1 Introduction

Let \(d\ge 2\) be an integer and \(\Omega := {\mathbb {T}}^d = ({\mathbb {R}}/{\mathbb {Z}}) ^d\). Assume that \(T>0\) and that \(U_t \subset \Omega \) is an open set with the smooth boundary \(M_t:=\partial U_t\) for any \(t \in [0,T)\). The family of the hypersurfaces \(\{ M_t \} _{t \in [0,T)}\) is called the volume preserving mean curvature flow if the normal velocity vector \(\vec {v}\) satisfies

$$\begin{aligned} \vec {v}=\vec {h} - \left( \frac{1}{{\mathscr {H}}^{d-1} (M_t)} \int _{M_t} \vec {h} \cdot \vec {\nu } \, d {\mathscr {H}}^{d-1} \right) \vec {\nu }, \quad \text {on} \ M_t, \ t \in (0,T). \end{aligned}$$
(1)

Here, \({\mathscr {H}}^{d-1}\) is the \((d-1)\)-dimensional Hausdorff measure, and \(\vec {h}\) and \(\vec {\nu }\) are the mean curvature vector and the inner unit normal vector of \(M_t\), respectively. Note that the solution \(\{M_t\} _{t \in [0,T)}\) to (1) satisfies

$$\begin{aligned} \frac{\textrm{d}}{\textrm{d}t}{\mathscr {H}}^{d-1} (M_t)\le & {} 0 \quad \text {and} \quad \nonumber \\ \frac{\textrm{d}}{\textrm{d}t} {\mathscr {L}}^{d} (U_t)= & {} - \int _{M_t} \vec {v} \cdot \vec {\nu } \, d {\mathscr {H}}^{d-1} =0, \quad t \in (0,T), \end{aligned}$$
(2)

where \({\mathscr {L}}^{d}\) is the d-dimensional Lebesgue measure. From (2), \(\{M_t\} _{t \in [0,T)}\) has the volume preserving property, that is, \({\mathscr {L}}^d (U_t)\) is constant with respect to t.

For when \(U_0\) is convex, Gage [15] and Huisken [19] proved that there exists a solution to (1) and it converges to a sphere as \(t\rightarrow \infty \). Escher and Simonett [10] showed the short time existence of the solution to (1) for smooth initial data \(M_0\) and they also proved that if \(M_0\) is sufficiently close to a sphere in the sense of the little Hölder norm \(h^{1+\alpha }\), then there exists a global solution and it converges to some sphere as \(t\rightarrow \infty \) (see also [3, 4, 29] for related results). Mugnai, Seis, and Spadaro [35] studied the minimizing movement for (1) and they proved the global existence of the flat flow, that is, there exist \(C=C(d,U_0) >0\) and a family of Caccioppoli sets \(\{ U_t \}_{t \in [0,\infty )}\) such that \({\mathscr {L}}^d (U_s \bigtriangleup U_t) \le C\sqrt{s-t}\) for any \(0\le t<s\), \({\mathscr {H}}^{d-1} (\partial ^*U_t)\) is monotone decreasing, and \({\mathscr {L}}^d (U_t)\) is constant. Here, \(\partial ^*U_t\) is the reduced boundary of \(U_t\). In addition, for \(d \le 7\), they proved the global existence of the weak solution to (1) in the sense of the distribution, under the reasonable assumption for the convergence, that is,

$$\begin{aligned} \lim _{k\rightarrow \infty } \int _0 ^T {\mathscr {H}}^{d-1} (\partial ^*U_t ^k) \, \textrm{d}t = \int _0 ^T {\mathscr {H}}^{d-1} (\partial ^*U_t) \, \textrm{d}t, \end{aligned}$$
(3)

where \(\{ U_t ^k \}_{t \in [0,T)}\) is the time-discretized approximate solution to (1). This kind of condition was introduced in [31] (see also [1, 26]). Laux and Swartz [28] showed the convergence of the thresholding schemes to the distributional BV-solutions of (1) under an assumption of the convergence similar to (3). Laux and Simon [27] also proved similar results in the case of the phase field method. On the other hand, the author [43] proved the existence of the weak solution (family of integral varifolds) to (1) in the sense of \(L^2\)-flow for \(2\le d \le 3\) without any such convergence assumption, via the phase field method studied by Golovaty [17]. Recently, Kim and Kwon [24] proved the existence of the viscosity solution to (1) for the case where \(U_0\) satisfies a geometric condition called \(\rho \)-reflection. Moreover, they also proved that the viscosity solution converges to some sphere uniformly as \(t \rightarrow \infty \).

Let \(\{\delta _i \}_{i=1} ^\infty \) be a positive sequence with \(\delta _i \rightarrow 0\) as \(i\rightarrow \infty \) and we denote \(\delta _i\) as \(\delta \) for simplicity. Suppose that \(U_t ^\delta \) is an open set with smooth boundary \(M_t ^\delta \) for any \(t \in [0,T)\). The approximate solutions studied in [24, 35] correspond to the following mean curvature flow \(\{ M_t ^\delta \}_{t \in [0,T)}\) with non-local term:

$$\begin{aligned} \vec {v} = \vec {h} - \lambda ^\delta \vec {\nu }, \quad \text {on} \ M_t ^\delta , \ t \in (0,T), \end{aligned}$$
(4)

where

$$\begin{aligned} \lambda ^\delta (t) =\frac{1}{\delta } ({\mathscr {L}}^d (U_{0} ^\delta ) -{\mathscr {L}}^d (U_t ^\delta )). \end{aligned}$$

One can check that (4) is a \(L^2\)-gradient flow of

$$\begin{aligned} E ^\delta (t)={\mathscr {H}}^{d-1} (M_t ^\delta ) +\frac{1}{2\delta }({\mathscr {L}}^d (U_{0}^\delta ) - {\mathscr {L}}^d (U_t ^\delta ))^2, \end{aligned}$$

that is,

$$\begin{aligned} \frac{\textrm{d}}{\textrm{d}t} E ^\delta (t) = - \int _{M_t ^\delta } \vert \vec {v} \vert ^2 \, d {\mathscr {H}}^{d-1} \le 0 \qquad \text {for any} \ t \in (0,T). \end{aligned}$$

Hence \(\{ M_t ^\delta \}_{t \in [0,T)}\) satisfies a relaxed volume preserving property, namely,

$$\begin{aligned} \begin{aligned} ( {\mathscr {L}}^d (U_0 ^\delta ) - {\mathscr {L}}^d (U_t ^\delta ) )^2 \le 2\delta E^\delta (t) \le 2\delta E ^\delta (0) =2\delta {\mathscr {H}}^{d-1} (M_0 ^\delta ). \end{aligned} \end{aligned}$$

Therefore \(\{ M_t ^\delta \}_{t \in [0, T)}\) converges to the solution \(\{ M_t \}_{t \in [0,T)}\) to (1) as \(\delta \rightarrow 0\) formally. Note that we cannot directly obtain the monotonicity of \({\mathscr {H}}^{d-1} (M_t)\) by the energy estimates above. However, if we have a natural energy estimate \(\sup _i \int _0 ^T \vert \lambda ^{\delta _i} (t) \vert ^2 \, dt \le C_T\) for some constant \(C_T>0\), we can expect the property in some sense, because

$$\begin{aligned} \liminf _{i\rightarrow \infty } \frac{1}{2\delta _i }({\mathscr {L}}^d (U_{0}^{\delta _i}) - {\mathscr {L}}^d (U_t ^{\delta _i}))^2 = \liminf _{i\rightarrow \infty } \frac{\delta _i}{2} \vert \lambda ^{\delta _i} (t) \vert ^2 = 0 \qquad \text {for a.e.} \ t \in [0,T), \end{aligned}$$

by Fatou’s lemma (see Proposition 10). The reason why the \(L^2\)-estimate is natural is because the non-local term of the solution to (1) satisfies it (see Proposition 19). Mugnai, Seis, and Spadaro [35] used a minimizing movement scheme corresponding to (4), and Kim and Kwon [24] used (4) to prove the existence of the viscosity solution to (1). Based on these results, in this paper we show the global existence of the weak solution to (1), via the phase field method corresponding to (4).

We denote \(W(a):= \dfrac{(1-a^2)^2}{2}\) and \(k(s) = \int _{0} ^s \sqrt{2W (a)} \, \textrm{d}a= s -\dfrac{1}{3} s^3\). Let \(\varepsilon \in (0,1)\), \(T>0\), and \(\alpha \in (0,1)\). With reference to [35] and [24], in this paper we consider the following Allen–Cahn equation with non-local term:

$$\begin{aligned} \left\{ \begin{array}{ll} \varepsilon \partial _t \varphi ^{\varepsilon } =\varepsilon \Delta \varphi ^{\varepsilon } -\dfrac{W' (\varphi ^{\varepsilon })}{\varepsilon }+ \lambda ^{\varepsilon } \sqrt{2W(\varphi ^\varepsilon )},&{}{} (x,t)\in \Omega \times (0, \infty ), \\ \varphi ^{\varepsilon } (x,0) = \varphi _0 ^{\varepsilon } (x), &{}{}x\in \Omega , \end{array} \right. \end{aligned}$$
(5)

where \(\lambda ^{\varepsilon }\) is given by

$$\begin{aligned} \lambda ^{\varepsilon }(t) =\frac{1}{\varepsilon ^\alpha } \left( \int _{\Omega } k(\varphi ^\varepsilon _0 (x))\, \textrm{d}x - \int _{\Omega } k(\varphi ^\varepsilon (x,t))\, \textrm{d}x \right) . \end{aligned}$$
(6)

Note that if \(\varphi _0 ^\varepsilon \) satisfies suitable assumptions, the standard PDE theories imply the global existence and uniqueness of the solution to (5) (see Remark 8). Set

$$\begin{aligned} \begin{aligned} E^{\varepsilon } (t) =&\, \,\int _{\Omega } \left( \frac{\varepsilon \vert \nabla \varphi ^\varepsilon (x,t) \vert ^2}{2} + \frac{W(\varphi ^\varepsilon (x,t))}{\varepsilon } \right) \, \textrm{d}x \\&\, + \frac{1}{2\varepsilon ^\alpha } \left( \int _{\Omega } k(\varphi ^\varepsilon _0 (x))\, \textrm{d}x - \int _{\Omega } k(\varphi ^\varepsilon (x,t))\, \textrm{d}x \right) ^2 \\ =:&E_S ^\varepsilon (t) + E_P ^\varepsilon (t). \end{aligned} \end{aligned}$$
(7)

As above, one can check that the solution \(\varphi ^{\varepsilon } \) to (5) satisfies

$$\begin{aligned} \frac{\textrm{d}}{\textrm{d}t} E^\varepsilon (t) = - \int _{\Omega } \varepsilon ( \partial _t \varphi ^{\varepsilon } )^2 \, \textrm{d}x \le 0 \qquad \text {for any} \ t \in (0,\infty ), \end{aligned}$$
(8)
$$\begin{aligned} E^\varepsilon (T) + \int _0 ^T \int _{\Omega } \varepsilon (\partial _t \varphi ^{\varepsilon } )^2 \, \textrm{d}x \textrm{d}t = E^\varepsilon (0) = E ^\varepsilon _S (0) \qquad \text {for any} \ T\ge 0, \end{aligned}$$
(9)

and

$$\begin{aligned} \left( \int _{\Omega } k(\varphi ^\varepsilon _0 (x))\, \textrm{d}x - \int _{\Omega } k(\varphi ^\varepsilon (x,t))\, \textrm{d}x \right) ^2 = 2 \varepsilon ^\alpha E^\varepsilon _P (t) \le 2 \varepsilon ^\alpha E ^\varepsilon _S (0) \end{aligned}$$
(10)

for any \(t \in [0,\infty )\). Assume \( \sup _{\varepsilon \in (0,1)} E^\varepsilon _S (0) <\infty \) (this assumption corresponds to \({\mathscr {H}}^{d-1} (M_0) <\infty \) for (1)). Then, we can expect that \(\varphi ^\varepsilon (x,t) \approx 1\) or \(-1\) when x is outside the neighborhood of the zero level set \(M_t ^\varepsilon =\{x \in \Omega \mid \varphi ^\varepsilon (x,t)=0 \}\) for sufficiently small \(\varepsilon \). Then we have \(\int _{\Omega } k(\varphi ^\varepsilon )\, \textrm{d}x \approx \frac{2}{3} \int _{\Omega } \varphi ^\varepsilon \, \textrm{d}x\) and thus we can regard (10) as a relaxed volume preserving property. The function \(\sqrt{2W(\varphi ^\varepsilon )}\) expresses that the non-local term is almost zero when x is outside the neighborhood of \(M_t ^\varepsilon \). In addition, \(\sqrt{2W(\varphi ^\varepsilon )}\) plays important roles in \(L^\infty \)-estimates and energy estimates (see Proposition 6 and Theorem 12).

The first main result of this paper is that there exists a global-in-time weak solution to (1) for any \(d\ge 2\) in the sense of \(L^2\)-flow, under the assumptions on the regularity of \(M_0\) (see Theorem 3). We employ (5) to construct the solution. Note that we do not require assumptions such as (3). The second main result is that, when \(M_0\) is \(C^1\) and the value \({\mathscr {H}}^{d-1} (M_0)/ ({\mathscr {L}}^d (U_0))^{\frac{d-1}{d}}\) is sufficiently close to that of a ball, there exists \(T_1 >0\) such that the flow has a unit density for a.e. \(t \in [0,T_1)\) and is also a distributional BV-solution up to \(t=T_1\) (see Theorem 4). To obtain the main results, we need to prove that the varifold \(V_t ^\varepsilon \) defined by the Modica–Mortola functional [33] converges to a integral varifold for a.e. \(t\ge 0\) (roughly speaking, the condition (3) corresponds to this convergence). For the standard Allen–Cahn equation without non-local term, this convergence was shown by Ilmanen [23] and Tonegawa [47]. Therefore we can expect the convergence for (5) if \(\lambda ^\varepsilon \) has suitable properties. In fact, \(\lambda ^\varepsilon \) can be regarded as an error term when we consider the parabolic rescaled equation of (5). We explain this more precisely. Define \({\tilde{\varphi }} ^\varepsilon (\tilde{x},{\tilde{t}}) = \varphi ^\varepsilon (\varepsilon {\tilde{x}}, \varepsilon ^2 {\tilde{t}})\). Then \({\tilde{\varphi }} ^\varepsilon \) satisfies

$$\begin{aligned} \partial _{{\tilde{t}}} {\tilde{\varphi }} ^\varepsilon = \Delta _{{\tilde{x}}} {\tilde{\varphi }} ^\varepsilon - W' ({\tilde{\varphi }} ^\varepsilon ) +\varepsilon \lambda ^\varepsilon (\varepsilon ^2 {\tilde{t}}) \sqrt{2W({\tilde{\varphi }} ^\varepsilon )}, \end{aligned}$$
(11)

where \(\Delta _{{\tilde{x}}} \) is a Laplacian with respect to \(\tilde{x}\). Assume \(\sup _{x} \vert \varphi ^\varepsilon _0 (x) \vert < 1\). Then Proposition 5 below yields \(\sup _{x,t} \vert \varphi ^\varepsilon (x,t) \vert < 1\). Thus we have

$$\begin{aligned} \sup _{{\tilde{t}} \ge 0} |\varepsilon \lambda ^\varepsilon (\varepsilon ^2 {\tilde{t}}) \sqrt{2W({\tilde{\varphi }} ^\varepsilon )} |\le \sup _{{\tilde{t}} \ge 0} |\varepsilon \lambda ^\varepsilon (\varepsilon ^2 {\tilde{t}}) |\le \frac{4}{3} {\mathscr {L}}^{d} (\Omega ) \varepsilon ^{1-\alpha } =\frac{4}{3} \varepsilon ^{1-\alpha }, \end{aligned}$$
(12)

where we used \(\max _{s \in [-1,1]} \vert k (s) \vert =\frac{2}{3}\). Therefore, broadly speaking, the non-local term \(\varepsilon \lambda ^\varepsilon (\varepsilon ^2 {\tilde{t}})\sqrt{2W({\tilde{\varphi }} ^\varepsilon )}\) is a perturbation (to the best of our knowledge, for (1), no phase field model with such a property has been known). Hence we can show the rectifiability and the integrality of the varifold \(V_t\) with arguments similar to that in [23, 47] (see also [45]). However, the proofs are not exactly the same as those, because the monotonicity formula for (5) is different from the standard one (see Proposition 8). Therefore we give the proofs in Section 4. In addition, as another good property of \(\lambda ^\varepsilon \), the \(L^2\)-norm can be controlled (see Lemma 1). This property is useful when proving the monotonicity formula and the rectifiability of \(V_t\).

In [23], to construct the weak solution to the mean curvature flow (Brakke flow), the simplest Allen–Cahn equation was considered. As generalizations of the result, the equations with external forces (see [30, 34, 44, 45]) and with Laplace-Beltrami operators (see [36, 37]) have been studied. The most well-known phase field model for (1) studied by Rubinstein and Sternberg [39] is the equation

$$\begin{aligned} \left\{ \begin{array}{ll} \varepsilon \partial _t \varphi ^{\varepsilon } =\varepsilon \Delta \varphi ^{\varepsilon } -\dfrac{W' (\varphi ^{\varepsilon })}{\varepsilon }+ \Lambda ^{\varepsilon }, &{} (x,t)\in \Omega \times (0,T), \\ \varphi ^{\varepsilon } (x,0) = \varphi _0 ^{\varepsilon } (x), &{}x\in \Omega , \end{array} \right. \end{aligned}$$
(13)

where \(\Lambda ^\varepsilon (t) = \frac{1}{{\mathscr {L}}^d (\Omega )}\int _\Omega \frac{W' (\varphi ^{\varepsilon } (x,t) )}{\varepsilon } \, dx\). As above, the solution to (13) has the volume preserving property \( \frac{d}{dt} \int _{\Omega } \varphi ^\varepsilon \, dx =0 \). Chen, Hilhorst, and Logak [8] proved that for the smooth solution \(\{M_t\} _{t \in [0,T)}\) to (1), there exists a family of functions \(\{ \varphi _0 ^{\varepsilon _i } \}_{i =1 } ^\infty \) with \(\varepsilon _i \rightarrow 0\) such that the level set \(M_t ^{\varepsilon _i} = \{ x \in \Omega \mid \varphi ^{\varepsilon _i} (x,t)=0 \}\) converges to \(M_t\), where \(\varphi ^\varepsilon \) is a solution to (13) with initial data \(\varphi ^\varepsilon _0\). In addition, as mentioned above, Laux and Simon [27] proved the convergence of the vector-valued version of (13) to the weak volume preserving multiphase mean curvature flow under an assumption that corresponds to (3). However, it is an open problem to show its convergence for (13) without such assumptions. One of the difficulties is that the boundedness of \(\sup _{\varepsilon >0} \int _0 ^T \vert \Lambda ^\varepsilon \vert ^2 \, dt\) proved by Bronsard and Stoth [7] does not immediately lead to

$$\begin{aligned} \sup _{\varepsilon >0} \int _0 ^T \int _{\Omega } \varepsilon \left( \Delta \varphi ^{\varepsilon } -\dfrac{W' (\varphi ^{\varepsilon })}{\varepsilon ^2 } \right) ^2 \,\textrm{d}x\textrm{d}t <\infty . \end{aligned}$$
(14)

Note that (14) corresponds to \(\int _0 ^T \int _{M_t} \vert \vec {h} \vert ^2 \, \textrm{d}{\mathscr {H}}^{d-1} \textrm{d}t <\infty \) of the solution to (1) and is important to show the rectifiability of the varifold (see Theorem 12 and Theorem 14). As another phase field method for (1), the study of Brassel and Bretin [6] is known (see [43, Section 1] for a comparison of these equations).

Recently, the weak-strong uniqueness of the mean curvature flow and the relationship between weak solutions have been well studied. In [12], they proved the weak-strong uniqueness of the BV solution to the multiphase mean curvature flow. More precisely, if both BV solution and strong solution with the same initial data exist, then the BV solution agrees with the strong solution while both exist. As a result strongly related to our study, Laux [25] obtained the weak-strong uniqueness for the volume preserving mean curvature flow (see Remark 5). Hensel and Laux [18] proposed a new weak solution to the mean curvature flow via varifolds and the De Giorgi type inequality. Moreover, they proved the existence of the varifold solution and the weak-strong uniqueness. In [41], Stuvard and Tonegawa showed that the multiphase Brakke flow which they construct is a \(L^2\)-flow. In addition, they proved that the flow is also a BV solution for a short time, under the suitable assumptions for initial data (therefore, the results of [12] can be used in this case).

The organization of the rest of this paper is as follows: in Section 2, we set our notations and state the main results. In Section 3, to obtain the existence theorem we prove the energy estimates and \(L^\infty \)-estimates for the solution to (5). In addition, for \(d=2\) or 3, we give a short proof for the integrality of the limit measure \(\mu _t\) constructed as a weak solution to (1). In Section 4, we show the integrality of \(\mu _t\) for any \(d \ge 2\). In Section 5, we prove the main results. In Section 6, we give some supplements for this paper.

2 Preliminaries and Main Results

2.1 Notations and definitions

For \(r>0\), \(d\in {\mathbb {N}}\), and \(x \in {\mathbb {R}}^d\), we denote \(B_r ^d (x):= \{ y \in {\mathbb {R}}^d \mid \vert x-y \vert <r \}\) (we often write this as \(B_r (x)\) for simplicity). We define \(\omega _d:= {\mathscr {L}}^d (B_1 ^d (0))\). For \(d\times d\) matrix \(A =(a_{ij})\) and \(B=(b_{ij})\), we denote the usual matrix multiplication by \(A \circ B\) and \( A\cdot B:= \sum _{i,j} a_{ij} b_{ij}. \) For \(a=(a_1,a_2,\dots , a_d) \in {\mathbb {R}}^d\), we define a \(d\times d\) matrix \(a\otimes a\) by \(a\otimes a:= (a_i a_j)\). Next we recall notations and definitions from geometric measure theory and refer to [11, 16, 40, 47] for more details. For a Caccioppoli set \(E \subset {\mathbb {R}}^d\), we denote the reduced boundary of E by \(\partial ^*E\). For the characteristic function \(\chi _{E}\), we denote the total variation measure of the distributional derivative \(\nabla \chi _E\) by \(\Vert \nabla \chi _E \Vert \). Let \(U \subset {\mathbb {R}}^d\) be an open set. We write the space of bounded variation functions on U as BV(U). For any Radon measure \(\mu \) on U and \(\phi \in C_c (U)\), we often write \(\int \phi \, d\mu \) as \(\mu (\phi )\). For \(p\ge 1\), we write \(f \in L^p (\mu )\) if f is \(\mu \)-measurable and \(\int \vert f \vert ^p \, \textrm{d}\mu <\infty \). For \(d,k \in {\mathbb {N}}\) with \(k<d\), let \({\mathbb {G}}(d,k)\) be the space of k-dimensional subspace of \({\mathbb {R}}^d\). For an open set \(U \subset {\mathbb {R}}^d\), let \(G_k (U):= U \times {\mathbb {G}} (d,k)\). We say V is a general k-varifold on U if V is a Radon measure on \(G_k (U)\). We denote the set of all general k-varifolds on U by \({\mathbb {V}} _k (U)\). For a general varifold \(V \in {\mathbb {V}} _k (U)\), we define the weight measure \(\Vert V\Vert \) by

$$\begin{aligned} \Vert V \Vert (\phi ):= \int _{G_k (U)} \phi (x) \, \textrm{d}V (x,S) \qquad \text {for any} \ \phi \in C_c (U). \end{aligned}$$

We call \(V \in {\mathbb {V}} _k (U)\) is rectifiable if there exist a \({\mathscr {H}}^k\)-measurable k-countably rectifiable set \(M \subset U\) and \(\theta \in L_{loc} ^1 ({\mathscr {H}}^{k}\lfloor _{M})\) such that \(\theta >0\) \({\mathscr {H}}^k\)-a.e. and

$$\begin{aligned} V(\phi ) =\int _M \phi (x, T_x M) \theta (x) \, \textrm{d} {\mathscr {H}}^k (x) \qquad \text {for any} \ \phi \in C_c (G_k(U)), \end{aligned}$$

where \(T_x M\) is the approximate tangent space of M at x. Note that such x exists for \({\mathscr {H}}^k\)-a.e. on M. If \(\theta \in {\mathbb {N}}\) \({\mathscr {H}}^k\)-a.e. on M, we call V is integral. In addition, if \(\theta =1\) \({\mathscr {H}}^k\)-a.e. on M, we say V has unit density.

For \(V \in {\mathbb {V}} _k (U)\) and a smooth diffeomorphism \(f:U\rightarrow U\), we define the push-forward of V by

$$\begin{aligned} f_{\#} V (\phi ):= \int _{G_ k (U)} \phi (x, \nabla f (x) \circ S) \vert \Lambda _k \nabla f (x) \circ S \vert \, d V(x,S) \end{aligned}$$
(15)

for any \(\phi \in C_c (G_k(U))\), where \(\vert \Lambda _k \nabla f (x) \circ S \vert \) is the Jacobian of the map.

For \(V \in {\mathbb {V}} _k (U)\), we define the first variation \(\delta V\) by

$$\begin{aligned} \delta V ( \vec {\phi } ):= \int _{G_k(U)} \nabla \vec \phi (x) \cdot S \, \textrm{d}V (x,S) \qquad \text {for any} \ \vec \phi \in C_c ^1 (U;{\mathbb {R}}^d). \end{aligned}$$

Here, we identify \(S \in {\mathbb {G}} (d,k)\) with the corresponding orthogonal projection of \({\mathbb {R}}^d\) onto S. When the total variation \(\Vert \delta V \Vert \) of \(\delta V\) is locally bounded and absolutely continuous with respect to \(\Vert V \Vert \), there exists a measurable vector field \(\vec h\) such that

$$\begin{aligned} \delta V (\vec \phi ) = - \int _U \vec \phi (x) \cdot \vec h(x) \, d \Vert V \Vert (x) \qquad \text {for any} \ \vec \phi \in C_c ^1 (U;{\mathbb {R}}^d). \end{aligned}$$

The vector valued function \(\vec h\) is called the generalized mean curvature vector of V. In addition, a Radon measure \(\mu \) is called k-rectifiable if there exists a k-rectifiable varifold such that \(\mu \) is represented by \(\mu = \Vert V \Vert \). Note that this V is uniquely determined, so the first variation and the generalized mean curvature vector of \(\mu \) is naturally determined by V. The definition of an integral Radon measure is determined in the same way.

The formulation of the following is similar to that of the Brakke flow [5, 48]:

Definition 1

(\(L^2\)-flow [34]) Let \(T>0\), \(U\subset {\mathbb {R}}^d\) be an open set, and \(\{\mu _t\} _{t \in [0,T)}\) be a family of Radon measures on U. Set \(d\mu := d\mu _t dt\). We call \(\{\mu _t\} _{t \in [0,T)}\) an \(L^2\)-flow with a generalized velocity vector \(\vec {v}\) if the following hold:

  1. 1.

    For a.e. \(t \in (0,T)\), \(\mu _t\) is \((d-1)\)-integral, and also has a generalized mean curvature vector \(\vec {h} \in L^2 (\mu _t; {\mathbb {R}}^d)\).

  2. 2.

