1 Introduction

Nonlinear Calderón-Zygmund theory for elliptic and parabolic partial differential equations have been extensively studied in recent decades. In particular, an interior \(W^{2,p}\)-estimate with \(p>d\) for a class of fully nonlinear elliptic equations, was first obtained by Caffarelli in [6] via the Aleksandrov-Bakel’man-Pucci estimate, the covering argument and the Harnack inequality by using an improvement of Krylov-Safonov’s technique. A similar interior estimate was extended by Wang in [26] to fully nonlinear parabolic equations by adapting the Aleksandrov-Bakel’man-Pucci-Krylov-Tso maximum principle and the compactness method. Later, Escauriaza in [11] further sharpened Caffarelli’s result in [6] to the range \(p>d-\varepsilon\) with a small \(\varepsilon >0\) depending on the ellipticity constants and d by considering the reverse Hölder inequality. Meanwhile, we also notice that a solvability in \(W^{2,1}_{p,loc}(\Omega _T) \cap C(\overline{\Omega }_T)\) for the initial-boundary problem of fully nonlinear parabolic equations was obtained by Crandall, Kocan and Świech in [7]. Winter in [27] further made use of the Alexandroff maximum principle and the weak Harnack inequality on the boundary setting to establish the boundary estimate as well as the \(W^{2,p}\)-solvability for the Dirichlet boundary problem associated nonlinear elliptic equations. Recently, an interesting work from Krylov [14] provided a unified approach to show the \(L^p\)-solvability for nondivergence elliptic and parabolic equations under the regular assumption of \(VMO_{x}\) principle coefficients. This approach mainly relies on the pointwise estimates of sharp functions for the spatial derivatives of solutions due to the Fefferman-Stein theorem. Furthermore, such an approach was developed by Dong, Kim and Krylov to attain the \(W^{2,p}\)-solvability for fully nonlinear elliptic and parabolic equations with the nonlinearities being VMO in independent variables. For examples, Krylov [15] showed the \(W^{2,p}\)-solvability with \(p>d\) to the Bellman’s equations with the VMO nonlinearity by his unified argument. Dong-Krylov-Li [8] demonstrated an interior \(W^{2,p}\)-solvability for \(p>d\) and \(W^{2,1}_{p}\)-solvability for \(p>d+1\), respectively, to fully nonlinear elliptic and parabolic equations. More generally, Krylov [16] studied the existence and uniqueness in \(W^{2,p}\) for \(p>d\) to the strong solution of fully nonlinear elliptic equations \(H(u,Du,D^{2}u,x)=0\). Dong and Krylov in [9] proved the solvability in \(W^{2,1}_{p}\) for \(p>d+1\) to the strong solutions of fully nonlinear parabolic equations \(\partial _{t} u(t,x)+H(u,Du,D^{2}u;t,x)=0\) under some relaxed convexity assumptions instead of requiring H to be convex or concave with respect to \(D^{2}u\). Very recently, Dong and Krylov [10] also derived the regularity in the mixed-norm Sobolev spaces to fully nonlinear elliptic and parabolic equations by improving the Fefferman-Stein theorem in the mixed-norm weighted Lebesgue spaces.

This paper is actually a continuation of Dong-Krylov-Li’s work in [8], which extends it in two folds: the Fefferman-Stein theorem in the Lorentz spaces and Lorentz regularity of fully nonlinear parabolic equations. Here, the Fefferman-Stein inequality in the Lorentz spaces is also inspired by Dong and Krylov’s paper [10] involving that in the mixed-norm form. Our main aim is to attain a global estimate in the Sobolev-Lorentz spaces by using a unified approach to fully nonlinear parabolic equations with the nonlinearities being small BMO oscillation in independent variable in a bounded \(C^{1,1}\) parabolic domain. To state our problem under consideration more precisely, let us recall some related notations. Let \(\Omega\) be a bounded domain of \(\mathbb {R}^{d}\) for \(d\ge 2\) with \(C^{1,1}\)-smooth boundary, and \(\Omega _{T}:= \Omega \times (0,T)\) for \(T>0\) with its usual parabolic boundary \(\partial \Omega _{T} := \Big ( \partial \Omega \times (0,T] \Big ) \cup \Big ( \Omega \times \{0\} \Big )\). We consider the following zero Cauchy-Dirichlet problem of a fully nonlinear parabolic equation:

$$\begin{aligned} \left\{ \begin{array}{ll} \displaystyle { \partial _{t} u(x,t) + H(D^{2}u, Du, u, x,t) = 0 } &{} \quad \text {in} \,\,\, \Omega _{T}, \\ u = 0 &{} \quad \text {on} \,\,\, \partial \Omega _{T}, \end{array} \right. \end{aligned}$$
(1.1)

where \(D^{2} u = \Big ( D_{ij} u\Big )_{d\times d}\) denotes the Hessian matrix of u, and \(D u=\Big ( D_{i} u \Big )_{d\times 1}\) denotes its gradient.

As usual, we use \(\mathring{W}^{1,\,2}_{2}(\Omega _{T})\) to denote the set of all functions \(u\in L^2(\Omega _{T})\) that their weak derivatives \(Du,D^2u,D_tu\) belong to \(L^2\)-spaces, and u vanishes on the parabolic boundary \(\partial \Omega _{T}\) in the trace sense. Therefore, a strong solution of (1.1) which is treated throughout this paper is a Sobolev function \(u \in \mathring{W}^{1,\,2}_{2}(\Omega _{T})\) satisfying (1.1) almost everywhere in \(\Omega _{T}\), for more details to see [18, Chapter VII]. In this context, we suppose that the nonlinearity \(H(D^{2}u, Du, u, x,t)\) can be decomposed into following two parts:

$$\begin{aligned} H(D^{2}u, Du, u, x,t) = F(D^{2}u, x,t) + G(D^{2}u, Du, u, x,t). \end{aligned}$$

Let \({\mathcal {S}}\) be the set of all symmetric \(d\times d\)-matrices, and they satisfy the following structural assumptions:

H1. We suppose that \(F(u'',x,t)\) is convex and positive homogeneous of degree one with respect to \(u'' \in {\mathcal {S}}\), and there exist two constants \(0<\lambda \le \Lambda < \infty\) such that

$$\begin{aligned} \lambda \, |\xi |^{2} \le F(u''+ \xi \xi ^{*}, x,t) - F(u'', x,t) \le \Lambda \, |\xi |^{2} \end{aligned}$$

for any \(u'' \in {\mathcal {S}}\), \(\xi \in \mathbb {R}^{d}\) and each \((x,t)\in \Omega _{T}\).

