1 Introduction

Let \({\mathcal{M}}\) and \({\mathcal{N}}\) be smooth compact Riemannian manifolds of dimension m and n with metric g and h, respectively. We assume that, by Nash’s embedding theorem, \({\mathcal{N}}\) is isometrically embedded into \(\mathrm{I}\!\mathrm{R}^l\) \((l > n)\). For a smooth map u from \({\mathcal{M}}\) to \({\mathcal{N}} \subset \mathrm{I}\!\mathrm{R}^l\), we consider the p-energy

$$\begin{aligned} \displaystyle { E (u) : = \int _{\mathcal{M}} \frac{1}{p} |D u|^p \, d {\mathcal{M}}, \quad \quad p \ge 2. } \end{aligned}$$
(1.1)

Here the unknown map \(u = \left( u^i \right) \), \(i = 1, \ldots , l\), is a vector-valued function, defined on \(\mathcal{M}\) with values into \({\mathcal{N}} \subset \mathrm{I}\!\mathrm{R}^l\). In a local coordinate \(x = (x_\alpha )\), \(\alpha = 1, \ldots , m\), on \({\mathcal{M}}\), the usual notation is used : \(g = \left( g_{\alpha \beta }\right) \), \(\left( g_{\alpha \beta }\right) ^{- 1} = \left( g^{\alpha \beta }\right) \), \(|g| = |\det \left( g_{\alpha \beta }\right) |\), and \(d {\mathcal{M}} = \sqrt{|g|} d x\) is a volume element with m-dimensional Lebesgue measure dx, and \(D_\alpha = \partial /\partial x_\alpha \), \(\alpha = 1, \ldots , m\), \(D u = \left( D_\alpha u^i\right) \) is the gradient of a map u, and \(|D u|^2 = \sum _{\alpha , \beta = 1}^m g^{\alpha \beta } D_\alpha u \cdot D_\beta u\) with an Euclidean inner product \(\cdot \) in \(\mathrm{I}\!\mathrm{R}^l\).

The p-harmonic map is a critical point of the p-energy and satisfies the Euler–Lagrange equation

$$\begin{aligned} \displaystyle { \left\{ \begin{array}{ll} \displaystyle { - \Delta _p u = |D u|^{p - 2} A (u)(D u, D u) } \\ \displaystyle { u \in {\mathcal{N}} } \end{array} \right. } \end{aligned}$$
(1.2)

where the p-Laplace operator is denoted by

$$\begin{aligned} \displaystyle { \Delta _p u = \frac{1}{\sqrt{|g|}} \sum _{\alpha , \beta = 1}^m D_\alpha \left( |D u|^{p - 2} \sqrt{|g|} g^{\alpha \beta } D_\beta u \right) } \end{aligned}$$
(1.3)

and the second fundamental form A(u)(DuDu) of \({\mathcal{N}} \subset \mathrm{I}\!\mathrm{R}^l\) is a vector field along the map \(u \in {\mathcal{N}}\) with values into the orthogonal complement of the tangent space of \({\mathcal{N}}\) at u (if necessary, the manifold \({\mathcal{N}}\) is assumed to be orientable).

An approach to look for p-harmonic maps is to exploit the gradient flow associated with the p-energy, called the p-harmonic flow, which are described by the evolutionary p-Laplacian system

$$\begin{aligned} \displaystyle { \left\{ \begin{array}{ll} \displaystyle { \partial _t u - \Delta _p u = |D u|^{p - 2} A (u)(D u, D u) } \\ \displaystyle { u \in {\mathcal{N}} } \end{array} \right. } \end{aligned}$$
(1.4)

where \(u = u (t, \, x)\) is defined on \({\mathcal{M}}_\infty = (0, \, \infty ) \times {\mathcal{M}}\) with values onto \(\mathrm{I}\!\mathrm{R}^l\), \(\partial _t u = \left( \partial _t u^i\right) \) is a partial derivative on time. In this paper we study a global existence and regularity of a solution to the Cauchy problem for the p-harmonic flow (1.4).

Let \(\mathrm{I}\!\mathrm{R}^l = \mathcal{T}_u {\mathcal{N}} \oplus (\mathcal{T}_u {\mathcal{N}})^\bot \) be the orthogonal decomposition of \(\mathrm{I}\!\mathrm{R}^l\) with respect to the tangent space \(\mathcal{T}_u {\mathcal{N}}\) at each \(u \in {\mathcal{N}}\). The corresponding orthonormal basis is \(\left( e_1 (u), \ldots , e_n (u)\right) \) of the tangent space \(\mathcal{T}_u {\mathcal{N}}\) and \(\left( e_{n + 1} (u), \ldots , e_l (u)\right) \) of its orthogonal complement \((\mathcal{T}_u {\mathcal{N}})^\bot \). Then we find an equivalent representation for the p-harmonic flow

$$\begin{aligned} \displaystyle { \partial _t u - \Delta _p u \, \bot \, \mathcal{T}_u {\mathcal{N}} \, \Longleftrightarrow \, \partial _t u - \Delta _p u = |D u|^{p - 2} A (u) (D u, D u). } \end{aligned}$$
(1.5)

In fact, there exists some vector-valued function \(\lambda = \left( \lambda ^j (u)\right) \), \(j = n + 1, \ldots , l\), such that

$$\begin{aligned} \displaystyle { \partial _t u - \Delta _p u \, \bot \, \mathcal{T}_u {\mathcal{N}} \, \Longleftrightarrow \, \partial _t u - \Delta _p u = \sum _{j = n + 1}^l \lambda ^j (u) e^j (u) } \end{aligned}$$

and, simply multiplying each of the orthonormal basis \(e_j (u)\), \(j = n + 1, \ldots , l\), by the second equation above, we have

$$\begin{aligned} \displaystyle { \lambda ^j (u) = |D u|^{p - 2} \sum _{\alpha , \beta = 1}^m \sqrt{|g|} g^{\alpha \beta } D_\beta u \cdot \big ( D_\alpha u \cdot D_u e_j (u) \big ), } \end{aligned}$$

where \(\partial _t u, \, D u \in \mathcal{T}_u {\mathcal{N}}\) because the map \(u = u (t, x)\) moves on \({\mathcal{N}}\), and thus, the usual Euclidean innner product in \(\mathrm{I}\!\mathrm{R}^l\) is taken, so that \(\partial _t u \cdot e_j (u) = 0\) and \(D u \cdot e_j (u) = 0\), for all \(j = n + 1, \ldots , l\). Here the last summation term in the equation above is nothing but the second fundamental form of \({\mathcal{N}}\) along the map u. Furthermore, the Euclidean inner product in \(\mathrm{I}\!\mathrm{R}^l\) of \(\partial _t u\) with the p-harmonic flow Eq. (1.4) leads the energy identity

$$\begin{aligned} \displaystyle { |\partial _t u|^2 - \frac{1}{\sqrt{|g|}} D_\alpha \left( |D u|^{p - 2} \sqrt{|g|} g^{\alpha \beta } D_\beta u\cdot \partial _t u \right) + \partial _t \frac{1}{p} |D u|^p = 0, } \end{aligned}$$

integrated in \({\mathcal{M}}\) yielding, through integration by parts,

$$\begin{aligned} \displaystyle { \frac{d}{d t} E (u (t)) = - \Vert \partial _t u (t)\Vert _2^2. } \end{aligned}$$
(1.6)

Thus, the p-energy E(u(t)) is decreasing along the solution u(t) of the p-harmonic flow and, in fact, the solution \(\left\{ u (t)\right\} \subset C^\infty ({\mathcal{M}}, {\mathcal{N}})\), \(0< t < \infty \), is the trajectory of negative direction gradient vector field of the p-energy

$$\begin{aligned} \displaystyle { \frac{d u}{d t} (t) }= & {} \displaystyle { - \nabla E (u (t)) } \\= & {} \displaystyle { \Delta _p u (t) + |D u (t)|^{p - 2} A (u(t)) (D u (t), D u(t)), } \end{aligned}$$

by the Euler–Lagrange equation (1.2), where \(\nabla E (u (t))\) is the G\(\hat{a}\)teaux derivative of \(E (\cdot )\) at \(u (t) \in C^\infty ({\mathcal{M}}, {\mathcal{N}})\). Therefore, a global in time solution to (1.4) for any initial data may converge to critical points of the p-energy, the p-harmonic maps, as time tends to \(\infty \). This heat flow method was originally realized by J. Eells and J. H. Sampson for the harmonic flow in the case \(p = 2\) under the condition that the sectional curvature of target manifold \({\mathcal{N}}\) is non-positive, in their pioneering work [15, 23]. This fundamental result in the harmonic flow case \(p = 2\) was also extends to hold similarly for the p-harmonic flow.

Theorem 1

 [16, 31] Suppose that the sectional curvature of the target manifold \({\mathcal{N}}\) is non-positive, \({ Sect} ({\mathcal{N}}) \le 0\). Then, for any smooth initial map from \({\mathcal{M}}\) to \({\mathcal{N}}\), there exists a unique global in time weak solution of the Cauchy problem on \({\mathcal{M}}\) for p-harmonic flow (1.4). The solution u and its gradient are Hölder continuous in time-space. The solution and its gradient uniformly converge to a weak solution and its gradient, respectively, of the p-harmonic map, as time tends to \(\infty \), respectively, which are Hölder continuous.

We call the weak solution which is locally continuous on time-space together with its gradient the regular solution. The curvature restriction on the target manifold in general is necessary for the global existence of regular solution of the p-harmonic flow. In fact, without any curvature restriction on the target manifold, we have some example of a blowing up solution at a finite time (see [5] in the case \(p = m = 3\)). But, a global in time weak solution may be exist.

Theorem 2

 [24] Let \(p = m \ge 2\) and the initial data be in the set of Sobolev maps \(W^{1, p} ({\mathcal{M}}, {\mathcal{N}})\) between two smooth, compact Riemannian manifolds \({\mathcal{M}}\) and \({\mathcal{N}}\). Then, there exists a global in time weak solution of Cauchy problem on \({\mathcal{M}}\) for the m-harmonic flow. The solution and its gradient are Hölder continuous on time-space, except for at most finitely many time slices.

In the case \(p = m = 2\), the global in time existence as above is also shown for the initial-boundary value problem of the two-dimensional harmonic flow. Moreover, the solution is smooth except for at most finitely many points [3, 38]. In the case \(p = m\), a nice Sobolev type inequality on time-space, referred as Ladyzhenskaya or Nash inequality, can be available and is crucial for regularity estimate in this case.

In the higher dimensional case \(m \ge 3\), M. Struwe et al. established the following fundamental result for global existence and regularity of the harmonic flow in the case \(p = 2\) in [8, 9, 39]

Theorem 3

 [8, 9, 39] Let \(p = 2\). Let initial and boundary data \(u_0\) be smooth map from \({\mathcal{M}}\) into \({\mathcal{N}}\). Then, there exists a global in time weak solution u of the harmonic flow (1.4). The solution u satisfies the energy inequality: letting \({\mathcal{M}}_\infty = (0, \infty ) \times {\mathcal{M}}\),

$$\begin{aligned} \Vert \partial _t u\Vert _{{\mathrm {L}}^2 ({\mathcal{M}}_\infty )}^2 + \sup _{0< t < \infty } E (u (t)) \le E (u_0). \end{aligned}$$
(1.7)

There exists a relatively closed subset \(\Sigma \subset (0, \, \infty ] \times {\mathcal{M}}\) such that the solution u is smooth in the complement of \(\Sigma \), \({\mathcal{M}}_\infty {\setminus } \Sigma \); \(\Sigma \) is of at most finite m-dimensional Hausdorff measure with respect to the usual parabolic metric in \({\mathcal{M}}_\infty \), and furthermore, for any time \(t_0 > 0\) and some positive \(C_0= C ({\mathcal{M}}, {\mathcal{N}}, t_0, E (u_0))\), \(\mathcal{H}^{m - 2} (\{t_0\} \times \Sigma ) \le C_0\); As time suitably tends to \(\infty \), the solution converges to a weakly harmonic map \(u_\infty \) weakly in Sobolev space \(W^{1, \, 2} ({\mathcal{M}}, \, \mathrm{I}\!\mathrm{R}^l)\). There exists a closed set \(\Sigma _\infty \subset {\mathcal{M}}\) such that \(u_\infty \) is smooth on \({\mathcal{M}} {\setminus } \Sigma _\infty \); \(\Sigma _\infty \) is of at most finite \((m - 2)\)-dimensional Hausdorff measure: For some positive \(C_0^\prime = C^\prime ({\mathcal{M}}, {\mathcal{N}}, t_0, E (u_0))\), \(\mathcal{H}^{m - 2} (\Sigma _\infty ) \le C_0^\prime \).

There also exist blowing up solutions at a finite time (see [4, 7, 11, 22]).

If the target manifold is the standard unit sphere, the global in time existence of weak solution to the p-harmonic flow is also shown by use of the special structure of the target standard unit sphere [6, 25, 27, 32].

In differential geometry, the regularity has been studied under a smallness of image of a solution, instead of curvature condition [18], and the everywhere regularity of a small solution of harmonic flow is shown in [19, 20, 37]. Such regularity of a small solution of p-harmonic flow remained open (refer to [28]).

Theorem 3 implies the global in time existence of weak solution of the harmonic flow in the case \(p = 2\), which is partial regular in the sense of regularity outside exceptional closed set. It has remained open whether or not the corresponding result holds for the p-harmonic flow, since the important result, Theorem 2, was obtained for the case \(p = m\).

A compactness for regular p-harmonic flows with uniform boundedness of p-energy has been recently proved by the author in [33, 34] (see [39, Theorem 6.1; its proof, pp. 494–497] for the harmonic flow). The compactness result will be the key ingredient for the global in time existence of p-harmonic flow (refer to [9] for the harmonic flow case).

Theorem 4

(A compactness of regular p-harmonic flows with uniformly bounded p-energy) Let \(p > 2\). Suppose that a family \(\{u_k\}\) of regular p-harmonic flows on \(\mathrm{I}\!\mathrm{R}^m_\infty = (0, \, \infty ) \times \mathrm{I}\!\mathrm{R}^m\) satisfies the p-energy boundedness with uniform positive constant C

$$\begin{aligned} p \, \Vert \partial _t u_k \Vert _{{\mathrm {L}}^2 (\mathrm{I}\!\mathrm{R}^m_\infty )}^2 + \sup _{0< t < \infty } \Vert D u_k (t)\Vert _{{\mathrm {L}}^p (\mathrm{I}\!\mathrm{R}^m)}^p \le C \end{aligned}$$
(1.8)

and converges to a limit map u in the sense

$$\begin{aligned}&u_k \longrightarrow u \quad \text{ weakly } * \text{ in } {\mathrm {L}}^\infty \left( 0, \, T; \, {\mathrm {W}}^{1, p} (\mathrm{I}\!\mathrm{R}^m_\infty , \, \mathrm{I}\!\mathrm{R}^l)\right) , \end{aligned}$$
(1.9)
$$\begin{aligned}&D u_k \longrightarrow D u \quad \text{ weakly } \text{ in } {\mathrm {L}}^p \left( \mathrm{I}\!\mathrm{R}^m_\infty , \, \mathrm{I}\!\mathrm{R}^{m l}\right) , \end{aligned}$$
(1.10)
$$\begin{aligned}&\partial _t u_k \longrightarrow \partial _t u \quad \text{ weakly } \text{ in } {\mathrm {L}}^2 \left( \mathrm{I}\!\mathrm{R}^m_\infty , \, \mathrm{I}\!\mathrm{R}^l\right) . \end{aligned}$$
(1.11)

Then, the limit map u is a global weak solution on \(\mathrm{I}\!\mathrm{R}^m_\infty \) of the p-harmonic flow such that \(u \in {\mathcal{N}}\) almost everywhere in \(\mathrm{I}\!\mathrm{R}^m_\infty \), and the p-energy boundedness is valid, replacing \(u_k\) by u in (1.8). Moreover, the limit map u is partial regular in the sense : There exists a relatively closed set \(\Sigma \) in \(\mathrm{I}\!\mathrm{R}^m_\infty \) such that u and its gradient Du are locally in time-space continuous in the complement \(\mathrm{I}\!\mathrm{R}^m_\infty {\setminus } \Sigma \), and the size of \(\Sigma \) is also estimated by the Hausdorff measure : For any positive number \(\gamma _0\), \(2< \gamma _0 < p\), the set \(\Sigma \) is of at most locally zero m-dimensional Hausdorff measure with respect to the time-space metric \(|t|^{1/\gamma _0} + |x|\), and, furthermore, for any positive time \(\tau < \infty \), the \((m - \gamma _0)\)-dimensional Hausdorff measure of \(\{\tau \} \times \Sigma \) with respect to the usual Euclidean metric is locally zero.

In this paper we show the global existence and regularity of a weak solution of the Cauchy problem for the p-harmonic flow (1.4) with an initial data \(u_0\)

$$\begin{aligned} \left\{ \begin{array}{ll} \displaystyle { \partial _t u - {\text{ div }} \left( |D u|^{p - 2} D u \right) = |D u|^{p - 2} A (u)(D u, D u) } &{} \text{ in } {\mathcal{M}}_\infty \\ u (0) = u_0 &{} \text{ on } {\mathcal{M}} \end{array} \right. \end{aligned}$$
(1.12)

and a convergence of the solution of p-harmonic flow to a p-harmonic map along a time sequence tending to infinity. The Sobolev space on \({\mathcal{M}}\) is usually defined as

Definition 1

$$\begin{aligned}&\displaystyle { {\mathrm {W}}^{1, \, p} \left( {\mathcal{M}}, \, \mathrm{I}\!\mathrm{R}^l\right) : = \left\{ v \in {\mathrm {L}}^p \left( {\mathcal{M}}, \, \mathrm{I}\!\mathrm{R}^l\right) \left| \, \exists \text{ a } \text{ weak } \text{ derivative } \, D v \in {\mathrm {L}}^p \left( {\mathcal{M}}, \, \mathrm{I}\!\mathrm{R}^{m l}\right) \right. \right\} ; }\nonumber \\&\displaystyle { {\mathrm {W}}^{1, \, p} \left( {\mathcal{M}}, \, {\mathcal{N}}\right) : = \left\{ v \in {\mathrm {W}}^{1, \, p} \left( {\mathcal{M}}, \, \mathrm{I}\!\mathrm{R}^l\right) \left| \, v \in {\mathcal{N}} \, \text{ almost } \text{ everywhere } \text{ in } {\mathcal{M}} \right. \right\} ; } \nonumber \\&\displaystyle { \Vert v\Vert _{{\mathrm {W}}^{1, \, p} ({\mathcal{M}})} : = \Vert v\Vert _{{\mathrm {L}}^p ({\mathcal{M}})}^p + \Vert D v\Vert _{{\mathrm {L}}^p ({\mathcal{M}})} } \end{aligned}$$
(1.13)

Definition 2

Let \(u_0 \in {\mathrm {W}}^{1, p} ({\mathcal{M}}, \, {\mathcal{N}})\). A map u is called a global weak solution of the Cauchy problem (1.12) if and only if u is a measurable vector-valued function defined on \({\mathcal{M}}_\infty := (0, \, \infty ) \times {\mathcal{M}}\) with values into \(\mathrm{I}\!\mathrm{R}^l\), satisfying the following four conditions :

  1. (D1)

    \(u \in {\mathrm {L}}^\infty (0, \infty ; \, {\mathrm {W}}^{1, p} ({\mathcal{M}}, \, \mathrm{I}\!\mathrm{R}^l))\), \(\partial _t u \in {\mathrm {L}}^2 ({\mathcal{M}}_\infty , \mathrm{I}\!\mathrm{R}^l);\)

  2. (D2)

    \(u \in {\mathcal{N}}\) almost everywhere in \({\mathcal{M}}_\infty ;\)

  3. (D3)

    u satisfies (1.4) in the sense of distributions, that is, for any smooth map \(\phi \) \(\in \) \(\mathrm {C}^\infty _0 ({\mathcal{M}}_\infty , \, \mathrm{I}\!\mathrm{R}^l)\),

    $$\begin{aligned} \displaystyle { \int _{{\mathcal{M}}_\infty } \{ \partial _t u \, \cdot \, \phi + |D u|^{p - 2} D u \cdot D \phi - |D u|^{p- 2} \phi \cdot A (u) (D u, D u) \} \, d z = 0; } \end{aligned}$$
  4. (D4)

    u attains the initial data continuously in the Sobolev space

    $$\begin{aligned} \left| u (t) - u_0 \right| _{{\mathrm {W}}^{1, p} ({\mathcal{M}}, \, \mathrm{I}\!\mathrm{R}^l)} \rightarrow 0 \quad \text{ as } t \rightarrow 0. \end{aligned}$$

Theorem 5

(A global existence and regularity for the p-harmonic flow) Let \(p > 2\). Let \(u_0 \in {\mathrm {W}}^{1, p} ({\mathcal{M}}, \, {\mathcal{N}})\). Then, there exists a global weak solution u of (1.12), satisfying the energy inequality

$$\begin{aligned} \displaystyle { \, \Vert \partial _t u \Vert _{{\mathrm {L}}^2 ({\mathcal{M}}_\infty )}^2 + \sup _{0< t < \infty } E (u (t)) \le E (u_0). } \end{aligned}$$
(1.14)

Moreover, the solution u is partial regular in the following sense : There exists a relatively closed set \(\Sigma \) in \({\mathcal{M}}_\infty = (0, \infty ) \times {\mathcal{M}}\) such that u and its gradient Du are locally in time-space continuous in the complement \({\mathcal{M}}_\infty {\setminus } \Sigma \), and the size of \(\Sigma \) is also estimated by the Hausdorff measure : For any positive number \(\gamma _0\), \(2< \gamma _0 < p\), the set \(\Sigma \) is of at most zero m-dimensional Hausdorff measure with respect to the time-space metric \(|t|^{1/\gamma _0} + |x|\), and, furthermore, for any positive time \(\tau < \infty \), the \((m - \gamma _0)\)-dimensional Hausdorff measure of \(\{\tau \} \times \Sigma \) with respect to the usual Euclidean metric is zero. As time suitably tends to \(\infty \), the solution converges to a weakly p-harmonic map \(u_\infty \) weakly in \(W^{1, \, p} ({\mathcal{M}}, \, \mathrm{I}\!\mathrm{R}^l)\). There exists a closed set \(\Sigma _\infty \subset {\mathcal{M}}\) such that \(u_\infty \) and its gradient \(D u_\infty \) are locally continuous on \({\mathcal{M}} {\setminus } \Sigma _\infty \); For any positive number \(\gamma _0\), \(2< \gamma _0 < p\), \(\Sigma _\infty \) is of at most zero \((m - \gamma _0)\)-dimensional Hausdorff measure.

Remark

Measuring by use of the time-space metric \(|t|^{1/p} + |x|\) on \({\mathcal{M}}_\infty \), the \((m + p - \gamma _0)\)-dimensional Hausdorff size of \(\Sigma \) is zero. The scale order in the estimate of singular set \(\Sigma \) is almost optimal, since the exponent \(\gamma _0\) can be as close to p as possible.

The contents of the paper are as follows :

  1. 1.

    Introduction

  2. 2.

    Penalty approximation

  3. 3.

    Small energy regularity estimate

  4. 3.1

    Preliminaries;    3.2 Local energy regularity estimate

  5. 4.

    Passing to the limit

  6. 5.

    Monotonicity estimate of a local scaled energy

  7. 6.

    Appendix

In Sect. 2, we introduce the so-called penalty approximation for the p-harmonic flow. In Sect. 3, some preliminary estimates for the penalty approximating solutions are derived, those proofs are given in “Appendix”, and then, the small energy regularity estimate is shown to hold uniformly for the penalty approximating solutions, and is applied for their convergence to a weak solution of the p-harmonic flow in Sect. 4, based on the compactness result, Theorem 4. The monotonicity estimate, Lemmata 12 and 13, is demonstrated in Sect. 5.

2 Penalty approximation

In this section we set the approximation scheme for the p-harmonic flow. We will approximate the p-harmonic flow by the solutions of the gradient flow for the so-called penalized functional, introduced in [9] for the harmonic flow case \(p = 2\) (also refer to [29, 40]).

Since the manifold \({\mathcal{N}}\) is smooth and compact, there exists a tubular neighborhood \(\mathcal{O}_{2 \delta _{\mathcal{N}}}\) with width \(2 \delta _{\mathcal{N}}\) of \({\mathcal{N}}\) in \(\mathrm{I}\!\mathrm{R}^l\) such that any point \(u \in \mathcal{O}_{2 \delta _{\mathcal{N}}}\) has a unique nearest point \(\pi _{\mathcal{N}} (u) \in {\mathcal{N}}\) satisfying \( \text{ dist } \left( u, \, \pi _{\mathcal{N}} (u)\right) = \text{ dist } \left( u, \, {\mathcal{N}}\right) \) for the Euclidean distance \( \text{ dist } \left( \cdot , \cdot \right) \), where the projection \(\pi _{\mathcal{N}} \, : \, \mathcal{O}_{2 \delta _{\mathcal{N}}} \rightarrow {\mathcal{N}}\) is smooth, since the manifold \(\mathcal{N}\) is smooth. The distance function \( \text{ dist } (u, \, {\mathcal{N}})\) is Lipschitz continuous on \(u \in \mathcal{O}_{2 \delta _{\mathcal{N}}}\).

Let \(\chi \) be a smooth, non-decreasing real-valued function defined on \([0, \, \infty )\) such that \(\chi (s) = s\) for \(s \le (\delta _{\mathcal{N}})^2\) and \(\chi (s) = 2 (\delta _{\mathcal{N}})^2\) for \(s \ge 4 (\delta _{\mathcal{N}})^2\). Then, the function \(\chi \left( \text{ dist }^2 (u, \, {\mathcal{N}})\right) \) is smooth on \(u \in \mathrm{I}\!\mathrm{R}^l\). Its gradient at \(u \in \mathcal{O}_{2 \delta _{\mathcal{N}}}\) is computed as

$$\begin{aligned}&D_u \chi \left( \text{ dist }^2 (u, \, {\mathcal{N}})\right) = 2 \chi ^\prime \left( \text{ dist }^2 (u, \, {\mathcal{N}})\right) \text{ dist } (u, \, {\mathcal{N}}) D_u \text{ dist } (u, \, {\mathcal{N}}); \\&D_u \text{ dist } (u, \, {\mathcal{N}}) = \frac{u - \pi _{\mathcal{N}} (u)}{|u - \pi _{\mathcal{N}} (u)|} \end{aligned}$$

parallel to the vector field \(u - \pi _{\mathcal{N}} (u)\) and orthogonal to \(\mathcal{T}_{\pi _{\mathcal{N}} (u)} {\mathcal{N}}\). We also have that, for any \(u \in {\mathcal{N}}\) and any tangent vector \(\tau \in \mathcal{T}_u {\mathcal{N}}\),

$$\begin{aligned} \displaystyle { \left| \tau ^i \tau ^j D_{u^i} D_{u^j} \text{ dist } (u, \, {\mathcal{N}}) \right| \le C({\mathcal{N}}) |\tau |^2 } \end{aligned}$$

(see [2, Theorem 3.1, pp. 704–705], [1, Theorem 2.1]).

For positive parameters \(1 \le K \nearrow \infty \) and \(1 > \epsilon \searrow 0\), we consider the Cauchy problem in \({\mathcal{M}}_\infty \) with initial data \(u_0\) for the gradient flow, called the penalized equation,

$$\begin{aligned} \displaystyle { \left\{ \begin{array}{ll} \partial _t u - \Delta _{p, \, \epsilon } u + C_0 \, K \, \chi ^\prime \left( \text{ dist }^2 (u, \, {\mathcal{N}})\right) \text{ dist } (u, \, {\mathcal{N}}) D_u \text{ dist } (u, \, {\mathcal{N}}) = 0 \\ u (0) = u_0 \end{array} \right. } \end{aligned}$$
(2.1)

associated with the penalized functional, defined by

$$\begin{aligned} \displaystyle { F_{K, \, \epsilon } (u) : = E_\epsilon (u) + C_0 \, \frac{K}{2} \int _{\mathcal{M}} \chi \left( \text{ dist }^2 (u, \, {\mathcal{N}})\right) \, d {\mathcal{M}}, } \end{aligned}$$
(2.2)

where the positive constant \(C_0\) will be stipulated later, depending only on p, \({\mathcal{M}}\) and \({\mathcal{N}}\) (see Lemma 9 and its proof in “Appendix B”). The partial differential operator \(\Delta _{p, \, \epsilon }\) and its corresponding energy, called the regularized p-Laplace operator and the regularized p-energy, respectively, are defined as

$$\begin{aligned}&\Delta _{p, \, \epsilon } u : = \frac{1}{\sqrt{|g|}} \sum _{\alpha , \beta = 1}^m D_\alpha \left( \big (\epsilon + |D u|^2\big )^{\frac{p - 2}{2}} \sqrt{|g|} g^{\alpha \beta } D_\beta u \right) ; \nonumber \\&\quad E_\epsilon (u) : = \int _{\mathcal{M}} \frac{1}{p} \left( \epsilon + |D u|^2\right) ^{\frac{p}{2}} \, d {\mathcal{M}} \end{aligned}$$
(2.3)

We now state the global existence for (2.1). For the proof see “Appendix A”.

Lemma 6

(Existence for the penalty approximation) Let \(p > 2\) and let \(u_0 \in {\mathrm {W}}^{1, p} \left( {\mathcal{M}}, \, {\mathcal{N}}\right) \). For each positive numbers K and \(\epsilon \), there exists a weak solution \(u = u_{K, \, \epsilon }\) of the Cauchy problem for the penalized equation (2.1) such that \(u = u_{K, \, \epsilon }\) satisfies the energy inequality

$$\begin{aligned} \displaystyle { \Vert \partial _t u \Vert _{ {\mathrm {L}}^2 ({\mathcal{M}}_\infty ) }^2 + \sup _{0< t < \infty } F_{K, \, \epsilon } (u) \le E_\epsilon (u_0) } \end{aligned}$$
(2.4)

and, that u, Du, \(\partial _t u\) and \(D^2 u\) are locally (Hölder) continuous on time and space (with some Hölder exponent) in \({\mathcal{M}}_\infty \) and u satisfies the penalized equation everywhere in \({\mathcal{M}}_\infty \).

3 Small energy regularity estimate

3.1 Preliminaries

In this section we show some regularity estimates for solutions \(u = u_{K, \, \epsilon }\) of the penalized equations (2.1). Those proofs are given in “Appendix”.

Lemma 7

(Energy inequality) Let \(u_0 \in {\mathrm {W}}^{1, \, p} \big ({\mathcal{M}}, \, {\mathcal{N}}\big )\) and \(u = u_{K, \, \epsilon }\) be a regular solution of (2.1). Then, it holds that

$$\begin{aligned} \displaystyle { \Vert \partial _t u \Vert _{ {\mathrm {L}}^2 ({\mathcal{M}}_\infty ) }^2 + \sup _{0< t < \infty } F_{K, \, \epsilon } (u) \le E_\epsilon (u_0). } \end{aligned}$$
(3.1)

A solution of the penalized equation is uniformly bounded, that is used in the regularity estimate.

Lemma 8

(Boundedness) Let \(u = u_{K, \, \epsilon }\) be a regular solution of (2.1). Then it holds that \(\sup _{ {\mathcal{M}}_\infty } |u| \le H\), where the positive number H is so large that \(B (H) \supset \mathcal{O}_{ 2 \delta _{\mathcal{N}} } ({\mathcal{N}})\) in \(\mathrm{I}\!\mathrm{R}^l\).

