1 Introduction

In this note, we continue to analyse potential singularities of axisymmetric solutions to the non-stationary Navier–Stokes equations. In the previous paper [24], it has been shown that an axially symmetric solution is smooth provided a certain scale-invariant energy quantity of the velocity field is bounded. By definition, a potential singularity with bounded scale-invariant energy quantities is called the Type I blowup. It is important to notice that the above result does not follow from the so-called \(\varepsilon \)-regularity theory developed in [2, 16], and [10], where regularity is coming out due to smallness of those scale-invariant energy quantities.

We consider the 3D Navier–Stokes system

$$\begin{aligned} \partial _tv+v\cdot \nabla v-\Delta v=-\nabla q,\qquad \mathrm{div}\,v=0 \end{aligned}$$
(1.1)

in the parabolic cylinder \(Q={\mathcal {C}}\times ]-1,0[\), where \({\mathcal {C}}=\{x=(x_1,x_2,x_3):\,x_1^2+x^2_2<1,\,-1<x_3<1\}\). A solution v and q is supposed to be a suitable weak one, which means the following:

Definition 1.1

Let \(\omega \subset {\mathbb {R}}^3\) and \(T_2>T_1\). The pair w and r is a suitable weak solution to the Navier–Stokes system in \(Q_*=\omega \times ]T_1,T_2[\) if:

  1. 1.

    \(w\in L_{2,\infty }(Q_*)\), \(\nabla w\in L_2(Q_*)\), \(r\in L_\frac{3}{2}(Q_*)\);

  2. 2.

    w and r satisfy the Navier–Stokes equations in \(Q_*\) in the sense of distributions;

  3. 3.

    for a.a. \(t\in [T_1,T_2]\), the local energy inequality

    $$\begin{aligned} \int \limits _\omega \varphi (x,t)|w(x,t)|^2dx+2 \int \limits _{T_1}^t\int \limits _\omega \varphi |\nabla w|^2dxdt'\le \int \limits _{T_1}^t\int \limits _\omega [|w|^2(\partial _t\varphi +\Delta \varphi ) \\ +w\cdot \nabla \varphi (|w|^2+2r)]dxdt' \end{aligned}$$

    holds for all non-negative \(\varphi \in C^1_0(\omega \times ]T_1,T_2+(T_2-T_1)/2[).\)

In our standing assumption, it is supposed that a suitable weak solution v and q to the Navier–Stokes equations in \(Q=\mathcal C\times ]-1,0[\) is axially symmetric with respect to the axis \(x_3\). The latter means the following: if we introduce the corresponding cylindrical coordinates \((\varrho ,\varphi ,x_3)\) and use the corresponding representation \(v=v_\varrho e_\varrho +v_\varphi e_\varphi +v_3e_3\), then \(v_{\varrho ,\varphi }=v_{\varphi ,\varphi }=v_{3,\varphi }=q_{,\varphi }=0\).

There are many papers on regularity of axially symmetric solutions. We cannot pretend to cite all good works in this direction. For example, let us mention papers: [3,4,5, 9, 11,12,13, 19,20,21, 26,27,31], and [15].

Actually, our note is inspired by the paper [20], where the regularity of solutions has been proved under a slightly supercritical assumption. We would like to consider a different supercritical assumption, to give a different proof and to get a better result.

To state our supercritical assumption, additional notation is needed. Given \(x=(x_1,x_2,x_3)\in {\mathbb {R}}^3\), denote \(x'=(x_1,x_2,0)\). Next, different types of cylinders will be denoted as \({\mathcal {C}}(r)=\{x:\,|x'|<r, |x_3|<r\}\), \(\mathcal C(x_0,r)={\mathcal {C}}(r)+x_0\), \(Q^{\lambda ,\mu }(r)={\mathcal {C}}(\lambda r)\times ]-\mu r^2,0[\), \(Q^{1,1}(r)= Q(r)\), \(Q^{\lambda ,\mu }(z_0,r)={\mathcal {C}}(x_0,\lambda r)\times ]t_0-\mu R^2,t_0[\). And, finally, we let

$$\begin{aligned} f(R):=\frac{1}{\sqrt{R}} \left( \int \limits ^0_{-R^2}\left( \int \limits _{{\mathcal {C}}(R)} |v|^3dx\right) ^\frac{4}{3}dt\right) ^\frac{1}{4} \end{aligned}$$

and

$$\begin{aligned} M(R):=\frac{1}{\sqrt{R}}\left( \int \limits _{Q(R)}|v|^\frac{10}{3}dz\right) ^\frac{3}{10} \end{aligned}$$

for any \(0<R\le 1\) and assume that:

$$\begin{aligned} f(R)+M(R)\le g(R):=c_*\ln ^\alpha \ln ^\frac{1}{2}(1/R) \end{aligned}$$
(1.2)

for all \(0<R\le 2/3\), where \(c_*\) and \(\alpha \) are positive constants and \(\alpha \) obeys the condition:

$$\begin{aligned} 0<\alpha \le \frac{1}{224}. \end{aligned}$$
(1.3)

Without loss of generality, one may assume that \(g(R)\ge 1\) for \(0<R\le \frac{2}{3}\). To ensure the above condition, it is enough to increase the constant \(c_*\) if necessary.

Our aim could be the following completely local statement.

Theorem 1.2

Assume that a pair v and q is an axially symmetric suitable weak solution to the Navier–Stokes equations in Q and conditions (1.2) and (1.3) hold. Then the origin \(z=0\) is a regular point of v.

However, in this paper, we shall prove a weaker result leaving Theorem 1.2 as a plausible conjecture. We shall return to a proof of Theorem 1.2 elsewhere. In the present paper, the following fact is going to be justified.

Theorem 1.3

Let v be an axially symmetric solution to the Cauchy problem for the Navier–Stokes equations (1.1) in \({\mathbb {R}}^3\times ]0,T[\) with initial divergence free field \(v_0\) from the Sobolev space \(H^2 =W^2_2({\mathbb {R}}^3)\) such that

$$\begin{aligned} \sup \limits _{0<t<T-\delta }\Vert \nabla v(\cdot ,t)\Vert _{L_2({\mathbb {R}}^3)}\le C(\delta )<\infty \end{aligned}$$

for all \(0<\delta <T\). Assume further that

$$\begin{aligned} \Sigma _0= \sup \limits _{x\in \mathbb R^3}|v_{02}(x)x_1-v_{01}(x)x_2|<\infty \end{aligned}$$
(1.4)

and

$$\begin{aligned} \sup \limits _{-\infty<h<\infty } f(R;(0,h,T))+M(R;(0,h,T))\le g(R) \end{aligned}$$
(1.5)

for all \(0<R\le 2/3\) (it is assumed for simplicity that \(T>1\)), with some positive constants \(c_*\) and \(\alpha \), satisfying (1.3), where

$$\begin{aligned} f(R;z_0):=\frac{1}{\sqrt{R}} \left( \int \limits ^{t_0}_{t_0-R^2}\left( \int \limits _{{\mathcal {C}}(x_0,R)} |v|^3dx\right) ^\frac{4}{3}dt\right) ^\frac{1}{4} \end{aligned}$$

and

$$\begin{aligned} M(R;z_0):=\frac{1}{\sqrt{R}}\left( \int \limits _{Q(z_0,R)}|v|^\frac{10}{3}dz\right) ^\frac{3}{10}. \end{aligned}$$

Then v is a strong solution to the above Cauchy problem in \({\mathbb {R}}^3\times ]0,T[\), i.e.,

$$\begin{aligned} \sup \limits _{0<t<T}\Vert \nabla v(\cdot ,t)\Vert _{L_2({\mathbb {R}}^3)}<\infty . \end{aligned}$$

Our proof is based on the analysis of the following scalar equation

$$\begin{aligned} \partial _t\sigma +\left( v+2\frac{x'}{|x'|^2}\right) \cdot \nabla \sigma -\Delta \sigma =0 \end{aligned}$$
(1.6)

in \(Q\setminus (\{x'=0\}\times ]-1,0[)\), where \(\sigma :=\varrho v_\varphi =v_2x_1-v_1x_2\).

Let us list some differentiability properties of \(\sigma \). Some of them follows from partial regularity theory developed by Caffarelli–Kohn–Nirenberg.