H2. For any fixed \(u''\in {\mathcal {S}}\), \(u'\in \mathbb {R}^{d}\) and \((x,t)\in \Omega _T\), we let \(G(u'', u', u, x,t)\) be a monotone nonincreasing function in \(u\in \mathbb {R}\). Moreover, there exists a positive constant K such that

$$\begin{aligned} \left| G(u'',u',u,x,t) - G(u'',v',v,x,t) \right| \le K \, \Big ( |u'-v'| + |u-v| \Big ) \end{aligned}$$

and

$$\begin{aligned} \left| G(u'',u',u,x,t)\right| \le \psi \left( |u''| \right) \, |u''| + K \, \Big ( |u'| + |u| \Big ) + g(x,t) \end{aligned}$$

for any \(u'' \in {\mathcal {S}}\), \(u',v'\in \mathbb {R}^{d}\), \(u,v\in \mathbb {R}\) and each \((x,t)\in \Omega _{T}\). In the above, the function \(\psi (\rho ): \overline{\mathbb {R}^{+}} \rightarrow \overline{\mathbb {R}^{+}}\) is bounded, monotonically decreasing and \(\psi (\rho ) \rightarrow 0\) as \(\rho \rightarrow 0\), while g(xt) is a given function specified later.

Before stating our main result, it is necessary to introduce some notations. For any \(r>0\) and \((x,t)\in \mathbb {R}^{d+1}\), we denote

$$\begin{aligned} B_{r}(x) = \left\{ y \in \mathbb {R}^{d}: \,\, |x-y|<r \right\} \quad \text {and} \quad Q_{r}(x,t) = B_{r}(x) \times (t, \, t+r^{2}). \end{aligned}$$

For convenience, we write \(B_{r}=B_{r}(0)\) and \(Q_{r}=Q_{r}(0,0)\) in the following context. An average of f(xt) over \(Q_{r}\) is denoted by

$$\begin{aligned} \bar{f}_{Q_{r}} = \fint _{Q_{r}} f(x,t) \, dxdt = \frac{1}{|Q_{r}|} \int _{Q_{r}} f(x,t) \, dxdt, \end{aligned}$$

where \(|Q_{r}|\) is \((d+1)\)-dimensional Lebesgue measure of \(Q_{r}\). To our aim, we need to impose the following \((\delta , R_{0})\)-vanishing assumption on the leading term in (1.1) for some \(\delta\) and \(R_{0}>0\).

Assumption 1.1

For any \(u''\in S\) with \(|u''|=1\), we say that \(F(u'',x,t)\) is \((\delta , R_{0})\)-vanishing of (xt). If there exists \(R_{0}>0\) and a function \(\widetilde{F}(u'')\) (independent of \(z=(x,t)\)) satisfying (H1), such that for any \(r\in (0, R_{0}]\) and \((s,y) \in \overline{\Omega }_{T}\), we have

$$\begin{aligned} \sup _{u''\in S, \,\,|u''|=1} \fint _{Q_{r}(s,y)} \left| F(u'', x,t) - \widetilde{F}(u'') \right| \, dxdt \le \delta , \end{aligned}$$

where \(F(u'', x,t)\) is zero extended from \(Q_{r}(s,y) \cap \Omega _{T}\) to \(Q_{r}(s,y)\) for (xt) close to the parabolic boundary, and the parameter \(\delta >0\) will be specified later.

We refer to [8] for the same assumptions as H1, H2 and Assumption 1.1, where Dong, Krylov and Li established the \(L^{p}\) estimate with \(p>d+1\) for the zero Cauchy-Dirichlet problems (1.1) in a bounded \(C^{1,1}\)-smooth domain. In the article, our aim is to extend it to a regularity in the scale of Lorentz space which is extremely useful when dealing with borderline existence as a refined version of Lebesgue spaces. So, it is necessary to recall the Lorentz space and Sobolev-Lorentz space since our Sobolev-Lorentz regularity for solutions is taken into consideration. The Lorentz space is a two-parameter scale of spaces which make the classical Lebesgue spaces refined in some sense, and the Sobolev-Lorentz space is similarly done from Sobolev space. More precisely, we have

Definition 1.2

  1. (i)

     Let U be an open subset in \(\mathbb {R}^{d+1}\). For \(p\in [1, \infty )\) and \(q\in (0, + \infty )\), the Lorentz space \(L^{(p,q)}(U)\) is the set of all measurable functions \(f: U \rightarrow \mathbb {R}^{d+1}\) such that

    $$\begin{aligned} \Vert f\Vert _{L^{(p,q)}(U)} := \Bigg ( \, p \int _{0}^{\infty }\Big ( \, \kappa ^{p}|\{(x,t)\in U: |f(x,t)|>\kappa \}| \, \Big )^{\frac{q}{p}}\frac{d\kappa }{\kappa } \, \Bigg )^{\frac{1}{q}} < + \infty . \end{aligned}$$
    (1.2)

    For \(q=\infty\), the space \(L^{(p,\infty )}(U)\) is set to be the usual Marcinkiewicz space with quasinorm

    $$\begin{aligned} \Vert f\Vert _{L^{(p,\infty )}(U)} := \sup _{\kappa>0} \Big ( \, \kappa ^{p} |\{(x,t)\in U: |f(x,t)|>\kappa \}| \, \Big )^{\frac{1}{p}} < + \infty . \end{aligned}$$
    (1.3)
  2. (ii)

     The Sobolev-Lorentz space \(W^{1, \, 2}_{(p,q)}(U)\) is defined by

    $$\begin{aligned} W^{1, \, 2}_{(p,q)}(U) = \left\{ \, f : \,\, f, \, \partial _{t}f, \, Df, \, D^{2}f \, \in \, L^{(p,q)}(U) \, \right\} , \end{aligned}$$

    where the quasinorm of f is

    $$\begin{aligned} \left\| f\right\| _{W^{1, \, 2}_{(p,q)}(U)} := \left\| f\right\| _{L^{(p,q)}(U)} + \left\| \partial _{t}f \right\| _{L^{(p,q)}(U)} + \left\| Df\right\| _{L^{(p,q)}(U)} + \left\| D^{2}f\right\| _{L^{(p,q)}(U)}. \end{aligned}$$