    The vector field \(\vec {v}\) belongs to \(L^2 (0,T; (L^2 (\mu _t))^d)\) and

    $$\begin{aligned} \vec {v}(x,t) \perp T_x \mu _t \quad \text {for} \ \mu \text {-a.e.} \ (x,t) \in U \times (0,T), \end{aligned}$$

    where \(T_x \mu _t \in {\mathbb {G}} (d,d-1)\) is the approximate tangent space of \(\mu _t\) at x.

  3. 3.

    There exists \(C_T>0\) such that

    $$\begin{aligned} \left| \int _0 ^T \int _U (\partial _t \eta + \nabla \eta \cdot \vec {v}) \, d\mu _t dt \right| \le C_T \Vert \eta \Vert _{C^0 (U\times (0,T))} \end{aligned}$$
    (16)

    for any \(\eta \in C_c ^1 (U\times (0,T))\).

Remark 1

If there exists a family of smooth hypersurfaces \(\{M_t\}_{t \in [0,T)}\) with the normal velocity vector \(\vec {w}\), then (16) holds with \(\vec {v} =\vec {w}\) and \(\mu _t = {\mathscr {H}} ^{d-1} \lfloor _{M_t}\). In addition, if \(\vec {v}\) satisfies (16) with \(\mu _t = {\mathscr {H}} ^{d-1} \lfloor _{M_t}\), then \(\vec {v} =\vec {w}\). This proof is almost identical to the proof in [48, Proposition 2.1].

The \(L^2\)-flow has the following property:

Proposition 1

(See Proposition 3.3 of [34]) Assume that \(\{ \mu _t \} _{t \in (0,T)}\) is an \(L^2\)-flow with the generalized velocity vector \(\vec {v}\) and set \(d\mu :=d\mu _t dt\). Then

$$\begin{aligned} (\vec {v}(x_0,t_0),1) \in T_{(x_0,t_0)} \mu \end{aligned}$$

at \(\mu \)-a.e. \( (x_0,t_0) \in \Sigma (\mu )\), where \(T_{(x_0,t_0)} \mu \in {\mathbb {G}} (d+1,d)\) is the approximate tangent space of \(\mu \) at \((x_0, t_0)\) and \(\Sigma (\mu )= \{ (x,t) \mid T_{(x,t)} \mu \ \text {exists at} \ (x,t) \}\).

2.2 Assumptions for initial data

Let \(U _0 \subset \subset (0,1)^d\) be a bounded open set with the following properties:

  1. 1.

    There exists \(D_0 >0\) such that

    $$\begin{aligned} \sup _{x \in (0,1) ^d, 0<R <1} \frac{{\mathscr {H}}^{d-1} (M_0 \cap B_r (x))}{\omega _{d-1} r^{d-1}} \le D_0, \end{aligned}$$
    (17)

    where \(M_0 = \partial U_0\).

  2. 2.

    There exists a family of open sets \(\{ U_0 ^i\}_{i=1} ^\infty \) such that \(U_0 ^i\) has a \(C^3\) boundary \(M_0 ^i = \partial U_0 ^i\) for any i and it holds that

    $$\begin{aligned} \lim _{i\rightarrow \infty } {\mathscr {L}}^d (U_0 \triangle U_0 ^i)=0 \ \text{ and } \ \lim _{i\rightarrow \infty } \Vert \nabla \chi _{U_0 ^i} \Vert = \Vert \nabla \chi _{U_0} \Vert \ \text{ as } \text{ Radon } \text{ measures }.\nonumber \\ \end{aligned}$$
    (18)

Note that the second assumption is satisfied when \(U_0\) is a Caccioppoli set, and both conditions are fulfilled when \(M_0\) is \(C^1\) (see [16]).

We denote \(q^\varepsilon (r):= \tanh (r/\varepsilon )\) for \(r \in {\mathbb {R}}\). Then denoting derivatives by subscript r, \(q^\varepsilon \) satisfies

$$\begin{aligned} \frac{\varepsilon ( q ^\varepsilon _r (r)) ^2 }{2} = \frac{W(q^\varepsilon (r) )}{\varepsilon } \qquad \text {for any} \ r \in {\mathbb {R}} \end{aligned}$$
(19)

and

$$\begin{aligned} q ^\varepsilon _{rr} (r) = \frac{W' (q^\varepsilon (r) )}{\varepsilon ^2} \qquad \text {for any} \ r \in {\mathbb {R}}. \end{aligned}$$
(20)

In addition, (19) yields

$$\begin{aligned} \int _{{\mathbb {R}}} \left( \frac{\varepsilon (q ^\varepsilon _r (r)) ^2 }{2} + \frac{W(q^\varepsilon (r) )}{\varepsilon } \right) \, \textrm{d}r = \int _{{\mathbb {R}}} \sqrt{ 2W (q^\varepsilon ) } q^\varepsilon _r \, \textrm{d}r = \int _{-1} ^1 \sqrt{ 2W (q) } \, \textrm{d}q =: \sigma . \end{aligned}$$

This means that the Radon measure \(\mu _t ^\varepsilon \) defined below needs to be normalized by \(\sigma \).

Next we extend \(U_0 ^i \) and \(M_0 ^i \) periodically to \({\mathbb {R}}^d\) with period \(\Omega ={\mathbb {T}}^d\) and define

$$\begin{aligned} r _i (x) = \left\{ \begin{array}{ll} \textrm{dist}\,(x, M_0 ^i), &{} \text {if} \ x \in U_0 ^i, \\ -\textrm{dist}\,(x,M_0 ^i), &{} \text {if} \ x \not \in U_0 ^i. \end{array} \right. \end{aligned}$$

Then \(\vert \nabla r _i (x) \vert \le 1\) for a.e. \(x \in {\mathbb {R}}^d\) and there exists \(b_i >0\) such that \(r_i \) is \(C^3\) on \(N_{b_i}:= \{ x \mid \textrm{dist}\,(x, M_0 ^i)<b_i \}\) (see [9]). Let \(d_i\) be a smooth monotone non-decreasing function such that

$$\begin{aligned} d _i (r) = \left\{ \begin{array}{ll} r, &{} \text {if} \ \vert r \vert< \frac{1}{4} b_i, \\ \frac{2}{3} b_i, &{} \text {if} \ r >\frac{3}{4} b_i,\\ -\frac{2}{3} b_i, &{} \text {if} \ r <-\frac{3}{4} b_i, \end{array} \right. \end{aligned}$$

and \(\vert \frac{d}{dr}d_i \vert \le 1\). Set \(\overline{r_i}:= d_i (r_i) \). Then \(\overline{r_i} \in C^3 (\Omega )\), \(\overline{r_i} =r_i\) on \(N_{b_i/4}\), and \(\vert \nabla \overline{r_i} (x) \vert \le 1\) for any \(x \in {\mathbb {R}}^d\). Let \(\{ \varepsilon _i \}_{i=1} ^\infty \) be a positive sequence with \(\varepsilon _i \rightarrow 0\) and \(\displaystyle \frac{ \varepsilon _i }{b_i ^2} \rightarrow 0\) as \(i \rightarrow \infty \), and

$$\begin{aligned} \sup _{x \in {\mathbb {R}}^d} \vert \nabla ^{j+1} \overline{r_i} (x) \vert \le \varepsilon _i ^{-j} \qquad \text {for any} \ i \in {\mathbb {N}}, \ j=1,2. \end{aligned}$$
(21)

Note that (21) corresponds to the condition (29) below. We define a periodic function \(\varphi _0 ^{\varepsilon _i} \in C^3 (\Omega ) \) by

$$\begin{aligned} \varphi _0 ^{\varepsilon _i} (x):= q ^{\varepsilon _i} ( \overline{r_i} (x) ) =\tanh \left( \frac{d_i (r_i (x))}{\varepsilon _i} \right) \qquad \text {for any} \ i \in {\mathbb {N}}. \end{aligned}$$
(22)

We define a Radon measure \(\mu _t ^{\varepsilon _i}\) by

$$\begin{aligned} \mu _t ^{\varepsilon _i} (\phi ){:=} \frac{1}{\sigma }\int _{\Omega } \phi \left( \frac{\varepsilon _i \vert \nabla \varphi ^{\varepsilon _i} (x,t) \vert ^2}{2} {+} \frac{W(\varphi ^{\varepsilon _i} (x,t))}{\varepsilon _i} \right) \, \textrm{d}x, \quad \phi \in C_c (\Omega ), \end{aligned}$$
(23)

where \(\varphi ^{\varepsilon _i}\) is the solution to (5) with initial data \(\varphi _0 ^{\varepsilon _i}\) defined by (22) and \(\sigma = \int _{-1} ^1 \sqrt{2W(s)} \, ds\).

Remark 2

In this paper we choose a typical function \(W(a)= \frac{(1-a^2)^2}{2}\) as a double-well potential for simplicity. Since more general potentials have been considered for the convergence of the standard Allen–Cahn equation (see [23, 45, 47]), generalizations of our results regarding W can be made as well.

For \(\varphi ^{\varepsilon _i} _0\) and \(\mu _0 ^{\varepsilon _i}\), we have the following properties (see [23, p. 423] and [30, Section 5]):

Proposition 2

There exists a subsequence \(\{ \varepsilon _i \}_{i=1} ^\infty \) (denoted by the same index and the subsequence is taken only for \(\{ \varepsilon _i\}_{i=1} ^\infty \), not for \(\{ M_0 ^i\} _{i=1} ^\infty \)) such that the following hold:

  1. 1.

    For any \(i \in {\mathbb {N}}\) and \(x \in \Omega \), we have \(\displaystyle \frac{\varepsilon _i \vert \nabla \varphi ^{\varepsilon _i} _0 (x) \vert ^2 }{2} \le \frac{W(\varphi ^{\varepsilon _i} _0 (x))}{\varepsilon _i} \).

  2. 2.

    There exists \(D_1= D_1 (D_0) >0\) such that

    $$\begin{aligned} \max \left\{ \sup _{i \in {\mathbb {N}}} \mu _0 ^{\varepsilon _i} (\Omega ), \sup _{i \in {\mathbb {N}}, \ x \in \Omega , \ r \in (0,1)} \frac{\mu _0 ^{\varepsilon _i} (B_r (x))}{\omega ^{d-1} r^{d-1}} \right\} \le D_1. \end{aligned}$$
    (24)
  3. 3.

    \(\mu _0 ^{\varepsilon _i} \rightharpoonup {\mathscr {H}}^{d-1} \lfloor _{M_0}\) as Radon measures, that is,

    $$\begin{aligned} \int _{\Omega } \phi \, \textrm{d} \mu _0 ^{\varepsilon _i} \rightarrow \int _{M_0} \phi \, d{\mathscr {H}}^{d-1} \qquad \text {for any} \ \phi \in C_c(\Omega ). \end{aligned}$$
  4. 4.

    For \(\psi ^{\varepsilon _i} =\frac{1}{2} ( \varphi ^{\varepsilon _i} +1)\), \(\lim _{i\rightarrow \infty } \psi ^{\varepsilon _i} = \chi _{U_0} \) in \(L^1\) and \(\lim _{i\rightarrow \infty } \Vert \nabla \psi ^{\varepsilon _i } \Vert = \Vert \nabla \chi _{U_0} \Vert \) as Radon measures.

Remark 3

The first property (1) is obtained from \( \vert \nabla \overline{r_i} \vert \le 1\) (see the proof of Proposition 6). The assumption \(\frac{ \varepsilon _i }{b_i ^2} \rightarrow 0\) is used to show \(\int _{\Omega {\setminus } N_{b_i /4} } \left( \frac{\varepsilon _i \vert \nabla \varphi ^{\varepsilon _i} _0 \vert ^2}{2} + \frac{W(\varphi ^{\varepsilon _i} _0 )}{\varepsilon _i} \right) \, \textrm{d}x\rightarrow 0 \).

2.3 Main results

We denote the approximate velocity vector \(\vec {v} ^{\, \varepsilon _i}\) by

$$\begin{aligned} \vec {v} ^{\, \varepsilon _i} = \left\{ \begin{array}{ll} \dfrac{- \partial _t \varphi ^{\varepsilon _i} }{\vert \nabla \varphi ^{\varepsilon _i} \vert } \dfrac{\nabla \varphi ^{\varepsilon _i}}{\vert \nabla \varphi ^{\varepsilon _i} \vert }, &{} \text {if} \ \vert \nabla \varphi ^{\varepsilon _i} \vert \not =0, \\ \qquad 0, &{} \text {otherwise}. \end{array} \right. \end{aligned}$$

The first main result of this paper is

Theorem 3

Suppose that \(d\ge 2\) and \(U_0\) satisfies (17) and (18). For any \(i \in {\mathbb {N}}\), let \(\varphi _0 ^{\varepsilon _i}\) be defined so that all the claims of Proposition 2 are satisfied and \(\varphi ^{\varepsilon _i} \) be a solution to (5) with initial data \(\varphi _0 ^{\varepsilon _i}\). Then there exists a subsequence \(\{ \varepsilon _i \}_{i=1} ^\infty \) (denoted by the same index) such that the following hold:

  1. (a)

    There exist a countable subset \(B \subset [0,\infty )\) and a family of \((d-1)\)-integral Radon measures \(\{\mu _t\}_{t \in [0,\infty )}\) on \(\Omega \) such that

    $$\begin{aligned} \mu _0 = {\mathscr {H}}^{d-1} \lfloor _{M_0}, \qquad \mu _t ^{\varepsilon _i} \rightharpoonup \mu _t \ \ \ \text {as Radon measures for any} \ t\ge 0, \end{aligned}$$

    and

    $$\begin{aligned} \mu _s (\Omega ) \le \mu _t (\Omega ) \qquad \text {for any} \ s,t \in [0,\infty ) \setminus B \ \text {with} \ 0 \le t<s <\infty . \end{aligned}$$
  2. (b)

    There exists \(\psi \in BV_{loc} (\Omega \times [0,\infty )) \cap C_{loc} ^{\frac{1}{2}} ([0,\infty ); L^1 (\Omega ))\) such that the following hold:

    1. (b1)

      \(\psi ^{\varepsilon _i} \rightarrow \psi \) in \(L^1 _{loc} (\Omega \times [0,\infty ))\) and a.e. pointwise, where \(\psi ^{\varepsilon _i}=\frac{1}{2} (\varphi ^{\varepsilon _i} +1)\).

    2. (b2)

      \(\psi \vert _{t=0}=\chi _{U_0}\) a.e. on \(\Omega \).

    3. (b3)

      For any \(t \in [0,\infty )\), \(\psi (x,t) =1\) or 0 for \({\mathscr {L}}^d\)-a.e. \(x \in \Omega \) and \(\psi \) satisfies the volume preserving property, that is,

      $$\begin{aligned} \int _{\Omega } \psi (x,t) \, \textrm{d}x = {\mathscr {L}}^{d} (U_0) \qquad \text {for all} \ t \in [0,\infty ). \end{aligned}$$
    4. (b4)

      For any \(t \in [0,\infty )\) and for any \(\phi \in C_c (\Omega ; [0,\infty ))\), we have \(\Vert \nabla \psi (\cdot ,t) \Vert (\phi ) \le \mu _t (\phi )\).

  3. (c)

    For \(\lambda ^{\varepsilon _i}\) given by (6), we have

    $$\begin{aligned} \sup _{i \in {\mathbb {N}}} \int _0 ^T \vert \lambda ^{\varepsilon _i} \vert ^2 \, dt <\infty \qquad \text {for any} \ T>0 \end{aligned}$$

    and there exists \(\lambda \in L_{loc} ^2 (0,\infty )\) such that \(\lambda ^{\varepsilon _i} \rightharpoonup \lambda \) weakly in \(L^2 (0,T)\) for any \(T>0\).

  4. (d)

    There exists \(\vec {f} \in L_{loc} ^2 ([0,\infty ); (L^2 (\mu _t))^d)\) such that

    $$\begin{aligned} \begin{aligned}&\lim _{i \rightarrow \infty } \frac{1}{\sigma } \int _0 ^\infty \int _{\Omega } -\lambda ^{\varepsilon _i} \sqrt{2W (\varphi ^{\varepsilon _i})} \nabla \varphi ^{\varepsilon _i} \cdot \vec {\phi } \, \textrm{d}x\textrm{d}t \\ =&\int _{0} ^\infty \int _{\Omega } \vec {f} \cdot \vec {\phi } \, d \mu _t \textrm{d}t = \int _{0} ^\infty \int _{\Omega } -\lambda \vec {\nu }\cdot \vec {\phi } \, d \Vert \nabla \psi (\cdot ,t) \Vert \textrm{d}t \end{aligned} \end{aligned}$$
    (25)

    for any \(\vec {\phi } \in C_c (\Omega \times [0,\infty ); {\mathbb {R}}^d)\), where \(\vec \nu \) is the inner unit normal vector of \(\{ \psi (\cdot ,t) =1 \}\) on \(\textrm{spt}\,\Vert \nabla \psi (\cdot ,t) \Vert \).

  5. (e)

    The family of Radon measures \(\{ \mu _t \}_{t \in [0,\infty )}\) is an \(L^2\)-flow with a generalized velocity vector \(\vec {v} = \vec {h} +\vec {f}\), where \(\vec {h} \in L_{loc} ^2 ([0,\infty ); (L ^2 (\mu _t) )^d) \) is the generalized mean curvature vector of \(\mu _t\). Moreover, for any \(\vec {\phi } \in C_c (\Omega \times [0,\infty ); {\mathbb {R}}^d)\),

    $$\begin{aligned} \lim _{i \rightarrow \infty } \int _0 ^\infty \int _{\Omega } \vec {v}^{ \, \varepsilon _i} \cdot \vec {\phi } \, \textrm{d}\mu _t ^{\varepsilon _i}\textrm{d}t = \int _{0} ^\infty \int _{\Omega } \vec {v} \cdot \vec {\phi } \, \textrm{d} \mu _t \textrm{d}t. \end{aligned}$$
    (26)

Remark 4

From (b4), (d), and (e), we have

$$\begin{aligned} \int _{0} ^\infty \int _{\Omega } \vec {v} \cdot \vec {\phi } \, \textrm{d} \mu _t \textrm{d}t =\int _{0} ^\infty \int _{\Omega } \left( \vec {h} - \lambda \frac{d \Vert \nabla \psi (\cdot ,t) \Vert }{\textrm{d}\mu _t} \vec {\nu } \right) \cdot \vec {\phi } \, \textrm{d} \mu _t \textrm{d}t \end{aligned}$$

for any \(\vec {\phi } \in C_c (\Omega \times (0,\infty ); {\mathbb {R}}^d)\), where \(\frac{d \Vert \nabla \psi (\cdot ,t) \Vert }{d\mu _t}\) is the Radon–Nikodym derivative. Hence we have \(\vec {v} = \vec {h} -\lambda \vec {\nu }\) in the sense of \(L^2\)-flow if \(\mu _t = \Vert \nabla \psi (\cdot ,t) \Vert \) for a.e. t (from Theorem 4 below, this is correct for a short time if the initial data is sufficiently close to a ball). Since \(\mu _t\) is integral for a.e. t, for such t, we have \(\left( \frac{d \Vert \nabla \psi (\cdot ,t) \Vert }{d\mu _t} \right) ^{-1} \in {\mathbb {N}}\) for \(\mu _t\)-a.e. \(x \in \Omega \), if \(\frac{d \Vert \nabla \psi (\cdot ,t) \Vert }{d\mu _t}\not =0\).

Set \(U_t:= \{ x \in \Omega \mid \psi (x,t)=1 \}\) for \(t>0\). Let \(B \subset \subset (0,1) ^d\) be an open ball. We also show that if \(U_0 \approx B\), then there exists \(T_1 >0\) such that \(\{\partial ^*U _t\} _{t \in [0,T_1)}\) is a distributional solution to (1) in the framework of BV functions.

Theorem 4

For any \(r \in (0,\frac{1}{4})\), there exists \(\delta _1 >0\) depending only on d and r with the following properties. Assume that \( U_0 \subset (\frac{1}{4},\frac{3}{4})^d \) satisfies \({\mathscr {L}}^d (U_0)={\mathscr {L}}^d (B_{r} (0))\) and has a \(C^1\) boundary \(M_0\) with \({\mathscr {H}}^{d-1} (M_0) \le 2 {\mathscr {H}}^{d-1} (\partial B_{r} (0))\) and

$$\begin{aligned} {\mathscr {H}}^{d-1} (M_0) - d \omega _d ^{\frac{1}{d}} ({\mathscr {L}}^d (U_0))^{\frac{d-1}{d}} \le \delta _1. \end{aligned}$$
(27)

Then there exists \(T_1 =T_1(d,r,M_0) >0\) such that the following hold:

  1. (a)

    For a.e. \(t \in [0,T_1)\), \( \mu _t = \Vert \nabla \psi (\cdot ,t) \Vert = {\mathscr {H}}^{d-1} \lfloor _{\partial ^{*} U_t} \), where \(\{ \mu _t\} _{t \in [0,\infty )}\) is the \(L^2\)-flow with initial data \(\mu _0 ={\mathscr {H}}^{d-1}\lfloor _{M_0}\), given by Theorem 3.

  2. (b)

    Let \(\vec {v}\), \(\vec {h}\), \(\vec {\nu }\), and \(\lambda \) be functions given by Theorem 3. Then \(\{ \partial ^*U_t \}_{t \in [0,T_1)}\) is a distributional solution to (1) with initial data \(\partial U_0 = M_0\) in the following sense.

    1. (b1)

      For any \(t \in [0,T_1)\), \({\mathscr {L}}^d (U_t) = {\mathscr {L}}^d (U_0)\).

    2. (b2)

      For a.e. \(t\in [0,T_1)\), \(\vec {h}\) is also a generalized mean curvature vector of \({\mathscr {H}}^{d-1} \lfloor _{\partial ^*U_t}\).

    3. (b3)

      For any \(\vec {\phi } \in C_c (\Omega \times [0,T_1);{\mathbb {R}}^d)\), we have

      $$\begin{aligned} \int _0 ^{T_1} \int _{\partial ^*U_t} \{ \vec {v} -\vec {h} + \lambda \vec {\nu } \} \cdot \vec {\phi } \, \textrm{d}{\mathscr {H}} ^{d-1} \textrm{d}t =0. \end{aligned}$$
    4. (b4)

      For any \(\phi \in C_c ^1 (\Omega \times (0,T_1))\), we have

      $$\begin{aligned} \int _0 ^{T_1} \int _{U_t} \partial _t \phi \, \textrm{d}x\textrm{d}t = \int _0 ^{T_1} \int _{\partial ^*U_t} \vec {v} \cdot \vec {\nu } \phi \, \textrm{d}{\mathscr {H}}^{d-1} \textrm{d}t. \end{aligned}$$
    5. (b5)

      (Additional volume preserving property). For a.e. \(t \in [0,T_1)\), we have

      $$\begin{aligned} \int _{\partial ^*U_t} \vec {v} \cdot \vec {\nu } \, d {\mathscr {H}}^{d-1} = \int _{\Omega } \vec {v} \cdot \vec {\nu } \, d \Vert \nabla \psi (\cdot ,t)\Vert =0. \end{aligned}$$
    6. (b6)

      For a.e. \(t \in [0,T_1)\), we have

      $$\begin{aligned} \lambda (t) = \frac{1}{{\mathscr {H}}^{d-1} (\partial ^*U_t)} \int _{\partial ^*U_t} \vec {h} \cdot \vec {\nu } \, d {\mathscr {H}}^{d-1}. \end{aligned}$$

Remark 5

Recently, Laux [25] showed that if both strong solution and BV solution to the volume preserving mean curvature flow with same initial data exist, then the strong solution matches the BV solution while both exist (see also [12, 18]).

Remark 6

The isoperimetric inequality tells us that \(d \omega _d ^{\frac{1}{d}} ({\mathscr {L}}^d (U))^{\frac{d-1}{d}} \le {\mathscr {H}}^{d-1} (\partial ^*U) \) for any Caccioppoli set \(U \subset {\mathbb {R}}^d\) with \({\mathscr {L}}^d (U) <\infty \) and the equality holds if and only if there exists a ball \(B \subset {\mathbb {R}}^d\) such that \({\mathscr {L}}^d (U \triangle B )=0\) (see [13, 46] and references therein). Moreover, by the quantitative isoperimetric inequality (see [14, Theorem 1.1]) and the assumption (27), we have

$$\begin{aligned} \min \left\{ \left. \frac{{\mathscr {L}}^d (U_0 \bigtriangleup ( B_r (x) ) )}{r^d} \right| x \in {\mathbb {R}}^d \right\} \le \frac{C \sqrt{\delta _1} }{\sqrt{{\mathscr {H}}^{d-1} (B_r (0))} }, \end{aligned}$$

where \(r>0\) is a constant given by Theorem 4 and \(C>0\) depends only on d (thus, \(U_0\) needs to be close to a sphere in the above sense). On the other hand, \(M_0\) does not have to be close to a sphere in \(C^0\) (for example, \(U_0\) does not have to be connected).

Remark 7

The property (b4) claims that \(\vec v\) is a normal velocity vector in a weak sense, since

$$\begin{aligned}\frac{\textrm{d}}{\textrm{d}t} \int _{U_t} \phi \, \textrm{d}x = \int _{U_t} \partial _t \phi \, \textrm{d}x - \int _{\partial U _t} \vec v\cdot \vec \nu \phi \, d{\mathscr {H}}^{d-1} \end{aligned}$$

holds for any \(\phi \in C_c ^1 (\Omega \times (0,T_1))\) and \(t \in (0,T_1)\), where \(\{ U_t \}_{t\in [0,T_1)}\) is a family of open sets and the smooth boundary \(\partial U_t\) moves by the normal velocity vector \(\vec {v}\). By (2), we can regard (b5) as a volume preserving property in a weak sense.

3 Energy and Pointwise Estimates

In this section we show standard estimates for (5) such as the uniform \(L^2\)-estimate for \(\lambda ^\varepsilon \) and the monotonicity formula.

3.1 Assumptions

Let \(\{ \varepsilon _i \}_{i=1} ^\infty \) be a positive sequence with \(\varepsilon _i \rightarrow 0\) as \(i \rightarrow \infty \). In this section, we assume that there exist \(D_1>0\) and \(\omega >0\) such that (24) and

$$\begin{aligned} \frac{2}{3} - \left| \int _{\Omega } k(\varphi ^{\varepsilon _i} _0 (x))\, \textrm{d}x \right|> \omega >0 \end{aligned}$$
(28)

hold for any \(i \in {\mathbb {N}}\). The set \(\{ x \in \Omega \mid \varphi ^{\varepsilon _i} _0 (x) = 0 \}\) corresponds to the initial data \(M_0\) of (1), and (28) yields that \({\mathscr {L}}^d (\{ x \in \Omega \mid \varphi ^{\varepsilon _i} _0 \approx 1 \}) >0\) formally, since \(\int _{\Omega } k(\pm 1)\, dx = \pm \frac{2}{3} {\mathscr {L}}^d(\Omega ) =\pm \frac{2}{3} \). For some \({{C}_1} >0\), we also assume that the initial data \(\varphi _0 ^{\varepsilon _i}\) of the solution to (5) satisfies

$$\begin{aligned} \varphi ^{\varepsilon _i} _0 \in C^3 (\Omega ), \quad \sup _{x \in \Omega } \vert \varphi ^{\varepsilon _i} _0 (x) \vert <1, \quad \text {and} \quad \varepsilon _i ^j \sup _{x \in \Omega } \vert \nabla ^j \varphi ^{\varepsilon _i} _0 (x) \vert \le C_1 \end{aligned}$$
(29)

for any \(i \in {\mathbb {N}}\) and \(j=1,2,3\). In addition, to control the discrepancy measure \(\xi _t ^\varepsilon \) defined below, we assume

$$\begin{aligned} \frac{\varepsilon _i \vert \nabla \varphi ^{\varepsilon _i} _0 (x) \vert ^2 }{2} \le \frac{W(\varphi ^{\varepsilon _i} _0 (x))}{\varepsilon _i} \qquad \text {for any} \ x \in \Omega \ \text {and} \ i\in {\mathbb {N}}. \end{aligned}$$
(30)

Note that the function \(\varphi _0 ^{\varepsilon _i}\) defined by (22) satisfies all the assumptions above, for sufficiently large i. Throughout this paper, we often write \(\varepsilon \) as \(\varepsilon _i\) for simplicity.