Indeed, since v and q are an axially symmetric suitable weak solution, there exists a closed set \(S^\sigma \) in Q, whose 1D-parabolic measure in \({\mathbb {R}}^3\times {\mathbb {R}}\) is equal to zero and \(x'=0\) for any \(z=(x,t)\in S^\sigma \), such that any spatial derivative of v (and thus of \(\sigma \)) is Hölder continuous in \(Q\setminus S^\sigma \).

Next, we observe that

$$\begin{aligned} |\partial _t\sigma (z)-\Delta \sigma (z)|\le (\sup \limits _{z=(x,t)\in P(\delta ,R;R)\times ]-R^2,0[}|v(z)|+2/\delta )|\nabla \sigma (z)| \end{aligned}$$

for any \(0<\delta<R<1\), where \(P(a,b;h)=\{x: \,a<|x'|<b,\,|x_3|<h\}\). Since v is axially symmetric, the first factor on the right hand side is finite. This fact, by iteration, yields

$$\begin{aligned} \sigma \in W^{2,1}_p(P(\delta ,R;R)\times ]-R^2,0[) \end{aligned}$$

for any \(0<\delta<R<1\) and for any finite \(p\ge 2\).

It follows from the above partial regularity theory that, for any \(-1<t<0\),

$$\begin{aligned} \sigma (x',x_3,t)\rightarrow 0 \quad \text{ as }\quad |x'|\rightarrow 0 \end{aligned}$$
(1.7)

for all \(x_3\in ]-1,1[\setminus S^\sigma _t\), where \(S^\sigma _t=\{x_3\in ]-1,1[:\,(0,x_3,t)\in S^\sigma \}\).

In the same way, as it has been done in [26] and [24], one can show that \(\sigma \in L_\infty (Q(R))\) for any \(0<R<1\).

The main part of the proof of Theorem 1.3 is the following fact.

Proposition 1.4

Let \(\sigma =\varrho v_\varphi \), then

$$\begin{aligned} \mathrm{osc}_{z\in Q(r)}\le e^{-c\left[ \ln ^\frac{1}{4}(1/(2r)) -\ln ^\frac{1}{4}(1/(2R))\right] } \mathrm{osc}_{z\in Q(2R)}\sigma (z) \end{aligned}$$
(1.8)

where c is a positive absolute constant and \(0<r<R\le R_*(c_*,\alpha )\le 1/6\).

Here, \(\mathrm{osc}_{z\in Q(r)}\sigma (z) =M_{r}-m_{r}\) and

$$\begin{aligned} M_{r}=\sup \limits _{z\in Q(r)}\sigma (z),\qquad m_{r}=\inf \limits _{z\in Q(r)}\sigma (z). \end{aligned}$$

The above statement is an improvement of the result in [20], where the bound for oscillations of \(\sigma \) contains a fixed power of logarithmic factor only.

Remark 1.5

It is not so difficult to see that all results of the paper remain to be true if we replace v with \({\overline{v}}=v_\varrho e_\varrho +v_3e_3\) in the definitions of quantities M and f, see conditions (1.2) and (1.5).

The proof of Proposition 1.4 is based on a technique developed in [18], see also references there. We also would like to mention interesting results for the heat equation with a divergence free drift, see [1, 6, 7, 25].

2 Auxiliary Facts

Define the class \({\mathcal {V}}\) of functions \(\pi :Q\rightarrow {\mathbb {R}}\) possessing the properties:

(i) there exists a closed set \(S^\pi \) in Q, whose 1D-parabolic measure \({\mathbb {R}}^3\times {\mathbb {R}}\) is equal to zero and \(x'=0\) for any \(z=(x',x_3,t)\in S^\pi \), such that any spatial derivative is Hölder continuous in \(Q\setminus S^\pi \);

(ii)

$$\begin{aligned} \pi \in W^{2,1}_2(P(\delta ,R;R)\times ]-R^2,0[)\cap L_\infty (Q(R)) \end{aligned}$$

for any \(0<\delta<R<1\).

We are going to use the following subclass \({\mathcal {V}}_0\) of the class \({\mathcal {V}}\), saying that \(\pi \in {\mathcal {V}}_0\) if and only if \(\pi \in {\mathcal {V}}\) and

$$\begin{aligned} \partial _t\pi +\left( v+2\frac{x'}{|x'|^2}\right) \cdot \nabla \pi -\Delta \pi = 0\end{aligned}$$
(2.1)

in \({\mathcal {C}}\setminus \{x'=0\}\times ]-1,0[\).

We shall also say that \(\pi \in {\mathcal {V}}_0\) has the property \(({\mathcal {B}}_R)\) in Q(2R) if there exists a number \(k_R>0\) such that \(\pi (0,x_3,t)\ge k_R\) for \(-(2R)^2\le t\le 0\), \(x_3\in ]-2R,2R[\setminus S^\pi _t\).

Remark 2.1

Let \(0<r\le R\) and \(\pi \in {\mathcal {V}}_0\) have the property \(({\mathcal {B}}_R)\) in Q(2R). Then \(\pi \) has the property \((\mathcal B_r)\) in Q(2r) with any constant less or equal to \(k_R\).

In what follows, we always suppose that \(0<R\le 1/6\).

Proposition 2.2

Let \(\pi \in {\mathcal {V}}_0\) have the property \((\mathcal B_R)\). Then, for any \(0<k\le k_R\), for any \(0<\tau _1<\tau <2\), and for any \(0<\gamma _1<\gamma <4\), the following inequality holds:

$$\begin{aligned} \sup \limits _{z\in Q^{\tau _1,\gamma _1}(R)}\sigma (z)\le c_1(\tau _1,\tau ,\gamma _1,\gamma ,M(2R))\left( \frac{1}{|Q^{\tau ,\gamma }(R)|}\int \limits _{Q^{\tau ,\gamma }(R)}\sigma ^\frac{10}{3}dz\right) ^\frac{3}{10}, \end{aligned}$$
(2.2)

where \(\sigma =(k-\pi )_+\),

$$\begin{aligned} c_1(\tau _1,\tau ,\gamma _1,\gamma ,M(2R))=\frac{c}{(\tau -\tau _1)^\frac{16}{3}}\left( 1+\frac{\tau -\tau _1}{\sqrt{\gamma -\gamma _1}}+\left( \frac{1}{\gamma _1\tau _1^3}\right) ^\frac{1}{10}M(2R)\right) ^3, \end{aligned}$$

and \(Q^{\tau ,\gamma }(R)={\mathcal {C}}(\tau R)\times ]-\gamma R^2,0[\).

Proof

Repeating arguments in [24], we can get the following estimate of \(h=\sigma ^m\):

$$\begin{aligned}&\left( \int \limits ^0_{t_2}\int \limits _{{\mathcal {C}}(r_2)}|h|^\frac{10}{3}dz\right) ^\frac{3}{10}\nonumber \\&\quad \le c\left( \int \limits ^0_{t_1}\int \limits _{\mathcal C(r_1)}|h|^\frac{5}{2}dz\right) ^\frac{2}{5} \frac{(r_1^3|t_1|)^\frac{1}{10}}{r_1-r_2} \left( 1+\frac{r_1-r_2}{\sqrt{t_2-t_1}}+\overline{M}(r_1,t_1)+\frac{r_1^\frac{13}{8}|t_1|^\frac{1}{36}}{(r_1-r_2)^\frac{7}{9}}\right) \end{aligned}$$
(2.3)

for any \(0<r_2<r_1<2R\) and \(-4R^2<t_1<t_2<0\), where

$$\begin{aligned} {{\overline{M}}}(r_1,t_1)=\left( \frac{1}{|t_1|r_1^3}\right) ^\frac{1}{10}\left( \int \limits ^0_{t_1}\int \limits _{{\mathcal {C}}(r_1)}|v|^\frac{10}{3}dz\right) ^\frac{3}{10}. \end{aligned}$$