We will put the setting for local estimates for the penalized Eq. (2.1). For this purpose we recall some standard geometrical settings. Let \(R_{\mathcal{M}} > 0\) be a lower bound for the injective radius of the exponential map on \({\mathcal{M}}\). Thus, for any positive number \(R < R_{\mathcal{M}}\) and any point \(x_0 \in {\mathcal{M}}\), the geodesic ball \(\mathcal{B} (R, \, x_0) \subset {\mathcal{M}}\) of radius R around \(x_0\) is well-defined and diffeomorphic to the Euclidean ball \(B (R, \, 0) \subset \mathrm{I}\!\mathrm{R}^m\), under the linear homeomorphism \(\mathcal{T}_{x_0} {\mathcal{M}} \cong \mathrm{I}\!\mathrm{R}^m\), through the exponential map

$$\begin{aligned} \displaystyle { \exp _{x_0} \cdot \, : \, \mathrm{I}\!\mathrm{R}^m \supset B (R, \, 0) \ni x \rightarrow \exp _{x_0} x \in \mathcal{B} (R, \, x_0) \subset {\mathcal{M}}. } \end{aligned}$$

For any \(t \in (0, \, \infty )\), the map

$$\begin{aligned} \displaystyle { u \left( t, \, \exp _{x_0} \cdot \right) \, : \, \mathrm{I}\!\mathrm{R}^m \supset B (R, \, 0) \ni x \rightarrow u \left( t, \, \exp _{x_0} x\right) \in \mathrm{I}\!\mathrm{R}^l } \end{aligned}$$
(3.2)

is well-defined. Hereafter let \(x_0 \in {\mathcal{M}}\) be arbitrarily taken and fixed. We abbreviate as \(B (R_{\mathcal{M}}) = B (R_{\mathcal{M}}, \, 0)\). We denote \(u \left( t, \, \exp _{x_0} x\right) \) by \(u (t, \, x)\) for any \((t, \, x) \in \left( B (R_{\mathcal{M}})\right) _\infty : = (0, \, \infty ) \times B (R_{\mathcal{M}})\) and, furthermore, by translation, regard u as a map defined on \(\left( B (R_{\mathcal{M}})\right) _\infty \) with values into \(\mathrm{I}\!\mathrm{R}^l\).

Let us denote the penalized energy density for a map u by

$$\begin{aligned} \displaystyle { e_{K, \, \epsilon } (u) : = \frac{1}{p} \big (\epsilon + |D u|^2\big )^{\frac{p}{2}} + \frac{K}{2} \chi \big ( { \text{ dist }}^2 (u, \, {\mathcal{N}}) \big ) }. \end{aligned}$$
(3.3)

We need the so-called Bochner type estimate for the penalized energy density. See “Appendix C” for the proof. Here the constant \(C_0\) in (2.1) is appropriately chosen.

Lemma 9

(Bochner type estimate) Let \(p > 2\) and \(u = u_{K, \, \epsilon }\) be a regular solution to (2.1). For brevity, put \(e (u) = e_{K, \, \epsilon } (u)\). Then, it holds in \((B_{R_{\mathcal{M}}})_\infty \) that

$$\begin{aligned}&\displaystyle { \partial _t e (u) - \frac{1}{\sqrt{|g|}} D_\alpha \left( \big ( \epsilon + |D u|^2 \big )^{\frac{p - 2}{2}} \sqrt{|g|} \mathcal{A}^{\alpha \beta } D_\beta e (u) \right) } \nonumber \\&\qquad \displaystyle { + \, C_1 \, \big ( \epsilon + |D u|^2 \big )^{\frac{p - 2}{2}} \big |D^2 u\big |^2 + C_2 \, \Big | 2^{- 1} \, K \, D_u \chi \Big ( \text{ dist }^2 \big (u, \, {\mathcal{N}}\big )\Big ) \Big |^2 } \nonumber \\&\quad \displaystyle { \le C_3 \, \left( 1 + e (u)^{\frac{2}{p}}\right) \, e (u)^{2 \left( 1 - \frac{1}{p}\right) },} \end{aligned}$$
(3.4)

where

$$\begin{aligned} \displaystyle { \mathcal{A}^{\alpha \beta } : = g^{\alpha \beta } + (p - 2) \frac{ g^{\alpha \gamma } g^{\beta \mu } D_\gamma u \cdot D_\mu u}{\epsilon + |D u|^2}, \quad \left| D^2 u\right| ^2 = g^{\alpha \beta } g^{\gamma \mu } D_\alpha D_\gamma u \cdot D_\beta D_\mu u, } \end{aligned}$$

the summation convention over repeated indices is used and the positive constants \(C_i\) \((i = 1, 2, 3)\) depend on p, \(\mathcal{M}\) and \(\mathcal{N}\).

Let \(\lambda _0\) be a positive number, R be a positive number such that \(R < \min \{1, \, R_{\mathcal{M}}/2, \, T^{1/\lambda _0}\}\) and \((t_0, \, x_0)\) in the parabolic like envelope \(\mathcal{P} : = \big \{ (t, \, x) \, : \, T - R^{\lambda _0}< t \le T, \, |x|^{\lambda _0} < t - (T - R^{\lambda _0}) \big \}.\) In the following we use time-space local cylinder. For \(r, \, \tau > 0\), \(Q \left( \tau , \, r\right) (t_0, \, x_0) = \left( t_0 - \tau , \, t_0\right) \times B (r, \, x_0)\), where \(B (r, \, x_0)\) is an open ball in \(B_{R_{\mathcal{M}}}\) with center \(x_0\) and radius r. For brevity, we put \(u = u_{K, \, \epsilon }\), \(e (u) = e_{K, \, \epsilon } (u)\) in (3.3) and abbreviate the time-space Lebesgue measure \(d t \, d {\mathcal{M}}\) as dz.

Lemma 10

(Gradient boundedness on a small region) For some \((t_0, \, x_0) \in \mathcal{P}\), let \(\rho _0 : = \left( (t_0 - (T - R^{\lambda _0}))^{1/\lambda _0} - |x_0| \right) /4\). Suppose that, for \(\lambda _0 > 0\), \(r_0 > 0\), \(C_1 > 0\) and \(L > 0\),

$$\begin{aligned} \displaystyle { r_0 \le \frac{\rho _0}{2}; \quad L^{2 - p} \, (r_0)^2 \le (\rho _0)^{\lambda _0}; \quad r_0 \, \sup _{ Q \left( L^{2 - p} (r_0)^2, \, r_0 \right) (t_0, \, x_0) } (e (u))^{\frac{1}{p}} \le C_1. } \end{aligned}$$
(3.5)

Let \(q > 2\) be a positive number. Then there exists a positive number C depending only on q, p, \({\mathcal{M}}\) and \({\mathcal{N}}\), but, independent of L, such that

$$\begin{aligned}&\sup _{ Q \left( L^{2 - p} (r_0/2)^2, \, r_0/2 \right) (t_0, \, x_0) } e (u) \, \mathcal{C}^q \nonumber \\&\quad \le \frac{C \, L^{2 - p}}{\left| Q \left( L^{2 - p} (r_0)^2, \, r_0 \right) \right| } \int \limits _{ Q \left( L^{2 - p} (r_0)^2, \, r_0 \right) (t_0, \, x_0) } (e (u))^{2 - \frac{2}{p}} \, \mathcal{C}^q \, d z + C \, L^p; \nonumber \\&\qquad \mathcal{C} (t, \, x) : = \left( ( t - ( T - R^{\lambda _0} ) )^{ \frac{1}{\lambda _0} } - |x| \right) _+. \end{aligned}$$
(3.6)

The detail of proof is presented in “Appendix D” (refer to [10, 12]).

3.2 Local regularity estimates

The partial regularity is based on the so-called small energy regularity estimate (refer to [39, Theorems 5.1, 5.3, 5.4; their proofs, pp. 491–494]). The small energy regularity estimate for the p-harmonic flow in the case \(p > 2\) has been recently established in [33, 34]. Our main task here is to demonstrate that the small energy regularity estimate holds uniformly for solutions of the penalized equations.

Theorem 11

(Small energy regularity) Let \(p > 2\). Let \(B_0\) and \(a_0\) be positive numbers satisfying the conditions

$$\begin{aligned} \displaystyle { \frac{6 p - 4}{p + 2}< B_0< p; \quad \frac{B_0 - 2}{p - 2} < a_0 \le 1. } \end{aligned}$$
(3.7)

Let \(u = u_{K, \, \epsilon }\) be a regular solution of (2.1) on \((B (R_{\mathcal{M}}))_T = (0, \, T) \times B (R_{\mathcal{M}}, \, 0)\) for a positive \(T < \infty \), satisfying the energy bound

$$\begin{aligned} \displaystyle { \Vert \partial _t u \Vert _{{\mathrm {L}}^2 ({\mathcal{M}}_T)}^2 + \sup _{0< t < T} F_{K, \, \epsilon } (u) \le C_1 } \end{aligned}$$
(3.8)

for a positive number \(C_1\) depending only on \({\mathcal{M}}\), p and \({\mathcal{N}}\). Then, there exists a small positive numeber \(R_0 < 1\), depending only on \({\mathcal{M}}\), \({\mathcal{N}}\), p, \(B_0\), \(a_0\) and \(C_1\), and the following holds true : Let \(\gamma _0\) be any positive number satisfying

$$\begin{aligned} \displaystyle { 2< \gamma _0 < p. } \end{aligned}$$

If, for some small positive \(R < \min \{R_{\mathcal{M}}, \, R_0, \, T^{1/B_0}\}\),

$$\begin{aligned} \displaystyle { \limsup _{r \searrow 0} r^{\gamma _0 - m} \int _{\{t = T - R^{B_0}\} \times B (r, \, 0)} e_{K, \, \epsilon } (u (t, \, x)) \, d {\mathcal{M}} \le 1, } \end{aligned}$$
(3.9)

then, there holds

$$\begin{aligned} \displaystyle { \sup _{ \left( T - (R/4)^{B_0}, \, T\right) \times B (R/4, \, 0) } e_{K, \, \epsilon } (u (t, \, x)) \le C_2 \, R^{- a_0 p}, } \end{aligned}$$
(3.10)

where the positive constant \(C_2\) depends only on \(\gamma _0\), \(B_0\), \(a_0\), p, \({\mathcal{M}}\), \({\mathcal{N}}\) and \(C_1\).

The novelty here is a new monotonicity type estimate of a localized scaled energy, which may be of its own interest. Let us define our localized scaled energy in the following way: Let \(T \ge 0\) be given, and \((t_0, \, x_0)\) in the parabolic like envelope

$$\begin{aligned} \displaystyle { \left\{ (t, \, x) \in (0, \infty ) \times B (R_{\mathcal{M}}) \, : \, \min \{(R_{\mathcal{M}})^{B_0}, \, 1\}> t - T \ge |x|^{B_0} \right\} ; \quad B_0 > 2. } \end{aligned}$$

The localized scaled energy is defined by

$$\begin{aligned}&E_{\pm } (r) = \frac{1}{\Lambda ^p} \int _{\{t = t_0 \pm \Lambda ^{2 - p} r^2\} \times B (R_{\mathcal{M}})} {\bar{e}}_{K, \, \epsilon } (u (t, \, x)) \, \mathcal{B}_{\pm } (t_0, x_0 ; t, x) \, \mathcal{C}^q (t, x) \, d {\mathcal{M}}; \nonumber \\&{\bar{e}}_{K, \, \epsilon } (u) : = \frac{1}{p} \big (\epsilon + |D u|^2\big )^{\frac{p}{2}} + \, C_0 \, \frac{K}{2} \chi \big ({ \text{ dist }}^2 (u, \, {\mathcal{N}}\big ) \end{aligned}$$
(3.11)

and \(\Lambda = \Lambda (r)\) is a function of a scale radius r, defined as

$$\begin{aligned} \displaystyle { \Lambda = \Lambda (r) = r^{ \frac{B_0 - 2}{2 - p} }; \quad p> B_0 > \frac{6 p - 4}{p + 2} } \end{aligned}$$
(3.12)

for any r, \(0< r < R_{\mathcal{M}}/2\), where we note that

$$\begin{aligned} \displaystyle { p> \frac{6 p - 4}{p + 2} \Longleftrightarrow (p - 2)^2 > 0. } \end{aligned}$$

The forward or backward in time Barenblatt like function, denoted by \(\mathcal{B}_-\) and \(\mathcal{B}_-\), respectively, are defined by

$$\begin{aligned} \displaystyle { \mathcal{B}_{\pm } (t_0, \, x_0; \, t, \, x) = \frac{1}{(\mp t_0 \pm t)^{\frac{m}{B_0}}} \left( 1 - \, \left( \frac{|x - x_0|}{2 \, (\mp t_0 \pm t)^{\frac{1}{B_0}}} \right) ^{\frac{p}{p - 1}} \right) ^{\frac{p - 1}{p - 2}}_{+}, \,\,\, \mp t < \mp t_0. }\qquad \end{aligned}$$
(3.13)

The localized function \(\mathcal{C}\) is defined and used as

$$\begin{aligned} \displaystyle { \mathcal{C} (t, \, x) : = \left( (t - T)^{1/B_0} - |x| \right) _+; \quad q > 2. } \end{aligned}$$
(3.14)

We call \(E_+ (r)\) and \(E_- (r)\) the forward and backward localized scaled p-energy, respectively.

Our monotonicity type estimate of a scaled energy is the following. The proof is postponed by Sect. 5

Lemma 12

(Monotonicity estimate for the backward localized scaled p-energy) Let \(p > 2\) and \(q > 2\). For any regular solution u to (2.1) the following estimate holds for all positive numbers \(r, \rho \), \(r^{B_0} = \Lambda (r)^{2 - p} r^2 < \rho ^{B_0} = \Lambda (\rho )^{2 - p} \rho ^2 \le \min \{1, \, (R_{\mathcal{M}})^{B_0}, \, (t_0 - T)/2\}\),

$$\begin{aligned} E_{-} (r)\le & {} E_{-} (\rho ) + \, C \, \left( \rho ^\mu - r^\mu \right) \nonumber \\&+ \, C \, \int \limits _{t_0 - \rho ^{B_0}}^{t_0 - r^{B_0}} \Vert \mathcal{C} ^{q - 2} (t) \, \big ( {\bar{e}}_{K, \, \epsilon } (u (t)) \big )^{\theta _0} \Vert _{ L^\infty \left( B ( (t_0 - t)^{1/B_0}, \, x_0 ) \right) } \, d t, \end{aligned}$$
(3.15)

where \(B_0\) as in (3.12), and the positive exponents \(\theta _0 \ge 2\) and \(\mu \) depend only on \(B_0\), p and \({\mathcal{N}}\), \({\mathcal{M}}\), p and \(B_0\), respectively, and the positive constant C depends only on the same ones as \(\mu \) and q.

Lemma 13

(Monotonicity estimate for the forward localized scaled p-energy) Let \(p > 2\) and \(q > 2\). For any regular solution u to (2.1) the following estimate holds for all positive numbers \(r, \rho \), \(r^{B_0} = \Lambda (r)^{2 - p} r^2 < \rho ^{B_0} = \Lambda (\rho )^{2 - p} \rho ^2 \le \min \{1, \, T - t_0 + (R_{\mathcal{M}})^{B_0}\}\)

$$\begin{aligned} \displaystyle { E_{+} (\rho ) }\le & {} \displaystyle { E_{+} (r) + \, C \, \left( \rho ^\mu - r^\mu \right) } \nonumber \\&+ \, C \, \int \limits _{t_0 + r^{B_0}}^{t_0 + \rho ^{B_0}} \Vert \mathcal{C} ^{q - 2} (t) \, \big ( {\bar{e}}_{K. \, \epsilon } (u (t)) \big )^{\theta _0} \Vert _{ L^\infty \left( B ( (t - t_0)^{1/B_0}, \, x_0 ) \right) } \, d t, \end{aligned}$$
(3.16)

where \(B_0\) as in (3.12), and the positive constants \(\theta _0 \ge 2\), \(\mu \) and C have the same dependence as those in Lemma 12.

From now on we show the validity of Theorem 11.

First of all we make parallel translation \(t^\prime = t - T\), \(x^\prime = x\) of the Eq. (1.4) and its solutions u on \((0, \, T) \times B (R_{\mathcal{M}})\) to those on \((- T, \, 0) \times B (R_{\mathcal{M}})\) with the same notation. The Eq. (1.4) is invariant under parallel transformation.

Under this setting the statement of Theorem 11 is rewritten as

Lemma 14

There exists a positive number \(R_0 < 1\), depending only on \(B_0\), p, \(\mathcal{M}\) and \(\mathcal{N}\), such that the following is valid : If

$$\begin{aligned} \displaystyle { \limsup _{r \searrow 0} r^{\gamma _0 - m} \int _{\{t = - R^{B_0}\} \times B (r, \, 0)} e_{K, \, \epsilon } (u (t, \, x)) \, d {\mathcal{M}} \le 1, } \end{aligned}$$
(3.17)

is satisfied for some small positive \(R \le R_0\) with

$$\begin{aligned} \displaystyle { \gamma _0 = \frac{p (B_0 - 2)}{p - 2}, } \end{aligned}$$
(3.18)

then, it holds that, for a positive constant \(C_2\) depending only on p, \(\mathcal{M}\), \({\mathcal{N}}\) and \(B_0\),

$$\begin{aligned} \displaystyle { \sup _{\left( - (R/4)^{B_0}, \, 0\right) \times B \left( R/4, \, 0 \right) } e_{K, \, \epsilon } (u) \le \, C_2 \, R^{- a_0 p}. } \end{aligned}$$
(3.19)

The proof of Lemma 14 consists of several steps, which are separately explained with those proofs. Our strategy of proof is based on a now classical argument similar to [9, 39], originally introduced by Schoen for the partial regularity of harmonic maps [36]. Here we carefully make local estimates under an intrinsic scaling to the evolutionary p-Laplace operator.

Hereafter in this section we put, for brevity,

$$\begin{aligned} u = u_{K, \, \epsilon }; \quad e (u) = e_{K, \, \epsilon } (u). \end{aligned}$$

Let positive numbers \(\lambda _0 > 2\) and \(a_0 < 1\) be determined later. According to \(\lambda _0\) and \(a_0\), we choose a positive number \(\epsilon \) such that

$$\begin{aligned} \displaystyle { 0< \epsilon < 2 \left( a_0 - \frac{\lambda _0 - 2}{p - 2}\right) , } \end{aligned}$$
(3.20)

where we should choose \(a_0\) as

$$\begin{aligned} \displaystyle { a_0 - \frac{\lambda _0 - 2}{p - 2}> 0 \Longleftrightarrow a_0 > \frac{\lambda _0 - 2}{p - 2}. } \end{aligned}$$
(3.21)

For t, \(- R^{\lambda _0} \le t \le 0\), we define a function f(t) as

(3.22)

where we notice by (3.20) that

$$\begin{aligned} \displaystyle { A_0 = 2 \left( 1 - \frac{\lambda _0 - 2}{a_0 (p - 2)} \right) - \frac{\epsilon }{a_0} > 0 \Longleftrightarrow \epsilon < 2 \left( a_0 - \frac{\lambda _0 - 2}{p - 2}\right) . } \end{aligned}$$

Now we also define a function g(t) as

(3.23)

It is readily seen that, for any t, \(- R^{\lambda _0} \le t \le 0\),

$$\begin{aligned} \left( f (t) \right) ^{\frac{1}{A_0}}= & {} \sup _{ - R^{\lambda _0} \le \tau \le t } \left( \sup _{x \in B \left( \left( \tau + R^{\lambda _0} \right) ^{ \frac{1}{\lambda _0} }, \, 0 \right) } \left( (\tau + R^{\lambda _0})^{\frac{1}{\lambda _0}} - |x| \right) ^{a_0} \, (e (u (\tau , \, x)))^{\frac{1}{p}} \right) \nonumber \\\ge & {} \sup _{x \in B \left( \left( t + R^{\lambda _0} \right) ^{ 1/\lambda _0 }, \, 0 \right) } \left( (t + R^{\lambda _0})^{\frac{1}{\lambda _0}} - |x| \right) ^{a_0} \, (e (u (t, \, x)))^{\frac{1}{p}} = \left( g (t) \right) ^{\frac{1}{A_0}};\nonumber \\&\quad 0 \le g (t) \le f (t). \end{aligned}$$
(3.24)

Let t, \(- R^{\lambda _0} < t \le 0\), be arbitrarily taken and fixed. Then we can choose some time-space points \((t_0, \, x_0)\) such that \(t_0 \in (- R^{\lambda _0}, \, t]\) and \(x_0 \in B \big ( (t_0 + R^{\lambda _0})^{1/\lambda _0}, \, 0 \big )\), and

$$\begin{aligned} \displaystyle { (f (t))^{\frac{1}{A_0}} }= & {} \displaystyle { \left( (t_0 + R^{\lambda _0})^{\frac{1}{\lambda _0}} - |x_0| \right) ^{a_0} (e (u (t_0, \, x_0)))^{\frac{1}{p}} } \nonumber \\= & {} \displaystyle { 4^{a_0} \, (\rho _0)^{a_0} \, (e (u (t_0, \, x_0)))^{\frac{1}{p}} } \end{aligned}$$
(3.25)

where we put

$$\begin{aligned} \displaystyle { \rho _0 : = \frac{ (t_0 + R^{\lambda _0})^{\frac{1}{\lambda _0}} - |x_0| }{4}. } \end{aligned}$$
(3.26)

Here, if \(t_0 = - R^{\lambda _0}\) or \(|x_0| = (t_0 + R^{\lambda _0})^{1/\lambda _0}\), then \(f (t) = 0\) and \(g (t) = 0\).

Refined gradient boundedness on a small region By Lemma 10, we make the gradient bounded by a local scaled energy on a small region. We divide our consideration into two cases.

Case 1.:

First we treat the case that \((\rho _0)^{a_0} \, (e (u (t_0, \, x_0)))^{\frac{1}{p}} \le 1\).

Then we have that

$$\begin{aligned}&\left( \frac{ (t_0 + R^{\lambda _0})^{\frac{1}{\lambda _0}} - |x_0| }{4} \right) ^{a_0} \, (e (u (t_0, \, x_0)))^{\frac{1}{p}} \le 1 \nonumber \\&\quad \Longleftrightarrow \displaystyle { \left( (t_0 + R^{\lambda _0})^{\frac{1}{\lambda _0}} - |x_0| \right) ^{a_0} \, (e (u (t_0, \, x_0)))^{\frac{1}{p}} \le 4^{a_0} } \nonumber \\&\quad \Longleftrightarrow \displaystyle { f (t) \le 4^{p \, a_0 \, A_0}.} \end{aligned}$$
(3.27)

By (3.24) and (3.27) we have

$$\begin{aligned} \displaystyle { g (t) \le f (t) \le 4^{p \, a_0 \, A_0}. } \end{aligned}$$
(3.28)
Case 2.:

Next we study the case that \((\rho _0)^{a_0} (e (u (t_0, \, x_0)))^{\frac{1}{p}} > 1\).

Then we have

$$\begin{aligned} \displaystyle { r_1 : = \left( \frac{1}{ (e (u (t_0, \, x_0)))^{\frac{1}{p}} } \right) ^{\frac{1}{a_0}} < \rho _0 \le 1. } \end{aligned}$$
(3.29)

Let L be

$$\begin{aligned} \displaystyle { L : = (r_1)^{\frac{\lambda _0 - 2}{2 - p}}. } \end{aligned}$$
(3.30)

It holds that

$$\begin{aligned} \displaystyle { L^{2 - p} \, (r_1)^2 \le (\rho _0)^{\lambda _0}, } \end{aligned}$$
(3.31)

because

$$\begin{aligned} \displaystyle { L^{2 - p} \, (r_1)^2 = (r_1)^{\lambda _0} \le (\rho _0)^{\lambda _0} \Longleftrightarrow r_1 \le \rho _0. } \end{aligned}$$

Under (3.31) we have

$$\begin{aligned} \displaystyle { r_1 \, \sup _{Q \left( L^{2 - p} (r_1)^2, \, r_1 \right) (t_0, \, x_0)} (e (u))^{\frac{1}{p}} \le C_1 := 2^{a_0}. } \end{aligned}$$
(3.32)

For the validity of (3.32), we observe from (3.22) and (3.25) that

$$\begin{aligned} \displaystyle { \sup _{ \left( t_0 - (\rho _0)^{\lambda _0}, \, t_0 \right) \times B \left( \rho _0, \, x_0\right) } (e (u))^{\frac{1}{p}} \le 2^{a_0} (e (u (t_0, \, x_0)))^{\frac{1}{p}}. } \end{aligned}$$
(3.33)

Then we find that, for L in (3.30),

$$\begin{aligned} \displaystyle { r_1 \, \sup _{ Q \left( L^{2 - p} (r_1)^2, \, r_1 \right) (t_0, \, x_0) } (e (u))^{\frac{1}{p}} }\le & {} \displaystyle { (r_1)^{a_0} \, \sup _{ Q \left( L^{2 - p} (r_1)^2, \, r_1 \right) (t_0, \, x_0) } (e (u))^{\frac{1}{p}} } \nonumber \\\le & {} \displaystyle { \frac{1}{ (e (u (t_0, \, x_0)))^{\frac{1}{p}} } \, \sup _{ \left( t_0 - (\rho _0)^{\lambda _0}, \, t_0 \right) \times B \left( \rho _0, \, x_0\right) } (e (u))^{\frac{1}{p}} } \nonumber \\\le & {} \displaystyle { 2^{a_0}, } \end{aligned}$$
(3.34)

where we choose \(a_0\) as

$$\begin{aligned} \displaystyle { 0 < a_0 \le 1. } \end{aligned}$$
(3.35)

Here we show the validity of (3.33), through (3.22) and (3.25). For any \(\tau \), \(t_0 - (\rho _0)^{\lambda _0} \le \tau \le t_0\), we find that

$$\begin{aligned}&\displaystyle { (2 \rho _0)^{a_0} \, \sup _{x \in B (\rho _0, \, x_0)} (e (u (\tau , \, x)))^{\frac{1}{p}} \le \sup _{x \in B (\rho _0, \, x_0)} \left( (\tau + R^{\lambda _0})^{1/\lambda _0} - |x| \right) ^{a_0} (e (u (\tau , \, x)))^{\frac{1}{p}}, } \nonumber \\ \end{aligned}$$
(3.36)

because it holds that for any \(\tau \), \(t_0 - (\rho _0)^{\lambda _0} \le \tau \le t_0\), and any \(x \in B (\rho _0, \, x_0)\)

$$\begin{aligned}&\displaystyle { \left( t_0 + R^{\lambda _0} - (\rho _0)^{\lambda _0} \right) ^{1/\lambda _0} \ge \left( t_0 + R^{\lambda _0}\right) ^{1/\lambda _0} - \rho _0 \quad ; } \nonumber \\&\displaystyle { \left( \tau + R^{\lambda _0} \right) ^{1/\lambda _0} - |x| \ge \left( t_0 - (\rho _0)^{\lambda _0} + R^{\lambda _0} \right) ^{1/\lambda _0} - \left( |x_0| + \rho _0\right) } \nonumber \\&\quad \displaystyle { \ge \left( t_0 + R^{\lambda _0} \right) ^{1/\lambda _0} - |x_0| - 2 \rho _0 = 2 \rho _0, } \end{aligned}$$
(3.37)

where we note the definition \(\rho _0\) in (3.26) and use the simple algebraic inequality for any positive number a and b

$$\begin{aligned} \displaystyle { a^{1/\lambda _0} + b^{1/\lambda _0} \ge \left( a + b\right) ^{1/\lambda _0}. } \end{aligned}$$

From (3.36) we obtain that

$$\begin{aligned}&\displaystyle { (2 \rho _0)^{a_0} \sup _{ \left( t_0 - (\rho _0)^{\lambda _0}, \, t_0 \right) \times B \left( \rho _0, \, x_0\right) } (e (u))^{\frac{1}{p}} } \\&\quad \displaystyle { \le \sup _{ t_0 - (\rho _0)^{\lambda _0}< \tau< t_0 } \left( \sup _{ x \in B \left( \rho _0, \, x_0\right) } \left\{ \left( (\tau + R^{\lambda _0})^{1/\lambda _0} - |x| \right) ^{a_0} (e (u (\tau , \, x)))^{\frac{1}{p}} \right\} \right) } \\&\quad \displaystyle { \le \sup _{ - R^{\lambda _0}< \tau < t } \left( \sup _{ x \in B \left( (\tau + R^{\lambda _0})^{1/\lambda _0}, \, 0 \right) } \left\{ \left( (\tau + R^{\lambda _0})^{1/\lambda _0} - |x| \right) ^{a_0} (e (u) (\tau , \, x))^{\frac{1}{p}} \right\} \right) } \\&\quad \displaystyle { = \left( (t_0 + R^{\lambda _0})^{1/\lambda _0} - |x_0| \right) ^{a_0} \, (e (u (t_0, \, x_0)))^{\frac{1}{p}} = (4 \rho _0)^{a_0} \, (e (u (t_0, \, x_0)))^{\frac{1}{p}}, } \end{aligned}$$

where we use that for any \(\tau \), \(t_0 - (\rho _0)^{\lambda _0} \le \tau \le t_0\)

$$\begin{aligned} \displaystyle { B (\rho _0, \, x_0) \subset B \left( (\tau + R^{\lambda _0})^{1/\lambda _0}, \, 0 \right) , } \end{aligned}$$

because by (3.37), for any \(\tau \), \(t_0 - (\rho _0)^{\lambda _0} \le \tau \le t_0\),

$$\begin{aligned} \displaystyle { \left( \tau + R^{\lambda _0} \right) ^{1/\lambda _0} \ge \left( t_0 + R^{\lambda _0} - (\rho _0)^{\lambda _0} \right) ^{1/\lambda _0} \ge |x_0| + \rho _0. } \end{aligned}$$

Thus, (3.33) is actually verified.

Under the choice of parameters \(\lambda _0 > 2\) and \(a_0\) in (3.21) and (3.35), we should have

$$\begin{aligned} \frac{\lambda _0 - 2}{p - 2}< a_0 \le 1&\Longleftarrow&\frac{\lambda _0 - 2}{p - 2} < 1 \end{aligned}$$
(3.38)
$$\begin{aligned}\Longleftrightarrow & {} \lambda _0 < p \end{aligned}$$
(3.39)

and, (3.32) which verifies the condition (3.5) with letting \(r_0 = r_1/2\). Thus, we can apply Lemma 10 and take the \(L^\infty \)-estimate of gradient (3.6), yielding

$$\begin{aligned}&L^{- p} e (u (t_0, \, x_0)) \, \mathcal{C}^q (t_0, \, x_0) \nonumber \\&\quad \le \displaystyle { L^{- p} \sup _{ Q \left( L^{2 - p} \, (r_1/4)^2, \, r_1/4 \right) (t_0, \, x_0) } e (u) \, \mathcal{C}^q } \nonumber \\&\quad \le \displaystyle { \frac{ C \, L^{- 2 p + 2} }{\left| Q \left( L^{2 - p} (r_1/2)^2, \, r_1/2 \right) \right| } \int \limits _{ Q \left( L^{2 - p} (r_1/2)^2, \, r_1/2 \right) (t_0, \, x_0) } (e (u))^{2 - \frac{2}{p}} \, \mathcal{C}^q \, d z + C, } \qquad \end{aligned}$$
(3.40)

where \(C > 0\) depends only on \(a_0\), p, \(\mathcal{M}\) and \({\mathcal{N}}\).