The regularity in the scale of Lorentz spaces concerning partial differential equations was originated from Talenti’s work [23] based on the symmetrization argument. In 2010, Mingione [21] developed some sharp estimates in the Lorentz spaces for nonlinear elliptic equations in divergence form by using the duality exponent. Using the geometric argument first introduced by Byun and Wang [4], Mengesha and Phuc [20] recently obtained a global weighted Lorentz estimate to quasilinear elliptic equations with small BMO coefficients in Reifenberg flat domains. In addition, Baroni [2] showed a local Lorentz estimate for parabolic systems of p-Laplacian type by improving the classic Hardy inequality and the reverse Hölder inequality, and a similar result was also obtained in [3] for parabolic obstacle problems of p-Laplacian type by using the same technique. A global extension in the weighted Lorentz spaces was established by Tian and Zheng in [24] to the zero Cauchy-Dirichlet problems of nonlinear parabolic equations with small BMO nonlinearities in Reifenberg domains. Recently, Bui and Duong [5] got a global weighted Lorentz estimate to the weak solutions of divergence nonlinear parabolic equations with nonstandard growth over Reifenberg domains. Very recently, Tian and Zheng [25] also proved a global estimate of weak solutions in variable Lorentz spaces to parabolic obstacle problems with nonstandard growth over a quasiconvex domain based on the large-M-inequality principle originated from Acerbi and Mingione’s work [1]. Zhang and Zheng in [28] also showed local Lorentz estimates for a class of fully nonlinear parabolic and elliptic equations by using the large-M-inequality principle.

We now summarize a global Sobolev-Lorentz estimate for fully nonlinear parabolic equation (1.1) as follows.

Theorem 1.3

Let \(\Omega\) be a bounded \(C^{1,1}\)-smooth domain and \(T>0\). For the Cauchy-Dirichlet problems (1.1) with assumptions H1 and H2, then there exists a small constant \(\delta =\delta (d,\lambda ,\Lambda ,p,q,\partial \Omega ) >0\) such that if the \(F(u'',x,t)\) is \((\delta , R_{0})\)-vanishing shown as Assumption 1.1for some \(R_{0}>0\), then the following is true:

  1. (i)

      For \(u\in W^{1,2}_{(p,q)}(\Omega _{T})\) satisfying (1.1) with \(g \in L^{(p,q)}(\Omega _{T})\) for \(p\in (d+1, \, +\infty )\) and \(q\in (0, \, +\infty ]\), we have

    $$\begin{aligned} \left\| u \right\| _{W^{1,2}_{(p,q)}(\Omega _{T})} \le N \left( \left\| g \right\| _{L^{(p,q)}(\Omega _{T})} + \bar{\rho } \left\| \psi \right\| _{L^{\infty }} \right) \end{aligned}$$
    (1.4)

    for some \(\bar{\rho }>0\), where N is a positive constant depending only on \(d,p,q,\lambda ,\Lambda ,\delta ,R_{0},K,T,|\Omega |\) and \(\Vert \partial \Omega \Vert _{C^{1,1}}\) while \(q<\infty\), and the constant is independent of q while \(q=\infty\).

  2. (ii)

      For any \(g \in L^{(p,q)}(\Omega _{T})\) with \(p\in (d+1, \, +\infty )\) and \(q\in (0, \, +\infty ]\), there is a unique solution \(u\in W^{1,2}_{(p,q)}(\Omega _{T})\) of (1.1). In particular, if \(\psi (0)=0\) in (H2), then there is a unique solution \(u\in W^{1,2}_{(p,q)}(\Omega _{T})\) of (1.1) with the estimate

    $$\begin{aligned} \left\| u \right\| _{W^{1,2}_{(p,q)}(\Omega _{T})} \le N \left\| g \right\| _{L^{(p,q)}(\Omega _{T})}. \end{aligned}$$
    (1.5)

It is also worth noting that a solvability theorem in the space \(W^{1,2}_{p,loc}(\Omega _T) \cap C(\overline{\Omega }_T)\) can be found in [7] for the initial boundary value problem for a class of fully nonlinear parabolic equations. The solvability in the Sobolev spaces \(W^{1,2}_p(\overline{\Omega }_T)\) for \(p > d +1\) of the terminal boundary value problem is proved by Dong, Krylov and Li in [8] for such fully nonlinear parabolic equations (1.1), in bounded cylindrical domains, in the case of VMO “coefficients”. Here, we further prove the solvability theorem 1.3 in the scale of Sobolev-Lorentz spaces under the assumptions of H1 and H2 along this line from Dong-Krylov’s argument. A key ingredient of our main proof is that we need to improve the Fefferman-Stein type theorem of sharp functions in the Lorentz spaces. We note that Krylov in [15] gave a modified Fefferman-Stein theorem in Lebesgue spaces by a filtration of partitions, which leads to that the \(L^p\)-solvability was obtained for Bellman’s equations with VMO coefficients in the whole space. Furthermore, Dong and Krylov [10] improved the Fefferman-Stein theorem in the mixed-norm weighted Lebesgue spaces, which yields a mixed-norm regularity for fully nonlinear elliptic and parabolic equations. Inspired by their works, we are going to establish a variant of the Fefferman-Stein theorem in the Lorentz spaces on the basis of the Calderón-Zygmund decomposition in the parabolic setting. Another important point is that the partitions of unity do not work for global estimates of the solutions in the Lorentz-Sobolev spaces, since the commuting transformation of integral does not hold in the Lorentz spaces. To this end, we make use of the idea of local \(L^{p}\)-estimate for nondivergence parabolic equations to handle the mean oscillation estimates of Hessian introduced by Krylov’s paper [15], which originated from Lin’s paper [19] for nondivergence elliptic equations.

