \hbox {in}\quad \Sigma . \end{aligned}$$

As a result, one can show in a standard way that \((\mathfrak {K}_a, P_{v, v}, d_l, d_r)\) satisfies system (6.6). Moreover, one also has \((\mathfrak {K}_a, P_{v, v}, d_l, d_r)\in \Sigma _h\) in the proof of Proposition 6.3. \(\quad \square \)

In the end, we go back to our water-waves system \(\text{(WW) }\). In fact, thanks to system (6.6) for \((\mathfrak {K}_a, P_{v, v}, d_l, d_r)\in \Sigma \), we derive the mean curvature \(\kappa \), which defines the free surface. Based on the knowledge of \(\Gamma _t\), we also obtain v by Section 6.2. Therefore, using similar arguments as in Section 6.4 [53] and thanks to discussions in Section 6.3, we can finally retrieve the solution (vP) to the water-waves system \(\text{(WW) }\).