1 Introduction

In this paper, we study the local boundedness and Hölder regularity of solutions to the inhomogeneous equation

$$\begin{aligned} u_t+ (-\Delta _p)^s u=f(x,t), \end{aligned}$$
(1.1)

where \(f \in L^r_{\textrm{loc}}(I; L^q_{\textrm{loc}}(\Omega ))\) with \(q\ge 1\), \(r\ge 1\), \(p\ge 2\) and \(s \in (0,1)\). Here, \( (- \Delta _p)^s \) is the fractional p-Laplacian, arising as the first variation of the Sobolev–Slobodeckiĭ seminorm

$$\begin{aligned} (-\Delta _p)^s u (x):=2\, \mathrm {P.V.} \int _{{\mathbb {R}}^n}\frac{|u(x)-u(y)|^{p-2}(u(x)-u(y))}{|x-y|^{n+s\,p}} \;\textrm{d}y. \end{aligned}$$

Nonlocal equations involving operators of the above type, with a singular kernel, were first considered in [31] to the best of our knowledge.

In this study, continuing the work in [7], we perform a perturbative argument to obtain Hölder continuity estimates, with explicit exponents for the equations with a right-hand side. Our approach closely follows the arguments in [47] and [6]. In such perturbative arguments, it is often possible to establish Hölder regularity results for bounded solutions using only \(L^\infty \) estimates for the equations with zero right-hand side. Here, to estimate the Hölder seminorms of certain functions in the proof of Theorem 1.2 as well as to prove Theorem 3.6, we are led to prove Proposition 3.4. As a by-product, by combining Proposition 3.4 with the existing local boundedness results we obtain an \(L^\infty \) bound for equations with right-hand sides. This is Theorem 1.1. The proof is inspired by the work [5].

Below, we state the main results. For the definition of the tail and relevant function spaces, see Sect. 2. We use the following notation of parabolic cylinders

$$\begin{aligned} Q_{R,r}(x,T):= B_R(x_0) \times (T-r,T]. \end{aligned}$$

The exponent \(p_s^\star = \frac{np}{n-sp}\) is the critical exponent for the Sobolev embedding theorem, see Proposition 2.5. We denote by \(p^\prime \), the Hölder conjugate of p, that is \(p^\prime = \frac{p}{p-1}\).

Theorem 1.1

Let \(\Omega \subset {\mathbb {R}}^n\) be a bounded and open set, \(I=(t_0,t_1]\), \(p\ge 2\), \(0<s<1\). Consider q and r such that

$$\begin{aligned} \frac{1}{r} + \frac{n}{spq} < 1. \end{aligned}$$

In addition, assume that \(r\ge p^\prime \),

$$\begin{aligned} q \ge 1 \quad \text {if } sp\ne n, \quad \text {and}\quad q>1 \quad \text {if } sp=n. \end{aligned}$$

Suppose u is a local weak solution of

$$\begin{aligned} u_t+(-\Delta _p)^s u=f \qquad \text{ in } \Omega \times I, \end{aligned}$$

such that

$$\begin{aligned} u\in L^p_{\textrm{loc}} (I; L_{sp}^{p-1}({\mathbb {R}}^n)) \quad \text {and} \quad f \in L^{r}_{\textrm{loc}}(I; L^{q}_{\textrm{loc}}(\Omega )). \end{aligned}$$

then u is locally bounded in \(\Omega \). More specifically, if \(Q_{2R,(2R)^{sp}(x_0,T_0)} \Subset \Omega \times I\), u is bounded in \(Q_{R/2,(R/2)^{sp}}(x_0,T_0)\), and in the case \(sp \ne n \), the estimate reads

where \(C=C(n,s,p)\), \(\nu = 1-\frac{1}{r} - \frac{n}{spq}\) and \( \vartheta = 1+\frac{sp\nu }{n} \).

In the case \(sp=n \), given any l such that \( \frac{p}{r^\prime }(1-\frac{1}{r} - \frac{1}{q})^{-1}<l < \infty \) we get

where \(C= C(n,s,p,q,l) \), \(\vartheta = 2-\frac{1}{r}- \frac{1}{q} - \frac{p}{lr^\prime }\) and \(\nu = 1-\frac{1}{r} - \frac{1}{q}\).

Theorem 1.2

Let \(\Omega \subset {\mathbb {R}}^n\) be a bounded and open set, \(I=(t_0,t_1]\), \(p\ge 2\), \(0<s<1\). Consider q and r such that

$$\begin{aligned} \frac{1}{r} + \frac{n}{spq} < 1. \end{aligned}$$

In addition, assume that \(r\ge p^\prime \),

$$\begin{aligned} q \ge 1 \quad \text {if } sp\ne n, \quad \text {and}\quad q>1 \quad \text {if } sp=n. \end{aligned}$$

Define the exponent

$$\begin{aligned} \Theta (s,p):=\left\{ \begin{array}{rl} \dfrac{s\,p}{p-1},&{} \text{ if } s<\dfrac{p-1}{p},\\ &{}\\ 1,&{} \text{ if } s\ge \dfrac{p-1}{p}. \end{array} \right. \end{aligned}$$
(1.2)

Suppose u is a local weak solution of

$$\begin{aligned} u_t+(-\Delta _p)^s u=f \qquad \text{ in } \Omega \times I, \end{aligned}$$

such that

$$\begin{aligned} u\in L^{\infty }_{\textrm{loc}}(I;L^\infty _{\textrm{loc}}(\Omega )) \cap L^\infty _{\textrm{loc}} (I; L_{sp}^{p-1}({\mathbb {R}}^n)), \quad \text {and} \quad f \in L^{r}_{\textrm{loc}}(I; L^{q}_{\textrm{loc}}(\Omega )). \end{aligned}$$

Then

$$\begin{aligned}{} & {} u\in C^{\alpha }_{x,\textrm{loc}}(\Omega \times I)\cap C^{\frac{\alpha }{sp- (p-2)\alpha }}_{t,\textrm{loc}}(\Omega \times I),\quad \text{ for } \text{ every } 0<\alpha \\{} & {} \quad \le \frac{r(spq-n)-spq}{q(r(p-1)-(p-2))} \text { such that } \alpha < \Theta . \end{aligned}$$

More precisely, given \(\alpha < \Theta \) satisfying

$$\begin{aligned} \alpha \le \frac{r(spq-n)-spq}{q(r(p-1)-(p-2))}, \end{aligned}$$

for every \(R>0\), \(x_0\in \Omega \) and \(T_0\) such that

$$\begin{aligned} Q_{R,2R^{s\,p}}(x_0,T_0)\Subset \Omega \times (t_0,t_1], \end{aligned}$$

there exists a constant \(C=C(n,s,p, q,r,\alpha )>0\) such that

$$\begin{aligned} \begin{aligned} |u(x_1,t_1)-&u(x_2,t_2)| \le C \,\left[ {\mathcal {M}}\Bigl (\frac{|x_2-x_1|}{R} \Bigr )^\alpha + {\mathcal {M}}^{p-1}\Bigl ( \frac{|t_2-t_1|}{R^{s\, p}} \Bigr )^{\frac{\alpha }{sp-(p-2)\alpha }} \right] \end{aligned}\nonumber \\ \end{aligned}$$
(1.3)

for any \((x_1,t_1),\,(x_2,t_2)\in Q_{R/2,(R/2)^{s\,p}}(x_0,T_0)\), with

$$\begin{aligned} {\mathcal {M}}= & {} 1+ \Vert u\Vert _{L^\infty (Q_{R,2R^{sp}}(x_0,T_0))} + \sup _{ T_0- 2R^{sp}\le t \le T_0}\textrm{Tail}_{p-1,sp} (u(\,\bullet \,,t);x_0,R) \\{} & {} + \bigl ( R^{sp-\frac{n}{q} - \frac{sp}{r}} \Vert f\Vert _{L^{q,r}(Q_{R,2R^{sp}}(x_0,T_0))} \bigr )^{\frac{1}{1+ \frac{p-2}{r^\prime }}}. \end{aligned}$$

1.1 Known results

Recently, there has been a growing interest in nonlocal problems of both elliptic and parabolic types. For studies of fractional p-Laplace operators with different (continuous) kernels, see [4]. Parabolic equations of the type (1.1) were first considered in [42] with a slightly different diffusion operator. See also [1, 39, 48] and [49] for studies of the existence, uniqueness and long time behavior of solutions.

A noteworthy area of investigation has been devoted to adapting the classical De Giorgi–Nash–Moser theory for nonlocal equations. Local boundedness, Hölder estimates and Harnack inequalities have been established in the elliptic case under general assumptions on the kernels; see, for instance, [19, 23, 24, 32].

Here we seize the opportunity to mention [16,17,18] and [51] which contain regularity results for parabolic nonlocal equations.

Local boundedness for parabolic nonlocal equations has been studied, for instance, in [11, 22, 33, 45]. In particular, the local boundedness of the solutions to equations modeled on (1.1) with zero right-hand side was obtained in [45]. The results concern operators of the form

$$\begin{aligned} L_K =P.V. \int _{R^n} K(x,y,t)|u(x)-u(y)|^{p-2}(u(x)-u(y)) \;\textrm{d}y, \end{aligned}$$
(1.4)

where K is a measurable kernel, which is symmetric in the space variables and satisfies the ellipticity condition

$$\begin{aligned} \frac{\Lambda ^{-1}}{|x-y|^{n+sp}} \le K(x,y,t) \le \frac{\Lambda }{|x-y|^{n+sp}}. \end{aligned}$$

Later in [22], local boundedness for certain right-hand sides of the form f(xtu) was established. See also [3] for a recent boundedness result in the setting of nonlocal kinetic Kolmogorov–Fokker–Planck equations. All the aforementioned local boundedness results have a particular unnatural assumption, \(u\in L^\infty (I;L_{sp}^{p-1}({\mathbb {R}}^n))\). It is more natural to assume \(u\in L^{p-1}(I;L_{sp}^{p-1}({\mathbb {R}}^n))\). This difficulty has been completely resolved in [34] when \(p=2\) and generalizes to the nonlinear setting in [10].

[46] contains a Harnack inequality for nonlinear parabolic equations with zero right-hand side, see also [34] for a full Harnack inequality with optimal tail assumption for \(p=2\). Hölder regularity has also been established in [13, 27] for \(p=2\) and for locally bounded solutions in [2] and [37] for all \(1<p<\infty \) for equations with zero right-hand side.

The question of higher regularity of solutions to nonlocal equations has also been a subject of intensive study during the past few years. For instance, see [28, 43] for a nonlocal Schauder-type theory. We also refer to [14, 15] for nonlocal analogs of Krylov–Safanov and Evans–Krylov theorems. We refer to [6, 7, 11, 12, 26, 40, 41] for studies of higher regularity in the variational setting. In particular, in [7] they prove Hölder continuity of the solutions to (1.1) with explicit exponents (for \(f=0\) and \(K= |x-y|^{-n-sp}\)). Recently in [29], the same type of result has been established for nonlocal equations with double phase that is for diffusion operators involving two different degrees of homogeneity and differentiability.

Perturbative arguments have been very successful in obtaining sharp boundedness and Hölder regularity estimate at least in the elliptic setting, see, for instance, [25, 35]. See also [36] for an overview of the local theory. In this study, continuing the work in [7], we perform a perturbative argument to obtain Hölder continuity estimates with explicit exponents for equations with a right-hand side. However, we have to say that the current work has some unnatural assumptions that have yet to be overcome.

1.1.1 Discussion of the results and comparison to some previous works

Our results contain an unnatural assumption \(r\ge p^\prime \), as well as the assumption \(u \in L^p(I;L_{sp}^{p-1}({\mathbb {R}}^n))\) in Theorem 1.1. We use these assumptions in two places. First and foremost these assumptions are needed to ensure the existence of a solution to (3.1), the so-called (sp)-caloric replacement of our solution. This limitation comes from the regularity assumption on the boundary condition in Theorem 2.12 which is essentially the same as [7, Theorem A.3] see Remark 2.13. We also use the assumption \(r\ge p^\prime \) in obtaining the estimates in Lemma 3.2. We believe it is possible to overcome this issue by an interpolation argument, see Remark 3.3. It has to be mentioned that we also use the assumption \(u \in L^p(I;L_{sp}^{p-1}({\mathbb {R}}^n))\) to justify testing the equation with powers of the solution in Appendix B. Having said this, it is reasonable to expect Theorem 1.1 to hold for any weak solution under the assumption

$$\begin{aligned} \frac{1}{r} + \frac{n}{spq} < 1, \end{aligned}$$
(1.5)

as this is the only assumption that appears in the estimates. The same extra assumptions on q and r are present in Theorem 3.6 due to the same reason as in Theorem 1.1. Furthermore, we assume our solutions to have bounded tail in time, that is, \(u\in L^\infty (I; L^{p-1}_{sp}({\mathbb {R}}^n))\). In light of the recent developments in [10, 34], one can actually weaken the assumptions on the tail. In particular, by using [10, Theorem 1.2] instead of [7, Theorem 1.2] in the proof of Theorem 3.6, with some small modifications in the argument one should be able to obtain the Hölder continuity of the solutions under the assumption \(u \in L^p(I;L^{p-1}_{sp}({\mathbb {R}}^n))\), and the same assumptions on qr as in Theorem 3.6. We also believe that it is possible to avoid using Proposition 3.4 in the proof of Theorem 3.6, by using [10, Theorem 1.1 and Theorem 1.2]. As improving upon this assumption does not improve our main result, Theorem 1.2, we do not complicate the article by going through the details of this issue. Furthermore, we actually expect the result to be true under the weaker assumption \(u \in L^l(I;L^{p-1}_{sp}({\mathbb {R}}^n))\) for some \(l>p-1\) and without the assumption \(r\ge p^\prime \), but the current restrictions in the article especially with respect to the existence of the (sp)-caloric replacement do not allow us to obtain such a result.

Let us also mention that the local boundedness and Hölder regularity results mentioned above hold for a more general class of equations with measurable coefficient \(u_t + L_k u =0\), where \(L_k\) is as in 1.4. Although we write our results for the equation \(u_t + (-\Delta _p)^s u = f\), the arguments in the proofs of Theorem 1.1 and Theorem 3.6 can be adapted to the equations \(u_t + L_k u = f\) with measurable, asymmetric, uniformly elliptic coefficients easily. The only difference is that a dependence on the ellipticity coefficients will appear in the constants. But the question of what assumption is needed on the kernel to get higher Hölder regularity is subtle. We refer to [11, 25, 40] for a study of this issue.

The equation \(u_t -\Delta _p u = f\) can be seen as a limit of the equation \(u_t +(1-s)c(n,p)(-\Delta _p)^s u = f\) as \(s \nearrow 1\). A relevant question is whether the estimates provided here in the article are stable with respect to s as \(s \nearrow 1\). We have to admit that we did not keep track of the dependence of the constants on s while writing this article, and we wrote the article for the operator \(\partial _t + (-\Delta _p)^s\) instead of \({\partial _t + (1-s) (-\Delta _p)^s}\). Still, we can say a few words on the dependence of our constants on s for those who might be interested in pursuing this question. The proofs of Theorem 1.1 and Theorem 3.6 are combinations of local boundedness estimates in [10, Theorem 1.1] and the Holder continuity result [7, Theorem 1.2] for the equations with zero right-hand side, together with the comparison estimates of Lemma 3.2 and Proposition 3.4. [7, Theorem 1.2] is stable as \(s \nearrow 1\) see [7, Remark 1.7], as for [10, Theorem 1.1] they did not specify the dependence of their constants on s in their article. In Lemma 3.2 and Proposition 3.4, the dependence of the constants on s comes from the Sobolev and Morrey inequalities. The constants in these inequalities behave like \(s(1-s)\) with respect to s, but it has to be mentioned that we update the constants to be greater than one in several places. It might be the case that if one considers the operator \(\partial _t +(1-s)(-\Delta _p)^s\) instead, the estimates in Lemma 3.2 and Proposition 3.4 would become robust as \(s \nearrow 1\). We cannot specify the dependence of the constant in Theorem 1.2 on s specifically. The main difficulty lies in the proof of Lemma 3.7, which is proved by contradiction.

Now we compare the main results of the article to some other works.

Local boundedness and continuity In the recent work [11], they address the issue of local boundedness when \(p=2\) for a more general class of operators by a direct proof. By avoiding the difficulty of the existence of the caloric replacement, their result does not contain the extra assumption \(r\ge p^\prime \), although they assume \(u \in L^\infty (I;L_{sp}^{p-1}({\mathbb {R}}^n))\).

We compare our boundedness result to [22]. Their result concerns more general right-hand sides depending on the solution as well. In the limiting case of \(s \rightarrow 1\), they reproduce the local boundedness result contained in [21] for the evolution p-Laplacian equation. To compare the results, if we restrict their result to right-hand sides that are u-independent, their assumption on the integrability becomes \(q,r >\frac{n+sp}{sp}(\frac{p(n+2\,s)}{2sp +(p-1)n}) \). Their analysis is done with the same integrability assumption in time and space. Our local boundedness result, Theorem 1.1, contains this range of exponents.

In the limiting case when s goes to 1, 1.5 become \(1-\frac{1}{r} - \frac{n}{pq} >0\). This is in accordance with the classical condition for boundedness of the evolution p-Laplace equation, see, for example, Remark 1 in [38], there they have a finer assumption formulated in terms of the Lorentz norm of the right-hand side, and moreover, they obtained estimates in terms of a parabolic version of Wolf potentials. It would be interesting to obtain finer estimates beyond \(L^p\) spaces, although we do not pursue this question in this article. If we assume the same integrability in time and space, the condition \(1-\frac{1}{r} -\frac{n}{spq}>0 \) reduces to \(f\in L^{{\hat{q}}}\) with \({\hat{q}}>\frac{n+p}{p}\). This matches the condition in [50].

Now we turn our attention to the nonlocal elliptic (time-independent) case. For \(r=\infty \), the condition for boundedness and basic Hölder continuity becomes

$$\begin{aligned} q> \frac{n}{sp}\,, \quad \text {if }\; sp<n, \quad \text {and} \quad q >1\,, \quad \text {if }\; sp=n, \quad \text {and} \quad q \ge 1\,, \quad \text {if }\; sp=n. \end{aligned}$$

In the case \(sp<n\), this is the same condition for local boundedness and continuity contained in [8, 35]. When \(sp>n\) and \(q\ge 1\), the boundedness and Hölder continuity for the time-independent equation is automatic using Morrey’s inequality. The question of whether the solutions are locally bounded under the equality case of (1.5) is subtle. On the one hand, if \(r=\infty \) even in the time-independent (elliptic) setting one requires the strict inequality \(q> \frac{n}{sp}\) to obtain boundedness; on the other hand, local boundedness is obtained in the case \(r=1\) and \(q=\infty \) in [34], see also [10].

There are actually local boundedness and Hölder continuity results available for the equations with zero right-hand side if \(p< 2\). One could try to prove local boundedness and basic Hölder regularity of the solutions for the solutions of the equations with right-hand side in the singular case \(p<2\) as well. We have to warn the reader that some of the arguments in this article do not carry over to the singular case as they are written here. We use the condition \(p\ge 2\) extensively, in particular in the Pointwise inequalities (2.1) and (2.3). We feel that it is better if we leave the study of the singular case to another work. We also have to mention that if one is only interested in local boundedness estimates, doing a nonperturbative argument is more suitable, as one can also deal with sub- and supersolutions.

Hölder continuity exponent: In the case \(r= \infty \), the critical Hölder continuity exponent

$$\begin{aligned} \min {\left\{ \Theta , \frac{r(spq-n)-spq}{q(r(p-1)-(p-2))} \right\} } =\min {\left\{ \Theta , sp\frac{1-\frac{1}{r}-\frac{n}{spq}}{p-1-\frac{p-2}{r}} \right\} }, \end{aligned}$$
(1.6)

reduces to \(\min { \lbrace \Theta ,\frac{sp}{p-1}(1-\frac{n}{spq})\rbrace } \) which matches the results in [6]. Although the results reported in [6] require a strict inequality \(\alpha < \min { \lbrace \Theta ,\frac{sp}{p-1}(1-\frac{n}{spq})\rbrace }\), an inspection of the proofs reveals that the strict inequality is only needed when the minimum corresponds to \(\Theta \). The assumptions needed for their proof are actually \(\alpha \le \frac{sp}{p-1}(1-\frac{n}{spq})\) and \(\alpha < \Theta \). Through a finer estimate in [25], they have addressed this issue further and proved that given \(\alpha \le \Theta \), if the right-hand side f belongs to the Marcinkiewicz space \( L^{\frac{n}{sp-\alpha (p-1)}, \infty }(\Omega )\) then the solution is \(C^\alpha _{\textrm{loc}}(\Omega )\).

Let us also compare our results to the local p-parabolic equation studied in [47] where precise Hölder continuity exponents are obtained. If we send s to 1, (1.6) becomes

$$\begin{aligned} \min {\left\{ 1, \frac{r(pq-n)-pq}{q(r(p-1)-(p-2))} \right\} } \,, \end{aligned}$$

which is in accordance with the result in [47].

In [29], explicit Hölder continuity exponents for the more general case of double phase nonlocal diffusion operators were obtained. The ideas explored there are similar to the ones in [7], but their result allows for a bounded right-hand side instead of just zero. Their result implies the Hölder continuity exponent that we get in the case of \(f\in L^\infty \), although with a slightly different estimate of the Hölder constants. Let us also mention that in the recent work [11] the conclusions of Theorem 1.2 have been obtained when \(p=2\), for a more general class of operators and kernels. Although [11, Theorem 1.2] does not contain the extra assumption \(r\ge p^\prime \), their argument is similar to our proof of Theorem 1.2 and the same difficulty regarding the existence of the (sp)-caloric replacement is present in their proof. This difficulty has not been addressed properly in their article. In the assumptions [11, (A.1)] for their existence theorem, the regularity assumption \(\xi _t \in L^2((0,T);W^{s,2}(\Omega ^\prime ))^\star \) is present. We are not able to verify this assumption when \(\xi \) is a solution of

$$\begin{aligned} \xi _t + (-\Delta )^s \xi = f, \end{aligned}$$

for \(f \in L^{q,r}\) such that \(\frac{n}{2sq}+ \frac{1}{r} <1\) as it is claimed by [11, Remark 6]. See Remark 2.13 for a possible strategy for resolving this issue.

Let us close this section with the question of how much regularity one should expect if the solution has a lower integrability of the tail in time. Namely given a weak solution u of the equation

$$\begin{aligned} u_t+(-\Delta _p)^s u=0, \end{aligned}$$

such that \(u \in L^{l}(L_{sp}^{p-1}({\mathbb {R}}^n))\) how much Hölder regularity does the solution have. Let us first mention that an example in [34, Example 5.2] shows that the assumption \(u \in L^{p-1}(L_{sp}^{p-1}({\mathbb {R}}^n))\) does not ensure the Hölder regularity of the solution. On the other hand, it is proved in [34] and [10] that if \(l>p-1\) then the solution is Hölder continuous, and the general strategy in these works is to treat the following nonlocal term

$$\begin{aligned} G(t)=\int _{{\mathbb {R}}^n\setminus B_R(x_0)} \frac{|u(y,t)|^{p-1}}{|x_0-y|^{n+sp}} \;\textrm{d}y, \end{aligned}$$

which appears in the Caccioppoli inequalities, as a right-hand side in \(L^{\frac{l}{p-1}}(I; L^\infty (B))\). See [10, Section 1.2] for more details. Following this general philosophy, one can expect the solution to be \(C^\alpha _x\) and \(C^{\frac{\alpha }{sp- (p-2)\alpha }}_t\) with

$$\begin{aligned} \alpha = \min \left\{ sp \frac{1-\frac{p-1}{l}}{p-1 - \frac{(p-2)(p-1)}{l}},1\right\} . \end{aligned}$$

However, at the moment we do not have a definite answer to how this can be shown rigorously.

1.2 Plan of the paper

In Sect. 2, we introduce some notations and preliminary lemmas. We also restate and adapt a result on the existence of solutions to our setting.