Multiplying the both sides of (3.40) by \(\left( L^{- p} e (u (t_0, \, x_0)) \right) ^{\frac{2}{p} - 1}\), we have

$$\begin{aligned}&\left( L^{- p} e (u (t_0, \, x_0)) \right) ^{ \frac{2}{p} } \, \mathcal{C}^q (t_0, \, x_0) \nonumber \\&\quad \le \frac{C \, L^{- p}}{\left| Q \left( L^{2 - p} \, (r_1/2)^2, \, r_1/2 \right) \right| } \int \limits _{ Q \left( L^{2 - p} \, (r_1/2)^2, \, r_1/2 \right) (t_0, \, x_0) } e (u) \, \mathcal{C}^q \, d z + C, \end{aligned}$$
(3.41)

where by \(r_1\) in (3.29), L in (3.30) and (3.33) we compute as

$$\begin{aligned}&\displaystyle { \left( L^{- p} e (u (t_0, \, x_0)) \right) ^{\frac{2}{p} - 1} = (r_1)^{ (p - 2) \left( a_0 - \frac{\lambda _0 - 2}{p - 2} \right) }; \quad L^{- p + 2} = (r_1)^{\lambda _0 - 2}; }\\&\displaystyle { \sup _{ Q \big ( L^{2 - p} (r_1/2)^2, \, r_1/2 \big ) } (e (u))^{1 - \frac{2}{p}} \le \big ( 2^{a_0} (e (u (t_0, x_0)))^{\frac{1}{p}} \big )^{p - 2} = 2^{a_0 (p - 2)} (r_1)^{- a_0 (p - 2)}; }\\&\displaystyle { \left( L^{- p} e (u (t_0, \, x_0)) \right) ^{\frac{2}{p} - 1} \, L^{- p + 2} \, \sup _{ Q \big ( L^{2 - p} (r_1/2)^2, \, r_1/2 \big ) } (e (u))^{1 - \frac{2}{p}} \le 2^{a_0 (p - 2)}. } \end{aligned}$$

Furthermore, we divide our estimations into two cases, depending on the size of \(r_1\).

The positive number \(q > 2\) is selected later. Recall that the positive number \(\epsilon \) is as in (3.20). Then \(q/\epsilon > 1\).

Case 2-1 : \(0 < r_1 \le (\rho _0)^{\frac{q}{\epsilon }}\) \(; \quad \) Case 2-2 : \((\rho _0)^{\frac{q}{\epsilon }}< r_1 < \rho _0\).

Case 2-1 \(0 < r_1 \le (\rho _0)^{\frac{q}{\epsilon }}\).

Lemma 15

Suppose that

$$\begin{aligned} \displaystyle { 0 < r_1 \le (\rho _0)^{\frac{q}{\epsilon }}. } \end{aligned}$$
(3.42)

Then there exists \(t_0^\prime \in \left[ t_0 - (r_1)^{\lambda _0}/4, \, t_0 \right] \) such that

$$\begin{aligned}&\displaystyle { \big ( e (u (t_0, \, x_0)) \big )^{ \frac{1}{p} \left( 2 \left( 1 - \frac{\lambda _0 - 2}{a_0 (p - 2)}\right) - \frac{\epsilon }{a_0} \right) } \le \frac{ C \,(r_1)^{ \frac{p (\lambda _0 - 2)}{p - 2}} }{ \left| B \left( r_1/2\right) \right| } \, \int \limits _{ \{t = t_0^\prime \} \times B \left( r_1/2, \, x_0\right) } e (u) \, \mathcal{C}^q (t) \, d {\mathcal{M}} + C, } \nonumber \\&\quad \displaystyle { \mathcal{C} (t, \, x) : = \left( (t + R^{\lambda _0})^{1/\lambda _0} - |x| \right) _+; \quad q > 1, } \end{aligned}$$
(3.43)

where the positive constant C depends only on \(a_0\), m, p and \(\mathcal{N}\).

Proof

We will estimate both sides of (3.41).

By \(\rho _0\) in (3.42), \(r_1\) in (3.29) and L in (3.30), the left hand side of (3.41) is computed as

$$\begin{aligned} \mathcal{C}^q (t_0, \, x_0) \, \left( L^{- p} e (u (t_0, \, x_0)) \right) ^{\frac{2}{p}}\ge & {} 4^q \, (r_1)^\epsilon \, \left( L^{- p} e (u (t_0, \, x_0)) \right) ^{\frac{2}{p}} \\= & {} 4^q \, \big ( e (u (t_0, \, x_0)) \big )^{ \frac{1}{p} \left( 2 \left( 1 - \frac{\lambda _0 - 2}{a_0 (p - 2)} \right) - \frac{\epsilon }{a_0} \right) }, \end{aligned}$$

where by (3.42),

$$\begin{aligned} \displaystyle { \mathcal{C}^q (t_0, \, x_0) = (4 \, \rho _0)^q \ge 4^q \, (r_1)^\epsilon } \end{aligned}$$
(3.44)

and the parameters \(a_0\) and \(\epsilon \) satisfy (3.38) and (3.20).

In the right hand side of (3.41), we take the supremum on time to have, by L and \(r_1\) in (3.30),

$$\begin{aligned}&\displaystyle { \frac{ C \, (r_1)^{ \frac{p (\lambda _0 - 2)}{p - 2} } }{\left| Q \left( L^{2 - p} (r_1/2)^2, \, r_1/2 \right) \right| } \int \limits _{ Q \left( L^{2 - p} (r_1/2)^2, \, r_1/2 \right) (t_0, \, x_0) } e (u) \, \mathcal{C}^q \, d z } \nonumber \\&\quad \displaystyle { \le \, C \, \sup _{ t_0 - L^{2 -p} (r_1/2)^2< s < t_0 } \frac{ (r_1)^{\frac{p (\lambda _0 - 2)}{p - 2}} }{\left| B (r_1/2) \right| } \int \limits _{ B (r_1/2, \, x_0) } e (u (s)) \, \mathcal{C}^q \, d {\mathcal{M}} } \nonumber \\&\quad =\displaystyle { \frac{ \, C \, (r_1)^{\frac{p (\lambda _0 - 2)}{p - 2}} }{\left| B (r_1/2) \right| } \int \limits _{ \{s = t_0^\prime \} \times B (r_1/2, \, x_0) } e (u (s)) \, \mathcal{C}^q \, d {\mathcal{M}}, } \end{aligned}$$
(3.45)

where by continuity of the gradient of solution, we choose some \(t_0^\prime \) such that

$$\begin{aligned} \displaystyle { t_0 - L^{2 -p} (r_1/2)^2 \le t_0^\prime \le t_0 \Longleftrightarrow t_0 - (r_1)^{\lambda _0}/4 \le t_0^\prime \le t_0, } \end{aligned}$$
(3.46)

at which the supremum of the second line is attained. \(\square \)

Case 2-2 \((\rho _0)^{ \frac{q}{\epsilon } }< r_1 < \rho _0\).

Lemma 16

Suppose that

$$\begin{aligned} \displaystyle { (\rho _0)^{\frac{q}{\epsilon }}< r_1 < \rho _0. } \end{aligned}$$
(3.47)

Then there exists \(t_0^\prime \in \left[ t_0 - (\rho _0)^{q \, \lambda _0/\epsilon }/4, \, t_0 \right] \) such that

$$\begin{aligned}&\big ( e (u (t_0, \, x_0)) \big )^{ \frac{1}{p} \left( 2 \left( 1 - \frac{\lambda _0 - 2}{a_0 (p - 2)}\right) - \frac{\epsilon }{a_0} \right) } \nonumber \\&\quad \le \frac{C \, (\rho _0)^{ \frac{q p (B_0 - 2)}{\epsilon (p - 2)} } }{ \left| B \left( (\rho _0)^{ q/\epsilon }/2 \right) \right| } \, \int \limits _{ \{t = t_0^\prime \} \times B \left( (\rho _0)^{ q/\epsilon }/2, \, x_0 \right) } e (u (t)) \, \mathcal{C}^q \, d {\mathcal{M}} + C, \end{aligned}$$
(3.48)

where the positive constant C depends only on \(a_0\), p, \(\mathcal{M}\) and \(\mathcal{N}\).

Proof

First we take a look at the inequality (3.43) in Case 2-1. For \(r_1\), \(0 < r_1 \le (\rho _0)^{\frac{q}{\epsilon }}\) it holds that

$$\begin{aligned}&\displaystyle { \big ( e (u (t_0, \, x_0)) \big )^{ \frac{1}{p} \left( 2 \left( 1 - \frac{\lambda _0 - 2}{a_0 (p - 2)}\right) - \frac{\epsilon }{a_0} \right) } \le \frac{ C \, (r_1)^{\frac{p (\lambda _0 - 2)}{p - 2}} }{\left| B \left( r_1/2\right) \right| } \int \limits _{ \{t = t_0^\prime \} \times B \left( r_1/2, \, x_0\right) } e (u (t)) \, \mathcal{C}^q \, d {\mathcal{M}} + C } \nonumber \\&\quad \Longleftrightarrow \displaystyle { (r_1)^{ 2 \left( \frac{\lambda _0 - 2}{p - 2} - a_0\right) + \epsilon } \le \frac{ C \, (r_1)^{\frac{p (\lambda _0 - 2)}{p - 2}} }{\left| B \left( r_1/2\right) \right| } \int \limits _{ \{t = t_0^\prime \} \times B \left( r_1/2, \, x_0\right) } e (u (t)) \, \mathcal{C}^q \, d {\mathcal{M}} + C, } \end{aligned}$$
(3.49)

where we use the definition of \(r_1\) in (3.29). In particular, (3.49) is valid for \(r_1 = (\rho _0)^{\frac{q}{\epsilon }}\) and the corresponding \(t_0^\prime \) as in (3.45) and (3.46)

$$\begin{aligned}&\displaystyle { t_0 - (\rho _0)^{q \, \lambda _0/\epsilon }/4 \le t_0^\prime \le t_0 \quad ; } \nonumber \\&\displaystyle { \Big ( (\rho _0)^{ \frac{q}{\epsilon } } \Big )^{ 2 \left( \frac{\lambda _0 - 2}{p - 2} - a_0 \right) + \epsilon } \le \frac{C \, (\rho _0)^{ \frac{q p (\lambda _0 - 2)}{\epsilon (p - 2)} } }{ \left| B \left( (\rho _0)^{ q/\epsilon }/2 \right) \right| } \, \int \limits _{ \{t = t_0^\prime \} \times B \left( (\rho _0)^{ q/\epsilon }/2, \, x_0 \right) } e (u (t)) \, \mathcal{C}^q \, d {\mathcal{M}} + C. } \nonumber \\ \end{aligned}$$
(3.50)

Thus, for \(r_1\), \((\rho _0)^{ \frac{q}{\epsilon } }< r_1 < \rho _0\), we simply have

$$\begin{aligned}&\displaystyle { (r_1)^{ 2 \left( \frac{\lambda _0 - 2}{p - 2} - a_0 \right) + \epsilon } \le \frac{C \, (\rho _0)^{ \frac{q p (\lambda _0 - 2)}{\epsilon (p - 2)} } }{ \left| B \left( (\rho _0)^{ q/\epsilon }/2 \right) \right| } \, \int \limits _{ \{t = t_0^\prime \} \times B \left( (\rho _0)^{ \lambda _0/B_0 }/2, \, x_0 \right) } e (u (t)) \, \mathcal{C}^q \, d {\mathcal{M}} + C, } \end{aligned}$$

because of (3.20) and (3.38) again. \(\square \)

Now we derive an ordinary differential inequality for g(t), \(- R^{\lambda _0} \le t \le 0\).

Lemma 17

Let \(\lambda _0\), \(B_0\), \(a_0\) and \(\epsilon \) be positive parameters satisfying the conditions

$$\begin{aligned}&\displaystyle { \frac{6 p - 4}{p + 2}< \lambda _0 = B_0 < p; } \end{aligned}$$
(3.51)
$$\begin{aligned}&\displaystyle { \frac{\lambda _0 - 2}{p - 2}< a_0 \le 1; \quad 0< \epsilon < 2 \left( a_0 - \frac{\lambda _0 - 2}{p - 2} \right) . } \end{aligned}$$
(3.52)

Then the differential inequality holds for any positive \(R < 1\) and any t, \(- R^{\lambda _0} \le t \le 0\)

$$\begin{aligned} \displaystyle { g (t) \le g_0 + C \, \int \limits _{- R^{\lambda _0}}^t \, \left( g (\tau ) \right) ^{ \frac{p \theta _0}{A_0} } \, d \tau , } \end{aligned}$$
(3.53)

where the initial data \(g_0\) is

$$\begin{aligned}&\displaystyle { g_0 : = 4^{a_0 A_0} + C \, R^{a_0 A_0} + \, C \, R^{a_0 A_0} \limsup _{\rho \searrow 0} \, \frac{ \rho ^{ \frac{p (B_0 - 2)}{p - 2} } }{|B (\rho )|} \, \int \limits _{ \{t = - R^{\lambda _0}\} \times B (\rho , \, x_0) } e (u (t)) \, d {\mathcal{M}} } \nonumber \\ \end{aligned}$$
(3.54)

and the positive constant C depends only on \(\lambda _0\), p, \(\mathcal{M}\) and \(\mathcal{N}\).

Proof

Simply saying, our desired inequality (3.53) in Lemma 17 is obtained from combining the gradient \(L^\infty \)-estimate on a small region in Lemmata 15 and 16, and the monotonicity estimate of local scaled energy in Lemmata 12 and 13. Here we observe the admissible range of two parameters \(B_0\) in Lemmata 12 and 13, and \(\lambda _0\) in Lemmata 15 and 16, to choose as \(\lambda _0 = B_0\). By (3.12) and (3.39) we have

$$\begin{aligned} \frac{6 p - 4}{p + 2}< B_0< p; \quad 2< \lambda _0 < p \end{aligned}$$

and thus, we can choose \(B_0\) and \(\lambda _0\) as in (3.51), because

$$\begin{aligned} \frac{6 p - 4}{p + 2} < p \Longleftrightarrow (p - 2)^2 > 0. \end{aligned}$$

The choice of \(a_0\) in (3.38) and \(\epsilon \) in (3.20) are as in (3.52).

By use of the monotonicity estimate in Lemmata 12 and 13. we estimate the local in space scaled integral of gradient in the right hand side of (3.43) in Lemma 15 and (3.48) in Lemma 16

Backward monotonicity estimate, Lemma 12   First we apply the backward monotonicity estimate, Lemma 12, for the local scaled energy in the right hand side of (3.43) in Lemma 15 and (3.48) in Lemma 16.

Let us choose the time-component \(t_0\) of the pole of Barenblatt function \(\mathcal{B}_-\) in (5.3) as follows: For each

$$\begin{aligned}&\displaystyle { t_0 - (r_1)^{\lambda _0}/4 \le t_0^\prime \le t_0 \, \text{ in } \text{ Lemma }~15,} \nonumber \\&\text{ or } \nonumber \\&\displaystyle { t_0 - (\rho _0)^{\frac{q \, \lambda _0}{\epsilon }}/4 \le t_0^\prime \le t_0 \, \text{ in } \text{ Lemma }16, \, \text{ with } r_1 \text{ replaced } \text{ by } (\rho _0)^{\frac{q}{\epsilon }}, } \end{aligned}$$
(3.55)

let \(t_0\) be as \(t_1\)

$$\begin{aligned} \displaystyle { t_1 : = t_0^\prime + (r_1)^{B_0} = t_0^\prime + (\Lambda (r_1))^{2 - p} (r_1)^2. } \end{aligned}$$
(3.56)

Then, the local scaled integral in the right hand side of (3.43) and (3.48) is estimated as

$$\begin{aligned}&\displaystyle { \frac{ (r_1)^{\frac{p (B_0 - 2)}{p - 2}} }{ \left| B (r_1/2)\right| } \int \limits _{\{t = t_0^\prime \} \times B (r_1/2, \, x_0)} e (u (t)) \, {\mathcal{C} (t)}^q \, d {\mathcal{M}} } \nonumber \\&\quad \displaystyle { \le \, \frac{C}{(\Lambda (r_1))^p} \int \limits _{ \left\{ t = t_1 - (r_1)^{B_0} \right\} \times B \big (r_1, \, x_0\big ) } e (u (t)) \, \mathcal{B}_- \left( t_1, \, x_0; \, t \right) \, {\mathcal{C} (t)}^q d {\mathcal{M}} }, \end{aligned}$$
(3.57)

because by (3.56) we have, for \(t : = t_0^\prime \),

$$\begin{aligned}&\displaystyle { t = t_1 - (\Lambda (r_1))^{2 - p} \, (r_1)^2 \Longleftrightarrow t_1 - t = (\Lambda (r_1))^{2 - p} (r_1)^2 = (r_1)^{B_0}; \quad x \in B (r_1/2, \, x_0) } \\&\quad \displaystyle { \Longrightarrow (r_1)^{- m} \left( 1 - 2^{- \frac{p}{p - 1}}\right) ^{\frac{p - 1}{p - 2}} \le \mathcal{B}_- (t_1, \, x_0; \, t, \, x) }. \end{aligned}$$

Let \(\rho ^\prime \) be a positive number, chosen as

$$\begin{aligned} \displaystyle { (\rho ^\prime )^{B_0} = \frac{ t_1 + R^{\lambda _0} }{2} } \end{aligned}$$
(3.58)

and then, the backward monotonicity estimate in Lemma 12 yields the upper-boundedness for (3.57) by

$$\begin{aligned}&\displaystyle { \frac{C}{(\Lambda (\rho ^\prime ))^p} \int \limits _{ \left\{ t = t_1 - (\rho ^\prime )^{B_0} \right\} \times B \big (\rho ^\prime , \, x_0\big ) } e (u (t)) \, \mathcal{B}_- \left( t_1 , \, x_0; \, t \right) \, {\mathcal{C} (t)}^q \, d {\mathcal{M}} } \nonumber \\&\quad \displaystyle { + \, \, C \, \left( (\rho ^\prime )^\mu - r^\mu \right) } \nonumber \\&\quad \displaystyle { + \, C \, \int \limits _{t_1 - (\rho ^\prime )^{B_0}} ^{t_1 - (r_1)^{B_0}} \Vert {\mathcal{C} (\tau )}^{q - 2} \, (e (u (\tau )))^{\theta _0} \Vert _{ L^\infty ( B ( (t_1 - \tau )^{1 / B_0}, \, x_0 ) ) } \, d \tau . } \end{aligned}$$
(3.59)

Forward monotonicity estimate, Lemma 13   Next, we use the forward monotonicty estmate in Lemma 13 for estimating the first scaled energy in (3.59).

By use of \(\mathcal{C}\), the first term of (3.59) is evaluated by the forward scaled energy

$$\begin{aligned}&\displaystyle { \frac{C}{(\Lambda (\rho ))^p} \int \limits _{ \left\{ t = t_1 - \rho ^{B_0} \right\} \times B \big (\rho , \, x_0\big ) } e (u (t)) \, \mathcal{B}_+ \left( - R^{\lambda _0} , \, 0 ; \, t \right) \, {\mathcal{C} (t)}^{q - \frac{p - 1}{p - 2}} \, d {\mathcal{M}}, } \end{aligned}$$
(3.60)

since by (3.58) we find that, for \(t : = t_1 - (\Lambda (\rho ^\prime ))^{2 - p} \, (\rho ^\prime )^2\),

$$\begin{aligned}&\displaystyle { t_1 - t = (\rho ^\prime )^{B_0} = \frac{t_1 + R^{\lambda _0}}{2} = t - (- R^{\lambda _0}); } \\&\displaystyle { \left( t_1 - t \right) ^{- \frac{m}{B_0}} = \left( \frac{t_1 + R^{\lambda _0}}{2} \right) ^{- \frac{m}{B_0}} = \left( t - (- R^{\lambda _0}) \right) ^{- \frac{m}{B_0}} } \end{aligned}$$

and the function \(\mathcal{C}\) can be evaluated above as

$$\begin{aligned} \mathcal{C} (t, \, x) = \left( (t + R^{\lambda _0})^{1 / \lambda _0} - |x| \right) _+= & {} (t + R^{\lambda _0})^{1/\lambda _0} \, \left( 1 - \frac{|x|}{(t + R^{\lambda _0})^{1/\lambda _0}} \right) _+ \\\le & {} \left( 1 - \Big ( \frac{|x|}{(t + R^{\lambda _0})^{1/\lambda _0}} \Big )^{\frac{p}{p - 1}} \right) _+; \\ \mathcal{C}^q\le & {} \mathcal{C}^{q - \frac{p - 1}{p - 2}} \, \left( 1 - \Big ( \frac{ |x| }{ (t + R^{\lambda _0})^{1/\lambda _0} } \Big )^{ \frac{p}{p - 1} } \right) _+^{ \frac{p - 1}{p - 2} }, \end{aligned}$$

because

$$\begin{aligned}&\displaystyle { \lambda _0 = B_0; \quad \frac{p}{p - 1} > 1; }\\&\displaystyle { (t + R^{\lambda _0})^{1/\lambda _0} \le R \le 1; \quad \text{ supp } (\mathcal{C} (t)) = B \left( (t + R^{\lambda _0})^{1/\lambda _0}, \, 0 \right) , } \end{aligned}$$

and thus, for \(t : = t_1 -(\Lambda (\rho ^\prime ))^{2 - p} (\rho ^\prime )^2\),

$$\begin{aligned} \displaystyle { \mathcal{B}_- \left( t_1 , \, x_0 ; \, t \right) \, {\mathcal{C} (t)}^q \le \mathcal{B}_+ \left( - R^{\lambda _0} , \, 0 ; \, t \right) \, {\mathcal{C} (t)}^{q - \frac{p - 1}{p - 2}}. } \end{aligned}$$

Also the third term of (3.59) is bounded above by

$$\begin{aligned} \displaystyle { C \, \int \limits _{t_1 - (\rho ^\prime )^{B_0}} ^{t_1 - (r_1)^{B_0}} \Vert {\mathcal{C} (\tau )}^{q - 2} \, (e (u (\tau )))^{\theta _0} \Vert _{ L^\infty ( B ( (\tau + R^{\lambda _0})^{1/B_0}, \, 0 ) ) } \, d \tau , } \end{aligned}$$
(3.61)

because by the support of \(\mathcal{C}\) the region of \(L^\infty \) norm on space is actually

$$\begin{aligned} \displaystyle { B ( (t_1 - \tau )^{1/B_0}, \, x_0 ) \, \cap \, B ( (\tau + R^{\lambda _0})^{1/B_0}, \, 0 ) \subset B ( (\tau + R^{\lambda _0})^{1/B_0}, \, 0 ). } \end{aligned}$$

Then, by the forward monotonicity estimate in Lemma 13 (3.60) is bounded by

$$\begin{aligned}&\displaystyle { \limsup _{\rho \searrow 0} \, \left( \frac{C}{(\Lambda (\rho ))^p} \int \limits _{ \left\{ t = \rho ^{B_0} - R^{\lambda _0} \right\} \times B \big (\rho , \, 0\big ) } e (u (t)) \, \mathcal{B}_+ \left( - R^{\lambda _0} , \, 0 ; \, t \right) \, {\mathcal{C} (t)}^{q - \frac{p - 1}{p - 2}} \, d {\mathcal{M}} \right) } \nonumber \\&\quad \displaystyle { + \, \, C \, (\rho ^\prime )^\mu + \, C \, \int \limits ^{(\rho ^\prime )^{B_0} - R^{\lambda _0}} _{- R^{\lambda _0}} \Vert {\mathcal{C} (\tau )}^{ q - \frac{p - 1}{p - 2} - 2 } \, (e (u (\tau )))^{\theta _0} \Vert _{ L^\infty ( B ( (\tau + R^{\lambda _0})^{1/B_0}, \, 0 ) ) } \, d \tau , } \nonumber \\ \end{aligned}$$
(3.62)

where again, we note that by (3.58)

$$\begin{aligned} \displaystyle { (\rho ^\prime )^{B_0} - R^{\lambda _0} = \frac{t_1 - R^{\lambda _0}}{2} = t_1 - (\rho ^\prime )^{B_0}. } \end{aligned}$$

The first scaled energy term above is estimated above as

$$\begin{aligned} \displaystyle { C \, \limsup _{\rho \searrow 0} \, \left( \frac{\rho ^{ \frac{p (B_0 - 2)}{p - 2} } }{|B (\rho )|} \int \limits _{ \left\{ \tau = \rho ^{\lambda _0} - R^{\lambda _0} \right\} \times B (\rho ) (0) } e (u (\tau )) \, d {\mathcal{M}} \right) . } \end{aligned}$$
(3.63)

Now we combine the estimations above, (3.43), (3.48), (3.57), (3.59), (3.60), (3.61), (3.62) and (3.63) to have

$$\begin{aligned}&\displaystyle { \big ( e (u (t_0, \, x_0)) \big )^{ \frac{1}{p} \left( 2 \left( 1 - \frac{\lambda _0 - 2}{a_0 (p - 2)}\right) - \frac{\epsilon }{a_0} \right) } } \nonumber \\&\quad \displaystyle { \le C \, \limsup _{\rho \searrow 0} \, \left( \frac{\rho ^{ \frac{p (B_0 - 2)}{p - 2} } }{|B (\rho )|} \int \limits _{ \left\{ \tau = \rho ^{B_0} - R^{\lambda _0} \right\} \times B (\rho ) (0) } e (u (\tau )) \, d {\mathcal{M}} \right) } \nonumber \\&\qquad \displaystyle { + \, C \, \left( (\rho ^\prime )^\mu - r^\mu \right) + C \, (\rho ^\prime )^\mu } \nonumber \\&\qquad \displaystyle { + \, C \, \int \limits ^{t_1 - (r_1)^{B_0}} _{- R^{\lambda _0}} \Vert {\mathcal{C} (\tau )}^{ q - \frac{3 p - 5}{p - 2} } \, (e (u (\tau )))^{\theta _0} \Vert _{ L^\infty ( B ( (\tau + R^{\lambda _0})^{1/B_0}, \, 0 ) ) } \, d \tau , } \end{aligned}$$
(3.64)

where the power exponent in the left hand side is positive by (3.52), and the second one in the right hand side is bounded as

$$\begin{aligned} \displaystyle { C \, \left( (\rho ^\prime )^\mu - r^\mu \right) + C \, (\rho ^\prime )^\mu \le C \, R^{\mu }, } \end{aligned}$$
(3.65)

where we recall \(t_1\) in (3.56) and \(\rho ^\prime \) in (3.58)

$$\begin{aligned}&\displaystyle { t_1 = t_0^\prime + \Lambda (r_1)^{2 - p} \, (r_1)^2; \quad (\Lambda (r_1))^{2 - p} \, (r_1)^2 = (r_1)^{B_0}; \quad - R^{\lambda _0}< t_0^\prime < 0; \quad \lambda _0 = B_0 }\\&\quad \displaystyle { \Longrightarrow t_1 \le R^{\lambda _0}; \quad \rho ^\prime = \left( \frac{t_1 + R^{\lambda _0}}{2} \right) ^{1 / B_0} \le \left( R^{\lambda _0} \right) ^{1/B_0} = R. } \end{aligned}$$

Differential inequality We gather (3.64) and (3.65) and then, multiply the resulting inequality by \(\left( \Big (t_0 + R^{\lambda _0}\Big )^{1/\lambda _0} - |x_0| \right) ^{a_0 \, A_0}\) to have

$$\begin{aligned}&\displaystyle { \left( \left( \Big (t_0 + R^{\lambda _0}\Big )^{\frac{1}{\lambda _0}} - |x_0| \right) ^{a_0} e (u (t_0, \, x_0))^{\frac{1}{p}} \right) ^{A_0} } \nonumber \\&\quad \displaystyle { \le \, C \, R^{a_0 A_0} \, \limsup _{\rho \searrow 0} \, \left( \frac{ \rho ^{ \frac{p (\lambda _0 - 2)}{p - 2} } }{|B (\rho )|} \, \int \limits _{ \{t = \rho ^{B_0} - R^{\lambda _0}\} \times B (\rho , \, 0) } e (u (t)) \, d {\mathcal{M}} \right) }\nonumber \\&\qquad \displaystyle { + \, C \, R^{a_0 A_0} \left( 1 + R^\mu \right) } \nonumber \\&\qquad \displaystyle { + \, C \, R^{a_0 A_0} \, \int \limits _{ - R^{\lambda _0} }^t \Vert {\mathcal{C} (\tau )}^{ q - \frac{3 p - 5}{p - 2} } \, (e (u (\tau )))^{\theta _0} \Vert _{ L^\infty \left( B \left( (\tau + R^{\lambda _0})^{1/\lambda 0}, \, 0 \right) \right) } \, d \tau }, \end{aligned}$$
(3.66)

where we note by (3.51) that \(B_0 = \lambda _0\).

Moreover, we will modify some terms in (3.66) for our demand. By (3.24) the left hand side of (3.66) is estimated below by g(t).

In the third term in the right hand side of (3.66) the integrand is bounded by

$$\begin{aligned} \displaystyle { \Vert \, \left( \big ( \tau + R^{\lambda _0} \big )^{\frac{1}{\lambda _0}} - |\cdot | \right) ^{a_0 p \theta _0} \, (e (u (\tau )))^{\theta _0} \, \Vert _{L^\infty B \left( (\tau + R^{\lambda _0})^{1/\lambda _0}, \, 0 \right) }, } \end{aligned}$$
(3.67)

since \(q > 2\) can be chosen to be large, comparing with \(a_0 p \theta _0\) and depending only on p and \(B_0\), in fact,

$$\begin{aligned}&\displaystyle { 0 < a_0 \le 1; \quad q - \frac{3 p - 5}{p - 2} \ge a_0 p \, \theta _0 \Longleftarrow q \ge p \, \theta _0 + \frac{3 p - 5}{p - 2}, } \end{aligned}$$

where \(\theta _0\) depends only on p and \(B_0\).

Finally, collecting (3.28) in Case 1, and (3.66) (3.67) in Case 2, we arrive at our desired estimation (3.53).

Here we observe that the principal integral quantity in (3.66) is rewritten as

$$\begin{aligned}&\displaystyle { \limsup _{\rho \searrow 0} \, \frac{ \rho ^{ \frac{p (B_0 - 2)}{p - 2} } }{|B (\rho )|} \, \int \limits _{ \{t = \rho ^{\lambda _0} - R^{\lambda _0}\} \times B (\rho , \, x_0) } e (u (t)) \, d {\mathcal{M}} } \nonumber \\&\quad \displaystyle { = \limsup _{\rho \searrow 0} \, \frac{ \rho ^{ \frac{p (B_0 - 2)}{p - 2} } }{|B (\rho )|} \, \int \limits _{ \{t = - R^{\lambda _0}\} \times B (\rho , \, x_0) } e (u (t)) \, d {\mathcal{M}}. } \end{aligned}$$
(3.68)

In fact, by time-space continuity of Du, we have the estimation for sufficiently small positive \(\rho \)

$$\begin{aligned} \displaystyle { \frac{ \rho ^{ \frac{p (B_0 - 2)}{p - 2} } }{|B (\rho )|} \, \int \limits _{ B (\rho , \, x_0) } \left| e \left( u \big (\rho ^{\lambda _0} - R^{\lambda _0}\big )\right) - e \left( \big (- R^{\lambda _0}\big )\right) \right| \, d {\mathcal{M}} \le C \, \rho ^{\frac{p (B_0 - 2)}{p - 2} }, } \end{aligned}$$

which converges to 0, by taking the \(\limsup \) on \(\rho \) tending to 0 in the both side. \(\square \)

We are now in position to show the validity of Lemma 14. We solve the differential inequality (3.53) and (3.54), yielding the uniform gradient bound (3.19).

Proof of Lemma 14

The differential inequality (3.53) and (3.54) can be easily solved as

$$\begin{aligned}&\displaystyle { g (t) \le g_0/\left( 1 - C \, (\beta - 1) \left( R^{\lambda _0} + t \right) \, (g_0)^{\beta - 1} \right) ^{\frac{1}{\beta - 1}}, \quad - R^{\lambda _0} \le t \le 0, } \end{aligned}$$
(3.69)

with the exponent

$$\begin{aligned} \displaystyle { \beta = \frac{p \, \theta _0}{A_0} > 1, } \end{aligned}$$

which is satisfied by \(\theta _0 > 1\) and choice in (3.52).