Next, we wish to iterate (2.3). To this end, let \(m=m_i=\Big (4/3\Big )^i\),

$$\begin{aligned} r_1=r_i=\tau _1R+(\tau -\tau _1)R2^{-i+1}, \qquad r_2=r_{i+1} , \\ t_1=t_i=-\gamma _1R^2-(\gamma -\gamma _1)R^24^{-i+1},\qquad t_2=t_{i+1}, \end{aligned}$$

where \(i=1,2,...\). Then, we can derive from (2.3) the following inequality

$$\begin{aligned} G_{i+1}\le \left( \frac{c2^{i+1}}{\tau -\tau _1}\right) ^\frac{1}{m_i}\left( 1+\frac{\tau -\tau _1}{\sqrt{\gamma -\gamma _1}}+{{\overline{M}}}(r_i,t_i)+\frac{2^{(i+1)\frac{7}{9}}}{(\tau -\tau _1)^\frac{7}{9}}\right) ^\frac{1}{m_i}G_i, \end{aligned}$$
(2.4)

where

$$\begin{aligned} G_i=\left( \frac{1}{|t_i|r_i^3}\int \limits ^0_{t_i}\int \limits _{{\mathcal {C}}(r_i)}\sigma ^\frac{5m_i}{2}dz\right) ^\frac{2}{5m_i}. \end{aligned}$$

Noticing that

$$\begin{aligned} {{\overline{M}}}(r_i,t_i)\le c\left( \frac{1}{\gamma _1\tau _1^3}\right) ^\frac{1}{10}M(2R), \end{aligned}$$

let us make use of (2.4) to obtain the estimate

$$\begin{aligned} G_{i+1} \le \left( \frac{c2^{i+1}}{\tau -\tau _1}\right) ^\frac{1}{m_i}\left( 1+\frac{\tau -\tau _1}{\sqrt{\gamma -\gamma _1}}+\frac{2^{(i+1)\frac{7}{9}}}{(\tau -\tau _1)^\frac{7}{9}}+\left( \frac{1}{\gamma _1\tau _1^3}\right) ^\frac{1}{10}M(2R)\right) ^\frac{1}{m_i}G_i, \end{aligned}$$
(2.5)

which, after iterations, gives the following

$$\begin{aligned} G_{i+1}\le \xi _iG_1, \end{aligned}$$
(2.6)

where

$$\begin{aligned} \xi _i=\prod \limits ^i_{k=1}\left( \frac{c2^{k+1}}{\tau -\tau _1}\right) ^\frac{1}{m_k}\left( 1+\frac{\tau -\tau _1}{\sqrt{\gamma -\gamma _1}}+\frac{2^{(k+1)\frac{7}{9}}}{(\tau -\tau _1)^\frac{7}{9}}+\left( \frac{1}{\gamma _1\tau _1^3}\right) ^\frac{1}{10}M(2R)\right) ^\frac{1}{m_k}. \end{aligned}$$

Obviously,

$$\begin{aligned} \xi _i\le \prod \limits ^i_{k=1}\left( \frac{c2^{k+1}}{\tau -\tau _1}\right) ^\frac{1}{m_k} \left( 1+\frac{2^{(k+1)\frac{7}{9}}}{(\tau -\tau _1)^\frac{7}{9}}\right) ^\frac{1}{m_k}\left( 1+\frac{\tau -\tau _1}{\sqrt{\gamma -\gamma _1}} +\left( \frac{1}{\gamma _1\tau _1^3}\right) ^\frac{1}{10}M(2R)\right) ^\frac{1}{m_k}. \end{aligned}$$

Next,

$$\begin{aligned} \ln \xi _i\le A_1+A_2+A_3, \end{aligned}$$

where

$$\begin{aligned} A_1= & {} \sum ^i_{k=1}\frac{1}{m_k}(\ln c +(k+1)\ln 2-\ln (\tau -\tau _1))\le \ln c-3\ln (\tau -\tau _1), \\ A_2= & {} \sum ^i_{k=1}\frac{1}{m_k}\ln \left( 1+\frac{2^{(k+1)\frac{7}{9}}}{(\tau -\tau _1)^\frac{7}{9}}\right) =\sum ^i_{k=1}\frac{1}{m_k}\ln \left( \frac{2^{(k+1)\frac{7}{9}}}{(\tau -\tau _1)^\frac{7}{9}}\right) \\&\quad +\frac{1}{m_k}\ln \left( 1+\frac{(\tau -\tau _1)^\frac{7}{9}}{2^{(k+1)\frac{7}{9}}}\right) \le \ln \frac{c}{(\tau -\tau _1)^\frac{7}{3}} \\&\quad +(\tau -\tau _1)^\frac{7}{9}\sum ^i_{k=1} \frac{1}{m_k}\frac{1}{2^{(k+1)\frac{7}{9}}}\le \ln \frac{c}{(\tau -\tau _1)^\frac{7}{3}}, \end{aligned}$$

and

$$\begin{aligned}&A_3=\ln \left( 1+\frac{\tau -\tau _1}{\sqrt{\gamma -\gamma _1}}+\left( \frac{1}{\gamma _1\tau _1^3}\right) ^\frac{1}{10}M(2R)\right) \sum ^i_{k=1}\frac{1}{m_k} \\&\quad \le \ln \left( 1+\frac{\tau -\tau _1}{\sqrt{\gamma -\gamma _1}}+\left( \frac{1}{\gamma _1\tau _1^3}\right) ^\frac{1}{10}M(2R)\right) ^3. \end{aligned}$$

So,

$$\begin{aligned} \xi _i\le \frac{c}{(\tau -\tau _1)^\frac{16}{3}} \left( 1+\frac{\tau -\tau _1}{\sqrt{\gamma -\gamma _1}}+\left( \frac{1}{\gamma _1\tau _1^3}\right) ^\frac{1}{10}M(2R)\right) ^3. \end{aligned}$$

Passing to the limit as \(i\rightarrow \infty \) in (2.6), we complete the proof the Proposition. \(\square \)

Remark 2.3

If we additionally assume that \(\pi (\cdot ,-\theta R^2) \ge k\) in B for some \(0<\theta \le 1\), then we do not need to use a cut-off in t. So, for \(0<\lambda <1\), we have

$$\begin{aligned} \sup \limits _{Q^{\lambda ,\theta }(R)}\sigma \le c'_1(\lambda ,\theta ,M(2R))\left( \frac{1}{|Q^{1,\theta }(R)|}\int \limits _{Q^{1,\theta }(R)}\sigma ^\frac{10}{3}dz\right) ^\frac{3}{10}, \end{aligned}$$

where

$$\begin{aligned} c'_1(\lambda ,\theta ,M(2R))=\frac{c}{(1-\lambda )^\frac{16}{3}}\left( 1+\left( \frac{1}{\theta \lambda ^3}\right) ^\frac{1}{10}M(2R)\right) ^3. \end{aligned}$$

Corollary 2.4

Let a non-negative function \(\pi \in {\mathcal {V}}_0\) have the property \(({\mathcal {B}}_R)\) in Q(2R) and let \(0<\lambda _1<\lambda <2\) and \(0<\theta \le 1\). Suppose that

$$\begin{aligned} |\{\pi<k\}\cap Q^{\lambda , \theta }((0,t_0),R)|<\mu |Q^{\lambda ,\theta }(R)| \end{aligned}$$
(2.7)

for some \(t_0>-4R^2\), for some \(0<k\le k_R\), and for some

$$\begin{aligned} 0<\mu \le \mu _*=\left( \frac{1}{2c_1(\lambda _1,\lambda ,\theta /2,\theta ,M(2R))}\right) ^\frac{10}{3}. \end{aligned}$$

Then \(\pi \ge \frac{k}{2}\) in \(Q^{\lambda _1,\theta /2}((0,t_0),R)\).

If, in addition, \(\pi (\cdot ,t_0-\theta R^2)>k\) in \(\mathcal C(\lambda R)\), then \(\pi \ge \frac{k}{2}\) in \(Q^{\lambda _1,\theta }((0,t_0),R)\).

Proof

The first statement can be proved ad absurdum with the help of inequality (2.2) and a suitable choice of the number \(\mu _*\). The second statement is proved in the same way but with the help of the inequality of Remark 2.3. Number \(\mu _*\) is defined by the constant \(c_1'\) instead of \(c_1\). \(\square \)

The two lemmas below are obvious modifications of the corresponding statements in the paper [18].

Lemma 2.5

Let \(0\le \pi \in {\mathcal {V}}_0\) have the property \(({\mathcal {B}}_R)\) in Q(2R). Given \(0<\delta _0\le 1\), there exists a positive number \(\theta _0(\delta _0, f(2R))\le 1\) such that if, for \(0<\theta \le \theta _0\), \(0<k_0\le k_R\), there holds

$$\begin{aligned} |\{\pi (\cdot ,t_0-\theta R^2)\ge k_0\}\cap {\mathcal {C}}(R)|>\delta _0|{\mathcal {C}}(R)|, \end{aligned}$$

then

$$\begin{aligned} |\{\pi (\cdot ,t)\ge \frac{\delta _0}{3}k_0\}\cap {\mathcal {C}}(R)|>\frac{\delta _0}{3}|{\mathcal {C}}(R)| \end{aligned}$$

for all \(t\in [t_0-\theta R^2,t_0]\).