In Sect. 3, we establish basic local Hölder regularity and boundedness for local weak solutions.

Section 4 is devoted to proving Theorem 1.2. A so-called tangential analysis is performed to get specific Hölder continuity exponents in terms of qrs and p.

The article is also accompanied by two appendices. In the first one, Appendix A, we work out the details for a modified version of [7, Theorem 1.1]. The aim is to bound the Hölder seminorm of the solution in terms of the tail quantity.

In Appendix B, we justify using certain test functions in the weak formulation of (1.1).

2 Preliminaries

2.1 Notation

We define the monotone function \(J_p: {\mathbb {R}}\rightarrow {\mathbb {R}}\) by

$$\begin{aligned} J_p(t)= |t|^{p-2}t . \end{aligned}$$

We use the notation \(B_R(x_0)\) for the open ball of radius R centered at \(x_0\). If the center is the origin, we simply write \(B_R\). We use the notation of \(\omega _n\) for the surface area of the unit n-dimensional ball. For parabolic cylinders, we use the notation \(Q_{r,T}(x_0,t_0):= B_r(x_0)\times (t_0-T, t_0]\). If the center is the origin, we write \(Q_{r,T}\).

We will work with the fractional Sobolev space extensively:

$$\begin{aligned} W^{\beta , q}({\mathbb {R}}^n):= \lbrace \psi \in L^q({\mathbb {R}}^n) \,:\, [\psi ]_{W^{\beta , q}({\mathbb {R}}^n)}< \infty \rbrace , \qquad 0< \beta<1,\quad 1\le q<\infty , \end{aligned}$$

where the seminorm \([\psi ]_{W^{s,p}({\mathbb {R}}^n)}\) is defined as below

$$\begin{aligned} {[}\psi ]_{W^{\beta ,q}({\mathbb {R}}^n)}^q = \iint _{{\mathbb {R}}^n \times {\mathbb {R}}^n } \frac{|\psi (x)-\psi (y)|^q}{|x-y|^{n+\beta q}} \;\textrm{d}x \;\textrm{d}y. \end{aligned}$$

We also need the space \(W^{\beta , q }(\Omega )\) for a subset \(\Omega \subset {\mathbb {R}}^n\), defined by

$$\begin{aligned} W^{\beta , q}(\Omega ):= \lbrace \psi \in L^q(\Omega ) \,:\, [\psi ]_{W^{\beta , q}(\Omega )}< \infty \rbrace , \qquad 0< \beta<1, \quad 1\le q<\infty , \end{aligned}$$

where

$$\begin{aligned} {[}\psi ]_{W^{\beta ,q}(\Omega )}^q = \iint _{\Omega \times \Omega } \frac{|\psi (x)-\psi (y)|^q}{|x-y|^{n+\beta q}} \;\textrm{d}x \;\textrm{d}y. \end{aligned}$$

In the following, we assume that \(\Omega \subset {\mathbb {R}}^n\) is a bounded open set in \({\mathbb {R}}^n\). We define the space of Sobolev functions taking boundary values \(g \in L^{q-1}_{sp}({\mathbb {R}}^n) \) by

$$\begin{aligned} X_g^{\beta ,q}(\Omega ,\Omega ^\prime ) = \lbrace \psi \in W^{\alpha ,q}(\Omega ^\prime )\cap L_{sp}^{q-1}({\mathbb {R}}^n) \,: \, \psi =g \; \text {on} \; {\mathbb {R}}^n \setminus \Omega \, \rbrace , \end{aligned}$$

where \(\Omega ^\prime \) is an open set such that \(\Omega \Subset \Omega ^\prime \).

We recall the definition of tail space

$$\begin{aligned} L^{q}_{\alpha }({\mathbb {R}}^n)=\left\{ u\in L^{q}_{\textrm{loc}}({\mathbb {R}}^n)\,:\, \int _{{\mathbb {R}}^n} \frac{|u|^q}{1+|x|^{n+\alpha }}\;\textrm{d}x < +\infty \right\} ,\qquad q \ge 1 \text { and } \alpha > 0, \end{aligned}$$

which is endowed with the norm

$$\begin{aligned} \Vert u\Vert _{L_\alpha ^{q}({\mathbb {R}}^n)} = \left( \int _{{\mathbb {R}}^n} \frac{|u|^q}{1+|x|^{n+\alpha }}\;\textrm{d}x\right) ^{\frac{1}{q}}. \end{aligned}$$

For every \(x_0\in {\mathbb {R}}^n\), \(R>0\) and \(u\in L^q_{\alpha }({\mathbb {R}}^n)\), the following quantity

$$\begin{aligned} \textrm{Tail}_{q,\alpha }(u;x_0,R)=\left[ R^{\alpha }\,\int _{{\mathbb {R}}^n\setminus B_R(x_0)} \frac{|u|^q}{|x-x_0|^{n+\alpha }}\;\textrm{d}x\right] ^\frac{1}{q} \end{aligned}$$

plays an important role in regularity estimates for solutions to fractional problems.

Let \(I \subset {\mathbb {R}}\) be an interval and let V be a separable, reflexive, Banach space endowed with a norm \(\Vert \,\bullet \,\Vert _V\). We denote by \(V^\star \) its topological dual space. Suppose that v is a mapping such that for almost every \(t \in I\), we have \(v(t) \in V\). If the function \(t \rightarrow \Vert v(t)\Vert _V\) is measurable on I and \(1 \le p \le \infty \), then v is an element of the Banach space \(L^p(I; V)\) if and only if

$$\begin{aligned} \int _I \Vert v(t)\Vert _V \;\textrm{d}t < \infty . \end{aligned}$$

By [44, Theorem 1.5], the dual space of \(L^p(I; V)\) can be characterized according to \((L^p(I; V ))^\star = L^{p^\prime }(I; V^\star ).\) We write \(v \in C(I; V )\) if the mapping \(t \rightarrow v(t)\) is continuous with respect to the norm on V.

2.2 Pointwise inequalities

We will need the following pointwise inequality: Let \(p \ge 2\), then for every \(A,B \in {\mathbb {R}}\) we have

$$\begin{aligned} |A-B|^p \le C \bigl ( J_p(A) - J_p(B)\bigr )(A-B). \end{aligned}$$
(2.1)

For a proof look at [7, Remark A.4], a close inspection of the proof reveals that the constant can be taken as \(C= 3 \cdot 2^{p-1}\). Before stating the next inequality, we recall [8, Lemma A.2].

Lemma 2.1

Let \(1< p < \infty \) and \(g:{\mathbb {R}}\rightarrow {\mathbb {R}}\) be an increasing function, and define

$$\begin{aligned} G(t)= \int _0^t g^\prime (\tau )^{\frac{1}{p}} \;\textrm{d}\tau , \quad t \in {\mathbb {R}}. \end{aligned}$$

Then

$$\begin{aligned} J_p(a-b) \bigl ( g(a)-g(b) \bigr ) \ge \bigl |G(a)-G(b)\bigr |^p. \end{aligned}$$

Lemma 2.2

For \(p\ge 2\) and \(\beta \ge 1\),

$$\begin{aligned}{} & {} \bigl ( J_p(a-b)- J_p(c-d) \bigr )\Bigl ( ((a-c)^{+}_M +\delta )^\beta -((b-d)^{+}_M+\delta )^\beta \Bigr ) \nonumber \\{} & {} \quad \qquad \ge \frac{1}{3 \cdot 2^{p-1}} \frac{\beta p^p}{(\beta +p-1)^p}\Bigl |((a-c)^{+}_M +\delta )^{\frac{\beta +p-1}{p}} - ((b-d)^{+}_M +\delta )^{\frac{\beta + p-1}{p}}\Bigr |^p, \nonumber \\ \end{aligned}$$
(2.2)

where \((t)^{+}_M:=\min {\lbrace \max {\lbrace t,0\rbrace },M\rbrace }\).

Proof

We consider three cases according to the sign of \(a-b-c+d\). If \(a-b-c+d=0\) both the left-hand side and the right-hand side of (2.2) are zero. Now we verify the inequality for \(a-b-c+d > 0\)

First notice that using (2.1) with \(A= a-b \; \text {and}\; B= c-d\):

$$\begin{aligned} 3\cdot 2^{p-1} (a-b -c+d)(J_p(a-b)- J_p(c-d))\ge \left| a-b -c+d\right| ^p , \end{aligned}$$

using the fact that \(a-b -c+d > 0\), we arrive at

$$\begin{aligned} J_p(a-b)- J_p(c-d) \ge \frac{1}{3 \cdot 2^{p-1}} \frac{|a-b -c+d|^p}{a-b -c+d} = \frac{1}{3 \cdot 2^{p-1}} J_p((a-c)- (b-d)).\nonumber \\ \end{aligned}$$
(2.3)

Now we use Lemma 2.1 with \(g(t)= ((t)^{+}_M +\delta )^\beta \). Then with \(G= \int _0^t g^\prime (\tau )^{\frac{1}{p}} \;\textrm{d}\tau \),

$$\begin{aligned} G(t)=\frac{p\beta ^{\frac{1}{p}}}{\beta +p-1}\Bigl ((t^{+}_M +\delta )^{\frac{\beta +p-1}{p}} - \delta ^{\frac{\beta +p-1}{p}} \Bigr ). \end{aligned}$$

By Lemma 2.1,

$$\begin{aligned} J_p \bigl ((a-c)-(b-d) \bigr ) \bigl (g(a-c)-g(b-d) \bigr ) \ge |G(a-c)-G(b-d)|^p. \end{aligned}$$

Hence,

$$\begin{aligned} \begin{aligned}&J_p \bigl ((a-c)-(b-d) \bigr ) \Bigl ( ((a-c)^{+}_M +\delta )^\beta -((b-d)^{+}_M +\delta )^\beta \Bigr ) \\&\quad \ge \frac{\beta p^p}{(\beta +p-1)}\Bigl |((a-c)^{+}_M +\delta )^{\frac{\beta +p-1}{p}} - ((b-d)^{+}_M +\delta )^{\frac{\beta + p-1}{p}}\Bigr |^p. \end{aligned} \end{aligned}$$

Using (2.3) in the above inequality concludes the proof. It only remains to verify the case \(a-b-c+d <0\), now we are in the previous position and can use with (badc) instead of (abcd) to obtain

$$\begin{aligned} \begin{aligned}&\bigl ( J_p(b-a)- J_p(d-c) \bigr )\Bigl ( ((b-d)^{+}_M+\delta )^\beta -((a-c)^{+}_M +\delta )^\beta \Bigr ) \\&\ge \frac{1}{3 \cdot 2^{p-1}} \frac{\beta p^p}{(\beta +p-1)^p}\left| ((b-d)^{+}_M +\delta )^{\frac{\beta + p-1}{p}} - ((a-c)^{+}_M +\delta )^{\frac{\beta +p-1}{p}} \right| ^p. \end{aligned} \end{aligned}$$

As

$$\begin{aligned} \begin{aligned} J_p(b-a)- J_p(d-c) = - \bigl ( J_p(a-b) - J_p(c-d) \bigr ), \end{aligned} \end{aligned}$$
$$\begin{aligned} \begin{aligned} ((b-d)^{+}_M+\delta )^\beta -((a-c)^{+}_M +\delta )^\beta = - \Bigl ( ((a-c)^{+}_M +\delta )^\beta - ((b-d)^{+}_M+\delta )^\beta \Bigr ), \end{aligned} \end{aligned}$$

and

$$\begin{aligned} \begin{aligned}&\left| ((b-d)^{+}_M +\delta )^{\frac{\beta + p-1}{p}} - ((a-c)^{+}_M +\delta )^{\frac{\beta +p-1}{p}} \right| \\&\quad = \left| ((a-c)^{+}_M +\delta )^{\frac{\beta +p-1}{p}} - ((b-d)^{+}_M +\delta )^{\frac{\beta + p-1}{p}} \right| \end{aligned} \end{aligned}$$

we obtain (2.2) \(\square \)

The following pointwise inequality is a direct consequence of the convexity of the mapping \(t \rightarrow |t|^{\alpha }\) for \(\alpha \ge 1\).

$$\begin{aligned} |a^\alpha - b^\alpha |\ge \alpha \min \lbrace a,b \rbrace ^{\alpha -1} |a-b|, \quad \text {for} \quad a,b \ge 0. \end{aligned}$$
(2.4)

2.3 Functional inequalities

We need the following basic inequalities for the tail.

Lemma 2.3

Let \(\alpha >0\), \(1\le q< \infty \), and \(u,\,v \in L_\alpha ^q({\mathbb {R}}^n)\) such that \(u=v\) on \({\mathbb {R}}^n \setminus B_R(x_0)\). Then for any \(\sigma <1\),

Proof

\(\square \)

For a proof of the following result, see [6, Lemma 2.3].

Lemma 2.4

Let \(\alpha >0\), \(0<q<\infty \). Suppose that \(B_r(x_0) \subset B_R(x_1)\). Then for every \(u\in L^q_\alpha ({\mathbb {R}}^n)\), we have

$$\begin{aligned} \textrm{Tail}_{q,\alpha }(u;x_0,r)^q {\le } \Bigl (\frac{r}{R}\Bigr )^\alpha \Bigl (\frac{R}{R- |x-x_0|} \Bigr )^{n + \alpha } \textrm{Tail}_{q, \alpha } (u; x_1, R)^q + r^{-n}\Vert u\Vert _{L^q(B_R(x_1))}^q. \end{aligned}$$

If in addition \(u \in L^m_{\textrm{loc}}({\mathbb {R}}^n)\) for some \(q <m \le \infty \), then

$$\begin{aligned} \textrm{Tail}_{q,\alpha }(u;x_0,r)^q{} & {} \le \Bigl (\frac{r}{R}\Bigr )^\alpha \Bigl (\frac{R}{R- |x-x_0|} \Bigr )^{n + \alpha } \textrm{Tail}_{q, \alpha } (u; x_1, R)^q\\{} & {} \quad + \Bigl ( \frac{(n \omega _n)m-q}{\alpha m+nq}\Bigr )^{\frac{m-q}{m}} r^{-\frac{qn}{m}} \Vert u\Vert _{L^m(B_R(x_1))}, \end{aligned}$$

where \(\omega _n\) is the measure of the n-dimensional open ball of radius 1.

We also recall the following Sobolev- and Morrey-type inequalities:

Proposition 2.5

Suppose \(1<p<\infty \) and \(0<s<1\). Let \(\Omega \subset {\mathbb {R}}^n\) be an open and bounded set. Define \(p_s^\star \) as

$$\begin{aligned} p_s^\star := \frac{np}{n-sp}. \end{aligned}$$
(2.5)

For every \(u \in W^{s,p}({\mathbb {R}}^n)\) vanishing almost everywhere in \({\mathbb {R}}^n {\setminus } \Omega \), we have

$$\begin{aligned}{} & {} \Vert u\Vert _{L^{p_s^\star }(\Omega )}^p \le C_1(n,s,p)\, [u]_{W^{s,p}({\mathbb {R}}^n)}^p, \quad if \quad sp<n \end{aligned}$$
(2.6)
$$\begin{aligned}{} & {} \Vert u\Vert _{L^\infty (\Omega )}^p \le C_2(n,s,p) |\Omega |^{\frac{sp}{n}-1} [u]_{W^{s,p}({\mathbb {R}}^n)}^p, \quad if \quad sp>n \end{aligned}$$
(2.7)
$$\begin{aligned}{} & {} \Vert u\Vert _{L^l(\Omega )}^p \le C_3(n,s,p,l) |\Omega |^{\frac{p}{l}} [u]_{W^{s,p}({\mathbb {R}}^n)}^p, \quad \text { for every }1\le l < \infty , \quad if \; sp=n. \nonumber \\ \end{aligned}$$
(2.8)

In particular, the following Poincaré inequality holds true

$$\begin{aligned} \Vert u\Vert _{L^p(\Omega )}^p \le C\; |\Omega |^{\frac{sp}{n}}[u]_{W^{s,p}({\mathbb {R}}^n)}, \end{aligned}$$
(2.9)

for some \(C=C(n,s,p)\). Furthermore, in the supercritical range of exponents functions in \(W^{s,p}({\mathbb {R}}^n)\) are Hölder continuous and the following inequality holds true:

$$\begin{aligned} {[}u]_{C^{s-\frac{n}{p}}(\Omega )} \le C_4(n,s,p)[u]_{W^{s,p}({\mathbb {R}}^n)}, \quad \text {if} \quad sp>n. \end{aligned}$$
(2.10)

Remark 2.6

The above Sobolev-type inequalities are also valid for functions \(u \in X_0^{s,p}(\Omega , \Omega ^\prime )\), where \(\Omega \) is a bounded open set and \(\Omega ^\prime \) is an open set such that \(\Omega \Subset \Omega ^\prime \). This can be seen using the fact that there is an extension domain containing \(\Omega \) and included in \(\Omega ^\prime \).

We will often use the following special application of Hölder’s inequality

$$\begin{aligned} \Vert u(x,t)\Vert _{L^{q_1,r_1}(\Omega \times J)}{} & {} \le \Vert |\Omega |^{\frac{1}{q_1}-\frac{1}{q_2}} \Vert u(\,\bullet \,,t)\Vert _{L^{q_2}(\Omega )}\Vert _{L^{r_1}(J)} \nonumber \\{} & {} \le |\Omega |^{\frac{1}{q_1}-\frac{1}{q_2}} |J|^{\frac{1}{r_1}-\frac{1}{r_2}} \Vert u\Vert _{L^{q_2,r_2}(\Omega \times J),} \end{aligned}$$
(2.11)

where \(q_1<q_2 \,, r_1\le r_2\). The following interpolation inequality (see, e.g., [5]) will be useful.

Lemma 2.7

If w is contained in \(L^{q_1,r_1}(\Omega \times J) \cap L^{q_2, r_2}(\Omega \times J)\), then w is contained in \(L^{{\tilde{q}}, {\tilde{r}}}(\Omega \times J)\), where

$$\begin{aligned} \frac{1}{{\tilde{r}}} = \frac{\lambda }{r_1} + \frac{1-\lambda }{r_2}, \quad \frac{1}{{\tilde{q}}}= \frac{\lambda }{q_1} + \frac{1- \lambda }{q_2} \;, \quad (0 \le \lambda \le 1). \end{aligned}$$

Moreover,

$$\begin{aligned} \Vert w\Vert _{L^{{\tilde{q}}, {\tilde{r}}}(\Omega \times J)} \le \Vert w\Vert _{L^{q_1,r_1}(\Omega \times J)}^{\lambda } \Vert w\Vert _{L^{q_2,r_2}(\Omega \times J)}^{1-\lambda }. \end{aligned}$$

The following three lemmas will be needed in the proof of our local boundedness result (Proposition 3.4).

Lemma 2.8

Let \(sp \ne n\) and assume that w is in \(L^{p}\bigl ( (T_0-R^{sp},T_0);W^{s,p}({\mathbb {R}}^n)\bigr ) \cap L^{p,\infty }(Q_{R,R^{sp}}(x_0,T_0))\) and w(xt) is zero for all \(x \in {\mathbb {R}}^n {\setminus } B_R(x_0)\), for almost every \(t \in (T_0-R^{sp},T_0]\). Then w is in \(L^{p q^\prime , p r^\prime } (Q_{R,R^{sp}}(x_0,T_0))\) as long as qr satisfy

$$\begin{aligned} 1-\frac{1}{r} - \frac{n}{spq} \ge 0. \end{aligned}$$

Moreover,

$$\begin{aligned}{} & {} \Vert w\Vert _{L^{pq^\prime , pr^\prime }(Q_{R,R^{sp}}(x_0,T_0))}^p \\{} & {} \quad \le C R^{sp(1-\frac{1}{r} - \frac{n}{spq})} \Bigl ( \Vert w\Vert _{L^{p,\infty }(Q_{R,R^{sp}}(x_0,T_0))}^p + \int _{T_0-R^{sp}}^{T_0} [w]_{W^{s,p}({\mathbb {R}}^n)}^p \;\textrm{d}t \Bigr ), \end{aligned}$$

where C depends on \(n,s \text { and } p\). In particular, in the case of \(\frac{1}{r} + \frac{n}{spq}=1\) we have

$$\begin{aligned}{} & {} \Vert w\Vert _{L^{pq^\prime , pr^\prime }(Q_{R,R^{sp}}(x_0,T_0))}^p \\{} & {} \quad \le C(n,s,p)\left( \Vert w\Vert _{L^{p,\infty }(Q_{R,R^{sp}}(x_0,T_0))}^p + \int _{T_0-R^{sp}}^{T_0} [w]_{W^{s,p}({\mathbb {R}}^n)}^p \;\textrm{d}t \right) . \end{aligned}$$

Proof

Consider a pair of exponents \({\tilde{r}} = (\frac{1}{r^\prime } - (1-\frac{1}{r} - \frac{n}{spq}))^{-1}= \frac{spq}{n}\), and \({\tilde{q}} = q^\prime \) such that \(\frac{1}{{\tilde{r}}^{\, \prime }} + \frac{n}{sp{\tilde{q}}^{\, \prime }} =1 \). Using Hölder’s inequality (2.11), we obtain

$$\begin{aligned} \begin{aligned} \Vert w\Vert _{L^{pq^\prime , pr^\prime }(Q_{R,R^{sp}}(x_0,T_0))}^p&\le (R^{sp})^{\frac{1}{r^\prime } - \frac{1}{{\tilde{r}}}} \Vert w\Vert _{L^{p{\tilde{q}}, p{\tilde{r}}}(Q_{R,R^{sp}}(x_0,T_0))}^p\\&= R^{sp(1-\frac{1}{r} - \frac{n}{spq})} \Vert w\Vert _{L^{p{\tilde{q}}, p{\tilde{r}}}(Q_{R,R^{sp}}(x_0,T_0))}^p. \end{aligned} \end{aligned}$$

Now we split the proof into two cases depending on whether \(sp<n\) or not.