We simply obtain from (3.69)

$$\begin{aligned} \displaystyle { g (t) \le 2^{\frac{1}{\beta - 1}} g_0, \quad - R^{\lambda _0} \le t \le 0, } \end{aligned}$$
(3.70)

under the choice of R such that

$$\begin{aligned} \displaystyle { 1 - C \, (\beta - 1) \left( R^{\lambda _0} + t \right) \, (g_0)^{\beta - 1} \ge 2^{- 1} \Longleftrightarrow \left( \frac{1}{g_0} \right) ^{ \beta - 1 } \frac{1}{2 C (\beta - 1)} \ge R^{\lambda _0} + t, } \end{aligned}$$

which is satisfied by

$$\begin{aligned} \displaystyle { \left( \frac{1}{C} \right) ^{ \frac{\beta - 1}{\lambda _0} } \left( \frac{1}{2 C (\beta - 1)} \right) ^{ \frac{1}{\lambda _0} } \ge R \Longleftarrow 0< g_0 < C; \quad R^{\lambda _0} + t \le R^{\lambda _0} } \end{aligned}$$
(3.71)

and so, let \(R_0\) be the positive number in the left hand side of the first inequality in (3.71). \(\square \)

4 Passing to the limit

In this section we present the proof of Theorem 5, based on Theorem 11. As before we abbreviate the time-space Lebesgue measure \(d t d {\mathcal{M}}\) as dz.

Let \(\{\epsilon _k\}\) and \(\{K_k\}\) be sequences such that \(\epsilon _k \searrow 0\) and \(K_k \nearrow \infty \) as \(k \rightarrow \infty \). Let \(u_{K_k, \, \epsilon _k}\), \(k = 1, 2, \ldots \), be a sequence of solutions of the Cauchy problem with initial data \(u_0\) for the penalized equations (2.1) with approximating numbers \(\epsilon = \epsilon _k\) and \(K = K_k\), obtained in Lemma 6. Hereafter we put \(u_k = u_{K_k, \, \epsilon _k}\) \(e_k (u_k) = e_{K_k, \, \epsilon _k} (u_{K_k, \, \epsilon _k})\), for brevity.

By the energy inequality (2.4), there exist a subsequence of \(\{u_k\}\), also denoted by the same notation, and the limit map u such that, as \(k \rightarrow \infty \),

$$\begin{aligned}&\displaystyle { u_k \longrightarrow u \quad } \text{ weakly } * \text{ in } L^\infty \left( 0, \, \infty ; \, W^{1, p} ({\mathcal{M}}, \, \mathrm{I}\!\mathrm{R}^l)\right) , \end{aligned}$$
(4.1)
$$\begin{aligned}&\displaystyle { \partial _t u_k \longrightarrow \partial _t u \quad } \text{ weakly } \text{ in } L^2\left( {\mathcal{M}}_\infty , \, \mathrm{I}\!\mathrm{R}^l\right) , \end{aligned}$$
(4.2)
$$\begin{aligned}&\displaystyle { D u_k \longrightarrow D u \quad } \text{ weakly } \text{ in } L^p_{\mathrm{loc}} \left( {\mathcal{M}}_\infty , \, \mathrm{I}\!\mathrm{R}^{m l}\right) , \end{aligned}$$
(4.3)
$$\begin{aligned}&\displaystyle { \chi ({ \text{ dist }}^2 (u_k, \, {\mathcal{N}})) \longrightarrow 0 \quad } \text{ strongly } \text{ in } L^2_{\mathrm{loc}} \left( {\mathcal{M}}_\infty , \, \mathrm{I}\!\mathrm{R}^l\right) , \end{aligned}$$
(4.4)
$$\begin{aligned}&\displaystyle { u_k \longrightarrow u \quad } \text{ strongly } \text{ in } L^q_{{\mathrm{loc}}} \left( {\mathcal{M}}_\infty , \, \mathrm{I}\!\mathrm{R}^l\right) \, \, \text{ for } \text{ any } q, 1 \le q < \frac{m p}{(m - p)_+}, \end{aligned}$$
(4.5)

where the strong convergence in (4.5) follows from (4.1) and (4.2) (see [6, Lemma 1.4, p. 28]). Thus, furthermore, for a subsequence \(\{u_k\}\) denoted by the same notation,

$$\begin{aligned} \displaystyle { u_k \longrightarrow u, \quad } \text{ dist } (u_k, \, {\mathcal{N}}) \longrightarrow 0 \quad \text{ almost } \text{ everywhere } \text{ in } {\mathcal{M}}_\infty . \end{aligned}$$
(4.6)

The use of convergence (4.3) and (4.2) in the energy inequality (3.1) for \(u_k\) also yields (1.14) for the limit map u.

We demonstrate that the limit map u is a partial regular weak solution of the p-harmonic flow, as in the statement of Theorem 5. The proof is divided to several steps and proceeded.

Let us define the regular set of the limit map u as

$$\begin{aligned} \displaystyle { \text{ Reg } (u) : = \left\{ \left. z_0 = (t_0, \, x_0) \in {\mathcal{M}}_\infty \, \right| \, u \text{ is } \text{ regular } \text{ in } \text{ a } \text{ neighborhood } \text{ of } z_0 \right\} } \end{aligned}$$

and thus, the singular set as the complement of \(\hbox {Reg}(u)\), \(\Sigma : = \hbox {Sing}(u) = {\mathcal{M}}_\infty {\setminus }\) \(\hbox {Reg}(u)\). By definition, \(\hbox {Reg}(u)\) is a relatively open set of \({\mathcal{M}}_\infty \) and \(\hbox {Sing}(u)\) is relatively closed in \({\mathcal{M}}_\infty \). Let \(R_0\) be a sufficient small positive number, determined in Theorem 11. For \(\tau \), \(0< \tau < \infty \), and R, \(0< R < \min \{R_0, \, \tau ^{1/B_0}\}\), we put two subsets in \(\mathcal{M}\) as

$$\begin{aligned}&\displaystyle { \mathcal{S} (\tau , \, R) : = \Big \{ x_0 \in {\mathcal{M}} \, : \, \limsup _{k \rightarrow \infty } \Big ( \limsup _{r \searrow 0} r^{\gamma _0 - m} \int \limits _{\{t = \tau - R^{B_0}\} \times B (r, \, x_0)} e_k (u_k (t, \, x)) \, d {\mathcal{M}} \Big ) \ge 1 \Big \}; } \nonumber \\&\displaystyle { \mathcal{T} (\tau , \, R) : = \bigcap _{l = 1}^\infty \bigcup _{k = l}^\infty \Big \{ x_0 \in {\mathcal{M}} \, : \, \limsup _{r \searrow 0} r^{\gamma _0 - m} \int \limits _{\{t = \tau - R^{B_0}\} \times B (r, \, x_0)} e_k (u_k (t, \, x)) \, d {\mathcal{M}} > 1/2 \Big \}.}\nonumber \\ \end{aligned}$$
(4.7)

Then, let us define as

$$\begin{aligned} \displaystyle { \mathcal{S} (\tau ) : = \bigcap _{ 0< R< \min \{R_0, \, \tau ^{1/B_0}\} } \mathcal{S} (\tau , \, R) \quad ; \quad \mathcal{S} : = \mathop {\otimes }_{0< \tau < \infty } \mathcal{S} (\tau ), } \end{aligned}$$
(4.8)

where \(\displaystyle {\mathop {\otimes }_{0< \tau < \infty }}\) means the direct product of sets on positive time \(\tau < \infty \).

Regularity of the limit map   We will prove that \(\Sigma = \text{ Sing } (u) \subset \mathcal{S}\). For this purpose, we now show the regularity of limit map u in the complement of \(\mathcal S\). Let \((t_0, \, x_0)\) be in the complement of \(\mathcal S\). Thus, there exist a positive \(R < \min \{R_0, \, (t_0)^{1/B_0}\}\) and an infinite family \(\{u_k\}\) of regular solutions such that

$$\begin{aligned} \displaystyle { \limsup _{r \searrow 0} r^{\gamma _0 - m} \int \limits _{\{t = t_0 - R^{B_0}\} \times B (r, \, x_0)} e_k (u_k (t, \, x)) \, d x < 1. } \end{aligned}$$

Then we can apply Theorem 11 for each \(u_k\) above to obtain

$$\begin{aligned} \displaystyle { \sup _{\left( t_0 - (R/4)^{B_0}, \, t_0\right) \times B (R/4, \, x_0)} e_k (u_k) \le C \, R^{- p a_0}, } \end{aligned}$$
(4.9)

where the positive constant C depends only on \(B_0\), p, \(\mathcal{M}\) and \({\mathcal{N}}\).

Put \(Q : = \Big (t_0 - (R/8)^{B_0}, \, t_0\Big ) \times B (R/8, \, x_0)\). From (4.9), there exists a subsequence of \(\{u_k\}\), denoted by the same notation, such that, as \(k \rightarrow \infty \),

$$\begin{aligned} \displaystyle { D u_k \longrightarrow D u \quad } \text{ weakly } * \text{ in } L^\infty \left( Q\right) ; \displaystyle { \quad \sup _Q |D u| \le C \, R^{- p a_0}. } \end{aligned}$$
(4.10)

Now we will show the uniform continuity of \(\{u_k\}\) in Q. For this purpose we will derive a local \(L^2\) estimate of derivative of the penalty term. For any smooth function \(\phi \) of compact support in Q, we multiply the Bochner type estimate (3.4) in Lemma 9 by \(\phi ^2 \, \sqrt{|g|}\) and integrate by parts in Q to have, letting \(K = K_k\), \(u = u_k\) and \(e (u) = e_k (u_k)\),

$$\begin{aligned}&\displaystyle { \int _Q \phi ^2 \left( \frac{C_1}{2} \, \left( \epsilon + |D u|^2\right) ^{\frac{p - 2}{2}} \left| D^2 u\right| ^2 + \, \frac{C_2}{2} \, \left| \frac{K}{2} D_u \chi \big ({ \text{ dist }}^2 (u, \, {\mathcal{N}}\big ) \right| ^2 \right) \, d z } \nonumber \\&\quad \le \, \int _Q \left( \phi \, |\partial _t \phi | \, e (u) + |D \phi |^2 \left( \frac{2 p}{C_1} \, e (u) + \frac{2}{C_2} \, e (u)^{\frac{2}{p}} \right) \right. \nonumber \\&\qquad \left. + C_3 \, \phi ~2 \left( 1 + e (u)^{\frac{2}{p}}\right) \, e (u)^{2 \left( 1 - \frac{1}{p}\right) } \right) \, d z, \end{aligned}$$
(4.11)

where we use the Cauchy inequality in the first inequality.

Let \((t_0, \, x_0) \subset Q\) be any point and \(r \le R/8\) be any positive number, and \(Q (r) = (t_0 - r^q, \, t_0) \times B (r, \, x_0)\) with \(q > 1\). In (4.11) we choose a smooth function \(\phi \) such that \(0 \le \phi \le 1\), \(\phi = 1\) in Q(r), \(\phi = 0\) outside Q(2r), and \(|D\phi | \le C/r\) and \(|\partial _t \phi | \le C/r^q\). Thus, from (4.9) and (4.11) we obtain

$$\begin{aligned}&\displaystyle { \int _{Q (r)} \left( \frac{C_1}{2} \, \left( \epsilon + |D u|^2\right) ^{\frac{p - 2}{2}} \left| D^2 u\right| ^2 + \, \frac{C_2}{2} \, \left| \frac{K}{2} D_u \chi \big ({ \text{ dist }}^2 (u, \, {\mathcal{N}}\big ) \right| ^2 \right) \, d z } \nonumber \\&\quad \displaystyle { \le C \, \left( r^m + r^{m + q - 2} + r^{m + q}\right) \le C \, r^m } \end{aligned}$$
(4.12)

We also need the Poincaré inequality of parabolic type : Let \(u = u_k\). There exists a positive constant C, depending only on \(\mathcal{M}\) and p, such that, for any \(Q (r) \subset Q\),

$$\begin{aligned} \Vert u - {\bar{u}}_{Q (r)} \Vert _{L^2 (Q (r))}^2\le & {} C \, \left( r^2 \, \Vert D u\Vert _{L^2 (Q (r))}^2 + r^{- m + q - 2} \, \Vert \big ( \epsilon + |D u|^2 \big )^{1/2} \Vert _{L^{p - 1} (Q (r))}^{2 (p - 1)} \right. \nonumber \\&\qquad \left. + r^{2 q} \, \Vert 2^{- 1} K D_u \chi ({ \text{ dist }}^2 (u, \, {\mathcal{N}}))\Vert _{L^2 (Q (r))}^2 \right) , \end{aligned}$$
(4.13)

where \({\bar{u}}_{Q (r)}\) is the integral mean of u in Q(r). For the proof refer to [28].

Substituting (4.9) and (4.12) into (4.13), we have, for any \((t_0, \, x_0) \subset Q\), any positive \(r \le R/8\), and \(Q (r) = (t_0 - r^q, \, t_0) \times B (r, \, x_0)\),

$$\begin{aligned} \displaystyle { \Vert u - {\bar{u}}_{Q (r)} \Vert _{L^2 (Q (r))}^2 \le C \, \left( r^{m + q + 2} + r^{m + 3 q - 2} + r^{m + 2 q} \right) } \end{aligned}$$
(4.14)

and thus, choosing \(q > 1\) in (4.14), we obtain from Campanato’s isomorphism theorem that \(\{u_k\}\) is uniformly Hölder continuous in Q with exponent \(\min \{1, \, q - 1, \, \frac{q}{2}\}\) on the metric \(|t|^{1/ q} + |x|\), uniformly on \(u_k\). Thus, we see that \(\{u_k\}\) is equicontinuous, and uniformly bounded in Q by Lemma 8. Therefore, by Arzela-Ascoli theorem we find for a subsequence of \(\{u_k\}\), denoted by the same notation, and the limit map u that, as \(k \rightarrow \infty \),

$$\begin{aligned}&\displaystyle { u_k \longrightarrow u \quad \text{ uniformly } \text{ in } Q } \end{aligned}$$
(4.15)

and that the limit map u is uniformly continuous in Q. From (4.9) and (4.15), we see that, as \(k \rightarrow \infty \),

$$\begin{aligned} \displaystyle { \chi ({ \text{ dist }}^2 (u_k, \, {\mathcal{N}})) \le C/K_k \longrightarrow 0 \quad \text{ uniformly } \text{ in } Q \quad \Longrightarrow u \in {\mathcal{N}} \quad \text{ in } Q } \end{aligned}$$
(4.16)

Now we will show that the limit map u satisfies the p-harmonic flow equation in Q. From (4.9) and (4.11) we also see that \(\left\{ (K_k/2) \left. D_u \chi \big ({ \text{ dist }}^2 (u, \, {\mathcal{N}}\big ) \right| _{u = u_k} \right\} \) is bounded in \(L^2 (Q, \mathrm{I}\!\mathrm{R}^l)\) and then, there exists a vector-valued function \(\nu \in L^2 (Q, \mathrm{I}\!\mathrm{R}^l)\) such that, as \(k \rightarrow \infty \),

$$\begin{aligned} \displaystyle { (K_k/2) \left. D_u \chi \big ( { \text{ dist }}^2 (u, \, {\mathcal{N}}) \big ) \right| _{u = u_k} \longrightarrow \nu \quad \text{ weakly } \text{ in } L^2 (Q). } \end{aligned}$$
(4.17)

Let \(\mathcal{P}_{\mathcal{N}} \big (u (Q)\big )\) be a neighborhood of u(Q) in \(\mathcal{N}\). Let \(\tau (v)\) be any smooth tangent vector field of \({\mathcal{N}}\) on \(\mathcal{P}_{\mathcal{N}} \big (u (Q)\big )\), \(\tau (v) \in \mathcal{T}_v {\mathcal{N}}\) for any \(v \in \mathcal{P}_{\mathcal{N}} \big (u (Q)\big )\). By (4.15), we can choose a sufficiently large \(k_0\) such that, for any \(k \ge k_0\), \(u_k \in \mathcal{O}_{\delta _{\mathcal{N}}}\) in Q and \(\pi _{\mathcal{N}} (u_k) \in \mathcal{P}_{\mathcal{N}} \big (u (Q)\big ) \subset {\mathcal{N}}\) and \(\tau (\pi _{\mathcal{N}} (u_k)) \in \mathcal{T}_{\pi _{\mathcal{N}} (u_k)} {\mathcal{N}}\) in Q, where \(\mathcal{O}_{\delta _{\mathcal{N}}}\) is a tubular neighborhood in \(\mathrm{I}\!\mathrm{R}^l\) of \({\mathcal{N}}\) with width \(\delta _{\mathcal{N}}\), and \(\pi _{\mathcal{N}}\) is the nearest point projection to \(\mathcal{N}\) from the tubular neighborhood of \(\mathcal{N}\). Thus, we have that

$$\begin{aligned} \displaystyle { \left. D_u \chi \big ( { \text{ dist }}^2 (u, \, {\mathcal{N}}) \big ) \right| _{u = u_k} \cdot \tau (\pi _{\mathcal{N}} (u_k)) }= & {} \displaystyle { 2 \chi ^\prime \text{ dist } (u, \, {\mathcal{N}}) \left. D_u \text{ dist } (u, \, {\mathcal{N}}) \right| _{u = u_k} \cdot \tau \big ( \pi _{\mathcal{N}} (u_k) \big ) } \\= & {} \displaystyle { 0 \quad \text{ in } Q, } \end{aligned}$$

because \(\left. D_u \text{ dist } (u, \, {\mathcal{N}}) \right| _{u = u_k (z)}\) is orthogonal to \(\mathcal{T}_{\pi _{\mathcal{N}} (u_k (z))} {\mathcal{N}}\) for any \(z \in Q\), and then,

$$\begin{aligned} \displaystyle { \int _Q \frac{K_k}{2} \left. D_u \chi \big ( { \text{ dist }}^2 (u, \, {\mathcal{N}}) \big ) \right| _{u = u_k} \cdot \tau ( \pi _{\mathcal{N}} (u_k) ) \, d z = 0. } \end{aligned}$$
(4.18)

By (4.15) and (4.17), we can take the limit as \(k \rightarrow \infty \) in (4.18) to have, for any smooth tangent vector field \(\tau (v)\) of \(\mathcal{N}\) on \(\mathcal{P}_{\mathcal{N}} \big (u (Q)\big ) \subset {\mathcal{N}}\), as \(k \rightarrow \infty \),

$$\begin{aligned}&\displaystyle { 0 = \int _Q \frac{K_k}{2} \left. D_u \chi \big ( { \text{ dist }}^2 (u, \, {\mathcal{N}}) \big ) \right| _{u = u_k} \cdot \tau (\pi _{\mathcal{N}} (u_k)) \, d z \longrightarrow \int _Q \nu \cdot \tau (u) \, d z } \nonumber \\&\quad \displaystyle { \Longrightarrow \, \int _Q \nu \cdot \tau (u) \, d z = 0 } \nonumber \\&\quad \displaystyle { \Longleftrightarrow \nu (z) \, \bot \, \mathcal{T}_{u (z)} {\mathcal{N}} \quad \text{ for } \text{ any } z \in Q } \end{aligned}$$
(4.19)

and thus, \(\nu (z)\) is a normal vector field along u(z) for any \(z \in Q\). In the weak form of (2.1), for any smooth map \(\phi \) with compact support in Q,

$$\begin{aligned} \displaystyle { \int _Q \left( \partial _t u_k \cdot \phi + \big (\epsilon _k + |D u_k|^2\big )^{\frac{p - 2}{2}} g^{\alpha \beta } D_\beta u_k \cdot D_\alpha \phi + \frac{K_k}{2} \left. D_u \chi \big ( { \text{ dist }}^2 (u, \, {\mathcal{N}}) \big ) \right| _{u = u_k} \cdot \phi \right) \, d z = 0, } \end{aligned}$$

we pass to the limit as \(k \rightarrow \infty \) to find that the limit map u satisfies

$$\begin{aligned} \displaystyle { \int _Q \left( \partial _t u \cdot \phi + |D u|^{p - 2} g^{\alpha \beta } D_\beta u \cdot D_\alpha \phi + \nu \cdot \phi \right) \, d z = 0, } \end{aligned}$$
(4.20)

where we use the convergence in the first line of (4.19) and, the strong convergence of gradients \(\{D u_k\}\), obtained from (2.1) with the convergence (4.1), (4.2) and (4.17) (see [6, Theorem 2.1 and its proof, pp. 31–33]). Therefore, we obtain that

$$\begin{aligned} \displaystyle { \partial _t u - \Delta _p u + \nu = 0 \quad \text{ almost } \text{ everywhere } \text{ in } Q \text{ as } L^2 (Q)-\text{ map }. } \end{aligned}$$
(4.21)

We now observe that

$$\begin{aligned} |\nu (z)|= & {} - |D u (z)|^{p - 2} g^{\alpha \beta } (x) D_\beta u (z) \cdot ( D_\alpha u (z) \cdot \left. D_u \gamma (u) \right| _{u = u (z)} )\nonumber \\&\text{ almost } \text{ every } z = (t, \, x) \in Q. \end{aligned}$$
(4.22)

Let \({\bar{z}} = ({\bar{t}}, \, {\bar{x}}) \in Q\) be arbitrarily taken and fixed. Let \(\gamma (v)\) be a smooth unit normal vector field of \({\mathcal{N}}\) in \(u (Q) \subset {\mathcal{N}}\) such that \(\gamma (v) \in (\mathcal{T}_v {\mathcal{N}})^\bot \), \(|\gamma (v)| = 1\) for any \(v \in u (Q)\) and \(\gamma (u ({\bar{z}})) = \nu ({\bar{z}}) / |\nu ({\bar{z}})|\). We take the composite map \(\gamma (u)\) of \(\gamma (\cdot )\) and the limit map u, and use a test function \(\sqrt{|g|} \gamma (u) \, \eta \) for any smooth real-valued function \(\eta \) with compact support in Q to have

$$\begin{aligned}&\displaystyle { \int _Q \left( \partial _t u \cdot \gamma (u) \eta + |D u|^{p - 2} g^{\alpha \beta } D_\beta u \cdot \big ( D_\beta \gamma (u) \eta + \gamma (u) D_\beta \eta \big ) + \nu \cdot \gamma (u) \, \eta \right) \, d z = 0; } \\&\displaystyle { \int _Q \left( |D u|^{p - 2} g^{\alpha \beta } D_\beta u \cdot D_\beta \gamma (u) + \nu \cdot \gamma (u) \right) \, \eta \, d z = 0 } \\&\quad \displaystyle { \Longrightarrow \, \nu \cdot \gamma (u) = - |D u|^{p - 2} g^{\alpha \beta } D_\beta u \cdot D_\beta \gamma (u) \quad \text{ almost } \text{ everywhere } \text{ in } Q, } \end{aligned}$$

where in the second line, we use that \(\partial _t u, D_\alpha u \in \mathcal{T}_u {\mathcal{N}}\), \(\alpha = 1, \ldots , m\), and \(\gamma (u) \in (\mathcal{T}_u {\mathcal{N}})^\bot \) in Q. The last line yields, at \(z = {\bar{z}}\),

$$\begin{aligned} \displaystyle { |\nu ({\bar{z}})| = - |D u ({\bar{z}})|^{p - 2} g^{\alpha \beta } ({\bar{x}}) D_\beta u ({\bar{z}}) \cdot \big ( D_\beta u ({\bar{z}}) \cdot \left. D_u \gamma (u) \right| _{u = u ({\bar{z}})} \big ). } \end{aligned}$$

Thus, (4.22) actually holds true.

Furthermore, there exists a positive constant C depending only on bounds of curvature of \({\mathcal{N}}\) and \((g^{\alpha \beta })\) such that

$$\begin{aligned} \displaystyle { |\nu | \le C \, |D u|^p \quad \text{ almost } \text{ everywhere } \text{ in } Q. } \end{aligned}$$
(4.23)

In fact, from (4.22) we obtain

$$\begin{aligned} \displaystyle { |\nu (z)| \le C \, \max _{v \in u (Q)} |D_v \gamma (v)| \, |D u (z)|^p \quad \text{ for } \text{ almost } \text{ every } z \in Q. } \end{aligned}$$

By (4.23) and (4.10) we have that

(4.24)

where for the last statement of gradient continuity, we refer to [12, Theorem 1.1, p. 245; Sect. 4, p. 291; Sect. 1-(ii), pp. 217–218] (also see [26]).

As a consequence, we have that \((t_0, \, x_0)\) is a regular point and thus, \(\Sigma \subset \mathcal{S}\). Furthermore, from (4.21) and (4.22) it follows that the limit map u satisfies the p-harmonic flow Eq. (1.4) almost everywhere in Q.

Size estimate of the singular set   We recall again that \(\Sigma = \text{ Sing } (u) \subset \mathcal{S}\). Let us estimate the size of \(\mathcal{S}\).

From the definition of limit supremum on k and (4.7), we see that, for every \(\tau \), \(0< \tau < \infty \), and R, \(0< R < \min \{R_0, \, \tau ^{1/B_0}\}\),

$$\begin{aligned} \displaystyle { \mathcal{S} (\tau , \, R) \subset \mathcal{T} (\tau , \, R). } \end{aligned}$$
(4.25)

Here we have the estimation of size (see [17, Theorem 2.2; its proof, pp. 101–103], [21] for the proof) : It holds that, for every \(\tau \), \(0< \tau < \infty \), and R, \(0< R < \min \{R_0, \, \tau ^{1/B_0}\}\),

$$\begin{aligned} \displaystyle { \mathcal{H}^{m - \gamma _0} (\mathcal{T} (\tau , \, R)) = 0 } \end{aligned}$$

and so, by (4.25),

$$\begin{aligned} \displaystyle { \mathcal{H}^{m - \gamma _0} (\mathcal{S} (\tau , \, R)) = 0; \quad \mathcal{H}^{m - \gamma _0} (\mathcal{S} (\tau )) = 0. } \end{aligned}$$

Thus, for any positive \(\tau < \infty \),

$$\begin{aligned} \displaystyle { \{\tau \} \times \Sigma \subset \mathcal{S} (\tau ); \quad \mathcal{H}^{m - \gamma _0} \big (\{\tau \} \times \Sigma \big ) = 0.} \end{aligned}$$

Then, the m-dimensional Hausdorff measure of \(\mathcal{S} \cap {\mathcal{M}}_\infty \) with respect to the time-space metric \(|t|^{1/\gamma _0} + |x|\) is locally zero : For any positive \(T < \infty \), letting \({\mathcal{M}}_T = (0, \, T) \times {\mathcal{M}}\),

$$\begin{aligned} \displaystyle { \mathcal{H}^m \Big ( \mathcal{S} \bigcap {\mathcal{M}}_T \Big ) = \int \limits _0^T \mathcal{H}^{m - \gamma _0} \big (\mathcal{S} (\tau )\big ) \, d \tau = 0.} \end{aligned}$$

Weak solution of the p-harmonic flow   Now we set the two sequences of real-numbers as follows : Let \(\Lambda _0\) be a positive number. Let \(\epsilon \) be any small positive number and \(R_0 < 1\) be a sufficient small positive number, which are sent to zero, later. For positive constants \(M > 1\) and \(\theta < 1\), we define two geometrical progressions as

$$\begin{aligned} \displaystyle { \Lambda _l = \Lambda _0 M^l; \quad R_l = R_0 \theta ^l, \quad l = 0, 1, 2, \ldots . } \end{aligned}$$
(4.26)

It is seen that \(\Lambda _l \nearrow \infty \) and \(R_l \searrow 0\) as \(l \rightarrow \infty \).

Let K be any time-space domain, \(K = (0, T) \times B (R_{\mathcal{M}}, \, x_0)\) for \(T > 0\) and a geodesic ball \(B (R_{\mathcal{M}}, \, x_0)\) in \(\mathcal{M}\). We set a family of sets \(\mathcal{S}_l\), \(l = 0, 1, 2, \ldots \), as

$$\begin{aligned}&\displaystyle { \mathcal{S}_0 = \Big \{z \in {\mathcal{M}}_\infty \, : \, |D u (z)| \le \Lambda _0\Big \} \bigcap \left( K \bigcap \mathcal{S}\right) ; } \nonumber \\&\displaystyle { \mathcal{S}_l = \Big \{z \in {\mathcal{M}}_\infty \, : \, \Lambda _{l - 1} < |D u (z)| \le \Lambda _l\Big \} \bigcap \left( K \bigcap \mathcal{S}\right) , \quad l = 1, 2, \ldots , } \end{aligned}$$
(4.27)

where \(\mathcal{S}\) is as in (4.8).

By the size of \(\mathcal{S}\) shown as before and the compactness of \(K \bigcap \mathcal{S}\), we can choose a covering of \(K \bigcap \mathcal{S}\) in the following way : There exist sequences of positive numbers \(\{r_{l \, i}\}\) and time-space points \(\{z_{l \, i}\}\), \(l = 0, 1, 2, \ldots \); \(i = 1, 2, \ldots , I (l)\) with finite integer I(l) depending on each l, such that, for each \(l = 0, 1, 2, \ldots \),

$$\begin{aligned}&\displaystyle { P (r_{l \, i}) (z_{l \, i}) = (t_{l \, i} - (r_{l \, i})^{\gamma _0}, \, t_{l \, i} + (r_{l \, i})^{\gamma _0}) \times B (r_{l, \, i}) (x_{l \, i}), } \nonumber \\&\displaystyle { z_{l \, i} = (t_{l \, i}, \, x_{l \, i}), \quad r_{l \, i} \le R_l, \quad i = 1, 2, \ldots I (l), } \end{aligned}$$
(4.28)

are a family of time-space cylinders and a covering of \(\mathcal{S}_l\) in the sense that

$$\begin{aligned}&\displaystyle { P_{l \, i} : = P (r_{l \, i}) (z_{l \, i}) \quad : \quad } \text{ disjoint } \text{ each } \text{ other } ; \nonumber \\&\displaystyle { P_{l \, i}^\prime : = P (5 \, r_{l \, i}) (z_{l \, i}), \quad \bigcup _{i = 1}^{I (l)} P_{l \, i}^\prime \supset \mathcal{S}_l; \quad \sum _{i = 1}^{I (l)} (5 \, r_{l \, i})^m \le \epsilon , } \end{aligned}$$
(4.29)

where \(\epsilon \) is firstly taken as small positive number.

Furthermore, by the compactness of \(K \bigcap \mathcal{S}\), we can take a covering of \(K \bigcap \mathcal{S}\) from \(\{P_{l \, i}^\prime \}\), obtained above, which consists of finitely many time-space cylinders \(P_{l \, i}^\prime \), \(l = 0, 1, 2, \ldots , L\) with finite integer L; \(i = 1, 2, \ldots , I (l)\), and has the properties

$$\begin{aligned}&\displaystyle { P_{l \, i} : = P (r_{l \, i}) (z_{l \, i}) \, : \, } \text{ disjoint } \text{ each } \text{ other }, \displaystyle { \quad l = 0, 1, 2, \ldots , L ; \, i = 1, 2, \ldots , I (l); } \nonumber \\&\displaystyle { \bigcup _{l =0}^L \bigcup _{i = 1}^{I (l)} P_{l \, i}^\prime \supset K \bigcap \mathcal{S}; \quad \sum _{l = 0}^L \sum _{i = 1}^{I (l)} (5 \, r_{l \, i})^m \le \epsilon . } \end{aligned}$$
(4.30)

Let \(\eta \) be a smooth function on \({\mathcal{M}}_\infty \) such that \(0 \le \eta \le 1\), \(\eta = 1\) in \(P (1) (0) := (- 1, \, 1) \times B (1) (0)\) and the support of \(\eta \) is contained in \(P (2) (0) : = \big (- 2^{\gamma _0}, \, 2^{\gamma _0}\big ) \times B (2) (0)\), and \(|\partial _t \eta | + |D \eta | \le C\) with a positive number C depending only on m and \(\gamma _0\). For \(l = 0, 1, 2, \ldots \); \(i = 1, 2, \ldots , I (l)\), we denote by \(\eta _{\, l \, i}\) the scaled function

$$\begin{aligned} \eta _{\, l \, i} (t, \, x) = \eta \Big ((t - t_{l \, i})/(5 \, r_{l \, i})^{\gamma _0}, \, (x - x_i)/5 \, r_{l \, i}\Big ) \end{aligned}$$

and then, \( \text{ the } \text{ support } \text{ of } \eta _{\, l \, i} \subset P_{l \, i}^{\prime \prime } : = P (10 \, r_{l \, i}) (z_{l \, i})\).