3.2 Pointwise estimates

The comparison principle implies

Proposition 5

The solution \(\varphi ^\varepsilon \) to (5) with (29) satisfies

$$\begin{aligned} \vert \varphi ^\varepsilon (x,t) \vert <1,\qquad x \in \Omega , \ t\ge 0. \end{aligned}$$
(31)

Remark 8

The estimate (31) implies \(\sqrt{2W (\varphi ^\varepsilon )} =1-(\varphi ^\varepsilon )^2 \). By a priori estimates including Proposition 6 below, standard PDE theories imply the global existence and uniqueness of the classical solution to (5) with initial data \(\varphi ^\varepsilon _0\) satisfying (29).

Proof

Suppose that \(t_0:= \inf \{ t \in [0,\infty ) \mid \sup _{x \in \Omega } \varphi ^\varepsilon (x,t) \ge 1 \} <\infty \). Then \(t_0 > 0 \) since \(\sup _{x \in \Omega } \varphi ^\varepsilon _0 (x) <1 \). We may assume that there exists \(t_1 \in (t_0, \infty )\) such that \(\sup _{x \in \Omega } \varphi ^\varepsilon (x,t) \le 2\) for any \(t <t_1\). Let \(\varphi ^\varepsilon _+ \) be a solution to

$$\begin{aligned} \varepsilon \partial _t(\varphi ^{\varepsilon } _+) =\varepsilon \Delta \varphi _+ ^{\varepsilon } -\dfrac{W' (\varphi ^{\varepsilon } _+ )}{\varepsilon }+ L^\varepsilon \sqrt{2W(\varphi ^\varepsilon _+)}, \qquad (x,t) \in \Omega \times (0,t_1) \end{aligned}$$
(32)

with initial data \(\varphi ^\varepsilon _+ (x,0) = \sup _{x \in \Omega } \varphi ^\varepsilon _0 (x) \), where \(L^\varepsilon := 2 \varepsilon ^{-\alpha } \max _{ \vert s \vert \le 2} \vert k(s) \vert \). Note that \(\sup _{ t \in (0, t_1)} \vert \lambda ^\varepsilon (t) \vert \le L^\varepsilon \), where \(\lambda ^\varepsilon \) is given by the solution \(\varphi ^\varepsilon \) to (5), and this implies that \(\varphi ^{\varepsilon } _+\) is a supersolution to (5) if we regard \(\lambda ^\varepsilon \) as a given function. Since the initial data is constant and \(W'(s), \sqrt{2W(s)} \rightarrow 0\) as \(s\rightarrow 1\), one can easily check that the solution \(\varphi _+ ^\varepsilon \) to (32) depends only on t and satisfies \(\varphi _+ ^\varepsilon (t) <1\) for any \(t \in (0,t_1)\). Therefore the comparison principle implies that \(\varphi ^\varepsilon (x,t) \le \varphi _+ ^\varepsilon (t) <1\) for any \((x,t) \in \Omega \times (0,t_1)\). This yields a contradiction. Hence \(\varphi ^\varepsilon (x,t) <1\) for any \((x,t) \in \Omega \times [0,\infty )\) and the remaining inequality \(\varphi ^\varepsilon (x,t)>-1\) can be proved similarly. \(\square \)

In addition, by Proposition 5 and the maximum principle we have the following proposition (see [23]):

Proposition 6

If the solution \(\varphi ^\varepsilon \) to (5) satisfies (29) and (30), then we have

$$\begin{aligned} \frac{\varepsilon \vert \nabla \varphi ^\varepsilon (x,t) \vert ^2 }{2} \le \frac{W(\varphi ^\varepsilon (x,t))}{\varepsilon },\qquad x \in \Omega , \ t\ge 0. \end{aligned}$$
(33)

Proof

By (31), we can define a function \(r^\varepsilon \) by

$$\begin{aligned} r^\varepsilon (x,t) = ( q^\varepsilon )^{-1} (\varphi ^\varepsilon (x,t)), \qquad x \in \Omega , \ t\ge 0, \end{aligned}$$

since \(q^\varepsilon : {\mathbb {R}}\rightarrow (-1, 1)\) is one to one and surjective. Denoting the derivatives of \(q ^\varepsilon \) by subscript r, we compute that

$$\begin{aligned} \begin{aligned} \varepsilon q ^\varepsilon _r \partial _t r ^\varepsilon&= \varepsilon q ^\varepsilon _r \Delta r^\varepsilon + \varepsilon q ^\varepsilon _{rr} \vert \nabla r ^\varepsilon \vert ^2 -\dfrac{W' (q ^\varepsilon )}{\varepsilon } + \lambda ^{\varepsilon } \sqrt{2W(q ^\varepsilon )}\\&= \sqrt{2W(q ^\varepsilon )} \Delta r^\varepsilon + \dfrac{W' (q ^\varepsilon )}{\varepsilon } ( \vert \nabla r ^\varepsilon \vert ^2 -1) + \lambda ^{\varepsilon } \sqrt{2W(q ^\varepsilon )}, \end{aligned} \end{aligned}$$

where we used (19) and (20). Then we obtain

$$\begin{aligned} \partial _t r^\varepsilon = \Delta r^\varepsilon - \dfrac{2 q ^\varepsilon (r^\varepsilon ) }{\varepsilon } ( \vert \nabla r ^\varepsilon \vert ^2 -1) + \lambda ^{\varepsilon }, \end{aligned}$$
(34)

where we used \(W' (q^\varepsilon ) / \sqrt{2W (q^\varepsilon )} =-2 q^\varepsilon \). We compute

$$\begin{aligned} \frac{1}{2} \partial _t \vert \nabla r^\varepsilon \vert ^2 = \frac{1}{2} \Delta \vert \nabla r^\varepsilon \vert ^2 - \vert \nabla r^\varepsilon \vert ^2 -\nabla r ^\varepsilon \cdot \nabla \left( \dfrac{2 q ^\varepsilon (r^\varepsilon ) }{\varepsilon } ( \vert \nabla r ^\varepsilon \vert ^2 -1) \right) , \end{aligned}$$

where we used \(\nabla \lambda ^\varepsilon =0\). Set \(w^\varepsilon = \vert \nabla r ^\varepsilon \vert ^2 -1\). Then \(w^\varepsilon \) satisfies

$$\begin{aligned} \partial _t w ^\varepsilon \le \Delta w^\varepsilon - \dfrac{4 q ^\varepsilon (r^\varepsilon ) }{\varepsilon } \nabla r^\varepsilon \cdot \nabla w^\varepsilon -2\left( \nabla r^\varepsilon \cdot \nabla \dfrac{2 q ^\varepsilon (r^\varepsilon )}{\varepsilon } \right) w ^\varepsilon . \end{aligned}$$

In addition, we have \(w ^\varepsilon (x,0) \le 0\), because

$$\begin{aligned} \frac{\varepsilon \vert \nabla \varphi ^\varepsilon _0 (x) \vert ^2 }{2} - \frac{W(\varphi ^\varepsilon _0 (x))}{\varepsilon } = \frac{W(q ^\varepsilon (r^\varepsilon (x,0)) )}{\varepsilon } (\vert \nabla r ^\varepsilon (x,0) \vert ^2 -1) \le 0 \end{aligned}$$

by (30). Hence the maximum principle implies \(w^\varepsilon (x,t) \le 0\) for any \(x \in \Omega \) and \(t \in [0,\infty )\), and we obtain (33) by \(\frac{\varepsilon \vert \nabla \varphi ^\varepsilon \vert ^2 }{2} - \frac{W(\varphi ^\varepsilon )}{\varepsilon } = \frac{W(q ^\varepsilon )}{\varepsilon } w^\varepsilon \le 0. \) \(\square \)

3.3 Energy estimates

By \(E_S ^\varepsilon (t)=\sigma \mu _t ^\varepsilon (\Omega )\), (8), (9), and (24), we can easily obtain the following estimates:

Proposition 7

For any \(\varepsilon >0\) and \(T>0\), we have

$$\begin{aligned} \mu _{T} ^\varepsilon (\Omega ) + \frac{1}{\sigma } \int _{0} ^{T} \int _{\Omega } \varepsilon (\partial _t \varphi ^{\varepsilon } )^2 \, \textrm{d}x\textrm{d}t \le \mu _{0} ^\varepsilon (\Omega ) \end{aligned}$$
(35)

and

$$\begin{aligned} \sup _{\varepsilon \in (0,1)} \mu _T ^\varepsilon (\Omega ) \le \sup _{\varepsilon \in (0,1)} \mu _0 ^\varepsilon (\Omega ) \le D_1. \end{aligned}$$
(36)

Remark 9

Generally, “\(\mu _{s} ^\varepsilon (\Omega ) \le \mu _{t} ^\varepsilon (\Omega )\) for any \(0\le t<s <\infty \)” can not be shown from the energy estimates above. However, there exists a countable set B such that \(\mu _{s} (\Omega ) \le \mu _{t} (\Omega )\) holds for any \(t,s \in [0,\infty ) {\setminus } B\) with \(t<s\), where \(\mu _{t} (\Omega ) =\lim _{\varepsilon \rightarrow 0} \mu _{t} ^\varepsilon (\Omega ) \) (see Proposition 10).

Set \(D' _1:= \sup _{\varepsilon \in (0,1)} \mu _0 ^\varepsilon (\Omega )\). Note that \(D' _1 \le D_1\). By an argument similar to that in [7], we have the following lemma:

Lemma 1

There exist constants \(C_2= C_2(\omega ,d,D' _1)>0\), \(C_3= C_3(\omega ,d,D' _1)>0\), and \(\varepsilon _1 =\varepsilon _1 (\omega , d, D' _1, \alpha ) >0\) such that

$$\begin{aligned} \int _ {0} ^{T} \vert \lambda ^\varepsilon (t) \vert ^2 \, \textrm{d}t \le C_2 (\mu _0 ^\varepsilon (\Omega ) -\mu _{T} ^\varepsilon (\Omega ) + T ) \quad \text {for any} \ \varepsilon \in (0,\varepsilon _1) \ \text {and} \ T>0, \nonumber \\ \end{aligned}$$
(37)

and

$$\begin{aligned} \sup _{\varepsilon \in (0,\varepsilon _1)} \int _ {t_1} ^{t_2} \vert \lambda ^\varepsilon (t) \vert ^2 \, \textrm{d}t \le C_3 (1+t_2 -t_1) \qquad \text {for any} \ 0 \le t_1< t_2 <\infty . \end{aligned}$$
(38)

Proof

Let \(\vec {\zeta }=(\zeta ^1,\zeta ^2,\dots , \zeta ^d): \Omega \times [0,\infty ) \rightarrow {\mathbb {R}}^d\) be a smooth periodic test function. By integration by parts, we have

$$\begin{aligned} \int _{\Omega } \sqrt{2W(\varphi ^\varepsilon )} \nabla \varphi ^\varepsilon \cdot \vec {\zeta } \, \text {d}x =\int _{\Omega } \nabla k(\varphi ^\varepsilon ) \cdot \vec {\zeta } \, \text {d}x = - \int _{\Omega } k ( \varphi ^\varepsilon ) \textrm{div}\, \vec {\zeta } \, \text {d}x \end{aligned}$$
(39)

and

$$\begin{aligned} \int _\Omega \Delta \varphi ^\varepsilon \nabla \varphi ^\varepsilon \cdot \vec \zeta \, \text {d}x = -\sum _{i,j=1} ^d \int _{\Omega } \partial _{x_i} \varphi ^\varepsilon \partial _{x_j} \varphi ^\varepsilon \partial _{x_i} \zeta ^j \, \text {d}x + \int _\Omega \frac{ \vert \nabla \varphi ^\varepsilon \vert ^2}{2} \textrm{div}\, \vec \zeta \, \text {d}x.\quad \quad \end{aligned}$$
(40)

Multiply (5) by \(\nabla \varphi ^\varepsilon \cdot \vec {\zeta }\) and integrate over \(\Omega \). Then, using (39) and (40), we have

$$\begin{aligned} \begin{aligned}&\int _{\Omega } \varepsilon \partial _t \varphi ^{\varepsilon } \nabla \varphi ^\varepsilon \cdot \vec \zeta \, \text {d}x + \sum _{i,j=1} ^d \int _{\Omega } \varepsilon \partial _{x_i} \varphi ^\varepsilon \partial _{x_j} \varphi ^\varepsilon \partial _{x_i} \zeta ^j \, \text {d}x \\ {}&- \int _{\Omega } \left( \frac{\varepsilon \vert \nabla \varphi ^\varepsilon \vert ^2}{2} +\frac{W(\varphi ^\varepsilon )}{\varepsilon }\right) \textrm{div}\, \vec \zeta \, \text {d}x = -\lambda ^\varepsilon \int _{\Omega } k(\varphi ^\varepsilon ) \textrm{div}\, \vec \zeta \, dx. \end{aligned} \end{aligned}$$
(41)

The Cauchy–Schwarz inequality, (35), and (36) imply

$$\begin{aligned} \begin{aligned}&\left| \int _{\Omega } \varepsilon \partial _t \varphi ^{\varepsilon } \nabla \varphi ^\varepsilon \cdot \vec \zeta \, dx + \sum _{i,j=1} ^d \int _{\Omega } \varepsilon \partial _{x_i} \varphi ^\varepsilon \partial _{x_j} \varphi ^\varepsilon \partial _{x_i} \zeta ^j \, \text {d}x \right. \\ {}&\left. - \int _{\Omega } \left( \frac{\varepsilon \vert \nabla \varphi ^\varepsilon \vert ^2}{2} +\frac{W(\varphi ^\varepsilon )}{\varepsilon }\right) \textrm{div}\, \vec \zeta \, \text {d}x \right| \\ \le&\, C_4 \Vert \vec \zeta (\cdot , t) \Vert _{C^1 (\Omega )} \left( (D'_1) ^\frac{1}{2} \left( \int _\Omega \varepsilon (\partial _t \varphi ^{\varepsilon } )^2 \, dx \right) ^{\frac{1}{2}} + D' _1 \right) , \end{aligned} \end{aligned}$$
(42)

where \(C_4>0\) depends only on d. Let \(\eta \in C ^\infty _c (B_1 (0))\) be a smooth non-negative function with \(\int _{B_1 (0)} \eta \, dx =1\) and define the standard mollifier \(\eta _\delta \) by \(\eta _\delta (x) = \delta ^{-d} \eta (x/\delta )\) for \(\delta >0\). Let \(u=u(x,t)\) be the periodic solution to

Note that

and there exists \(C>0\) depending only on \({\mathscr {L}}^d (\Omega )\) such that

where we used \(\Vert \varphi ^\varepsilon \Vert _{L^\infty } \le 1\). Therefore the standard PDE arguments imply the existence and uniqueness of the solution u and

$$\begin{aligned} \Vert u(\cdot , t) \Vert _{C^{2,\beta } (\Omega )} \le C_5, \qquad t\ge 0, \end{aligned}$$

where \(\beta \in (0,1)\) and \(C_5>0\) depends only on \(\beta \), d, and \(\delta \). Set \(\vec \zeta (x,t) = \nabla u(x,t)\). Then, by (41) and (42), we have

$$\begin{aligned} \begin{aligned} \vert \lambda ^\varepsilon \vert \left| \int _{\Omega } k(\varphi ^\varepsilon ) \textrm{div}\, \vec \zeta \, \text {d}x \right| \le C_4 C_5 \left( (D' _1) ^\frac{1}{2} \left( \int _\Omega \varepsilon (\partial _t \varphi ^{\varepsilon } )^2 \, dx \right) ^{\frac{1}{2}} + D' _1 \right) . \end{aligned} \end{aligned}$$
(43)

We compute

(44)

By \((k(s))^2 -\frac{4}{9} \ge -W(s)\) for any \(s \in [-1,1]\), we have

$$\begin{aligned} \int _{\Omega } (k (\varphi ^\varepsilon )) ^2 -\frac{4}{9} \, dx \ge -\varepsilon \sigma \mu _t ^\varepsilon (\Omega ) \ge -\varepsilon \sigma D' _1. \end{aligned}$$
(45)

By using

$$\begin{aligned} \int _{\Omega } \vert \nabla (k(\varphi ^\varepsilon )) \vert \, \textrm{d}x = \int _{\Omega } \sqrt{2W (\varphi ^\varepsilon )} \vert \nabla \varphi ^\varepsilon \vert \, \textrm{d}x \le \sigma \mu _t ^\varepsilon (\Omega ) \le \sigma D' _1, \end{aligned}$$

\(\Vert \varphi ^\varepsilon \Vert _{L^\infty } \le 1\), and Proposition 18, we have

$$\begin{aligned} \left| \int _{\Omega } k (\varphi ^\varepsilon ) \{ k(\varphi ^\varepsilon ) *\eta _\delta -k(\varphi ^\varepsilon ) \} \, \textrm{d}x \right| \le C_6 \delta , \end{aligned}$$
(46)

and

$$\begin{aligned} \left| \frac{1}{{\mathscr {L}}^d (\Omega )} \int _{\Omega } k (\varphi ^\varepsilon ) \, \textrm{d}x \left( \int _{\Omega } k (\varphi ^\varepsilon ) \, \textrm{d}x -\int _{\Omega } k (\varphi ^\varepsilon ) *\eta _\delta \, \textrm{d}x \right) \right| \le C_6 \delta , \end{aligned}$$
(47)

where \(C_6 >0\) depends only on \(D' _1\) and \({\mathscr {L}}^d (\Omega )\). Set \(\delta = \frac{\omega ^2}{4C_6 {\mathscr {L}}^d (\Omega )}\). By (10), (28), (44), (45), (46), and (47), there exists \(\varepsilon _1 >0\) depending only on \(\alpha \), \(D' _1\), \({\mathscr {L}}^d (\Omega )\), and \(\omega \) such that

$$\begin{aligned} \begin{aligned}&-\int _{\Omega } k(\varphi ^\varepsilon ) \textrm{div}\, \vec \zeta \, \text {d}x \\ \ge&\, \frac{4}{9} {\mathscr {L}}^d (\Omega ) -\frac{1}{{\mathscr {L}}^d (\Omega )} \left( \int _{\Omega } k (\varphi ^\varepsilon ) \, \text {d}x \right) ^2 -\varepsilon \sigma D' _1 -2 C_6 \delta \\ \ge&\, \frac{1}{{\mathscr {L}}^d (\Omega )} \left( \omega ^2 -\frac{4\sqrt{2}}{3} \varepsilon ^{\frac{\alpha }{2}} (D' _1)^{\frac{1}{2}} \right) -\varepsilon \sigma D' _1 -2 C_6 \delta \\ \ge&\, \frac{1}{4{\mathscr {L}}^d (\Omega )} \omega ^2 \end{aligned} \end{aligned}$$
(48)

holds for any \(\varepsilon \in (0,\varepsilon _1)\), where we used \((\int _{\Omega } k (\varphi _0 ^\varepsilon ) \, \textrm{d}x)^2 -(\int _{\Omega } k (\varphi ^\varepsilon ) \, dx)^2 \le \frac{4\sqrt{2}}{3} \varepsilon ^{\frac{\alpha }{2}} (D' _1)^{\frac{1}{2}}\) by (10). From (35), (36), (43), and (48), we obtain (37) and (38). \(\square \)

Remark 10

For the classical solution to the volume preserving mean curvature flow, we can obtain a similar estimate for the non-local term (see Proposition 19).

We define the discrepancy measure \(\xi _t ^\varepsilon \) on \(\Omega \) by

$$\begin{aligned} \xi _t ^\varepsilon ( \phi ):= \frac{1}{\sigma } \int _{\Omega } \phi (x) \xi _\varepsilon (x,t) \, \textrm{d}x, \qquad \phi \in C_c (\Omega ), \end{aligned}$$
(49)

where

$$\begin{aligned} \xi _\varepsilon (x,t) = \frac{\varepsilon \vert \nabla \varphi ^\varepsilon (x,t) \vert ^2 }{2} - \frac{W(\varphi ^\varepsilon (x,t))}{\varepsilon }. \end{aligned}$$

Proposition 6 implies the following lemma.

Lemma 2

Assume (30). Then \(\xi _\varepsilon (x,t) \le 0\) for any \( (x,t) \in \Omega \times [0,\infty )\). In addition, \(\xi _t ^\varepsilon \) is a non-positive measure for any \(t \ge 0\).

We denote the backward heat kernel \(\rho =\rho _{(y,s)} (x,t)\) by

$$\begin{aligned} \rho _{(y,s)} (x,t) = \frac{1}{(4\pi (s-t)) ^{\frac{d-1}{2}}} e ^{-\frac{\vert x-y\vert ^2}{4(s-t)}}, \qquad x,y \in {\mathbb {R}}^d, \ 0\le t <s. \end{aligned}$$

With exactly the same proof as in [43, p. 2028], we obtain the following estimates similar to the monotonicity formula obtained by Huisken [20] and Ilmanen [23] (for convenience, we call thus the monotonicity formula)

Proposition 8

(See [43]) Let \(\xi _\varepsilon (x,0) \le 0\) for any \( x \in \Omega \). Assume (30). Then

$$\begin{aligned} \begin{aligned}&\frac{\textrm{d}}{\textrm{d}t} \int _{{\mathbb {R}}^d} \rho _{(y,s)} (x,t) \, \textrm{d}\mu _t ^\varepsilon (x) \\ \le&\, \frac{1}{2(s-t)} \int _{{\mathbb {R}}^d} \rho _{(y,s)} (x,t) \, \textrm{d}\xi _t ^\varepsilon (x) + \frac{1}{2} (\lambda ^\varepsilon )^2 \int _{{\mathbb {R}}^d} \rho _{(y,s)} (x,t) \, \textrm{d}\mu _t ^\varepsilon (x) \end{aligned} \end{aligned}$$
(50)

holds for any \(0\le t<s<\infty \) and for any \(y \in {\mathbb {R}}^d\). Here, \(\mu _t ^\varepsilon \) and \(\xi ^\varepsilon _t\) are extended periodically to \({\mathbb {R}}^d\). In addition, we have

$$\begin{aligned} \begin{aligned} \int _{{\mathbb {R}}^d} \rho _{(y,s)} (x,t) \, d\mu _t ^\varepsilon (x) \Big \vert _{t=t_2} \le&\, \left( \int _{{\mathbb {R}}^d} \rho _{(y,s)} (x,t) \, \textrm{d}\mu _t ^\varepsilon (x) \Big \vert _{t=t_1} \right) e^{\frac{1}{2} \int _{t_1} ^{t_2} \vert \lambda ^{\varepsilon } \vert ^2 \, \textrm{d}t } \\ \le&\, \left( \int _{{\mathbb {R}}^d} \rho _{(y,s)} (x,t) \, \textrm{d}\mu _t ^\varepsilon (x) \Big \vert _{t=t_1} \right) e^{C_3 (t_2 -t_1 +1)} \end{aligned} \end{aligned}$$
(51)

for any \(y \in {\mathbb {R}}^d\), \(0\le t_1< t_2 <\infty \), and \(\varepsilon \in (0,\varepsilon _1)\).

Remark 11

Ilmanen [23] proved the monotone decreasing of \(\int _{{\mathbb {R}}^d} \rho _{(y,s)} (x,t) \, \textrm{d}\mu _t ^\varepsilon (x)\) with respect to t, for the solution to the Allen–Cahn equation without the non-local term under suitable assumptions. In general, one can show that the Brakke flow with smooth initial data has unit density for a short time by using the monotonicity formula see [45]). However, in order to show a similar conclusion for our problem, it is necessary that \(\mu _0 ^\varepsilon (\Omega ) -\mu _ {t} ^\varepsilon (\Omega )\) is small enough, due to (37) (see Lemma 12 below).

As a corollary of the monotonicity formula, we can obtain the following upper bounds of the densities of \(\mu _t ^\varepsilon \):

Corollary 1

(See [23, 42]) There exists \(0<D_2 <\infty \) depending only on d, \(C_3\), \(D_1\), and T such that

$$\begin{aligned} \mu _t ^\varepsilon (B_R (y)) \le D_2 R^{d-1} \end{aligned}$$
(52)

for all \(y \in {\mathbb {R}}^d\), \(R \in (0,1)\), \(\varepsilon \in (0,\varepsilon _1)\), and \(t \in [0, T]\).