Case \(sp<n\): We use Lemma 2.7 with the choice

$$\begin{aligned} \frac{1}{p{\tilde{r}}} = \frac{\lambda }{p} \qquad \text {and} \qquad \frac{1}{p{\tilde{q}}} = \frac{\lambda }{p_s^\star } + \frac{1-\lambda }{p} \,,\quad (0\le \lambda \le 1). \end{aligned}$$

This yields

$$\begin{aligned} \Vert w\Vert _{L^{p{\tilde{q}}, p{\tilde{r}}}(Q_{R,R^{sp}}(x_0,T_0))} \le \Vert w\Vert _{L^{p_s^\star , p}(Q_{R,R^{sp}}(x_0,T_0))}^\lambda \Vert w\Vert _{L^{p,\infty }(Q_{R,R^{sp}}(x_0,T_0))}^{1-\lambda }. \end{aligned}$$

The above relations hold for \(\lambda = \frac{1}{{\tilde{r}}} = \frac{n}{sp{\tilde{q}}^{\, \prime }}\), and using Sobolev’s inequality 2.6, we arrive at

$$\begin{aligned} \begin{aligned} \Vert w\Vert _{L^{p{\tilde{q}}, p{\tilde{r}}}(Q_{R,R^{sp}}(x_0,T_0))}^p&\le \Vert w\Vert _{L^{p_s^\star , p}(Q_{R,R^{sp}}(x_0,T_0))}^{p \lambda } \Vert w\Vert _{L^{p,\infty }(Q_{R,R^{sp}}(x_0,T_0))}^{p(1-\lambda )} \\&\le C \Vert w\Vert _{L^{p,\infty }(Q_{R,R^{sp}}(x_0,T_0))}^{p(1-\lambda )} \left( \int _{T_0-R^{sp}}^{T_0} [w]_{W^{s,p}({\mathbb {R}}^n)}^p \;\textrm{d}t\right) ^{\lambda }. \end{aligned} \end{aligned}$$

By using Young’s inequality, we get

$$\begin{aligned} \begin{aligned}&\Vert w\Vert _{L^{p{\tilde{q}}, p{\tilde{r}}}(Q_{R,R^{sp}}(x_0,T_0))}^p\\&\quad \le C \Vert w\Vert _{L^{p, \infty }(Q_{R,R^{sp}}(x_0,T_0))}^{p (1-\lambda )} \left( \int _{T_0-R^{sp}}^{T_0} [w]_{W^{s,p}({\mathbb {R}}^n)}^p \;\textrm{d}t\right) ^{\lambda } \\&\quad \le C \left( (1-\lambda )\Vert w\Vert _{L^{p,\infty }(Q_{R,R^{sp}}(x_0,T_0))}^{p} +\lambda \int _{T_0-R^{sp}}^{T_0} [w]_{W^{s,p}({\mathbb {R}}^n)}^p \;\textrm{d}t \right) \\&\quad \le C(n,s,p)\left( \Vert w\Vert _{L^{p,\infty }(Q_{R,R^{sp}}(x_0,T_0))}^{p} +\int _{T_0-R^{sp}}^{T_0} [w]_{W^{s,p}({\mathbb {R}}^n)}^p \;\textrm{d}t \right) . \end{aligned} \end{aligned}$$

Case \(sp>n\): In this case, we use the following interpolation between Hölder and \(L^p\) spaces:

$$\begin{aligned} \Vert w\Vert _{L^\infty (B_{R}(x_0))}\le C \Vert w\Vert _{L^p(B_{R}(x_0))}^\chi [w]_{C^\alpha (B_{R}(x_0))}^{1-\chi }, \quad \text {with} \quad \chi =\frac{\alpha }{\alpha + \frac{n}{p}}. \end{aligned}$$

See [9, Lemma 2.2] for a proof. In light of the Morrey-type inequality (2.10), for almost every \(t \in (T_0-R^{sp}, T_0) \) we arrive at

$$\begin{aligned} \Vert w(\,\bullet \,,t)\Vert _{L^\infty (B_R(x_0))}{} & {} \le C \Vert w(\,\bullet \,,t)\Vert _{L^p (B_R(x_0))}^{1-\frac{n}{sp}} [w(\,\bullet \,,t)]_{C^{s-\frac{n}{p}}(B_R(x_0))}^{\frac{n}{sp}}\nonumber \\{} & {} \le C \Vert u(\,\bullet \,,t)\Vert _{L^p (B_R(x_0))}^{1-\frac{n}{sp}} [w(\,\bullet \,,t)]_{W^{s,p}({\mathbb {R}}^n)}^{\frac{n}{sp}}. \end{aligned}$$
(2.12)

Now we interpolate once more between \(L^p\) and \(L^\infty \) to obtain

$$\begin{aligned} \begin{aligned} \Vert w(\,\bullet \,,t)\Vert _{L^{p {\tilde{q}}}(B_R(x_0))}&\le \Vert w(\,\bullet \,,t)\Vert _{L^p(B_R(x_0))}^{\frac{1}{{\tilde{q}}}} \Vert w(\,\bullet \,,t)\Vert _{L^\infty (B_R(x_0))}^{\frac{1}{{\tilde{q}}^\prime }}\\ \text {by } (2.12)\quad&\le C \Vert w(\,\bullet \,,t)\Vert _{L^p(B_R(x_0))}^{\frac{1}{{\tilde{q}}}} \Vert w(\,\bullet \,,t)\Vert _{L^p(B_R(x_0))}^{\frac{1}{{\tilde{q}}^\prime }(1-\frac{n}{sp})} [w(\,\bullet \,,t)]_{W^{s,p}({\mathbb {R}}^n)}^{\frac{n}{sp {\tilde{q}}^\prime }}\\&\le C \Vert w(\,\bullet \,,t)\Vert _{L^p(B_R(x_0))}^{1-\frac{n}{sp{\tilde{q}}^\prime }} [w(\,\bullet \,,t)]_{W^{s,p}({\mathbb {R}}^n)}^{\frac{n}{sp {\tilde{q}}^\prime }}. \end{aligned} \end{aligned}$$

We raise both sides to the power \(p{\tilde{r}}\) and integrate with respect to t. Recalling that \(\frac{1}{{\tilde{r}}}=\frac{n}{sp {\tilde{q}}^\prime }\), we obtain

$$\begin{aligned} \begin{aligned} \Vert w\Vert _{L^{p{\tilde{q}}, p{\tilde{r}}}(Q_{R,R^{sp}}(x_0,T_0))}^{p {\tilde{r}}}&\le C^{p{\tilde{r}}} \int _{T_0-R^{sp}}^{T_0} \Vert w(\,\bullet \,,t)\Vert _{L^p(B_R(x_0))}^{p({\tilde{r}}-1)} [w(\,\bullet \,,t)]_{W^{s,p}({\mathbb {R}}^n)}^{p} \;\textrm{d}t \\&\le C^{p{\tilde{r}}} \sup _{T_0-R^{sp}\le t\le T_0}\Vert w(\,\bullet \,,t)\Vert _{L^p(B_R(x_0))}^{p({\tilde{r}}-1)}\\&\quad \times \int _{T_0-R^{sp}}^{T_0} [w(\,\bullet \,,t)]_{W^{s,p}({\mathbb {R}}^n)}^{p} \;\textrm{d}t. \end{aligned} \end{aligned}$$

Taking the \({\tilde{r}}\) root and applying Young’s inequality, we obtain the desired estimate:

$$\begin{aligned} \begin{aligned}&\Vert w\Vert _{L^{p{\tilde{q}}, p{\tilde{r}}}(Q_{R,R^{sp}}(x_0,T_0))}^p \le C^p \Vert w\Vert _{L^{p, \infty }(Q_{R,R^{sp}}(x_0,T_0))}^{p (\frac{{\tilde{r}}-1}{{\tilde{r}}})} \left( \int _{T_0-R^{sp}}^{T_0} [w]_{W^{s,p}({\mathbb {R}}^n)}^p \;\textrm{d}t\right) ^{\frac{1}{{\tilde{r}}}} \\&\quad \le C(n,s,p)\left( \Vert w\Vert _{L^{p,\infty }(Q_{R,R^{sp}}(x_0,T_0))}^{p} +\int _{T_0-R^{sp}}^{T_0} [w]_{W^{s,p}({\mathbb {R}}^n)}^p \;\textrm{d}t \right) . \end{aligned} \end{aligned}$$

\(\square \)

Lemma 2.9

Let \(sp=n,\, q\ge 1, \text { and } r\ge 1 \) such that

$$\begin{aligned} 1 - \frac{1}{r} - \frac{1}{q} > 0. \end{aligned}$$

Assume that \(w \in L^{l,p}(Q_{R,R^{sp}}(x_0,T_0))\cap L^{p,\infty }(Q_{R,R^{sp}}(x_0,T_0))\) for some l such that

$$\begin{aligned} l= \frac{p}{r^\prime }\left( 1-\frac{1}{r} - \frac{1}{q}\right) ^{-1}. \end{aligned}$$

Then w belongs to \(L^{pq^\prime , p r^\prime }(Q_{R,R^{sp}}(x_0,T_0))\) and

$$\begin{aligned} \Vert w\Vert _{L^{pq^\prime , p r^\prime }(Q_{R,R^{sp}}(x_0,T_0))}^p {\le } R^{\frac{np}{lr^\prime }} \Bigl ( \Vert w\Vert _{L^{p,\infty }(Q_{R,R^{sp}}(x_0,T_0))}^p {+} R^{\frac{-np}{l}}\Vert w\Vert _{L^{l,p}(Q_{R,R^{sp}}(x_0,T_0))}^p \Bigr ). \end{aligned}$$

Proof

We use Lemma 2.7 with the choice

$$\begin{aligned} \frac{1}{p r^\prime } = \frac{\lambda }{p} \quad \text {and} \quad \frac{1}{p q^\prime }= \frac{\lambda }{l} + \frac{1-\lambda }{p}, \quad (0\le \lambda \le 1). \end{aligned}$$

Due to the assumption \(\frac{1}{l}= \frac{r^\prime }{p}(1-\frac{1}{r}-\frac{1}{q})\), the above equalities hold for \(\lambda = \frac{1}{r^\prime }\). Hence, we get

$$\begin{aligned} \Vert w\Vert _{L^{p q^\prime , p r^\prime }(Q_{R,R^{sp}}(x_0,T_0))} \le \Vert w\Vert _{L^{l,p}(Q_{R,R^{sp}}(x_0,T_0))}^\lambda \Vert w\Vert _{L^{p, \infty }(Q_{R,R^{sp}}(x_0,T_0))}^{1-\lambda }. \end{aligned}$$

Therefore, recalling that \(\lambda = \frac{1}{r^\prime }\)

$$\begin{aligned} \begin{aligned}&R^{\frac{- n p}{l r^\prime }} \Vert w\Vert _{L^{p q^\prime , p r^\prime }(Q_{R,R^{sp}}(x_0,T_0))}^p =R^{\frac{-\lambda n p}{l}} \Vert w\Vert _{L^{p q^\prime , p r^\prime }(Q_{R,R^{sp}}(x_0,T_0))}^p \\&\le \Bigl ( R^{\frac{-np}{l}} \Vert w\Vert _{L^{l,p}(Q_{R,R^{sp}}(x_0,T_0))}^p \bigr )^\lambda \Bigl (\Vert w\Vert _{L^{p, \infty }(Q_{R,R^{sp}}(x_0,T_0))}^p \Bigr )^{1-\lambda }. \end{aligned} \end{aligned}$$

Using Young’s inequality for the right-hand side, we can conclude

$$\begin{aligned} \Vert w\Vert _{L^{p q^\prime , p r^\prime }(Q_{R,R^{sp}}(x_0,T_0))}^p {\le } R^{\frac{np}{l r^\prime }} \Bigl ( \Vert w\Vert _{L^{p, \infty }(Q_{R,R^{sp}}(x_0,T_0))}^p {+} R^{\frac{-np}{l}} \Vert w\Vert _{L^{l,p}(Q_{R,R^{sp}}(x_0,T_0))}^p \Bigr ). \end{aligned}$$

\(\square \)

2.4 Weak solutions

Definition 2.10

For any \(t_0,t_1\in {\mathbb {R}}\) with \(t_0<t_1\), we define \(I=(t_0,t_1]\). Let

$$\begin{aligned} f\in \left( L^p(I;X^{s,p}_0 ({\mathcal {K}}\,, \Omega ))\cap L^\infty (I;L^2({\mathcal {K}}))\right) ^{\star }, \end{aligned}$$

for any open \({\mathcal {K}}\) such that \({\mathcal {K}}\Subset \Omega \). We say that u is a local weak solution to the equation

$$\begin{aligned} \partial _t u + (-\Delta _p)^su = f,\qquad \text{ in } \Omega \times I, \end{aligned}$$

if for any closed interval \(J=[T_0,T_1]\subset I\), the function u is such that

$$\begin{aligned} u\in L^p(J;W_{{\textrm{loc}}}^{s,p}(\Omega ))\cap L^{p-1}(J;L_{s\,p}^{p-1}({\mathbb {R}}^n))\cap C(J;L_{{\textrm{loc}}}^2(\Omega )), \end{aligned}$$

and it satisfies

$$\begin{aligned} \begin{aligned}&-\int _J\int _\Omega u(x,t)\,\partial _t\varphi (x,t)\;\textrm{d}x\;\textrm{d}t+ \int _J\iint _{{\mathbb {R}}^n\times {\mathbb {R}}^n}\\&\quad \frac{J_p(u(x,t)-u(y,t))\,(\varphi (x,t)-\varphi (y,t))}{|x-y|^{n+s\,p}}\,\;\textrm{d}x \;\textrm{d}y \;\textrm{d}t \\&\quad = \int _\Omega u(x,T_0)\,\varphi (x,T_0)\;\textrm{d}x -\int _\Omega u(x,T_1)\,\varphi (x,T_1)\;\textrm{d}x +\langle f,\varphi \rangle , \end{aligned}\nonumber \\ \end{aligned}$$
(2.13)

for any \(\varphi \in L^p(J;W^{s,p}(\Omega ))\cap C^1(J;L^2(\Omega ))\) which has spatial support compactly contained in \(\Omega \). In equation (2.13), the symbol \(\langle \,\bullet \,,\,\bullet \,\rangle \) stands for the duality pairing between \(W^{s,p}(\Omega )\) and its dual space \((W^{s,p}(\Omega ))^*\).

Now, we define the notion of a weak solution to an initial boundary value problem.

Definition 2.11

Let \(I=[t_0,t_1]\), \(p\ge 2\), \(0<s<1\), and \(\Omega \Subset \Omega ^\prime \), where \(\Omega ^\prime \) is a bounded open set in \({\mathbb {R}}^n\). Assume that the functions \(u_0,f\) and g satisfy

$$\begin{aligned}{} & {} u_0\in L^2(\Omega ),\\{} & {} \quad f\in \left( L^p(I;X^{s,p}_0 (\Omega \,, \Omega ^\prime ))\cap L^\infty (I;L^2(\Omega ))\right) ^{\star },\\{} & {} \quad g\in L^p(I;W^{s,p}(\Omega '))\cap L^{p-1}(I;L_{s\,p}^{p-1}({\mathbb {R}}^n)). \end{aligned}$$

We say that u is a weak solution of the initial boundary value problem

$$\begin{aligned} \left\{ \begin{array}{rcll} \partial _tu + (-\Delta _p)^su&{}=&{}f,&{} \text{ in } \Omega \times I,\\ u&{}=&{}g,&{} \text{ on } ({\mathbb {R}}^n\setminus \Omega )\times I,\\ u(\,\bullet \,,t_0) &{}=&{} u_0,&{}\text { on }\Omega , \end{array}\right. \end{aligned}$$
(2.14)

if the following properties are verified:

  • \(u\in L^p(I;W^{s,p}(\Omega '))\cap L^{p-1}(I;L_{sp}^{p-1}({\mathbb {R}}^n))\cap C(I;L^2(\Omega ))\);

  • \(u\in X_{{\textbf{g}}(t)}(\Omega ,\Omega ')\) for almost every \(t\in I\), where \(({\textbf{g}}(t))(x)=g(x,t)\);

  • \(\lim _{t\rightarrow t_0}\Vert u(\,\bullet \,,t) - u_0\Vert _{L^2(\Omega )}=0\);

  • for every \(J=[T_0,T_1]\subset I\) and every \(\varphi \in L^{p}(J;X_0^{s,p}(\Omega ,\Omega '))\cap C^1(J;L^2(\Omega ))\)

    $$\begin{aligned} \begin{aligned}&\quad -\int _J\int _\Omega u(x,t)\,\partial _t\varphi (x,t)\;\textrm{d}x\;\textrm{d}t\\&\quad + \int _J\iint _{{\mathbb {R}}^n\times {\mathbb {R}}^n}\frac{J_p(u(x,t)-u(y,t))\,(\varphi (x,t)-\varphi (y,t))}{|x-y|^{n+sp}}\;\textrm{d}x\;\textrm{d}y\;\textrm{d}t \\&= \int _\Omega u(x,T_0)\,\varphi (x,T_0)\;\textrm{d}x -\int _\Omega u(x,T_1)\,\varphi (x,T_1)\;\textrm{d}x \\&\quad + \langle f,\varphi \rangle . \end{aligned} \end{aligned}$$

Let us mention that given a local weak solution in a cylinder \(I \times \Omega ^\prime \), where \(I=(t_0,t_1]\) and \(\Omega ^\prime \) is a bounded, open subset of \({\mathbb {R}}^n\), by considering a smaller cylinder \(J \times \Omega \) such that \(\Omega \Subset \Omega ^\prime \) and J is a closed interval compactly contained in I we end up a weak solution in the smaller cylinder \(J \times \Omega \).

Throughout the article, we work with right-hand sides \(f\in L^{p^\prime } (I;L^{(p^\star _s)^\prime }(\Omega ))\), where \(p_s^\star \) is the Sobolev exponent and we consider it to be infinity if \(sp>n\). An application of Hölder’s inequality together with the Sobolev–Morrey inequalities ensures that \(f\in f\in L^{p'}(I;(X_0^{s,p}(\Omega ,\Omega '))^*) \subset \left( L^p(I;X^{s,p}_0 (\Omega \,, \Omega ^\prime ))\cap L^\infty (I;L^2(\Omega ))\right) ^{\star }\) with the duality pairing

$$\begin{aligned} \langle f,\varphi \rangle = \int _I \int _\Omega f(x,t) \varphi (x,t) \;\textrm{d}x \;\textrm{d}t. \end{aligned}$$

Theorem 2.12

Let \(p\ge 2\), let \(I = (T_0,T_1]\) and suppose that g satisfies

$$\begin{aligned}&g\in L^p(I;W^{s,p}(\Omega '))\cap L^{p}(I;L^{p-1}_{s\,p}({\mathbb {R}}^n)) \cap C(I;L^2(\Omega )),\\&\partial _t g\in L^{p'}(I;(X_0^{s,p}(\Omega ,\Omega '))^*),\\&\lim _{t\rightarrow t_0}\Vert g(\,\bullet \,,t)-g_0\Vert _{L^2(\Omega )} = 0,\qquad \text { for some }g_0\in L^2(\Omega ). \end{aligned}$$

Suppose also that

$$\begin{aligned} f\in L^{p'}(I;(X_0^{s,p}(\Omega ,\Omega '))^*). \end{aligned}$$

Then for any initial datum \(g_0\in L^2(\Omega )\), there exists a unique weak solution u to problem

$$\begin{aligned} {\left\{ \begin{array}{ll} u_t + (-\Delta _p)^s u= f \quad &{} in \;\; \Omega \times I \\ u=g \quad &{} in \;\; ({\mathbb {R}}^n \setminus \Omega ) \times I \\ u(x,T_0) = g(x,T_0) \quad &{} in \;\; \Omega . \end{array}\right. } \end{aligned}$$
(2.15)

Proof

In [7, Theorem A.3], the same result is proved with a stronger condition \(g_t \in L^{p^\prime }(I;W^{s,p}(\Omega ^\prime )^\star )\). The stronger condition is not needed in the proof. This condition can be replaced with \(g_t \in L^{p^\prime }(I,X^{s,p}_0 (\Omega ; \Omega ^\prime )^\star )\) in all of the steps in the proof, except that the construction gives us a \(C(I;L^2(\Omega ))\) solution. There, the stronger assumption is used only to show that the boundary condition is in \(C(I;L^2(\Omega ))\), which we assume here. \(\square \)

Remark 2.13

The condition \(\partial _t g \in L^{p'}(I;(X_0^{s,p}(\Omega ,\Omega '))^*)\) is too strong. This condition forces us to assume \(r \ge p^\prime \), \(q \ge (p_s^\star )^\prime \) and \(u \in L^p(I; L^{p-1}_{sp}({\mathbb {R}}^n))\) in Proposition 3.1 and hence in all our results. A more natural condition would be to assume \(\partial _t g \in \left( L^p(I;X^{s,p}_0 (\Omega \,, \Omega ^\prime ))\cap L^\infty (I;L^2(\Omega ))\right) ^{\star }\). We believe it is possible to overcome this difficulty by pursuing an approximation procedure in the spirit of [35, Theorem 1.1 and Lemma 4.1].

3 Basic Hölder regularity and stability

Throughout the rest of the article, we assume \(0<s<1\) and \(2\le p < \infty \).

Here, we argue that the norm of the (sp)-caloric replacement of u is close to u if f is small enough. By the (sp)-caloric replacement of u in a cylinder \(B_\rho (x_0) \times I\), we mean the solution to the following

$$\begin{aligned} {\left\{ \begin{array}{ll} v_t + (-\Delta _p)^s v=0 \quad &{} in \;\; B_\rho (x_0) \times I \\ v=u \quad &{} in \;\; ({\mathbb {R}}^n \setminus B_\rho (x_0) ) \times I \\ v(x,\tau _0) = u(x,\tau _0) \quad &{} in \;\; B_\rho (x_0). \end{array}\right. } \end{aligned}$$
(3.1)

Here \(\tau _0\) is the initial point of the interval I. First we show the existence of a (sp)-caloric replacement using Theorem 2.12

Proposition 3.1

Let u be a local weak solution of \( u_t + (-\Delta _p)^s u=f \) in the cylinder \(B_\sigma \times J\), for some interval \(J=(t_1,t_2]\) with \(f \in L^{q,r}_{\textrm{loc}}(B_\sigma \times J)\) such that \(r\ge p^\prime \),

$$\begin{aligned} q\ge (p_s^\star )^\prime \quad \text {if}\quad sp<n,\quad q \ge 1 \quad \text {if} \quad sp>n, \quad \text {and}\quad q>1 \quad \text {if} \quad sp=n. \end{aligned}$$

In addition, we assume that \(u \in L^p(J;L_{sp}^{p-1}(R^n))\). Then for any \(0<\rho < \sigma \), and closed interval \(I \Subset J\), the (sp)-caloric replacement of u in \(B_\rho (x_0)\times I\) (weak solution to (3.1)) exists.