Let \(\mathcal{L} : = \{0, 1, 2, \ldots , L\}\), \(\mathcal{I} (l) : = \{1, 2, \ldots , I (l)\}\). Let \(\phi \) be any smooth map defined on \({\mathcal{M}}_\infty \) with values into \(\mathrm{I}\!\mathrm{R}^l\) with compact support in K. From (4.21) we obtain

$$\begin{aligned} \displaystyle { 0 }= & {} \displaystyle { \int \limits _{K} \left( \partial _t u - \Delta _p u + \nu \right) \cdot \phi \inf _{l \in \mathcal{L} ; \, i \in \mathcal{I} (l)} (1 - \eta _{\, l \, i}) \, d z } \nonumber \\= & {} \displaystyle { \int \limits _{K} \left( \partial _t u \cdot \phi + |D u|^{p - 2} g^{\alpha \beta } D_\beta u \cdot D_\alpha \phi + \nu \cdot \phi \right) \inf _{l \in \mathcal{L} ; \, i \in \mathcal{I} (l)} (1 - \eta _{\, l \, i}) \, d z } \nonumber \\&\quad \displaystyle { - \int \limits _{K} |D u|^{p - 2} g^{\alpha \beta } D_\beta u \cdot \sup _{l \in \mathcal{L} ; \, i \in \mathcal{I} (l)} (\phi \, D_\alpha \eta _{\, l \, i}) \, d z. } \end{aligned}$$
(4.31)

We note that the number of overlaps of \(\left\{ P_{l \, i}^{\prime \prime } \right\} , l \in \mathcal{L} ;\) \(\, i \in \mathcal{I} (l)\), is at most finite and thus, there exists a subfamily \(\left\{ Q_{l \, i} \right\} \) of \(\left\{ P_{l \, i}^{\prime \prime } \right\} \) such that

$$\begin{aligned} \displaystyle { \left| D \eta _{\, l \, i} (z)\right| = \sup _{l \in \mathcal{L}, \, i \in \mathcal{I} (l)} \left| D \eta _{\, l \, i} (z)\right| \quad } \text{ for } Q_{l \, i} : = \exists P_{l \, i}^{\prime \prime } \ni z \end{aligned}$$

for any \(z \in \bigcup _{l \in \mathcal{L}} \bigcup _{i \in \mathcal{I} (l)} \left( \text{ supp } \left( D \eta _{\, l \, i}\right) \bigcap P_{l \, i}^{\prime \prime } \bigcap \mathcal{S}_l \right) \). Thus, the last error term in (4.31) is estimated above by

$$\begin{aligned} \displaystyle { \int \limits _K |D u|^{p - 1} |\phi | \sup _{l \in \mathcal{L} ; \, i \in \mathcal{I} (l)} |D \eta _{\, l \, i}| \, d z }= & {} \displaystyle { \int \limits _{ \bigcup _{l \in \mathcal{L}} \bigcup _{i \in \mathcal{I} (l)} \left( P_{l \, i}^{\prime \prime } \bigcap \mathcal{S}_l \right) } |D u|^{p - 1} |\phi | \sup _{l \in \mathcal{L} ; \, i \in \mathcal{I} (l)} |D \eta _{\, l \, i}| \, d z }\\= & {} \displaystyle { \int \limits _{ \bigcup _{l \in \mathcal{L}} \bigcup _{i \in \mathcal{I} (l)} \left( Q_{l \, i} \bigcap \mathcal{S}_l \right) } |D u|^{p - 1} |\phi | |D \eta _{\, l \, i}| \, d z }\\\le & {} \displaystyle { \sup _K |\phi | \, \sum _{l = 0}^L \sum _{i = 1}^{I (l)} \left( C \, (r_{l \, i})^{- 1} \int \limits _{ Q_{l \, i} \bigcap \mathcal{S}_{l} } |D u|^{p - 1} \, d z \right) , } \end{aligned}$$

of which the last integral is bounded by

$$\begin{aligned} \displaystyle { (r_{l \, i})^{- 1} \, \left| Q_{l \, i} \right| \, (\Lambda _{l})^{p - 1} }= & {} \displaystyle { (r_{l \, i})^{- 1} \, \left| P_{l \, i}^{\prime \prime } \right| \, (\Lambda _{l})^{p - 1} }\\\le & {} \displaystyle { C \, (\Lambda _{l})^{p - 1} \, (r_{l \, i})^{m + \gamma _0 - 1} } \\\le & {} \displaystyle { C \, (\Lambda _{l})^{p - 1} \, (R_{l})^{\gamma _0 - 1} \, (r_{l \, i})^m, } \end{aligned}$$

with a positive constant C depending only on m, where we use that \(\gamma _0 > 1\) and that, by (4.27) and (4.28), for each \(l = 0, 1, 2, \ldots \),

$$\begin{aligned} \displaystyle { |D u| \le \Lambda _l \quad } \text{ in } \mathcal{S}_l \displaystyle {; \quad r_{l \, i} \le R_l, \quad i = 1, 2, \ldots , I (l). } \end{aligned}$$

Thus, it holds that

$$\begin{aligned} \displaystyle { \int \limits _K |D u|^{p - 1} |\phi | \sup _{ l \in \mathcal{L} ; \, i \in \mathcal{I} (l) } |D \eta _{\, l \, i}| \, d z }\le & {} \displaystyle { C \, \sup _K |\phi | \, \sum _{l = 0}^L (\Lambda _l)^{p - 1} (R_l)^{\gamma _0 - 1} \, \sum _{i = 1}^{I (l)} (r_{l \, i})^m } \nonumber \\\le & {} \displaystyle { C^\prime \, \epsilon \, \sum _{l = 0}^\infty (\Lambda _l)^{p - 1} (R_l)^{\gamma _0 - 1}, } \end{aligned}$$
(4.32)

where we use (4.29) and the positive constant \(C^\prime \) depends only on m, \(\gamma _0\) and \(\sup _K |\phi |\). For summation on l, we choose the ratios \(M > 1\) and \(\theta < 1\) in (4.26) as

$$\begin{aligned} \displaystyle { 0< \theta < 1; \quad M = \theta ^{- a} \quad \text{ for } \text{ some } a > 0 \text{ chosen } \text{ later } } \end{aligned}$$
(4.33)

and compute

$$\begin{aligned} \displaystyle { (\Lambda _l)^{p - 1} (R_l)^{\gamma _0 - 1} = (\Lambda _0)^{p - 1} \, (R_0)^{\gamma _0 - 1} \, \theta ^{l (- a (p - 1) + \gamma _0 - 1)} } \end{aligned}$$

and thus,

$$\begin{aligned} \displaystyle { \sum _{l = 0}^\infty (\Lambda _l)^{p - 1} (R_l)^{\gamma _0 - 1} }\le & {} \displaystyle { (\Lambda _0)^{p - 1} \, (R_0)^{\gamma _0 - 1} \, \sum _{l = 0}^\infty \theta ^{l (- a (p - 1) + \gamma _0 - 1)} } \nonumber \\= & {} \displaystyle { \frac{ (R_0)^{\gamma _0 - 1} \, (\Lambda _0)^{p - 1} }{ 1 - \theta ^{- a (p - 1) + \gamma _0 - 1} }, } \end{aligned}$$
(4.34)

provided that

$$\begin{aligned} \displaystyle { - a (p - 1) + \gamma _0 - 1> 0 \, \Longleftrightarrow \, 0< a < \frac{\gamma _0 - 1}{p - 1}; \quad \gamma _0 > 1. } \end{aligned}$$

Finally, we see from (4.29) and definition of \(\eta _{\, l \, i}\) that

$$\begin{aligned} \displaystyle { \inf _{l \in \mathcal{L} ; \, i \in \mathcal{I}} (1 - \eta _{\, l \, i}) \rightarrow 1 \quad } \text{ almost } \text{ everywhere } \text{ in } K \text{ as } R_0 \searrow 0 \end{aligned}$$
(4.35)

and thus, we can take the limit as \(R_0 \searrow 0\) in (4.32) and (4.34), and use the Lebesgue convergence theorem with (4.35) in the second line of (4.31) to find that the limit map u satisfies the p-harmonic flow equation in the weak sense. \(\square \)

Convergence to the p-harmonic map at a time-infinity   We will present the convergence of u to a p-harmonic map as time tends to infinity. By (1.14) we choose a sequence of time \(\{\tau _l\}\), \(\tau _l \nearrow \infty \), and a limit map \(u_\infty \) such that, as \(l \rightarrow \infty \),

$$\begin{aligned}&\displaystyle { u (\tau _l) \longrightarrow u_\infty \quad } \text{ weakly } \text{ in } W^{1, p} \big ({\mathcal{M}}, \, \mathrm{I}\!\mathrm{R}^l\big ) \end{aligned}$$
(4.36)
$$\begin{aligned}&\displaystyle { D u (\tau _l) \longrightarrow D u_\infty \quad } \text{ weakly } \text{ in } L^p \big ({\mathcal{M}}, \, \mathrm{I}\!\mathrm{R}^{m l}\big ) \end{aligned}$$
(4.37)
$$\begin{aligned}&\displaystyle { \partial _t u (\tau _l) \longrightarrow 0 \quad } \text{ strongly } \text{ in } L^2 \big ({\mathcal{M}}, \, \mathrm{I}\!\mathrm{R}^l\big ), \end{aligned}$$
(4.38)

where from (1.14) we obtain that, for some time-sequence \(\{t_l\}\), \(t_l \nearrow \infty \) as \(l \rightarrow \infty \),

$$\begin{aligned} \displaystyle { \Vert \partial _t u (\tau _l)\Vert _{ L^2 ({\mathcal{M}}) }^2 = \Vert \partial _t u\Vert _{L^2 ( (t_{l - 1}, \, t_l) \times {\mathcal{M}} )}^2 \longrightarrow 0. } \end{aligned}$$

Then, from the convergence (4.1), (4.2) and (4.3), there exists a subsequence of \(\{u_k (\tau _k)\}\), satisfying the same convergence as in (4.36), (4.37) and (4.38) with \(u (\tau _l)\) replaced by \(u_k (\tau _k)\), as \(k \nearrow \infty \).

Let us define the regular set of \(u_\infty \) as

$$\begin{aligned} \displaystyle { \text{ Reg } (u_\infty ) : = \left\{ x_0 \in {\mathcal{M}} \, : \, u_\infty \, \text{ is } \text{ regular } \text{ in } \text{ a } \text{ neghborhood } \text{ of } \, x_0 \right\} } \end{aligned}$$
(4.39)

and the singular set \(\hbox {Sing}(u_\infty )\) as the complement of \(\hbox {Reg}(u_\infty )\), \(\Sigma _\infty : = \hbox {Sing}(u_\infty ) = {\mathcal{M}} {\setminus }\) \(\hbox {Reg}(u_\infty )\). By definition, \(\hbox {Reg}(u_\infty )\) is relatively open in \(\mathcal{M}\) and \(\hbox {Sing}(u_\infty )\) is relatively closed in \({\mathcal{M}}\).

Let us put, for \(0< R < R_0\),

$$\begin{aligned}&\displaystyle { \mathcal{S}_\infty (R) : = \left\{ x_0 \in {\mathcal{M}} \, : \, \limsup _{k \rightarrow \infty } \left( \limsup _{r \searrow 0} \, r^{\gamma _0 - m} \int _{ \{t = \tau _k - R^{B_0}\} \times B (r, \, x_0) } e_k (u_k (t, \, x)) \, d {\mathcal{M}} \right) \ge 1 \right\} ; } \nonumber \\&\displaystyle { \mathcal{S}_\infty : = \bigcap _{0< R < R_0} \mathcal{S}_\infty (R). } \end{aligned}$$
(4.40)

Then, similarly as in Size estimate of the singular set before, we have that, for any positive \(R < R_0\),

$$\begin{aligned} \displaystyle { \mathcal{H}^{m - \gamma _0} \big (\mathcal{S}_\infty (R)\big ) = 0; \quad \mathcal{H}^{m - \gamma _0} \big (\mathcal{S}_\infty \big ) = 0. } \end{aligned}$$
(4.41)

We will show that \(\Sigma _\infty \subset \mathcal{S}_\infty \). Let \(x_0\) be in the complement of \(\mathcal{S}_\infty \) and then, there exist a positive \(R < R_0\), a subsequence of \(\{u_k (\tau _k)\}\), denoted by the same notation as before, such that

$$\begin{aligned} \displaystyle { \limsup _{r \searrow 0} \, r^{\gamma _0 - m} \int \limits _{ \{t = \tau _k - R^{B_0}\} \times B (r, \, x_0) } e_k (u_k (t, \, x)) \, d {\mathcal{M}} < 1. } \end{aligned}$$

Then, by Theorem 11, we have

$$\begin{aligned} \displaystyle { \sup _{ (\tau _k - (R/4)^{B_0}, \, \tau _k) \times B (R/4, \, x_0) } e_k (u_k) \le C \, R^{- p a_0}, } \end{aligned}$$
(4.42)

where the positive constant C depends only on p, \(\mathcal{M}\) and \(\mathcal{N}\). Based on (4.42), we can proceed the same limit process as in (4.11)–(4.24) to find that \(u_\infty \) is regular in \(B (R/8, \, x_0)\) and thus, \(x_0 \in \text{ Reg } (u_\infty ) = {\mathcal{M}} {\setminus } \Sigma _\infty \). Therefore, the complement of \(\mathcal{S}_\infty \) is contained in that of \(\Sigma _\infty \), \( \text{ Reg } (u_\infty )\), and thus, \(\Sigma _\infty \subset \mathcal{S}_\infty \). By use of the size estimate of \(\mathcal{S}_\infty \) in (4.41), we can adopt the similar argument as in (4.26)–(4.35), where time-space regions used are replaced by the corresponding space regions, and thus, find that \(u_\infty \) is a weak solution of the p-harmonic map. \(\square \)

5 Monotonicity estimate of a local scaled energy

We now prove the monotonicity type estimate.

We make parallel translation on time of the Eq. (2.1) and its solutions u on \((T, \, \infty ) \times B (R_{\mathcal{M}})\) to those on \((0, \, \infty ) \times B (R_{\mathcal{M}})\) with the same notation.

Hereafter we assume that the metric \(g = \big (g_{\alpha \beta }\big )\) is the identity matrix. In the general case with \(\big (g_{\alpha \beta }\big )\), the lower order terms containing the derivatives of \(g_{\alpha \beta }\) only appear and controlled well as in the following estimations.

Let \((t_0, \, x_0)\) in the parabolic like envelope \(\displaystyle { \left\{ (t, \, x) \, : \, \min \{1, \, (R_{\mathcal{M}})^{B_0}\} > t \ge |x|^{B_0} \right\} , }\) \(\displaystyle { B_0 > 2. }\)

First we prove the backward monotonicity estimate, Lemma 12. Our localized scaled penalized energy is defined as

$$\begin{aligned}&\displaystyle { E (r) = \frac{1}{\Lambda ^p} \int _{\{t = t_0 - \Lambda ^{2 - p} \, r^2\} \times B (R_{\mathcal{M}})} {\bar{e}} (u (t, \, x)) \, \mathcal{B} (t_0, x_0 ; \, t, x) \, \mathcal{C}^q (t, \, x) \, d x; } \end{aligned}$$
(5.1)
$$\begin{aligned}&\displaystyle { {\bar{e}} (u) : = {\bar{e}}_{K, \, \epsilon } (u) = \frac{1}{p} \big (\epsilon + |D u|^2\big )^{\frac{p}{2}} + \, C_0 \, \frac{K}{2} \chi \big ({ \text{ dist }}^2 (u, \, {\mathcal{N}}\big ); } \nonumber \\&\displaystyle { \Lambda = \Lambda (r) = r^{ \frac{B_0 - 2}{2 - p} }; \quad p> B_0 > \frac{6 p - 4}{p + 2}; \quad 0 < r \le \min \{1, \, R_{\mathcal{M}}, \, (t_0)^{1/B_0}\} } \end{aligned}$$
(5.2)

with weight

$$\begin{aligned}&\displaystyle { \mathcal{B} (t_0, \, x_0 ; \, t, \, x) = \frac{1}{(t_0 - t)^{\frac{m}{B_0}}} \left( 1 - \, \left( \frac{|x - x_0|}{(t_0 - t)^{\frac{1}{B_0}}} \right) ^{\frac{p}{p - 1}} \right) ^{\frac{p - 1}{p - 2}}_{+}, \qquad t < t_0; } \nonumber \\&\displaystyle { \mathcal{C} (t, \, x) = \left( t^{1/B_0} - |x| \right) _+; \quad q > 2. } \end{aligned}$$
(5.3)

Hereafter, for brevity, we use the notation as above.

Lemma 18

Let \(p > 2\) and \(q > 2\). For any regular solution u to (2.1) the following estimate is valid for any positive number \(r < \rho \le \min \{1, \, R_{\mathcal{M}}, \, \left( t_0/2\right) ^{1/B_0}\}\)

$$\begin{aligned} E (r)\le & {} \, E (\rho ) + \, C \, \left( \rho ^\mu - r^\mu \right) \nonumber \\&\quad \displaystyle { + \, C \,\int \limits ^{t_0 - r^{B_0}} _{t_0 - \rho ^{B_0}} \Vert \mathcal{C}^{q - 2} (t) \, ({\bar{e}} (u (t)))^{\theta _0} \Vert _{ L^\infty \left( B ( (t_0 - t)^{1/B_0}, \, x_0 ) \right) } \, d t, } \end{aligned}$$
(5.4)

where

$$\begin{aligned} \displaystyle { \Lambda = \Lambda (r) = r^{\frac{B_0 - 2}{2 - p}}; \quad (\Lambda (r))^{2 - p} \, r^2 = r^{B_0} } \end{aligned}$$

and the positive exponents \(\theta _0 \ge 2\) and \(\mu \) depend only on \(B_0\), p and \(\mathcal{N}\), m, p and \(B_0\), respectively, and the positive constant C depends only on the same ones as \(\mu \) and q.

The proof is proceeded similarly as in [33, Lemmta 5 and 6]. Here we will study how to control well the approximating term, the derivative of penalty term.

Proof of Lemma 18

As before, let

$$\begin{aligned} \displaystyle { \Lambda = r^{ \frac{B_0 - 2}{2 - p} }, \quad p> B_0 > \frac{6 p - 4}{p + 2} } \end{aligned}$$

and let r be any positive number in the range \(0 < r \le \min \left\{ 1, \, R_{\mathcal{M}}, \, \left( t_0/2\right) ^{1/B_0}\right\} \). First we make a scaling transformation intrinsic to the evolutionary p-Laplace operator

$$\begin{aligned} \displaystyle { t = t_0 + \Lambda ^{2 - p} r^2 \, s; \quad x = x_0 + r \, y; \quad v (s, \, y) = \frac{ u (t_0 + \Lambda ^{2 - p} r^2 \, s, \, x_0 + r \, y) }{\Lambda \, r} } \end{aligned}$$
(5.5)

and, under the scaling transformation

$$\begin{aligned} \displaystyle { t = t_0 - \Lambda ^{2 - p} r^2 \, \Longleftrightarrow \, s = - 1. } \end{aligned}$$

Then the scaled solution v is a solution of the scaled equation on \(\{s = - 1\} \times \{y \in \mathrm{I}\!\mathrm{R}^m \, : \, x_0 + r y \in B(R_{\mathcal{M}})\}\)

$$\begin{aligned} \displaystyle { \partial _s v - \text{ div } \left( \big (\Lambda ^{- 2} \epsilon + |D v|^2\big )^{\frac{p - 2}{2}} D v \right) = - C_0 \, \frac{K/\Lambda ^p}{2} D_v \chi \left( \text{ dist }^2 \big (\Lambda \, r \, v, \, {\mathcal{N}}\big )\right) } \end{aligned}$$
(5.6)

and we put the notation

$$\begin{aligned}&\displaystyle { {\bar{\epsilon }} = \Lambda ^{- 2} \epsilon ; } \displaystyle { \quad {\bar{K}} = \Lambda ^{- p} K; } \\&\displaystyle { f = f (v) : = \frac{1}{p} \big ({\bar{\epsilon }} + |D v|^2\big )^{\frac{p}{2}}; } \displaystyle { \quad g = g (v) : = \frac{{\bar{K}}}{2} \chi \big ({ \text{ dist }}^2 (\Lambda r v, \, {\mathcal{N}})\big ); } \\&\displaystyle { \quad \Delta _p v = \text{ div } \left( \big (p \, f\big )^{1 - \frac{2}{p}} D v \right) ; } \displaystyle { \quad {\bar{e}} = {\bar{e}} (v) = f (v) + \, C_0 \, g (v). } \end{aligned}$$

The scaled penalized energy is rewritten as

$$\begin{aligned}&\displaystyle { E (r) = \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } {\bar{e}} (v (s, \, y)) \, \mathcal{B} (s, \, y) \, \mathcal{C}^q (s, \, y) \, d y; } \nonumber \\&\displaystyle { \mathcal{B} (s, \, y) = \frac{1}{ (- s)^{ \frac{m}{B_0} } } \left( 1 - \, \left( \frac{|y|}{ (- s)^{ \frac{1}{B_0} } } \right) ^{ \frac{p}{p - 1} } \right) _+^{ \frac{p - 1}{p - 2} }; \, \, \mathcal{C} (s, \, y) = \left( (t_0 + r^{B_0} \, s)^{1/B_0} - |x_0 + r \, y| \right) _+, }\nonumber \\ \end{aligned}$$
(5.7)

where the integral in (5.7) is well-defined by \(\text{ supp } (\mathcal{C})\) and \(\text{ supp } (\mathcal{B})\) and we simply compute as

$$\begin{aligned}&\displaystyle { D v (s, \, y) = \frac{1}{\Lambda } \, D_x u (t, \, x); \quad {\bar{e}} (v) = \frac{1}{\Lambda ^p} \, {\bar{e}} (u) }\\&\displaystyle { \Lambda = r^{\frac{B_0 - 2}{2 - p}} \Longleftrightarrow \Lambda ^{\frac{p - 2}{B_0}} \, r^{\frac{B_0 - 2}{B_0}} = 1; \quad \mathcal{B} (s, \, y) \, d y = \mathcal{B} \left( t_0, \, x_0 ; \, t, \, x \right) \, d x. } \end{aligned}$$

Our main task in monotonicity estimate is to derive appropriate values of parameter such that

$$\begin{aligned} \displaystyle { p> B_0 > \frac{6 p - 4}{p + 2}. } \end{aligned}$$
(5.8)

Step 1 : differentiation of E(r) on r. We compute differentiation of E(r) on r.

$$\begin{aligned} \displaystyle { \frac{d}{d r} E (r) }= & {} \displaystyle { \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } \, \frac{d}{d r} {\bar{e}} (v) \, \mathcal{B} \, \mathcal{C}^q \, d y + \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } \, {\bar{e}} (v) \, \mathcal{B} \, \frac{d}{d r} \mathcal{C}^q \, d y } \nonumber \\= & {} \displaystyle { \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } \, \left( (p \, f)^{1 - \frac{2}{p}} D v \cdot \frac{d}{d r} D v + C_0 \, \frac{d v}{d r} \cdot D_v g (v) \right) \, \mathcal{B} \, \mathcal{C}^q d y } \nonumber \\&\displaystyle { + \, \frac{B - 2}{r \, (p - 2)} \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } \, \left( \bar{\epsilon } \, (p \, f)^{1 - \frac{2}{p}} + p \, C_0 \, g (v) \right) \, \mathcal{B} \, \mathcal{C}^q \, d y } \nonumber \\&\displaystyle { + \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } \, {\bar{e}} (v) \, \mathcal{B} \, \frac{d}{d r} \mathcal{C}^q \, d y } \nonumber \\= & {} \displaystyle { : I + II + III, } \end{aligned}$$
(5.9)

since

$$\begin{aligned} \displaystyle { \frac{d {\bar{e}} (v)}{d r} = (p \, f)^{1 - \frac{2}{p}} \left( \frac{\bar{\epsilon } (B - 2)}{r \, (p - 2)} + D v \cdot \frac{d}{d r} D v \right) + \frac{C_0 \, p (B - 2)}{r \, (p - 2)} g (v) + C_0 \, \frac{d v}{d r} \cdot D_v g (v). } \end{aligned}$$

Estimations of II and III. The term II is nonnegative.

The term III is estimated by Young’s inequality as

$$\begin{aligned} III= & {} \displaystyle { q \, \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } \, {\bar{e}} (v) \, \mathcal{B} \, \mathcal{C}^{q - 1} \, \frac{d \mathcal{C}}{d r} \, d y } \nonumber \\\ge & {} \displaystyle { - \, \frac{C}{r} \, \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } \, {\bar{e}} (v) \, \mathcal{C}^{q - 1} \, \mathcal{B} \, d y } \nonumber \\\ge & {} \displaystyle { - \, \frac{C}{r^{1 + \delta }} \, \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } \, \left( {\bar{e}} (v)\right) ^{\frac{2 (p - 1)}{p}} \, \mathcal{C}^q \, \mathcal{B} \, d y \, - \, \frac{C \, r^{\frac{\delta p}{p - 2}}}{r} \, \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } \mathcal{C}^{q - \frac{2 (p - 1)}{p - 2}} \, \mathcal{B} \, d y, } \nonumber \\ \end{aligned}$$
(5.10)

where \(\mathcal{C} (s, \, y) : = \Big ( (t_0 + r^{B_0} \, s)^{1/B_0} - |x_0 + r \, y| \Big )_+\) is a Lipschitz function and the derivative of \(\mathcal C\) on r is computed as

$$\begin{aligned} \displaystyle { \left| \frac{d}{d r} \mathcal{C} (s, \, y) \right| }= & {} \displaystyle { \left| \frac{d}{d r} \left( (t_0 + r^{B_0} s)^{1/B_0} - |x_0 + r y| \right) _+ \right| } \\= & {} \displaystyle { \chi _{ \{ |x_0 + r y| \le (t_0 + r^{B_0} s)^{1/B_0} \} } \left| (t_0 + r^{B_0} s)^{1/B_0} \frac{ r^{B_0 - 1} s }{ (t_0 + r^{B_0} s) } - \frac{x_0 + r y}{|x_0 + r y|} \cdot y \right| } \end{aligned}$$

and thus, on the support \(\{y \in \mathrm{I}\!\mathrm{R}^m \, : \, |y| < 1\}\) of \(\mathcal{B} (-1, \, y)\)

$$\begin{aligned} \displaystyle { \left. \left| \frac{d}{d r} \mathcal{C} \right| \right| _{s = - 1} \le 2 \chi _{ \{ |x_0 + r y| \le (t_0 + r^{B_0})^{1 / B_0} \} } \, r^{- 1} } \end{aligned}$$

because of the conditions

$$\begin{aligned} \displaystyle { 0 < t_0 \le 1; \quad \frac{r^{B_0}}{t_0 - r^{B_0}} \le 1 \Longleftrightarrow r^{B_0} \le \frac{t_0}{2}. } \end{aligned}$$

The 1st term of (5.10) is scaled back and bounded below by

$$\begin{aligned}&\displaystyle { - \, \frac{C}{r^{1 + \delta }} \frac{1}{\Lambda ^{2 (p - 1)}} \, \left. \Vert \left( {\bar{e}} (u (t)) \right) ^{\frac{2 (p - 1)}{p}} \, \mathcal{C}^q (t) \Vert _{L^\infty \left( \text{ supp } (\mathcal{B} (t)) \right) } \right| _{t = t_0 - r^{B_0}} \, \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } \mathcal{B} \, d y, } \end{aligned}$$

where we use \(\Lambda ^{2 - p} \, r^2 = r^{B_0}\),

$$\begin{aligned} \displaystyle { \int _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } \mathcal{B} \, d y = \int _{\mathrm{I}\!\mathrm{R}^m} \left( 1 - |y|^{\frac{p}{p - 1}} \right) _+^{\frac{p - 1}{p - 2}} \, d y < \infty } \end{aligned}$$

and the notation

$$\begin{aligned} \displaystyle { {\bar{e}} (u) : = f (u) + C_0 \, g (u) : = \frac{1}{p} \left( \epsilon + |D u|^2\right) ^{\frac{p}{2}} + C_0 \, \frac{K}{2} \chi \left( { \text{ dist }}^2 (u, \, {\mathcal{N}})\right) . } \end{aligned}$$

Each term of I is separately estimated in the following.

Estimation of I. By integration by parts, we have

(5.11)

where the generator of dilation is computed as

$$\begin{aligned} \displaystyle { \frac{d v}{d r} }= & {} \displaystyle { r^{- 1} \, \left( - \big ( 1 + r \, \Lambda ^{- 1} \, \Lambda ^\prime \big ) \, v + \big ( (2 - p) \, r \, \Lambda ^{- 1} \, \Lambda ^\prime + 2 \big ) \, s \, \partial _s v + y \cdot D v \right) . } \end{aligned}$$
(5.12)

Estimation of \(I_1\).

where \({\bar{v}}\) is a weighted integral mean as in (5.14) below, and

$$\begin{aligned}&{\displaystyle \left. - \frac{1}{r} \, \left( 1 + r \, \frac{\Lambda ^\prime }{\Lambda } \right) \, {\bar{v}} \right| _{s = - 1} \cdot \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } \text{ div } \left( (p \, f)^{1 - \frac{2}{p}} D v \, \mathcal{B} \, \mathcal{C}^{q/2} \right) \, d y }\\&\quad = 0, \end{aligned}$$

because of Gauss’s divergence theorem and the compactness of support of \(\mathcal{B}\) and \(\mathcal{C}\).

Each term \(I_{11}\) and \(I_{12}\) is separately estimated in the following.

Estimation of \(I_{11}\). \(I_{11}\) is computed as

(5.13)

Now we will estimate each of three terms in (5.13).