Let us now consider the following Dirichlet problem of fully nonlinear elliptic equations:

$$\begin{aligned} \left\{ \begin{array}{ll} \displaystyle { H(D^{2}u, Du, u, x) = 0 } &{} \quad \text {in} \,\,\, \Omega , \\ u = 0 &{} \quad \text {on} \,\,\, \partial \Omega . \end{array} \right. \end{aligned}$$
(1.6)

As a direct consequence of main Theorem , we immediately have the following solvability regarding the elliptic setting.

Corollary 1.4

Let \(\Omega\) be a bounded \(C^{1,1}\)-smooth domain. For the Dirichlet problem (1.6) with assumptions H1 and H2, then there exists a small constant \(\delta =\delta (d,\lambda ,\Lambda ,p,q,\partial \Omega ) >0\) such that if the \(F(u'',x)\) is \((\delta , R_{0})\)-vanishing for some \(R_{0}>0\), then the following is true:

  1. (i)

      For \(u\in W^{2}_{(p,q)}(\Omega )\) satisfying (1.6) with \(g \in L^{(p,q)}(\Omega )\) with \(p\in (d, \, +\infty )\) and \(q\in (0, \, +\infty ]\), we have

    $$\begin{aligned} \left\| u \right\| _{W^{2}_{(p,q)}(\Omega )} \le N \left( \left\| g \right\| _{L^{(p,q)}(\Omega )} + \bar{\rho } \left\| \psi \right\| _{L^{\infty }} \right) \end{aligned}$$
    (1.7)

    for some \(\bar{\rho }>0\), where N is a positive constant depending only on \(d,p,q,\lambda ,\Lambda ,\delta ,R_{0},K,|\Omega |\) and \(\Vert \partial \Omega \Vert _{C^{1,1}}\) if \(q<\infty\), while the constant is not dependent of q if \(q=\infty\). Specifically, if \(\psi (0)=0\) in (H2), then we have

    $$\begin{aligned} \left\| u \right\| _{W^{2}_{(p,q)}(\Omega )} \le N \left\| g \right\| _{L^{(p,q)}(\Omega )}. \end{aligned}$$
    (1.8)
  2. (ii)

       For any \(g \in L^{(p,q)}(\Omega )\) with \(p\in (d, \, +\infty )\) and \(q\in (0, \, +\infty ]\), there is a unique solution \(u\in W^{2}_{(p,q)}(\Omega )\) of (1.6).

The remainder of this paper is organized as follows. In Sect. 2, we present the Fefferman-Stein type theorem in the Lorentz spaces, and some necessary preliminary results concerning the Gagliardo-Nirenberg interpolation inequality and Alexandrov’s estimate in the Lorentz spaces. In Sect. 3 we complete the proof of Theorem 1.3. Moreover, as a corollary of main Theorem 1.3, a global Sobolev-Lorentz estimate of the solution is proved for parabolic Bellman’s equations with small BMO coefficients in a bounded \(C^{1,1}\)-smooth domain.

2 Preliminary results

This section is devoted to some related notations and basic facts which will be useful in the main proof. Throughout this paper, we denote \(N(d,p,q,\lambda ,\Lambda ,\cdots )\) or \(N_{i}(d,p,q,\lambda ,\Lambda ,\cdots )\) for \(i=0, 1,2,\cdots\), a universal constant depending only on prescribed quantities and possibly varying from line to line. First, we give a generalized version of the Fefferman-Stein theorem of sharp functions in the Lorentz spaces. Let \(\mathfrak {C}\) be a set of all \({\mathcal {C}}_{(R^{2}, \, R)}(x,t)\) for any \(R>0\) with the fixed point \((x,t) \in \mathbb {R}^{d+1}\). When f is locally integrable on \(\mathbb {R}^{d+1}\), we define the parabolic maximal function \({\mathcal {M}}f\) by

$$\begin{aligned} {\mathcal {M}}f(x,t) = \sup _{{\mathcal {C}} \in \mathfrak {C}} \fint _{\mathcal {C}} |f(s,y)| \, dyds. \end{aligned}$$
(2.1)

For \(\gamma \in (0,1]\), we set

$$\begin{aligned} f^{\sharp }_{\gamma }(x,t) = \sup _{\mathcal {C} \in \mathfrak {C}} \left( \fint _{\mathcal {C}} \fint _{\mathcal {C}} |f(s,y) - f(\tau ,z)|^{\gamma } \,\, dyds dzd\tau \right) ^{\frac{1}{\gamma }}. \end{aligned}$$
(2.2)

Remark 2.1

While \(\gamma =1\), we observe that \(f\in BMO\) is equivalent to \(f^{\sharp }_{1} \in L^{\infty }\). We also note that \(BMO\cong \mathcal {L}^{p,d}\)(the Campanato spaces) for any \(1\le p< \infty\) (cf. [22, Chapter IV]), which leads to this fact that the norms \(\Vert f^{\sharp }_{\gamma }\Vert _{L^{\infty }}\) and \(\Vert f^{\sharp }_{1}\Vert _{L^{\infty }}\) are equivalent for any \(\gamma \in (0,1]\).

The Hardy-Littlewood maximal function theorem states that if \(f\in L^{p}\) for \(1<p\le \infty\), then \(\mathcal {M}f \in L^{p}\) and \(\Vert \mathcal {M}f \Vert _{L^{p}} \le N(p,d) \, \Vert f \Vert _{L^{p}}\), for instance, see [22, Chapter I]. By interpolating argument, Mengesha and Phuc in [20, Lemma 3.11] extended the above boundedness of the Hardy-Littlewood maximal function from the Lebesgue spaces to the weighted Lorentz spaces. Similarly, if \(f^{\sharp }_{\gamma } \in L^{p}\) for \(1<p<\infty\), then \(\mathcal {M}f \in L^{p}\) and \(\Vert \mathcal {M}f \Vert _{L^{p}} \le N(d,p,\gamma ) \, \Vert f^{\sharp }_{\gamma }\Vert _{L^{p}}\) as given by Krylov in [15, Theorem 5.3]. Therefore, the interpolation method is also applied to the Fefferman-Stein sharp function in the Lorentz spaces just as what Mengesha and Phuc did. Indeed, for any fixed \(1<p<\infty\), we take \(p_1 = \frac{p+1}{2}\) to obtain