Remark 2.6

There is a formula for \(\theta _0\):

$$\begin{aligned} \theta _0=\left( \frac{c\delta _0^6}{1+\delta _0^2f(2R)}\right) ^\frac{4}{3}. \end{aligned}$$

Lemma 2.7

Let \(0\le \pi \in {\mathcal {V}}_0\) have the property \(({\mathcal {B}}_R)\) in Q(2R). Let, for any \(t\in [t_0-\theta _1R^2,t_0]\),

$$\begin{aligned} |\{\pi (\cdot ,t)\ge k_1\}\cap {\mathcal {C}}(R)|\ge \delta _1|{\mathcal {C}}(R)| \end{aligned}$$

for some \(0<k_1\le k_R\) and for some \(0<\delta _1\le 1\) and \(0<\theta _1\le 1\).

Then, for any \(\mu _1\in ]0,1[\), the following inequality is valid:

$$\begin{aligned} |\{\pi <2^{-s}k_1\}\cap Q^{1,\theta _1}((0,t_0),R)|\le \mu _1|Q^{1,\theta _1}(R)| \end{aligned}$$

with the integer number s defined as

$$\begin{aligned} s=\mathrm{entier} \left( \frac{c}{\delta _1^2\mu _1^2\theta _1}(1+f(2R))\right) +1. \end{aligned}$$

Corollary 2.8

Let \(0\le \pi \in {\mathcal {V}}_0\) have the property \(({\mathcal {B}}_R)\) in Q(2R). If \(\pi (\cdot ,{{\overline{t}}})\ge k_2\) in \({\mathcal {C}}(R)\), then, for any \(\sigma \in ]0,1[\), the inequality \(\pi \ge \beta _2k_2\) holds in \(Q^{\sigma ,\theta _0}((0,t_0),R)\) with \(\theta _0=(c/g(2R))^\frac{4}{3}\) and \(t_0={{\overline{t}}}+\theta R^2\), where

$$\begin{aligned} \beta _2=\frac{1}{6} 2^{-c(1-\sigma )^{-40}\sigma ^{-6}g^{25}(2R)} \end{aligned}$$

provided \(R\le R_{*}(c_*,\alpha )\). It is supposed that \(0<k_2\le k_R\).

Proof

We apply Lemma 2.5 with \(\delta _0=1\) and \(k_0=k_2\). Then, for \(\sigma =4/(27c)\), we have

$$\begin{aligned} \theta _0=\left( \frac{\frac{4}{27}\sigma }{\frac{1}{\sigma }+f(2R)}\right) ^\frac{4}{3}\ge \left( \frac{c}{g(2R)}\right) ^\frac{4}{3} \end{aligned}$$

for all \(0<R\le R_{*}(c_*,\alpha )\) and state that the following inequality holds:

$$\begin{aligned} |\{\pi (\cdot ,t)>\frac{k_0}{3}\}\cap {\mathcal {C}}(R)|\ge \frac{1}{3}|{\mathcal {C}}(R)| \end{aligned}$$

for any \(t\in [t_0-\theta _0R^2,t_0]\), where \(t_0={{\overline{t}}}+\theta _0R^2\). In what follows, we are going to use the quantity \((c/(g(2R)))^\frac{4}{3}\) as a new number \(\theta _0\) instead of \(\theta _0(1,f(2R))\).

Now, we are going to apply Lemma 2.7 with another set of parameters \(k_1=\frac{1}{3}k_2\), \(\theta _1=\theta _0\), \(\delta _1=\frac{1}{3}\), and

$$\begin{aligned}&\mu _1=\mu _*=\left( \frac{1}{2c'_1}\right) ^\frac{10}{3}, \qquad c'_1=\frac{c}{(1-\sigma )^\frac{16}{3}}\left( 1+\left( \frac{1}{\theta _0\sigma ^3}\right) ^\frac{1}{10}M(2R)\right) ^3\\&\quad \le \frac{c}{(1-\sigma )^\frac{16}{3}}\left( \frac{1}{\theta _0\sigma ^3}\right) ^\frac{3}{10}g^3(2R). \end{aligned}$$

Lemma (2.7) gives us:

$$\begin{aligned} |\{\pi<2^{-s}k_1\}\cap Q^{1,\theta _1}((0,t_0),R)|<\mu _1| Q^{1,\theta _1}(R)|, \end{aligned}$$

where

$$\begin{aligned} s=\mathrm{entier} \left( \frac{c}{\delta _1^2\mu _1^2\theta _1}(1+f(2R))\right) +1 \end{aligned}$$

and \( \mathrm{entier}(x)\) is the largest integer not exceeding x. But we know that

$$\begin{aligned} \pi (\cdot ,t_0-\theta _0 R^2)\ge k_2>2^{-s}k_1=2^{-s}\frac{k_2}{3}. \end{aligned}$$

Then, from Corollary 2.4, it follows that \(\pi >\frac{1}{2}2^{-s}k_1=\beta _2k_2\) with \(\beta _2=\frac{1}{2}2^{-s}\frac{1}{3}\) in \(Q^{\sigma ,\theta _0}((0,t_0),R)\). \(\square \)

Given \(\theta \in ]0,1]\), we can find an number \(0<R_{*1}(c_*,\alpha ,\theta )\le R_*(c_*,\alpha )\) so that \(\Big (\frac{c}{g(2r)}\Big )^\frac{4}{3}\le \theta \) for all \(0<r\le R_{*1}\).

Lemma 2.9

Let \(0\le \pi \in {\mathcal {V}}_0\) have the property \(({\mathcal {B}}_R)\) in Q(2R), assuming that \(R\le R_{*1}(c_*,\alpha ,\theta )\) for some \(0<\theta \le 1\). Suppose further that, for some \(0<k\le k_R\) and for some \(-R^2\le {{\overline{t}}}\le -\theta R^2\), there holds \(\pi (\cdot ,{{\overline{t}}})\ge k\) in \({\mathcal {C}}(R)\). Then \(\pi \ge \beta _0k\) in \({{\widehat{Q}}}:={\mathcal {C}}(\frac{2}{3}R)\times [\overline{t},0]\), where

$$\begin{aligned} \beta _0\ge \ln ^{-\frac{1}{2}}( 1/R) \end{aligned}$$

for \(R\le R_{*2}(c_*,\alpha ,\theta ).\)

Proof

Let

$$\begin{aligned} N=\mathrm{entier}\left( \frac{9}{8}\frac{ |{{\overline{t}}}|}{{{\tilde{\theta }}}_0R^2}\right) +1, \end{aligned}$$

where \({{\tilde{\theta }}}_0=(c/g(\frac{2}{3}2R))^\frac{4}{3} \le \theta \). Next, we introduce

$$\begin{aligned} {\hat{\theta }}_0=\frac{ |{{\overline{t}}}|}{(\frac{8N}{9}+\frac{1}{2N})R^2}\le {{\tilde{\theta }}}_0. \end{aligned}$$

Step 1. By Corollary 2.8, the inequality \(\pi \ge \beta ^{(1)}_2k\) holds at least in \({\mathcal {C}}((1-\frac{1}{3N})R)\times [{{\overline{t}}}_1,{{\overline{t}}}_1+{\hat{\theta }}_0R^2]\), where \(\overline{t}_1={{\overline{t}}}\), \({{\overline{t}}}_2= {{\overline{t}}}_1+{\hat{\theta }}_0R^2\), \(\sigma =1-1/(3N)\ge 2/3\), \(1-\sigma =1/(3N)\), and

$$\begin{aligned} \ln \beta ^{(1)}_2=-\ln 6-cN^{40}g^{25}(2R) \end{aligned}$$

Step 2. Here, we are going to use Corollary 2.8 with \(R(1-1/(3N))\) instead of R and with \(\sigma =(1-2(3N))/(1-1/(3N))\). As a result, we have the estimate \(\pi \ge \beta ^{(2)}_2\beta ^{(1)}_2k\) at least in \(\mathcal C((1-2/(3N))R)\times [{{\overline{t}}}_2,{{\overline{t}}}_2+{\hat{\theta }}_0 (1-1/(3N))^2R^2]\), \({{\overline{t}}}_3={{\overline{t}}}_2+{\hat{\theta }}_0 (1-1/(3N))^2R^2\), and

$$\begin{aligned} \ln \beta ^{(2)}_2=-\ln 6-cN^{40}g^{25}(2(1-1/(3N))R). \end{aligned}$$

So, \(\pi \ge \beta ^{(2)}_2\beta ^{(1)}_2k\) in \({\mathcal {C}}((1-2(3N))R)\times [{{\overline{t}}}, {{\overline{t}}}_3]\).