For estimation of \(I_{111}\) we use the Poincaré type inequality with weight of Barenblatt like function [35, Theorem 5.3.4, p. 134]. Let \({\bar{v}}\) be a weighted integral mean

$$\begin{aligned} \displaystyle { {\bar{v}} = \int _{\{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m} (v \, \mathcal{C}^{q/2}) \, \mathcal{B} \, d y/\int _{\{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m} \mathcal{B} \, d y. } \end{aligned}$$
(5.14)

Lemma 19

(Poincaré inequality)

$$\begin{aligned} \displaystyle { \int \limits _{\{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m} \, \Big | (v \, \mathcal{C}^{q/2}) - {\bar{v}} \Big |^2 \, \mathcal{B} \, d y \le C \, \int \limits _{\{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m} \, \Big | D (v \, \mathcal{C}^{q / 2}) \Big |^2 \, \mathcal{B} \, d y. } \end{aligned}$$
(5.15)

\(I_{111}\) of (5.13). \(I_{111}\) is estimated by Cauchy’s inequality for small \(c > 0\) as

$$\begin{aligned}&\displaystyle { I_{111} \ge - \frac{c}{2 \, r} \, \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } |\partial _s v|^2 \, \mathcal{B} \, \mathcal{C}^q \, d y - \frac{1}{2 c \, r} \, \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } \, \Big | (v \, \mathcal{C}^{q/2}) - {\bar{v}} \Big |^2 \, \mathcal{B} \, d y, }\nonumber \\ \end{aligned}$$
(5.16)

where by definition of \(\Lambda \), \(1 + r \, \Lambda ^{- 1} \, \Lambda ^\prime = (p - B_0) (p -2)^{- 1}.\) The 1st time-derivative term is absorbed into that of (5.24) below, later. By the Poincaré inequality (5.15) and Young’s inequality with \(\delta > 0\), the 2nd term is bounded below by

$$\begin{aligned} \displaystyle { - \frac{C}{2 c \, r} \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } \, \Big | D (v \, \mathcal{C}^{q/2}) \Big |^2 \, \mathcal{B} \, d y }\ge & {} - \, \frac{C}{r^{1 + \delta }} \, \int \limits _{\{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m} \big (p \, f\big )^{2 \left( 1 - \frac{1}{p}\right) } \, \mathcal{B} \, \mathcal{C}^q \, d y \nonumber \\&\qquad - \, \frac{C}{ r^{ 1 - \frac{\delta }{p - 2} } } \, \int \limits _{\{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m} \mathcal{B} \, \mathcal{C}^q \, d y \nonumber \\&\qquad - \, \frac{C}{r \, \Lambda ^2} \, \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } \mathcal{C}^{q - 2} \, \mathcal{B} \, d y, \end{aligned}$$
(5.17)

where the last term is obtained from the derivative of \(\mathcal C\) on y, scaling back, a boundedness of the map u with a bound H depending only on \({\mathcal{N}}\) in Lemma 8

$$\begin{aligned} \displaystyle { \left| D_y \mathcal{C} (t, \, x) \right| }= & {} \displaystyle { \chi _{ \{ |x_0 + r y| \le (t_0 + r^{B_0} s)^{1/B_0} \} } \, \left| - \frac{x_0 + r y}{|x_0 + r y|} \, r \right| } \nonumber \\\le & {} \displaystyle { r \, \chi _{ \{ |x_0 + r y| \le (t_0 + r^{B_0} s)^{1/B_0} \} }; } \nonumber \\ \displaystyle { |v|^2 \, |D \mathcal{C}|^2 }\le & {} \displaystyle { \frac{|u|^2}{\Lambda ^2 \, r^2} \, r^2 = \Lambda ^{- 2} \, H^2. } \end{aligned}$$
(5.18)

\(I_{112}\) of (5.13). By Cauchy’s inequality,

$$\begin{aligned}&\displaystyle { I_{112} \ge - \, \frac{C}{r} \, \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } \big (p \, f\big )^{1 - \frac{1}{p}} \, \Big | (v \, \mathcal{C}^{q/2}) - {\bar{v}} \Big | \, |y|^{\frac{1}{p - 1}} \, \left( 1 - \, |y|^{ \frac{p}{p - 1} } \right) _+^{ \frac{1}{p - 2} } \, \mathcal{C}^{q/2} \, d y } \nonumber \\&\quad \displaystyle { \ge - \, \frac{C}{r} \, \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } \, \Big | (v \, \mathcal{C}^{q/2}) - {\bar{v}} \Big |^2 \, \mathcal{B} \, d y } \nonumber \\&\displaystyle { \qquad - \, \frac{C}{r} \, \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } \, \big (p \, f\big )^{2 - \frac{2}{p}} |y|^{\frac{2}{p - 1}} \, \left( 1 - \, |y|^{ \frac{p}{p - 1} } \right) _+^{ \frac{3 - p}{p - 2} } \, \mathcal{C}^q \, d y, } \end{aligned}$$
(5.19)

where the 1st one is the same as the 2nd term in (5.16) and bounded below for \(\delta > 0\) as in (5.17) and, the 2nd one of (5.19), together with the 1st one of (5.17), is estimated below by

$$\begin{aligned} \displaystyle { - \frac{ C \, (r^{- \delta } + 1) }{ r \, \Lambda ^{2 (p - 1)} } \, \left. \Vert {{ \mathcal C} (t) }^q \, \big (p \, f (u (t))\big )^{ 2 \left( 1 - \frac{1}{p}\right) } \Vert _{L^\infty \left( \text{ supp } \, (\mathcal{B} (t)) \right) } \right| _{t = t_0 - r^{B_0}} }, \end{aligned}$$
(5.20)

where we make a scaling back and compute as

$$\begin{aligned} \displaystyle { \int \limits _{\mathrm{I}\!\mathrm{R}^m} \, |y|^{\frac{2}{p - 1}} \left( 1 - \, |y|^{\frac{p}{p - 1}} \right) _+^{\frac{3 - p}{p - 2}} \, d y < \infty ; \quad \frac{3 - p}{p - 2}> - 1 \Longleftrightarrow 3 > 2.} \end{aligned}$$

\(I_{113}\) of (5.13). By the boundedness (5.18) of derivative of \(\mathcal C\) and Cauchy’s inequality,

$$\begin{aligned} I_{113}\ge & {} - \frac{q (p - B_0)}{2 (p - 2) \, r} \int \limits _{\{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m} \big (p \, f\big )^{1 - \frac{1}{p}} \, \Big | v \, \mathcal{C}^{q/2} - {\bar{v}} \Big | \, \mathcal{C}^{q/2 - 1} \, |D \mathcal{C}| \, \mathcal{B} \, d y \\\ge & {} - \, C \, \int \limits _{\{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m} \big (p \, f\big )^{2 - \frac{2}{p}} \, \mathcal{C}^{q - 2} \, \mathcal{B} \, d y - \, C \, \int \limits _{\{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m} \Big | v \, \mathcal{C}^{q/2} - {\bar{v}} \Big |^2 \, \mathcal{B} \, d y, \end{aligned}$$

of which the 1st term is estimated, similarly as in (5.20), below by

$$\begin{aligned} \displaystyle { - \frac{C \, (r^{- \delta } + 1)}{r \, \Lambda ^{2 (p - 1)}} \, \left. \Vert {\mathcal{C} (t)}^{q - 2} \, \big (p \, f (u (t))\big )^{2 \left( 1 - \frac{1}{p}\right) } \Vert _{L^\infty \left( \text{ supp } \, (\mathcal{B} (t)) \right) } \right| _{t = t_0 - r^{B_0}} } \end{aligned}$$
(5.21)

and the 2nd term is bounded below as in (5.17).

Estimation of \(I_{12}\). By Cauchy’s inequality with \(\delta > 0\), we estimate as

$$\begin{aligned}&\displaystyle { r^{- 1} \, \left| {\bar{v}} \cdot \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } D_v g \, \mathcal{B} \, \mathcal{C}^{q/2} \, d y \right| \le r^{- 1} \, |{\bar{v}}| \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } |D_v g| \, \mathcal{B} \, \mathcal{C}^{q/2} \, d y } \\&\quad \displaystyle { \le \frac{C (\Vert u_0\Vert _{L^\infty (\mathrm{I}\!\mathrm{R}^m)}, \, {\mathcal{N}})}{\Lambda \, r^2} \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } |D_v g| \, \mathcal{B} \, \mathcal{C}^{q/2} \, d y } \\&\quad \displaystyle { \le C \, r^{\delta - 1} \, \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } \mathcal{B} \, d y + \frac{C}{ \Lambda ^2 \, r^{3 + \delta } } \, \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } |D_v g|^2 \, \mathcal{B} \, \mathcal{C}^q \, d y, } \end{aligned}$$

where we use a boundedness of u with \(H > 0\) depending only on \({\mathcal{N}}\) in Lemma 8 to have

$$\begin{aligned} \displaystyle { |v (s)| = \frac{|u|}{\Lambda \, r} \le \frac{H}{\Lambda \, r}; \quad |{\bar{v}}| \le \int \limits _{\{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m} |v (s)| \mathcal{B} \, d y/\int \limits _{\{s = -1\} \times \mathrm{I}\!\mathrm{R}^m} \mathcal{B} \, d y \le \frac{H}{\Lambda \, r}. } \end{aligned}$$

Estimation of \(I_2\). As before by Cauchy’s inequality

$$\begin{aligned} I_2\ge & {} - \frac{q (p - B_0)}{2 (p - 2) \, r} \, \int \limits _{\{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m} \big (p \, f\big )^{1 - \frac{1}{p}} \, |v| \, \mathcal{C}^{q/2 - 1} |D \mathcal{C}| \, \mathcal{C}^{q/2} \, \mathcal{B} \, d y \nonumber \\\ge & {} - \, \frac{C}{r} \, \int \limits _{\{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m} (p \, f)^{2 \left( 1 - \frac{1}{p}\right) } \, \mathcal{C}^q \, \mathcal{B} \, d y - \, \frac{C}{r \Lambda ^2} \, \int \limits _{\{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m} \mathcal{C}^{q - 2} \, \mathcal{B} \, d y, \end{aligned}$$
(5.22)

of which the 1st term is estimated below by (5.20).

Estimation of \(I_3\). \(I_3\) is treated as

(5.23)

Moreover each term of (5.23) is arranged as

$$\begin{aligned}&\displaystyle { \frac{1}{r} \, \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } \, (- s) \, \left( (2 - p) \, r \, \Lambda ^{- 1} \, \Lambda ^\prime + 2 \right) \, |\partial _s v|^2 \, \mathcal{B} \, \mathcal{C}^q \, d y } \nonumber \\&\qquad \displaystyle { - \, \frac{1}{r} \, \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } \left( y \cdot D v\right) \cdot \partial _s v \, \mathcal{B} \, \mathcal{C}^q \, d y } \nonumber \\&\displaystyle { \qquad + \, \frac{p}{r \, (p - 2)} \, \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } \, \left\{ \big (p \, f\big )^{1 - \frac{2}{p}} \, \left| y \cdot D v\right| ^2 \right. } \nonumber \\&\displaystyle { \left. \qquad \qquad \qquad \qquad + \, \left( (2 - p) \, r \, \Lambda ^{- 1} \, \Lambda ^\prime + 2 \right) \, \big (p \, f\big )^{1 - \frac{2}{p}} \left( y \cdot D v\right) \cdot \left( s \, \partial _s v\right) \right\} \, \mathcal{C}^q \times } \nonumber \\&\displaystyle { \qquad \qquad \qquad \qquad \quad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \times \, |y|^{- \frac{p - 2}{p - 1}} \, \left( 1 - \, |y|^{ \frac{p}{p - 1} } \right) _+^{ \frac{1}{p - 2} } \, \, d y } \nonumber \\&\displaystyle { \qquad - \, \frac{1}{r} \, \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } \, \left( \left( (2 - p) \, r \, \Lambda ^{- 1} \, \Lambda ^\prime + 2 \right) \, s \, \partial _s v + y \cdot D v \right) \, \big (p \, f\big )^{1 - \frac{2}{p}} D v \cdot D \mathcal{C}^q \, \mathcal{B} \, \, d y. } \nonumber \\&\quad \displaystyle { = : I_{31} + I_{32} +I_{33} + I_{34} + I_{35}. } \end{aligned}$$
(5.24)

Now each term in (5.24) is separately estimated.

\(I_{31}\) of (5.24). \(I_{31} \ge 0\) by the positivity of the coefficient. In fact, by definition of \(\Lambda \) and \(s = - 1\)

$$\begin{aligned} \displaystyle { (- s) \, \left( (2 - p) \, r \, \Lambda ^{- 1} \, \Lambda ^\prime + 2 \right) = B_0> 0 \Longleftrightarrow \Lambda = r^{(B_0 - 2)/(2 - p)}, \quad B_0 > 0. } \end{aligned}$$

\(I_{32}\) of (5.24). By Cauchy’s inequality for small \(c > 0\),

$$\begin{aligned} \displaystyle { I_{32} \ge \frac{c}{2 \, r} \, \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } \, |\partial _s v|^2 \, \mathcal{B} \, \mathcal{C}^q \, d y - \frac{1}{2 c \, r} \, \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } |y|^2 \, |D v|^2 \, \mathcal{B} \, \mathcal{C}^q \, d y. } \end{aligned}$$

The time-derivative term is absorbed into \(I_{31}\). By Young’s inequality the 2nd term is estimated below for \(\delta > 0\) by

$$\begin{aligned} \displaystyle { - \, \frac{C}{r^{1 + \delta }} \, \int \limits _{\{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m} |D v|^{2 (p - 1)} \, \mathcal{B} \, \mathcal{C}^q \, d y - \, \frac{C}{r^{1 - \frac{\delta }{p - 2}}} \, \int \limits _{\{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m} \mathcal{B} \, \mathcal{C}^q \, d y, } \end{aligned}$$
(5.25)

of which the 1st term is bounded below by (5.20).

\(I_{33}\) of (5.24). Clearly, \(I_{33} \ge 0\).

\(I_{34}\) of (5.24).

By Cauchy’s inequality for small \(c > 0\),

$$\begin{aligned} I_{34}\ge & {} - \, \frac{p \, B_0}{r \, (p - 2)} \, \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } \, |\partial _s v| \, \big (p \, f\big )^{1 - \frac{1}{p}} \, |y|^{\frac{1}{p - 1}} \, \left( 1 - \, |y|^{ \frac{p}{p - 1} } \right) _+^{ \frac{1}{p - 2} } \, \mathcal{C}^q \, d y \nonumber \\\ge & {} - \frac{c}{2 \, r} \, \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } \, |\partial _s v|^2 \, \mathcal{B} \, \mathcal{C}^q \, d y \nonumber \\&\displaystyle { - \frac{C}{2 c \, r} \, \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } \, \big (p \, f\big )^{2 \left( 1 - \frac{1}{p}\right) } \, |y|^{\frac{2}{p - 1}} \, \left( 1 - \, |y|^{ \frac{p}{p - 1} } \right) _+^{ \frac{3 - p}{p - 2} } \, \mathcal{C}^q \, d y, } \end{aligned}$$
(5.26)

where in the 1st inequality \((2 - p) \, r \, \Lambda ^{- 1} \, \Lambda ^\prime + 2 = B_0\) as before. The 1st term of (5.26) is absorbed into \(I_{31}\). The 2nd term of (5.26) is estimated below by (5.20).

\(I_{35}\) of (5.24). By Young’s inequality and the estimation (5.18) of derivative of \(\mathcal{C}\),

$$\begin{aligned}&\displaystyle { I_{35} \ge - \, \frac{c}{2} \, \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } \, |\partial _s v|^2 \, \mathcal{C}^q \, \mathcal{B} \, d y - \, \frac{1}{2 c} \, \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } \big (p \, f\big )^{2 \left( 1 - \frac{1}{p}\right) } \, \mathcal{C}^{q - 2} \, \mathcal{B} \, d y } \nonumber \\&\displaystyle { \quad - \, C \, \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } \big (p \, f\big )^{2 \left( 1 - \frac{1}{p}\right) } \, \mathcal{C}^{ \frac{2 (q - 1) (p - 1)}{p} } \, \mathcal{B} \, d y - \, C \, \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } \mathcal{B} \, d y, } \end{aligned}$$
(5.27)

where the 1st term can be absorbed into \(I_{31}\) and the 2nd and 3rd terms are bounded below by

$$\begin{aligned} \displaystyle { - \frac{C \, (r^{- \delta } + 1)}{r \, \Lambda ^{2 (p - 1)}} \, \left. \Vert {\mathcal{C} (t)}^{q - 2} \, \big (p \, f (u (t))\big )^{2 \left( 1 - \frac{1}{p}\right) } \Vert _{L^\infty \left( \text{ supp } \, (\mathcal{B} (t)) \right) } \right| _{t = t_0 - r^{B_0}}, } \end{aligned}$$
(5.28)

because \(2 (q - 1) (p - 1)/p > q - 2\) \(\Longleftarrow \) \(q > 2\).

Resulting estimation. Combining all of the estimations above we have

$$\begin{aligned} \displaystyle { \frac{d}{d r} E (r) }= & {} \displaystyle { I + II + III } \nonumber \\\ge & {} \displaystyle { J - \, C \, \left( r^{- 1 + \frac{\delta }{p - 2}} + r^{- 1 + \delta } \right) - \, \frac{C}{r^{3 + \delta } \, \Lambda ^2} \, \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } |D_v g (v)|^2 \, \mathcal{B} \, \mathcal{C}^q \, d y } \nonumber \\&\displaystyle { \, \left. - \, C \, \frac{r^{- \delta } + 1}{r \, \Lambda ^{2 (p - 1)}} \, \Vert \mathcal{C} (t)^{q - 2} \, {\bar{e}} (u (t))^{\frac{2 (p - 1)}{p}} \Vert _{L^\infty \left( \text{ supp } \, (\mathcal{B} (t)) \right) } \right| _{t = t_0 - r^{B_0}} } \end{aligned}$$
(5.29)

with \(\Lambda = r^{(B_0 - 2)/(2 - p)}\), according to (5.20), (5.21) and (5.28), and

$$\begin{aligned} \displaystyle { J }= & {} \displaystyle { \frac{B_0}{2 \, r} \, \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } \, (- s) \, |\partial _s v|^2 \, \mathcal{B} \, \mathcal{C}^q \, d y } \\&\displaystyle { + \frac{p}{r \, (p - 2)} \, \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } \, |D v|^{p - 2} \, \left| y \cdot D v\right| ^2 \, |y|^{- \frac{p - 2}{p - 1}} \, \left( 1 - \, |y|^{ \frac{p}{p - 1} } \right) _+^{ \frac{1}{p - 2} } \, \mathcal{C}^q \, d y. } \end{aligned}$$

The term J is clearly nonnegative. From (5.29) integrated on the interval \(\left( r, \, \rho \right) \) we derive

$$\begin{aligned}&\displaystyle { E (\rho ) - E (r) } \nonumber \\&\quad \displaystyle { \ge - \, C \, \int _r^\rho \, \left( r^{- 1 + \frac{\delta }{p - 2}} + r^{- 1 + \delta } \right) \, d r \, - \, \int _r^\rho \, \frac{C}{r^{3 + \delta } \, \Lambda ^2} \, \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } |D_v g (v)|^2 \, \mathcal{B} \, \mathcal{C}^q \, d y \, d r } \nonumber \\&\qquad \displaystyle { \, - \, C \, \int _r^\rho \, \frac{r^{- \delta } + 1}{r \, \Lambda ^{2 (p - 1)}} \, \left. \Vert {\mathcal{C} (t)}^{q - 2} {\bar{e}} (u (t))^{\frac{2 (p - 1)}{p}} \Vert _{L^\infty \left( \text{ supp } (\mathcal{B} (t)) \right) } \right| _{t = t_0 - \Lambda ^{2 - p} \, r^2} \, d r. } \nonumber \\&\quad \displaystyle { = : C \, \big (U_1 + U_2 + U_3\big ). } \end{aligned}$$
(5.30)

Step 2 : a uniform bound. We will make a bound of each term \(U_i\), \(i = 1, 2, 3\), in the right hand side of (5.30).

\(U_1\) of (5.30). The 1st integrals on r in the 2nd line of (5.30) are computed as

$$\begin{aligned}&\displaystyle { \int _r^\rho \, r^{- 1 + \frac{\delta }{p - 2}} \, d r = \frac{p - 2}{\delta } \, \left( \rho ^{ \frac{\delta }{p - 2} } - r^{ \frac{\delta }{p - 2} } \right) ; \quad \int _r^\rho \, r^{- 1 + \delta } \, d r = \frac{1}{\delta } \, \left( \rho ^{ \delta } - r^{ \delta } \right) , } \nonumber \\ \end{aligned}$$
(5.31)

\(U_3\) of (5.30). \(- U_3\) is computed as

$$\begin{aligned}&\int _r^\rho r^{- 1} \, \left( - B_0 \, \Lambda ^{2 - p} \, r \right) ^{- 1}\times \nonumber \\&\qquad \times \frac{1}{ \Lambda ^{2 (p - 1)} } \Lambda ^\delta \Vert {\mathcal{C} (t)}^{q - 2} \, {\bar{e}} (u (t))^{\frac{2 (p - 1)}{p}} \Vert _{ L^\infty \left( \text{ supp } \, (\mathcal{B} (t)) \right) } \, \left( - B_0 \, \Lambda ^{2 - p} \, r \right) \, d t \nonumber \\&\quad = \frac{1}{B_0} \, \int \limits ^{t_0 - \left( \Lambda (r)\right) ^{2 - p} \, r^2} _{t_0 - \left( \Lambda (\rho )\right) ^{2 - p} \, \rho ^2} \, \left( t_0 - t \right) ^{ - 1 + \frac{2 (p - 1) (B_0 - 2)}{B_0 (p - 2)} - \frac{\delta (B_0 - 2)}{B_0 (p - 2)} }\times \nonumber \\&\qquad \times \Vert {\mathcal{C} (t)}^{q - 2} \, {\bar{e}} (u (t))^{\frac{2 (p - 1)}{p}} \Vert _{ L^\infty \left( \text{ supp } \, (\mathcal{B} (t)) \right) } \, d t, \end{aligned}$$
(5.32)

where by definition of \(\Lambda \)

$$\begin{aligned} \displaystyle { \Lambda = r^{(B_0 - 2)/(2 - p)} \, \Longleftrightarrow \, \left( \Lambda (r)\right) ^{2 - p} \, r^2 = r^{B_0} } \end{aligned}$$

and, in the last term we make a changing of variable

$$\begin{aligned}&\displaystyle { t = t_0 - \Lambda ^{2 - p} \, r^2 \, \Longleftrightarrow \, t_0 - t = \Lambda ^{2 - p} \, r^2 = r^{B_0}; }\\&\displaystyle { \frac{d t}{d r} = - B_0 \, \Lambda ^{2 - p} \, r \, \Longleftrightarrow \, d t = - B_0 \, \Lambda ^{2 - p} \, r \, d r. } \end{aligned}$$

Here the exponent of power of \(\left( t_0 - t\right) \) in (5.32) is estimated as

$$\begin{aligned}&\displaystyle { - 1 + \frac{2 (p - 1) (B_0 - 2)}{B_0 (p - 2)}> 0 \Longleftrightarrow B_0> \frac{4 (p - 1)}{p} } \\&\quad \displaystyle { \Longleftarrow B_0> \frac{6 p - 4}{p + 2}; \quad \frac{6 p - 4}{p + 2}> \frac{4 (p - 1)}{p} \, \Longleftrightarrow \, (p - 2)^2 > 0; } \\&\qquad \displaystyle { - 1 + \frac{2 (p - 1) (B_0 - 2)}{B_0 (p - 2)} - \frac{\delta (B_0 - 2)}{B_0 (p - 2)} \ge 0 \, \Longleftrightarrow \, 0 < \delta \le 2 (p - 1) - \frac{B_0 (p - 2)}{B_0 - 2} } \end{aligned}$$

and then,

$$\begin{aligned}&\displaystyle { t_0 - \left( \Lambda (\rho )\right) ^{2 - p} \, \rho ^2 \le t \le t_0 - \left( \Lambda (r)\right) ^{2 - p} \, r^2 \, \Longleftrightarrow \, r^{B_0} \le t_0 - t \le \rho ^{B_0}, }\\&\displaystyle { \left( t_0 - t \right) ^{ - 1 + \frac{2 (p - 1) (B_0 - 2)}{B_0 (p - 2)} - \frac{\delta (B_0 - 2)}{B_0 (p - 2)} } \le \rho ^{ B_0 \left( - 1 + \frac{2 (p - 1) (B_0 - 2)}{B_0 (p - 2)} - \frac{\delta (B_0 - 2)}{B_0 (p - 2)} \right) }\le 1 } \end{aligned}$$

and thus, the right hand side of (5.32) is bounded above by

$$\begin{aligned} \displaystyle { \frac{1}{B_0} \, \int \limits ^{t_0 - r^{B_0}} _{t_0 - \rho ^{B_0}} \, \Vert {\mathcal{C} (t)}^{q - 2} \, {\bar{e}} (u (t))^{\frac{2 (p - 1)}{p}} \Vert _{ L^\infty \left( \text{ supp } \, (\mathcal{B} (t))\right) } \, d t. } \end{aligned}$$

\(U_2\) of (5.30). \(U_2\) is given by the approximation term, the derivative of penalty term in (2.1) and our task is to control \(U_2\) well in the appropriate way. \(U_2\) is evaluated by use of the Bochner type estimate for the penalty term

$$\begin{aligned} \displaystyle { \partial _s g - \text{ div } \big ( (p \, f)^{1 - \frac{2}{p}} \, D g \big ) + \, C_0^\prime \, |D_v g|^2 \le C \, (\Lambda r)^2 \, {\bar{e}}^2 } \end{aligned}$$
(5.33)

with positive constants \(C_0^\prime \) and C depending only on p, \({\mathcal{M}}\) and \({\mathcal{N}}\). The derivation of (5.33) is done similarly as in “Appendix C”, under the scaling settings (5.5) and (5.6), by using (6.8) below.

Let \(\bar{r}\) be as \(r \le {\bar{r}} < \rho \) and chosen later. In the following we will replace r by \(\bar{r}\), \(r \le {\bar{r}} < \rho \), and proceed to the similar estimations as for \(U_2\).

Multiplying a test function \(\mathcal{B} \, \mathcal{C}^q\) in (5.33) and then, integrating the resulting inequality on y in \(\{s = -1\} \times \mathrm{I}\!\mathrm{R}^m\) and on r in a interval \(({\bar{r}}, \, \rho )\), the estimation for \(U_2\) is done as

$$\begin{aligned} \displaystyle { - \left( C_0^\prime - \frac{c}{2}\right) \, U_2 }&= \displaystyle { \left( C_0^\prime - \frac{c}{2}\right) \, \int _{\bar{r}}^{\rho } \frac{1}{r^{3 + \delta } \, \Lambda ^2} \, \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } |D_v g|^2 \, \mathcal{B} \, \mathcal{C}^q \, d y \, d r } \nonumber \\&\le \displaystyle { - \, \int _{\bar{r}}^{\rho } \frac{1}{r^{3 + \delta } \, \Lambda ^2} \, \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } \partial _s g \, \mathcal{B} \, \mathcal{C}^q \, d y \, d r } \nonumber \\&\qquad \displaystyle { + \, \int _{\bar{r}}^{\rho } \frac{1}{r^{3 + \delta } \, \Lambda ^2} \, \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } \frac{1}{2 c} \, \left( (p \, f)^{2 - \frac{2}{p}} \, \left( \mathcal{B}^\prime \, \mathcal{C}^q + \mathcal{B} \, q^2 \mathcal{C}^{q - 2} |D \mathcal{C}|^2 \right) \right. } \nonumber \\&\quad \displaystyle { \left. + \, C \, (\Lambda r)^2 \, {\bar{e}}^2 \, \mathcal{B} \, \mathcal{C}^q \right) \, d y \, d r } \nonumber \\&= : U_{21} + U_{22}, \end{aligned}$$
(5.34)

where by Cauchy’s inequality with a small \(c > 0\), the integrand term in the 3rd line is obtained from

$$\begin{aligned} \begin{aligned} \displaystyle { \left| (p \, f)^{1 - \frac{2}{p}} \, D g \cdot D \big (\mathcal{B} \, \mathcal{C}^q\big ) \right| }&= \displaystyle { (p \, f)^{1 - \frac{1}{p}} \, |D_v g| \left( |D \mathcal{B}| \mathcal{C}^q + \mathcal{B} |D \mathcal{C}^q| \right) . } \\&\le \displaystyle { \frac{c}{2} \, |D_v g|^2 \, \mathcal{B} \, \mathcal{C}^q + \frac{1}{2 c} \, (p \, f)^{\frac{2 (p - 1)}{p}} \, \left( \mathcal{B}^\prime \, \mathcal{C}^q + \mathcal{B} \, q^2 \mathcal{C}^{q - 2} |D \mathcal{C}|^2 \right) ; } \\ \displaystyle { \mathcal{B}^\prime }&: = \displaystyle { |y|^{\frac{2}{p - 1}} \left( 1 - \, |y|^{\frac{p}{p - 1}} \right) _+^{\frac{3 - p}{p - 2}}. } \end{aligned} \end{aligned}$$

Each term in the right hand side of (5.34) is separately treated in the following.

\(U_{22}\) in (5.34) The 3rd and 4th lines in the right hand side of (5.34) , \(U_{22}\), is scaled back and

$$\begin{aligned} U_{22}\le & {} C \, \int _{\bar{r}}^{\rho } \frac{1}{r^{3 + \delta } \, \Lambda ^{2 p}} \, \Vert \mathcal{C} (t)^{q - 2} \big ( \epsilon + |D u (t)|^2 \big )^{p - 1} + \nonumber \\&\qquad \left. + \, \mathcal{C} (t)^q \, r^2 \, {\bar{e}} (u (t))^2 \Vert _{ L^\infty (\mathrm{supp} (\mathcal{B} (t))) } \right| _{ t = t_0 - r^{B_0} } \, d r \, \times \nonumber \\&\quad \times \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } \left( \mathcal{B}^\prime \, \mathcal{C}^2 + \mathcal{B} \, |D \mathcal{C}|^2 + \mathcal{B} \right) \, d y, \end{aligned}$$
(5.35)

where in the 2nd line we compute as

$$\begin{aligned} (\Lambda r)^2 \, \Lambda ^{- 2} = r^2 \end{aligned}$$

and, the integral on y in the 3rd line is bounded by a constant as before, since

$$\begin{aligned} \displaystyle { \mathcal{C} + |D \mathcal{C}| \le 2; \quad \int \limits _{\mathrm{I}\!\mathrm{R}^m} \, ( \mathcal{B} + \mathcal{B}^\prime ) \, d y < \infty , \quad \frac{3 - p}{p - 2}> - 1 \Longleftrightarrow 3 > 2. } \end{aligned}$$

The integral on r in the 1st and 2nd lines is transformed into that on time by changing a variable \(t = t_0 - \Lambda (r)^{2 - p} r^2 = t_0 - r^{B_0}\)

$$\begin{aligned}&\displaystyle { C \, \int \limits _{t_0 - \rho ^{B_0}}^{t_0 - {\bar{r}}^{B_0}} (t_0 - t)^{ \frac{1}{B_0} \left( - 4 - \delta + \frac{(p + 2) (B_0 - 2)}{p - 2} \right) } \Vert \mathcal{C} (t)^{q - 2} \big ( {\bar{e}} (u (t))^{\frac{2 (p - 1)}{p}} + {\bar{e}} (u (t))^2 \big ) \Vert _{ L^\infty (\mathrm{supp} (\mathcal{B} (t)) } \, d t }\\&\quad \displaystyle { \le \, C \, \int \limits _{t_0 - \rho ^{B_0}}^{t_0 - {\bar{r}}^{B_0}} \Vert \mathcal{C} (t)^{q - 2} \big ( {\bar{e}} (u (t))^{\frac{2 (p - 1)}{p}} + \, {\bar{e}} (u (t))^2 \big ) \Vert _{ L^\infty (\mathrm{supp} (\mathcal{B} (t)) } \, d t, } \end{aligned}$$

where the power exponents of scale radius are computed as

$$\begin{aligned}&\displaystyle { r^2 \le 1 \, \Longleftarrow 0< r \le \rho \le 1; }\\&\displaystyle { \frac{ \Lambda ^{p - 2} \, r^{- 1} }{ r^{3 + \delta } \, \Lambda ^{2 p} } = r^{- 4 - \delta + \frac{(p + 2) (B_0 - 2)}{p - 2}}; } \\&\displaystyle { - 4 - \delta + \frac{(p + 2) (B_0 - 2)}{p - 2} \ge 0 \Longleftrightarrow 0 < \delta \le - 4 + \frac{(p + 2) (B_0 - 2)}{p - 2} }\\&\displaystyle { \Longleftarrow - 4 + \frac{(p + 2) (B_0 - 2)}{p - 2}> 0 \Longleftrightarrow B_0 > \frac{6 p - 4}{p + 2}. } \end{aligned}$$

\(U_{21}\) in (5.34) \(- U_{21}\) is computed as

$$\begin{aligned} \displaystyle { - U_{21} }&= \int _{\bar{r}}^{\rho } \frac{1}{r^{3 + \delta } \, \Lambda ^2} \, \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } \partial _s \big (g \, \mathcal{C}^q\big ) \, \mathcal{B} \, d y \, d r \nonumber \\&\qquad - \int _{\bar{r}}^{\rho } \frac{1}{r^{3 + \delta } \, \Lambda ^2} \, \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } g \, \partial _s \mathcal{C}^q \, \mathcal{B} \, d y \, d r \nonumber \\&= : \displaystyle { U_{211} + U_{212} } \end{aligned}$$
(5.36)

Each term in (5.36) is separately estimated in the following.