Proof

Using \( \int _0 ^1 \left( \log \frac{1}{k} \right) ^{\frac{d-1}{2}} \, dk =\Gamma (\frac{d-1}{2} +1) =\pi ^{\frac{d-1}{2}} /\omega _{d-1} \) and the same calculation as (155) below, we have

$$\begin{aligned} \int _{{\mathbb {R}}^d} \rho _{(y,s)} (x,0) \, \textrm{d} \mu _0 ^\varepsilon \le \frac{D_1 \omega _{d-1}}{\pi ^{\frac{d-1}{2}}} \int _0 ^1 \left( \log \frac{1}{k} \right) ^{\frac{d-1}{2}} \, \textrm{d}k =D_1 \end{aligned}$$
(53)

for any \(s>0\) and \(y \in {\mathbb {R}}^d\). By (51) and (53),

$$\begin{aligned} \int _{{\mathbb {R}}^d} \rho _{(y,s)} (x,t) \, \textrm{d} \mu _t ^\varepsilon \le D_1 e^{C_3 (T+1)} \end{aligned}$$
(54)

for any \(t \in [0,T)\) with \(0<t<s\) and \(y \in {\mathbb {R}}^d\). Set \(R=2\sqrt{s-t}\). We compute

$$\begin{aligned} \int _{{\mathbb {R}}^d} \rho _{(y,s)} (x,t) \, \textrm{d} \mu _t ^\varepsilon =\frac{1}{\pi ^{\frac{d-1}{2} } R^{d-1}} \int _{{\mathbb {R}}^d} e^{- \frac{\vert x-y \vert ^2}{R^2}} \, \textrm{d} \mu _t ^\varepsilon \ge \frac{1}{\pi ^{\frac{d-1}{2} } R^{d-1}} \int _{B_R (y)} e^{- 1} \, \textrm{d} \mu _t ^\varepsilon . \end{aligned}$$
(55)

Therefore we have (52) by (54) and (55). \(\square \)

By integration by parts, we have the following estimate:

Lemma 3

For any non-negative test function \(\phi \in C_c ^2 (\Omega )\), there exists \(C_7 >0\) depending only on \(D_1\), \(\Vert \phi \Vert _{C^2 (\Omega )}\), \(\omega \), and d such that

$$\begin{aligned} \int _0 ^T \left| \frac{\textrm{d}}{\textrm{d}t} \mu _t ^\varepsilon (\phi ) \right| \, \textrm{d}t \le C_7(1+T) \qquad \text {for any} \ \varepsilon \in (0,\varepsilon _1) \ \text {and} \ T>0. \end{aligned}$$
(56)

Proof

By integration by parts, we have

$$\begin{aligned} \begin{aligned}&\frac{\textrm{d}}{\textrm{d}t} \int _{\Omega } \phi \left( \frac{\varepsilon \vert \nabla \varphi ^\varepsilon \vert ^2}{2} + \frac{W(\varphi ^\varepsilon )}{\varepsilon } \right) \, \textrm{d}x \\ =&\, - \int _{\Omega } \varepsilon \phi (\partial _t \varphi ^{\varepsilon } ) ^2 \, \textrm{d}x +\lambda ^\varepsilon \int _{\Omega } \phi \sqrt{2 W (\varphi ^\varepsilon )} \partial _t \varphi ^{\varepsilon } \, \textrm{d}x - \int _{\Omega } \varepsilon (\nabla \phi \cdot \nabla \varphi ^\varepsilon ) \partial _t \varphi ^{\varepsilon } \, \textrm{d}x. \end{aligned} \end{aligned}$$
(57)

By Cauchy’s mean-value theorem, there exists \(c=c(d) >0\) such that \(\sup _{\phi >0} \frac{\vert \nabla \phi \vert ^2}{\phi } \le c(d) \Vert \nabla ^2 \phi \Vert _{C^0 (\Omega )}\). Hence,

$$\begin{aligned} \begin{aligned}&\left| \frac{\textrm{d}}{\textrm{d}t} \int _{\Omega } \phi \left( \frac{\varepsilon \vert \nabla \varphi ^\varepsilon \vert ^2}{2} + \frac{W(\varphi ^\varepsilon )}{\varepsilon } \right) \, \textrm{d}x \right| \\ \le&\, \vert \lambda ^\varepsilon \vert ^2 \int _{\Omega } \phi \frac{2W(\varphi ^\varepsilon )}{\varepsilon } \, \textrm{d}x + \int _{\Omega } \frac{\vert \nabla \phi \vert ^2}{\phi } \varepsilon \vert \nabla \varphi ^\varepsilon \vert ^2 \, \textrm{d}x + 2 \int _{\Omega } \varepsilon \phi ( \partial _t \varphi ^{\varepsilon } ) ^2 \, \textrm{d}x \\ \le&\, C \sigma D_1 \left( 1+ \vert \lambda ^\varepsilon \vert ^2 + \int _{\Omega } \varepsilon ( \partial _t \varphi ^{\varepsilon } ) ^2 \, \textrm{d}x \right) , \end{aligned} \end{aligned}$$
(58)

where \(C>0\) depends only on \(\Vert \phi \Vert _{C^2 (\Omega )}\) and d. Thus (35), (38), and (58) imply (56). \(\square \)

With an argument similar to that of [23], we can show the following proposition by using (56):

Proposition 9

There exist a subsequence \(\{ \varepsilon _{i_j} \}_{j=1} ^\infty \) and a family of Radon measures \(\{ \mu _t \}_{t\ge 0}\) such that

$$\begin{aligned} \mu _t ^{\varepsilon _{i_j}} \rightharpoonup \mu _t \qquad \text {as Radon measures on} \ \Omega \end{aligned}$$
(59)

for any \(t \in [0,\infty )\) and for any \(d\ge 2\). In addition, there exists a countable set \(B \subset [0,\infty )\) such that \(\mu _t (\Omega )\) is continuous on \([0,\infty ) {\setminus } B\).

Proof

Let \(\{ \phi _k \} _{k=1} ^\infty \subset C_c (\Omega ) \) be a dense subset with \(\phi _k \in C_c ^2 (\Omega )\) for any k, and for any \(x \in \Omega \cap {\mathbb {Q}}^d\) and \(r \in (0,1) \cap {\mathbb {Q}}\), there exists \(k \in {\mathbb {N}}\) such that \(\phi _k \in C_c ^2 (B_r (x))\). Let \(f _k ^i (t)=\mu _0 ^{\varepsilon _i} (\phi _k) + \int _0 ^t \left( \frac{d}{ds} \mu _s ^{\varepsilon _i} (\phi _k) \right) _+ \, ds\) and \( g _k ^i (t) = \int _0 ^t \left( \frac{d}{ds} \mu _s ^{\varepsilon _i} (\phi _k) \right) _- \, ds \). Then

$$\begin{aligned} \mu _t ^{\varepsilon _i} (\phi _k) = f_k ^i (t) - g_k ^i (t), \end{aligned}$$

and \(f _k ^i (t)\) and \(g _k ^i (t)\) are non-decreasing functions with

$$\begin{aligned} 0 \le f_k ^i (t) \le C _k \qquad \text {and} \qquad 0 \le g_k ^i (t) \le C _k \end{aligned}$$

for any i and \(t \in [0,T)\), where \(C_k >0\) depends only on \(D_1\), \(\Vert \phi _k \Vert _{C^2 (\Omega )}\), \(\omega \), and d, by (56). Then Helly’s selection theorem implies that there exist a subsequence \(\varepsilon _i \rightarrow 0\) (denoted by the same index), \(f_k,g_k:[0,T) \rightarrow [0,\infty )\) such that \(\lim _{i\rightarrow \infty } f_k ^i (t) = f_k (t)\) and \(\lim _{i\rightarrow \infty } g_k ^i (t) = g_k (t)\) for any \(t \in [0,T)\). Therefore we have

$$\begin{aligned} \lim _{i \rightarrow \infty } \mu _t ^{\varepsilon _i} (\phi _k) = f_k (t) - g_k (t) \end{aligned}$$
(60)

for any \(t \in [0,T)\). By this and the diagonal argument, we can choose a subsequence such that (60) holds for any \(t \in [0,T)\) and \(k \in {\mathbb {N}}\). On the other hand, for any \(t \in [0,T)\), the compactness of Radon measures yields that there exist \(\mu _t \) and a subsequence \(\varepsilon _i \rightarrow 0\) (depending on t) such that \(\mu _t ^{\varepsilon _i} \rightharpoonup \mu _t\) as Radon measures. However, \(\mu _t\) is uniquely determined by (60). Hence we obtain (59) for any \(t \in [0,T)\). By the diagonal argument with \(T\rightarrow \infty \), we have (59) for any \(t \in [0,\infty )\).

From a similar argument as to that above, there exists monotone increasing functions f and g such that \(\mu _t (\Omega ) = f(t) - g(t)\). By the monotonicity, there exists a countable set B such that f and g are continuous on \([0,\infty ) \setminus B\). This concludes the proof. \(\square \)

Proposition 10

Let B be the countable set given by Proposition 9. For any \(t,s \in [0,\infty ) {\setminus } B\) with \(t<s\), we have \(\mu _s (\Omega ) \le \mu _t (\Omega )\).

Proof

From Proposition 9, we may assume that \(\mu _t ^{\varepsilon _i} (\Omega ) \rightarrow \mu _t (\Omega )\) for any \(t \in [0,\infty )\). We recall that \(E_S ^{\varepsilon _i}\) and \(E_P ^{\varepsilon _i}\) are energies defined by (7). By (8) and \(E ^{\varepsilon _i} (t) \le E_S ^{\varepsilon _i} (0) \le D_1\), Helly’s selection theorem yields that there exist a subsequence \(\varepsilon _i \rightarrow 0\) (denoted by the same index) and a monotone decreasing function E(t) such that \(E^{\varepsilon _i} (t) \rightarrow E (t)\) for any \(t \in [0,\infty )\). For any \(T>0\), the estimate (38) and Fatou’s lemma imply

$$\begin{aligned} \int _0 ^T \liminf _{i\rightarrow \infty } E_P ^{\varepsilon _i} (t) \, \textrm{d}t \le \liminf _{i\rightarrow \infty } \int _0 ^T E_P ^{\varepsilon _i} (t) \, \textrm{d}t = \liminf _{i\rightarrow \infty } \int _0 ^T \frac{\varepsilon _i ^\alpha }{2} \vert \lambda ^{\varepsilon _i} \vert ^2 \, \textrm{d}t=0. \end{aligned}$$

Therefore \(\liminf _{i\rightarrow \infty } E_P ^{\varepsilon _i} (t)=0\) a.e. \(t \ge 0\) and hence \( E(t) = \sigma \mu _t (\Omega ) \) for a.e. \(t\ge 0\). By this, the monotonicity of E(t), and the continuity of \(\mu _t (\Omega ) \) on \([0,\infty ) {\setminus } B\), we obtain the claim. \(\square \)

We define a Radon measure \(\mu \) on \(\Omega \times [0,\infty )\) by \(d\mu :=d\mu _t dt\). By the boundedness of \(\sup _i \mu _t ^{\varepsilon _i} (\Omega )\), the dominated convergence theorem implies

$$\begin{aligned} \lim _{i\rightarrow \infty } \int _0 ^T \int _{\Omega } \phi \, \textrm{d}\mu _t ^{\varepsilon _i} dt = \int _{\Omega \times [0,T)} \phi \, \textrm{d}\mu \qquad \text {for any} \ \phi \in C_c (\Omega \times [0,T)). \end{aligned}$$

For our measures \(\mu \) and \(\mu _t\), we have the following property:

Proposition 11

There exists a countable set \({\tilde{B}} \subset [0,\infty )\) such that

$$\begin{aligned} \textrm{spt}\,\mu _t \subset \{ x \in \Omega \mid (x,t) \in \textrm{spt}\,\mu \} \end{aligned}$$
(61)

for any \(t \in (0,\infty ) \setminus {\tilde{B}}\).

Proof

Let \(f_k\) and \(g_k\) be monotone increase functions given by Proposition 9. Then there exists a countable set \({\tilde{B}}\) such that \(f_k\) and \(g_k\) are continuous on \([0,\infty ) {\setminus } {\tilde{B}}\) for any k. Suppose that there exists \(t_0 \in [0,\infty ) {\setminus } {\tilde{B}}\) such that \(x \in \textrm{spt}\,\mu _{t_0}\) and \((x,t_0) \not \in \textrm{spt}\,\mu \). Then we may assume that there exists k such that \(x \in \textrm{spt}\,\phi _k \) and \(\mu (\phi _k \times (t_0 -\delta , t_0 +\delta ))=0\) for sufficiently small \(\delta >0\), where \(\phi _k\) is a function given by Proposition 9. From \(x \in \textrm{spt}\,\mu _{t_0}\), \(\mu _{t_0 } (\phi _k)>0\) and there exists \(\delta ' >0\) such that \( \mu _{t } (\phi _k)>0\) for any \(t \in (t_0 -\delta ', t_0 +\delta ')\) by the continuity of \(f_k\) and \(g_k\). However, this contradicts \(\mu (\phi _k \times (t_0 -\delta , t_0 +\delta ))=0\). Therefore we obtain (61) for \(t \in [0,\infty ) {\setminus } {\tilde{B}}\). \(\square \)

3.4 Integrality of \(\mu _t\) for \(d \le 3\)

In the case of \(d\le 3\), we can use the results of [38]. For \(d\ge 4\), we employ the arguments of [23, 32, 45] in Section 4 below.

Theorem 12

Assume that \(d=2\) or 3 and (59). Then \(\mu _t\) is integral for a.e. \(t \ge 0\).

Proof

The estimates (35), (36), and (38) imply

$$\begin{aligned} \begin{aligned}&\, \int _0 ^T \int _{\Omega } \varepsilon \left( \Delta \varphi ^{\varepsilon } -\dfrac{W' (\varphi ^{\varepsilon })}{\varepsilon ^2 } \right) ^2 \,\textrm{d}x\textrm{d}t \\ \le&\, \int _0 ^T \int _{\Omega } \varepsilon ( \partial _t \varphi ^{\varepsilon } )^2 \,\textrm{d}x\textrm{d}t + \int _0 ^T \vert \lambda ^\varepsilon \vert ^2 \int _{\Omega } \dfrac{2 W (\varphi ^{\varepsilon })}{\varepsilon } \,\textrm{d}x\textrm{d}t \\ \le&\, \sigma \mu _0 ^\varepsilon (\Omega ) + 2 D_1 C_3 (1+T) \le \sigma D_1 + 2 D_1 C_3 (1+T) \end{aligned} \end{aligned}$$
(62)

for any \(T>0\). Then Fatou’s lemma yields

$$\begin{aligned} \begin{aligned}&\int _0 ^T \liminf _{i\rightarrow \infty } \int _{\Omega } \varepsilon _i \left( \Delta \varphi ^{\varepsilon _i} -\dfrac{W' (\varphi ^{\varepsilon _i})}{\varepsilon ^2 _i } \right) ^2 \,\textrm{d}x \textrm{d}t \\ \le&\, \liminf _{i\rightarrow \infty } \int _0 ^T \int _{\Omega } \varepsilon _i \left( \Delta \varphi ^{\varepsilon _i} -\dfrac{W' (\varphi ^{\varepsilon _i})}{\varepsilon ^2 _i } \right) ^2 \,\textrm{d}x\textrm{d}t <\infty . \end{aligned} \end{aligned}$$

Therefore

$$\begin{aligned} \liminf _{i\rightarrow \infty } \int _{\Omega } \varepsilon _i \left( \Delta \varphi ^{\varepsilon _i} -\dfrac{W' (\varphi ^{\varepsilon _i})}{\varepsilon ^2 _i } \right) ^2 \,dx < \infty \qquad \text {for a.e. } \ t\ge 0. \end{aligned}$$

By this, \(2\le d \le 3\), and (36), \(\mu _t\) is integral for a.e. \(t \ge 0\) (see [38, Theorem 5.1]). \(\square \)

4 Rectifiability and Integrality of \(\mu _t\)

We already proved the rectifiability and integrality of \(\mu _t\) with \(d\le 3\) in Theorem 12. Next we consider the case of \(d\ge 2\) and basically follow [23, 32, 45].

4.1 Assumptions

We assume (24) and (2830) again in this section. Let \(\{ \varepsilon _i \}_{i=1} ^\infty \) be a positive sequence such that \(\varepsilon _i \rightarrow 0 \) as \(i \rightarrow \infty \). By the weak compactness of the Radon measures and Proposition 9, we may assume that there exist Radon measures \(\mu \), \(\vert \xi \vert \) and a family of Radon measures \(\{ \mu _t\} _{t \in [0,T)}\) such that

$$\begin{aligned} \mu ( \phi ) = \lim _{i \rightarrow \infty } \int _ 0 ^T \mu _t ^{\varepsilon _i} (\phi ) \, \textrm{d}t, \qquad \vert \xi \vert ( \phi ) = \lim _{i \rightarrow \infty } \int _ 0 ^T \int _{\Omega } \sigma ^{-1} \vert \xi _{\varepsilon _i} \vert \phi \, \textrm{d}x\textrm{d}t \end{aligned}$$

for \(\phi \in C_c (\Omega \times (0,T))\), and

$$\begin{aligned} \mu _t ( \phi ) = \lim _{i \rightarrow \infty } \mu _t ^{\varepsilon _i} (\phi ), \qquad \phi \in C_c (\Omega ), \ t \in [0,T). \end{aligned}$$

Remark 12

In the discussion above, we proved that there exists \(\mu _t = \lim _{\varepsilon \rightarrow 0} \mu _t ^\varepsilon \) for any \(t\ge 0\), however such a property does not necessarily hold for \(\xi _t ^\varepsilon \), because we do not know anything about the monotonicity of \(\xi _t ^\varepsilon \) (which was the key to the argument for \(\mu _t ^\varepsilon \)).

By the standard PDE theories and the rescaling arguments, we obtain the following lemma. The proof is almost the same as [45, Lemma 4.1]. So, we skip this.

Lemma 4

There exists \(C_8>0\) depending only on d and \(C_1\) such that

$$\begin{aligned} \sup _{\Omega \times [0,T) } \varepsilon \vert \nabla \varphi ^\varepsilon \vert + \sup _{x,y \in \Omega , \ t \in [0,T)} \frac{\varepsilon ^{\frac{3}{2}} \vert \nabla \varphi ^\varepsilon (x,t) - \nabla \varphi ^\varepsilon (y,t) \vert }{\vert x-y \vert ^{\frac{1}{2}}} \le C_8 \end{aligned}$$
(63)

for any \(\varepsilon \in (0,1)\).

4.2 Vanishing of \(\xi \)

First we show \(\vert \xi \vert =0\) for any \(d \ge 2\).

Lemma 5

Assume \((x',t') \in \textrm{spt}\,\mu \) and \(\alpha _1 \in (0,1)\). Then there exist a sequence \(\{ (x_j,t_j) \}_{j=1} ^\infty \) and a subsequence \(\{ \varepsilon _{i_j} \}_{j=1} ^\infty \) such that \(\vert (x_j,t_j)-(x', t') \vert < \frac{1}{j}\) and \(\vert \varphi ^{\varepsilon _{i_j}} (x_j,t_j) \vert <\alpha _1 \) for all j.

Proof

Define \(Q_r = \overline{ B_r (x') \times (t' -r, t' +r)}\) for \(r>0\). If the claim is not true, then there are \(r>0\) and \(N>1\) such that \(\inf _{Q_r} \vert \varphi ^{\varepsilon _i } \vert \ge \alpha _1 \) for any \(i >N\). Without loss of generality, we may assume that \(\inf _{Q_r} \varphi ^{\varepsilon _i } \ge \alpha _1 \) for any \(i >N\). For \(s \in [\alpha _1,1)\), we have \(W(s) = \frac{1}{4\,s}W'(s) (s^2 -1) \le \frac{1}{4\alpha _1 }W'(s) (s^2 -1)\). Note that \( W' (s) (s^2 -1) \ge 0\) for \(s \in [\alpha _1,1)\). Assume that \(\phi \in C_c ^\infty (B_r (x'))\) satisfies \(0 \le \phi \le 1\) and \(\phi =1 \) on \(B_{r/2} (x')\). We compute

$$\begin{aligned} \begin{aligned} \int _{Q_r} \phi ^2 \frac{W(\varphi ^\varepsilon )}{\varepsilon ^2} \, \textrm{d}x \textrm{d}t&\le \, \frac{1}{4\alpha _1} \int _{Q_r} \phi ^2 \frac{W'(\varphi ^\varepsilon )}{\varepsilon ^2} ( (\varphi ^\varepsilon )^2 - 1 ) \, \textrm{d}x \textrm{d}t \\&= \, \frac{1}{4\alpha _1} \int _{Q_r} \phi ^2 \left( -\partial _t \varphi ^{\varepsilon } +\Delta \varphi ^{\varepsilon } + \lambda ^{\varepsilon } \frac{\sqrt{2W(\varphi ^\varepsilon )}}{\varepsilon } \right) \\&\qquad \times ( (\varphi ^\varepsilon )^2 - 1 ) \, \textrm{d}x \textrm{d}t. \end{aligned} \end{aligned}$$

Now we estimate the three terms on the right hand side above. We compute

$$\begin{aligned} \begin{aligned} \left| \int _{Q_r} \phi ^2 \partial _t \varphi ^{\varepsilon } ( (\varphi ^\varepsilon )^2 - 1 ) \, \textrm{d}x \textrm{d}t\right| = \left| \int _{t'- r} ^{t' +r} \frac{\textrm{d}}{\textrm{d}t} \int _{B_r (x')} \phi ^2 \left( \frac{1}{3} (\varphi ^\varepsilon )^3 - \varphi ^\varepsilon \right) \, \textrm{d}x \textrm{d}t\right| \le C, \end{aligned} \end{aligned}$$

where \(C>0\) depends only on r. Here we used \(\Vert \varphi ^\varepsilon \Vert _{L^\infty } \le 1\) and \(0\le \phi \le 1\). By \(\inf _{Q_r} \varphi ^{\varepsilon _i } \ge \alpha _1 \), integration by parts, and Young’s inequality,

$$\begin{aligned} \begin{aligned}&\int _{Q_r} \phi ^2 \Delta \varphi ^\varepsilon ( (\varphi ^\varepsilon )^2 - 1 ) \, \textrm{d}x \textrm{d}t \\ =&\, \int _{Q_r} -2\phi (\nabla \phi \cdot \nabla \varphi ^\varepsilon ) ( (\varphi ^\varepsilon )^2 - 1 ) -2 \phi ^2 \vert \nabla \varphi ^\varepsilon \vert ^2 \varphi ^\varepsilon \, \textrm{d}x \textrm{d}t \\ \le&\, \int _{Q_r} \alpha _1 \phi ^2 \vert \nabla \varphi ^\varepsilon \vert ^2 + \frac{1}{\alpha _1} \vert \nabla \phi \vert ^2 ( (\varphi ^\varepsilon )^2 - 1 )^2 -2 \phi ^2 \vert \nabla \varphi ^\varepsilon \vert ^2 \alpha _1 \, \textrm{d}x \textrm{d}t \\ \le&\, \frac{1}{\alpha _1} \int _{Q_r} \vert \nabla \phi \vert ^2 \, \textrm{d}x \textrm{d}t \le C, \end{aligned} \end{aligned}$$

where \(C>0\) depends only on \(\alpha _1\), r, and \(\Vert \nabla \phi \Vert _{L^\infty }\). Here we used \(0\le \phi \le 1\). By (36), (38), and \(\sqrt{2W (\varphi ^\varepsilon )} ( (\varphi ^\varepsilon )^2 - 1 ) =-2 W (\varphi ^\varepsilon ) \),

$$\begin{aligned} \begin{aligned}&\left| \int _{Q_r} \phi ^2 \lambda ^{\varepsilon } \frac{\sqrt{2W(\varphi ^\varepsilon )}}{\varepsilon } ( (\varphi ^\varepsilon )^2 - 1 ) \, \textrm{d}x \textrm{d}t\right| \le \int _{Q_r} \phi ^2 \vert \lambda ^{\varepsilon } \vert \frac{2W(\varphi ^\varepsilon )}{\varepsilon } \, \textrm{d}x \textrm{d}t \\ \le&\, \int _{t' -r} ^{t' +r} \vert \lambda ^{\varepsilon } \vert \int _{B_r (x')} \frac{2W(\varphi ^\varepsilon )}{\varepsilon } \, \textrm{d}x \textrm{d}t \\ \le&\, 2 \sigma D_1 \sqrt{2r} \left( \int _{t' -r} ^{t' +r} \vert \lambda ^{\varepsilon } \vert ^2 \, \textrm{d}t\right) ^{\frac{1}{2}} \le 2 \sigma D_1 \sqrt{2r} C_3 ^{\frac{1}{2}} (1+ r)^{\frac{1}{2}}. \end{aligned} \end{aligned}$$

Therefore there exists \(C>0\) depending only on \(\alpha _1\), r, \(C_3\), \(\Vert \nabla \phi \Vert _{L^\infty }\), and \(D_1\) such that

$$\begin{aligned} \int _{Q_r} \phi ^2 \frac{W(\varphi ^\varepsilon )}{\varepsilon ^2} \, \textrm{d}x \textrm{d}t \le C. \end{aligned}$$

By (33), \(\mu _t ^\varepsilon (B_{r/2} (x')) \le 2\sigma ^{-1} \int _{B_{r/2} (x')} \frac{W}{\varepsilon } \, \textrm{d}x\). Thus

$$\begin{aligned} \int _{t' -r} ^{t'+r} \mu _t ^\varepsilon (B_{r/2} (x')) \, \textrm{d}t \le 2 \sigma ^{-1} \int _{Q_r} \phi ^2 \frac{W(\varphi ^\varepsilon )}{\varepsilon } \, \textrm{d}x \textrm{d}t \le 2 \sigma ^{-1} \varepsilon C, \end{aligned}$$

where \(C>0\) depends only on \(\alpha _1\), r, \(C_3\), \(\Vert \nabla \phi \Vert _{L^\infty }\), and \(D_1\). However, this implies \((x',t') \not \in \textrm{spt}\,\mu \). This is a contradiction. \(\square \)

Set

$$\begin{aligned} \rho _{y} ^r (x):= \frac{1}{( \sqrt{2 \pi } r ) ^{d-1} }e^{-\frac{\vert x-y \vert ^2}{2 r^2}}, \qquad r>0, \ x,y\in {\mathbb {R}}^d. \end{aligned}$$
(64)

Note that \(\rho _{(y,s)}(x,t)=\rho _y ^r (x)=\rho _x ^r (y)\) for \(r=\sqrt{2(s-t)}\).

Lemma 6

There exist \(\gamma _1, \eta _1, \eta _2 \in (0,1)\) depending only on d, W, T, \(D_2\), and \(C_3\) such that the following hold. For \(t,s \in [0,T/2)\) with \(0<s-t\le \eta _1\), we denote \(r = \sqrt{2(s-t)}\) and \(t' = s+ r^2/2\). If \(x \in \Omega \) satisfies

$$\begin{aligned} \int _{{\mathbb {R}}^d } \rho ^r _{x} (y) \, \textrm{d} \mu _s (y) = \int _{{\mathbb {R}}^d } \rho _{(x,t')} (y,s) \, \textrm{d} \mu _s (y) <\eta _2, \end{aligned}$$
(65)

then \((B_{\gamma _1 r} (x) \times \{ t' \}) \cap \textrm{spt}\,\mu = \emptyset \).