After N steps, we shall have \({{\overline{t}}}_N=0\) and

$$\begin{aligned} \pi \ge \beta ^{(N)}_2...\beta ^{(1)}_2k=\beta _0(R)k \end{aligned}$$

in \({\mathcal {C}}(\frac{2}{3}R)\times [{{\overline{t}}},0]\), where

$$\begin{aligned} \ln \beta ^{(i+1)}_2=-\ln 6-cN^{40}g^{25}(2(1-i/(3N))R) \end{aligned}$$

for \(i=0,1,...,N-1\).

Next, according to assumption (1.2), we can have

$$\begin{aligned} \ln \beta _0\ge -N\ln 6-cN^{40}\sum \limits ^{N-1}_{k=1} c_*^{25}\ln ^\gamma \ln ^\frac{1}{2}\left( \frac{1}{2(1-i/(3N))R}\right) , \end{aligned}$$

where \(25\alpha <1\). Since

$$\begin{aligned} \ln \frac{1}{1-x}\le 2x \end{aligned}$$

provided \(0\le x\le 1/2\), we find, assuming that \(R\le 1/6\), the following:

$$\begin{aligned}&\ln ^\gamma \ln ^\frac{1}{2}\left( \frac{1}{2(1-i/(3N))R}\right) \le \ln ^\gamma \left( \ln \frac{1}{2R} +\frac{i}{N}\right) ^\frac{1}{2} \\&\quad \le \ln ^\gamma \left( \ln ^\frac{1}{2} \frac{1}{2R} +\left( \frac{i}{N}\right) ^\frac{1}{2}\right) = \ln ^\gamma \ln ^\frac{1}{2} \frac{1}{2R}\left( 1 +\left( \frac{i}{N\ln \frac{1}{2R}}\right) ^\frac{1}{2}\right) \\&\quad \le \ln ^\gamma \ln ^\frac{1}{2} \frac{1}{2R}\left( 1 +\left( \frac{i}{N}\right) ^\frac{1}{2}\right) =\left( \ln \left( \ln ^\frac{1}{2} \frac{1}{2R}\right) +\ln \left( 1 +\left( \frac{i}{N}\right) ^\frac{1}{2}\right) \right) ^\gamma \\&\quad \le \left( \ln \left( \ln ^\frac{1}{2} \frac{1}{2R}\right) +\left( \frac{i}{N}\right) ^\frac{1}{2}\right) ^\gamma \le \ln ^\gamma \left( \ln ^\frac{1}{2} \frac{1}{2R}\right) +\left( \frac{i}{N}\right) ^\frac{\gamma }{2}. \end{aligned}$$

From the latter inequality, one can deduce the bound

$$\begin{aligned}&\ln \beta _0\ge -N\ln 6-cc_*^{25}N^{40}\left( N\ln ^\gamma \ln ^\frac{1}{2} \frac{1}{2R}+\sum \limits ^{N-1}_{i=0}\left( \frac{i}{N}\right) ^\frac{\gamma }{2}\right) \\&\quad \ge -N\ln 6-cc_*^{25}N^{41}\ln ^\gamma \ln ^\frac{1}{2} \frac{1}{2R}, \end{aligned}$$

which is valid for \(0<R\le R_{*3}(\alpha )\le 1/6\). Taking into account that \(N\le c(g(2R))^\frac{4}{3}\), we conclude

$$\begin{aligned} \ln \beta _0\ge -c_1(c_*)\ln ^\frac{239\alpha }{3}\sqrt{\ln \frac{1}{R}}. \end{aligned}$$

It remains to find \(R_{*4}(c_*,\alpha )\le 1\) such that

$$\begin{aligned} c_1(c_*)\ln ^{\frac{239\alpha }{3}-1}\sqrt{\ln \frac{1}{R}}\le 1 \end{aligned}$$

for all \(0<R\le R_{*4}\). So, we have the required inequality provided \(0<R\le R_{*2}=\min \{R_{*1},R_{*3},R_{*4}\}\).\(\square \)

3 Proof of Proposition 1.4

Now, we can state an analog of Lemma 4.2 of [18] for the class \({\mathcal {V}}\).

Lemma 3.1

Let \(0\le \pi \in {\mathcal {V}}_0\) possess the property \(({\mathcal {B}}_R)\) in Q(2R).

Suppose further that

$$\begin{aligned} \pi \le M_0k_R\end{aligned}$$
(3.1)

in Q(2R) for some \(M_0\ge 1\). Then, there exists \({{\overline{t}}}\in [-R^2,-\frac{3}{4}R^2]\) such that

$$\begin{aligned} |e_{\kappa _0}({{\overline{t}}})|\ge \delta _0|{\mathcal {C}}(R)| \end{aligned}$$
(3.2)

Here, \(\kappa _0=\kappa _0(f(2R))= c/(1+f(2R))\), \(e_\kappa (t):=\{x\in {\mathcal {C}}(R): \pi (x,t)\ge \kappa k_R\}\), and

$$\begin{aligned} \delta _0(M_0,f(2R))=\Big (\frac{c}{M_0(1+f(2R))}\Big )^\frac{9}{4}. \end{aligned}$$

Proof

Here, we follow arguments of the paper [18]. They are based on the identity:

$$\begin{aligned}&\int \limits _{Q}(-\pi \partial _t\eta - \pi \Delta \eta -(v+2x'/|x'|^2)\cdot \nabla \eta \pi )dxdt \nonumber \\&\quad =4\pi _0 \int \limits _{-1}^0\int \limits _{-1}^1 \pi (0,x_3,t)\eta (0,x_3,t) dx_3dt, \end{aligned}$$
(3.3)

which is valid for any non-negative test function \(\eta \) supported in Q. Here, \(\pi _0=3.14...\). Although a similar statement has been proven in [18] under the assumption that \(\pi \) is Lipschitz, it remains to be true for functions \(\pi \) from the class \({\mathcal {V}}_0\) as well. Indeed, take a smooth cut-off function \(\psi =\psi (x')\) so that \(\psi (x')=\Psi (|x'|)\), \(0\le \psi \le 1\), \(\psi (x')=0\) if \(|x'|\le \varepsilon /2\), \(\psi (x')=1\) if \(|x'|\ge \varepsilon \), \(\Psi '(\varrho )\le c/\varrho \) and \(\Psi ''(\varrho )\le c/\varrho ^2\) for some positive constant c. Then, it follows from (2.1) that:

$$\begin{aligned} \int \limits _{Q} \Big (\pi \partial _t(\eta \psi )+\pi (u+b)\cdot \nabla (\eta \psi )+\pi \Delta (\eta \psi )\Big )dz=0. \end{aligned}$$

There are two difficult terms for passing to the limit as \(\varepsilon \rightarrow 0\). The first one is as follows:

$$\begin{aligned} I_1:=\int \limits _{Q}\pi \eta \Delta \psi dx dt=J_1+J_2, \end{aligned}$$

where

$$\begin{aligned} J_1:=\int \limits _{Q}(\pi \eta -(\pi \eta )|_{x'=0})\Delta \psi dx dt, \end{aligned}$$

For \(J_2\), we find

$$\begin{aligned} J_2:=\int \limits _{Q}(\pi \eta )|_{x'=0}\Delta \psi dx dt=\int \limits ^0_{-1}\int \limits ^1_{-1}(\pi \eta )|_{x'=0}dx_3dt\int \limits _{|x'|<1}\Delta \psi (x')dx' \end{aligned}$$

and

$$\begin{aligned} \int \limits _{|x'|<1}\Delta \psi (x')dx'=2\pi _0\int \limits ^\varepsilon _\frac{\varepsilon }{2}\frac{1}{\varrho }\frac{\partial }{\partial \varrho }\Big (\varrho \Psi '(\varrho ) \Big )\varrho d\varrho =2\pi _0\varrho \Psi '(\varrho )\Big |^\varepsilon _\frac{\varepsilon }{2}=0. \end{aligned}$$