$$\begin{aligned} \big \Vert \mathcal {M}f \big \Vert _{L^{p_1}(\mathbb {R}^{d+1})} \le N_1 \, \big \Vert f^{\sharp }_{\gamma } \big \Vert _{L^{p_1}(\mathbb {R}^{d+1})} \end{aligned}$$

for any \(f^{\sharp }_{\gamma } \in L^{p_1}(\mathbb {R}^{d+1})\); while we take \(p_2 = p+1\) to show

$$\begin{aligned} \big \Vert \mathcal {M}f \big \Vert _{L^{p_2}(\mathbb {R}^{d+1})} \le N_2 \, \big \Vert f^{\sharp }_{\gamma } \big \Vert _{L^{p_2}(\mathbb {R}^{d+1})} \end{aligned}$$

for any \(f^{\sharp }_{\gamma } \in L^{p_2}(\mathbb {R}^{d+1})\). Note that for any \(1<p<\infty\), \(1<p_1< p<p_2<+\infty\). By taking \(\vartheta =1-\frac{1}{p}\) such that \(\frac{1}{p}=\frac{1-\vartheta }{p_1}+\frac{\vartheta }{p_2}\), we use the interpolation theorem for Lorentz spaces (cf. [12, Theorem 1.4.19]) to obtain the following conclusion

$$\begin{aligned} \big \Vert \, \mathcal {M}f \, \big \Vert _{L^{(p,q)}(\mathbb {R}^{d+1})} \le N_0N_1^{1-\vartheta }N^{\vartheta }_2 \, \big \Vert \, f^{\sharp }_{\gamma } \, \big \Vert _{L^{(p,q)}(\mathbb {R}^{d+1})}. \end{aligned}$$

In summary, we have the following conclusion.

Lemma 2.2

Let \(1<p<\infty\), \(0<q\le \infty\) and \(\gamma \in (0,1]\). Then, for any \(f^{\sharp }_{\gamma } \in L^{(p,q)}(\mathbb {R}^{d+1})\) there exists a constant N depending only on dpq and \(\gamma\) such that

$$\begin{aligned} \big \Vert \, \mathcal {M}f \, \big \Vert _{L^{(p,q)}(\mathbb {R}^{d+1})} \le N \, \big \Vert \, f^{\sharp }_{\gamma } \, \big \Vert _{L^{(p,q)}(\mathbb {R}^{d+1})}. \end{aligned}$$

Furthermore, we can also show the improved Fefferman-Stein Theorem (Lemma 2.2) in a half space via the odd or even extensions, see Lemma 5 in [8]. To this end, we write

$$\begin{aligned} \mathbb {R}^{d+1}_{+} = \left\{ \, (x,t) = (x,t^{1}, \cdots ,x^{d}) \in \mathbb {R}^{d+1}: \,\, x^{1}>0 \, \right\} . \end{aligned}$$

Lemma 2.3

Let \(1<p<\infty\), \(0<q\le \infty\) and \(\gamma \in (0,1]\). Then, for any \(f^{\sharp }_{\gamma } \in L^{(p,q)}(\mathbb {R}^{d+1}_{+})\) there exists a constant N depending only on dpq and \(\gamma\) such that

$$\begin{aligned} \big \Vert \, \mathcal {M}f \, \big \Vert _{L^{(p,q)}(\mathbb {R}^{d+1}_{+})} \le \, N \, \big \Vert \, f^{\sharp }_{\gamma } \, \big \Vert _{L^{(p,q)}(\mathbb {R}^{d+1}_{+})}. \end{aligned}$$

Next let us extend the Gagliardo–Nirenberg interpolation inequality from the Lebesgue spaces to the Lorentz spaces. We first recall the conclusion by Dong and Krylov in [10, Lemma 6.5 (ii)]. For a fix \(\rho >0\) and \(f(x,t)\in L^1_{loc}(\mathbb {R}^{d+1})\), we introduce the confined maximal function and the confined sharp function:

$$\begin{aligned} \mathbb {M}_{\rho }f(x,t)= & {} \sup _{\left\{ \begin{array}{c} \mathcal {C} \in \mathfrak {C} \,: \\ R \in [\rho ,\infty ) \end{array}\right\} } \fint _{\mathcal {C}} |f(s,y)| \, dyds, \\ f^{\sharp }_{1, \,\rho }(x,t)= & {} \sup _{\left\{ \begin{array}{c} \mathcal {C} \in \mathfrak {C} \,: \\ R \in (0,\rho ] \end{array}\right\} } \fint _{\mathcal {C}} \fint _{\mathcal {C}} |f(s,y) - f(\tau ,z)| \, dyds dzd\tau . \end{aligned}$$

Lemma 2.4

For any \(\rho >0\), \(p\in [1,\infty )\), and \(f\in W^{1,2}_{p,\,\text {loc}}(\mathbb {R}^{d+1})\), we have

$$\begin{aligned} \mathbb {M}_{\rho } \left( |Df|^{p} \right) \le N \, \mathbb {M}_{\rho }^{\frac{1}{2}} \left( |D^{2}f|^{p} \right) \mathbb {M}_{\rho }^{\frac{1}{2}} \left( |f|^{p} \right) + N \,\rho ^{-p} \,\mathbb {M}_{\rho } \left( |f|^{p} \right) , \end{aligned}$$

where \(N=N(d,p)\) is a positive constant.

Indeed, Lemma 2.4 yields the following interpolation inequality in the Lorentz space.

Lemma 2.5

Let \(1<p<\infty\) and \(0<q\le \infty\). Then for any \(f \in W^{1,\,2}_{(p,q),\,\text {loc}}(\mathbb {R}^{d+1})\), we have

$$\begin{aligned} \left\| \,Df\,\right\| _{L^{(p,q)}(\mathbb {R}^{d+1})} \le N \left( \varepsilon \left\| \,D^{2}f\,\right\| _{L^{(p,q)}(\mathbb {R}^{d+1})} + \varepsilon ^{-1} \left\| \,f\,\right\| _{L^{(p,q)}(\mathbb {R}^{d+1})} \right) \end{aligned}$$
(2.3)

for any \(\varepsilon >0\), where \(N=N(p,q,d)\) is a positive constant.