Proof

We shall check the conditions in Theorem 2.12. If they are satisfied, there exists a unique weak solution \(v \in L^p(I, W^{s,p}(B_{\sigma })) \cap L^{p}(I;L_{sp}^{p-1}({\mathbb {R}}^n)) \cap C(I;L^2(B_\rho ))\) to the problem (3.1). The only condition on u that is not immediate from the fact that u is weak solution is \(\partial _t u \in L^{p^\prime }(I;X^{s,p}_0 (B_\rho \,, B_{\sigma })^\star ) \). We have to show that for every function \(\psi \in L^p(I;X^{s,p}_0 (B_\rho \,, B_{\sigma }))\)

$$\begin{aligned} \Bigl |\int _I \langle u_t,\psi \rangle \;\textrm{d}x \;\textrm{d}t\Bigr | \le C \int _{I} \Vert \psi \Vert _{W^{s,p}(B_{\sigma })}^p \;\textrm{d}t. \end{aligned}$$
(3.2)

Here we only write the proof for the case \(sp< n\), the case of \(sp \ge n\) is similar, except that one has to use the critical case of Sobolev inequality and the Morrey inequality instead of using the Sobolev inequality. We shall verify (3.2) for test functions belonging to the dense subspace, \(\psi \in L^p(I;X^{s,p}_0 (B_\rho \,, B_{\sigma }))\cap C^1_{0}(I; L^2(B))\). We use the equation to do so. We have

$$\begin{aligned} \begin{aligned} \int _I \langle u_t,\psi \rangle \;\textrm{d}x \;\textrm{d}t&= \int _I \int _{B_\rho }u \psi _t \;\textrm{d}x \;\textrm{d}t \\&= -\int _I \iint _{{\mathbb {R}}^n \times {\mathbb {R}}^n}\! \frac{J_p(u(x,t){-}u(y,t))(\psi (x,t) {-}\psi (y,t))}{|x-y|^{n+sp}} \;\textrm{d}x \;\textrm{d}y \;\textrm{d}t \\&\quad + \int _I \int _{B_{r}} f(x,t) \psi (x,t) \;\textrm{d}x \;\textrm{d}t \\&= - \int _I \iint _{B_{\sigma } \times B_{\sigma }}\! \frac{J_p(u(x,t){-}u(y,t))(\psi (x,t) {-}\psi (y,t))}{|x-y|^{n+sp}} \;\textrm{d}x \;\textrm{d}y \;\textrm{d}t\\&\quad - 2 \int _I \int _{{\mathbb {R}}^n\setminus B_{\sigma }} \int _{B_\rho } \frac{J_p(u(x,t)-u(y,t))\psi (x,t)}{|x-y|^{n+sp}} \;\textrm{d}x \;\textrm{d}y \;\textrm{d}t \\&\quad + \int _I \int _{B_\rho } f(x,t)\psi (x,t) \;\textrm{d}x \;\textrm{d}t. \end{aligned} \end{aligned}$$

By Hölder’s inequality, we have

$$\begin{aligned}{} & {} \int _I \iint _{B_{\sigma } \times B_{\sigma }} \frac{|J_p(u(x,t)-u(y,t))(\psi (x,t) -\psi (y,t))|}{|x-y|^{n+sp}} \;\textrm{d}x \;\textrm{d}y \;\textrm{d}t \nonumber \\{} & {} \quad \le \int _I \Bigl \Vert \frac{J_p(u(x,t)-u(y,t))}{|x-y|^{\frac{n}{p^\prime } +s(p-1)}}\Bigr \Vert _{L^{p^\prime }(B_{\sigma }\times B_{\sigma })} \Bigl \Vert \frac{\psi (x,t)-\psi (y,t)}{|x-y|^{\frac{n}{p}+s}}\Bigr \Vert _{L^p(B_{\sigma }\times B_{\sigma })} \;\textrm{d}t \nonumber \\{} & {} \quad \le [u]_{L^p(I;W^{s,p}(B_{\sigma }))}^{p-1} [\psi ]_{L^p(I;W^{s,p}(B_{\sigma }))}. \end{aligned}$$
(3.3)

For the other nonlocal term, we note that for every \(x \in B_\rho \) and \(y \in {\mathbb {R}}^n\setminus B_{\sigma }\) we have \(|y| \le \frac{\sigma }{\sigma -\rho } |x-y|\). Hence,

$$\begin{aligned} \begin{aligned}&\int _{{\mathbb {R}}^n\setminus B_{\sigma }} \frac{|J_p(u(x,t)-u(y,t))|}{|x-y|^{n+sp}} \;\textrm{d}y \\&\quad \le \left( \frac{\sigma }{\sigma - \rho }\right) ^{n+sp} C(p) \int _{{\mathbb {R}}^n\setminus B_{\sigma }} \frac{|u(x,t)|^{p-1}+ |u(y,t)|^{p-1}}{|y|^{n+sp}} \;\textrm{d}y \\&\quad \le C(\sigma , \rho , s,p, n ) \Bigl ( |u(x,t)|^{p-1} + \Vert u(\,\bullet \,,t)\Vert ^{p-1}_{L_{sp}^{p-1}} \Bigr ). \end{aligned} \end{aligned}$$

Therefore,

$$\begin{aligned} \begin{aligned}&\int _I \int _{B_\rho } \int _{{\mathbb {R}}^n\setminus B_{\sigma }} \frac{|J_p(u(x,t)-u(y,t))\psi (x,t)|}{|x-y|^{n+sp}} \;\textrm{d}y \;\textrm{d}x \;\textrm{d}t \\&\quad \le C(\sigma , \rho , s,p, n ) \Bigl ( \int _I \int _{B_\rho } |\psi (x,t)| |u(x,t)|^{p-1} \;\textrm{d}x \;\textrm{d}t \\&\qquad + \int _I \Vert \psi (\,\bullet \,,t)\Vert _{L^1(B_\rho )} \Vert u(\,\bullet \,,t)\Vert ^{p-1}_{L_{sp}^{p-1}({\mathbb {R}}^n)} \;\textrm{d}t \Bigr ). \end{aligned} \end{aligned}$$

By Hölder’s inequality, we have

$$\begin{aligned} \int _I \int _{B_\rho } |\varphi (x,t)| |u(x,t)|^{p-1} \;\textrm{d}x \;\textrm{d}t\le & {} \int _I \Vert \psi (\,\bullet \,, t)\Vert _{L^p(B_\rho )} \Vert u(\,\bullet \,,t)\Vert _{L^{p}(B_\rho )}^{p-1} \nonumber \\{} & {} \le \Vert \psi \Vert _{L^p(I;L^p(B_\rho ))} \Vert u\Vert _{L^p(I;L^p(B_\rho ))}^{p-1}. \end{aligned}$$
(3.4)

For the other term,

$$\begin{aligned} \int _I \Vert \psi (\,\bullet \,,t)\Vert _{L^1(B_\rho )} \Vert u(\,\bullet \,,t)\Vert ^{p-1}_{L_{sp}^{p-1}({\mathbb {R}}^n)} \;\textrm{d}t \le \Vert \psi \Vert _{L^p(I;L^1(B_\rho ))} \Vert u(\,\bullet \,,t)\Vert _{L^p(I;L_{sp}^{p-1}({\mathbb {R}}^n))}^{p-1}. \nonumber \\ \end{aligned}$$
(3.5)

Since \(f \in L^{p^\prime }(I; L^{(p_s^\star )^\prime }(B_\rho ))\), by Hölder’s inequality and Sobolev’s inequality we obtain

$$\begin{aligned} \begin{aligned} \int _{I} \int _{B_{\rho }} |f \psi | \;\textrm{d}x \;\textrm{d}t&\le \int _{I} \Vert f\Vert _{L^{(p_s^\star )^\prime }(B_{\rho })} \Vert \psi \Vert _{L^{p_s^\star }(B_{\rho })} \;\textrm{d}t \\&\le \int _{I} \Vert f\Vert _{L^{(p_s^\star )^\prime }(B_{\rho })} \Vert \psi \Vert _{W^{s,p}(B_{\sigma })} \;\textrm{d}t \\&\le \Vert f\Vert _{L^{((p_s^\star )^\prime ,p^\prime )}(B_{\rho } \times I)} \Vert \psi \Vert _{L^p(I;W^{s,p}(B_\sigma ))}. \end{aligned}\nonumber \\ \end{aligned}$$
(3.6)

Therefore, combining with (3.3), (3.4), and (3.5) we obtain

$$\begin{aligned} \Bigl |\int _I \langle v_t,\psi \rangle \;\textrm{d}t\Bigr | \le C(\sigma , \rho , s,p,n,u,f) \Vert \psi \Vert _{L^p(I;W^{s,p}(B_\sigma ))}. \end{aligned}$$

\(\square \)

Lemma 3.2

Assume that \(f\in L^{q,r}_{\textrm{loc}}(Q_{\sigma ,\sigma ^{sp}}(x_0,T_0))\) with \(r\ge p^\prime \),

$$\begin{aligned} q\ge (p_s^\star )^\prime \quad \text {if}\quad sp<n,\quad q \ge 1 \quad \text {if} \quad sp>n, \quad \text {and}\quad q>1 \quad \text {if} \quad sp=n. \end{aligned}$$

Let u be a local weak solution of \(\partial _t u + (-\Delta _p)^s u= f\) in \(Q_{\sigma ,\sigma ^{sp}}(x_0,T_0)\), such that \(u \in L^p_{\textrm{loc}}((T_0-\sigma ^{sp},T_0]; L^{p-1}_{sp}({\mathbb {R}}^n))\). Let \(\rho < \sigma \) and consider v to be the (sp)-caloric replacement of u in \(Q_{\rho ,\rho ^{sp}}(x_0,T_0)\). Then we have

(3.7)

and

$$\begin{aligned} \Vert u-v\Vert _{L^{q^\prime ,r^\prime }(Q_{\rho ,\rho ^{sp}}(x_0,T_0))} \le C \rho ^{\xi + n} \Vert f\Vert _{L^{q,r}(Q_{\rho ,\rho ^{sp}}(x_0,T_0))}^{\frac{1}{p-1}}, \end{aligned}$$
(3.8)

with \(\xi = spp^\prime (1- \frac{1}{r} - \frac{n}{spq})\) and \(C= C(n,s,p)\), in the case \(sp\ne n\). In the case \(sp=n\), we can take \(\xi = spp^\prime (1 - \frac{1}{r} - \frac{1}{q})\), with \(C=C(n,s,p,q)\) also depending on q.

Proof

Let \(J:=[T_0-\rho ^{sp},T_0]\), throughout the proof, we drop the dependence of the balls on the center and write \(B_\rho \) instead of \(B_\rho (x_0)\), and \(Q_{\rho ,\rho ^{sp}}\) instead of \(Q_{\rho ,\rho ^{sp}}(x_0,T_0)\).

By subtracting the weak formulation of the equations (2.13) for u and v with the same test function \(\varphi (x,t) \in L^{p}(J;X_0^{s,p}(B_{\rho },B_{\sigma }))\cap C^1(J;L^2(B_\rho ))\), we get

$$\begin{aligned} \begin{aligned}&- \int _J \int _{B_\rho } (u(x,t)- v(x,t))\frac{\partial }{\partial t} \varphi (x,t) \;\textrm{d}x \;\textrm{d}t \\&\quad + \int _J \int _{{\mathbb {R}}^n} \int _{{\mathbb {R}}^n} \frac{\bigl [J_p \bigl (u (x,t) - u(y,t)\bigr ) - J_p\bigl ( v (x,t) - v(y,t)\bigr )\bigr ](\varphi (x,t) - \varphi (y,t))}{|x-y|^{n+sp}} \;\textrm{d}x \;\textrm{d}y \;\textrm{d}t \\&= \int _{B_\rho } ((u(x,T_0-\rho ^{sp}) - v(x,T_0-\rho ^{sp}))\varphi (x,T_0-\rho ^{sp}) \;\textrm{d}x \\&\quad - \int _{B_\rho } ((u(x,T_0) - v(x,T_0))\varphi (x,T_0) \;\textrm{d}x \\&\quad + \int _J \int _{B_\rho } f(x,t)\varphi (x,t) \;\textrm{d}x \;\textrm{d}t. \end{aligned} \end{aligned}$$

Now we take \(\varphi := u-v\), which belongs to \(L^p(J; X_0^{s,p}(B_{\rho };B_\sigma ))\), but it may not be in \(C^1(J;L^2(B_\rho ))\). We justify taking this as a test function in Appendix B. By Proposition 6.1 with \(F(t)=t\), we get

$$\begin{aligned} \int _J&\iint _{{\mathbb {R}}^n\times {\mathbb {R}}^n} \frac{\bigl [J_p \bigl (u (x,t) {-} v(x,t)\bigr ) {-} J_p\bigl ( u (y,t) {-} v(y,t)\bigr )\bigr ]\bigl [\bigl (u (x,t) {-} u(y,t)\bigr ) {-}\bigl ( v (x,t) {-} v(y,t)\bigr ) \bigr ]}{|x-y|^{n+sp}} \;\textrm{d}x \;\textrm{d}y \;\textrm{d}t \nonumber \\&= \int _J \int _{B_\rho } f(x,t)(u(x,t) -v(x,t)) \;\textrm{d}x\;\textrm{d}t \nonumber \\&\qquad - \frac{1}{2}\int _{B_\rho } ((u(x,T_0) - v(x,T_0))^2 -((u(x,T_0- \rho ^{sp}) - v(x,T_0 - \rho ^{sp}))^2 \;\textrm{d}x \nonumber \\&= \int _J \int _{B_\rho } f(x,t)(u(x,t) -v(x,t)) \;\textrm{d}x\;\textrm{d}t - \frac{1}{2}\int _{B_\rho } ((u(x,T_0) - v(x,T_0))^2 \;\textrm{d}x \nonumber \\&\le \int _J \int _{B_\rho } |f(x,t)(u(x,t) -v(x,t))| \;\textrm{d}x\;\textrm{d}t, \end{aligned}$$
(3.9)

where in the third line we have used \(u(x,T_0-\rho ^{sp}) = v(x,T_0-\rho ^{sp})\). The left-hand side is essentially the \(W^{s,\,p}\) seminorm. By the pointwise inequality (2.1),

$$\begin{aligned} \begin{aligned}&\int _J [u-v]_{W^{s,p}({\mathbb {R}}^n)}^p \;\textrm{d}t \\&\quad = \int _J \iint _{{\mathbb {R}}^n\times {\mathbb {R}}^n} \frac{|u(x,t)-v(x,t) - (u(y,t)-v(y,t))|^p}{|x-y|^{n+sp}} \;\textrm{d}x \;\textrm{d}y \;\textrm{d}t \\&\quad \le C(p) \int _J \iint _{{\mathbb {R}}^n\times {\mathbb {R}}^n}\\&\qquad \times \frac{\bigl [J_p \bigl (u (x,t) - u(y,t)\bigr ) - J_p\bigl ( v (x,t) - v(y,t)\bigr )\bigr ]\bigl [u (x,t) - u(y,t) -\bigl ( v (x,t) - v(y,t)\bigr ) \bigr ]}{|x-y|^{n+sp}} \;\textrm{d}x \;\textrm{d}y \;\textrm{d}t. \end{aligned} \end{aligned}$$

Therefore, by (3.9) and Hölder’s inequality

$$\begin{aligned} \int _J [u-v]_{W^{s,p}({\mathbb {R}}^n)}^p \;\textrm{d}t\le & {} C(p) \int _J \int _{B_\rho } |f(x,t)(u(x,t) -v(x,t))| \;\textrm{d}x\;\textrm{d}t \nonumber \\\le & {} C(p) \int _J \Vert f(\;\bullet ,t)\Vert _{L^q(B_\rho )}\Vert (u-v)(\;\bullet ,t)\Vert _{L^{q\prime }(B_\rho )} \;\textrm{d}t \nonumber \\\le & {} C(p) \Vert f\Vert _{L^{q,r}(Q_{\rho ,\rho ^{sp}})} \Vert u-v\Vert _{L^{q\prime ,r \prime }(Q_{\rho ,\rho ^{sp}})}. \end{aligned}$$
(3.10)

Now we consider three cases: \(sp<n\), \(sp>n\) and \(sp=n\).

Case \(sp < n\). By Hölder’s inequality (2.11) and Sobolev’s inequality (2.6), we have

$$\begin{aligned} \begin{aligned} \Vert u-v\Vert _{L^{q\prime ,r \prime }}&\le |B_\rho |^{ \frac{1}{q\prime } - \frac{1}{p_s^\star }} \Bigl ( \int _J \Vert u-v\Vert _{L^{p_s^\star }(B_\rho )}^{r\prime } \;\textrm{d}t \Bigr )^{\frac{1}{r \prime }}\\&\le C(n,s,p) |B_\rho |^{ \frac{1}{q\prime } - \frac{1}{p_s^\star }} \Bigl ( \int _J [u-v]_{W^{s,p}({\mathbb {R}}^n)}^{r\prime } \;\textrm{d}t \Bigr )^{\frac{1}{r\prime }} \\&\le C(n,s,p) |B_\rho |^{ \frac{1}{q\prime } - \frac{1}{p_s^\star }} |J|^{\frac{1}{r\prime } - \frac{1}{p}} \Bigl ( \int _J [u-v]_{W^{s,p}({\mathbb {R}}^n)}^p \;\textrm{d}t\Bigr )^{\frac{1}{p}}. \end{aligned} \nonumber \\ \end{aligned}$$
(3.11)

Combined with (3.10), this yields

$$\begin{aligned} \begin{aligned} \Bigl ( \int _J [u-v]_{W^{s,p}({\mathbb {R}}^n)}^p \;\textrm{d}t\Bigr )^{\frac{p-1}{p}}&\le C |B_\rho |^{ \frac{1}{q\prime } - \frac{1}{p_s^\star }} |J|^{\frac{1}{r\prime } - \frac{1}{p}} \Vert f\Vert _{L^{q,r}(Q_{\rho ,\rho ^{sp}})}\\&= C |B_\rho |^{ \frac{1}{q\prime } - \frac{n-sp}{np}} |J|^{\frac{1}{r\prime } - \frac{1}{p}} \Vert f\Vert _{L^{q,r}(Q_{\rho ,\rho ^{sp}})}, \end{aligned} \nonumber \\ \end{aligned}$$
(3.12)

where \(C= C(n,s,p)\). By the Poincaré inequality,

Also from (3.12) and (3.11), we get

$$\begin{aligned} \Vert u-v\Vert _{L^{q\prime , r\prime }(Q_{\rho ,\rho ^{sp}})} \le C(n,s,p)|B_\rho |^{ \frac{p\prime }{q\prime } - p\prime \frac{n-sp}{np}} |J|^{\frac{p\prime }{r\prime } - \frac{p\prime }{p}} \Vert f\Vert _{L^{q,r}(Q_{\rho ,\rho ^{sp}})}^{\frac{1}{p-1}}. \end{aligned}$$

Case \(sp>n\). In this case, we use Morrey’s inequality (2.7) and Hölder’s inequality and obtain

$$\begin{aligned} \begin{aligned} \Vert u-v\Vert _{L^{q\prime ,r \prime }(Q_{\rho ,\rho ^{sp}})}&\le C |B_\rho |^{ \frac{1}{q\prime }} \Bigl ( \int _J \Vert u-v\Vert _{L^{\infty }(B_\rho )}^{r\prime } \;\textrm{d}t \Bigr )^{\frac{1}{r \prime }} \\&\le C |B_\rho |^{\frac{1}{q^\prime }}|J|^{\frac{1}{r\prime } - \frac{1}{p}} \Bigl ( \int _J \Vert u-v\Vert _{L^{\infty }(B_\rho )}^{p} \;\textrm{d}t \Bigr )^{\frac{1}{p}}\\&\le C |B_\rho |^{ \frac{1}{q\prime } + \frac{sp-n}{np}} |J|^{\frac{1}{r\prime } - \frac{1}{p}} \Bigl ( \int _J [u-v]_{W^{s,p}({\mathbb {R}}^n)}^p \;\textrm{d}t\Bigr )^{\frac{1}{p}}. \end{aligned} \nonumber \\ \end{aligned}$$
(3.13)

Together with (3.10), this implies

$$\begin{aligned} \Bigl ( \int _J [u-v]_{W^{s,p}({\mathbb {R}}^n)}^p \;\textrm{d}t\Bigr )^{\frac{p-1}{p}} \le C |B_\rho |^{ \frac{1}{q\prime } - \frac{n-sp}{np}} |J|^{\frac{1}{r\prime } - \frac{1}{p}} \Vert f\Vert _{L^{q,r}(Q_{\rho ,\rho ^{sp}})}. \end{aligned}$$
(3.14)

By the Poincaré inequality,

Combining (3.13) and (3.14), we get

$$\begin{aligned} \Vert u-v\Vert _{L^{q\prime , r\prime }(Q_{\rho ,\rho ^{sp}})} \le C(n,s,p)|B_\rho |^{ \frac{p\prime }{q\prime } - p\prime \frac{n-sp}{np}} |J|^{\frac{p\prime }{r\prime } - \frac{p\prime }{p}} \Vert f\Vert _{L^{q,r}(Q_{\rho ,\rho ^{sp}})}^{\frac{1}{p-1}}. \end{aligned}$$

Case \(sp=n\). In this case, we use the critical case of Sobolev’s inequality (2.8) for \( l= q^\prime \) and obtain

$$\begin{aligned} \Vert u-v\Vert _{L^{q^\prime }(B_\rho )}^p \le C(n,s,p,q) |B_\rho |^{\frac{p}{q^\prime }} [u-v]_{W^{s,p}({\mathbb {R}}^n)}^p. \end{aligned}$$

Hence, using Hölder’s inequality, we have for any \(r\ge p^\prime \)

$$\begin{aligned} \begin{aligned} \Vert u-v\Vert _{L^{q\prime ,r \prime }(Q_{\rho ,\rho ^{sp}})}&= \Bigl ( \int _J \Vert u-v\Vert _{L^{q^\prime }(B_\rho )}^{r\prime } \;\textrm{d}t \Bigr )^{\frac{1}{r \prime }}\\&\le C |B_\rho |^{ \frac{1}{q\prime } } \Bigl ( \int _J [u-v]_{W^{s,p}({\mathbb {R}}^n)}^{r\prime } \;\textrm{d}t \Bigr )^{\frac{1}{r\prime }} \\&\le C |B_\rho |^{ \frac{1}{q\prime }} |J|^{\frac{1}{r\prime } - \frac{1}{p}} \Bigl ( \int _J [u-v]_{W^{s,p}({\mathbb {R}}^n)}^p \;\textrm{d}t\Bigr )^{\frac{1}{p}}. \end{aligned} \end{aligned}$$

The above constant \(C=C(n,s,p,q)\) does blow up as q goes to 1. In a similar way as in the prior cases, we get for \(q>1\) and \(r\ge p^\prime \)

and

$$\begin{aligned} \Vert u-v\Vert _{L^{q\prime , r\prime }(Q_{\rho ,\rho ^{sp}})} \le C(n,s,p,q)|B_\rho |^{ \frac{p\prime }{q\prime } } |J|^{\frac{p\prime }{r\prime } - \frac{p\prime }{p}} \Vert f\Vert _{L^{q,r}(Q_{\rho ,\rho ^{sp}})}^{\frac{1}{p-1}}. \end{aligned}$$

Using that \(|B_\rho |\sim \rho ^{n}\) and \(|I| \sim \rho ^{sp}\), we can conclude that

and

$$\begin{aligned} \Vert u-v\Vert _{L^{q^\prime ,r^\prime }(Q_{\rho ,\rho ^{sp}})} \le C \rho ^{\xi + n} \Vert f\Vert _{L^{q,r}(Q_{\rho ,\rho ^{sp}})}^{\frac{1}{p-1}}. \end{aligned}$$

Here in the case of \(sp\ne n\),

$$\begin{aligned} \begin{aligned} \xi&= \frac{n p^\prime }{q^\prime } -p^\prime \frac{n-sp}{p} +sp -n + \frac{spp^\prime }{r^\prime } - \frac{spp^\prime }{p} -sp\\&= p^\prime \left( \frac{n}{q^\prime } - \frac{n}{p^\prime } - \frac{n - sp}{p} + \frac{sp}{r^\prime } - \frac{sp}{p}\right) \\&= p^\prime \left( \frac{n}{q^\prime } - n + \frac{sp}{r^\prime } \right) = p^\prime \left( \frac{sp}{r^\prime } - \frac{n}{q}\right) = spp^\prime \left( 1- \frac{1}{r} - \frac{n}{spq}\right) , \end{aligned} \end{aligned}$$

and in the case \(sp=n\),

$$\begin{aligned} \begin{aligned} \xi = p^\prime \left( \frac{n}{q^\prime } + \frac{sp}{r^\prime } - \frac{sp}{p} -\frac{sp}{p^\prime }\right)&= sp p^\prime \left( \frac{1}{q^\prime } +\frac{1}{r^\prime } - \frac{1}{p}-\frac{1}{p^\prime }\right) \\&= sp p^\prime \left( 1-\frac{1}{q} + 1-\frac{1}{r}-1\right) \\&=spp^\prime \left( 1 - \frac{1}{r} - \frac{1}{q}\right) . \end{aligned} \end{aligned}$$

\(\square \)

Remark 3.3

In Lemma 3.2, we assume the same conditions as in Proposition 3.1. These assumptions are used in the proof not only to ensure the existence of the (sp)-caloric replacement but also to derive (3.11) and (3.13). As mentioned in Remark 2.13, one can expect the existence of the (sp)-caloric replacement under a more general condition for the right-hand side. If such an existence theorem is available, one can expect the estimates in Lemma 3.2 to hold true for more general right-hand sides. In the proof of Lemma 3.2, we only used the diffusion term in (3.9), but the stronger estimate

$$\begin{aligned}{} & {} \sup _{t \in J} \Vert (u-v)(\,\bullet \,,t)\Vert _{L^2(B_\rho (x_0))}^2 + \int _{J}[(u-v)(\,\bullet \,, t)]_{W^{s,p}({\mathbb {R}}^n)}^p \;\textrm{d}t \\{} & {} \quad \le C \int _J \int _{B_\rho (x_0)} |f(x,t)(u-v)(x,t)| \;\textrm{d}x \;\textrm{d}t, \end{aligned}$$

holds true. It might be possible to utilize an interpolation argument similar to Lemma 2.8 to replace the equations (3.11) and (3.13) and relax the assumptions on q and r. See also [12, Lemma 2.2]. However, the nonhomogeneity of the equation is for sure a challenge in pursuing this line of reasoning.

Next, we perform a Moser iteration to get an \(L^\infty \) bound for the difference between the solution and its (sp)-caloric replacement.