Now, we wish to show that

$$\begin{aligned} J_1:=\int \limits _{Q}\xi \Delta \psi dx dt \rightarrow 0 \end{aligned}$$

as \(\varepsilon \rightarrow 0\), where, \(\xi :=\pi \eta -(\pi \eta )|_{x'=0}.\) To this end, let us introduce the function

$$\begin{aligned} H_\varepsilon (x_3,t):=\int \limits _{\frac{\varepsilon }{2}<\varrho <\varepsilon }\xi \Delta \psi dx'. \end{aligned}$$

It can be bounded from above and from below

$$\begin{aligned} |H_\varepsilon (x_3,t)|\le c \sup \limits _{\mathrm{spt} \eta }\pi \sup \limits _{|x'|<1}\eta (x',x_3,t)\frac{1}{\varepsilon ^2}\int \limits _{\frac{\varepsilon }{2}}^\varepsilon \varrho d\varrho =:h(x_3,t) \end{aligned}$$

provided \(\varepsilon <1\). The function h is supported in \(]-1,1[ \times ]-1,0[\) and thus

$$\begin{aligned} \int \limits _{-1}^1\int \limits ^0_{-1}h(x_3,t)dx_3dt<\infty . \end{aligned}$$

Now, let \((0,x_3,t)\) be a regular point of \(\pi \), i.e., \((0,x_3,t)\notin S^\pi \). Then, \(\xi (x',x_3,t)\rightarrow 0\) as \(|x'|\rightarrow 0\) and thus for any \(\delta >0\) there exists a number \(\tau (x_3,t)>0\) such that \(|\xi (x',x_3,t)|<\delta \) provided \(|x'|<\tau \). So,

$$\begin{aligned} |H_\varepsilon (x_3,t)|<c\frac{\delta }{\varepsilon ^2}\int \limits ^\varepsilon _\frac{\varepsilon }{2}\varrho d\varrho =c\frac{\delta }{2} \end{aligned}$$

provided \(\varepsilon <\tau \). Therefore, \(H_\varepsilon (x_3,t)\rightarrow 0\) as \(\varepsilon \rightarrow 0\) and by the Lebesgue theorem on dominated convergence, we find that

$$\begin{aligned} J_1=\int \limits _{-1}^1\int \limits _{-1}^0H_\varepsilon (x_3,t)dx_3dt\rightarrow 0 \end{aligned}$$

as \(\varepsilon \rightarrow 0\).

Similar arguments work for the second difficult term:

$$\begin{aligned} I:=\int \limits _{Q}\pi \eta b\cdot \nabla \psi dz=J_1+J_2, \end{aligned}$$

where

$$\begin{aligned} J_1=\int \limits _{Q}\xi b\cdot \nabla \psi dz \end{aligned}$$

and

$$\begin{aligned}&J_2:=\int \limits _{Q}(\pi \eta )|_{x'=0}b\cdot \nabla \psi dx dt=\int \limits ^0_{-1}\int \limits ^1_{-1}(\pi \eta )|_{x'=0}dx_3dt2\pi _0\int \limits _\frac{\varepsilon }{2}^\varepsilon \frac{2}{\varrho }\Psi '(\varrho )\varrho d\varrho \\&\quad =4\pi _0\int \limits ^0_{-1}\int \limits ^1_{-1}(\pi \eta )|_{x'=0}dx_3dt. \end{aligned}$$

The fact that \(J_1\rightarrow 0\) as \(\varepsilon \rightarrow 0\) can be justified in the same way as above, replacing \(H_\varepsilon \) with the function

$$\begin{aligned} G_\varepsilon (x_3,t):=\int \limits _{\frac{\varepsilon }{2}<|\ x'| <\varepsilon }\xi b\cdot \nabla \psi dx'. \end{aligned}$$

Other terms can be treated in a similar way and even easier. So, the required identity (3.3) has been proven.

Now, let us select the test function \(\eta \) in (3.3), using the following notation

$$\begin{aligned} Q^{\lambda ,\theta }(z_0,R):={\mathcal {C}}(x_0,\lambda R)\times ]t_0-\theta R^2,t_0[, \end{aligned}$$

so that \(\eta =1\) in \(Q^{\frac{1}{2},\frac{1}{8}}((0,-\frac{13}{16}R^2),R)\), \(\eta =0\) out of \(Q^{1,\frac{1}{4}}((0,-\frac{3}{4}R^2),R)\) and \(|\partial _t\eta |+|\nabla \eta |^2+|\nabla ^2\eta |\le c/R^2\). Taking into account that \(\pi \) has the property \((\mathcal B_R)\), we find

$$\begin{aligned} \frac{\pi _0}{2}k_RR^2\le \frac{c}{R^2}\int \limits _{Q^{1,\frac{1}{4}}(z_R,R)}\pi dz+\frac{c}{R}\int \limits _{Q^{1,\frac{1}{4}}(z_R,R)}\pi |v|dz+\frac{c}{R}\int \limits _{Q^{1,\frac{1}{4}}(z_R,R)}\frac{\pi }{|x'|}dz, \end{aligned}$$

where \(z_R=(0,-\frac{3}{4}R^2)\).

Setting \( E_\kappa =\{(x,t): t\in ]-R^2,-\frac{3}{4}R^2[, x\in e_\kappa (t)\},\) we can deduce from the latter inequality

$$\begin{aligned}&\frac{\pi _0}{2}k_RR^3 \le \frac{c}{R^2}\int \limits _{Q^{1,\frac{1}{4}}(z_R,R)\setminus E_\kappa }\pi dz+\frac{c}{R}\int \limits _{Q^{1,\frac{1}{4}}(z_R,R)\setminus E_\kappa }\pi |v|dz+\frac{c}{R}\int \limits _{Q^{1,\frac{1}{4}}(z_R,R)\setminus E_\kappa }\frac{\pi }{|x'|}dz \\&\quad + \frac{c}{R^2}\int \limits _{Q^{1,\frac{1}{4}}(z_R,R)\cap E_\kappa }\pi dz+\frac{c}{R}\int \limits _{Q^{1,\frac{1}{4}}(z_R,R)\cap E_\kappa }\pi |v|dz+\frac{c}{R}\int \limits _{Q^{1,\frac{1}{4}}(z_R,R)\cap E_\kappa }\frac{\pi }{|x'|}dz. \end{aligned}$$

Applying (3.1) and recalling definitions of the sets \(e_\kappa (t)\) and \(E_\kappa \), we can get

$$\begin{aligned}&\frac{\pi _0}{2}k_RR^3 \le \frac{c\kappa k_R}{R^2}\Big \{|Q^{1,\frac{1}{4}}(R)| +R\int \limits _{Q^{1,\frac{1}{4}}(z_R,R)\setminus E_\kappa } |v|dz+R\int \limits _{Q^{1,\frac{1}{4}}(z_R,R)\setminus E_\kappa }\frac{1}{|x'|}dz\Big \} \\&\quad +\frac{cM_0k_R}{R^2}\Big \{|E_\kappa |+R\int \limits _{Q^{1,\frac{1}{4}}(z_R,R)\cap E_\kappa } |v|dz+R\int \limits _{Q^{1,\frac{1}{4}}(z_R,R)\cap E_\kappa }\frac{1}{|x'|}dz\Big \}. \end{aligned}$$

We need to estimate integrals in the above inequality. First, for integrals, containing v, Holder inequality gives

$$\begin{aligned}&\int \limits _{Q^{1,\frac{1}{4}}(z_R,R)\setminus E_\kappa }|v|dx\le \Vert {\mathbb {I}}\Vert _{\frac{3}{2},\frac{4}{3},Q^{1,\frac{1}{4}}(R)}\left( \int \limits ^{-\frac{3}{4}R^2}_{-R^2}\left( \int \limits _{{\mathcal {C}}(R)}|v|^3dx\right) ^\frac{4}{3}dt\right) ^\frac{1}{4} \\&\quad \le f(2R)R^\frac{1}{2}\Vert {\mathbb {I}}\Vert _{\frac{3}{2},\frac{4}{3},Q^{1,\frac{1}{4}}(R)}\le f(2R)R^4 \end{aligned}$$

and similarly

$$\begin{aligned} \int \limits _{Q^{1,\frac{1}{4}}(z_R,R)\cap E_\kappa } |v|dz\le f(2R)R^\frac{1}{2}\Vert {\mathbb {I}}\Vert _{\frac{3}{2},\frac{4}{3},E_\kappa }. \end{aligned}$$