Proof

First we remark that \(0\le t< s<t' <T\), \(s=\frac{t+t'}{2}\), and \(r=\sqrt{2(s-t)}=\sqrt{2(t'-s)}\). Assume that \(x \in \Omega \) satisfies (65), \((x',t') \in \textrm{spt}\,\mu \), and \(x' \in B_{\gamma _1 r} (x)\). We choose \(\gamma _1\), \(\eta _1\), and \(\eta _2\) later. Let \(\alpha _1 \in (0,1)\) be a constant. By Lemma 5, there exist a sequence \(\{ (x_j,t_j) \}_{j=1} ^\infty \) and a subsequence \(\varepsilon _j \rightarrow 0\) such that \(\lim _{j\rightarrow \infty } (x_j,t_j) = (x',t' )\) and \( \vert \varphi ^{\varepsilon _j} (x_j,t_j) \vert <\alpha _1\) for all j. Then we may assume that for \(\alpha ' = (\alpha _1 + 1)/2 >\alpha _1\), there exists \(\gamma _2 =\gamma _2 (W,\alpha _1) >0\) such that \( \frac{W(\varphi ^{\varepsilon _j} (y,t_j) )}{\varepsilon _j} \ge \frac{W(\alpha ')}{\varepsilon _j} \) for any j and for any \(y \in B_{\gamma _2 \varepsilon _j } (x_j)\), because \(W(\alpha _1) >W(\alpha ')\) and

$$\begin{aligned} \vert \varphi ^{\varepsilon _j} (y,t_j) -\varphi ^{\varepsilon _j} (x_j,t_j) \vert\le & {} \sup _{z \in \Omega } \Vert \nabla \varphi ^{\varepsilon _j} (z,t_j) \Vert \vert y-x_j \vert \\\le & {} \varepsilon _j ^{-1} W(0) \vert y-x_j \vert \le W(0) \gamma _2 \end{aligned}$$

for any \(y \in B_{\gamma _2 \varepsilon _j } (x_j)\), where we used (33). Thus, there exists \(\eta _3 = \eta _3 (d, \gamma _2)>0\) such that

$$\begin{aligned} \begin{aligned} \eta _3 \le&\, \int _{B_{\gamma _2 \varepsilon _j (x_j)}} \frac{W(\alpha ')}{\varepsilon _j} \rho _{(x_j, t_j +\varepsilon _j ^2)} (y,t_j) \, \textrm{d}y \\ \le&\, \int _{B_{\gamma _2 \varepsilon _j (x_j)}} \frac{W(\varphi ^{\varepsilon _j} (y,t_j) )}{\varepsilon _j} \rho _{(x_j, t_j +\varepsilon _j ^2)} (y,t_j) \, \textrm{d}y. \end{aligned} \end{aligned}$$

Here we used

$$\begin{aligned} \inf _{y \in B_{\gamma _2 \varepsilon } (x_j)} \rho _{x_j, t_j +\varepsilon _j ^2} (y, t_j)> C_9 \varepsilon _j ^{1-d} >0, \end{aligned}$$

where \(C_9>0\) depends only on d and \(\gamma _2\). By the monotonicity formula, we have

$$\begin{aligned} \begin{aligned} \eta _3 \le \int _{{\mathbb {R}}^d} \rho _{(x_j, t_j +\varepsilon _j ^2)} (y,t_j) \, \textrm{d}\mu _{t_j} ^{\varepsilon _j} (y) \le e^{C_3 (T +1)} \int _{{\mathbb {R}}^d} \rho _{(x_j, t_j +\varepsilon _j ^2)} (y,s) \, \textrm{d}\mu _{s} ^{\varepsilon _j}(y). \end{aligned} \end{aligned}$$

Choose \(\eta _2 = \eta _2 (d,\gamma _2, T, C_3) >0\) such that

$$\begin{aligned} \begin{aligned} 2 \eta _2 \le \int _{{\mathbb {R}}^d} \rho _{(x_j, t_j +\varepsilon _j ^2)} (y,s) \, \textrm{d}\mu _{s} ^{\varepsilon _j} (y) \end{aligned} \end{aligned}$$

and letting \(j\rightarrow \infty \), we have

$$\begin{aligned} \begin{aligned} 2 \eta _2 \le \int _{{\mathbb {R}}^d} \rho _{(x', t')} (y,s) \, \textrm{d}\mu _{s} (y). \end{aligned} \end{aligned}$$
(66)

Changing the center of the backward heat kernel by using (153), we have

$$\begin{aligned} \begin{aligned} \eta _2 \le \int _{{\mathbb {R}}^d} \rho _{(x, t')} (y,s) \, \textrm{d}\mu _{s}(y) \end{aligned} \end{aligned}$$

when \(\vert x-x' \vert \le \gamma _1 r\). Here \(\gamma _1\) depends only on \(\eta _2\) and \(D_2\). This is a contradiction to (65). Therefore \((x', t') \not \in \textrm{spt}\,\mu \). \(\square \)

We can also show the following using the estimate (66):

Lemma 7

There exists \(C_{10} \le 0\) depending only on d, T, \(C_3\), and \(D_2\) such that

$$\begin{aligned} {\mathscr {H}}^{d-1} (\textrm{spt}\,\mu _t \cap U) \le C_{10} \liminf _{r \downarrow 0} \mu _{t-r^2} (U) \end{aligned}$$
(67)

for any \(t \in (0,T) \setminus {\tilde{B}}\) and for any open set \(U \subset \Omega \), where \({\tilde{B}}\) is the countable set given by Proposition 11.

Proof

We only need to prove (67) for any compact set \(K \subset U\). Let \(X_t:= \{ x \in K \mid (x,t) \in \textrm{spt}\,\mu \}\) with \(t \in (0,T) {\setminus } {\tilde{B}}\). For any \(x \in X_t\), by (66), we have

$$\begin{aligned} \begin{aligned} 2 \eta _2 \le \int _{{\mathbb {R}}^d} \rho _{(x, t)} (y,t-r^2) \, \textrm{d}\mu _{t-r^2} (y) \end{aligned} \end{aligned}$$

for sufficiently small \(r>0\). By (154), we deduce that

$$\begin{aligned} \begin{aligned}&\int _{{\mathbb {R}}^d} \rho _{(x, t)} (y,t-r^2) \, \textrm{d}\mu _{t-r^2} (y) \\ \le&\, \int _{B _{Lr} (x)} \rho _{(x, t)} (y,t-r^2) \, \textrm{d}\mu _{t-r^2} (y) +2^{d-1} e^{- \frac{3L^2}{8}} D_2 \end{aligned} \end{aligned}$$

for any \(L>0\). Therefore for sufficiently large \(L>0\) depending only on d, \(\gamma _2\), T, \(C_3\), and \(D_2\), we have

$$\begin{aligned} \begin{aligned} \eta _2 \le \int _{B _{Lr} (x)} \rho _{(x, t)} (y,t-r^2) \, d\mu _{t-r^2} (y) \le (4\pi )^{-\frac{d-1}{2}} r^{1-d} \mu _{t-r^2} (B_{Lr} (x)), \end{aligned} \end{aligned}$$

where we used \(\rho _{(x,t)} (y,t-r^2) \le (4\pi )^{-\frac{d-1}{2}} r^{1-d} \). Hence there exists \(C_{11} >0\) depending only on d, \(\gamma _2\), T, \(C_3\), and \(D_2\) such that

$$\begin{aligned} \omega _{d-1} r^{d-1} \le C_{11} \mu _{t-r^2} (B_{Lr} (x)) \end{aligned}$$
(68)

holds for any sufficiently small \(r >0\). Set \({\mathcal {B}}:=\{ {\overline{B}} _{Lr} (x) \subset U \mid x \in X_t \}\). By the Besicovitch covering theorem, there exists a finite sub-collection \({\mathcal {B}}_1\), \({\mathcal {B}}_2\), ..., \({\mathcal {B}}_{N(d)}\) such that each \({\mathcal {B}}_i\) is a family of the disjoint closed balls and

$$\begin{aligned} X_t \subset \cup _{i=1} ^{N(d)} \cup _{{\overline{B}} _{Lr} (x_j) \in {\mathcal {B}}_i} {\overline{B}} _{Lr} (x_j). \end{aligned}$$
(69)

Let \({\mathscr {H}}^{d-1} _\delta \) be defined in [11, Chapter 2]. Note that \({\mathscr {H}}^{d-1} = \lim _{\delta \downarrow 0} {\mathscr {H}}^{d-1} _\delta \). By (68) and (69), we compute

$$\begin{aligned} {\mathscr {H}}^{d-1} _{2Lr} (X_t) \le&\, \sum _{i=1} ^{N(d)} \sum _{{\overline{B}} _{Lr} (x_j) \in {\mathcal {B}}_i} \omega _{d-1} (Lr) ^{d-1} \\ \le&\, \sum _{i=1} ^{N(d)} \sum _{{\overline{B}} _{Lr} (x_j) \in {\mathcal {B}}_i} L^{d-1} C_{11} \mu _{t-r^2} ({\overline{B}}_{Lr} (x_j)) \\ \le&\, \sum _{i=1} ^{N(d)} L^{d-1} C_{11} \mu _{t-r^2} (U) = N(d) L^{d-1} C_{11} \mu _{t-r^2} (U). \end{aligned}$$

Letting \(r\downarrow 0\), we have \({\mathscr {H}}^{d-1} (X_t) \le N(d) L^{d-1} C_{11} \liminf _{r\downarrow 0} \mu _{t-r^2} (U)\). By this and (61), we obtain (67). \(\square \)

By Lemma 6, we obtain

Lemma 8

(see [23, 32]) For \(T \in [1,\infty )\), let \(\eta _2\) be a constant as in Lemma 6. Set

$$\begin{aligned} Z_T:= \left\{ (x,t) \in \textrm{spt}\,\mu \mid 0\le t \le T/2, \ \limsup _{s\downarrow t} \int _{{\mathbb {R}}^d} \rho _{(y,s)} (x,t) \, \textrm{d} \mu _s (y) \le \eta _2 /2 \right\} . \end{aligned}$$

Then \(\mu (Z_T) =0\) holds.

Proof

Let \(\eta _1\), \(\eta _2\), and \(\gamma _1\) be constants as in Lemma 6. For \(\tau \in (0, \eta _1)\), we denote

$$\begin{aligned} Z^{\tau }:= \left\{ (x,t) \in \textrm{spt}\,\mu \mid 0\le t \le T/2, \ \int _{{\mathbb {R}}^d} \rho _{(y,s)} (x,t) \, \textrm{d} \mu _s (y) < \eta _2, \ \forall \ s \in (t, t+\tau ] \right\} . \end{aligned}$$

Let \(\{ \tau _m \}_{m=1} ^\infty \) be a positive sequence with \(\tau _m \rightarrow 0\) as \(m \rightarrow \infty \). Then \(Z_T \subset \cup _{m=1} ^{\infty } Z ^{\tau _m}\). Therefore we need only show \(\mu (Z ^\tau ) =0\) for any \(\tau \in (0,\eta _1)\). Set

$$\begin{aligned} P_\tau (x,t):= \{ (x', t') \mid \tau> \vert t-t' \vert >\gamma _1 ^{-2} \vert x-x' \vert ^2\}, \quad x \in \Omega , \ t \in [0, T/2). \end{aligned}$$

We now show that if \((x,t) \in Z^\tau \), then

$$\begin{aligned} P_\tau (x,t) \cap Z^\tau =\emptyset . \end{aligned}$$
(70)

Assume that \((x', t') \in P_\tau (x,t) \cap Z^\tau \) for a contradiction. First we consider the case of \(t' >t\). Set \(s=\frac{t'+t}{2}\) and \(r=\sqrt{t' -t} = \sqrt{2(s-t)}\). Since \((x,t) \in Z^\tau \),

$$\begin{aligned} \int _{{\mathbb {R}}^d} \rho _{x} ^r (y) \, \textrm{d} \mu _s (y) = \int _{{\mathbb {R}}^d} \rho _{(y,s)} (x,t) \, \textrm{d} \mu _s (y) < \eta _2. \end{aligned}$$

Therefore Lemma 6 yields \((x', t') \not \in \textrm{spt}\,\mu \), because \(x' \in B_{\gamma _1 r} (x)\) by the definition of \(P_\tau (x,t)\). This yields a contradiction. In the case of \(t' <t\), we can show \((x,t) \not \in \textrm{spt}\,\mu \) similarly. This is a contradiction. Therefore (70) holds.

For \((x_0, t_0) \in \Omega \times [\tau /2, T/2]\), we denote

$$\begin{aligned} Z^{\tau ,x_0,t_0} = Z^\tau \cap \left( B_{\frac{\gamma _1}{2} \sqrt{\tau }} (x_0) \times (t_0 -\tau /2, t_0 +\tau /2)\right) . \end{aligned}$$

We can choose a countable set \(\{ (x_j, t_j ) \}_{j=1} ^\infty \) such that \( Z^\tau \subset \cup _{j=1} ^\infty Z^{\tau ,x_j,t_j} \). Thus we only need to prove \(\mu (Z^{\tau ,x_0,t_0}) =0\). Let \(P: {\mathbb {R}}^{d+1} \rightarrow {\mathbb {R}}^d\) be a projection such that \(P(x,t)=x\). For \(\rho \in (0,1)\) and \(r \le \rho \), let \(\{{\overline{B}}_{r/5} (x_\lambda ) \}_{\lambda \in \Lambda } \) be a covering of \(P(Z^{\tau ,x_0,t_0}) \subset B_{\frac{\gamma _1}{2} \sqrt{\tau }} (x_0)\). Then, we may choose a countable covering \({\mathcal {F}}=\{ {\overline{B}}_{r} (x_i) \}_{i=1} ^\infty \) of \(P(Z^{\tau ,x_0,t_0})\) with \((x_i, t_i) \in Z^{\tau ,x_0,t_0}\) for some \(t_i\), by Vitali’s covering theorem. Let A be a set of centers of all balls in \(\{ {\overline{B}}_{r} (x_i) \}_{i=1} ^\infty \). Then, by Besicovitch’s covering theorem, there exist N(d) and subcollections \({\mathcal {F}} _1\), \({\mathcal {F}} _2\), ..., \({\mathcal {F}} _{N(d)} \subset {\mathcal {F}}\) of disjoint balls such that

$$\begin{aligned} A \subset \cup _{k=1} ^{N (d)} \cup _{B_{k,i} \in {\mathcal {F}}_k} B_{k,i}. \end{aligned}$$
(71)

Note that \({\mathcal {F}}_k\) is finite (\({\mathcal {F}} _k = \{ B_{k,1}, \dots ,B_{k,n_k} \}\)) and

$$\begin{aligned} {\mathscr {L}} ^d (\cup _{i=1} ^{n_k} B_{k,i}) = \sum _{i=1} ^{n_k} {\mathscr {L}} ^d (B_{k,i}) \le {\mathscr {L}}^d ( {\overline{B}}_{\frac{\gamma _1}{2} \sqrt{\tau } +\rho } (x_0) ) \end{aligned}$$

since each balls in \({\mathcal {F}} _k\) are disjoint and \(B_{k,i} \subset {\overline{B}}_{\frac{\gamma _1}{2} \sqrt{\tau } +\rho } (x_0)\). Therefore

$$\begin{aligned} \sum _{k=1} ^{N(d)} \sum _{i=1} ^{n_k} \omega _d r^d =\sum _{k=1} ^{N(d)} \sum _{B_{k,i} \in {\mathcal {F}} _k} {\mathscr {L}} ^d (B_{k,i}) \le N(d) {\mathscr {L}}^d ( {\overline{B}}_{\frac{\gamma _1}{2} \sqrt{\tau } +\rho } (x_0) ) =:N'.\nonumber \\ \end{aligned}$$
(72)

If \((x,t) \in Z^{\tau ,x_0,t_0}\), then there exists \(B_{k,i} = {\overline{B}}_r (x_{k,i}) \in {\mathcal {F}}_k\) for some k and i such that \(x \in {\overline{B}}_{2r} (x_{k,i})\) and \( \vert t_{k,i} -t \vert \le \gamma _1 ^{-1} \vert x_{k,i} -x \vert ^2 \le 4 \gamma _1 ^{-1} r ^2\) by (70) and (71) (note that we should change the radius because A is not a covering of \(Z^{\tau ,x_0,t_0}\)). Hence, we have

$$\begin{aligned} \begin{aligned} Z^{\tau ,x_0,t_0} \subset \cup _{k=1} ^{N(d)} \cup _{i=1} ^{n_k} {\overline{B}}_{2r} (x_{k,i}) \times (t_{k,i} - 4 r ^2 \gamma _1 ^{-2}, t_{k,i} + 4 r ^2 \gamma _1 ^{-2}) \end{aligned} \end{aligned}$$

By this, (52), and (72) we obtain

$$\begin{aligned} \begin{aligned} \mu (Z^{\tau ,x_0,t_0}) \le&\, \sum _{k=1} ^{N(d)} \sum _{i=1} ^{n_k} \mu ({\overline{B}}_{2r} (x_i) \times (t_i - 4 r ^2 \gamma _1 ^{-2}, t_i + 4 r ^2 \gamma _1 ^{-2})) \\ \le&\, \sum _{k=1} ^{N(d)} \sum _{i=1} ^{n_k} D_2 (2r) ^{d-1} \times 8 \gamma _1 ^{-2} r ^2 \le 2^{d+2} \gamma _1 ^{-2} \omega _d ^{-1} N' D_2 \rho . \end{aligned} \end{aligned}$$

Letting \(\rho \rightarrow 0\), we have \(\mu (Z^{\tau ,x_0,t_0})=0\). Thus \(\mu (Z_T) =0\) holds. \(\square \)

Theorem 13

(see [23]) We see that \(\vert \xi \vert =0\) and \(\lim _{i \rightarrow \infty } \vert \xi _t ^{\varepsilon _i} \vert (\Omega )=0\) for a.e. \(t\in [0,T)\).

Proof

First we show that

$$\begin{aligned} \int _{\Omega \times (0,s) } \frac{\rho _{(y,s)} (x,t) }{s-t} \, d \vert \xi \vert (x,t) \le C \end{aligned}$$
(73)

for some \(C>0\). By (33) and (51), integrating (50) on \((0, s-\delta )\) with \(\delta >0\), we obtain

$$\begin{aligned} \begin{aligned}&\int _0 ^{s-\delta } \frac{1}{2\sigma (s-t)} \int _{{\mathbb {R}}^d} \rho _{(y,s)} (x,t) \left| \frac{W(\varphi ^\varepsilon )}{\varepsilon } - \frac{\varepsilon \vert \nabla \varphi ^\varepsilon \vert ^2 }{2} \right| \, \textrm{d}x \textrm{d}t \\ \le&\, \left( 1+ e^{C_3 (s+1)} \frac{1}{2} \int _{0} ^{s-\delta } \vert \lambda ^\varepsilon \vert ^2 \, \textrm{d}t \right) \int _{{\mathbb {R}}^d} \rho _{(y,s)} (x,0) \, d \mu _0 ^\varepsilon . \end{aligned} \end{aligned}$$

Letting \(\delta \rightarrow 0\) and \(\varepsilon \rightarrow 0\), we obtain (73). Next, integrating (73) on \(\Omega \times (0,T)\) by \(d\mu _s ds\) we have

$$\begin{aligned} \int _{\Omega \times (0,T)} \left( \int _{\Omega \times (t,T)} \frac{\rho _{(y,s)} (x,t) }{s-t} \, \textrm{d}\mu _s (y) \textrm{d}s \right) \textrm{d} \vert \xi \vert (x,t) \le CD_1 T, \end{aligned}$$

where we used Fubini’s theorem. Then this boundedness implies

$$\begin{aligned} \int _{\Omega \times (t,T)} \frac{\rho _{(y,s)} (x,t) }{s-t} \, \textrm{d}\mu _s (y) \textrm{d}s <\infty \qquad \text {for} \ \vert \xi \vert \text {-a.e.} \ (x,t) \in \Omega \times (0,T). \end{aligned}$$
(74)

Next we claim

$$\begin{aligned} a(x,t):= \limsup _{s\downarrow t} \int _{\Omega } \rho _{(y,s)} (x,t) \, d\mu _s (y)=0 \qquad \text {for} \ \vert \xi \vert \text {-a.e.} \ (x,t) \in \Omega \times (0,T).\nonumber \\ \end{aligned}$$
(75)

Define \(\beta := \log (s-t)\) and

$$\begin{aligned} h (s):= \int _{\Omega } \rho _{(y,s)} (x,t) \, \textrm{d} \mu _s (y). \end{aligned}$$

Assume that (xt) satisfies (74). Then

$$\begin{aligned} \int _{-\infty } ^{\log (T-t)} h (t+ e^\beta ) \, \textrm{d} \beta <\infty . \end{aligned}$$
(76)

Let \(\theta \in (0,1]\) and \(\{ \beta _i \} _{i=1} ^\infty \) be a negative monotone decreasing sequence such that

$$\begin{aligned} \beta _i \downarrow -\infty , \quad 0< \beta _i - \beta _{i+1} \le \theta , \quad \text {and} \quad h(t + e^{\beta _i}) \le \theta . \end{aligned}$$

For any \(\beta \in (-\infty , \beta _1)\), choose i such that \(\beta \in [\beta _{i},\beta _{i-1})\) holds. One can check that

$$\begin{aligned} \sup _{y \in B_{Mr} (x)} \frac{\rho _{(y,t+2 e^\beta - e^{\beta _i })} (x,t )}{ \rho _{(y,t+ e^{\beta _i })} (x,t )} \le e^{M^2 (1-e ^{\beta -\beta _i})} \le e^{M^2 (1-e ^{\theta })} \end{aligned}$$
(77)

for \(M>0\), where \(r=\sqrt{2(2 e ^{\beta } -e ^{\beta _i})}\). We compute

$$\begin{aligned} \begin{aligned}&h(t+e^\beta ) = \int _{\Omega } \rho _{(y,t+e^\beta )} (x,t) \, \textrm{d}\mu _{t+e^\beta } (y) = \int _{\Omega } \rho _{(y,t+2 e^\beta )} (x,t + e^\beta ) \, \textrm{d} \mu _{t+e^\beta } (y)\\ \le&\, e^{C_3 (\beta -\beta _i +1) } \int _{\Omega } \rho _{(y,t+2 e^\beta )} (x,t + e^{\beta _i }) \, \textrm{d} \mu _{t+e^{\beta _i}} (y) \\ \le&\, e^{2 C_3 } \int _{\Omega } \rho _{(y,t+2 e^\beta - e^{\beta _i })} (x,t ) \, \textrm{d} \mu _{t+e^{\beta _i}} (y) \\ \le&\, e^{2 C_3 } \int _{B_{Mr} (x)} \rho _{(y,t+2 e^\beta - e^{\beta _i })} (x,t ) \, \textrm{d} \mu _{t+e^{\beta _i}} (y) + e^{2 C_3 } 2^{d-1} e^{-\frac{3M^2}{8}} D_2\\ \le&\, e^{2 C_3 } e^{M^2 (1-e ^{\theta })} \int _{B_{Mr} (x)} \rho _{(y,t+ e^{\beta _i })} (x,t ) \, \textrm{d} \mu _{t+e^{\beta _i}} (y) + e^{2 C_3 } 2^{d-1} e^{-\frac{3M^2}{8}} D_2\\ \le&\, e^{2 C_3 } e^{M^2 (1-e ^{\theta })} \theta + e^{2 C_3 } 2^{d-1} e^{-\frac{3M^2}{8}} D_2, \end{aligned} \end{aligned}$$
(78)

where we used (51), (154), and

$$\begin{aligned} \begin{aligned} \int _{\Omega } \rho _{(y,t+e^{\beta _i })} (x,t ) \, d \mu _{t+e^{\beta _i}} (y) =h(t+e^{\beta _i}) \le \theta . \end{aligned} \end{aligned}$$

Thus, for any \(\delta >0\), we can choose \(\theta \in (0,1]\) and \(M>0\) such that \(h (t + e^{\beta }) \le \delta \) for any \(\beta < \beta _1\). This proves (75). Set

$$\begin{aligned}{} & {} A:= \{ (x,t) \in \Omega \times (0,T) \mid a(x,t)=0 \} \ \ \text {and}\\{} & {} B:=\{ (x,t) \in \Omega \times (0,T) \mid a(x,t)>0 \}. \end{aligned}$$

Then \(\Omega \times (0,T)= A\cup B\) and \(\vert \xi \vert (B)=0\) by (75). Moreover, Lemma 8 and (154) imply \(\mu (A) =0\) and thus \(\vert \xi \vert (A) =0\), because \(\vert \xi \vert \) is absolute continuous with respect to \(\mu \). Therefore \(\vert \xi \vert (\Omega \times (0,T)) =0\). The rest of the claim can be shown from the dominated convergence theorem. \(\square \)

4.3 Rectifiability

Next we show the rectifiability of \(\mu _t\).

Definition 2

For \(\phi \in C_c (G_{d-1} (\Omega ))\), we define \(V_t ^\varepsilon \in {\mathbb {V}} _{d-1} (\Omega )\) by

$$\begin{aligned} V_t ^\varepsilon (\phi ):= \int _{\Omega \cap \{ \vert \nabla \varphi ^\varepsilon (x,t) \vert \not =0 \}} \phi \left( x, I -\frac{\nabla \varphi ^\varepsilon (x,t) }{ \vert \nabla \varphi ^\varepsilon (x,t) \vert } \otimes \frac{\nabla \varphi ^\varepsilon (x,t) }{\vert \nabla \varphi ^\varepsilon (x,t) \vert } \right) \, \textrm{d}\mu _t ^\varepsilon (x). \end{aligned}$$
(79)

Here, \(\varphi ^\varepsilon \) is a solution to (5).

Note that the first variation of \(V_t ^\varepsilon \) is given by

$$\begin{aligned} \begin{aligned} \delta V_t ^\varepsilon (\vec {\phi }) =&\, \int _{G_{d-1} (\Omega )} \nabla \vec {\phi } (x) \cdot S \, \textrm{d}V_t ^\varepsilon (x,S) \\ =&\, \int _{\Omega \cap \{ \vert \nabla \varphi ^\varepsilon (x,t) \vert \not =0 \}} \nabla \vec {\phi } (x) \cdot \left( I -\frac{\nabla \varphi ^\varepsilon (x,t) }{\vert \nabla \varphi ^\varepsilon (x,t) \vert } \otimes \frac{\nabla \varphi ^\varepsilon (x,t) }{\vert \nabla \varphi ^\varepsilon (x,t) \vert } \right) \, \textrm{d}\mu _t ^\varepsilon (x) \end{aligned} \end{aligned}$$

for \(\vec {\phi } \in C_c ^1 (\Omega ; {\mathbb {R}}^d)\). By integration by parts, we have

$$\begin{aligned} \begin{aligned}&\delta V_t ^\varepsilon (\vec {\phi }) \\ =&\, \int _{\Omega } (\vec {\phi } \cdot \nabla \varphi ^\varepsilon ) \left( \varepsilon \Delta \varphi ^\varepsilon -\frac{W'(\varphi ^\varepsilon )}{\varepsilon } \right) \, \text {d}x -\int _{\Omega \cap \{ \vert \nabla \varphi ^\varepsilon (x,t) \vert \not =0 \}} \frac{W(\varphi ^\varepsilon )}{\varepsilon } \textrm{div}\, \vec {\phi } \, \text {d}x \\ {}&+ \int _{\Omega \cap \{ \vert \nabla \varphi ^\varepsilon (x,t) \vert \not =0 \}} \nabla \vec {\phi } \cdot \left( \frac{\nabla \varphi ^\varepsilon (x,t) }{\vert \nabla \varphi ^\varepsilon (x,t) \vert } \otimes \frac{\nabla \varphi ^\varepsilon (x,t) }{\vert \nabla \varphi ^\varepsilon (x,t) \vert } \right) \xi _\varepsilon \, \text {d}x. \end{aligned} \end{aligned}$$
(80)

Note that the second and third terms of the right hand side converges to 0 for a.e. \(t \in [ 0,T)\) by Theorem 13. By (35) and (38), we have

$$\begin{aligned} \sup _{i \in {\mathbb {N}}} \int _0 ^T \int _{\Omega } \varepsilon _i \left( \Delta \varphi ^{\varepsilon _i } -\dfrac{W' (\varphi ^{\varepsilon _i})}{\varepsilon _i ^2 } \right) ^2 \,\textrm{d}x\textrm{d}t \le C \end{aligned}$$

for some \(C>0\) (see the proof of Theorem 12). Thus Fatou’s lemma implies

$$\begin{aligned} \liminf _{i \rightarrow \infty } \int _\Omega \varepsilon _i \left( \Delta \varphi ^{\varepsilon _i} -\frac{W'(\varphi ^{\varepsilon _i})}{\varepsilon _i ^2} \right) ^2 \, \textrm{d}x < \infty \end{aligned}$$
(81)

for a.e. \(t \in [ 0, T)\). Hence, by (80), (81), and the Cauchy–Schwarz inequality, we have

$$\begin{aligned} \begin{aligned}&\liminf _{i\rightarrow \infty } \vert \delta V_t ^{\varepsilon _i} (\vec {\phi }) \vert \\ \le&\, \liminf _{i\rightarrow \infty } \left( \int _{\Omega } \varepsilon _i \vert \nabla \varphi ^{\varepsilon _i} \vert ^2 \, \textrm{d}x \right) ^{\frac{1}{2}} \left( \int _\Omega \varepsilon _i \left( \Delta \varphi ^{\varepsilon _i} -\frac{W'(\varphi ^{\varepsilon _i})}{\varepsilon _i ^2} \right) ^2 \, \textrm{d}x \right) ^{\frac{1}{2}} \\ \le&\, D_1 ^{\frac{1}{2}} \liminf _{i\rightarrow \infty } \left( \int _\Omega \varepsilon _i \left( \Delta \varphi ^{\varepsilon _i} -\frac{W'(\varphi ^{\varepsilon _i})}{\varepsilon _i ^2} \right) ^2 \, \textrm{d}x \right) ^{\frac{1}{2}} <\infty \end{aligned} \end{aligned}$$
(82)

for a.e. \(t \in [ 0, T)\) and for any \(\vec {\phi } \in C_c ^1 (\Omega ;{\mathbb {R}}^d)\) with \(\sup \vert \vec {\phi } \vert \le 1\). Let \(t \in [ 0, T) {\setminus } {\tilde{B}}\) satisfy (82), where \({\tilde{B}}\) is given by Proposition 11. Taking a subsequence \(i_j \rightarrow \infty \) (note that the subsequence depends on t), there exists a varifold \(V_t\) such that \(V_t ^{\varepsilon _{i_j}} \rightharpoonup V_t\) as Radon measures and \(\delta V_t\) is a Radon measure by (82). In addition, Proposition 11, Lemma 7, and the standard measure theoretic argument imply

$$\begin{aligned} V_t =V_t \lfloor _{\{ x \in \Omega \mid \limsup _{r\downarrow 0} r^{1-d} \Vert V_t \Vert (B_r (x)) >0 \} \times {\mathbb {G}}(d,d-1)}. \end{aligned}$$

Therefore Allard’s rectifiability theorem yields the following theorem.