Proposition 3.4

Let u be a local weak solution of

$$\begin{aligned} \partial _t u + (-\Delta _p)^s u=f, \quad \text {in} \quad Q_{\sigma ,\sigma ^{sp}}(x_0,T_0), \end{aligned}$$

with \(f \in L^{q,r}_{\textrm{loc}} (Q_{\sigma ,\sigma ^{sp}}(x_0,T_0))\) such that

$$\begin{aligned} \frac{1}{r} + \frac{n}{spq} < 1. \end{aligned}$$

In addition, assume that \(u\in L^p_{\textrm{loc}} \bigl ( (T_0-\sigma ^{sp},T_0];L_{sp}^{p-1}({\mathbb {R}}^n)\bigr )\), \(r\ge p^\prime \),

$$\begin{aligned} q \ge 1 \quad \text {if}\quad sp\ne n, \quad \text {and}\quad q>1 \quad \text {if} \quad sp=n. \end{aligned}$$

Let v be the (sp)-caloric replacement of u in \(Q_{R,R^{s p}}(x_0,T_0)\), with \(R< \sigma \). Then in the case of \(sp \ne n\), we have

$$\begin{aligned} \Vert (u-v)^+\Vert _{L^\infty (Q_{RR^{s p}}(x_0,T_0))} \le C(n,s,p) \vartheta ^{\frac{(p-1)\vartheta }{(\vartheta -1 )^2}} \Bigl ( 1 + R^{sp\nu }\Vert f\Vert _{L^{q,r}\left( Q_{R,R^{sp}}(x_0,T_0)\right) } \Bigr ), \end{aligned}$$

where \(\nu = 1-\frac{1}{r} - \frac{n}{spq}\) and

$$\begin{aligned} \vartheta = 1+\frac{sp\nu }{n}. \end{aligned}$$

In the case of \(sp=n\), given any l such that \( \frac{p}{r^\prime }(1-\frac{1}{r} - \frac{1}{q})^{-1}<l < \infty \) we get

$$\begin{aligned} \Vert (u{-}v)^+\Vert _{L^\infty (Q_{R,R^{s p}}(x_0,T_0))}{\le } C(n,s,p,q,l) \vartheta ^{\frac{(p-1)\vartheta }{(\vartheta - 1)^2}} \Bigl ( 1 + R^{sp\nu }\Vert f\Vert _{L^{q,r}\left( Q_{R,R^{sp}}(x_0,T_0)\right) } \Bigr ), \end{aligned}$$

where \(\vartheta = 2-\frac{1}{r}- \frac{1}{q} - \frac{p}{lr^\prime }\) and \(\nu = 1-\frac{1}{r} - \frac{1}{q}\).

Proof

Throughout the proof, we write \(Q_{R,R^{sp}}\) instead of \(Q_{R,R^{s,p}}(x_0,T_0)\) and \(B_R\) instead of \(B_R(x_0)\). We also define the interval J to be \(J:= (T_0-R^{sp},T_0]\). First, we verify that our assumptions ensure that the (sp)-caloric replacement of u exists. If \(sp \ge n\), we have explicitly assumed what is needed to use Proposition 3.1. If \(sp< n\), we have to verify that \(q\ge (p_s^\star )^\prime \). This follows from the assumption \(\frac{1}{r} + \frac{n}{spq} < 1\). Indeed

$$\begin{aligned} \frac{n}{sp}\le q\left( 1- \frac{1}{r}\right) \le q, \end{aligned}$$

and it is straightforward to verify that \(\frac{n}{sp}\ge (p_s^\star )^\prime \). This shows that v, the (sp)-caloric replacement of u exist. Let us also mention that the assumptions in Lemma 3.2 are the same as in Proposition 3.1, and we can use this lemma. Now, we test the equations with powers of \(u-v\) and perform a Moser iteration. Using Proposition 6.1 with

$$\begin{aligned} F(t)=( \min { \lbrace t^{+} \,,M \rbrace } +\delta )^\beta - \delta ^\beta , \end{aligned}$$

and

$$\begin{aligned} \delta = \max { \left\{ 1, R^{sp\nu }\Vert f\Vert _{L^{q,r}(Q_{R,R^{sp}})} \right\} }, \end{aligned}$$
(3.15)

we get

$$\begin{aligned}{} & {} \sup _{t \in J} \int _{B_R} {\mathcal {F}}(u-v) \;\textrm{d}x + \int _J \iint _{{\mathbb {R}}^n \times {\mathbb {R}}^n} \frac{J_{p}(u(x,t)-u(y,t)) - J_p(v(x,t)-v(y,t))}{|x-y|^{n+sp}} \nonumber \\{} & {} \qquad \times (F(u(x,t)-v(x,t)) - F(u(y,t)-v(y,t))) \;\textrm{d}x \;\textrm{d}y \;\textrm{d}t\nonumber \\{} & {} \quad \le \int _J \int _{B_R} |f(x,t)| F(u(x,t)-v(x,t)) \nonumber \\{} & {} \quad \le \Vert f\Vert _{L^{q,r}(B_R \times J)} \Vert ((u-v)^{+}_M + \delta )^{\beta }\Vert _{L^{q^\prime , r^\prime }(B_R \times J)}. \end{aligned}$$
(3.16)

In the last line, we have used Hölder’s inequality. Here \({\mathcal {F}}(t) = \int _0^t F(t) \;\textrm{d}t\) is

$$\begin{aligned} {\mathcal {F}}(t) = {\left\{ \begin{array}{ll} \qquad \qquad 0 \quad &{} \text {if} \quad t\le 0, \\ \quad \frac{1}{\beta + 1}(t+\delta )^{\beta +1} -\frac{\delta ^{\beta +1}}{\beta + 1} - t \delta ^\beta \quad &{}\text {if} \quad 0\le t \le M, \\ \frac{1}{\beta + 1}(M +\delta )^{\beta +1} -\frac{\delta ^{\beta +1}}{\beta + 1} - t \delta ^\beta + (t-M)(M+\delta )^\beta &{} \text {if} \quad t \ge M. \end{array}\right. } \end{aligned}$$

Notice that by Young’s inequality, for \(t \ge 0\)

$$\begin{aligned} \frac{(t+\delta )^{\beta + 1}}{2(\beta + 1)} + \frac{\beta }{\beta + 1} 2\delta ^{\beta + 1} \ge \frac{t+\delta }{2^{\frac{1}{\beta + 1} }} 2^{\frac{\beta }{\beta + 1}} \delta ^\beta \ge t \delta ^\beta \,. \end{aligned}$$

In particular, for \(0\le t \le M\)

$$\begin{aligned} {\mathcal {F}}(t) \ge \frac{(t+\delta )^{\beta + 1}}{2(\beta + 1)} - \frac{2\beta + 1}{\beta +1} \delta ^{\beta + 1} \ge \frac{(t+\delta )^{\beta + 1}}{2(\beta + 1)} - 2\delta ^{\beta + 1}, \end{aligned}$$

and for \(t\ge M\)

$$\begin{aligned}{} & {} \frac{1}{\beta + 1}(M +\delta )^{\beta +1} -\frac{\delta ^{\beta +1}}{\beta + 1} - t \delta ^\beta + (t-M)(M+\delta )^\beta \\{} & {} \quad = \frac{1}{\beta + 1}(M +\delta )^{\beta +1} -\frac{\delta ^{\beta +1}}{\beta + 1} - M \delta ^\beta + (t-M)\bigl ((M+\delta )^\beta -\delta ^\beta \bigr ) \ge {\mathcal {F}}(M)\\{} & {} \quad \ge \frac{(M+\delta )^{\beta +1}}{2(\beta + 1)} - 2\delta ^{\beta + 1}. \end{aligned}$$

Hence,

$$\begin{aligned} {\mathcal {F}}(t) \ge \frac{(t^{+}_M +\delta )^{\beta +1}}{2(\beta + 1)} - 2\delta ^{\beta + 1}. \end{aligned}$$
(3.17)

Using Lemma 2.2 for the second term in the left-hand side of (3.16) and (3.17) in the first term, we obtain

$$\begin{aligned}&\frac{1}{2(\beta + 1)} \sup _{t \in J} \int _{B_R} ((u-v)^{+}_M + \delta )^{\beta +1} \;\textrm{d}x\nonumber \\&\qquad + \frac{1}{3\cdot 2^{p-1}}\frac{\beta p^p}{(\beta +p -1)^p}\int _J \left[ ((u-v)^{+}_M + \delta )^{\frac{\beta +p-1}{p}}\right] _{W^{s, p}({\mathbb {R}}^n)}^p \;\textrm{d}t \nonumber \\&\quad \le \sup _{t \in J} \int _{B_R} {\mathcal {F}}(u-v) \;\textrm{d}x + 2\delta ^{\beta +1} |B_R| \nonumber \\&\qquad + \int _J \iint _{{\mathbb {R}}^n \times {\mathbb {R}}^n} \frac{J_{p}(u(x,t)-u(y,t)) - J_p(v(x,t)-v(y,t))}{|x-y|^{n+sp}} \nonumber \\&\qquad \times (F(u(x,t)-v(x,t)) - F(u(y,t)-v(y,t))) \;\textrm{d}x \;\textrm{d}y \;\textrm{d}t \nonumber \\&\quad \le \Vert f\Vert _{L^{q,r}(B_R \times J)} \Vert ((u-v)^{+}_M+ \delta )^{\beta }\Vert _{L^{q^\prime , r^\prime }(B_R \times J)} + 2\delta ^{\beta +1} |B_R|. \end{aligned}$$
(3.18)

Let \(w(x,t) = ((u-v)^{+}_M +\delta )^{\frac{\beta }{p}}\). Since \(\delta \le (u-v)^{+}_M + \delta \), we see that

$$\begin{aligned} \delta ^\beta \le \frac{\Vert w\Vert _{L^{pq^\prime , p r^\prime }(Q_{R,R^{s p}})}^p}{|B_R|^{1-\frac{1}{q}} |J|^{1- \frac{1}{r}}}. \end{aligned}$$
(3.19)

Using (3.19) in (3.18), we get

$$\begin{aligned}&\frac{\delta }{2(\beta + 1)} \Vert w\Vert _{L^{p,\infty }(B_R \times J)}^p \, + \frac{1}{3\cdot 2^{p-1}}\frac{\beta p^p}{(\beta +p -1)^p}\int _J \left[ w^{\frac{\beta +p-1}{\beta }}\right] _{W^{s, p}({\mathbb {R}}^n)}^p \;\textrm{d}t\nonumber \\&\quad \le \Vert f\Vert _{L^{q,r}(B_R \times J)} \Vert w\Vert _{L^{p q^\prime , p r^\prime }(B_R \times J)}^p \, + 2\delta |B_R| \frac{ \Vert w\Vert _{L^{p q^\prime , p r^\prime }(B_R \times J)}^p}{|B_R|^{1-\frac{1}{q}}|J|^{1-\frac{1}{r}}}. \end{aligned}$$
(3.20)

By (2.4), we have

$$\begin{aligned} \begin{aligned}{}[w^{\frac{\beta +p-1}{\beta }}]_{W^{s, p}({\mathbb {R}}^n)}^p&= \iint _{{\mathbb {R}}^n \times {\mathbb {R}}^n} \frac{|w(x)^{\frac{\beta +p-1}{\beta }}- w(y)^{\frac{\beta +p-1}{\beta }}|^p}{|x-y|^{n+sp}} \;\textrm{d}x \;\textrm{d}y \\&\ge \left( \frac{\beta +p-1}{\beta } \right) ^p \min \lbrace w(x),w(y)\rbrace ^{p(\frac{\beta +p-1}{\beta }-1)}\\&\quad \times \iint _{{\mathbb {R}}^n \times {\mathbb {R}}^n} \frac{|w(x)- w(y)|^p}{|x-y|^{n+sp}} \;\textrm{d}x \;\textrm{d}y \\&\ge \left( \frac{\beta +p-1}{\beta } \right) ^p \delta ^{p-1}[w]_{W^{s,p}({\mathbb {R}}^n)}^p. \end{aligned} \end{aligned}$$

Using this in (3.20) and since J has length \(R^{sp}\), we arrive at

$$\begin{aligned} \begin{aligned} \frac{\delta }{2(\beta + 1)}&\Vert w\Vert _{L^{p,\infty }(Q_{R,R^{sp}})}^p \, + \frac{1}{3\cdot 2^{p-1}}\frac{\delta ^{p-1} p^p}{\beta ^{p-1}}\int _J [w]_{W^{s, p}({\mathbb {R}}^n)}^p \;\textrm{d}t \\&\le \Vert w\Vert _{L^{p q^\prime , p r^\prime }(Q_{R,R^{sp}})}^p \left( \Vert f\Vert _{L^{q,r}(Q_{R,R^{sp}})} + \frac{2(n \omega _n)^{\frac{1}{q}}\delta R^n}{R^{n(1-\frac{1}{q}) + sp(1-\frac{1}{r})}} \right) . \end{aligned} \end{aligned}$$

Upon multiplying both sides by \(\frac{3 \cdot 2^{p-1} \cdot \beta ^{p-1}}{\delta }\), this implies

$$\begin{aligned} \begin{aligned} 3&\cdot 2^{p-2}\frac{\beta ^{p-1}}{\beta +1 }\Vert w\Vert _{L^{p,\infty }(Q_{R,R^{s p}})}^p + \delta ^{p-2} p^p \int _J [w]_{W^{s, p}({\mathbb {R}}^n)}^{p} \;\textrm{d}t \\&\le 3 \cdot 2^{p-1}\times \frac{\beta ^{p-1}}{\delta } \Vert w\Vert _{L^{pq^\prime ,pr^\prime }(Q_{R,R^{sp}})}^p \bigl ( \Vert f\Vert _{L^{q,r}(Q_{R,R^{sp}})} +2(n \omega _n)^{\frac{1}{q}}\delta R^{-sp\nu }\bigr ). \end{aligned} \end{aligned}$$
(3.21)

Since \(\delta \ge 1\) and \(p \ge 2\), for \(\beta \ge 1\) we have

$$\begin{aligned} \begin{aligned} \Vert w\Vert&_{L^{p,\infty }(Q_{R,R^{s p}})}^p + \int _J [w]_{W^{s, p}({\mathbb {R}}^n)}^p \;\textrm{d}t \\&\le 3 \cdot 2^{p-2}\frac{\beta ^{p-1}}{\beta +1 }\Vert w\Vert _{L^{p,\infty }(Q_{R,R^{s p}})}^p + \delta ^{p-2} p^p \int _J [w]_{W^{s, p}({\mathbb {R}}^n)}^{p} \;\textrm{d}t. \end{aligned} \end{aligned}$$

Using this in (3.21), we get

$$\begin{aligned}{} & {} \Vert w\Vert _{L^{p,\infty }(Q_{R,R^{s p}})}^p + \int _J [w]_{W^{s, p}({\mathbb {R}}^n)}^p \;\textrm{d}t \nonumber \\{} & {} \quad \le 3\cdot 2^{p-1}\beta ^{p-1} \Vert w\Vert _{L^{p q^\prime , p r^\prime }(Q_{R,R^{s p}})}^p \Bigl ( \frac{\Vert f\Vert _{L^{q,r}(Q_{R,R^{s p}})}}{\delta } + 2(n \omega _n)^{\frac{1}{q}} R^{-sp\nu }\Bigr ) \nonumber \\{} & {} \quad \le C \beta ^{p-1} \Vert w\Vert _{L^{p q^\prime , p r^\prime }(Q_{R,R^{s p}})}^p \Bigl ( \frac{\Vert f\Vert _{L^{q,r}(Q_{R,R^{sp}})}}{\delta } + R^{-sp\nu } \Bigr ), \end{aligned}$$
(3.22)

where \(C=C(n,p)\). Now we consider two cases depending on whether \(sp \ne n\) or \(sp=n\).

Case \(sp \ne n\): Notice that since \(\nu > 0 \), if we take \(\vartheta = 1+ \frac{sp\nu }{n}\), the exponents \((\vartheta r^\prime )^\prime , (\vartheta q^\prime )^\prime \) satisfy the condition of Lemma 2.8. Indeed,

$$\begin{aligned}{} & {} 1- \frac{1}{(\vartheta r^\prime )^\prime } - \frac{n}{sp (\vartheta q^\prime )^\prime } = \frac{1}{\vartheta r^\prime } + \frac{n}{sp \vartheta q^\prime } - \frac{n}{sp}\\{} & {} \quad = \frac{1}{\vartheta }\left( \frac{1}{r^\prime } +\frac{n}{spq^\prime }- \frac{\vartheta n}{sp}\right) = \frac{1}{\vartheta }\left( \nu +\frac{n}{sp} -\frac{\vartheta n}{sp}\right) =0. \end{aligned}$$

As \(w- \delta ^{\frac{\beta }{p}}\) does vanish in \(B_{R}(x_0)^c\), using Lemma 2.8 for the exponents \((\vartheta q^\prime )^\prime \) and \((\vartheta r^\prime )^\prime \) we get

$$\begin{aligned}{} & {} \Vert w -\delta ^{\frac{\beta }{p}}\Vert _{L^{\vartheta p q^\prime , \vartheta p r^\prime }(Q_{R,R^{sp}})}^p \nonumber \\{} & {} \quad \le C(n,s,p) \left( \Vert w-\delta ^{\frac{\beta }{p}}\Vert _{L^{p,\infty }(Q_{R,R^{s p}})}^p + \int _{T_0-R^{sp}}^{T_0} [w-\delta ^{\frac{\beta }{p}}]_{W^{s,p}({\mathbb {R}}^n)}^p \;\textrm{d}t \right) \nonumber \\{} & {} \quad \le C(n,s,p) \left( \Vert w\Vert _{L^{p,\infty }(Q_{R,R^{s p}})}^p + \int _{T_0-R^{sp}}^{T_0} [w]_{W^{s,p}({\mathbb {R}}^n)}^p \;\textrm{d}t \right) \nonumber \\ \text {by }(3.22){} & {} \le C(n,s,p)\beta ^{p-1} \Vert w\Vert _{L^{p q^\prime , p r^\prime }(Q_{R,R^{s p}})}^p \Bigl ( \frac{\Vert f\Vert _{L^{q,r}(Q_{R,R^{sp}})}}{\delta } + R^{-sp\nu } \Bigr ). \end{aligned}$$
(3.23)

Here we have used that \(w- \delta ^{\frac{\beta }{p}}\) is nonnegative as well as the fact that \([w]_{W^{s,p}({\mathbb {R}}^n)}\) does not change by subtracting a constant from w. Hence, by (3.19) and (3.23) we obtain

$$\begin{aligned}{} & {} \Vert w^\vartheta \Vert _{L^{ p q^\prime , p r^\prime }(Q_{R,R^{sp}})}^{\frac{p}{\vartheta }} \nonumber \\{} & {} \quad = \Vert w\Vert _{L^{\vartheta p q^\prime , \vartheta p r^\prime }(Q_{R,R^{sp}})}^p \le \left( \Vert w- \delta ^{\frac{\beta }{p}}\Vert _{L^{\vartheta p q^\prime , \vartheta p r^\prime }(Q_{R,R^{sp}})}+ \delta ^{\frac{\beta }{p}} R^{\frac{n}{\vartheta p q^\prime }+ \frac{sp}{\vartheta p r^\prime }} \right) ^p\nonumber \\{} & {} \quad \le 2^{p-1} \left( \Vert w- \delta ^{\frac{\beta }{p}}\Vert _{L^{\vartheta p q^\prime , \vartheta p r^\prime }(Q_{R,R^{sp}})}^p + \delta ^\beta R^{\frac{n}{\vartheta q^\prime }+ \frac{sp}{\vartheta r^\prime }} \right) \nonumber \\{} & {} \quad \le C\beta ^{p-1} \Vert w\Vert _{L^{p q^\prime , p r^\prime }(Q_{R,R^{s p}})}^p \left( \frac{\Vert f\Vert _{L^{q,r}(Q_{R,R^{sp}})}}{\delta } + R^{-sp\nu } \right) \nonumber \\{} & {} \qquad + 2^{p-1}\delta ^\beta R^{\frac{n}{\vartheta q^\prime }+ \frac{sp}{\vartheta r^\prime }}\nonumber \\{} & {} \quad \le C\beta ^{p-1} \Vert w\Vert _{L^{p q^\prime , p r^\prime }(Q_{R,R^{s p}})}^p \left( \frac{\Vert f\Vert _{L^{q,r}(Q_{R,R^{sp}})}}{\delta } + R^{-sp\nu } \right) \nonumber \\{} & {} \qquad + 2^{p-1}(n \omega _n)^{\frac{1}{q}-1}\frac{\Vert w\Vert _{L^{p q^\prime , p r^\prime }(Q_{R,R^{s p}})}^p }{R^{\frac{n}{q^\prime }+\frac{sp}{r^\prime } -\frac{n}{\vartheta q^\prime }- \frac{sp}{\vartheta r^\prime }}} \nonumber \\{} & {} \quad \le C\beta ^{p-1} \Vert w\Vert _{L^{p q^\prime , p r^\prime }(Q_{R,R^{s p}})}^p \left( \frac{\Vert f\Vert _{L^{q,r}(Q_{R,R^{sp}})}}{\delta } + R^{-sp\nu } + R^{(\frac{1}{\vartheta }-1)(sp\nu + n)} \right) .\nonumber \\ \end{aligned}$$
(3.24)

Observe that \((\frac{1}{\vartheta }-1)(sp\nu + n)= - \frac{\vartheta -1}{\vartheta }(sp\nu +n) = -sp\nu \). Furthermore, recalling the definition of \(\delta \) (3.15) whenever \(\delta > 1\) we have

$$\begin{aligned} \frac{\Vert f\Vert _{L^{q,r}(Q_{R,R^{sp}})}}{\delta } = R^{-sp\nu }. \end{aligned}$$

When \(\delta = 1 \), it is straightforward to verify that

$$\begin{aligned} \frac{\Vert f\Vert _{L^{q,r}(Q_{R,R^{sp}})}}{\delta } \le R^{-sp\nu }. \end{aligned}$$

Inserting these into (3.24), we arrive at

$$\begin{aligned} \Vert w^\vartheta \Vert _{L^{ p q^\prime , p r^\prime }(Q_{R,R^{sp}})}^{\frac{p}{\vartheta }} \le C \beta ^{p-1} R^{-sp\nu }\Vert w\Vert _{L^{p q^\prime , p r^\prime }(Q_{R,R^{s p}})}^p. \end{aligned}$$
(3.25)

Now we iterate this inequality with the following choice of exponents

$$\begin{aligned} \beta _0 = 1,\qquad \beta _{m+1} = \vartheta \beta _m = \vartheta ^{m+1}. \end{aligned}$$

With the notation

$$\begin{aligned} \varphi _m:= \Vert ((u-v)^{+}_M +\delta )^\frac{\beta _m}{p} \Vert _{L^{p q^\prime , pr^\prime }(Q_{R,R^{sp}})}^\frac{p}{\beta _m} =\Vert (u-v)^{+}_M + \delta \Vert _{L^{\beta _m q^\prime , \beta _m r^\prime }(Q_{R,R^{sp}})}, \end{aligned}$$

(3.25) reads

$$\begin{aligned} \varphi _{m+1} \le \bigl (C \, R^{-sp\nu } \bigr )^\frac{1}{\vartheta ^m}\vartheta ^\frac{(p-1)m}{\vartheta ^m} \varphi _m. \end{aligned}$$

Iterating this yields

$$\begin{aligned} \varphi _{m+1} \le \bigl (C \, R^{-sp\nu } \bigr )^{\sum _{j=0}^m \vartheta ^{-j}} \vartheta ^{(p-1) \sum _{j=0}^m j \vartheta ^{-j}} \varphi _0. \end{aligned}$$
(3.26)

Since \(\vartheta >1 \), we have the following convergent series

$$\begin{aligned} \sum _{j=0}^\infty \vartheta ^{-j} = \frac{\vartheta }{\vartheta -1}=\frac{n+sp\nu }{sp\nu }, \end{aligned}$$

and

$$\begin{aligned} \sum _{j=0}^\infty j \vartheta ^{-j} = \frac{\vartheta }{(\vartheta -1)^2} = \frac{n^2 + nsp\nu }{s^2 p^2 \nu ^2}. \end{aligned}$$

By (3.8) in Lemma 3.2,

$$\begin{aligned} \varphi _0= & {} \Vert (u-v)^{+}_M {+}\delta \Vert _{L^{q^\prime , r^\prime }(Q_{R,R^{sp}})} {\le } C(n,s,p)R^{spp^\prime \nu {+}n} \Vert f\Vert _{L^{q,r}(Q_{R,R^{sp}})}^{\frac{1}{p-1}} {+} \delta (n \omega _n)^{\frac{1}{q^\prime }} R^{\frac{n}{q^\prime } {+} \frac{sp}{r^\prime }} \nonumber \\= & {} C(n,s,p) R^{n+sp\nu } (R^{\frac{sp\nu }{p-1}}\Vert f\Vert _{L^{q,r}(Q_{R,R^{sp}})}^{\frac{1}{p-1}} +\delta ) \le C(n,s,p) R^{n+sp\nu } \left( \delta ^\frac{1}{p-1} + \delta \right) \nonumber \\\le & {} C(n,s,p)R^{n+sp\nu } \delta . \end{aligned}$$
(3.27)

In the last line, we have used that \(p-1\ge 1\) and \(\delta \ge 1\). Inserting (3.27) to (3.26) and sending m to infinity, we obtain

$$\begin{aligned} \begin{aligned} \Vert (u-v)^{+}_M +\delta \Vert _{L^\infty (Q_{R,R^{sp}})}&\le C \vartheta ^{\frac{(p-1)\vartheta }{(\vartheta -1 )^2}} R^{-n - sp \nu } R^{sp \nu + n} \delta \\&= C \vartheta ^{\frac{(p-1)\vartheta }{(\vartheta -1 )^2}} \delta \\&\le C \vartheta ^{\frac{(p-1)\vartheta }{(\vartheta -1 )^2}} \max {\lbrace 1, R^{sp\nu }\Vert f\Vert _{L^{q,r}(Q_{R,R^{sp}})} \rbrace } \\&\le C(n,s,p) \vartheta ^{\frac{(p-1)\vartheta }{(\vartheta -1 )^2}} \bigl ( 1 + R^{sp\nu } \Vert f\Vert _{L^{q,r}(Q_{R,R^{sp}})} \bigr ). \end{aligned} \end{aligned}$$

Since the above estimate is independent of M, we get

$$\begin{aligned} \begin{aligned}&\Vert (u-v)^{+}\Vert _{L^\infty (Q_{R,R^{sp}})} \le C(n,s,p) \vartheta ^{\frac{(p-1)\vartheta }{(\vartheta -1 )^2}} \Bigl ( 1 + R^{sp\nu } \Vert f\Vert _{L^{q,r}(Q_{R,R^{sp}})} \Bigr ), \end{aligned} \end{aligned}$$

which is the desired result.