\(U_{212}\) in (5.36) \(U_{212}\) is estimated by using

$$\begin{aligned}&\displaystyle { \partial _s \mathcal{C} = \chi _{ \left\{ |x_0 + r \, y|^{B_0} \le (t_0 + r^{B_0} \, s) \right\} } \, \frac{1}{B_0} (t_0 + r^{B_0} \, s)^{\frac{1}{B_0} - 1} r^{B_0} \le \frac{r}{B_0}; } \\&\displaystyle { r^{B_0} \le t_0/2, \quad s = - 1; } \\&\displaystyle { t_0 - r^{B_0} \ge t_0/2, \quad (t_0 - r^{B_0})^{\frac{1}{B_0} - 1} \le (t_0/2)^{\frac{1}{B_0} - 1} \le r^{1 - B_0}; } \\&\displaystyle { \frac{1}{r^{3 + \delta } \, \Lambda ^2} \, \left| g \, \partial _s \mathcal{C}^q \right| \le \frac{q}{B_0} \, \frac{1}{r^{2 + \delta } \, \Lambda ^2} \, \left| g\right| \, \mathcal{C}^{q - 1} \le \frac{q}{B_0} \, \frac{1}{r^{2 + \delta } \, \Lambda ^{2 + p}} \, \left| g (u)\right| \, \mathcal{C}^{q - 1} } \end{aligned}$$

and thus, by scaling back and a changing of variable \(t = t_0 - \Lambda (r)^{2 - p} r^2 = t_0 - r^{B_0}\),

$$\begin{aligned} \displaystyle { g (u (t)) }&: = \displaystyle { \frac{K}{2} \chi \left( { \text{ dist }}^2 \big (u (t, \, x_0 + (t_0 - t)^{1/B_0} y), \, {\mathcal{N}}\big ) \right) ; } \nonumber \\ \displaystyle { |U_{212}| }&\le \displaystyle { \frac{q}{2 B_0} \, \int _{\bar{r}}^{\rho } \left. \frac{1}{r^{2 + \delta } \, \Lambda ^{2 + p}} \, \int \limits _{ \mathrm{I}\!\mathrm{R}^m } g (u (t)) \, \mathcal{C}^{q - 1} (t) \, \mathcal{B} \, d y \right| _{ t = t_0 - r^{B_0} } \, d r } \nonumber \\&\le \displaystyle { \frac{q}{B_0} \, \int \limits _{t_0 - \rho ^{B_0}}^{t_0 - {\bar{r}}^{B_0}} (t_0 - t)^{\frac{1}{B_0} \left( - 2 - \delta + \frac{(p + 2) (B_0 - 2)}{p - 2}\right) } \, \Vert e (u (t)) \, \mathcal{C}^{q - 1} (t) \Vert _{L^\infty \big ( \mathrm{supp} ( \mathcal{B} (t) ) \big )} \, d t \, \times } \nonumber \\&\displaystyle { \times \int _{\mathrm{I}\!\mathrm{R}^m} \mathcal{B} \, d y } \nonumber \\&\le \displaystyle { C \, \int \limits _{t_0 - \rho ^{B_0}}^{t_0 - {\bar{r}}^{B_0}} \, \Vert \mathcal{C}^{q - 1} (t) \, {\bar{e}} (u (t)) \Vert _{L^\infty \big (\mathrm{supp} (\mathcal{B} (t))\big )} \, d t, } \end{aligned}$$
(5.37)

where \(\displaystyle { \mathcal{B} = \left. \mathcal{B} (s, \, y)\right| _{s = - 1} = \left( 1 - |y|^\frac{p}{p - 1}\right) _+^{\frac{p - 1}{p - 2}} }\) and the exponents are computed as

$$\begin{aligned}&\displaystyle { - 2 - \delta + \frac{(p + 2) (B_0 - 2)}{p - 2} \ge 0 \Longleftrightarrow 0 < \delta \le - 2 + \frac{(p + 2) (B_0 - 2)}{p - 2} }\\&\quad \displaystyle { \Longleftarrow - 2 + \frac{(p + 2) (B_0 - 2)}{p - 2}> 0 \Longleftrightarrow B_0> \frac{4 p}{p + 2} } \\&\quad \displaystyle { \Longleftarrow B_0> \frac{6 p - 4}{p + 2} \quad ; \quad \frac{6 p - 4}{p + 2}> \frac{4 p}{p + 2} \Longleftrightarrow p > 2. } \end{aligned}$$

\(U_{211}\) in (5.36) \(U_{211}\) is transformed into an integral on time by scaling back.

$$\begin{aligned} \displaystyle { U_{211} }= & {} \displaystyle { \int _{\bar{r}}^{\rho } \frac{1}{r^{3 + \delta } \, \Lambda ^2} \, \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } \partial _s \big (g \, \mathcal{C}^q\big ) \, \mathcal{B} \, d y \, d r } \nonumber \\= & {} \displaystyle { \int _{\bar{r}}^{\rho } \frac{1}{r^{3 + \delta } \, \Lambda ^2} \, \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } \, \Lambda ^{2 - p} \, r^2 \, \left. \frac{\bar{K}}{2} \partial _t {\widetilde{h}} (t, \, y, \, r)\right| _{t = t_0 + \Lambda ^{2 - p} r^2 s } \, \mathcal{B} \, d y \, d r, }\qquad \quad \end{aligned}$$
(5.38)

where we put

$$\begin{aligned} \displaystyle { {\widetilde{h}} (t, \, y, \, r) } : = \displaystyle { \frac{\bar{K}}{2} \chi \left( { \text{ dist }}^2 \big ( u (t, \, x_0 + r y), \, {\mathcal{N}} \big ) \right) \, \left( t^{\frac{1}{B_0}} - |x_0 + r y| \right) _+^q. } \end{aligned}$$

By changing a variable \(t = t_0 - \Lambda (r)^{2 - p} r^2 = t_0 - r^{B_0}\), we have

$$\begin{aligned}&\displaystyle { t = t_0 - r^{B_0} \Longleftrightarrow r = (t_0 - t)^{1/ B_0}; \quad {\bar{K}} = K \Lambda ^{- p} = K r^{ \frac{p (B_0 - 2)}{p - 2} } = K (t_0 - t)^{ \frac{p (B_0 - 2)}{B_0 (p - 2)} }; }\\&\displaystyle { d t = - B_0 r^{B_0 - 1} d r= - B_0 \Lambda ^{2 - p} r d r; }\\&\displaystyle { \frac{1}{r^{2 + \delta } \, \Lambda ^2} = (t_0 - t)^{- c_0}; \quad c_0 : = \frac{1}{B_0} \left( \frac{2 (p - B_0)}{p - 2} + \delta \right) } \end{aligned}$$

and an elementary computation

$$\begin{aligned} \begin{aligned} \displaystyle { h (t, \, y) }&: = \displaystyle { {\widetilde{h}} (t, \, y, \, (t_0 - t)^{1/B_0}) } \\&= \displaystyle { \frac{\bar{K}}{2} \chi \left( { \text{ dist }}^2 \big ( u (t, \, x_0 + (t_0 - t)^{1/B_0} y), \, {\mathcal{N}} \big ) \right) \, \left( t^{\frac{1}{B_0}} - |x_0 + (t_0 - t)^{1/B_0} y| \right) _+^q; }\\ \displaystyle { \partial _t h (t, \, y) }&= \displaystyle { - \frac{p (B_0 - 2)}{B_0 (p - 2)} (t_0 - t)^{- 1} h (t, \, y) - \frac{1}{B_0} (t_0 - t)^{- 1} y \cdot D_y h (t, \, y) }\\&\displaystyle { \qquad + \frac{\bar{K}}{2} \left. \partial _\tau {\widetilde{h}} (\tau , \, y, \, (t_0 - t)^{1/B_0}) \right| _{\tau = t}.} \end{aligned} \end{aligned}$$

Thus, we have

$$\begin{aligned} \displaystyle { U_{211} }&= \displaystyle { \frac{1}{B_0} \, \int \limits _{t_0 - \rho ^{B_0}}^{t_0 - {\bar{r}}^{B_0}} \, (t_0 - t)^{- c_0} \, \frac{d P}{d t} \, d t + \frac{p (B_0 - 2)}{B_0^2 (p - 2)} \, \int \limits _{t_0 - \rho ^{B_0}}^{t_0 - {\bar{r}}^{B_0}} \, (t_0 - t)^{- c_0 - 1} \, P (t) \, d t } \nonumber \\&\qquad \displaystyle { + \frac{1}{B_0^2} \, \int \limits _{t_0 - \rho ^{B_0}}^{t_0 - {\bar{r}}^{B_0}} \, (t_0 - t)^{- c_0 - 1} \, \int \limits _{ \mathrm{I}\!\mathrm{R}^m } y \cdot D_y h (t, \, y) \, \mathcal{B} (y) \, d y \, d t } \nonumber \\&= : U_{2111} + U_{2112} + U_{2113}, \end{aligned}$$
(5.39)

where we put

$$\begin{aligned}&\displaystyle { P (t) : = \int \limits _{ \mathrm{I}\!\mathrm{R}^m } h (t, \, y) \, \mathcal{B} (y) \, d y, \quad \mathcal{B} (y) = \left. \mathcal{B} (s, \, y)\right| _{s = - 1} = \left( 1 - |y|^\frac{p}{p - 1}\right) _+^{\frac{p - 1}{p - 2}}. } \end{aligned}$$

We will estimate each term in (5.39).

Now, we set \(\bar{r}\) as

$$\begin{aligned} \displaystyle { \exists \, {\bar{r}}, \, \, r \le {\bar{r}} \le \rho \quad : \quad \max _{t_0 - \rho ^{B_0} \le t \le t_0 - r^{B_0}} P (t) = P (t_0 - {\bar{r}}^{B_0}). } \end{aligned}$$
(5.40)

Then, by integration by parts in the integral on t, we have

$$\begin{aligned} \displaystyle { B_0 \times U_{2111} }= & {} \displaystyle { \int \limits _{t_0 - \rho ^{B_0}}^{t_0 - {\bar{r}}^{B_0}} \, (t_0 - t)^{- c_0} \, \frac{d P}{d t} \, d t } \nonumber \\= & {} \displaystyle { \left. (t_0 - t)^{- c_0} \, P (t) \right| _{t_0 - \rho ^{B_0}}^{t_0 - {\bar{r}}^{B_0}} - \int \limits _{t_0 - \rho ^{B_0}}^{t_0 - {\bar{r}}^{B_0}} \, P (t) \, c_0 \, (t_0 - t)^{- c_0 - 1} \, d t } \nonumber \\\ge & {} \displaystyle { {\bar{r}}^{- c_0 \, B_0} \, P (t_0 - {\bar{r}}^{B_0}) - {\rho }^{- c_0 \, B_0} \, P (t_0 - {\rho }^{B_0}) } \nonumber \\&\displaystyle { - P (t_0 - {\bar{r}}^{B_0}) \left( {\bar{r}}^{- c_0 \, B_0} - {\rho }^{- c_0 \, B_0} \right) } \nonumber \\= & {} \displaystyle { {\rho }^{- c_0 \, B_0} \left( P (t_0 - {\bar{r}}^{B_0}) - P (t_0 - {\rho }^{B_0}) \right) \ge 0. } \end{aligned}$$
(5.41)

Clearly, \(U_{2112} \ge 0\). By integration by parts in the integral on y, we also have

$$\begin{aligned} \displaystyle { U_{2113} }= & {} \displaystyle { \frac{1}{B_0^2} \, \int \limits _{t_0 - \rho ^{B_0}}^{t_0 - {\bar{r}}^{B_0}} \, (t_0 - t)^{- c_0 - 1} \, \int \limits _{ \mathrm{I}\!\mathrm{R}^m } h (t, \, y) \, \left( - m \, \mathcal{B} + \frac{p \big ( 1 - |y|^{\frac{p}{p - 1}} \big )_+^{\frac{1}{p - 2}} \, |y|^{\frac{p}{p - 1}} }{p - 2} \right) \, d y \, d t } \\\ge & {} \displaystyle { - \frac{m}{B_0^2} \, \int \limits _{t_0 - \rho ^{B_0}}^{t_0 - {\bar{r}}^{B_0}} \, (t_0 - t)^{- c_0 - 1} \, \int \limits _{ \mathrm{I}\!\mathrm{R}^m } h (t, \, y) \, \mathcal{C}^q (t) \, \mathcal{B} \, d y \, d t. } \end{aligned}$$

The last integral is estimated below as

$$\begin{aligned} \mathcal{C}^q (t)&: = \mathcal{C}^q (t, \, x_0 + (t_0 - t)^{1/B_0} y); \nonumber \\&\qquad \displaystyle { - \frac{m}{B_0^2} \, \int \limits _{t_0 - \rho ^{B_0}}^{t_0 - {\bar{r}}^{B_0}} \, (t_0 - t)^{b_0 - 1} \, \, \int \limits _{ \mathrm{I}\!\mathrm{R}^m } g (u (t)) \mathcal{C}^q (t) \, \mathcal{B} \, d y \, d t } \nonumber \\&\ge - \frac{m}{B_0^2} \, \int \limits _{ \mathrm{I}\!\mathrm{R}^m } \mathcal{B} \, d y \, \times \nonumber \\&\qquad \qquad \times \, \left( \, \int \limits _{t_0 - \rho ^{B_0}}^{t_0 - {\bar{r}}^{B_0}} \, (t_0 - t)^{\alpha _0 (b_0 - 1)} \, d t \right. \nonumber \\&\qquad \qquad \left. + \, \int \limits _{t_0 - \rho ^{B_0}}^{t_0 - {\bar{r}}^{B_0}} \left\| \frac{K}{2} \, \chi \big ( { \text{ dist }}^2 \big ( u (t, \, \cdot ), \, {\mathcal{N}} \big ) \big ) \, \mathcal{C}^q (t) \right\| _{L^\infty \big ( \text{ supp } \mathcal{B} (t) \big )}^{ \frac{\alpha _0}{\alpha _0 - 1} } \, d t \right) , \end{aligned}$$
(5.42)

where we use Young’s inequality with an exponent \(\alpha _0 > 1\) and compute as

$$\begin{aligned}&\displaystyle { t = t_0 - r^{B_0} \Longleftrightarrow r = (t_0 - t)^{1/B_0}; }\\&\displaystyle { \Lambda = r^{\frac{B_0 - 2}{2 - p}}, \quad h (u (t)) = \Lambda ^{- p} g (u (t)); }\\&\displaystyle { (t_0 - t)^{- c_0 - 1} \, \Lambda ^{- p} = (t_0 - t)^{b_0 - 1}; \quad b_0 : = \frac{B_0 (p + 2) - 4 p}{B_0 (p - 2)} - \frac{\delta }{B_0} } \end{aligned}$$

and we choose \(\alpha _0 > 1\) as

$$\begin{aligned}&\displaystyle { 1< \alpha _0< \frac{1}{1 - b_0} \Longleftarrow \alpha _0 (b_0 - 1)> - 1; \quad b_0< 1 \Longleftarrow B_0< p; } \nonumber \\&\displaystyle { \frac{1}{1 - b_0}> 1 \Longleftrightarrow b_0> 0 \Longleftrightarrow 0< \delta < \frac{B_0 (p + 2) - 4 p}{p - 2}; } \nonumber \\&\displaystyle { \frac{B_0 (p + 2) - 4 p}{p - 2}> 0 \Longleftrightarrow B_0> \frac{4 p}{p + 2} } \nonumber \\&\displaystyle { \Longleftarrow B_0> \frac{6 p - 4}{p + 2}; \frac{6 p - 4}{p + 2}> \frac{4 p}{p + 2} \Longleftrightarrow p > 2.} \end{aligned}$$
(5.43)

By (5.41) and (5.42) substituted into (5.39), we have

$$\begin{aligned} \displaystyle { U_{211} }\ge & {} \displaystyle { - \frac{m}{B_0^2} \int \limits _{ \mathrm{I}\!\mathrm{R}^m } \mathcal{B} \, d y \, \times } \nonumber \\&\displaystyle { \qquad \times \, \left( \frac{ \rho ^{\alpha _0 (b_0 - 1) + 1} - {\bar{r}}^{\alpha _0 (b_0 - 1) + 1} }{\alpha _0 (b_0 - 1) + 1} + \int \limits _{t_0 - \rho ^{B_0}}^{t_0 - {\bar{r}}^{B_0}} \Vert {\bar{e}} (u (t)) \, \mathcal{C}^q (t) \Vert _{L^\infty \big ( \text{ supp } \mathcal{B} (t) \big )}^{ \frac{\alpha _0}{\alpha _0 - 1} } \right) , } \nonumber \\ \end{aligned}$$
(5.44)

yielding, with (5.36) and (5.37) for \(U_{212}\),

$$\begin{aligned} \displaystyle { - U_{21} }\ge & {} \displaystyle { U_{211} +U_{212} }\\\ge & {} \displaystyle { - \, C \, \left( \rho ^{\alpha _0 (b_0 - 1) + 1} - {\bar{r}}^{\alpha _0 (b_0 - 1) + 1} \right) } \\&\displaystyle { - \, C \, \int \limits _{t_0 - \rho ^{B_0}}^{t_0 - {\bar{r}}^{B_0}} \Vert {\bar{e}} (u (t)) \, \mathcal{C}^q (t) \Vert _{L^\infty \big ( \text{ supp } \mathcal{B} (t) \big )}^{ \frac{\alpha _0}{\alpha _0 - 1} } } \\&\displaystyle { - \, C \, \int \limits _{t_0 - \rho ^{B_0}}^{t_0 - {\bar{r}}^{B_0}} \, \Vert \mathcal{C}^{q - 1} (t) \, {\bar{e}} (u (t)) \Vert _{L^\infty \big (\mathrm{supp} (\mathcal{B} (t))\big )} \, d t. } \end{aligned}$$

By definition of P(t) in (5.39) and \(\bar{r}\) in (5.40), P(t) is the local scaled integral of the penalty term, because by changing a variable \(x = x_0 + (t_0 - t)^{1/B_0} y\)

$$\begin{aligned}&\displaystyle { h (t, \, y) = \frac{\bar{K}}{2} \chi \left( { \text{ dist }}^2 \big (u (t, \, x_0 + (t_0 - t)^{1/B_0} y), \, {\mathcal{N}}\big ) \right) \, \left( t^{1/B_0} - |x_0 + (t_0 - t)^{1/B_0} y| \right) _+^q; }\\&\displaystyle { P (t) = \int \limits _{ \mathrm{I}\!\mathrm{R}^m } h (t, \, y) \, \mathcal{C}^q (t) \, \mathcal{B} \, d y = \frac{1}{\Lambda ^p} \int \limits _{ \mathrm{I}\!\mathrm{R}^m } \frac{K}{2} \chi \big ( { \text{ dist }}^2 (u (t, \, x), \, {\mathcal{N}}) \big ) \, \mathcal{C}^q (t, \, x) \, \mathcal{B} (t, \, x) \, d x } \end{aligned}$$

and it holds that

$$\begin{aligned} \displaystyle { P (t_0 - r^{B_0}) \le P (t_0 - {\bar{r}}^{B_0}) \quad \text{ for } {\bar{r}} \text{ in } (5.40)}. \end{aligned}$$
(5.45)

Collecting the estimations for \(U_1\), \(U_2\) and \(U_3\) above in (5.30), we have, for \({\bar{r}}\) in (5.40),

$$\begin{aligned} \displaystyle { E (\rho ) - E ({\bar{r}}) }\ge & {} \displaystyle { - \, C \, \left( \rho ^\mu - {\bar{r}}^\mu \right) } \nonumber \\&- \, C \, \int \limits _{t_0 - \rho ^{B_0}}^{t_0 - {\bar{r}}^{B_0}} \Vert \mathcal{C}^{q - 2} (t) \, {\bar{e}} (u (t))^{\frac{2 (p - 1)}{p}} \nonumber \\&+ \mathcal{C}^q (t) \, \big ( {\bar{e}} (u (t))^2 + {\bar{e}} (u (t))^{ \frac{\alpha _0}{\alpha _0 - 1} } \big ) \Vert _{L^\infty \left( \text{ supp } (\mathcal{B} (t)) \right) } \, d t. \end{aligned}$$
(5.46)

Let us put, for \(\alpha _0 > 1\) in (5.43),

$$\begin{aligned} \theta _0 = \max \left\{ 2, \, \frac{\alpha _0}{\alpha _0 - 1}\right\} . \end{aligned}$$
(5.47)

From (5.46), our desired monotonicity estimate is shown to hold true in the range of scale radius \([{\bar{r}}, \, \rho ]\). Also (5.45) is the monotonicity estimate in the range \([r, \, {\bar{r}}]\) of the local scaled integral of the penalty term. Therefore, it remains to estimate the local scaled p-energy in the range of scale radius \([r, \, {\bar{r}}]\).

Step 3 : Monotonicity of the scaled p-energy. We now show a monotonicity estimate for the scaled p-energy without the penalty term. Under the same notation as before we denote the scaled p-energy by

$$\begin{aligned} \displaystyle { \quad F (r) = \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } f (v (s, \, y)) \, \mathcal{B} (s, \, y) \, \mathcal{C}^q (s, \, y) \, d y ; \quad f = f (v) : = \frac{1}{p} \big ({\bar{\epsilon }} + |D v|^2\big )^{\frac{p}{2}} } \nonumber \\ \end{aligned}$$
(5.48)

and compute the differentiation of F(r) on a scale radius r

(5.49)

Clearly, \(H_2 \ge 0\). \(H_3\) is similarly estimated as III of (5.9) in (5.10), and

$$\begin{aligned} \displaystyle { H_1 }= & {} \displaystyle { \frac{1}{r} \, \left( 1 + \frac{r \, \Lambda ^\prime }{\Lambda } \right) \, \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } \left\{ \text{ div } \big ( (p \, f)^{1 - \frac{2}{p}} D v \, \mathcal{B} \, \mathcal{C}^{q/2} \big ) \cdot \big ( v \, \mathcal{C}^{q/2} - {\bar{v}} \big ) \right. } \nonumber \\&\qquad \displaystyle { \left. \, + \, (p \, f)^{1 - \frac{2}{p}} D v \cdot \Big ( v \, D \mathcal{C}^{q/2} \Big ) \, \mathcal{B} \, \mathcal{C}^{q/2} \right\} \, \, d y } \nonumber \\&\quad \displaystyle { - \, \frac{1}{r} \, \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } \text{ div } \big ( (p \, f)^{1 - \frac{2}{p}} D v \, \mathcal{B} \, \mathcal{C}^q \big ) \cdot \left( \Big ( \frac{(2 - p) \, r \, \Lambda ^\prime }{\Lambda } + 2 \Big ) \, s \, \partial _s v + y \cdot D v \right) \, d y } \nonumber \\= & {} \displaystyle { : H_{11} + H_{12} + H_{13}, } \end{aligned}$$
(5.50)

where we use an integration by parts and the dilation derivative (5.12).

Estimation of \(H_{11}\) We have

In the bracket of the right hand side, the 1st time-derivative term, the 3rd term and the 4th term are the same ones as in \(I_{111}\), \(I_{112}\) and \(I_{113}\) of \(I_{11}\) in (5.13), respectively. These terms are estimated as for \(I_{111}\), \(I_{112}\) and \(I_{113}\).

The 2nd term containing the derivative of penalty term is estimated in the following.

$$\begin{aligned}&\displaystyle { \frac{1}{r} \Big | \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } C_0 \, \big ( (v \, \mathcal{C}^{q/2} - {\bar{v}} \big ) \cdot D_v g \, \mathcal{B} \, \mathcal{C}^{q/2} \, d y \Big | } \\&\quad \displaystyle { \le \frac{(C_0)^2}{2 r} \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } \big | (v \, \mathcal{C}^{q/2}) - {\bar{v}} \big |^2 \, \mathcal{B} \, d y + \frac{1}{2 r} \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } \big | D_v g \big |^2 \, \mathcal{B} \, \mathcal{C}^q \, d y. } \end{aligned}$$

The 1st term is the same as \((- 1) \times \)2nd one in (5.16) and thus, estimated above by \((- 1) \times \)the right hand side of (5.17). The 2nd term is estimated in the following. Multiplying a test function \(\mathcal{B} \, \mathcal{C}^q\) by (5.33), we have, by Cauchy’s inequality with a small \(c > 0\),

$$\begin{aligned} \displaystyle { (C_0^\prime - \frac{1}{c}) \, |D_v g (v)|^2 \, \mathcal{B} \, \mathcal{C}^q }\le & {} \displaystyle { \frac{c}{2} \, |\partial _s v|^2 \, \mathcal{B} \, \mathcal{C}^q + \text{ div } \big ( (p \, f)^{1 - \frac{2}{p}} \, D g \, \mathcal{B} \, \mathcal{C}^q \big ) } \nonumber \\&\displaystyle { + \, \frac{c}{2} \, (p \, f)^{2 - \frac{2}{p}} \, \left( \mathcal{B}^\prime \, \mathcal{C}^q + \mathcal{B} \, \mathcal{C}^{q - 2} |D \mathcal{C}|^2 \right) } \nonumber \\&\displaystyle { + \, C \, (\Lambda r)^2 \, {\bar{e}} (v)^2 \, \mathcal{B} \, \mathcal{C}^q, } \end{aligned}$$
(5.51)

where \(\displaystyle { \mathcal{B}^\prime = |y|^{\frac{2}{p - 1}} \left( 1 - \, |y|^{\frac{p}{p - 1}} \right) _+^{\frac{3 - p}{p - 2}} }\), and

$$\begin{aligned} \displaystyle { \partial _s g (v) = \partial _s v \cdot D_v g (v); \quad D g (v) = D v \cdot D_v g (v). } \end{aligned}$$

The inequality (5.51) is integrated on y and then, estimated by integration by parts as

$$\begin{aligned}&\displaystyle { \frac{1}{r} \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } |D_v g|^2 \, \mathcal{B} \, \mathcal{C}^q \, d y \le \frac{c}{2 r} \, \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } |\partial _s v|^2 \, \mathcal{B} \, \mathcal{C}^q \, d y } \nonumber \\&\quad \displaystyle { + \, \frac{C}{r} \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } \left( (\Lambda r)^2 \, {\bar{e}}^2 \, \mathcal{B} \, \mathcal{C}^q \, + \, (p \, f)^{2 - \frac{2}{p}} \left( \mathcal{B}^\prime \, \mathcal{C}^q + \mathcal{B} \, q^2 \mathcal{C}^{q - 2} |D \mathcal{C}|^2 \right) \right) \, d y. } \nonumber \\ \end{aligned}$$
(5.52)

Estimations of \(H_{12}\) and \(H_{13}\) \(H_{12}\) is the same as \(I_2\) in (5.11) and thus, estimated as in (5.22) and (5.20).

\(H_{13}\) is the same as \(I_3\) in (5.11) except the derivative term of the penalty term and thus, is estimated similarly as for \(I_{3i}\), \(i = 1, \ldots , 5\), and the estimation (5.52) for the derivative of the penalty term in \(H_{11}\).

Gathering the estimations above and scaling back, we have, for \(\delta > 0\),

$$\begin{aligned} \displaystyle { \frac{d}{d r} F(r) }\ge & {} \displaystyle { J - C \, \left( r^{- 1 + \frac{\delta }{p - 2}} + r^{- 1+ \delta } \right) }\\&\displaystyle { - \left. \frac{C}{r \, \Lambda ^{2 (p - 1)}} \, \Vert {\bar{e}} (u (t))^{\frac{2 (p - 1)}{p}} \, \mathcal{C}^{q - 2} (t) + {\bar{e}} (u (t))^2 \, \mathcal{C}^q (t) \Vert _{L^\infty \left( \text{ supp } \, (\mathcal{B} (t) \right) } \right| _{t = t_0 - r^{B_0}}, } \end{aligned}$$

where in the 2nd line we estimate as

$$\begin{aligned} \displaystyle { (\Lambda r)^2 \Lambda ^{- 2} = r^2 \le 1 \Longleftarrow 0< r \le \rho \le 1; \quad \int \limits _{ \{s = - 1\} \times \mathrm{I}\!\mathrm{R}^m } \left( \mathcal{B} + \mathcal{B}^\prime \, \mathcal{C}^2 + \mathcal{B} \, |D \mathcal{C}|^2 \right) \, d y < \infty , } \end{aligned}$$

and, integrated on r in \((r, \, \rho )\), yielding

$$\begin{aligned} \displaystyle { F (\rho ) - F (r) }\ge & {} \displaystyle { - \, C \, \int _r^\rho \left( r^{- 1 + \frac{\delta }{p - 2}} + r^{- 1+ \delta } \right) \, d r } \nonumber \\&- \int _r^\rho \frac{C}{r \, \Lambda ^{2 (p - 1)}} \, \Vert {\bar{e}} (u (t))^{\frac{2 (p - 1)}{p}} \, \mathcal{C}^{q - 2} (t)+ \nonumber \\&\left. + \,{\bar{e}} (u (t))^2 \, \mathcal{C}^q (t) \Vert _{L^\infty \left( \text{ supp } \, (\mathcal{B} (t) \right) } \right| _{t = t_0 - r^{B_0}} \, d r. \end{aligned}$$
(5.53)

The 1st term in the right hand side is the same as \(U_1\) in (5.30) and estimated as in (5.31). The term in the 2nd and 3rd lines is, by changing a variable \(t = t_0 - \Lambda ^{2 - p} r^2 = t_0 - r^{B_0}\), computed as

(5.54)

where the power exponent of scale radius is evaluated as

$$\begin{aligned}&\displaystyle { r^{- 1} \frac{1}{\Lambda ^{2 (p - 1)}} \Lambda ^{p - 2} r^{- 1} = (t_0 - t)^{ \frac{1}{B_0} \left( - 2 + \frac{p (B_0 - 2)}{p - 2} \right) } \, \Longleftarrow \, t = t_0 - r^{B_0}; }\\&\displaystyle { - 2 + \frac{p (B_0 - 2)}{p - 2} \ge 0 \Longleftrightarrow B_0> \frac{4 (p - 1)}{p} } \\&\displaystyle { \Longleftarrow B_0> \frac{6 p - 4}{p + 2}; \quad \frac{6 p - 4}{p + 2}> \frac{4 (p - 1)}{p} \Longleftrightarrow (p - 2)^2 > 0. } \end{aligned}$$

Finally, we collect the estimations (5.45), (5.46) in Step 2, and (5.53), (5.54) in Step 3 to complete the proof of (5.4). \(\square \)

Now we show the validity of the forward monotonicity estimate, Lemma 13.