Theorem 14

For a.e. \(t \ge 0\), \(\mu _t\) is rectifiable. In addition, for a.e. \(t \ge 0\), \(\mu _t\) has a generalized mean curvature vector \(\vec {h} (\cdot ,t)\) with

$$\begin{aligned} \delta V_t (\vec {\phi }) =- \int _{\Omega } \vec {\phi } \cdot h (\cdot ,t) \, \textrm{d}\mu _t =\lim _{i \rightarrow \infty } \int _{\Omega } (\vec {\phi } \cdot \nabla \varphi ^{\varepsilon _i}) \left( \varepsilon _i \Delta \varphi ^{\varepsilon _i} -\frac{W'(\varphi ^{\varepsilon _i})}{\varepsilon _i} \right) \, \textrm{d}x \end{aligned}$$

and

$$\begin{aligned} \int _\Omega \phi \vert \vec {h} \vert ^2 \, \textrm{d}\mu _t \le \frac{1}{\sigma } \liminf _{i\rightarrow \infty } \int _{\Omega } \varepsilon _i \phi \left( \Delta \varphi ^{\varepsilon _i} -\frac{W'(\varphi ^{\varepsilon _i})}{\varepsilon _i ^2} \right) ^2 \, \textrm{d}x <\infty \end{aligned}$$

for any \(\phi \in C_c (\Omega ;[0,\infty ))\) and \(\vec \phi \in C_c (\Omega ;{\mathbb {R}}^d)\).

Detailed proof of this is in [23, 45], so we omit it (however, the essential part has already been discussed above). Note that (59) and \(\mu _t =\Vert V_t\Vert \) imply that \(V_t\) does not depend on the choice of subsequence \(\{V_t ^{\varepsilon _{i_j}} \}_{j=1} ^\infty \) above.

4.4 Integrality

To prove the integrality, we mainly follow [22, 45, 47]. The propositions that are directly applicable to our problem are in Appendix for readers’ convenience. Let \(\{ r _i\} _{i=1} ^\infty \) be a positive sequence with \(r _i \rightarrow 0\) and \(\frac{\varepsilon _i}{r_i} \rightarrow 0\) as \(i \rightarrow \infty \). Set \(u ^{{\tilde{\varepsilon }}} ({\tilde{x}}, {\tilde{t}}) = \varphi ^\varepsilon (x,t)\) and \(g ^{{\tilde{\varepsilon }}} ({\tilde{t}}) = r \lambda ^\varepsilon (t)\) for \({\tilde{x}}= \frac{x}{r}\), \({\tilde{t}}= \frac{t}{r^2}\), and \({\tilde{\varepsilon }} = \frac{\varepsilon }{r}\). Then, \(u ^{{\tilde{\varepsilon }}}\) is a solution to

$$\begin{aligned} {\tilde{\varepsilon }} \partial _{{\tilde{t}}} u ^{{\tilde{\varepsilon }}} ={\tilde{\varepsilon }} \Delta _{{\tilde{x}}} u ^{{\tilde{\varepsilon }}} -\dfrac{W' (u^{{\tilde{\varepsilon }}})}{{\tilde{\varepsilon }} } + g ^{{\tilde{\varepsilon }}} \sqrt{2W(u ^{{\tilde{\varepsilon }}})}. \end{aligned}$$
(83)

We remark that the monotonicity formula (51) and the upper bound of the density (52) hold for \(d\tilde{\mu }_{{\tilde{t}}} ^{{\tilde{\varepsilon }}} ({\tilde{x}})= \sigma ^{-1} (\frac{{\tilde{\varepsilon }} \vert \nabla _{{\tilde{x}}} u ^{\tilde{\varepsilon }} \vert ^2}{2} + \frac{W(u ^{{\tilde{\varepsilon }}})}{\tilde{\varepsilon }} )\, d{\tilde{x}}\), because the value

$$\begin{aligned} \int _{{\mathbb {R}}^d} \rho _{(y,s)} (x,t) \, d\mu _t ^\varepsilon (x) \end{aligned}$$

is invariant under this rescaling, and for any \(s>0\) we have

$$\begin{aligned} \begin{aligned}&\frac{1}{s^{d-1}} \int _{B_s (0)} \left( \frac{{\tilde{\varepsilon }} \vert \nabla _{{\tilde{x}}} u ^{{\tilde{\varepsilon }}} \vert ^2}{2} + \frac{W(u ^{{\tilde{\varepsilon }}})}{{\tilde{\varepsilon }}} \right) \, \textrm{d}{\tilde{x}} \\ =&\, \frac{1}{(sr)^{d-1}}\int _{B_{sr} (0)} \left( \frac{\varepsilon \vert \nabla \varphi ^{\varepsilon }\vert ^2}{2} + \frac{W(\varphi ^{\varepsilon })}{\varepsilon } \right) \, \textrm{d}x \le \sigma D_2 \end{aligned} \end{aligned}$$

by (52). We subsequently drop \({\tilde{\cdot }}\) for simplicity. First we consider the energy estimate on \(\{x \in B_1 (0) \mid \vert u ^\varepsilon (x,t) \vert \ge 1 -b \}\).

Proposition 15

(See [47]) For any \(s >0\) and \(a \in (0,T)\), there exist positive constants b and \(\varepsilon _2\) depending only on \(D_1\), \(D_2\), \(C_3\), a, \(\alpha \), and s such that

$$\begin{aligned} \int _{\{ x \in B_1 (0) \mid \vert u ^\varepsilon (x,t) \vert \ge 1-b \}} \frac{W( u ^\varepsilon (x,t))}{\varepsilon } \, \textrm{d}x \le s \end{aligned}$$

for all \(t \in (a, T)\) whenever \(\varepsilon \in (0,\varepsilon _2)\).

To prove Proposition 15, we prepare following two lemmas:

Lemma 9

(See [47]) For any \(\delta \in (0,T)\), there exist positive constants \(C_{12}\) and \(\varepsilon _3\) depending only on d, \(\delta \), \(\alpha \), and \(C_1\) with the following property. assume that there exist \((x_0,t_0) \in B_1 (0) \times (\delta ,T)\) and \(\gamma \in (0,\frac{2}{3}]\) such that

$$\begin{aligned} u ^\varepsilon (x_0,t_0) <1 -\varepsilon ^\gamma \qquad (\text {or} \ u ^\varepsilon (x_0,t_0) > -1 + \varepsilon ^\gamma ) \end{aligned}$$
(84)

and

$$\begin{aligned} 1\le {\tilde{r}}:= C_{12} \gamma \vert \log \varepsilon \vert \le \varepsilon ^{-1} \min \left\{ \sqrt{\frac{\delta }{2}}, \ \frac{1}{2} \right\} . \end{aligned}$$
(85)

Then

$$\begin{aligned} \inf _{B_{\varepsilon {\tilde{r}}} (x_0) \times (t_0 -\varepsilon ^2 {\tilde{r}}^2, t_0)} u ^\varepsilon < \frac{1}{2} \qquad \left( \text {resp.} \ \sup _{B_{\varepsilon {\tilde{r}}} (x_0) \times (t_0 -\varepsilon ^2 {\tilde{r}}^2, t_0)} u ^\varepsilon > - \frac{1}{2} \right) \end{aligned}$$

for any \(\varepsilon \in (0,\varepsilon _3)\).

Proof

We may assume that \(B_{\varepsilon {\tilde{r}}} (x_0) \times (t_0 -\varepsilon ^2 {\tilde{r}}^2, t_0) \subset B_2 (0) \times (0,T)\) by (85). We consider the rescaling of (83) by \({\tilde{x}}= \frac{x-x_0}{\varepsilon }\) and \({\tilde{t}}= \frac{t-t_0}{\varepsilon ^2}\). Then we obtain

$$\begin{aligned} \partial _{{\tilde{t}}} {\tilde{u}} ^{\varepsilon } = \Delta _{{\tilde{x}}} {\tilde{u}} ^{\varepsilon } - W' ({\tilde{u}} ^{\varepsilon }) + \varepsilon {\tilde{g}} ^{\varepsilon } \sqrt{2W({\tilde{u}} ^\varepsilon )}, \qquad (x,t) \in B_{{\tilde{r}}} (0) \times (-{\tilde{r}} ^2,0), \end{aligned}$$
(86)

where \({\tilde{u}} ^{\varepsilon } ({\tilde{x}},{\tilde{t}}) = u ^\varepsilon (x,t)\) and \({\tilde{g}} ^{\varepsilon } ({\tilde{t}}) = g ^\varepsilon (t)\). Note that (12) and \({\mathscr {L}}^d (\Omega )=1\) yield

$$\begin{aligned} \Vert \varepsilon {\tilde{g}} ^{\varepsilon } \Vert _{L^\infty } \le \frac{4}{3} \varepsilon ^{1- \alpha } \end{aligned}$$
(87)

for \(\alpha \in (0,1)\). Let \(\psi \) be a function with

$$\begin{aligned} \left\{ \begin{array}{ll} \partial _{\tilde{t}} \psi \, &{} \ge \Delta _{\tilde{x}} \psi -\frac{1}{10} \psi \qquad \text {on} \ {\mathbb {R}}^d \times (-\infty ,0), \\ \psi ({\tilde{x}},{\tilde{t}}) \, &{} \ge e^{\frac{\vert {\tilde{x}} \vert +\vert {\tilde{t}} \vert }{C_{13}}} \qquad \text {on} \ ({\mathbb {R}}^d \times (-\infty ,0)) \setminus B_1 (0,0), \\ \psi (0,0) \, &{} = 1, \end{array} \right. \end{aligned}$$
(88)

for some constant \(C_{13}>0\). For example, \(\psi = e^{-\frac{{\tilde{t}}}{100} -1} e^{\frac{1}{100d}\sqrt{1+\vert {\tilde{x}} \vert ^2}}\) satisfies (88). Set \({\tilde{r}}:= C_{13} \gamma \vert \log \varepsilon \vert \). We may assume that \({\tilde{r}}\ge 1\) for sufficiently small \(\varepsilon \). Note that

$$\begin{aligned} 1- \varepsilon ^\gamma e^{\frac{{\tilde{r}}}{C_{13}}} =0. \end{aligned}$$
(89)

The assumption (84) is equivalent to

$$\begin{aligned} {\tilde{u}} ^\varepsilon (0,0) <1 -\varepsilon ^\gamma . \end{aligned}$$
(90)

For a contradiction, we assume that

$$\begin{aligned} \inf _{B_{{\tilde{r}}} (0) \times (-{\tilde{r}} ^2, 0)} {\tilde{u}} ^\varepsilon \ge \frac{1}{2}. \end{aligned}$$
(91)

Set \(\phi ^\varepsilon := 1 - \varepsilon ^\gamma \psi \). Then (88) and (90) imply

$$\begin{aligned} \partial _{{\tilde{t}}} \phi ^\varepsilon \le \Delta _{{\tilde{x}}} \phi ^\varepsilon +\frac{1}{10} (1- \phi ^\varepsilon ) \qquad \text {on} \ {\mathbb {R}}^d \times (-\infty ,0) \end{aligned}$$

and

$$\begin{aligned} \phi ^\varepsilon (0,0) =1- \varepsilon ^\gamma \psi (0,0) =1 -\varepsilon ^\gamma > {\tilde{u}} ^\varepsilon (0,0). \end{aligned}$$
(92)

Moreover, by \({\tilde{r}} \ge 1\),

$$\begin{aligned} \psi \ge e^{\frac{\vert {\tilde{x}} \vert + \vert {\tilde{t}} \vert }{C_{13}}} \ge e^{\frac{\tilde{r}}{C_{13}}} \qquad \text{ on } \ \partial ( B_{{\tilde{r}}} (0) \times (-{\tilde{r}} ^2,0)). \end{aligned}$$

Therefore

$$\begin{aligned} \phi ^\varepsilon = 1 - \varepsilon ^\gamma \psi \le 1- \varepsilon ^\gamma e^{\frac{{\tilde{r}}}{C_{13}}} = 0 <\frac{1}{2} \le {\tilde{u}} ^\varepsilon \qquad \text {on} \ \partial ( B_{{\tilde{r}}} (0) \times (-{\tilde{r}} ^2,0)) \end{aligned}$$
(93)

by (89) and (91). We consider a function \(w = \phi ^\varepsilon - {\tilde{u}}^\varepsilon \) on \(B_{{\tilde{r}}} (0) \times (-{\tilde{r}} ^2,0)\). By (92) and (93), w attains its positive maximum at an interior point \((x',t') \in B_{{\tilde{r}}} (0) \times (-{\tilde{r}} ^2,0)\), and hence \(\partial _{{\tilde{t}}} w - \Delta _{{\tilde{x}}} w \ge 0\) and \(w >0\) at \((x',t')\). At \((x',t')\), we compute that

$$\begin{aligned} \begin{aligned} 0 \le&\, \partial _{{\tilde{t}}} w - \Delta _{{\tilde{x}}} w \le \frac{1}{10} (1- \phi ^\varepsilon ) + W' ({\tilde{u}} ^\varepsilon ) - \varepsilon {\tilde{g}} ^\varepsilon \sqrt{2W ({\tilde{u}} ^\varepsilon )} \\ =&\, \frac{1}{10} (1- \phi ^\varepsilon ) -2 {\tilde{u}} ^\varepsilon (1- ({\tilde{u}} ^\varepsilon )^2) - \varepsilon {\tilde{g}} ^\varepsilon (1- ({\tilde{u}} ^\varepsilon )^2) \\ \le&\, \frac{1}{10} (1- \phi ^\varepsilon ) +(-1 + \frac{8}{3} \varepsilon ^{1-\alpha }) (1- ({\tilde{u}} ^\varepsilon )^2) \\ \le&\, \frac{1}{10} (1- \phi ^\varepsilon ) + \frac{3}{2} (-1 + \frac{8}{3} \varepsilon ^{1-\alpha }) (1- \phi ^\varepsilon ) <0 \end{aligned} \end{aligned}$$

for sufficiently small \(\varepsilon \), where we used (87) and \(1>\phi ^\varepsilon > {\tilde{u}} ^\varepsilon \ge \frac{1}{2}\) at \((x',t')\). This is a contradiction. The other case can be proved similarly. \(\square \)

Lemma 10

(See [47]) For any \(\delta \in (0,T)\), there exist positive constants \(C_{14}\) and \(\varepsilon _4\) depending only on \(\delta \), \(\alpha \), d, \(C_3\), and \(D_2\) such that the following holds. For \(t \in (\delta , T)\) and \(r \in (0,\frac{1}{2})\), set

$$\begin{aligned} Z_{r,t_0}:= \left\{ x_0 \in B_1 (0) \mid \inf _{B_r (x_0) \times ( t_0 -r^2,t_0)} \vert u ^\varepsilon \vert <\frac{1}{2} \right\} . \end{aligned}$$

Then for any \(\varepsilon \in (0,\varepsilon _4)\), we have

$$\begin{aligned} {\mathscr {L}}^{d} (Z_{r,t_0}) \le C_{14} r, \qquad \varepsilon \le r <\frac{1}{2}. \end{aligned}$$
(94)

Proof

First we claim that there exist some constants \(\varepsilon _4\), \(C_{15}\), and \(C_{16}\) such that if \(x_0 \in Z_{r, t_0}\) and \(\varepsilon \in (0,\varepsilon _4)\) then

$$\begin{aligned} \sigma \mu _{t_0 -2r^2} ^\varepsilon (B_{C_{15} r } (x_0)) = \left. \int _{B_{C_{15} r } (x_0)} \frac{\varepsilon \vert \nabla u ^\varepsilon \vert ^2}{2} + \frac{W(u ^\varepsilon )}{\varepsilon } \, \text {d}x \right| _{t=t_0-2r^2} \ge C_{16} r^{d-1} \end{aligned}$$
(95)

holds for any \(r \in [\varepsilon ,\frac{1}{2})\). We may assume that \((x_1, t_1) \in B_r (x_0) \times (t_0 -r^2, t_0)\) with \(\vert u ^\varepsilon (x_1,t_1) \vert <\frac{1}{2}\). By the monotonicity formula (51), for any \(\varepsilon \in (0,\varepsilon _1)\) we have

$$\begin{aligned}{} & {} \int _{{\mathbb {R}}^d} \rho _{(x_1, t_1 +\varepsilon ^2)} (x,t) \, \textrm{d}\mu _t ^\varepsilon (x) \Big \vert _{t=t_1} \nonumber \\{} & {} \quad \le \left( \int _{{\mathbb {R}}^d} \rho _{(x_1, t_1 +\varepsilon ^2)} (x,t) \, \textrm{d}\mu _t ^\varepsilon (x) \Big \vert _{t=t_0 -2r^2} \right) e^{C_3 (3r^2 +1)}. \end{aligned}$$
(96)

By \(\vert u ^\varepsilon (x_1,t_1)\vert <\frac{1}{2}\), repeating the proof of Lemma 6, there exists \( \eta =\eta (\alpha ,d) >0\) such that

$$\begin{aligned} \eta \le \int _{{\mathbb {R}}^d} \rho _{(x_1, t_1 +\varepsilon ^2)} (x,t) \, \textrm{d}\mu _t ^\varepsilon (x) \Big \vert _{t=t_1}. \end{aligned}$$
(97)

Then (96), (97), and (154) imply

$$\begin{aligned} \eta ' \le \int _{B_R (x_1) } \rho _{(x_1, t_1 +\varepsilon ^2)} (x,t) \, \textrm{d}\mu _t ^\varepsilon (x) \Big \vert _{t=t_0 -2r^2} + 2^{d-1} e^{-\frac{3R^2}{16(t_1 +\varepsilon ^2 -t_0 +2 r^2)}} D_2, \end{aligned}$$

where \(\eta ' =\eta '(\alpha ,d,C_3) >0\). By \(\vert t_1 -t_0 \vert <r^2\) and \(\varepsilon \le r\), we have \(e^{-\frac{3R^2}{16(t_1 +\varepsilon ^2 -t_0 +2 r^2)}} \le e^{-\frac{3R^2}{64 r^2}}\). Thus there exists \(\gamma >0\) depending only on \(\alpha ,d,C_3\), and \(D_2\) such that

$$\begin{aligned} \frac{\eta '}{2} \le \int _{B_{\gamma r} (x_1) } \rho _{(x_1, t_1 +\varepsilon ^2)} (x,t_0 -2r^2) \, d\mu _{t_0 -2r^2} ^\varepsilon (x). \end{aligned}$$

Note that since \(t_1 +\varepsilon ^2 - (t_0 -2r^2) \ge 2r^2\) there exists \(C_{17}>0\) depending only on d such that

$$\begin{aligned} \rho _{(x_1, t_1 +\varepsilon ^2)} (x,t_0 -2r^2) \le \frac{C}{r^{d-1}}. \end{aligned}$$

Hence we obtain (95) for some \(C_{15}\) and \(C_{16}\).

Finally we prove (94). The inequality (95) yields that there exists \(C_{17}>0\) depending only on \(\alpha ,d,C_3\), and \(D_2\) such that

$$\begin{aligned} {\mathscr {L}} ^d ({\overline{B}}_{C_{15} r} (x_0)) \le r C_{17} \mu _{t_0 -2 r^2} ^\varepsilon ({\overline{B}}_{C_{15} r} (x_0)) \end{aligned}$$
(98)

for any \(x _0 \in Z_{r,t_0}\) and \(r \in [\varepsilon ,\frac{1}{2})\). Set \({\tilde{r}}:= C_{15} r\). By an argument similar to that in the proof of Lemma 8, there exist \({\mathcal {F}}_1,\dots ,{\mathcal {F}} _{N(d)}\) such that N(d) depends only on d, \({\mathcal {F}} _k= \{ {\overline{B}}_{{\tilde{r}}} (x_{k,1}), \dots , {\overline{B}}_{{\tilde{r}}} (x_{k,n_k}) \}\) is a family of disjoint closed balls for any k, and

$$\begin{aligned} Z_{r, t_0} \subset \cup _{k=1} ^{N(d)} \cup _{i=1} ^{n_k} {\overline{B}}_{2{\tilde{r}}} (x_{k,i}), \qquad x_{k,i} \in Z_{r,t_0} \quad \text {for any} \ k \ \text {and} \ i. \end{aligned}$$

Therefore

$$\begin{aligned} \begin{aligned} {\mathscr {L}}^d (Z_{r,t_0}) \le&\, \sum _{k=1} ^{N(d)} \sum _{i=1} ^{n_k} {\mathscr {L}}^d ( {\overline{B}} _{2{\tilde{r}}} (x_{k,i}) ) = 2^d \sum _{k=1} ^{N(d)} \sum _{i=1} ^{n_k} {\mathscr {L}}^d ( {\overline{B}} _{{\tilde{r}}} (x_{k,i}) ) \\ \le&\, 2^d \sum _{k=1} ^{N(d)} \sum _{i=1} ^{n_k} r C_{17} \mu _{t_0 -2 r^2} ^\varepsilon ({\overline{B}}_{{\tilde{r}}} (x_{k,i})) \\ =&2^d r C_{17} \sum _{k=1} ^{N(d)} \mu _{t_0 -2 r^2} ^\varepsilon \left( \cup _{i=1} ^{n_k} {\overline{B}}_{{\tilde{r}}} (x_{k,i}) \right) \\ \le&\, 2^d r C_{17} \sum _{k=1} ^{N(d)} \mu _{t_0 -2 r^2} ^\varepsilon (B_{1+ C_{15/2}} (0)) \le 2^d r C_{17} N(d) D_2, \end{aligned} \end{aligned}$$

where we used (52), (98), and the property that \({\mathcal {F}} _k\) is a family of disjoint balls. Hence, we obtain (94). \(\square \)

Proof of Proposition 15

First, we restrict \(b\in (0,1)\) to be small enough so that

$$\begin{aligned} 1-\sqrt{b} >\frac{1}{2}, \qquad \log \sqrt{b} \le -1, \qquad C_{12} \vert \log b \vert \ge 1. \end{aligned}$$
(99)

and restrict \(\varepsilon \) to be small enough to use Lemmas 9 and 10. We choose a positive integer J such that

$$\begin{aligned} \varepsilon ^{\frac{1}{2^{J+1}}} \in (b, \sqrt{b} ]. \end{aligned}$$
(100)

Then, (85), (99), and (100) imply

$$\begin{aligned} 1\le C_{12} \vert \log b \vert \le \frac{1}{2^J}C_{12} \vert \log \varepsilon \vert . \end{aligned}$$
(101)

Set \(t_0 \in (\delta ,T)\) and

$$\begin{aligned} A_j:= \left\{ x \in B_1 (0) \mid 1- \varepsilon ^{\frac{1}{2^{j+1}}} \le \vert u ^\varepsilon (x, t_0) \vert \le 1- \varepsilon ^{\frac{1}{2^{j}}} \right\} , \qquad j=1,\dots , J. \end{aligned}$$

For \(x_0 \in A_j\), we use Lemma 9 with \(\gamma =\frac{1}{2^j}\). Note that (85) holds with \({\tilde{r}} = \frac{1}{2^j} C_{12} \vert \log \varepsilon \vert \) by (101). Then we obtain

$$\begin{aligned} \inf _{B_{\varepsilon {\tilde{r}}} (x_0) \times (t_0 -\varepsilon ^2 {\tilde{r}}^2, t_0)} \vert u ^\varepsilon \vert < \frac{1}{2} \end{aligned}$$

and hence

$$\begin{aligned} A_j \subset Z_{\varepsilon {\tilde{r}},t_0}. \end{aligned}$$
(102)

By (85), we have \(\varepsilon \le \varepsilon {\tilde{r}} <\frac{1}{2}\) for sufficiently small \(\varepsilon \). Therefore (94) and (102) yield

$$\begin{aligned} {\mathscr {L}}^{d} (A_j) \le {\mathscr {L}}^{d} (Z_{\varepsilon \tilde{r},t_0}) \le \frac{1}{2^j} C_{12} C_{14} \varepsilon \vert \log \varepsilon \vert \end{aligned}$$
(103)

for any \(j=1,\dots ,J\). On the other hand, since \(\vert u ^\varepsilon (x,t_0) \vert \ge 1- \varepsilon ^{\frac{1}{2^{j+1}}}\) for any \(x \in A_j\), we obtain

$$\begin{aligned} \frac{W (u ^\varepsilon (x,t_0))}{\varepsilon } \le \frac{W (1- \varepsilon ^{\frac{1}{2^{j+1}}})}{\varepsilon } \le C_{18} \varepsilon ^{\frac{1}{2^{j}} -1} \end{aligned}$$
(104)

for some constant \(C_{18}\) depending only on W. We define \( Y:= \{ x \in B_1(0) \mid 1-b \le \vert u ^\varepsilon (x,t_0) \vert \le 1-\sqrt{\varepsilon } \} \). Note that

$$\begin{aligned} Y \subset \cup _{j=1} ^J A_j \end{aligned}$$
(105)

by (100). Set \(p(t) = 2^{-t} \varepsilon ^{2^{-t}}\). Then p satisfies

$$\begin{aligned} p' (t) = -(\log 2) 2^{-t} \varepsilon ^{2^{-t}} (1 + 2^{-t} \log \varepsilon ) >0 \quad \text {for any} \ t \in [1, J +1], \end{aligned}$$
(106)

because \(2^{-J-1} \log \varepsilon \le \log \sqrt{b} \le -1\) by (99) and (100). Set \(C_{19} = C_{12} C_{14} C_{18}\). Then from (100), (103), (104), (105), and (106) we have

$$\begin{aligned} \begin{aligned}&\int _Y \frac{W ( u ^\varepsilon (x,t_0))}{\varepsilon } \, \text {d}x \le \sum _{j=1} ^J \int _{A_j} \frac{W (u ^\varepsilon (x,t_0))}{\varepsilon } \, \text {d}x \le C_{19} \vert \log \varepsilon \vert \sum _{j=1} ^J 2^{-j} \varepsilon ^{2^{-j}} \\ \le&\, C_{19} \vert \log \varepsilon \vert \int _1 ^{J+1} 2^{-t} \varepsilon ^{2^{-t}} \, \text {d}t = C_{19} \frac{ \varepsilon ^{\frac{1}{2^{J+1}}} - \sqrt{\varepsilon } }{\log 2} \le C_{19} \frac{\sqrt{b}}{\log 2}. \end{aligned} \end{aligned}$$
(107)

Using the same argument above, we can show that

$$\begin{aligned} \begin{aligned}&\int _{\{ x\in B_1 (0) \mid 1-\sqrt{\varepsilon } \le \vert u^\varepsilon (x,t_0) \vert \le 1 -\varepsilon ^{\frac{2}{3}} \}} \frac{W(u ^\varepsilon )}{\varepsilon } \, \text {d}x \\ \le&\, C_{18} {\mathscr {L}} ^d (\{ x \in B_1 (0) \mid 1-\sqrt{\varepsilon } \le \vert u^\varepsilon (x,t_0) \vert \le 1 -\varepsilon ^{\frac{2}{3}} \}) \\ \le&\, \frac{2}{3} C_{19} \varepsilon \vert \log \varepsilon \vert , \end{aligned} \end{aligned}$$
(108)

where we used Lemma 9 with \(\gamma = \frac{2}{3}\). Since \(\vert u ^\varepsilon \vert \le 1\), we have

$$\begin{aligned} \begin{aligned}&\int _{\{ x \in B_1 (0) \mid 1-\varepsilon ^{\frac{2}{3}} \le \vert u^\varepsilon (x,t_0) \vert \}} \frac{W( u ^\varepsilon )}{\varepsilon } \, \textrm{d}x \\ \le&\, \frac{W(1-\varepsilon ^{\frac{2}{3}})}{\varepsilon } {\mathscr {L}}^d (\{ x \in B_1 (0) \mid 1-\varepsilon ^{\frac{2}{3}} \le \vert u^\varepsilon (x,t_0) \vert \})\\ \le&\, \varepsilon ^{\frac{1}{3}} {\mathscr {L}}^d (B_1 (0)). \end{aligned} \end{aligned}$$
(109)

By (107), (108), and (109), Proposition 15 holds for sufficiently small b and \(\varepsilon \). \(\square \)

Now we prove the integrality of \(\mu _t\).