Case sp=n. Here we use the critical case of Sobolev–Morrey inequality, (2.8) with

$$\begin{aligned} \max { \left\{ \frac{p}{r^\prime }\left( 1-\frac{1}{r} - \frac{1}{q}\right) ^{-1}, q^\prime \right\} }< l < \infty . \end{aligned}$$
(3.28)

This applied for the second term in the left-hand side of (3.22) implies

$$\begin{aligned} \begin{aligned}&\Vert w\Vert _{L^{p,\infty }(Q_{R,R^{s p}})}^p \, + \left( C(n,s,p,l)R^{\frac{np}{l}}\right) ^{-1} \left[ \Vert w\Vert _{L^{l, p}(Q_{R,R^{s p}}))}^p - \delta ^{\beta } |B_R|^{\frac{p}{l}} |J| \right] \\&\le \Vert w\Vert _{L^{p,\infty }(Q_{R,R^{s p}}))}^p \, + \left( C(n,s,p,l)R^{\frac{np}{l}}\right) ^{-1} \int _J \Vert w - \delta ^{\frac{\beta }{p}}\Vert _{L^{l}(B_R)}^p \\&\le \Vert w\Vert _{L^{p,\infty }(Q_{R,R^{s p}})}^p + \int _J [w-\delta ^{\frac{\beta }{p}}]_{W^{s, p}({\mathbb {R}}^n)}^p \;\textrm{d}t \\&= \Vert w\Vert _{L^{p,\infty }(Q_{R,R^{s p}})}^p + \int _J [w]_{W^{s, p}({\mathbb {R}}^n)}^p \;\textrm{d}t \\&\le C(n,p) \beta ^{p-1} \Vert w\Vert _{L^{p q^\prime , p r^\prime }(Q_{R,R^{s p}})}^p \Bigl ( \frac{\Vert f\Vert _{L^{q,r}(Q_{R,R^{sp}})}}{\delta } + R^{-sp\nu } \Bigr ). \end{aligned} \end{aligned}$$

We replace the constant C(nspl) with \( \max {\lbrace 1, C(n,s,p,l) \rbrace }\), and multiply both sides with it to arrive at

$$\begin{aligned} \begin{aligned}&\Vert w\Vert _{L^{p,\infty }(Q_{R,R^{s p}})}^p + R^{\frac{-np}{l}}\Vert w\Vert _{L^{l, p}(Q_{R,R^{s p}})}^p \\&\le C(n,s,p,l) \beta ^{p-1}\Vert w\Vert _{L^{p q^\prime , p r^\prime }(Q_{R,R^{s p}})}^p \Bigl (\frac{\Vert f\Vert _{L^{q,r}(Q_{R,R^{sp}})}}{\delta } {+} R^{-sp\nu } \Bigr ) {+} C(n,p,l)\delta ^\beta |J| \\ \text {(using }\beta \ge 1)&\le C(n,s,p,l) \beta ^{p-1} \left( \Vert w\Vert _{L^{p q^\prime , p r^\prime }(Q_{R,R^{s p}})}^p \Bigl ( \frac{\Vert f\Vert _{L^{q,r}(Q_{R,R^{sp}})}}{\delta } + R^{-sp\nu } \Bigr ) + \delta ^\beta R^{sp} \right) \\ \text {using }(3.19)&\le C(n,s,p,l)\beta ^{p-1} \Vert w\Vert _{L^{p q^\prime , p r^\prime }(Q_{R,R^{s p}})}^p \left( \frac{\Vert f\Vert _{L^{q,r}(Q_{R,R^{sp}})}}{\delta } + R^{-sp\nu } + \frac{R^{sp}}{R^{\frac{n}{q^\prime } + \frac{sp}{r^\prime }}} \right) \\&\le C(n,s,p,l)\beta ^{p-1} \Vert w\Vert _{L^{p q^\prime , p r^\prime }(Q_{R,R^{s p}})}^p R^{-sp\nu }. \end{aligned} \end{aligned}$$
(3.29)

In the last line, we have used that since \(sp=n\) we have

$$\begin{aligned} sp- \frac{n}{q^\prime } - \frac{sp}{r^\prime } = sp - n\left( 1 -\frac{1}{q} \right) -sp\left( 1- \frac{1}{r} \right) = sp\left( \frac{1}{r} + \frac{1}{q} - 1\right) = -sp\nu . \end{aligned}$$

We have also used the following inequality which we have discussed in the case \(sp\ne n\):

$$\begin{aligned} \frac{\Vert f\Vert _{L^{q,r}(Q_{R,R^{sp}})}}{\delta } \le R^{-sp\nu }. \end{aligned}$$

Now we choose \(\vartheta = 2-\frac{1}{r}- \frac{1}{q} - \frac{p}{lr^\prime }\). Notice that due to the choice of l, (3.28), we have \(\vartheta > 1\). Then the exponents \((\vartheta r^\prime )^\prime \) and \((\vartheta q^\prime )^\prime \) satisfy

$$\begin{aligned} 1 - \frac{1}{(\vartheta r^\prime )^\prime } - \frac{1}{(\vartheta q^\prime )^\prime } = \frac{p}{l \vartheta r^\prime }. \end{aligned}$$

Therefore, we can apply Lemma 2.9 with the exponents \((\vartheta r^\prime )^\prime \) and \((\vartheta q^\prime )^\prime \) to (3.29) to arrive at

$$\begin{aligned} \begin{aligned} \Vert w^\vartheta \Vert _{L^{pq^\prime ,pr^\prime }(Q_{R,R^{sp}})}^{\frac{p}{\vartheta }}&= \Vert w\Vert _{L^{p \vartheta q^\prime , p\vartheta r^\prime }(Q_{R,R^{sp}})}^p \le R^{sp(1-\frac{1}{(\vartheta r^\prime )^\prime } - \frac{1}{(\vartheta q^\prime )^\prime })}\\&\quad \times \left( \Vert w\Vert _{L^{p,\infty }(Q_{R,R^{s p}})}^p + R^{\frac{-np}{l}}\Vert w\Vert _{L^{l, p}(Q_{R,R^{s p}})}^p \right) \\&= R^{\frac{np}{l \vartheta r^\prime }} \left( \Vert w\Vert _{L^{p,\infty }(Q_{R,R^{s p}})}^p + R^{\frac{-np}{l}}\Vert w\Vert _{L^{\infty , p}(Q_{R,R^{s p}})}^p \right) \\&\le C(n,s,p,l)\beta ^{p-1}R^{\frac{np}{l \vartheta r^\prime }-sp\nu } \Vert w\Vert _{L^{pq^\prime ,p r^\prime }(Q_{R,R^{sp}})}^{p}. \end{aligned}\nonumber \\ \end{aligned}$$
(3.30)

We apply (3.30) with the exponents

$$\begin{aligned} \beta _0 = 1,\qquad \beta _{m+1} = \vartheta \beta _m = \vartheta ^{m+1}. \end{aligned}$$

Let

$$\begin{aligned} \varphi _m:= \Vert ((u-v)^{+}_M +\delta )^\frac{\beta _m}{p} \Vert _{L^{p q^\prime , pr^\prime }(Q_{R,R^{sp}})}^\frac{p}{\beta _m} =\Vert (u-v)^{+}_M + \delta \Vert _{L^{\beta _m q^\prime , \beta _m r^\prime }(Q_{R,R^{sp}})}. \end{aligned}$$

Then (3.30) reads

$$\begin{aligned} \varphi _{m+1} \le \bigl (C \, R^{\frac{np}{l\vartheta r^\prime } - sp\nu } \bigr )^\frac{1}{\vartheta ^m}\theta ^\frac{(p-1)m}{\vartheta ^m} \varphi _m. \end{aligned}$$

By iterating the above inequality, we get

$$\begin{aligned} \varphi _{m+1} \le \bigl (C \, R^{\frac{np}{l \vartheta r^\prime } - sp\nu } \bigr )^{\sum _{j=0}^m \vartheta ^{-j}} \vartheta ^{(p-1) \sum _{j=0}^m j \vartheta ^{-j}} \varphi _0. \end{aligned}$$
(3.31)

Since \(\vartheta > 1\), we have the following convergent series

$$\begin{aligned} \sum _{j=0}^\infty \vartheta ^{-j} = \frac{\vartheta }{\vartheta -1} \end{aligned}$$

and

$$\begin{aligned} \sum _{j=0}^\infty j \vartheta ^{-j} = \frac{\vartheta }{(\vartheta -1)^2}. \end{aligned}$$

By (3.8) in Lemma 3.2, we obtain

$$\begin{aligned}{} & {} \varphi _0 = \Vert (u-v)^{+}_M\Vert _{L^{q^\prime , r^\prime }(Q_{R,R^{sp}})} \le C(n,s,p,q)R^{spp^\prime \nu + n } \Vert f\Vert _{L^{q,r}(Q_{R,R^{sp}})}^{\frac{1}{p-1}} + \delta R^{\frac{n}{q^\prime } + \frac{sp}{r^\prime }} \\{} & {} \quad \le C(n,s,p,q)R^{n+sp\nu } \delta . \end{aligned}$$

Inserting this into (3.30), and sending m to infinity, we get

$$\begin{aligned} \begin{aligned} \Vert (u-v)^{+}_M+\delta \Vert _{L^\infty (Q_{R,R^{sp}})}&\le C(n,s,p,q,l) \vartheta ^{\frac{(p-1)\vartheta }{(\vartheta - 1)^2}} R^{\frac{\vartheta }{\vartheta -1}\left( \frac{np}{l\vartheta r^\prime } - sp\nu \right) } R^{n+sp\nu } \delta \\&= C(n,s,p,q,l) \vartheta ^{\frac{(p-1)\vartheta }{(\vartheta - 1)^2}} R^{\frac{np}{(\vartheta -1)l r^\prime }-\frac{sp\nu }{\vartheta -1} +n}\delta \\&= C(n,s,p,q,l) \vartheta ^{\frac{(p-1)\vartheta }{(\vartheta - 1)^2}} R^{\frac{n}{(\vartheta -1)} (\vartheta -1 + \frac{p}{l r^\prime }) - \frac{sp\nu }{\vartheta -1} }\delta \\&= C(n,s,p,q,l) \vartheta ^{\frac{(p-1)\vartheta }{(\vartheta - 1)^2}} R^{\frac{n \nu }{(\vartheta -1)} - \frac{sp\nu }{\vartheta -1} }\delta \\&\le C(n,s,p,q,l) \vartheta ^{\frac{(p-1)\vartheta }{(\vartheta - 1)^2}} \bigl ( 1 + R^{sp\nu }\Vert f\Vert _{L^{q,r}(Q_{R,R^{sp}})} \bigr ). \end{aligned} \end{aligned}$$

Hence, we arrive at the desired estimate

$$\begin{aligned} \Vert (u-v)^{+}\Vert _{L^\infty (Q_{R,R^{sp}})} \le C(n,s,p,q,l) \vartheta ^{\frac{(p-1)\vartheta }{(\vartheta - 1)^2}} \bigl ( 1 + R^{sp\nu }\Vert f\Vert _{L^{q,r}(Q_{R,R^{sp}})} \bigr ). \end{aligned}$$

\(\square \)

Notice that \(-u\) is a solution to the same type of problem, and we can apply the above proposition to \(-u\). Since \(-v\) is the (sp)-caloric replacement of \(-u\), we get the same bound on \(\Vert (-u+v)^+\Vert _{L^\infty (Q_{R,R^{sp}})}\); as a result, we get a bound on the \(\Vert u-v\Vert _{L^\infty (Q_{R,R^{sp}})}\).

Corollary 3.5

Let u be a solution of \(\partial _t u + (-\Delta _p)^s u=f\) in \(Q_{\sigma ,\sigma ^{sp}}(x_0,T_0)\) with \(f \in L^{q,r}_{\textrm{loc}} (Q_{\sigma ,\sigma ^{sp}}(x_0,T_0))\) such that

$$\begin{aligned} \frac{1}{r} + \frac{n}{spq} < 1. \end{aligned}$$

In addition, assume that \(r\ge p^\prime \),

$$\begin{aligned} q \ge 1 \quad \text {if } sp\ne n, \quad \text {and}\quad q>1 \quad \text {if } sp=n. \end{aligned}$$

Let v be the (sp)-caloric replacement of u in \(Q_{R,R^{s p}}(x_0,T_0)\), with \(R< \sigma \).

If \(sp \ne n\), then

$$\begin{aligned} \Vert u-v\Vert _{L^\infty (Q_{R,R^{s p}}(x_0,T_0))} \le C(n,s,p) \vartheta ^{\frac{(p-1)\vartheta }{(\vartheta -1 )^2}} \bigl ( 1 + R^{sp\nu }\Vert f\Vert _{L^{q,r}(Q_{R,R^{sp}}(x_0,T_0))}\bigr ), \end{aligned}$$

where \(\nu = 1-\frac{1}{r} - \frac{n}{spq}\) and \( \vartheta = 1+\frac{sp\nu }{n}. \)

If \(sp=n\), then for any l such that \( \frac{p}{r^\prime }(1-\frac{1}{r} - \frac{1}{q})^{-1}<l < \infty \), we have

$$\begin{aligned} \Vert u-v\Vert _{L^\infty (Q_{R,R^{s p}}(x_0,T_0))}\le C(n,s,p,q,l) \vartheta ^{\frac{(p-1)\vartheta }{(\vartheta - 1)^2}} \bigl ( 1 + R^{sp\nu }\Vert f\Vert _{L^{q,r}(Q_{R,R^{sp}}(x_0,T_0))} \bigr ), \end{aligned}$$

where \(\vartheta = 2-\frac{1}{r}- \frac{1}{q} - \frac{p}{lr^\prime }\) and \(\nu = 1-\frac{1}{r} - \frac{1}{q}\).

Now we combine the local boundedness results for the equations with zero right-hand side (see [45] and also [22]) with Proposition 3.4 to prove local boundedness for the equation with nonzero right-hand side.

By [10, Theorem 1.1] with \(q=p\) and \(\sigma = \frac{1}{2}\), we have

where \(\beta =\frac{2\,s+3n-\frac{2n}{p}}{n+s} \) and C depends on ns and p. By Hölder’s inequality, we have

As \(p\ge 2\), we have \(\frac{2}{p^2}\le \frac{1}{p}\) and \(\frac{\beta }{(\beta -1)p^2}\le \frac{1}{p}\). Hence, we arrive at

(3.32)

Proof of Theorem 1.1

For u, a local weak solution of

$$\begin{aligned} \partial _t u + (-\Delta _p)^su = f(x,t), \qquad in \; Q_{2R,(2R)^{sp}}(x_0,T_0), \end{aligned}$$

we consider v to be the (sp)-caloric replacement in \(Q_{R,R^{sp}}(x_0,T_0)\),

$$\begin{aligned} {\left\{ \begin{array}{ll} v_t + (-\Delta _p)^s v=0 \quad &{} in \;\; Q_{R,R^{sp}}(x_0,T_0), \\ v=u \quad &{} in \;\; ({\mathbb {R}}^n \setminus B_R(x_0) ) \times [T_0-R^{sp},T_0], \\ v(x,T_0-R^{sp}) = u(x,T_0- R^{sp}) \quad &{} in \;\; B_R(x_0). \end{array}\right. } \end{aligned}$$

As mentioned in the proof of Proposition 3.5, our assumptions ensure that we can use Proposition 3.1 and v exists. Using (3.32), we arrive at

(3.33)

Using Lemma 2.3 in (3.33), we arrive at

(3.34)

where \(C=C(n,s,p)\). Finally, using Proposition 3.4 to estimate the term \(\Vert u-v\Vert _{L^\infty (Q_{R,R^{sp}})}\), in (3.34) we get the desired result. Here the estimate is written in the case \(sp\ne n\)

\(\square \)

Theorem 3.6

Let \(f \in L^{q,r}(Q_{R_1,R_1^{sp}}(z,T_1))\) with

$$\begin{aligned} \frac{1}{r} + \frac{n}{spq} < 1. \end{aligned}$$

In addition, assume that \(r\ge p^\prime \),

$$\begin{aligned} q \ge 1 \quad \text {if}\quad sp\ne n, \quad \text {and}\quad q>1 \quad \text {if} \quad sp=n. \end{aligned}$$

If u is a weak solution of the equation

$$\begin{aligned} \partial _t u + (-\Delta _p)^s u =f \qquad in \; Q_{R_1,R_1^{sp}}(z,T_1), \end{aligned}$$

such that

$$\begin{aligned} u\in & {} L^p(I; W^{s,p}(B_{R_1}(z)))\cap C(I;L^2(B_{R_1}(z)))\cap L^\infty (I;L^{p-1}_{sp}({\mathbb {R}}^n)) \\{} & {} \cap L^\infty (Q_{R_1,R_1^{sp}}(z,T_1)), \end{aligned}$$

then u is locally Hölder continuous in time and space. In particular, there exists a \(\zeta >0\), such that for \(\sigma <1\), \((x_1,t_1), \; (x_2,t_2) \in Q_{\sigma R_1,(\sigma R_1)^{sp}}(z,T_1) \), there holds

$$\begin{aligned} |u(x_1,t_1)-u(x_2,t_2)| \le C {\mathcal {M}}(|x_1-x_2|^\zeta + |t_1-t_2|^{\frac{\zeta }{sp}} ), \end{aligned}$$

with C depending on nsp and \(\sigma \), and

$$\begin{aligned} \begin{aligned} {\mathcal {M}}&:= \Biggl [ 1 + \Vert u\Vert _{L^\infty (Q_{ R_1,R_1^{sp}}(z,T_1))} + \vartheta ^{\frac{(p-1)\vartheta }{(\vartheta -1 )^2}} \bigl ( 1 + d^{sp\nu } \Vert f\Vert _{L^{q,r}(Q_{ R_1,R_1^{sp}}(z,T_1))} \bigr ) \\&\quad + \sup _{T_1-R_1^{sp}<t\le T_1} \textrm{Tail}_{p-1,sp}(u(\,\bullet \,,t);z,R_1) \Biggr ]^{p-1} \\&+\min \{1,d\}^{-1}\Vert u\Vert _{L^\infty (Q_{R_1,R_1^{sp}}(z,T_1))} + \Vert f\Vert _{L^{q,r}(Q_{ R_1,R_1^{sp}}(z,T_1))}^{\frac{1}{p-1}}. \end{aligned} \end{aligned}$$

Proof

Take a cylinder \(Q_{\sigma R_1,(\sigma R_1)^{sp}}(z,T_1) \subset Q_{R_1,R_1^{sp}}(z,T_1)\) and let \(d:=\min { \lbrace R_1(1-\sigma ),R_1(1-\sigma ^{sp})^{\frac{1}{sp}}\rbrace } > 0 \). For any point, \((x_0,T_0) \in Q_{\sigma R_1,(\sigma R_1)^{sp}}(z,T_1) \) consider the (sp)-caloric replacement of u in the cylinder \(Q_{R,R^{sp}}(x_0,T_0)\) with \(R \le \min { \lbrace 1, d \rbrace }\). The choice of d implies that \(Q_{R,R^{sp}}(x_0,T_0)\subset Q_{ R_1,R_1^{sp}}(z,t) \). First, we observe that:

(3.35)

For \(\rho \le \frac{R}{2}\), v is Hölder continuous in \(Q_{\rho ,\rho ^{sp}}(x_0,T_0)\) by Theorem 5.1, and by the mean value theorem, there is a point \(({\tilde{x}}_0,{\tilde{t}}_0)\in Q_{\rho ,\rho ^{sp}}\) such that \({\bar{v}}_{x_0,t_0}= v({\tilde{x}}_0,{\tilde{t}}_0)\). With the notation

$$\begin{aligned} {\mathcal {M}}:= 1+ \Vert v\Vert _{L^\infty (Q_{R,(R)^{sp}})} + \sup _{T_0-R^{sp} < t \le T_0}\textrm{Tail}_{p-1, sp}(v(\,\bullet \,,t); x_0,R), \end{aligned}$$

Theorem 5.1 implies:

$$\begin{aligned} \begin{aligned} |v(x,t)-{\bar{v}}_{(x_0,t_0),\rho }|&\le C\Bigl ( {\mathcal {M}}\Bigl (\frac{x-{\tilde{x}}_0}{R}\Bigr )^{\frac{\Theta }{2}} + {\mathcal {M}}^{p-1} \Bigl ( \frac{t-{\tilde{t}}_0}{R^{sp}}\Bigr )^{\frac{\Gamma }{2}} \Bigr )\\&\le C {\mathcal {M}}^{p-1} \Bigl ( \bigl (\frac{2\rho }{R} \bigr )^{\frac{\Theta }{2}} {+} \bigl ((\frac{\rho }{R} )^{sp}\bigr )^{\frac{\Gamma }{2}} \Bigr ),\quad \text {for } \; (x,t) {\in } Q_{\rho ,\rho ^{sp}}(x_0,T_0) \end{aligned} \end{aligned}$$

with \(C=C(n,s,p)\). Therefore,

(3.36)

where the constants C depends on ns and p, and we have defined \(\delta := \min { \bigl \lbrace \frac{\Theta }{2}, \frac{\Gamma }{2}\bigr \rbrace }\).