To evaluate the last two integrals, let us take into account the fact:

$$\begin{aligned} \frac{1}{|x'|}\in L_{\frac{9}{5},\infty }(Q^{1,\frac{1}{4}}(z_R,R)). \end{aligned}$$

Then,

$$\begin{aligned}&\int \limits _{Q^{1,\frac{1}{4}}(z_R,R)\setminus E_\kappa }\frac{1}{|x'|}dz\le \Vert \frac{1}{|x'|}\Vert _{\frac{9}{5},\infty ,Q^{1,\frac{1}{4}}(z_R,R)}\Vert {\mathbb {I}}\Vert _{\frac{9}{4},1,Q^{1,\frac{1}{4}}(R)} \\&\quad \le cR^\frac{2}{3}R^\frac{10}{3}=cR^4 \end{aligned}$$

and

$$\begin{aligned} \int \limits _{Q^{1,\frac{1}{4}}(z_R,R)\cap E_\kappa }\frac{1}{|x|}dz\le c R^\frac{2}{3}\Vert {\mathbb {I}}\Vert _{\frac{9}{4},1,E_\kappa }. \end{aligned}$$

Hence, we have

$$\begin{aligned}&\frac{\pi _0}{2}k_RR^3\le c\kappa k_RR^3(1+f(2R)) \\&\quad +\frac{cM_0k_R}{R^2}\Big [|E_\kappa |+f(2R)R^\frac{3}{2}\Vert {\mathbb {I}}\Vert _{\frac{3}{2},\frac{4}{3},E_\kappa }+R^\frac{5}{3}\Vert {\mathbb {I}}\Vert _{\frac{9}{4},1,E_\kappa }\Big ]. \end{aligned}$$

So,

$$\begin{aligned} \frac{\pi _0}{2}\le c\kappa (1+f(2R)) +\frac{cM_0}{R^5}\Big [|E_\kappa |+f(2R)R^\frac{3}{2}\Vert \mathbb I\Vert _{\frac{3}{2},\frac{4}{3},E_\kappa }+R^\frac{5}{3}\Vert {\mathbb {I}}\Vert _{\frac{9}{4},1,E_\kappa }\Big ]. \end{aligned}$$

Now, one can find \(\kappa =\kappa _0(f(2R))=c/(1+f(2R))\) such that

$$\begin{aligned} \frac{cM_0}{R^5}\Big [|E_{\kappa _0}|+f(2R)R^\frac{3}{2}\Vert {\mathbb {I}}\Vert _{\frac{3}{2},\frac{4}{3},E_{\kappa _0}}+R^\frac{5}{3}\Vert {\mathbb {I}}\Vert _{\frac{9}{4},1,E_{\kappa _0}}\Big ]\ge 1. \end{aligned}$$

It remains to estimate two integrals on the left hand side of the latter inequality:

$$\begin{aligned} \Vert {\mathbb {I}}\Vert _{\frac{3}{2},\frac{4}{3},E_{\kappa _0}}=\left( \int \limits ^{-\frac{3}{4}R^2}_{-R^2}|e_\kappa (t)|^\frac{8}{9}dt\right) ^\frac{3}{4}\le c|E_{\kappa _0}|^\frac{2}{3}R^\frac{1}{6} \end{aligned}$$

and

$$\begin{aligned} \Vert {\mathbb {I}}\Vert _{\frac{9}{4},1,E_\kappa }\le c|E_{\kappa _0}|^\frac{4}{9}R^\frac{10}{9}. \end{aligned}$$

Letting \(A=|E_{\kappa _0}|/R^5\), we arrive at the following inequality

$$\begin{aligned} f(A):=A+A^\frac{4}{9}+f(2R)A^\frac{2}{3}\ge \frac{1}{cM_0}. \end{aligned}$$

Since \(f'(A)>0\) for \(A>0\), we can state that the last inequality implies

$$\begin{aligned} \frac{|E_{\kappa _0}|}{|{\mathcal {C}}(R)|\frac{1}{4}R^2}\ge \delta _0=\Big (\frac{c}{M_0(1+f(2R))}\Big )^\frac{9}{4}. \end{aligned}$$

It is not so difficult to show the exisence of \({{\overline{t}}}\in [-R^2,-\frac{3}{4}R^2]\) with the property:

$$\begin{aligned} |e_{\kappa _0}({{\overline{t}}})|\frac{1}{4}R^2\ge |E_{\kappa _0}|. \end{aligned}$$

So, it is proven that there exists \({{\bar{t}}}\in [-R^2,-3R^2/4]\) such that

$$\begin{aligned} |\{x\in {\mathcal {C}}(R): \pi (x,{{\bar{t}}})>\kappa _0 k_R\}|\ge \delta _0 |{\mathcal {C}}(R)|, \end{aligned}$$
(3.4)

which completes the proof of the lemma. \(\square \)

Now, we are able to prove Proposition 1.4.

Assume that the function \(\pi \) meets all the conditions of Lemma 3.1 and according to it, we can claim that:

$$\begin{aligned} |e_{\kappa _0}({{\overline{t}}})|=|\{x\in {\mathcal {C}}(R): \pi (x,{{\overline{t}}})\ge \kappa _0 k_R\}|\ge \delta _0|{\mathcal {C}}(R)| \end{aligned}$$

for some \({{\overline{t}}}\in [-R^2,-\frac{3}{4}R^2]\), \(\kappa _0=c/g(2R)\), and \(\delta _0=c(M_0)/g^\frac{9}{4}(2R)\). Now, we can calculate

$$\begin{aligned}&\theta (\delta _0(M_0,f(2R)),f(2R))\ge c\Big (\frac{\delta _0^6}{1+\delta _0^2f(2R)}\Big )^\frac{4}{3} \\&\quad \ge c(M_0)\Big (\frac{1}{g(2R)}\Big )^{18}, \end{aligned}$$

apply Lemma 2.5, and find

$$\begin{aligned} |\{\pi (\cdot ,t)\ge \delta _0\kappa _0k_R/3\}\cap {\mathcal {C}}(R)|>\delta _0/3|{\mathcal {C}}(R)| \end{aligned}$$

for all \(t\in [{{\overline{t}}},t_0]\) with \(t_0={{\overline{t}}}+\theta _0R^2\) and \(\theta _0=c(M_0)(g(2R))^{-18}\).

Next, it follows from Lemma 2.7 that:

$$\begin{aligned} |\{\pi <2^{-s}\delta _0\kappa _0k_R/3\}\cap Q^{1,\theta _0}((0,t_0),R)|\le \mu _* |Q^{1,\theta _0}(R)|, \end{aligned}$$

where

$$\begin{aligned} s=\mathrm{entier}\Big (\frac{c}{\delta _0^2\mu _*^2\theta _0}(1+f(2R))\Big )+1 \end{aligned}$$

and \(\mu _*\) is the number that appears in Corollary 2.4, see also Proposition 2.2. In our case,

$$\begin{aligned} \mu _*=\Big (\frac{1}{2c_1(3/4,1,\theta _0/2,\theta _0,M(2R))}\Big )^\frac{10}{3} \end{aligned}$$

and, moreover

$$\begin{aligned} c_1(3/4,1,\theta _0/2,\theta _0,M(2R))\le c\theta _0^{-\frac{3}{2}}g^3(2R)\le c(M_0)(g(2R))^{30}. \end{aligned}$$

Then, Corollary 2.4 implies the bound

$$\begin{aligned} \pi \ge 2^{-s}\delta _0\kappa _0k_R/6={\hat{\beta }}_2\kappa _0k_R \end{aligned}$$

in \(Q^{\frac{3}{4},\frac{1}{2}\theta _0}((0,t_0),R)\). So, combining previous estimates, we find the following:

$$\begin{aligned} {\hat{\beta }}_2=\frac{1}{6}2^{-s}\delta _0\ge e^{-sln2-\ln 6}\delta _0\ge e^{-cs}\delta _0, \end{aligned}$$

where

$$\begin{aligned}&s\le \frac{2g(2R)}{\delta _0^2\mu _*^2\theta _0}\le c(M_0)g(2R)(g(2R))^\frac{9}{2}(g(2R))^{18}{c_1}^\frac{20}{3} \\&\quad \le c(M_0)(g(2R))^\frac{47}{2}(g(2R))^{30})^\frac{20}{3}\le c(M_0) (g(2R))^{224}. \end{aligned}$$