As before by parallel transformation let the Eq. (2.1) and its solutions u be defined on \((0, \, \infty ) \times \mathrm{I}\!\mathrm{R}^m\) with the same notation.

Let \((t_0, \, x_0)\) in the parabolic like envelope \(\displaystyle { \left\{ (t, \, x) \, : \, \min \{1, \, (R_{\mathcal{M}})^{B_0}\} > t \ge |x|^{B_0} \right\} }\), \(\displaystyle { B_0 > 2 }\).

The forward localized scaled penalized energy is

$$\begin{aligned}&\displaystyle { E (r) = \frac{1}{\Lambda ^p} \int _{\{t = t_0 + \Lambda ^{2 - p} \, r^2\} \times B (R_{\mathcal{M}})} \frac{1}{p} {\bar{e}} (u (t, \, x)) \, \mathcal{B} (t_0, x_0 ; \, t, x) \, \mathcal{C}^q (t, \, x) \, d x; } \qquad \end{aligned}$$
(5.55)
$$\begin{aligned}&\displaystyle { \Lambda = \Lambda (r) = r^{ \frac{B_0 - 2}{2 - p} }; \quad p> B_0 > \frac{6 p - 4}{p + 2}; \quad 0 < r \le \min \{1, \, (R_{\mathcal{M}})^{1/ B_0}\} } \end{aligned}$$
(5.56)

with weight

$$\begin{aligned}&\displaystyle { \mathcal{B} (t_0, \, x_0 ; \, t, \, x) = \frac{1}{(t - t_0)^{\frac{m}{B_0}}} \left( 1 - \, \left( \frac{|x - x_0|}{(t - t_0)^{\frac{1}{B_0}}} \right) ^{\frac{p}{p - 1}} \right) ^{\frac{p - 1}{p - 2}}_{+}, \qquad t> t_0; } \nonumber \\&\displaystyle { \mathcal{C} (t, \, x) = \left( t^{1/B_0} - |x| \right) _+; \quad q > 2. } \end{aligned}$$
(5.57)

The notation as above is used.

Lemma 20

Let \(p > 2\) and \(q > 2\). For any regular solution u to (2.1) the following estimate is valid for any positive number \(r < \rho \le \min \{1, \, \big ((R_{\mathcal{M}})^{B_0} - t_0\big )^{1/B_0}\}\)

$$\begin{aligned} E (\rho )\le & {} \, E (r) + \, C \, \left( \rho ^\mu - r^\mu \right) \nonumber \\&\displaystyle { + \, C \,\int \limits _{t_0 + r^{B_0}} ^{t_0 + \rho ^{B_0}} \Vert \mathcal{C}^{q - 2} (t) \, ({\bar{e}} (u (t)))^{\theta _0} \Vert _{ L^\infty \left( B ( (t_0 - t)^{1/B_0}, \, x_0 ) \right) } \, d t, } \end{aligned}$$
(5.58)

where

$$\begin{aligned} \displaystyle { \Lambda = \Lambda (r) = r^{\frac{B_0 - 2}{2 - p}}, \quad (\Lambda (r))^{2 - p} \, r^2 = r^{B_0} } \end{aligned}$$

and the positive exponents \(\theta _0 \ge 2\) and \(\mu \) depend only on \(B_0\), p and \({\mathcal{N}}\), m, p and \(B_0\), respectively, and the positive constant C depends only on the same ones as \(\mu \) and q.

Proof of Lemma 20

As before we put

$$\begin{aligned} \displaystyle { \Lambda = r^{ \frac{B_0 - 2}{2 - p} }, \quad p> B_0 > \frac{6 p - 4}{p + 2} } \end{aligned}$$

and let r any positive number in the range \(0 < r \le \min \{1, \, \big ((R_{\mathcal{M}})^{B_0} - t_0\big )^{1/B_0}\}\). We make a scaling transformation intrinsic to the evolutionary p-Laplace operator

$$\begin{aligned} \displaystyle { t = t_0 + \Lambda ^{2 - p} r^2 \, s; \quad x = x_0 + r \, y; \quad v (s, \, y) = \frac{ u (t_0 + \Lambda ^{2 - p} r^2 \, s, \, x_0 + r \, y) }{\Lambda \, r} } \qquad \end{aligned}$$
(5.59)

and, under the scaling transformation it holds that

$$\begin{aligned} \displaystyle { t = t_0 + \Lambda ^{2 - p} r^2 \, \Longleftrightarrow \, s = + 1. } \end{aligned}$$

The scaled solution v is a solution of the scaled equation on \(\{s = 1\} \times \mathrm{I}\!\mathrm{R}^m\)

$$\begin{aligned} \displaystyle { \partial _s v - \text{ div } \left( \big (\Lambda ^{- 2} \epsilon + |D v|^2\big )^{\frac{p - 2}{2}} D v \right) = - C_0 \, \frac{K/\Lambda ^p}{2} \frac{d}{d v} \chi \left( \text{ dist }^2 \big (\Lambda \, r \, v, \, {\mathcal{N}}\big )\right) . } \qquad \end{aligned}$$
(5.60)

Hereafter we use the same notation as in (5.6).

Similarly as the backward case, the scaled energy is rewritten as

$$\begin{aligned}&\displaystyle { E (r) = \int \limits _{ \{s = 1\} \times \mathrm{I}\!\mathrm{R}^m } {\bar{e}} ((v (s, \, y)) \, \mathcal{B} (s, \, y) \, \mathcal{C}^q (s, \, y) \, d y; } \nonumber \\&\displaystyle { \mathcal{B} (s, \, y) = \frac{1}{s^{m/B_0}} \left( 1 - \, \left( \frac{|y|}{s^{1/B_0}} \right) ^{ \frac{p}{p - 1} } \right) _+^{ \frac{p - 1}{p - 2} }; \,\, \mathcal{C} (s, \, y) = \left( (t_0 + r^{B_0} \, s)^{1/B_0} - |x_0 + r \, y| \right) _+, }\nonumber \\ \end{aligned}$$
(5.61)

where the integral in (5.61) is well-defined by \(\text{ supp } (\mathcal{C})\) and \(\text{ supp } (\mathcal{B})\).

The computation and estimation are similar as in those for the backward monotonicity estimate. In the following we indicate only the part of estimations, different from the backward monotonicity. In the following the integral region on y is changed to \(\{s = 1\} \times \mathrm{I}\!\mathrm{R}^m\).

Similarly as in (5.9) in the backward case, we make differentiation of E(r) on r

$$\begin{aligned} \displaystyle { \frac{d}{d r} E (r) }= & {} \displaystyle { \int \limits _{ \{s = 1\} \times \mathrm{I}\!\mathrm{R}^m } \, \left( (p \, f)^{1 - \frac{2}{p}} D v \cdot \frac{d}{d r} D v + C_0 \, \frac{d v}{d r} \cdot D_v g (v) \right) \, \mathcal{B} \, \mathcal{C}^q d y } \nonumber \\&\displaystyle { + \, \frac{B - 2}{r \, (p - 2)} \int \limits _{ \{s = 1\} \times \mathrm{I}\!\mathrm{R}^m } \, \left( \bar{\epsilon } \, (p \, f)^{1 - \frac{2}{p}} + p \, C_0 \, g (v) \right) \, \mathcal{B} \, \mathcal{C}^q \, d y } \nonumber \\&\displaystyle { \qquad + \int \limits _{ \{s = 1\} \times \mathrm{I}\!\mathrm{R}^m } \, {\bar{e}} (v) \, \mathcal{B} \, \frac{d}{d r} \mathcal{C}^q \, d y } \nonumber \\=: & {} \displaystyle { I + II + III. } \end{aligned}$$
(5.62)

Estimation of II and III. By Young’s inequality and \(0< \bar{\epsilon } = \Lambda ^{- 2} \epsilon \le 1\) the term II is bounded with \(\delta > 0\) by

$$\begin{aligned} \displaystyle { II \le \frac{C}{r^{1 + \delta }} \int \limits _{\{s = 1\} \times \mathrm{I}\!\mathrm{R}^m} \left( {\bar{e}} (v)\right) ^{\frac{2 (p - 1)}{p}} \, \mathcal{C}^q \, \mathcal{B} \, d y + C \, \left( \frac{r^{\frac{\delta (p - 2)}{p}}}{r} + \frac{r^{\frac{\delta p}{p - 2}}}{r} \right) \int \limits _{\{s = 1\} \times \mathrm{I}\!\mathrm{R}^m} \mathcal{C}^q \, \mathcal{B} \, d y. } \end{aligned}$$

For estimation of III the derivative of \(\mathcal{C}\) on r is computed as

$$\begin{aligned} \displaystyle { \left| \frac{d}{d r} \mathcal{C} (t, \, x) \right| = \chi _{ \{ |x_0 + r y| \le (t_0 + r^{B_0} s)^{1/B_0} \} } \left| (t_0 + r^{B_0} s)^{1/B_0} \frac{ r^{B_0 - 1} s }{ t_0 + r^{B_0} s } - \frac{x_0 + r y}{|x_0 + r y|} \cdot y \right| } \end{aligned}$$

and thus, on the support \(\{y \in \mathrm{I}\!\mathrm{R}^m \, : \, |y| < 1\}\) of \(\mathcal{B} (1, \, y)\)

$$\begin{aligned} \displaystyle { \left. \left| \frac{d}{d r} \mathcal{C} \right| \right| _{s = 1} \le 3 \, \chi _{ \{ |x_0 + r y| \le (t_0 - r^{B_0})^{1/B_0} \} } \, r^{- 1} } \end{aligned}$$

because of the conditions

$$\begin{aligned} \displaystyle { 0 < t_0 \le 1; \quad \frac{r^{B_0}}{t_0 + r^{B_0}} \le 1. } \end{aligned}$$

Thus, exactly as (5.10) in the backward case, we have

$$\begin{aligned} \displaystyle { III \le \frac{C}{r^{1 + \delta }} \, \int \limits _{ \{s = 1\} \times \mathrm{I}\!\mathrm{R}^m } \, ( \bar{e} (v) )^{\frac{2 (p - 1)}{p}} \, \mathcal{C}^q \, \mathcal{B} \, d y \, + \, \frac{C \, r^{\frac{\delta p}{p - 2}}}{r} \int \limits _{ \{s = 1\} \times \mathrm{I}\!\mathrm{R}^m } \mathcal{C}^{q - \frac{2 (p - 1)}{p - 2}} \, \mathcal{B} \, d y. } \end{aligned}$$

The estimation of I is exactly same as (5.11) in the backward case. The terms corresponding to \(I_1\) are bounded above by \((-1) \times \)the terms (5.20), (5.21) and some controllable integral terms containing \(\mathcal{B}\), \(\mathcal{C}\) and their derivatives, where the integral region is replaced by \(\{s = 1\} \times \mathrm{I}\!\mathrm{R}^m\). The term corresponding to \(I_2\) is estimated above by \((- 1) \times \)the right hand side of (5.22) with the integral region replaced by \(\{s = 1\} \times \mathrm{I}\!\mathrm{R}^m\).

\(I_3\) is computed exactly as (5.23) and (5.24) with the integral region replaced by \(\{s = 1\} \times \mathrm{I}\!\mathrm{R}^m\). In the term corresponding to \(I_{31}\) we note by \(s = 1\) that

$$\begin{aligned} \displaystyle { (- s) \, \left( (2 - p) \, r \, \Lambda ^{- 1} \, \Lambda ^\prime + 2 \right) = - B_0 < 0 \Longleftrightarrow \Lambda = r^{(B_0 - 2)/(2 - p)}, \quad B_0 > 0. } \end{aligned}$$

The term corresponding to \(I_{33}\) is estimated above by

$$\begin{aligned}&\displaystyle { \frac{p}{r \, (p - 2)} \, \int \limits _{ \{s = 1\} \times \mathrm{I}\!\mathrm{R}^m } \left( p \, f\right) ^{1 - \frac{2}{p}} \, \left| y \cdot D v\right| ^2 \, \mathcal{C}^q \, |y|^{- \frac{p - 2}{p - 1}} \, \left( 1 - \, |y|^{ \frac{p}{p - 1} } \right) _+^{ \frac{1}{p - 2} } \, d y }\\&\quad \displaystyle { \le \, \frac{p}{r \, (p - 2)} \, \int \limits _{ \{s = 1\} \times \mathrm{I}\!\mathrm{R}^m } \, p \, f \, \mathcal{C}^q \, |y|^{2 - \frac{p - 2}{p - 1}} \, \left( 1 - \, |y|^{ \frac{p}{p - 1} } \right) _+^{ \frac{1}{p - 2} } \, d y }\\&\quad \displaystyle { \le \, \frac{C \, r^{- \delta }}{r} \, \int \limits _{ \{s = 1\} \times \mathrm{I}\!\mathrm{R}^m } \, f^{\frac{2 (p - 1)}{p}} \, \mathcal{C}^q \, |y|^{2 - \frac{p - 2}{p - 1}} \, \left( 1 - \, |y|^{ \frac{p}{p - 1} } \right) _+^{ \frac{1}{p - 2} } \, d y } \\&\qquad \displaystyle { + \, \frac{C \, r^{ \frac{\delta \, p}{p - 2} }}{r} \, \int \limits _{ \{s = 1\} \times \mathrm{I}\!\mathrm{R}^m } \mathcal{C}^q \, |y|^{2 - \frac{p - 2}{p - 1}} \, \left( 1 - \, |y|^{ \frac{p}{p - 1} } \right) _+^{ \frac{1}{p - 2} } \, d y. } \end{aligned}$$

The other terms corresponding to \(I_{3i}\), \(i = 2, 4, 5\), are bounded above by \((- 1) \times \)the terms of the right hand side of (5.25), (5.26), (5.27) and (5.28).

Combining all of the estimations above we have

$$\begin{aligned} \displaystyle { \frac{d}{d r} E (r) }\le & {} J + \, C \, \left( r^{- 1 + \delta } + r^{- 1 + \frac{\delta }{p - 2}} + r^{- 1 + \frac{\delta p}{p - 2}} \right) \, \nonumber \\&+ \frac{C}{r^{3 + \delta } \, \Lambda ^2} \int \limits _{\{s = 1\} \times \mathrm{I}\!\mathrm{R}^m} \left| D_v g (v)\right| ^2 \, \mathcal{B} \, \mathcal{C}^q \, d y \nonumber \\&\displaystyle { + \, \left. C \, \frac{ r^{- \delta } + 1 }{ r \, \Lambda ^{2 (p - 1)} } \, \Vert \mathcal{C}^{q - 2} (t) \, ({\bar{e}} (u (t)))^{\frac{2 (p - 1)}{p}} \Vert _{L^\infty \left( \text{ supp } \, (\mathcal{B} (t)) \right) } \right| _{t = t_0 + r^{B_0}} },\qquad \end{aligned}$$
(5.63)

where \(\Lambda = r^{(B_0 - 2)/(2 - p)}\), but

$$\begin{aligned} \displaystyle { J = - \frac{1}{2} \, B_0 \, r^{- 1} \, \int \limits _{ \{s = 1\} \times \mathrm{I}\!\mathrm{R}^m } \, |\partial _s v|^2 \, \mathcal{B} \, \mathcal{C}^q \, d y. } \end{aligned}$$

The term J is clearly nonpositive. From (5.63) integrated on the interval \(\left( r, \, \rho \right) \) we derive

$$\begin{aligned}&\displaystyle { E (\rho ) - E (r) } \nonumber \\&\quad \le \, C \, \int _r^\rho \, \left( r^{- 1 + \delta } + r^{- 1 + \frac{\delta }{p - 2}} + r^{- 1 + \frac{\delta p}{p - 2}} \right) \, d r \, \nonumber \\&\qquad + \, \int _r^\rho \, \frac{C}{r^{3 + \delta } \, \Lambda ^2} \int \limits _{\{s = 1\} \times \mathrm{I}\!\mathrm{R}^m} \left| D_v g (v)\right| ^2 \, \mathcal{B} \, \mathcal{C}^q \, d y \, d r \nonumber \\&\qquad \displaystyle { + \, C \, \int _r^\rho \, \left. \frac{ r^{- \delta } + 1 }{ r \, \Lambda ^{2 (p - 1)} } \, \Vert \mathcal{C}^{q - 2} (t) \, ({\bar{e}} (u (t)))^{\frac{2 (p - 1)}{p}} \Vert _{L^\infty \left( \text{ supp } \mathcal{B} (t) \right) } \right| _{\tau = t_0 + r^{B_0}} \, d r } \nonumber \\&\quad \displaystyle { = : C \, (U_1 + U_2 + U_3). } \end{aligned}$$
(5.64)

The terms in the right hand side of (5.64) correspond to those in (5.30). Note that \(U_1\), \(U_2\) and \(U_3\) are just \((- 1) \times \) the corresponding terms in (5.30). \(U_1\) and \(U_3\) can be estimated exactly similarly as the corresponding terms in (5.30).

We have to care the estimation of \(U_2\). Under the scaling setting (5.59) and (5.60) in the forward case now, we also have the Bochner type estimate (5.33) for the penalty term. We can proceed to the estimations, similarly as in (5.34), to obtain

The estimation for \(U_{22}\) is the same as in (5.35) in the backward case.

\(U_{21}\) is also computed as in (5.36) in the backward case

$$\begin{aligned} \begin{aligned} \displaystyle { - U_{21} }&= \displaystyle { \int _r^{\rho } \frac{1}{r^{3 + \delta } \, \Lambda ^2} \, \int \limits _{ \{s = 1\} \times \mathrm{I}\!\mathrm{R}^m } \partial _s \big (g \, \mathcal{C}^q\big ) \, \mathcal{B} \, d y \, d r - \int _r^{\rho } \frac{1}{r^{3 + \delta } \, \Lambda ^2} \, \int \limits _{ \{s = 1\} \times \mathrm{I}\!\mathrm{R}^m } g \, \partial _s \mathcal{C}^q \, \mathcal{B} \, d y \, d r }\\&= : \displaystyle { U_{211} + U_{212}. } \end{aligned} \end{aligned}$$

The estimation for \(|U_{212}|\) is done in the same way as in (5.37) in the backward case. The estimation for \(U_{211}\) is performed in the following. By changing a variable \(t = t_0 + \Lambda (r)^{2 - p} r^2 = t_0 + r^{B_0}\), we have

$$\begin{aligned} \begin{aligned}&\displaystyle { t = t_0 + r^{B_0} \Longleftrightarrow r = (t - t_0)^{1/ B_0}; \quad {\bar{K}} = K \Lambda ^{- p} = K r^{ \frac{p (B_0 - 2)}{p - 2} } = (t - t_0)^{ \frac{p (B_0 - 2)}{B_0 (p - 2)} } ; }\\&\displaystyle { d t = B_0 r^{B_0 - 1} d r= B_0 \Lambda ^{2 - p} r d r; }\\&\displaystyle { \frac{1}{r^{2 + \delta } \, \Lambda ^2} = (t - t_0)^{- c_0}; \quad c_0 : = \frac{1}{B_0} \left( \frac{2 (p - B_0)}{p - 2} + \delta \right) } \end{aligned} \end{aligned}$$

and a computation

Thus, we have

$$\begin{aligned} \displaystyle { U_{211} }&= \displaystyle { \frac{1}{B_0} \, \int \limits ^{t_0 + \rho ^{B_0}}_{t_0 + {\bar{r}}^{B_0}} \, (t - t_0)^{- c_0} \, \frac{d P}{d t} \, d t - \frac{p (B_0 - 2)}{B_0^2 (p - 2)} \, \int \limits ^{t_0 + \rho ^{B_0}}_{t_0 + {\bar{r}}^{B_0}} \, (t - t_0)^{- c_0 - 1} \, P (t) \, d t } \nonumber \\&\qquad \displaystyle { - \frac{1}{B_0^2} \, \int \limits ^{t_0 + \rho ^{B_0}}_{t_0 + {\bar{r}}^{B_0}} \, (t - t_0)^{- c_0 - 1} \, \int \limits _{ \mathrm{I}\!\mathrm{R}^m } y \cdot D_y h (t, \, y) \, \mathcal{B} (y) \, d y \, d t } \nonumber \\&= : U_{2111} + U_{2112} + U_{2113}, \end{aligned}$$
(5.65)

where we put

$$\begin{aligned} \displaystyle { P (t) : = \int \limits _{ \mathrm{I}\!\mathrm{R}^m } h (t, \, y) \, \mathcal{B} (y) \, d y, \quad \mathcal{B} (y) = \left. \mathcal{B} (s, \, y)\right| _{s = 1} = \left( 1 - |y|^\frac{p}{p - 1}\right) _+^{\frac{p - 1}{p - 2}}. } \end{aligned}$$

We will estimate each term in (5.65).

Now, we set \(\bar{r}\) as

$$\begin{aligned} \displaystyle { \exists \, {\bar{r}}, \, \, r \le {\bar{r}} \le \rho \quad : \quad \min _{ t_0 + r^{B_0} \le t \le t_0 + \rho ^{B_0} } P (t) = P (t_0 + {\bar{r}}^{B_0}). } \end{aligned}$$
(5.66)

Then, by integration by parts in the integral on t, we have

$$\begin{aligned} \displaystyle { B_0 \times U_{2111} }= & {} \displaystyle { \int \limits ^{t_0 + \rho ^{B_0}}_{t_0 + {\bar{r}}^{B_0}} \, (t - t_0)^{- c_0} \, \frac{d P}{d t} \, d t } \nonumber \\= & {} \displaystyle { \left. (t - t_0)^{- c_0} \, P (t) \right| ^{t_0 + \rho ^{B_0}}_{t_0 + {\bar{r}}^{B_0}} + \int \limits ^{t_0 + \rho ^{B_0}}_{t_0 + {\bar{r}}^{B_0}} \, P (t) \, c_0 \, (t - t_0)^{- c_0 - 1} \, d t } \nonumber \\\ge & {} \displaystyle { {\rho }^{- c_0 \, B_0} \, P (t_0 + {\rho }^{B_0}) - {\bar{r}}^{- c_0 \, B_0} \, P (t_0 + {\bar{r}}^{B_0}) } \nonumber \\&\quad \displaystyle { + P (t_0 + {\bar{r}}^{B_0}) \left( {\bar{r}}^{- c_0 \, B_0} - {\rho }^{- c_0 \, B_0} \right) } \nonumber \\= & {} \displaystyle { {\rho }^{- c_0 \, B_0} \left( P (t_0 + {\rho }^{B_0}) - P (t_0 + {\bar{r}}^{B_0}) \right) \ge 0. } \end{aligned}$$
(5.67)

By integration by parts in the integral on y, we also have

$$\begin{aligned} \displaystyle { U_{2113} }= & {} \frac{1}{B_0^2} \, \int \limits ^{t_0 + \rho ^{B_0}}_{t_0 + {\bar{r}}^{B_0}} \, (t - t_0)^{- c_0 - 1} \times \\&\quad \times \int \limits _{ \mathrm{I}\!\mathrm{R}^m } h (t, \, y) \, \left( m \, \mathcal{B} - \frac{p}{p - 2} \big ( 1 - |y|^{\frac{p}{p - 1}} \big )_+^{\frac{1}{p - 2}} \, |y|^{\frac{p}{p - 1}} \right) \, d y \, d t; \\ \displaystyle { U_{2112} + U_{2113} }\ge & {} \displaystyle { \frac{1}{B_0^2} \left( m - \frac{p (B_0 - 2)}{p - 2} \right) \, \int \limits ^{t_0 + \rho ^{B_0}}_{t_0 + {\bar{r}}^{B_0}} \, (t - t_0)^{- c_0 - 1} \, P (t) \, d t }\\&- \frac{p}{B_0^2 (p - 2)} \, \int \limits ^{t_0 + \rho ^{B_0}}_{t_0 + {\bar{r}}^{B_0}} \, (t - t_0)^{- c_0 - 1} \\&\quad \int \limits _{ \mathrm{I}\!\mathrm{R}^m } h (t, \, y) \, \mathcal{C}^q (t) \, \big ( 1 - |y|^{\frac{p}{p - 1}} \big )_+^{\frac{1}{p - 2}} \, |y|^{\frac{p}{p - 1}} \, d y \, d t. \end{aligned}$$

The 1st term is nonnegative, since \(m - \frac{p (B_0 - 2)}{p - 2} > m - p \ge 0\), and the last integral is estimated below as

$$\begin{aligned}&\displaystyle { g (u (t)) : = \frac{K}{2} \chi \left( { \text{ dist }}^2 \big ( u (t, \, x_0 + (t - t_0)^{1/B_0} y), \, {\mathcal{N}} \big ) \right) , } \nonumber \\&\displaystyle { \mathcal{C}^q (t) : = \mathcal{C}^q (t, \, x_0 + (t - t_0)^{1/B_0} y); } \nonumber \\&\qquad \displaystyle { - \frac{p}{B_0^2 (p - 2)} \, \int \limits ^{t_0 + \rho ^{B_0}}_{t_0 + {\bar{r}}^{B_0}} \, (t - t_0)^{b_0 - 1} \, \, \int \limits _{ \mathrm{I}\!\mathrm{R}^m } g (u (t)) \, \mathcal{C}^q (t) \, \big ( 1 - |y|^{\frac{p}{p - 1}} \big )_+^{\frac{1}{p - 2}} \, |y|^{\frac{p}{p - 1}} \, d y \, d t } \nonumber \\&\quad \displaystyle { \ge - \frac{p}{B_0^2 (p - 2)} \, \int \limits _{ \mathrm{I}\!\mathrm{R}^m } \big ( 1 - |y|^{\frac{p}{p - 1}} \big )_+^{\frac{1}{p - 2}} \, |y|^{\frac{p}{p - 1}} \, d y \, \times } \nonumber \\&\qquad \quad \displaystyle { \times \left( \int \limits ^{t_0 + \rho ^{B_0}}_{t_0 + {\bar{r}}^{B_0}} (t - t_0)^{\alpha _0 (b_0 - 1)} \, d t + \, \int \limits ^{t_0 + \rho ^{B_0}}_{t_0 + {\bar{r}}^{B_0}} \Vert \frac{K}{2} \chi \big ( { \text{ dist }}^2 \big ( u (t, \, \cdot ), \, {\mathcal{N}} \big ) \big ) \, \mathcal{C}^q (t) \Vert _{L^\infty \big ( \text{ supp } \mathcal{B} (t) \big )}^{ \frac{\alpha _0}{\alpha _0 - 1} } dt\right) , }\nonumber \\ \end{aligned}$$
(5.68)

where \(\alpha _0 > 1\) is as in (5.43) in the backward case. Therefore we have

$$\begin{aligned} \displaystyle { - U_{21} }\ge & {} \displaystyle { - \, C \, \left( \rho ^{\alpha _0 (b_0 - 1) + 1} - {\bar{r}}^{\alpha _0 (b_0 - 1) + 1} \right) }\\&\displaystyle { - \, C \, \int \limits ^{t_0 + \rho ^{B_0}}_{t_0 + {\bar{r}}^{B_0}} \Vert {\bar{e}} (u (t)) \, \mathcal{C}^q (t) \Vert _{L^\infty \big ( \text{ supp } \mathcal{B} (t) \big )}^{ \frac{\alpha _0}{\alpha _0 - 1} } } \\&\displaystyle { - \, C \, \int \limits ^{t_0 + \rho ^{B_0}}_{t_0 + {\bar{r}}^{B_0}} \, \Vert \mathcal{C}^{q - 1} (t) \, {\bar{e}} (u (t)) \Vert _{L^\infty \big (\mathrm{supp} (\mathcal{B} (t))\big )} \, d t. } \end{aligned}$$

By definition of P(t) in (5.65) and \(\bar{r}\) in (5.66), P(t) is the local scaled integral of the penalty term, because by changing a variable \(x = x_0 + (t - t_0)^{1/B_0} y\)

$$\begin{aligned}&\displaystyle { h (t, \, y) = \frac{\bar{K}}{2} \chi \left( { \text{ dist }}^2 \big (u (t, \, x_0 + (t - t_0)^{1/B_0} y), \, {\mathcal{N}}\big ) \right) \, \left( t^{1/B_0} - |x_0 + (t - t_0)^{1/B_0} y| \right) _+^q; }\\&\displaystyle { P (t) = \int \limits _{ \mathrm{I}\!\mathrm{R}^m } h (t, \, y) \, \mathcal{B} \, d y = \frac{1}{\Lambda ^p} \int \limits _{ \mathrm{I}\!\mathrm{R}^m } \frac{K}{2} \chi \big ( { \text{ dist }}^2 (u (t, \, x), \, {\mathcal{N}}) \big ) \, \mathcal{C}^q (t, \, x) \, \mathcal{B} (t, \, x) \, d x } \end{aligned}$$

and it holds that

$$\begin{aligned} \displaystyle { P (t_0 + {\bar{r}}^{B_0}) \le P (t_0 + r^{B_0}) \quad \text{ for } {\bar{r}} \text{ in } (5.66). } \end{aligned}$$
(5.69)

Collecting the estimations for \(U_1\), \(U_2\) and \(U_2\) above in (5.64), we have, for \({\bar{r}}\) in (5.66),

$$\begin{aligned} \displaystyle { E (\rho ) - E ({\bar{r}}) }\le & {} \displaystyle { \, C \, \left( \rho ^\mu - {\bar{r}}^\mu \right) } \nonumber \\&+ \, C \, \int \limits ^{t_0 + \rho ^{B_0}}_{t_0 + {\bar{r}}^{B_0}} \Vert \mathcal{C}^{q - 2} (t) \, {\bar{e}} (u (t))^{\frac{2 (p - 1)}{p}} +\nonumber \\&+ \mathcal{C}^q (t) \, \big ( {\bar{e}} (u (t))^2 + {\bar{e}} (u (t))^{\frac{\alpha _0}{\alpha _0 - 1}} \big ) \Vert _{L^\infty \left( \text{ supp } (\mathcal{B} (t)) \right) } \, d t. \end{aligned}$$
(5.70)

Let \(\theta _0\) be as in (5.43) and (5.47) in the backward case. By (5.70), our desired monotonicity estimate holds true in the range of scale radius \([{\bar{r}}, \, \rho ]\), and (5.69) is the monotonicity estimate in the range \([r, \, {\bar{r}}]\) of the local scaled integral of the penalty term. Therefore, it remains to estimate the local scaled p-energy in the range of scale radius \([r, \, {\bar{r}}]\). The monotonicity estimate of the local scaled p-energy in the range of scale radius \([r, \, {\bar{r}}]\) is estimated exactly as Step 3 in the backward case. In fact, letting as in (5.48)

$$\begin{aligned} \displaystyle { \quad F (r) = \int \limits _{ \{s = 1\} \times \mathrm{I}\!\mathrm{R}^m } f (v (s, \, y)) \, \mathcal{B} (s, \, y) \, \mathcal{C}^q (s, \, y) \, d y; \quad f = f (v) : = \frac{1}{p} \big ({\bar{\epsilon }} + |D v|^2\big )^{\frac{p}{2}}, } \end{aligned}$$

we arrive at the estimate corresponding to (5.53)

$$\begin{aligned} \displaystyle { F (\rho ) - F (r) }\le & {} \displaystyle { C \, \left( \rho ^\mu - r^\mu \right) }\\&+ \int _r^\rho \frac{C}{r \, \Lambda ^{2 (p - 1)}} \, \Vert {\bar{e}} (u (t))^{\frac{2 (p - 1)}{p}} \, \mathcal{C}^{q - 2} (t)+ \\&\left. + {\bar{e}} (u (t))^2 \, \mathcal{C}^q (t) \Vert _{L^\infty \left( \text{ supp } \, (\mathcal{B} (t) \right) } \right| _{t = t_0 + r^{B_0}} \, d r, \end{aligned}$$

where the last term is controlled as in (5.54). \(\square \)