Moreover, by Lemma 3.2

(3.37)

where \(\xi \) is defined in Lemma 3.2. Notice that \(\xi >0\) by our assumptions on q and r. Inserting (3.37) and (3.36) in (3.35), we arrive at

Using Corollary 3.5, we get:

with \(\vartheta \) and \(\nu \) defined in Corollary 3.5; here, the estimate is only written in the case \(sp \ne n\) for simplicity. Since \(Q_{R,R^{sp}}(x_0,T_0)\subset Q_{ R_1,R_1^{sp}}(z,T_1)\), the above expression is less than

$$\begin{aligned} \begin{aligned}&\le C(n,s,p)\left( \frac{R}{\rho }\right) ^{n+sp} R^\xi \Vert f\Vert _{L^{q,r}(Q_{ R_1,R_1^{sp}}(z,T_1))}^{p^\prime } \\&\quad + C(n,s,p)\left( \frac{\rho }{R}\right) ^{\delta p} \Bigl [ 1+ \Vert u\Vert _{L^\infty (Q_{ R_1,R_1^{sp}}(z,T_1))}^{p} \\&\quad + \sup _{T_0-R^{sp}<t\le T_0} \textrm{Tail}_{p-1,sp}(u(\,\bullet \,,t);x_0,R)^{p} \\&\quad +\Bigl ( \vartheta ^{\frac{(p-1)\vartheta }{(\vartheta -1 )^2}} \bigl ( 1 + d^{sp\nu }\Vert f\Vert _{L^{q,r}(Q_{ R_1,R_1^{sp}}(z,T_1))} \bigr ) \Bigr )^p \Bigr ]^{p-1}. \end{aligned} \end{aligned}$$

Concerning the tail term, since \(B_{R}(x_0) \subset B_{R_1}(z)\), using Lemma 2.4 we have

$$\begin{aligned} \begin{aligned}&\textrm{Tail}_{p-1,sp} (u(\,\bullet \,,t);x_0,R)^{p-1} \\&\quad {\le } \Bigl ( \frac{R}{R_1} \Bigr )^{sp}\!\Bigl ( \frac{R_1}{R_1-|x_0-z|} \Bigr )^{n +s\,p} \textrm{Tail}_{p-1,sp}(u(\,\bullet \,,t);z,R_1)^{p-1} {+} \Vert u(\,\bullet \,,t)\Vert _{L^{\infty }(B_{R_1}(z))}^{p-1}, \end{aligned} \end{aligned}$$
(3.38)

and by the choice of the radii, we have

$$\begin{aligned} \frac{R}{R_1}< \frac{d}{R_1} <1-\sigma \quad and \quad \frac{R_1}{R_1- |x_0-z|} \le \frac{R_1}{R_1-\sigma R_1} \le \frac{1}{1-\sigma }. \end{aligned}$$

Hence, taking the supremum in time and using Minkowski’s inequality in (3.38), we arrive at

$$\begin{aligned} \begin{aligned}&\sup _{T_0-R^{sp}<t\le T_0} \textrm{Tail}_{p-1,sp}(u;x_0,R)^{p} \\&\le C\frac{1}{(1-\sigma )^n} \Bigl ( \sup _{T_0-R^{sp}<t\le T_0} \textrm{Tail}_{p-1,sp}(u(\,\bullet \,,t);z,R_1)^{p} + \Vert u\Vert _{L^\infty ([T_0-R^{sp},T_0]\times B_{R_1}(z)}^p \Bigr ) \\&\le C\frac{1}{(1-\sigma )^n} \Bigl ( \Vert u\Vert _{L^\infty (Q_{R_1,R_1^{sp}}(z,T_1))}^p +\sup _{T_1-R_1^{sp}<t\le T_1} \textrm{Tail}_{p-1,sp}(u(\,\bullet \,,t);z,R_1)^{p} \Bigr ), \end{aligned} \end{aligned}$$

where the above constant C depends on \(n, s \text {and }p\). In conclusion,

Now we make the choice \(\rho =\frac{R^\theta }{2}\) with

$$\begin{aligned} \theta := 1+ \frac{\xi }{\delta p + n + sp}. \end{aligned}$$

This yields

for any \(0< \rho < \frac{\min {\lbrace 1,d \rbrace }^\theta }{2}\), where

$$\begin{aligned} \zeta = \frac{\xi \delta }{n +sp + \delta p + \xi }. \end{aligned}$$

For values of \(\rho \ge \frac{\min {\lbrace 1,d \rbrace }^\theta }{2}\),

We can then conclude that for any cylinder of arbitrary size we have

with C depending on

$$\begin{aligned}{} & {} n,\;s,\; p,\; R_1,\; \sigma ,\; \sup _{T_1-R_1^{sp}<t\le T_1} \textrm{Tail}_{p-1,sp}(u(\,\bullet \,,t);z,R_1),\; \Vert f\Vert _{L^{q,r}(Q_{ R_1,R_1^{sp}}(z,T_1))} \text {,}\\{} & {} \text {and }\Vert u\Vert _{L^\infty (Q_{R_1,R_1^{s p}}(z,T_1))}. \end{aligned}$$

In particular, one can obtain

(3.39)

Now we use the characterization of the Campanato spaces in \({\mathbb {R}}^{n+1}\) with a general metric in [30], see also [20]. Our setting does not fit directly in the context considered there, since we only work with cylinders that are one-sided in the time direction that is \((t-r^{sp},t]\times B_r(x)\) instead of \((t-r^{sp},t+r^{sp})\times B_r(x)\). Still, if you follow the proof in [30] with small modifications, you can also conclude the result in this setting.

In the case of \(sp \ge 1\), using [30, Theorem 3.2] we get the Hölder continuity of u with exponent \(\zeta \) in \(Q_{\sigma R, (\sigma R)^{sp}}\) with respect to the metric

$$\begin{aligned} d((x,\tau _1),(y,\tau _2))= \max {\lbrace |x-y|, |\tau _2 - \tau _1|^{\frac{1}{sp}}\rbrace }, \end{aligned}$$

for which the balls of radius r are of the form \((t-r^{sp},t+r^{sp})\times B_r(x)\), which means

$$\begin{aligned} |u(x_1,t_1)-u(x_2,t_2)|\le & {} C {\mathcal {M}}\bigl ( |x_1-x_2| + |t_1 - t_2|^{\frac{1}{sp}} \bigr )^\zeta \\\le & {} C {\mathcal {M}}\bigl ( |x_1-x_2|^\zeta + |t_1-t_2|^{\frac{\zeta }{sp}} \Bigr ). \end{aligned}$$

Here C depends on nsp and \(\sigma \). In the case of \(sp< 1\), we use the metric

$$\begin{aligned} d((x,\tau _1),(y,\tau _2))= \max {\lbrace |x-y|^{sp}, |\tau _2 - \tau _1| \rbrace }. \end{aligned}$$

The balls of radius r are of the form \((t-r,t+r)\times B_{r^{\frac{1}{sp}}}(x)\). Hence, we have a decay of order \(r^{\frac{\xi }{sp} p}\) of the average of u on the half balls. [30, Theorem 3.2] implies the following Hölder continuity on \(Q_{\sigma R_1,(\sigma R_1)^{sp}}\)

$$\begin{aligned} |u(x_1,t_1)-u(x_2,t_2)| {\le } C \bigl ( |x_1-x_2|^{sp} {+} |t_1 {-} t_2| \bigr )^{\frac{\zeta }{sp}} {\le } C \bigl ( (|x_1-x_2|^{\zeta } + |t_1-t_2|^{\frac{\zeta }{sp}} \Bigr ). \end{aligned}$$

\(\square \)

Lemma 3.7

(Stability in \(L^\infty \)) Let \(f \in L^{q,r}_{\textrm{loc}}(Q_{2R,(2R)^{sp}})\) with

$$\begin{aligned} \frac{1}{r} + \frac{n}{spq} < 1. \end{aligned}$$

In addition, assume that \(r\ge p^\prime \),

$$\begin{aligned} q \ge 1 \quad \text {if}\quad sp\ne n, \quad \text {and}\quad q>1 \quad \text {if} \quad sp=n. \end{aligned}$$

Let u be a local weak solution to the equation

$$\begin{aligned} u_t +(-\Delta _p)^s u = f \quad in \; Q_{2R,(2R)^{sp}}, \end{aligned}$$

with

$$\begin{aligned} \Vert u\Vert _{L^\infty (Q_{R,R^{sp}})} + \sup _{-R^{sp} < t \le 0} \textrm{Tail}_{p-1,sp}(u(\,\bullet \,,t);0,R) \le M, \end{aligned}$$

and

$$\begin{aligned} \Vert f\Vert _{L^{q,r}(Q_{R,R^{sp}})}\le \omega . \end{aligned}$$

Consider the (sp)-caloric replacement

$$\begin{aligned} {\left\{ \begin{array}{ll} \varphi _t + (-\Delta _p)^s \varphi =0 \quad &{} in \;\; Q_{R,R^{s \, p}} \\ \varphi =u \quad &{} in \;\; ({\mathbb {R}}^n \setminus B_R) \times [-R^{s\,p},0] \\ \varphi (x,-R^{s\,p}) = u(x,-R^{sp}) \quad &{} in \;\; B_R. \end{array}\right. } \end{aligned}$$

Then for \(\sigma < 1\), there is a \(\delta _{M,R,\sigma }(\omega )\) such that

$$\begin{aligned} \Vert u - \varphi \Vert _{L^\infty (Q_{\sigma R,(\sigma R)^{s\,p}})} < \delta _{M,R,\sigma }(\omega ), \end{aligned}$$

and \(\delta _{M,R, \sigma }(\omega )\) converges to 0 as \(\omega \) goes to 0.

Proof

The existence of such a bound follows immediately from Corollary 3.5.

To show the convergence of \(\sigma _{M,R,\sigma }\) to zero, we argue by contradiction, suppose that there is a sequence \(f_n \in L^{q,r}(Q_{R,R^{s \, p}})\) and \(u_n\) such that

$$\begin{aligned}{} & {} \Vert u_n\Vert _{L^\infty (Q_{R,R^{s \, p}})} + \sup _{T_0 - R^{s \, p} \le t \le T_0} \textrm{Tail}_{p-1,sp}(u_n(\,\bullet \,,t);0,R) \le M\\{} & {} \text {and} \quad \Vert f_n\Vert _{L^{q,r}(Q_{R,R^{s \, p}})} \rightarrow 0, \end{aligned}$$

but

$$\begin{aligned} \Vert u_n - \varphi _n\Vert _{L^\infty (Q_{\sigma R, (\sigma R)^{s \, p}})}> \varepsilon >0. \end{aligned}$$
(3.40)

Using (3.12) from Lemma 3.2, we have

$$\begin{aligned} \lim _{n \rightarrow \infty } \int _{- R^{sp}}^{0} [u_n-\varphi _n]_{W^{s,p}({\mathbb {R}}^n)}^p \;\textrm{d}t \le C(n,s,p,q,r,R) \lim _{n \rightarrow \infty } \Vert f_n\Vert _{L^{q,r}(Q_{R,R^{s \, p}})}^{p^\prime } =0.\nonumber \\ \end{aligned}$$
(3.41)

By assumption, \(u_n\) is uniformly bounded in \(L^\infty (Q_{ R, R^{sp}})\). Now we show that \(\varphi _n\) is also uniformly bounded in \(L^\infty (Q_{R, R^{sp}})\).

$$\begin{aligned} \begin{aligned} \Vert \varphi _n\Vert _{L^\infty (Q_{ R, R^{s\,p}})}&\le \Vert u_n\Vert _{L^\infty (Q_{ R,(\sigma R)^{s\,p}})} + \Vert u_n - \varphi _n\Vert _{L^\infty (Q_{R, R^{s\,p}})} \\&\le M+ \Vert u_n-\varphi _n\Vert _{L^\infty (Q_{R,R^{s \, p}})}. \end{aligned} \end{aligned}$$
(3.42)

By Corollary 3.5,

$$\begin{aligned} \begin{aligned} \Vert u_n-\varphi _n\Vert _{L^\infty (Q_{R,R^{s \, p}})} \le C(n,s,p) \vartheta ^{\frac{(p-1)\vartheta }{(\vartheta -1 )^2}} \bigl ( 1 + R^{sp\nu }\Vert f_n\Vert _{L^{q,r}(Q_{R,R^{sp}})} \bigr ). \end{aligned} \end{aligned}$$
(3.43)

Since \(\Vert f_n\Vert _{L^{q,r}(Q_{R,R^{s \, p}})}^{p^\prime }\) is uniformly bounded, (3.43) and (3.42) give us a uniform bound on \( \Vert \varphi _n\Vert _{L^\infty (Q_{ R, R^{s\,p}})} \).

Now we are in a position to use Theorem 3.6 for both of the sequences \(u_n\) and \(\varphi _n\), which gives us a uniform bound on the Hölder seminorms of \(u_n\) and \(\varphi _n\) in \(Q_{\sigma R,(\sigma R)^{s p}}\). Therefore, by Arzela–Ascoli’s theorem \(u_n - \varphi _n\) has a uniformly convergent subsequence in \(Q_{\sigma R,(\sigma R)^{sp}}\). By (3.41), the limit is 0, contradicting (3.40). \(\square \)

4 Improved Hölder regularity for nonhomogeneous equation

Proposition 4.1

Let \(f\in L^{q,r}(Q_{1,2})\) with qr satisfying

$$\begin{aligned} \frac{1}{r} + \frac{n}{spq} < 1. \end{aligned}$$

In addition, assume that \(r\ge p^\prime \),

$$\begin{aligned} q \ge 1 \quad \text {if}\quad sp\ne n, \quad \text {and}\quad q>1 \quad \text {if} \quad sp=n. \end{aligned}$$

Let u be a weak solution of \(u_t + (-\Delta _p)^s u=f\) in \(Q_{1,2}\) that satisfies

$$\begin{aligned} \Vert u\Vert _{L^\infty (Q_{1,2})} \le 1 \;, \quad \sup _{-2\le t \le 0}\textrm{Tail}_{p-1,sp}(u;0,1) \le 1. \end{aligned}$$

Then there exists \(\omega \) such that if

$$\begin{aligned} \Vert f\Vert _{L^{q,r}(Q_{1,2})} \le \omega (n,s,p,q,r,\alpha ), \end{aligned}$$

u is locally Hölder continuous in \(Q_{\frac{1}{2},\frac{1}{2^{sp}}}\) with exponents \(\alpha \) in space and \(\frac{\alpha }{sp-(p-2)\alpha }\) in time, as long as

$$\begin{aligned} \alpha \le \frac{r(spq-n)-spq}{q(r(p-1)-(p-2))} \quad \text {and} \quad \alpha < \Theta . \end{aligned}$$
(4.1)

Recall that \(\Theta = \min { \left\{ \frac{sp}{p-1},1 \right\} }\).

More precisely, for \((x_1,t_1),\; (x_2,t_2) \in Q_{\frac{1}{2},\frac{1}{2^{sp}}}\) we have

$$\begin{aligned} |u(x_2,t_2) - u(x_1,t_1)|\le C(n,s,p,q,r,\alpha ) \left( |x_2 - x_1|^\alpha + |t_2 - t_1|^{\frac{\alpha }{sp-(p-2)\alpha }} \right) . \end{aligned}$$

Proof

Step 1: Decay at the origin.

For this part, we prove a decay at the origin for u under the assumptions

$$\begin{aligned} \Vert u\Vert _{L^\infty (Q_{1,1})} \le 1 \;, \quad \sup _{-1\le t \le 0}\textrm{Tail}_{p-1,sp}(u;0,1) \le 1, \quad \text {and} \;\; \Vert f\Vert _{L^{q,r}(Q_{1,1})} \le \omega . \end{aligned}$$
(4.2)

Here \(\omega >0\) is a small number to be determined later which depends on nsp and \(\alpha \). We introduce the parabolic cylinder

$$\begin{aligned} G_r:=B_r(0)\times (-r^\beta ,0], \end{aligned}$$

with \(\beta =sp-(p-2)\alpha \). We show that for any exponent \(\alpha \) satisfying (4.1), the following holds for \(r< 1\)

$$\begin{aligned} \Vert u(x,t)-u(0,0)\Vert _{L^\infty (G_{r})} \le C r^{\alpha }. \end{aligned}$$

It is enough to prove the inequality for a sequence of radii \( (r_k)_{k=0}^\infty ,\; r_k= \lambda ^k \), for some \(\lambda < 1\). Without loss of generality, we assume \(u(0,0)=0\). Consider the rescaled functions

$$\begin{aligned} v_k(x,t):= \frac{u(\lambda ^k x, \lambda ^{k\beta }t)}{\lambda ^{\alpha k}}, \end{aligned}$$

with \(\lambda \) small enough to be determined later. We will prove the following by induction,

$$\begin{aligned} \Vert v_k(x,t)\Vert _{L^\infty (G_1)} \le 1 \qquad \text {and } \qquad \sup _{-1\le t \le 0} \int _{{\mathbb {R}}^n \setminus B_1} \frac{|v_k(x,t)|}{|x|^{n+ s\, p}} \;\textrm{d}x\le 1. \end{aligned}$$
(4.3)

For \(k=0\), (4.3) follows from our assumptions (4.2).

Observe that

$$\begin{aligned} {\left\{ \begin{array}{ll} \frac{\partial v_k(x,t)}{\partial t} =\lambda ^{\beta k - \alpha k} u_t(\lambda ^k x, \lambda ^{\beta k}t) &{} \\ (-\Delta _p)^s v_k(x,t) = \lambda ^{k[sp-(p-1)\alpha ]}(-\Delta _p)^s u(\lambda ^k x, \lambda ^{\beta k}t). &{} \\ \end{array}\right. } \end{aligned}$$

With \(\beta = sp - (p-2)\alpha \), \(v_k(x,t)\) solves

$$\begin{aligned} \frac{\partial v_k}{\partial t} + (-\Delta _p)^s v_k= \lambda ^{k[sp- (p-1)\alpha ]}f(\lambda ^k x, \lambda ^{\beta k}t) =: f_k(x,t) \qquad \text {in} \; Q_{\frac{1}{\lambda ^k}, \frac{1}{\lambda ^{\beta k}}}. \end{aligned}$$

Moreover,

$$\begin{aligned} \begin{aligned} \Vert f_{k}\Vert _{L^{q,r}(G_1)}^r&= \int _{-1}^0 \Bigl ( \int _{B_1}|f_k(x,r)|^q \;\textrm{d}x \Bigr )^{\frac{r}{q}} \;\textrm{d}t \\&= \int _{-1}^0 \Bigl ( \int _{B_{\lambda ^k}} \lambda ^{kq[sp - (p-1) \alpha ]-kn}|f(x,\lambda ^{\beta k}t)|^q \;\textrm{d}x \Bigr )^\frac{r}{q} \;\textrm{d}t \\&=\int _{-1}^0 \lambda ^{rk[sp-(p-1)\alpha ] -\frac{krn}{q}} \Bigl ( \int _{B_{\lambda ^k}} |f(x,\lambda ^{\beta k} t)|^q \;\textrm{d}x \Bigr )^\frac{r}{q} \;\textrm{d}t \\&= \lambda ^{rk[sp-(p-1)\alpha ] -\frac{krn}{q} - \beta k} \Vert f\Vert _{L^{q,r}(G_{\lambda ^k})}. \end{aligned} \end{aligned}$$

Since \(\lambda <1\), and the exponent of \(\lambda \) is nonnegative by (4.1), we get \(\Vert f_{k}\Vert _{L^{q,r}(G_1)} \le \omega \).

Assume that (4.3) holds for k. Now we prove that it holds for \(k+1\). Consider the (sp)-caloric replacement of \(v_k(x,t)\) in \(Q_{1,1}\), say \(\varphi _k(x,t)\). Then

$$\begin{aligned} |v_k(x,t)| \le |v_k(x,t) - \varphi _k(x,t)| + |\varphi _k(x,t) -\varphi _k(0,t)| + |\varphi _k(0,t)-v_k(0,t)|. \end{aligned}$$

By Theorem 5.1, \(\varphi _k\) is locally Hölder continuous in \(Q_{1,1}\), and for \((x,t) \in Q_{\frac{1}{2}, \frac{1}{2^{sp}}}\),

$$\begin{aligned} |\varphi _k(x,t) - \varphi _k(0,0)| \le C_1 |x|^{\Theta - \varepsilon } + C_2 |t|^{\Gamma - \frac{\varepsilon }{\beta }}. \end{aligned}$$

Here we take \(\varepsilon = \frac{\Theta -\alpha }{2}\). Since \(\Vert f_k\Vert _{L^{q,r}(Q_{1,1})} \le \omega \), Lemma 3.7 implies

$$\begin{aligned} |v_k(x,t)| \le 2\delta (\omega ) + C_1 |x|^{\Theta - \varepsilon } + C_2 |t|^{\Gamma - \frac{\varepsilon }{\beta }}, \qquad \text {in} \quad Q_{\frac{1}{4}, \frac{1}{4^{sp}}}. \end{aligned}$$
(4.4)

In Theorem 5.1, the Hölder constants are bounded by

$$\begin{aligned} \begin{aligned} (C_2)^{\frac{1}{p-1}} \le C_1&\le C\bigl (1+ \Vert \varphi _k\Vert _{L^\infty (Q_{1,1})} + \sup _{-\frac{1}{2^{sp}}\le t \le 0} \textrm{Tail}_{p-1,sp}(\varphi _k;0,1)\bigr ) \\&\le C \bigl ( 1+ \Vert \varphi _k\Vert _{L^\infty (Q_{1,1})} + \sup _{-1 \le t \le 0} \textrm{Tail}_{p-1,sp}(v_k;0,1) \bigr )\\&\le C \bigl ( 1+ \Vert v_k - \varphi _k\Vert _{L^\infty (Q_{1,1})} + \Vert v_k\Vert _{L^\infty (Q_{1,1})} \\&\quad + \sup _{-1 \le t \le 0} \textrm{Tail}_{p-1,sp}(v_k;0,1) \bigr ). \end{aligned} \end{aligned}$$

Therefore, by (4.3) we have

$$\begin{aligned} (C_2)^{\frac{1}{p-1}} \le C_1 \le C(n,s,p,\alpha )( 3 + \Vert v_k - \varphi _k\Vert _{L^\infty (Q_{1,1})}). \end{aligned}$$

By Corollary 3.5,

$$\begin{aligned} C_1 \le C\left( 3 + C(n,s,p,q,r) (1 + \Vert f_k\Vert _{L^{q,r}(Q_{1,1})}) \right) \le C( 3 + C(n,s,p,q,r)(1+ \omega )). \end{aligned}$$

This is a bound independent of k. We can take \(\omega \) to be less than 1 and take \(C_1=C(n,s,p)( 3 + 2C(n,s,p,q,r),\) with the C(nspqr) coming from Corollary 3.5, so that the constants \(C_1,\; C_2\) are independent of \(\omega \) as well.