Theorem 16

For a.e. \(t>0\), there exist a countably \((d-1)\)-rectifiable set \(M_t\) and \({\mathscr {H}}^{d-1}\)-measurable function \(\theta _t: M_t \rightarrow {\mathbb {N}}\) with \(\theta _t \in L_{loc} ^{1} ({\mathscr {H}}^{d-1} \lfloor _{M_t})\) such that \(\mu _t = \theta _t {\mathscr {H}}^{d-1} \lfloor _{M_t}\) holds.

Proof

Set \(H ^\varepsilon := \Delta \varphi ^\varepsilon -\frac{W' (\varphi ^\varepsilon )}{\varepsilon ^2}\). Then for a.e. \(t _0 > 0\), we can choose a subsequence \(\{ V_{t_0} ^{\varepsilon _{i_j}} \} _{j=1} ^\infty \) such that \(V_{t_0} ^{\varepsilon _{i_j}} \rightharpoonup V_{t_0}\),

$$\begin{aligned} \lim _{j\rightarrow \infty } \int _{\Omega } \vert \xi _{\varepsilon _{i_j}} (x,t_0) \vert \, \textrm{d}x =0, \end{aligned}$$
(110)

and

$$\begin{aligned} c_H (t_0):= \sup _{j \in {\mathbb {N}}} \int _{\Omega } \varepsilon _{i_j} \vert H ^{\varepsilon _{i_j}} \nabla \varphi ^{\varepsilon _{i_j}} \vert (x,t_0) \, \textrm{d}x <\infty \end{aligned}$$
(111)

hold by Theorem 13 and (81). Note that \(V_{t_0}\) is a countably \((d-1)\)-rectifiable varifold and determined by \(\mu _{t_0}\) uniquely from Theorem 14. We fix such \(t_0 >0\) and show the claim for \(\mu _{t_0}\). In this proof, even if we take a subsequence \(\varepsilon _{i_j}\), we always abbreviate \(\varepsilon _{i_j}\) by \(\varepsilon _i\) for simplicity. Set

$$\begin{aligned} A_{i,m}:= \left\{ x \in \Omega \mid \int _{B_r (x)} \varepsilon _i \vert H^{\varepsilon _i} \nabla \varphi ^{\varepsilon _i} \vert (x,t_0) \, \textrm{d}x \le m \mu _{t_0} ^{\varepsilon _i} (B_r (x)) \ \ \text {for any} \ r \in \left( 0, \frac{1}{2}\right) \right\} \end{aligned}$$

and

$$\begin{aligned} A_m:= \{ x \in \Omega \mid \text {there exists} \ x_i \in A_{i,m} \ \text {for any} \ i \in {\mathbb {N}}\ \text {such that} \ x_i \rightarrow x \} \end{aligned}$$

for any \(m \in {\mathbb {N}}\). Then the Besicovitch covering theorem implies

$$\begin{aligned} \mu _{t_0} ^{\varepsilon _i} (\Omega \setminus A_{i,m}) \le \frac{c(d) c_H (t_0)}{m}, \end{aligned}$$
(112)

where \(c(d) >0\) is a constant depending only on d. Set \( A:= \cup _{m=1} ^\infty A_m. \) Next we prove

$$\begin{aligned} \mu _{t_0} (\Omega \setminus A)=0. \end{aligned}$$
(113)

If (113) is not true, there exists a compact set \(K \subset \Omega {\setminus } A\) with \(\mu _{t_0} (K)>\frac{1}{2} \mu _{t_0} (\Omega {\setminus } A) >0\). Since \(A_1 \subset A_2 \subset A_3 \subset \cdots \), we have \(K \subset \Omega {\setminus } A_m\) for any \(m \in {\mathbb {N}}\). For any \(x \in K\), there is a neighborhood \(B_r (x)\) such that \(B_r (x) \cap A_{i,m} =\emptyset \) for sufficiently large i, by the definition of \(A_m\). This and the compactness of K imply that there exist an open set \(O_m\) and \(i _0 \in {\mathbb {N}}\) such that \(K \subset O_m\) and \(O_m \cap A_{i,m} =\emptyset \) for any \(i \ge i_0\). Let \(\phi _m \in C_c (O_m)\) be a non-negative test function such that \(0\le \phi _m \le 1\) and \(\phi _m =1\) on K. We compute

$$\begin{aligned} \begin{aligned} \mu _{t_0} (K) \le \int _\Omega \phi _m \, \textrm{d} \mu _{t_0} =&\, \lim _{i\rightarrow \infty } \int _\Omega \phi _m \, \textrm{d} \mu _{t_0} ^{\varepsilon _i} =\lim _{i\rightarrow \infty } \int _{\Omega \setminus A_{k,m}} \phi _m \, \textrm{d} \mu _{t_0} ^{\varepsilon _i}\\ \le&\, \liminf _{i\rightarrow \infty } \mu _{t_0} ^{\varepsilon _i} (\Omega \setminus A_{k,m}) \end{aligned} \end{aligned}$$
(114)

for any \(k \ge i_0\). Combining (112) and (114), we obtain \(\mu _{t_0} (K) =0\). Therefore we have proved (113).

By the rectifiability of \(\mu _{t_0}\) and (113), for \(\mu _{t_0}\) a.e. \(x \in \textrm{spt}\,\mu _{t_0}\), it has an approximate tangent space P and \(x \in A_m\) for some m. Fix such x. We may assume that \(x=0\) and \(P= \{ x \in {\mathbb {R}}^d \mid x_d=0 \}\) by a parallel translation and a rotation. Set \(\theta := \lim _{r \downarrow 0} \frac{\mu _{t_0} (B_r (0))}{\omega _{d-1} r^{d-1}}\). We only need to prove \(\theta \in {\mathbb {N}}\). Let \(\Phi _{r} (x) = \frac{x}{r}\) for \(r>0\) and \((\Phi _r)_\# V_{t_0}\) be the push-forward of the varifold (see (15)). Then for any positive sequence \(r_i \rightarrow 0\), we have \(\lim _{i\rightarrow \infty } (\Phi _{r_i})_\# V_{t_0} = \theta \vert P \vert \), where \(\vert P \vert \) is the unit density varifold generated by P. By the assumption \(0 \in A_m\), there exists \(\{ x_i \}_{i=1} ^\infty \) such that \(x_i \in A_{i,m}\) and \(x_i \rightarrow 0\) as \(i\rightarrow \infty \). Passing to a subsequence if necessary, we may assume that

$$\begin{aligned} \lim _{i\rightarrow \infty } \frac{x_i}{r_i}=0, \qquad \lim _{i\rightarrow \infty } \frac{\varepsilon _i}{r_i}=0, \end{aligned}$$
(115)

and

$$\begin{aligned} \lim _{i\rightarrow \infty } (\Phi _{r_i})_\# V_{t_0} ^{\varepsilon _i} = \theta \vert P \vert . \end{aligned}$$
(116)

Set \(u ^{{\tilde{\varepsilon }}_i } ({\tilde{x}}, {\tilde{t}}) = \varphi ^{\varepsilon _i} (x,t)\) and \(g ^{{\tilde{\varepsilon }}_i } ({\tilde{t}}) = r_i \lambda ^{\varepsilon _i} (t)\) for \({\tilde{x}}= \frac{x}{r_i}\), \({\tilde{t}}= \frac{t-t_0}{r^2 _i}\), and \({\tilde{\varepsilon }}_i = \frac{\varepsilon _i}{r_i}\) (another functions \({\tilde{\xi }} _{\tilde{\varepsilon }_i}\) and \({\tilde{H}} ^{{\tilde{\varepsilon }} _i}\) are defined in the same way). Note that \({\tilde{x}}_i:= \frac{x_i}{r_i} \rightarrow 0\) and \({\tilde{\varepsilon }} _i \rightarrow 0\) by (115) and \(u ^{\tilde{\varepsilon }_i }\) is a solution to (83) with \(\tilde{\varepsilon }_i\) instead of \({\tilde{\varepsilon }}\). We compute

$$\begin{aligned} \int _{B_3 (0)} \vert {\tilde{\xi }} _{{\tilde{\varepsilon }} _i} (\tilde{x},0) \vert \, \textrm{d} {\tilde{x}} = \frac{1}{r_i ^{d-1}} \int _{B_{3r_i} (0)} \vert \xi _{\varepsilon _i} (x,0) \vert \, \textrm{d}x. \end{aligned}$$

Thus, by (110) we may assume that

$$\begin{aligned} \lim _{i\rightarrow \infty } \int _{B_3 (0)} \vert {\tilde{\xi }} _{\tilde{\varepsilon }_i} ({\tilde{x}},0) \vert \, \textrm{d} {\tilde{x}}=0, \end{aligned}$$
(117)

passing to a subsequence if necessary. We compute

$$\begin{aligned} \begin{aligned}&{\tilde{\varepsilon }} _i \int _{B_3 (0)} \vert {\tilde{H}} ^{{\tilde{\varepsilon }} _i} \nabla _{{\tilde{x}}} {\tilde{u}} ^{{\tilde{\varepsilon }}_i} ({\tilde{x}}, 0 )\vert \, \textrm{d} {\tilde{x}} \\ =&\, \frac{\varepsilon _i}{r_i ^{d-2}} \int _{B_{3r_i} (0)} \vert H ^{ \varepsilon _i} \nabla \varphi ^{\varepsilon _i} (x, t_0 ) \vert \, \textrm{d} x \le \frac{m}{r_i ^{d-2}} \mu _{t_0} ^{\varepsilon _i} (B_{4r_i} (x_i)) \\ \le&\, 4^{d-1} m \omega _{d-1} D _2 r_i \rightarrow 0 \quad \text {as} \ i\rightarrow \infty , \end{aligned} \end{aligned}$$
(118)

where we used (52), (115), and \(x_i \in A_{i,m}\). Let \({\tilde{V}}_{{\tilde{t}}} ^{{\tilde{\varepsilon }} _i} \) be a varifold defined by (79) with \(u ^{{\tilde{\varepsilon }}_i}\) instead of \(\varphi ^\varepsilon \). Then \({\tilde{V}}_0 ^{\tilde{\varepsilon }_i} = (\Phi _{r_i})_\# V_{t_0} ^{\varepsilon _i}\). Next we show

$$\begin{aligned} \lim _{i\rightarrow \infty } \int _{B_3(0)} (1 - (\nu _d ^i) ^2 ) \tilde{\varepsilon }_i \vert \nabla _{{\tilde{x}}} u ^{{\tilde{\varepsilon }} _i} \vert ^2 \, \textrm{d} {\tilde{x}} \Big \vert _{{\tilde{t}}=0}=0, \end{aligned}$$
(119)

where \(\nu ^i = (\nu _1 ^i,\nu _2 ^i,\dots , \nu _d ^i) = \frac{\nabla _{{\tilde{x}}} u ^{{\tilde{\varepsilon }} _i}}{\vert \nabla _{{\tilde{x}}} u ^{{\tilde{\varepsilon }} _i} \vert }\). For \(S \in {\mathbb {G}}(d,d-1)\), set \(\psi (S):=1- \nu _d ^2\), where \(\nu \in {\mathbb {S}}^{d-1}\) be one of the unit normal vectors to S. Then \(\psi :{\mathbb {G}}(d,d-1) \rightarrow {\mathbb {R}}\) is well-defined, continuous, and \(\psi (P)=0\). Hence, for any \(\phi \in C_c ({\mathbb {R}}^d)\), \(\phi \psi \in C_c (G_{d-1}({\mathbb {R}}^d) )\) and

$$\begin{aligned} \begin{aligned} \lim _{i\rightarrow \infty } {\tilde{V}}_0 ^{{\tilde{\varepsilon }} _i} (\phi \psi ) =&\, \int \phi ({\tilde{x}} ) (1 - ( \nu _d ^i) ^2) \, d \Vert {\tilde{V}}_0 ^{{\tilde{\varepsilon }} _i} \Vert ({\tilde{x}})\\ =&\, \lim _{i\rightarrow \infty } (\Phi _{r_i})_\# V_{t_0} ^{\varepsilon _i} (\phi \psi ) = \theta \vert P \vert (\phi \psi ) \\ =&\, \theta \int \phi ({\tilde{x}}) \psi (P) \, d {\mathscr {H}} ^{d-1} ({\tilde{x}})=0, \end{aligned} \end{aligned}$$
(120)

where we used (116) and \(\psi (P)=0\). Thus (120) proves (119). We subsequently fix the subsequence and drop \({\tilde{\cdot }}\) and time variable (for example, we write \(u^{\tilde{\varepsilon }_i} ({\tilde{x}}, 0)\) as \(u^{\varepsilon _i} \)) for simplicity. We assume that \(N \in {\mathbb {N}}\) is the smallest positive integer grater than \(\theta \), namely,

$$\begin{aligned} \theta \in [N-1,N). \end{aligned}$$
(121)

Let \(s >0\) be an arbitrary number. Then Proposition 15 and (33) imply that there exists \(b >0\) such that

$$\begin{aligned} \int _{\{ x \in B_3 (0) \mid \vert u ^{\varepsilon _i} (x) \vert \ge 1-b \}} \frac{\varepsilon _i \vert \nabla u ^{\varepsilon _i} \vert ^2}{2} + \frac{W( u ^{\varepsilon _i})}{\varepsilon _i } \, dx \le s \end{aligned}$$
(122)

for sufficiently large i. Note that we may use Proposition 15 with \(t=0\) since \({\tilde{t}} = \frac{t -t_0}{r_i ^2}\) in this proof. For these \(s >0\), \(b>0\), and \(c>0\) given by Lemma 4, we choose \(\varrho \) and L given by Propositions 20 and 21 in the Appendix with \(R=2\) (we may restrict \(\varrho \) to be small if necessary). We choose \(a=L\varepsilon _i\) as a constant in Proposition 20. Set

$$\begin{aligned} \begin{aligned} G_i:=&\, B_2 (0) \cap \{ \vert u^{\varepsilon _i} \vert \le 1-b \} \\&\, \cap \{ x \mid \int _{B_r(x)} \varepsilon _i \vert H^{\varepsilon _i} \nabla u ^{\varepsilon _i} \vert + \vert \xi _{\varepsilon _i} \vert +(1-\nu _d ^2) \varepsilon _i \vert \nabla u ^{\varepsilon _i} \vert ^2 \, \textrm{d}x \le \varrho \mu _0 ^{\varepsilon _i} (B_r (x)) \\&\qquad \text {if} \ \varepsilon _i L \le r \le 1 \} \end{aligned}\nonumber \\ \end{aligned}$$
(123)

for sufficiently large i. The Besicovitch covering theorem, (117), (118), and (119) yield

$$\begin{aligned} \begin{aligned}&\mu _0 ^{\varepsilon _i} ( (B_2 \cap \{ \vert u^{\varepsilon _i} \vert \le 1-b \}) \setminus G_i ) \\ \le&\, \frac{c(d)}{\varrho } \int _{B_3 (0)} \varepsilon _i \vert H^{\varepsilon _i} \nabla u ^{\varepsilon _i} \vert + \vert \xi _{\varepsilon _i}\vert +(1-\nu _d ^2) \varepsilon _i \vert \nabla u ^{\varepsilon _i} \vert ^2 \, dx \rightarrow 0 \qquad \text {as} \ i\rightarrow \infty . \end{aligned} \end{aligned}$$
(124)

Next we show that for sufficiently large i

$$\begin{aligned} \begin{aligned} \frac{\mu _0 ^{\varepsilon _i} (B_r (x))}{ \omega _{d-1} r^{d-1} } \ge 1- 2s, \qquad \text {for any} \ x \in G_i \ \text {and} \ r \in [L\varepsilon _i, 1]. \end{aligned} \end{aligned}$$
(125)

Note that all the assumptions in Proposition 21 are satisfied by Lemma 4, (33), and (123). Thus we have (125) with \(r=L \varepsilon _i\). By integration by parts, we have

$$\begin{aligned} \begin{aligned} \frac{\textrm{d}}{\textrm{d}\tau } \left\{ \frac{1}{\tau ^{d-1}} \int _{B_\tau (x)} e_{\varepsilon _i} \, \textrm{d}y \right\} + \frac{1}{\tau ^{\textrm{d}}} \int _{B_\tau (x)} (\xi _{\varepsilon _i} +\varepsilon _i H^{\varepsilon _i} (y\cdot \nabla u ^{\varepsilon _i})) \, \textrm{d}y \\ -\frac{\varepsilon _i}{\tau ^{d+1}} \int _{\partial B_\tau (x)} (y\cdot \nabla u ^{\varepsilon _i}) ^2 d {\mathscr {H}}^{d-1} (y) =0. \end{aligned} \end{aligned}$$

Thus we can compute

$$\begin{aligned} \begin{aligned} \frac{1}{ \sigma \tau ^{d-1}} \int _{B_\tau (x)} e_{\varepsilon _i} \, \textrm{d}y \Big \vert _{\tau =L\varepsilon _i} ^r \ge&\,- \int _{L\varepsilon _i} ^r \frac{1}{\sigma \tau ^{d}} \int _{B_\tau (x)} \varepsilon _i H^{\varepsilon _i} (y\cdot \nabla u^{\varepsilon _i}) \, \textrm{d}y \textrm{d}\tau \\ \ge&\, - \int _{L\varepsilon _i} ^1 \frac{1}{\sigma \tau ^{d}} \int _{B_\tau (x)} \varepsilon _i \tau \vert H^{\varepsilon _i} \nabla u^{\varepsilon _i} \vert \, \textrm{d}y \textrm{d}\tau \\ \ge&\, -\frac{\varrho D_2}{\sigma }, \end{aligned} \end{aligned}$$

where we used (52), (123), and \(\xi _{\varepsilon _i} \le 0\). Therefore we obtain (125) for sufficiently large i by restricting \(\varrho \) to be small. Let \(\delta >0\) and \(\phi \in C_c (B_3 (0))\) be a non-negative test function such that \(\phi =1\) on \(B_2(0) \cap \{ \vert x_d \vert > \delta \}\). Then there exists \(i_0 \ge 1\) such that

$$\begin{aligned} \mu _0 ^{\varepsilon _i} (\phi ) \le (1-2s) \omega _{d-1} \frac{\delta ^{d-1}}{2}, \qquad \text {for any} \ i \ge i_0, \end{aligned}$$
(126)

since \(\mu _0 ^{\varepsilon _i} =\Vert V_0 ^{\varepsilon _i} \Vert \rightharpoonup \theta {\mathscr {H}}^{d-1} \lfloor _{P}\). Assume that \(x \in G_i \cap \{ \vert x_d \vert > 2 \delta \}\) for \(i \ge i_0\). Then (125) and (126) imply

$$\begin{aligned}{} & {} (1-2s) \omega _{d-1} \delta ^{d-1} \le \mu _0 ^{\varepsilon _i} (B _{\delta _1} (x)) \le \mu _0 ^{\varepsilon _i} (\phi ) \le (1-2s) \omega _{d-1} \frac{\delta ^{d-1}}{2}, \\{} & {} \qquad \text {for any} \ i \ge i_0. \end{aligned}$$

This is a contradiction. Thus

$$\begin{aligned} \textrm{dist}\,(P, G_i) \rightarrow 0 \qquad \text {as} \ i \rightarrow \infty . \end{aligned}$$
(127)

Set \(Y:=P^{-1} (x) \cap G_i \cap \{ x \mid u ^{\varepsilon _i} (x) =l \}\) for \(x \in P \cap B_1 (0)\). Next we show that for sufficiently large i

$$\begin{aligned} \# Y \le N-1, \quad \text {for any} \ x\in P \cap B_1 (0) \ \text {and} \ \vert l \vert \le 1-b. \end{aligned}$$
(128)

For a contradiction, assume that \(\# Y \ge N\) and choose \(y_j \in Y\) for \(j=1,2,\dots , N\). We use Proposition 20 with \(R=1\), \(a=L\varepsilon _i\) and \(Y' =\{ y_j \}_{j=1} ^N\) instead of Y. Note that the smallness of \(\textrm{diam}\,Y'\) is true from (127) and \(\vert y_j -y_k \vert >3 L\varepsilon _i\) for any \(1\le j < k \le N\) holds by (159). Then (158) yields

$$\begin{aligned} \sum _{j=1} ^N \frac{1}{(L\varepsilon _i )^{d-1}} \mu _0 ^{\varepsilon _i} (B_{L\varepsilon _i} (y_j)) \le s+ (1+s) \mu _0 ^{\varepsilon _i} (\{ z \mid \textrm{dist}\,(z,Y') <1 \}) \end{aligned}$$
(129)

for sufficiently large i. By (127) and \(\mu _0 ^{\varepsilon _i} =\Vert V_0 ^{\varepsilon _i} \Vert \rightharpoonup \theta {\mathscr {H}}^{d-1} \lfloor _{P}\),

$$\begin{aligned} \limsup _{i\rightarrow \infty } \mu _0 ^{\varepsilon _i} (\{ z \mid \textrm{dist}\,(z,Y') <1 \}) \le \theta {\mathscr {H}} ^{d-1} \lfloor _{P} (\overline{B_1 (0)}) = \theta \omega _{d-1}. \end{aligned}$$

By this, \(\# Y' = N\), (125), and (129) we have

$$\begin{aligned} N\omega _{d-1} (1-2s) \le s+(1+s) \theta \omega _{d-1}. \end{aligned}$$

However, this contradicts (121) by restricting s to be small. Thus (128) holds for sufficiently large i.

Finally, we complete the proof. Set \({\hat{V}}_0 ^{\varepsilon _i}:= V_0 ^{\varepsilon _i} \lfloor _{\{ \vert x_d \vert \le 1 \} \times {\mathbb {G}}(d,d-1)}\). We regard P as a diagonal matrix \((p_{jk})\) with \(p_{kk}=1\) for \(1\le k\le d-1\) and \(p_{dd}=0\). Then the push-forward of \({\hat{V}}_0 ^{\varepsilon _i}\) by P is given by

$$\begin{aligned} \begin{aligned} P_{\#} {\hat{V}}_0 ^{\varepsilon _i} (\phi ) =&\, \int _{\{ \vert x_d \vert \le 1 \}} \phi (Px, \nabla Px \circ (I -\nu ^i \otimes \nu ^i)) \vert \Lambda _{d-1} \nabla Px \circ (I -\nu ^i \otimes \nu ^i) \vert \, \textrm{d} \mu _0 ^{\varepsilon _i} \\ =&\, \int _{\{ \vert x_d \vert \le 1 \}} \phi (Px, P \circ (I -\nu ^i \otimes \nu ^i)) \vert \nu ^i _d \vert \, \textrm{d} \mu _0 ^{\varepsilon _i} \end{aligned} \end{aligned}$$

for any \(\phi \in C_c (P \cap B_2 (0) \times {\mathbb {G}}(d,d-1))\). Here \(\vert \Lambda _{d-1} \nabla Px \circ (I -\nu ^i \otimes \nu ^i) \vert \) is the Jacobian and \(\nu ^ i _d = \frac{\partial _{x_d} u ^{\varepsilon _i}}{\vert \nabla u ^{\varepsilon _i} \vert }\). Due to (116), \(P_\# {\hat{V}}_0 ^{\varepsilon _i} \rightharpoonup P_\# (\theta {\mathscr {H}}^{d-1} \lfloor _{P}) =\theta {\mathscr {H}}^{d-1} \lfloor _{P}\) as \(i\rightarrow \infty \). By (117),

$$\begin{aligned} \lim _{i\rightarrow \infty } \int _{B_3 (0)} \left| \frac{\varepsilon _i \vert \nabla u ^{\varepsilon _i} \vert ^2}{2} + \frac{W( u ^{\varepsilon _i})}{\varepsilon _i } - \sqrt{2W(u^{\varepsilon _i})} \vert \nabla u^{\varepsilon _i}\vert \right| \, \textrm{d}x =0 \end{aligned}$$
(130)

holds (see (133) below). We compute

$$\begin{aligned} \begin{aligned} \omega _{d-1} \theta =&\, {\mathscr {H}}^{d-1} \lfloor _{P} (B_1 (0)) \le \liminf _{i\rightarrow \infty } \Vert P_\# {\hat{V}}_0 ^{\varepsilon _i} \Vert (B_1 (0)) = \liminf _{i\rightarrow \infty } \int _{B_1 (0)} \vert \nu _d ^i \vert \, \textrm{d} \mu _0 ^{\varepsilon _i} \\ \le&\, \liminf _{i\rightarrow \infty } \int _{B_1 (0) \cap G_i } \vert \nu _d ^i \vert \, \textrm{d} \mu _0 ^{\varepsilon _i} + 2s \\ \le&\, \liminf _{i\rightarrow \infty } \frac{1}{\sigma } \int _{B_1 (0) \cap G_i } \vert \nu _d ^i \vert \sqrt{2W(u^{\varepsilon _i})} \vert \nabla u^{\varepsilon _i} \vert \, \textrm{d}x + 2s, \end{aligned} \end{aligned}$$
(131)

where we used (122), (124), and (130). By the co-area formula and the area formula, we have

$$\begin{aligned} \begin{aligned}&\int _{B_1 (0) \cap G_i } \vert \nu _d ^i \vert \sqrt{2W(u^{\varepsilon _i})} \vert \nabla u^{\varepsilon _i} \vert \, \textrm{d}x \\ =&\, \int _{-1+b} ^{1-b} \sqrt{2W(\tau )} \int _{ B_1 (0) \cap G_i \cap \{ u ^{\varepsilon _i } =\tau \} } \vert \Lambda _{d-1} \nabla Px \circ (I -\nu ^i \otimes \nu ^i) \vert \, d{\mathscr {H}}^{d-1} (x) \textrm{d}\tau \\ =&\, \int _{-1+b} ^{1-b} \sqrt{2W(\tau )} \int _{B_1 (0) \cap \{ x_d=0 \} } {\mathscr {H}}^{0} ( \{ u ^{\varepsilon _i} =\tau \} \cap G_i \cap P^{-1} (x) ) \, d{\mathscr {H}}^{d-1} (x) \textrm{d}\tau \\ \le&\, \int _{-1+b} ^{1-b} \sqrt{2W(\tau )} \int _{B_1 (0) \cap \{ x_d=0 \} } (N-1) \, \textrm{d}{\mathscr {H}}^{d-1} (x) \textrm{d}\tau \\ \le&\, \sigma (N-1) \omega _{d-1}, \end{aligned}\nonumber \\ \end{aligned}$$
(132)

where we used (128) and \(\sigma = \int _{-1} ^{1} \sqrt{2W (\tau )} \, d\tau \). Hence \( \theta \le N-1 \) due to (131) and (132) and the arbitrariness of s. By this and (121), \(\theta =N-1\). \(\square \)

5 Proofs of Main Theorems

In this section we prove Theorem 3 and Theorem 4 on the existence of the weak solution in the sense of \(L^2\)-flow and distributional BV-solution.