Proof

Indeed, we note that for any \(\rho >0\),

$$\begin{aligned} (Df)^{\sharp }_{1}(x,t) \le (Df)^{\sharp }_{1, \,\rho }(x,t) + 2 \mathbb {M}_{\rho }(|Df|)(x,t) \quad \text {in} \quad \mathbb {R}^{d+1}. \end{aligned}$$
(2.4)

By the Poincaré inequality we get

$$\begin{aligned} (Df)^{\sharp }_{1, \,\rho }(x,t) \le N \rho \mathcal {M}(|D^{2}f|)(x,t). \end{aligned}$$

Following Lemma 2.4 with \(p=1\) it yields

$$\begin{aligned} \mathbb {M}_{\rho }(|Df|)(x,t)\le & {} N \rho \mathcal {M}(|D^{2}f|)(x,t) + N \rho ^{-1} \,\mathcal {M}\left( |f| \right) (x,t). \end{aligned}$$

Putting the above two estimates into (2.4) deduces

$$\begin{aligned} (Df)^{\sharp }_{1}(x,t) \le N \rho \mathcal {M}(|D^{2}f|)(x,t) + N \rho ^{-1} \,\mathcal {M}\left( |f| \right) (x,t) \end{aligned}$$

for any \((x,t) \in \mathbb {R}^{d+1}\).

Now, we make use of Lemma 2.2 and the Hardy-Littlewood maximal function theorem in the Lorentz spaces to obtain the required result by taking \(\varepsilon =\rho\). This completes the proof. \(\square\)

In what follows, we consider the Cauchy-Dirichlet problem of linear parabolic equations in nondivergence form:

$$\begin{aligned} \left\{ \begin{array}{ll} \partial _{t} u + Lu := \partial _{t} u + a^{ij}(x,t) D_{ij} u + b^{i}(x,t)D_{i}u - c(x,t)u = \varphi (x,t) &{} \quad \text {in} \quad \Omega _{T}, \\ u=0 &{} \quad \text {on} \quad \partial \Omega _{T}. \end{array} \right. \end{aligned}$$
(2.5)

Here, we suppose that all coefficients \(a^{ij}(x,t)\), \(b^{i}(x,t)\) and c(xt) are bounded measurable functions defined on \(\Omega _{T}\) and the symmetric coefficient matrix \(A(x,t)=\Big (a^{ij}(x,t)\Big )\) satisfies an uniformly ellipticity, which means that there exist positive constants \(0<{\lambda }\le {\Lambda }< \infty\) such that

$$\begin{aligned} |b^{i}(x,t)| \le {\Lambda }, \quad i=1, \cdots , d; \quad |c(x,t)|\le {\Lambda } \end{aligned}$$
(2.6)

and

$$\begin{aligned} {\lambda } |\xi |^{2} \le A(x,t) \, \xi \xi ^{*} \le {\Lambda } |\xi |^{2} \end{aligned}$$
(2.7)

for any \(\xi \in \mathbb {R}^{d}\) and almost all \((x,t)\in \Omega _{T}\).

In the following we recall the classical Aleksandrov estimate to the Cauchy-Dirichlet problem (2.5), see [13, Lemma 3].

Lemma 2.6

Let \(p\ge d+1\). Assume that \(u \in \mathring{W}^{1,\, 2}_{p}(\Omega _{T})\) is a solution of (2.5) with (2.6) and (2.7). If \(\varphi (x,t) \in L^{p}(\Omega _{T})\) and \(c(x,t) \ge 0\) in \(\Omega _{T}\), then

$$\begin{aligned} {\text {ess}\sup }_{(x,t) \in \Omega _{T}}|u(x,t)| \le N \Big \Vert |c|^{\frac{d+1-p}{p}} |\det A|^{\frac{1}{p}} \varphi \, \Big \Vert _{L^{p}(\Omega _{T})}, \end{aligned}$$

where \(N=N(d,p,{\lambda },{\Lambda },|\Omega |)\).

Combining Lemma 2.6 with the imbedding inequality in the Lorentz spaces, it yields the following \(L^{(p,q)}\)-estimate to the problem (2.5).

Lemma 2.7

For \(d+1<p<\infty\) and \(0<q\le \infty\), let \(u \in \mathring{W}^{1,\, 2}_{(p,q)}(\Omega _{T})\) be the solution of (2.5) with (2.6) and (2.7). If \(\varphi (x,t) \in L^{(p,q)}(\Omega _{T})\) and \(c(x,t) \ge 0\) in \(\Omega _{T}\), then

$$\begin{aligned} \left\| \,u\,\right\| _{L^{(p,q)}(\Omega _{T})} \le N \left\| \,\varphi \,\right\| _{L^{(p,q)}(\Omega _{T})}, \end{aligned}$$
(2.8)

where \(N=N(d,p,q,{\lambda },{\Lambda },T,|\Omega |)\).

Proof

We first use Lemma 2.6 with \(p=d+1\), and the assumptions (2.6, 2.7) to obtain

$$\begin{aligned} |u(x,t)| \le N \left\| \varphi \, \right\| _{L^{d+1}(\Omega _{T})} \end{aligned}$$

for almost all \((x,t) \in \Omega _{T}\). Furthermore, by the imbedding inequality in the Lorentz spaces (cf. [12]), for any \(p>d+1\) it yields

$$\begin{aligned} \left\| \varphi \right\| _{L^{d+1}(\Omega _{T})} \le N \, |\Omega _{T}|^{\frac{1}{d+1}-\frac{1}{p}} \left\| \varphi \right\| _{L^{(p,q)}(\Omega _{T})}, \end{aligned}$$

where \(N=N(p,q,d)\). Putting the above two estimates together leads to the required result. This completes the proof. \(\square\)

3 Proof of Theorem 1.3

This section is devoted to our main proof. We first consider the following parabolic equation

$$\begin{aligned} \partial _{t} u(x,t) + F(D^{2}u, x,t) =h(x,t) \qquad \text {in} \quad \mathbb {R}^{d+1}. \end{aligned}$$
(3.1)

Hereafter, we suppose that Assumption 1.1 holds with \(\mathbb {R}^{d+1}\) instead of \(\Omega _{T}\). To obtain the Sobolev-Lorentz estimate of (3.1) in the whole space, for any fixed point \((t_{0}, x_{0})\in \mathbb {R}^{d+1}\), we begin with establishing the following mean oscillation estimate on \(Q_{r}(t_{0},x_{0})\) for \(r>0\).