So,

$$\begin{aligned}&{\hat{\beta }}_2\ge e^{-c(M_0)(g(2R))^{224}}c(M_0)(g(2R))^{-\frac{9}{4}}\ge e^{-2c(M_0)(g(2R))^{224}} \\&\quad \ge e^{-c(M_0,c_*)\ln ^{224\alpha }\sqrt{\ln \frac{1}{R}}}. \end{aligned}$$

Obviously, there exists a number \(0<R_{*5}(M_0,c_*,\alpha )\le \min \{1/6,R_{*2}\}\) such that

$$\begin{aligned} 2c(M_0,c_*)\ln ^{224\alpha -1}\sqrt{\ln \frac{1}{R}}\le \frac{1}{2} \end{aligned}$$

and

$$\begin{aligned} c(M_0,c_*)\ln ^{224\alpha }\sqrt{\ln \frac{1}{R}}\ge \ln g(2R) \end{aligned}$$

for \(0<R\le R_{*5}(M_0,c_*,\alpha )\) and thus

$$\begin{aligned}&-c(M_0,c_*)\ln ^{224\alpha }\sqrt{\ln \frac{1}{R}} \\&\quad =-2c(M_0,c_*)\ln ^{224\alpha }\sqrt{\ln \frac{1}{R}}++c(M_0,c_*)\ln ^{224\alpha }\sqrt{\ln \frac{1}{R}} \\&\quad -\ge \Big (2c(M_0,c_*)\ln ^{224\alpha -1}\sqrt{\ln \frac{1}{R}}\Big )\ln \sqrt{\ln \frac{1}{R}}+\ln g(2R) \\&\quad \ge -\ln \Big (\ln \frac{1}{R}\Big )^\frac{1}{4}+\ln g(2R) . \end{aligned}$$

Now, the number \({\hat{\beta }}_2\) is estimated as follows:

$$\begin{aligned} {\hat{\beta }}_2\ge \Big (\ln \frac{1}{R} \Big )^{-\frac{1}{4}}g(2R) \end{aligned}$$
(3.5)

for \(0<R\le R_{*5}(M_0,c_*,\alpha )\).

Since

$$\begin{aligned} -R^2\le \overline{t}+\theta _0/2R^2=t_0-\theta _0/2R^2<t_0={{\overline{t}}}+\theta _0R^2\le -\frac{3}{4}R^2+\frac{1}{4}R^2=-\frac{1}{2}R^2, \end{aligned}$$

there is \({{\overline{t}}}_1\in [-R^2,-\frac{1}{2}R^2]\) such that

$$\begin{aligned} \pi (\cdot ,{{\overline{t}}}_1)>{\hat{\beta }}_2\kappa _0k_R \end{aligned}$$

in \(\mathcal C(\frac{3}{4}R)\). It allows us to apply Lemma 2.9 with \(\theta =1/2\), with \(\frac{3}{4}R\) instead of R, with \({{\overline{t}}}_1\) instead of \({{\overline{t}}}\), and with \({\hat{\beta }}_2\kappa _0k_R\) instead of k. According to Lemma 2.9, the inequality

$$\begin{aligned} \pi \ge \beta _0{\hat{\beta }}_2\kappa _0k_R \end{aligned}$$

holds in Q(R/2). It follows from Lemma 2.9 and from (3.5) that

$$\begin{aligned} \pi \ge \frac{c k_R}{\ln ^\frac{3}{4} (\frac{1}{R})} =\beta (2R)k_R \end{aligned}$$

in Q(R/2).

By our assumption imposed on function \(\sigma \), we can put \(k_R=\frac{1}{2}\mathrm{osc}_{z\in Q(2R)}\sigma (z)\). Then, either \(\pi =\sigma -m_{2R}\) or \(\pi =M_{2R}-\sigma (z)\) satisfies all the conditions of the proposition with \(M_0=2\). Simple arguments show that

$$\begin{aligned} \mathrm{osc}_{z\in Q(R/2)}\sigma (z)\le \Big (1-\frac{1}{2}\beta (2R)\Big )\mathrm{osc}_{z\in Q(2R)}\sigma (z). \end{aligned}$$

Now, after iterations of the latter inequality, we arrive at the following bound

$$\begin{aligned}&\mathrm{osc}_{z\in Q(R/2^{2k+1})}\le \prod \limits ^{k}_{i=0}\Big ((1-\frac{1}{2}\beta (R/2^{2i-1})\Big )\mathrm{osc}_{z\in Q(2R)}\sigma (z)= \\&\quad =\eta _k\mathrm{osc}_{z\in Q(2R)}\sigma (z) \end{aligned}$$

being valid for any natural number k.

In order to evaluate \(\eta _k\), take \(\ln \) of it. As a result,

$$\begin{aligned}&\ln \eta _k=\sum \limits ^k_{i=0}\ln \Big ((1-\frac{1}{2}\beta (R/2^{2i-1})\Big )\le -\sum \limits ^k_{i=0}\frac{1}{2}\beta (R/2^{2i-1}) \\&\quad \le -c\sum \limits ^k_{i=0}(\ln ( 2^{2i-1}/R))^{-\frac{3}{4}} =-c\sum \limits ^k_{i=0}\frac{1}{(i\ln 4+\ln (1/{(2R)}))^\frac{3}{4} } \\&\quad \le -c\int \limits ^{k+1}_0\frac{dx}{(x\ln 4+\ln ( 1/(2R)) )^\frac{3}{4}} \\&\quad =-\frac{4c}{\ln 4}(x\ln 4+\ln (1/(2R)))^\frac{1}{4}\Big |^{k+1}_0 \\&\quad =-c\Big (\ln ^\frac{1}{4}(2^{2k+1}/R) -\ln ^\frac{1}{4}(1/(2R))\Big ) \Big ). \end{aligned}$$

Hence, we have

$$\begin{aligned} \mathrm{osc}_{z\in Q(R/2^{2k+1})}\le e^{-c\Big [\ln ^\frac{1}{4}(2^{2k+1}/R) -\ln ^\frac{1}{4}(1/(2R))\Big ]} \mathrm{osc}_{z\in Q(2R)}\sigma (z) \end{aligned}$$

and thus

$$\begin{aligned} \mathrm{osc}_{z\in Q(r)}\le e^{-c\Big [\ln ^\frac{1}{4}(1/(2r)) -\ln ^\frac{1}{4}(1/(2R))\Big ]} \mathrm{osc}_{z\in Q(2R)}\sigma (z) \end{aligned}$$

for all \(0<r<R\le R_*(c_*,\alpha )= R_{*5}(2,c_*,\alpha )\). So, (1.8) follows. The proof of Proposition 1.4 is complete.

4 Proof of Theorem 1.3

By the maximimum principle, we have \(|\sigma | =|\varrho v_\varphi |\le \Sigma _0\) in \({\mathbb {R}}^3\times ]0,T[\). From Proposition 1.4, it follows that

$$\begin{aligned} |\sigma (\varrho ,x_3,t)|\le e^{-c\Big [\ln ^\frac{1}{4}(1/(2\varrho )) -\ln ^\frac{1}{4}(1/(2R_*))\Big ]} 2\Sigma _0 \end{aligned}$$

fo all \(0<\varrho \le R_*(c_*,\alpha )\), for all \(x_3\in {\mathbb {R}}\), and for \(t\in ]T-R_*^2,T[\). Obviously, it remains true \(\varrho >R_*\) as well. So, we have for all \(x\in {\mathbb {R}}^3\) and for all \(t\in ]T-R_*^2,T[\)

$$\begin{aligned} |\sigma (\varrho ,x_3,t)|\le C(c_*,\alpha )e^{-c\ln ^\frac{1}{4}(1/(2\varrho ))}\Sigma _0\le C(c_*,\alpha )\Sigma _0\frac{m!}{c^m\ln ^{\frac{m}{4}}(1/(2\varrho ))} \end{aligned}$$

for all natural numbers m.

Now, let us notice that \(v(\cdot , T-R_*^2)\in H^2\). Therefore, one can use the results of papers [12, 30], and [15], see also [17, 4], on the Cauchy problem for the Navier–Stokes system (1.1) in \({\mathbb {R}}^3\times ]T-R_*^2,T[\) and conclude that v is a strong solution in the interval ]0, T[.