Proof of Theorem 3

Let \(\{ \varphi ^{\varepsilon _i} _0 \} _{i=1} ^\infty \) be a family of functions such that all the claims of Proposition 2 are satisfied. Then one can check that all the assumptions in Sections 3 and 4 are fulfilled. Therefore (a) holds by Propositions 2, 9, and 10. By taking a subsequence \(\varepsilon _i \rightarrow 0\), we obtain (b) (the proof is standard and is exactly the same as that in [43], so we omit it). By Lemma 1 and the weak compactness of \(L^2 (0,T)\), we may take a subsequence \(\varepsilon _i \rightarrow 0\) such that (c) holds (for the weak convergence for all \(T>0\), we only need to use the diagonal argument).

Next we show (d). We compute

$$\begin{aligned} \frac{\varepsilon \vert \nabla \varphi ^\varepsilon \vert ^2}{2} + \frac{W(\varphi ^\varepsilon )}{\varepsilon } - \sqrt{2W(\varphi ^\varepsilon )} \vert \nabla \varphi ^\varepsilon \vert =\left( \frac{\sqrt{\varepsilon } \vert \nabla \varphi ^\varepsilon \vert }{\sqrt{2}} - \frac{\sqrt{W(\varphi ^\varepsilon )}}{\sqrt{\varepsilon }} \right) ^2 \le \vert \xi _\varepsilon \vert .\nonumber \\ \end{aligned}$$
(133)

Set \(\textrm{d} {\hat{\mu }} ^\varepsilon := \frac{1}{\sigma }\sqrt{2W(\varphi ^\varepsilon )} \vert \nabla \varphi ^\varepsilon \vert \, \textrm{d}x\textrm{d}t\). By (133), Proposition 9, and Theorem 13, we have

$$\begin{aligned} {\hat{\mu }} ^\varepsilon \rightharpoonup \mu \qquad \text {as Radon measures}, \end{aligned}$$
(134)

where \(\textrm{d}\mu := \textrm{d}\mu _t \textrm{d}t\). By (35), (38), (133), and (134), we obtain

$$\begin{aligned} \sup _{i \in {\mathbb {N}}} \int _{\Omega \times (0,T)} \vert \lambda ^{\varepsilon _i} \vert ^2 \, \textrm{d} {\hat{\mu }} ^{\varepsilon _i} \le \sup _{i \in {\mathbb {N}}} \int _{\Omega \times (0,T)} \vert \lambda ^{\varepsilon _i} \vert ^2 \, \textrm{d} {\mu } ^{\varepsilon _i} _t dt \le D_1 C_3 (1+T). \end{aligned}$$

Then there exist \(\vec {f} \in (L_{loc} ^2 (\mu ))^d\) and the subsequence \(\varepsilon _i \rightarrow 0\) such that

$$\begin{aligned} \begin{aligned} \frac{1}{\sigma } \int _{\Omega \times (0,T) } - \lambda ^{\varepsilon _i} \sqrt{2W(\varphi ^{\varepsilon _i} )} \nabla \varphi ^{\varepsilon _i} \cdot \vec {\phi } \, \textrm{d}x\textrm{d}t =&\, \int _{\Omega \times (0,T) \cap \{ \vert \nabla \varphi ^{\varepsilon _i}\vert \not = 0 \} } - \lambda ^{\varepsilon _i} \frac{\nabla \varphi ^{\varepsilon _i}}{\vert \nabla \varphi ^{\varepsilon _i}\vert } \cdot \vec {\phi } \, \textrm{d}{\hat{\mu }} ^{\varepsilon _i}\\ \rightarrow&\, \int _{\Omega \times (0,T)} \vec {f} \cdot \vec {\phi } \, \textrm{d} \mu _t \textrm{d}t \end{aligned} \end{aligned}$$

for any \(\vec {\phi } \in C_c (\Omega \times [0,T); {\mathbb {R}}^d)\) (see [21, Theorem 4.4.2]). Moreover, if \(\vec {\phi } \) is smooth, we have

$$\begin{aligned} \begin{aligned}&\int _{\Omega \times (0,T)} \vec {f} \cdot \vec {\phi } \, \text {d} \mu = \lim _{i\rightarrow \infty } \frac{1}{\sigma } \int _{\Omega \times (0,T) } - \lambda ^{\varepsilon _i} \sqrt{2W(\varphi ^{\varepsilon _i} )} \nabla \varphi ^{\varepsilon _i} \cdot \vec {\phi } \, \text {d}x\text {d}t \\ =&\, \lim _{i\rightarrow \infty } \frac{1}{\sigma } \int _{\Omega \times (0,T) } \lambda ^{\varepsilon _i} k( \varphi ^{\varepsilon _i}) \textrm{div}\, \vec {\phi } \, \text {d}x\text {d}t = \int _0 ^T \lambda \int _{\Omega } \psi \textrm{div}\, \vec {\phi } \, \text {d}x\text {d}t \\ =&\, - \int _0 ^T \lambda \int _{\Omega } \vec {\phi } \cdot \nu \, \text {d} \Vert \nabla \psi (\cdot , t)\Vert \text {d}t, \end{aligned} \end{aligned}$$

where we used (c), \(k'(s) =\sqrt{2W(s)}\), \(\lim _{i\rightarrow \infty } k(\varphi ^{\varepsilon _i}) = \sigma (\psi -\frac{1}{2}) \) for a.e. (xt), and the dominated convergence theorem. Hence we have (25).

Now we prove (e). By replacing \(d{{\hat{\mu }}} ^{\varepsilon }\) with \(d {\tilde{\mu }}^\varepsilon := \frac{\varepsilon }{\sigma } \vert \nabla \varphi ^{\varepsilon }\vert ^2 \, dxdt\), the convergence (26) is obtained in the same way as (d). In addition, for any \(\vec {\phi } \in C_c (\Omega \times [0,T); {\mathbb {R}}^d)\), we compute

$$\begin{aligned} \begin{aligned}&\int _{0} ^T \int _{\Omega } \vec {v} \cdot \vec {\phi } \, \textrm{d} \mu _t \textrm{d}t = \lim _{i \rightarrow \infty } \int _0 ^T \int _{\Omega } \vec {v}^{ \, \varepsilon _i} \cdot \vec {\phi } \, \textrm{d}\mu _t ^{\varepsilon _i} \textrm{d}t = \lim _{i \rightarrow \infty } \int _0 ^T \int _{\Omega } \vec {v}^{ \, \varepsilon _i} \cdot \vec {\phi } \, d {\tilde{\mu }} ^{\varepsilon _i}\\ =&\, \lim _{i \rightarrow \infty } \int _0 ^T \int _{\Omega } -\varepsilon _i \left( \Delta \varphi ^{\varepsilon _i} -\frac{W' (\varphi ^{\varepsilon _i})}{\varepsilon _i ^2} + \lambda ^{\varepsilon _i} \frac{\sqrt{2W(\varphi ^{\varepsilon _i})}}{\varepsilon _i} \right) \nabla \varphi ^{\varepsilon _i} \cdot \vec {\phi } \, \textrm{d}x\textrm{d}t \\ =&\, \int _0 ^T \int _{\Omega } (\vec {h} +\vec {f}) \cdot \vec {\phi } \, \textrm{d}\mu _t \textrm{d}t, \end{aligned} \end{aligned}$$

where we used Theorem 14 and (d). Thus \(\vec {v}=\vec {h} +\vec {f}\). One can check that \(\{ \mu _t \}_{t \in [0,\infty )}\) is an \(L^2\)-flow with the generalized velocity vector \(\vec {v}\) (see [43, Proposition 4.3] for the inequality (16) and [34, Lemma 6.3] for the perpendicularity). \(\square \)

To prove Theorem 4, we use the next proposition and lemmas. In the original proof of the proposition, \(2\le d\le 3\) is assumed to use the results of [38, Proposition 4.9, Theorem 5.1]. however, we already know that \(\vert \xi \vert =0\) and \(\mu _t\) is integral for a.e. t, so we can show the claim in the same way.

Proposition 17

(See Proposition 4.5 of [34]) Let \(\psi \), \(\vec {v}\), and \(\vec {\nu }\) are given by Theorem 3. Then \(\int \vert \vec {v} \cdot \vec {\nu } \vert \textrm{d} \Vert \nabla \psi (\cdot , t)\Vert \textrm{d}t <\infty \) and

$$\begin{aligned} \int _0 ^T \int _{\Omega } \phi \vec {v} \cdot \vec {\nu } \, \textrm{d} \Vert \nabla \psi (\cdot , t)\Vert \textrm{d}t = \int _0 ^T \int _{\Omega } \partial _t \phi \psi \, \textrm{d}x\textrm{d}t \end{aligned}$$
(135)

for any \(\phi \in C_c ^1 (\Omega \times (0,T))\) and for any \(T>0\).

Proof

Set \(\nabla _{x,t} = (\nabla , \partial _t)\) in the sense of BV. One can check that \(\Vert \nabla _{x,t} \psi \Vert \ll \mu \), \( \mu \lfloor _{\partial ^*\{ \psi =1 \}}\) is rectifiable, \(\int \vert \vec {v} \cdot \vec {\nu } \vert \, d \Vert \nabla \psi (\cdot , t)\Vert dt <\infty \), and the approximate tangent space coincides with that of \(\Vert \nabla _{x,t} \psi \Vert \) for \(\mu \)-a.e. and \(\Vert \nabla _{x,t} \psi \Vert \)-a.e. (see [34, Proposition 8.1–8.3] and [2, Proposition 2.85]). By this and Proposition 1, we have

$$\begin{aligned} \begin{aligned} 0= & {} \int _0 ^T \int _{\Omega } \phi (\vec {v},1) \cdot \vec \nu _{x,t} \, \textrm{d}\Vert \nabla _{x,t} \psi \Vert = \int _0 ^T \int _{\Omega } \phi \vec {v} \cdot \vec {\nu } \, d \Vert \nabla \psi (\cdot , t)\Vert \textrm{d}t \\{} & {} \quad + \int _0 ^T \int _{\Omega } \phi \partial _t \psi \, \textrm{d}x\textrm{d}t \end{aligned} \end{aligned}$$

for any \(\phi \in C_c ^1 (\Omega \times (0,T))\), where \(\vec \nu _{x,t}\) is the inner unit normal vector of \(\{ (x,t) \mid \psi (x,t)=1 \}\). Therefore we obtain (135). \(\square \)

Lemma 11

Let \(\gamma \) and \(\delta \) be positive constants with \(\delta < \gamma \). Under the same assumptions of Theorem 4, there exist \(T_2 \in ( 0,1)\) and \(\varepsilon _5 \in (0,1)\) depending only on \(\gamma \), \(\delta \), and \(C_3(\omega ,d,D' _1)\) with the following property. Let \(g: {\mathbb {R}}\rightarrow [0,\infty )\) be a smooth even function such that \(g(0)=0\), \(0 \le g'' (s) \le 2\) for any \(s \in {\mathbb {R}}\), and \(g(s)=\vert s \vert -\frac{1}{2}\) if \(\vert s \vert \ge 1\). Set \(g^\delta (s):= \delta g(s/\delta )\) and define \(\tilde{r}^{\varepsilon ,\delta } \in C^\infty ({\mathbb {R}}^d \times [0,\infty ))\) by

$$\begin{aligned} {\tilde{r}}^{\varepsilon ,\delta } (x,t):= g^\delta (x_1) + \int _0 ^t \lambda ^\varepsilon (\tau ) \, \textrm{d}\tau + 2\delta ^{-1} t - \gamma , \end{aligned}$$

where \(\lambda ^\varepsilon \) is given by (6). Set \({\tilde{\phi }}^{\varepsilon ,\delta }:= q^\varepsilon (\tilde{r}^{\varepsilon ,\delta } )\) and assume that \({\tilde{\phi }} ^\varepsilon (x,0) \ge \varphi ^\varepsilon _0 (x) \) for any \(x \in {\mathbb {R}}^d\). Then

$$\begin{aligned} {\tilde{\phi }} ^{\varepsilon ,\delta } \ge \varphi ^\varepsilon \qquad \text {in} \ {\mathbb {R}}^d \times [0,T_2) \end{aligned}$$
(136)

for any \(\varepsilon \in (0,\varepsilon _5)\).

Proof

We denote \({\tilde{r}} ={\tilde{r}}^{\varepsilon ,\delta }\) for simplicity. By (34) and the comparison principle, we only need to prove

$$\begin{aligned} \partial _t {\tilde{r}} \ge \Delta {\tilde{r}} - \dfrac{2 q ^\varepsilon ({\tilde{r}}) }{\varepsilon } ( \vert \nabla {\tilde{r}} \vert ^2 -1) + \lambda ^\varepsilon \qquad \text {in} \ {\mathbb {R}}^d \times (0,T_2) \end{aligned}$$
(137)

for sufficiently small \(T_2>0\) and \(\varepsilon >0\), since \(\tilde{\phi }^{\varepsilon ,\delta } \ge \varphi ^\varepsilon \) if and only if \({\tilde{r}} \ge r ^\varepsilon \). In the case of \(\vert x_1 \vert \ge \delta \), (137) holds by \(\partial _t {\tilde{r}} = 2\delta ^{-1} + \lambda ^\varepsilon \), \(\vert \nabla {\tilde{r}} \vert =1\), and \(\Delta {\tilde{r}} =0\). Next we consider the case of \(\vert x_1 \vert \le \delta \). Set \(O_\delta :=\{ x \in {\mathbb {R}}^d \mid \vert x_1 \vert \le \delta \}\). By \({\tilde{r}} (x,0) \le -\gamma +\frac{\delta }{2} \) on \(O_\delta \) and

$$\begin{aligned} \vert {\tilde{r}} (x,t) -{\tilde{r}} (x,0) \vert \le \int _0 ^t \vert \lambda ^\varepsilon (\tau ) \vert \, d\tau + 2\delta ^{-1} t \le \sqrt{C_3 (1+t)} \sqrt{t} + 2\delta ^{-1} t, \end{aligned}$$

there exists \(T_2 >0\) such that

$$\begin{aligned} {\tilde{r}} (x,t) \le -\frac{\gamma }{4}<0 \qquad \text {for any} \ (x,t) \in O_\delta \times [0,T_2). \end{aligned}$$
(138)

By (138) and \(\vert \nabla {\tilde{r}} \vert \le 1\), \(\dfrac{2 q ^\varepsilon ({\tilde{r}}) }{\varepsilon } ( \vert \nabla {\tilde{r}} \vert ^2 -1) \ge 0\). By using this, for any \((x,t) \in O_\delta \times [0,T_2)\),

$$\begin{aligned} \partial _t {\tilde{r}} - \Delta {\tilde{r}} + \dfrac{2 q ^\varepsilon ({\tilde{r}}) }{\varepsilon } ( \vert \nabla {\tilde{r}} \vert ^2 -1) - \lambda ^\varepsilon \ge 2\delta ^{-1} - \Delta {\tilde{r}} \ge 0, \end{aligned}$$
(139)

where we used \(\Delta {\tilde{r}} \le \delta ^{-1} g'' (x_1 /\delta ) \le 2 \delta ^{-1}\). Therefore we obtain (137). \(\square \)

Lemma 12

Let \(r \in (0,\frac{1}{4})\). Then there exists \(T_3 >0\) depending only on d and r with the following property. Let \(U_0 \subset \subset (\frac{1}{4},\frac{3}{4})^d\) satisfies \({\mathscr {L}}^d (U_0) ={\mathscr {L}}^d (B_{r} (0))\) and has a \(C^1\) boundary \(M_0\) with (27) for \(\delta _1 >0\). In addition, we assume \({\mathscr {H}}^{d-1} (M_0) \le 2 {\mathscr {H}}^{d-1} (\partial B_{r} (0))\). Then we have

$$\begin{aligned} 0\le \mu _0 (\Omega ) - \mu _t ( \Omega ) \le \delta _1 \qquad \text {for any} \ t \in [0,T_3), \end{aligned}$$
(140)

where \(\mu _t\) is a weak solution to (1) with initial data \(M_0\).

Proof

First we claim that there exists \(T_3 >0\) depending only on d and r such that

$$\begin{aligned} U_t = \{ x \in (0,1)^d \mid \psi (x,t)=1 \} \subset \left( \frac{1}{10}, \frac{9}{10}\right) ^d \end{aligned}$$
(141)

for any \( t \in [0,T_3)\), where \(\psi = \lim _{i\rightarrow \infty } \psi ^{\varepsilon _i} = \lim _{i\rightarrow \infty } \frac{1}{2} ( \varphi ^{\varepsilon _i} +1)\). Let \({\tilde{\phi }} ^{\varepsilon ,\delta }\) be a function given by Lemma 11 with \(\gamma = \frac{1}{10}\) and \(\delta = \frac{1}{20}\). By (136) and (138), one can check that there exists \(T_3=T_3 (C_3(\omega ,d,D' _1)) >0\) such that \(\lim _{ i\rightarrow \infty }\varphi ^{\varepsilon _i} (x,t) = -1 \) on \(\{ x \in {\mathbb {R}}^d \mid \vert x_1 \vert \le \frac{1}{10} \}\) for any \(t \in [0,T_3)\). Note that \(\omega \) and \(D' _1\) depend only on r by \({\mathscr {L}}^d (U_0) ={\mathscr {L}}^d (B_{r} (0))\) and \({\mathscr {H}}^{d-1} (M_0) \le 2 {\mathscr {H}}^{d-1} (\partial B_{r} (0))\). Hence \(T_3\) depends only on d and r. Therefore \(U_t \cap \{ x \in {\mathbb {R}}^d \mid \vert x_1 \vert \le \frac{1}{10} \} =\emptyset \) for any \(t \in [0,T_3)\). Similarly we have (141). Thus \(\partial ^*(U_t \cap (0,1)^d) = \partial ^*U_t\) for any \(t \in [0,T_3)\). Hence, by using the isoperimetric inequality for Caccioppoli sets (see [13, 46]), and (b3) and (b4) of Theorem 3, we have

$$\begin{aligned}{} & {} d \omega _d ^{\frac{1}{d}} ({\mathscr {L}}^d (U_0))^{\frac{d-1}{d}}= d \omega _d ^{\frac{1}{d}} ({\mathscr {L}}^d (U_t))^{\frac{d-1}{d}} \le {\mathscr {H}}^{d-1} (\partial ^*U_t) \le \mu _t (\Omega )\nonumber \\{} & {} \quad \text {for any} \ t \in [0,T_3). \end{aligned}$$
(142)

By (27) and (142), we obtain (140). \(\square \)

Finally we prove Theorem 4.

Proof of Theorem 4

First we show (a). From (37),

$$\begin{aligned} \int _ {0} ^{T} \vert \lambda ^\varepsilon (t) \vert ^2 \, dt \le C_2 (\mu _0 ^\varepsilon (\Omega ) -\mu _{T} ^\varepsilon (\Omega ) + T ) \end{aligned}$$

for any \(T>0\) and for any \(\varepsilon \in (0,\varepsilon _1)\). By this and (140), we can choose \(\delta _1 = \delta _1 (C_2(\omega ,d,D' _1)) >0\) so that

$$\begin{aligned} \limsup _{i\rightarrow \infty } e^{\frac{1}{2} \int _0 ^{T_4} \vert \lambda ^{\varepsilon _i} \vert ^2 \, dt} \le \limsup _{i \rightarrow \infty } e^{ \frac{1}{2} C_2 \delta _1 } e^{\frac{1}{2} C_2 T_4} \le \frac{10}{9}, \end{aligned}$$
(143)

where \(T_4 = T_4 (d,r) = \min \{ T_3, \frac{2}{ C_2 } \log \frac{100}{99} \} >0\) and \(\delta _1\) also depends only on d and r since \({\mathscr {L}}^d (U_0) ={\mathscr {L}}^d (B_{r} (0))\) and \({\mathscr {H}}^{d-1} (M_0) \le 2 {\mathscr {H}}^{d-1} (\partial B_{r} (0))\). Then (51) and (143) imply

$$\begin{aligned} \int _{{\mathbb {R}}^d} \rho _{(y,s)} (x,t) \, \textrm{d}\mu _t (x) \le \frac{10}{9} \int _{{\mathbb {R}}^d} \rho _{(y,s)} (x,0) \, \textrm{d}\mu _0 (x) \end{aligned}$$
(144)

for any \(y \in {\mathbb {R}}^d\), \(t\in [0,T_4)\), and \(s>0\) with \(0\le t <s\). Recall that \(\rho _{(y,s)} (x,0)\) converges to \((d-1)\)-dimensional delta function at y as \(s\downarrow 0\). Therefore, since \(M_0\) is \(C^1\), we may assume that there exists \(s_0>0\) depending only on \(M_0\) such that

$$\begin{aligned} \int _{{\mathbb {R}}^d} \rho _{(y,s)} (x,0) \, d \mu _0 \le \frac{3}{2} \end{aligned}$$
(145)

for any \(y\in {\mathbb {R}}^d\) and \(s \in (0,s_0)\). Set \(T_1= T_1 (d,r,M_0):= \min \{ T_4, s_0 \}\). Let \(t_0 \in (0,T_1)\) be a number such that \(\mu _{t_0}\) is integral. Then there exist a countably \((d-1)\)-rectifiable set \(M_{t_0}\) and \(\theta _{t_0} \in L ^1 _{loc} ({\mathscr {H}}^{d-1} \lfloor _{M_{t_0}})\) such that \(\mu _{t_0} = \theta _{t_0} {\mathscr {H}}^{d-1} \lfloor _{M_{t_0}}\). Assume that there exist \(x_0 \in M_{t_0}\) and \(N\ge 2\) such that \(M_{t_0}\) has an approximate tangent space at \(x_0\) and \(\displaystyle \lim _{r\rightarrow 0} \frac{ \mu _{t_0} (B_r (x_0)) }{\omega _{d-1} r^{d-1}} =\theta _{t_0} (x_0) =N\). Set \(r=\sqrt{2(s-t_0)}\) for \(t_0<s <T_1\). Using the same calculation as (155), for any \(\delta \in (0,1)\), we obtain

$$\begin{aligned} \begin{aligned}&\int _{{\mathbb {R}}^d} \rho _{(x_0,s)} (x,t) \, d\mu _{t_0} \ge \frac{1}{(\sqrt{2\pi } r)^{d-1}} \int _\delta ^1 \mu _{t_0} (B_{\sqrt{2r^2 \log \frac{1}{k} }} (x_0)) \, \textrm{d}k \\ \rightarrow&\, \frac{N \omega _{d-1}}{\pi ^{\frac{d-1}{2}}} \int _\delta ^1 \left( \log \frac{1}{k} \right) ^{\frac{d-1}{2}} \, \textrm{d}k \qquad \text {as} \ r\rightarrow 0 \ (s\downarrow t_0). \end{aligned} \end{aligned}$$

By this and \( \int _0 ^1 \left( \log \frac{1}{k} \right) ^{\frac{d-1}{2}} \, \textrm{d}k =\Gamma (\frac{d-1}{2} +1) =\pi ^{\frac{d-1}{2}} /\omega _{d-1} \), we have

$$\begin{aligned} \begin{aligned} \lim _{s\downarrow t_0} \int _{{\mathbb {R}}^d} \rho _{(x_0,s)} (x,t) \, \textrm{d}\mu _{t_0} =N. \end{aligned} \end{aligned}$$

Then we have a contradiction by (144) and (145). Therefore \(\theta _{t_0} (x)=1\) \({\mathscr {H}}^{d-1}\)-a.e. on \(M_{t_0}\). By an argument similar to that in [45, Theorem 2.2 (2d)], we have \({\mathscr {H}}^{d-1} (\partial ^*U_{t_0} {\setminus } M_{t_0})=0\) and \({\mathscr {H}}^{d-1} (M_{t_0} {\setminus } \partial ^*U_{t_0})=0\) because \(\theta _{t_0} (x)\) is an even integer for \({\mathscr {H}}^{d-1}\)-a.e. \(x \in M_{t_0} {\setminus } \partial ^*U_{t_0}\). Hence we obtain (a).

The claim (b1) and (b2) are clear and (b3) is also obvious by Remark 4 and \(\mu _t= \Vert \nabla \psi (\cdot ,t)\Vert \) for a.e. \(t \in (0,T_1)\). By (133), we have (b4).

Next we prove (b5). By (135), for any \(\phi \in C_c ^1 ((0,T))\), we compute

$$\begin{aligned} \begin{aligned} \int _0 ^{T_1} \phi \int _{\Omega } \vec v \cdot \vec \nu \, \textrm{d} \Vert \nabla \psi (\cdot ,t) \Vert \textrm{d}t =\int _0 ^{T_1} \partial _t \phi \int _{\Omega } \psi \, \textrm{d}x \textrm{d}t =0, \end{aligned} \end{aligned}$$

where we used (b3) of Theorem 3. Thus \(\int _{\Omega } \vec v \cdot \vec \nu \, d \Vert \nabla \psi (\cdot ,t) \Vert =0\) for a.e. \(t \in (0,T_1)\).

Now we prove (b6). Set \(d\nu :=d{\mathscr {H}}^{d-1} \lfloor _{\partial ^*U_t} dt\). Since the space \(C_c (\Omega )\) is dense in \(L^2 (\nu )\), for any \(\eta \in C_c ((0,T_1))\) we have

$$\begin{aligned} \begin{aligned} 0 =&\, \int _0 ^{T_1} \int _{\partial ^*U_t} \{ \vec v -\vec h +\lambda \vec \nu \} \cdot \vec \nu \eta \, d {\mathscr {H}}^{d-1} \textrm{d}t \\ =&\, \int _0 ^{T_1} \left( - \int _{\partial ^*U_t} \vec h\cdot \vec \nu \, d {\mathscr {H}}^{d-1} + \lambda {\mathscr {H}}^{d-1} (\partial ^*U_t) \right) \eta \, \textrm{d}t, \end{aligned} \end{aligned}$$

where we used (b3) and (b5). Hence we obtain (b6). \(\square \)