Lemma 3.1

For \(\kappa \ge 2\), \(r>0\), \(\beta \in (1, \infty )\) and \(\alpha \in (0,1)\), let \(u \in W^{1,\, 2}_{\beta (d+1)}(\mathbb {R}^{d+1})\) be the solution of (3.1). Then, there exists a small constant \(\delta _0=\delta _0(d,\lambda ,\Lambda ,\alpha ,\beta ,\gamma ,R_0)>0\) with some \(R_{0}>0\) such that the following holds: if Assumption1.1holds for any \(\delta \in (0,\delta _0)\), then we have

$$\begin{aligned}&\fint _{Q_{r}(t_{0}, x_{0})} \fint _{Q_{r}(t_{0}, x_{0})} \left| D^{2}u(x,t) - D^{2}u(s,y)\right| ^{\gamma } \, dxdtdyds \nonumber \\&\quad \le N \kappa ^{d+2} \left( \fint _{Q_{\kappa r}(t_{0}, x_{0})}|h|^{d+1} \, dxdt \right) ^{\frac{\gamma }{d+1}}\nonumber \\&\qquad + N \kappa ^{d+2} \delta ^{\frac{\gamma }{(d+1)\, \beta '}} \left( \fint _{Q_{\kappa r}(t_{0}, x_{0})} |D^{2}u|^{(d+1) \, \beta } \, dxdt \right) ^{\frac{\gamma }{(d+1)\, \beta }} \nonumber \\&\qquad + N \kappa ^{- \alpha \gamma } \left( \fint _{Q_{\kappa r}(t_{0}, x_{0})} |D^{2}u|^{d+1} \, dxdt \right) ^{\frac{\gamma }{d+1}} \nonumber \\&\qquad + N \kappa ^{d+2} R_{0}^{-2\gamma } \left( \fint _{Q_{\kappa r}(t_{0}, x_{0})} |u|^{d+1} \, dxdt\right) ^{\frac{\gamma }{d+1}} \nonumber \\&\qquad + N \kappa ^{d+2} R_{0}^{-\gamma } \left( \fint _{Q_{\kappa r}(t_{0}, x_{0})} |Du|^{d+1} \, dxdt\right) ^{\frac{\gamma }{d+1}} \end{aligned}$$
(3.2)

for some \(\gamma \in (0,1]\), where \(\beta ' = {\beta }/{(\beta -1)}\) and \(N=N(d,\delta ,\lambda ,\Lambda ,\beta )\).

Proof

The proof is divided into two folds.

Case 1. If \(0 < \kappa r \le R_{0}\), in the Assumption 1.1 we take

$$\begin{aligned} \widetilde{F}(D^{2}u) = \bar{F}_{Q_{\kappa r}(t_{0}, x_{0})} (D^{2}u). \end{aligned}$$

Consider

$$\begin{aligned} \partial _{t} u(x,t) + \widetilde{F}(D^{2}u) = \tilde{h}(x,t) \qquad \text {in} \quad \mathbb {R}^{d+1}, \end{aligned}$$
(3.3)

where

$$\begin{aligned} \tilde{h}(x,t) = h(x,t) + \widetilde{F}(D^{2}u) - F(D^{2}u, x,t). \end{aligned}$$

Let us take \(\displaystyle {\hat{u}'' = \frac{D^{2}u}{|D^{2}u|}}\), then \(\hat{u}''\in S\) and \(|\hat{u}''|=1\). With (H1) we get

$$\begin{aligned}&\fint _{Q_{\kappa r}(t_{0}, x_{0})} \left| F(D^{2}u, x,t) - \widetilde{F}(D^{2}u) \right| ^{d+1} \, dxdt \nonumber \\&\quad \le \sup _{u''\in S, \,\, |u''|=1} \fint _{Q_{\kappa r}(t_{0}, x_{0})} \left| F(u'', x,t) - \widetilde{F}(u'') \right| ^{d+1} |D^{2}u|^{d+1} \, dxdt \nonumber \\&\quad \le N \delta ^{\frac{1}{\beta '}} \left( \fint _{Q_{\kappa r}(t_{0}, x_{0})} |D^{2}u|^{(d+1)\beta } \, dxdt \right) ^{\frac{1}{\beta }}, \end{aligned}$$

where \(\delta\) is a small positive constant shown as in Assumption 1.1 and \(N=N(d, \lambda , \Lambda )\). It combined with Lemma 5.6 in [8] yields

$$\begin{aligned}&\fint _{Q_{r}(t_{0}, x_{0})} \fint _{Q_{r}(t_{0}, x_{0})} \left| D^{2}u(x,t) - D^{2}u(s,y)\right| ^{\gamma } \, dxdtdyds \nonumber \\&\quad \le N \kappa ^{d+2} \left( \fint _{Q_{\kappa r}(t_{0}, x_{0})} |h|^{d+1} \, dxdt \right) ^{\frac{\gamma }{d+1}}\nonumber \\&\qquad + N \kappa ^{d+2} \delta ^{\frac{\gamma }{(d+1)\, \beta '}} \left( \fint _{Q_{\kappa r}(t_{0}, x_{0})} |D^{2}u|^{(d+1) \, \beta } \, dxdt \right) ^{\frac{\gamma }{(d+1)\, \beta }} \nonumber \\&\qquad + N \kappa ^{- \alpha \gamma } \left( \fint _{Q_{\kappa r}(t_{0}, x_{0})} |D^{2}u|^{d+1} \, dxdt \right) ^{\frac{\gamma }{d+1}} \end{aligned}$$
(3.4)

for some \(\gamma \in (0,1]\).