Now we proceed and prove (4.3) for \(k+1\). First, we state our choice of \(\lambda \)

$$\begin{aligned} \lambda\le & {} \min \Bigl \lbrace \frac{1}{4}, \frac{1}{4^{\frac{sp}{\beta }}}, \nonumber \\{} & {} \times \frac{1}{(2 C_1+ 2C_2)^{\frac{2}{\Theta -\alpha }}}, \Bigl (1+ \frac{\omega _n(4^{sp} - 1)}{sp}+ \frac{(1+C_1 +C_2)^{p-1}}{(p-1)(\Theta -\alpha )/2} \Bigr )^{\frac{2}{(p-1)(\Theta - \alpha )}} \Bigr \rbrace .\nonumber \\ \end{aligned}$$
(4.5)

Since \(\lambda < \frac{1}{4}\), and \(\lambda ^\beta < \frac{1}{4^{s\, p}}\), \(Q_{\lambda , \lambda ^{\beta }} \subset Q_{\frac{1}{4},\frac{1}{4^{s \, p}}}\). Therefore, from (4.4) we obtain

$$\begin{aligned} \Vert v_k(x,t)\Vert _{L^\infty (G_ \lambda )} \le \delta (\omega ) + C_1 \lambda ^{\Theta - \varepsilon } + C_2 \lambda ^{\beta (\Gamma - \frac{\varepsilon }{\beta })}. \end{aligned}$$

Notice that \(\beta \Gamma \ge \Theta \), by the above choice of \(\beta \). Thus,

$$\begin{aligned} \Vert v_k(x,t)\Vert _{L^\infty (G_\lambda )} \le \delta (\omega ) + (C_1 + C_2) \lambda ^{\Theta - \varepsilon }. \end{aligned}$$
(4.6)

Recall that \(\varepsilon = \frac{\Theta -\alpha }{2}\) and by the assumption (4.5)

$$\begin{aligned} (C_1 +C_2) \lambda ^{\Theta - \varepsilon } < \frac{1}{2} \lambda ^\alpha . \end{aligned}$$

Now we choose \(\omega \) so that

$$\begin{aligned} 2\delta (\omega ) \le \frac{1}{2} \lambda ^{\Theta } \le \frac{1}{2} \lambda ^\alpha . \end{aligned}$$

This is possible since \(\delta (\omega )\) converges to zero as \(\omega \rightarrow 0\). Then, (4.6) implies

$$\begin{aligned} \Vert v_k(x,t)\Vert _{L^\infty (G_\lambda )} \le \lambda ^\alpha , \end{aligned}$$

which translates to

$$\begin{aligned} \Vert v_{k+1}(x,t)\Vert _{L^\infty (G_1)} = \Bigl \Vert \frac{v_k(\lambda x, \lambda ^\beta t)}{\lambda ^\alpha }\Bigr \Vert _{L^\infty (G_1)} \le 1, \end{aligned}$$
(4.7)

which is the first part of (4.3). For the second part, we want to show

$$\begin{aligned} \sup _{-1<t<0} \int _{{\mathbb {R}}^n \setminus B_1} \frac{|v_{k+1}(x,t)|^{p-1}}{|x|^{n+sp}} \;\textrm{d}x \le 1. \end{aligned}$$

We split the integral into three parts. Using the induction hypothesis,

$$\begin{aligned} \begin{aligned} \sup _{-1<t<0} \int _{{\mathbb {R}}^n \setminus B_{\frac{1}{\lambda }}} \frac{|v_{k+1}(x,t)|^{p-1}}{|x|^{n+sp}} \;\textrm{d}x&\le \sup _{-\lambda ^{-\beta } \le t\le 0} \int _{{\mathbb {R}}^n \setminus B_{\frac{1}{\lambda }}} \frac{|v_{k+1}(x,t)|^{p-1}}{|x|^{n+sp}} \;\textrm{d}x \\&= \lambda ^{sp -\alpha (p-1)} \sup _{-1<t<0} \int _{{\mathbb {R}}^n \setminus B_1} \frac{|v_{k}(x,t)|^{p-1}}{|x|^{n+sp}} \;\textrm{d}x \\ \left( \text {using } \Theta \le \frac{sp}{p-1}\right) \quad&\le \lambda ^{(p-1)(\Theta -\alpha )}. \end{aligned} \end{aligned}$$

Moreover, \(\Vert v_k\Vert _{L^\infty (G_1)} \le 1\), and hence,

$$\begin{aligned} \begin{aligned} \sup _{-1<t<0} \int _{B_{\frac{1}{\lambda }} \setminus B_{\frac{1}{4\lambda }}} \frac{|v_{k+1}(x,t)|^{p-1}}{|x|^{n+sp}} \;\textrm{d}x&\le \lambda ^{sp-\alpha (p-1)} \sup _{-\lambda ^\beta< t<0} \int _{B_{1} \setminus B_{\frac{1}{4}}} \frac{|v_{k}(x,t)|^{p-1}}{|x|^{n+sp}} \;\textrm{d}x \\&\le \lambda ^{sp-\alpha (p-1)} \int _{B_{1} \setminus B_{\frac{1}{4}}} \frac{1}{|x|^{n+sp}}\;\textrm{d}x \\&\le \lambda ^{(p-1)(\Theta - \alpha )} \frac{\omega _n(4^{sp} - 1)}{sp}:= C_3 \lambda ^{2(p-1)\varepsilon }. \end{aligned} \end{aligned}$$

For remaining part, we transfer the estimate (4.4) to \(v_{k+1}\) and obtain

$$\begin{aligned} |v_{k+1}(x,t)| \le \delta (\omega ) \lambda ^{-\alpha } + C_1 \lambda ^{\Theta -\varepsilon -\alpha } |x|^{\Theta -\varepsilon } + C_2 \lambda ^{\beta \Gamma - \varepsilon -\alpha } |t|^{\Gamma -\frac{\varepsilon }{\beta }} \qquad \text {in} \quad Q_{\frac{1}{4\lambda }, \frac{1}{4^{sp}\lambda ^\beta }}. \end{aligned}$$

In particular, since \(\lambda ^\beta \le \frac{1}{4^{sp}}\), \(Q_{\frac{1}{4\lambda },1} \subset Q_{\frac{1}{4\lambda }, \frac{1}{4^{sp}\lambda ^\beta }}\), and \(\delta (\omega ) \le \lambda ^{\Theta } \le \lambda ^{\Theta -\varepsilon }\) we get

$$\begin{aligned} \sup _{-1\le t\le 0} |v(x,t)| \le \lambda ^{\Theta - \varepsilon - \alpha } (1+ C_2\lambda ^{\beta \Gamma - \Theta } + C_1 |x|^{\Theta -\varepsilon }). \end{aligned}$$

Therefore,

$$\begin{aligned} \begin{aligned}&\sup _{-1<t<0} \int _{B_{\frac{1}{4\lambda }} \setminus B_1} \frac{|v_{k+1}(x,t)|^{p-1}}{|x|^{n+sp}} \;\textrm{d}x \\&\le \lambda ^{(p-1)(\Theta - \varepsilon - \alpha )} \int _{B_{\frac{1}{4\lambda }} \setminus B_1} \frac{|1+ C_2\lambda ^{\beta \Gamma - \Theta } + C_1 |x|^{\Theta -\varepsilon }|^{p-1}}{|x|^{n+sp}} \;\textrm{d}x \\&(\text {using}\;\; |x|\ge 1) \le (1+C_2\lambda ^{\beta \Gamma - \Theta } + C_1 )^{p-1} \lambda ^{(p-1)(\Theta - \varepsilon - \alpha )} \\&\quad \times \int _ {B_{\frac{1}{4\lambda }} \setminus B_1} \frac{1}{|x|^{n+sp-(p-1)(\Theta -\varepsilon )}} \;\textrm{d}x \\&(\text {using} \;\; sp\ge (p-1)\Theta ) \le (1+ C_2 + C_1)^{p-1} \lambda ^{(p-1)(\Theta - \varepsilon - \alpha )} \int _{{\mathbb {R}}^n \setminus B_1} \frac{1}{|x|^{n+\varepsilon (p-1)}} \;\textrm{d}x \\&\le \frac{(1+C_1+C_2)^{p-1}}{\varepsilon (p-1)} \lambda ^{(p-1)(\Theta - \varepsilon - \alpha )}:= C_4 \lambda ^{(p-1)\varepsilon }. \end{aligned} \end{aligned}$$

Hence,

$$\begin{aligned}{} & {} \sup _{-1<t<0} \int _{{\mathbb {R}}^n \setminus B_1} \frac{|v_{k+1}(x,t)|^{p-1}}{|x|^{n+sp}} \;\textrm{d}x \le \lambda ^{2(p-1)\varepsilon } + C_3\lambda ^{2(p-1)\varepsilon } + C_4 \lambda ^{(p-1)\varepsilon } \\{} & {} \quad \le \lambda ^{(p-1)\varepsilon }\left( 1+C_3 + \frac{(1+C_1+C_2)^{p-1}}{\varepsilon (p-1)}\right) . \end{aligned}$$

Using the assumption (4.5) on \(\lambda \), we obtain

$$\begin{aligned} \sup _{-1<t<0} \int _{{\mathbb {R}}^n \setminus B_1} \frac{|v_{k+1}(x,t)|^{p-1}}{|x|^{n+sp}} \;\textrm{d}x \le 1. \end{aligned}$$

Step 2: Regularity in a cylinder. We choose \(\alpha \) as in (4.1) and let \(\omega \) be as in Step 1. For a point \((x_0,t_0) \in Q_{\frac{1}{2}, \frac{1}{2^{sp}}}\), define

$$\begin{aligned} {\tilde{u}}(x,t) = \frac{1}{L} u(\frac{x}{2} + x_0, L^{2-p} \frac{1}{2^{sp}}t + t_0 ), \end{aligned}$$

where \(L = 2^\frac{n}{p-1}(1+|B_1|)^{\frac{1}{p-1}}\). Then \({\tilde{u}}\) is a solution of

$$\begin{aligned} \partial _t {\tilde{u}} + (- \Delta _p)^s {\tilde{u}} = \frac{L^{-(p-1)}}{2^{sp}} f \Bigl (\frac{x}{2} +x_0, L^{2-p} \frac{1}{2^{sp}}t +t_0 \Bigr ):= {\tilde{f}} \quad \text {in} \; Q_{1,2^{sp-1}L^{p-2}}. \end{aligned}$$

By the choice of L, \({\tilde{u}}\) satisfies the conditions (4.2) in Step 1. Since \(L\ge 1\), we immediately have

$$\begin{aligned} \Vert {\tilde{u}}\Vert _{L^\infty (Q_{1,1}(0,0))} \le \frac{1}{L}\Vert u\Vert _{ L^\infty (Q_{\frac{1}{2}, \frac{L^{2-p}}{2^{sp}}}(x_0,t_0))} \le \Vert u\Vert _{L^\infty (Q_{1,2})} \le 1, \end{aligned}$$

since \(Q_{\frac{1}{2},\frac{L^{2-p}}{2^{sp}}}(x_0,t_0) \subset Q_{1,2}\). As for the \(L^{q,r}\) norm of \({\tilde{f}}\), we have

$$\begin{aligned} \begin{aligned} \qquad \Vert {\tilde{f}}\Vert _{L^{q,r}(Q_{1,1})}&= \frac{L^{-(p-1)}}{2^{sp}} \bigl ( 2^{\frac{n}{q}+\frac{sp}{r}} L^\frac{p-2}{r} \Vert f\Vert _{L^{q,r}(Q_{\frac{1}{2},\frac{L^{2-p}}{2^{sp}}}(x_0,t_0))}\bigr ) \\&\le L^{-(p-1)} (1/2)^{sp(1-\frac{1}{r} - \frac{n}{spq})} \Vert f\Vert _{Q_{1,2}} \\&\le L^{-(p-1)} (1/2)^{sp(1-\frac{1}{r} - \frac{n}{spq})} \omega \le \omega . \end{aligned} \end{aligned}$$

Here we have used \(1-\frac{1}{r}-\frac{n}{spq} > 0\). Notice that in the case of \(sp \ge n\), we are assuming \(1-\frac{1}{r} -\frac{1}{q} > 0\) which is a stronger assumption. Now we verify the assumption on the tail.

$$\begin{aligned} \begin{aligned} \sup _{-1\le t \le 0} \int _{{\mathbb {R}}^n \setminus B_1} \frac{|{\tilde{u}}|^{p-1}}{|x|^{n+sp}} \;\textrm{d}x&= \frac{2^{-sp}}{L^{p-1}} \sup _{t_0 - \frac{L^{2-p}}{2^{sp}}\le t \le t_0} \int _{{\mathbb {R}}^n \setminus B_{\frac{1}{2}(x_0)}} \frac{|u(y)|^{p-1}}{|y-x_0|^{n+sp}} \;\textrm{d}y \\&\le \frac{1}{L^{p-1}} \sup _{-2\le t \le 0} \textrm{Tail}_{p-1,sp}(u(\,\bullet \,,t);x_0, \frac{1}{2})^{p-1} \\&\le \frac{1}{L^{p-1}} (\frac{1}{2})^{sp}\left( \frac{1}{1-|x-x_0|}\right) ^{n+sp} \\&\quad \times \sup _{-2\le t\le 0}\textrm{Tail}_{p-1,sp}(u(\,\bullet \,,t);0,1)^{p-1} \\&\quad + \frac{2^n}{L^{p-1}} \sup _{-2\le t\le 0} \Vert u(\,\bullet \,,t)\Vert _{L^{p-1}(B_1(0))}^{p-1} \\&\le \frac{2^n}{L^{p-1}} \bigl ( 1+ |B_1| \Vert u\Vert _{L^\infty (Q_{1,2})} \bigr ) \le \frac{2^n(1+|B_1|)}{L^{p-1}} \le 1. \end{aligned} \end{aligned}$$

Now we can apply Step 1 to \({\tilde{u}}\) and we get the decay

$$\begin{aligned} \Vert {\tilde{u}}-{\tilde{u}}(0,0)\Vert _{L^\infty (G_r)} \le Cr^\alpha , \quad \text {for} \;\; 0<r<1 \end{aligned}$$

or in other words

$$\begin{aligned} |{\tilde{u}}(x,t)-{\tilde{u}}(0,0)| \le C (|x|^\alpha + |t|^{\frac{\alpha }{\beta }}), \quad \text {for} \;\; (x,t) \in Q_{1,1}. \end{aligned}$$

In terms of u, this means

$$\begin{aligned}{} & {} |u(x,t)-u(x_0,t_0)|\le CL (2^\alpha |x-x_0|^{\alpha } + (2^{sp}L^{p-2})^{\frac{\alpha }{\beta }} |t-t_0|^{\frac{\alpha }{\beta }}), \quad \text {for} \nonumber \\{} & {} \;\; (x,t) \in Q_{\frac{1}{2},\frac{1}{2^{sp} L^{p-2}}}(x_0,t_0). \end{aligned}$$
(4.8)

Now take two points \((x_1,t_1)\,, \, (x_2,t_2) \in Q_{\frac{1}{2}, \frac{1}{2^{sp}}}\) and split the line joining them into \(1+ [L^{p-2}]\) pieces, say \((y_i,\tau _i)_{i=0}^{1+[L^{p-2}]}\) with \((x_1,t_1)=(y_0,\tau _0)\), \((x_2,t_2)=(y_{1+[L^{p-2}]},\tau _{1+[L^{p-2}]})\), \(|y_{i+1}-y_i|= \frac{|x_2-x_1|}{1+[L^{p-2}]} < \frac{1}{2}\) and \(|\tau _{i+1}- \tau _i|= \frac{|t_2-t_1|}{1+[L^{p-2}]} < \frac{1}{2^{sp}L^{p-2}}\) so that \((y_{i+1},\tau _{i+1}) \in Q_{\frac{1}{2}, \frac{1}{2^{sp}L^{p-2}}}(y_i,\tau _i)\). By (4.8) applied in each of \(Q_{\frac{1}{2}, \frac{1}{2^{sp}L^{p-2}}}(y_i,\tau _i)\) obtain

$$\begin{aligned} \begin{aligned} |u(x_2,t_2) - u(x_1,t_1) |&\le \sum _{i=0}^{[L^{p-2}]} |u(y_{i+1},\tau _{i+1})-u(y_i,\tau _i)| \\&\le CL\sum _{i=0}^{[L^{p-2}]} 2^\alpha |y_{i+1} -y_i|^\alpha + (2^{sp} L^{p-2})^{\frac{\alpha }{\beta }} |\tau _{i+1}-\tau _i|^{\frac{\alpha }{\beta }} \\&\le C(1{+}L)^{p-1} \Bigl ( \Bigl (2\frac{|x_2{-}x_1|}{1{+}[L^{p-2}]}\Bigl )^\alpha {+} \Bigl (2^{sp}L^{p-2} \frac{|t_2-t_1|}{1{+}\lfloor L^{p-2} \rfloor }\Bigl )^{\frac{\alpha }{\beta }} \Bigr )\\&\le C(n,s,p,q,r,\alpha ) (|x_2 - x_1|^\alpha + |t_2 - t_1|^{\frac{\alpha }{\beta }} ). \end{aligned} \end{aligned}$$

\(\square \)

Now we prove the Hölder regularity at any scale.

Proof of Theorem 1.2

We will consider the rescaled functions

$$\begin{aligned} {\tilde{u}}_{\iota }(x,t) = \frac{1}{\mu } u(Rx + x_0, \mu ^{2-p}R^{sp} t + \iota + T_0) \end{aligned}$$

with

$$\begin{aligned} \begin{aligned} \mu =&1+ \Vert u\Vert _{L^\infty (Q_{R,2R^{sp}}(x_0,T_0))} + \sup _{ T_0- 2R^{sp}\le t \le T_0}\textrm{Tail}_{p-1,sp} (u(\,\bullet \,,t);x_0,R) \\&\quad + \left( \frac{R^{sp-\frac{n}{q} - \frac{sp}{r}} \Vert f\Vert _{L^{q,r}(Q_{R,2R^{sp}}(x_0,T_0))}}{\omega } \right) ^{\frac{1}{p-1+ \frac{p-2}{r}}}, \end{aligned} \end{aligned}$$

where \(\omega = \omega (n,s,p,q,r,\alpha )\) is the same as in the proof of Proposition 4.1 and \(\iota \in [ -(R/2)^{sp} (1-\mu ^{2-p}),0]\). The interval \([ -(R/2)^{sp} (1-\mu ^{2-p}),0]\) is chosen so that the cylinders \(Q_{\frac{R}{2},\frac{\mu ^{2-p} R^{sp}}{2^{sp}}}(x_0,T_0+\iota )\) cover all of \(Q_{\frac{R}{2},(\frac{R}{2})^{sp}}(x_0,T_0)\) by varying \(\iota \) over. Note that for these choices of \(\iota \) we have \(Q_{R,2\mu ^{2-p}R^{sp}}(x_0,T_0+\iota )\subset Q_{R,2R^{sp}}(x_0,T_0)\). Then \({\tilde{u}}\) is a solution of

$$\begin{aligned} \partial _t {\tilde{u}}_\iota + (-\Delta _p^s){\tilde{u}}_\iota = R^{sp}\frac{f(Rx +x_0,\mu ^{2-p} R^{sp} t + \iota + T_0)}{\mu ^{p-1}}, \quad \text {in} \quad Q_{1,2}. \end{aligned}$$

We now verify that \({\tilde{u}}_\iota \) satisfies the conditions of Proposition 4.1. The \(L^{q,r}\) norm of the right-hand side is

$$\begin{aligned} \begin{aligned} \Bigl \Vert R^{sp} \frac{f(Rx,\mu ^{2-p} R^{sp} t + \iota )}{\mu ^{p-1}}\Bigr \Vert _{L^{q,r}(Q_{1,2})}&= \frac{\mu ^{\frac{p-2}{r}}}{2^{\frac{1}{r}}\mu ^{(p-1)}} R^{sp-\frac{n}{q}- \frac{sp}{r}} \Vert f\Vert _{L^{q,r}(Q_{R,2\mu ^{2-p}R^{sp}}(x_0,T_0+ \iota ))} \\&\le \frac{R^{sp-\frac{n}{q} - \frac{sp}{r}} \Vert f\Vert _{L^{q,r}(Q_{R,2R^{sp}}(x_0,T_0))}}{2^{\frac{1}{r}}\mu ^{p-1 -\frac{p-2}{r}}} \\&\le \frac{\omega }{2^{\frac{1}{r}}} < \omega . \end{aligned} \end{aligned}$$

The \(L^\infty \) norm of \({\tilde{u}}_\iota \) satisfies

$$\begin{aligned} \Vert {\tilde{u}}_\iota \Vert _{L^\infty (Q_{1,2}(0,0))} = \frac{1}{\mu } \Vert u\Vert _{L^\infty (Q_{R,2\mu ^{2-p}R^{sp}(x_0,T_0+\iota )})} \le \frac{1}{\mu } \Vert u\Vert _{L^\infty (Q_{R,2R^{sp}})} \le 1. \end{aligned}$$

Similarly

$$\begin{aligned} \begin{aligned} \sup _{ -2 \le t \le 0}\textrm{Tail}_{p-1,sp} ({\tilde{u}}(\,\bullet \,,t);0,1)&\le \frac{1}{\mu } \sup _{ T_0+\iota - 2\mu ^{2-p} R^{sp}\le t \le T_0+\iota }\textrm{Tail}_{p-1,sp} (u(\,\bullet \,,t);x_0,R) \\&\le \frac{1}{\mu }\sup _{ T_0- 2R^{sp}\le t \le T_0}\textrm{Tail}_{p-1,sp} (u(\,\bullet \,,t);x_0,R) \le 1. \end{aligned} \end{aligned}$$

Hence, using Proposition 4.1 for \({\tilde{u}}_\iota \), we get

$$\begin{aligned}{} & {} |{\tilde{u}}_\iota ({\tilde{x}}_2,{\tilde{t}}_2) - {\tilde{u}}_\iota ({\tilde{x}}_1,{\tilde{t}}_1)|\le C(|{\tilde{x}}_2 - {\tilde{x}}_1|^\alpha + |{\tilde{t}}_2 - {\tilde{t}}_1|^{\frac{\alpha }{sp-(p-2)\alpha }})\\{} & {} \text { for } ({\tilde{x}}_1,{\tilde{t}}_1),\; ({\tilde{x}}_2,{\tilde{t}}_2) \in Q_{\frac{1}{2},\frac{1}{2^{s \, p}}}(0,0), \end{aligned}$$

with \(C= C(n,s,p,q,r,\alpha ) \). This translates to

$$\begin{aligned} |u(x_2,\tau _2) - u(x_1,\tau _1)| \le \mu C \Bigl [ \Bigl (\frac{|x_2-x_1|}{R} \Bigr )^\alpha + \Bigl (\frac{|\tau _2-\tau _1|}{R^{s\, p} \mu ^{2-p}} \Bigr )^{\frac{\alpha }{sp-(p-2)\alpha }} \Bigr ], \end{aligned}$$
(4.9)

for \((x_1,\tau _1), \; (x_2,\tau _2) \in Q_{\frac{R}{2}, \frac{R^{s \, p}\mu ^{2-p}}{2^{s \, p}}}(x_0,T_0+\iota )\). Now we vary \(\iota \) to obtain an estimate in the whole \(Q_{\frac{R}{2}, (\frac{R}{2})^{s \, p}}\). Specifically we split the interval \([t_1,t_2]\) into \(1+ \lfloor \mu ^{p-2} \rfloor \) pieces, say \([\tau _{i+1},\tau _i]\), with \(\tau _i - \tau _{i+1}=\frac{|t_2-t_1|}{1+\lfloor \mu ^{p-2} \rfloor }\), \(\tau _0=t_2\), and \(\tau _{\lfloor 1+\mu ^{p-2} \rfloor } = t_1\). Using (4.9), we obtain

$$\begin{aligned} \begin{aligned} |u(x_2,t_2) - u(x_1,t_1)|&\le |u(x_2,t_1) - u(x_1,t_1)| + |u(x_2,t_2) - u(x_2,t_1)| \\&\le \mu C \Bigl (\frac{|x_2-x_1|}{R} \Bigr )^\alpha + \sum _{i=0}^{\lfloor \mu ^{p-2} \rfloor } |u(x_2,\tau _{i}) - u(x_2,\tau _{i+1})| \\&\le \mu C \Bigl [ \Bigl (\frac{|x_2-x_1|}{R} \Bigr )^\alpha + \sum _{i=0}^{\lfloor \mu ^{p-2} \rfloor } \Bigl ( \frac{|\tau _i-\tau _{i+1}|}{R^{s\, p} \mu ^{2-p}} \Bigr )^{\frac{\alpha }{sp-(p-2)\alpha }} \Bigr ] \\&= \mu C \Bigl [ \Bigl (\frac{|x_2{-}x_1|}{R} \Bigr )^\alpha {+} \sum _{i=0}^{\lfloor \mu ^{p-2} \rfloor } \Bigl ( \frac{|t_2{-}t_1|}{R^{s\, p} \mu ^{2-p}(1+ \lfloor \mu ^{p-2} \rfloor )} \Bigr )^{\frac{\alpha }{sp-(p-2)\alpha }} \Bigr ] \\&\le \mu C \Bigl [ \Bigl (\frac{|x_2-x_1|}{R} \Bigr )^\alpha + \sum _{i=0}^{\lfloor \mu ^{p-2} \rfloor } \Bigl ( \frac{|t_2-t_1|}{R^{s\, p}} \Bigr )^{\frac{\alpha }{sp-(p-2)\alpha }} \Bigr ] \\&\le \mu C \Bigl [ \Bigl (\frac{|x_2-x_1|}{R} \Bigr )^\alpha + 2 \mu ^{p-2}\Bigl ( \frac{|t_2-t_1|}{R^{s\, p}} \Bigr )^{\frac{\alpha }{sp-(p-2)\alpha }} \Bigr ], \end{aligned} \end{aligned}$$

which concludes the desired result. \(\square \)