Case 2. If \(\kappa r > R_{0}\), we remark that

$$\begin{aligned}&\fint _{Q_{r}(t_{0}, x_{0})} \fint _{Q_{r}(t_{0}, x_{0})} \left| D^{2}u(x,t) - D^{2}u(s,y)\right| ^{\gamma } \, dxdtdyds \nonumber \\&\quad \le 2 \fint _{Q_{r}(t_{0}, x_{0})} |D^{2}u(x,t)|^{\gamma } \, dxdt. \end{aligned}$$
(3.5)

It is a well-known fact that condition (H1) implies that there is an operator \(L_{0} = a^{ij}(x,t) D_{ij}\) with assumption (2.7) such that \(\partial _{t} u(x,t) + F(D^{2}u, x,t) = \partial _{t} u(x,t) + L_{0}u(x,t)\). Therefore, we use Corollary 4.2 in [15] with dilations and the standard cut-off argument to obtain

$$\begin{aligned} \fint _{Q_{\kappa r/2}(t_{0}, x_{0})} |D^{2}u(x,t)|^{\gamma } \, dxdt\le\, &\, {} N \Bigg ( \left( \fint _{Q_{\kappa r}(t_{0}, x_{0})} |h|^{d+1} \, dxdt\right) ^{\frac{\gamma }{d+1}} \nonumber \\&+ \frac{1}{(\kappa r)^{2\gamma }} \left( \fint _{Q_{\kappa r}(t_{0}, x_{0})} |u|^{d+1} \, dxdt\right) ^{\frac{\gamma }{d+1}} \\&+ \frac{1}{(\kappa r)^{\gamma }} \left( \fint _{Q_{\kappa r}(t_{0}, x_{0})} |Du|^{d+1} \, dxdt\right) ^{\frac{\gamma }{d+1}} \Bigg ) \end{aligned}$$

for some \(\gamma \in (0,1)\). With (3.5) and \(\kappa \ge 2\) it leads to that

$$\begin{aligned}&\fint _{Q_{r}(t_{0}, x_{0})} \fint _{Q_{r}(t_{0}, x_{0})} \left| D^{2}u(x,t) - D^{2}u(s,y)\right| ^{\gamma } \, dxdtdyds \nonumber \\&\quad \le N \kappa ^{d+2} \Bigg ( \left( \fint _{Q_{\kappa r}(t_{0}, x_{0})} |h|^{d+1} \, dxdt\right) ^{\frac{\gamma }{d+1}} \nonumber \\&\qquad + R_{0}^{-2\gamma } \left( \fint _{Q_{\kappa r}(t_{0}, x_{0})} |u|^{d+1} \, dxdt\right) ^{\frac{\gamma }{d+1}} \nonumber \\&\qquad + R_{0}^{-\gamma } \left( \fint _{Q_{\kappa r}(t_{0}, x_{0})} |Du|^{d+1} \, dxdt\right) ^{\frac{\gamma }{d+1}} \Bigg ). \end{aligned}$$
(3.6)

Putting the above two cases together completes the proof of the lemma. \(\square\)

Using Lemma 3.1 and the modified Fefferman-Stein Lemma 2.2, we get the following theorem.

Theorem 3.2

For \(p\in (d+1, \, +\infty )\) and \(q\in (0, \, +\infty ]\), let \(u \in W^{1,\, 2}_{(p,q)}(\mathbb {R}^{d+1})\) satisfying (3.1) with \(h\in L^{(p,q)}(\mathbb {R}^{d+1})\). Then there exists a small constant \(\delta _0 = \delta _0(d,\lambda ,\Lambda ,R_{0}) > 0\) with some \(R_{0}>0\) such that if Assumption 1.1holds for any \(\delta \in (0,\delta _0)\), then we have

$$\begin{aligned}&\left\| D^{2}u \right\| _{L^{(p,q)}(\mathbb {R}^{d+1})} + \left\| \partial _{t} u \right\| _{L^{(p,q)}(\mathbb {R}^{d+1})} \nonumber \\&\quad \le N \left( \left\| h \right\| _{L^{(p,q)}(\mathbb {R}^{d+1})} + \left\| u \right\| _{L^{(p,q)}(\mathbb {R}^{d+1})} \right) , \end{aligned}$$
(3.7)

where \(N=N(d,p,q,\lambda ,\Lambda ,\delta ,R_{0})\).

Proof

By Lemma 3.1 with \(\alpha \in (0,1)\), we obtain

$$\begin{aligned}&\left( D^{2}u \right) ^{\sharp }_{\gamma }(t_{0}, x_{0}) \\&\quad \le N \kappa ^{\frac{d+2}{\gamma }} \mathcal {M}^{\frac{1}{d+1}}\left( |h|^{d+1} \right) (t_{0}, x_{0})\nonumber \\&\qquad + N \kappa ^{\frac{d+2}{\gamma }} \delta ^{\frac{1}{(d+1)\, \beta '}} \mathcal {M}^{\frac{1}{(d+1)\, \beta }}\left( |D^{2}u|^{(d+1) \, \beta } \right) (t_{0}, x_{0}) \\&\qquad + N \kappa ^{- \alpha } \mathcal {M}^{\frac{1}{d+1}}\left( |D^{2}u|^{d+1} \right) (t_{0}, x_{0})\\&\qquad + N \kappa ^{\frac{d+2}{\gamma }}R_{0}^{-2} \mathcal {M}^{\frac{1}{d+1}}\left( |u|^{d+1} \right) (t_{0}, x_{0}) \\&\qquad + N \kappa ^{\frac{d+2}{\gamma }}R_{0}^{-1} \mathcal {M}^{\frac{1}{d+1}}\left( |Du|^{d+1} \right) (t_{0}, x_{0}) \end{aligned}$$

for some \(\gamma \in (0,1]\). Therefore, with the modified Fefferman-Stein Lemma 2.2 we get

$$\begin{aligned} \left\| D^{2}u \right\| _{L^{(p,q)}(\mathbb {R}^{d+1})}\le & {} N \left( \kappa ^{\frac{d+2}{\gamma }} \delta ^{\frac{1}{(d+1)\, \beta '}} + \kappa ^{- \alpha } \right) \left\| D^{2}u \right\| _{L^{(p,q)}(\mathbb {R}^{d+1})} \nonumber \\&+ N \kappa ^{\frac{d+2}{\gamma }} \left( \left\| h \right\| _{L^{(p,q)}(\mathbb {R}^{d+1})} \right. \nonumber \\&+ \left. \left\| u \right\| _{L^{(p,q)}(\mathbb {R}^{d+1})} + \left\| Du \right\| _{L^{(p,q)}(\mathbb {R}^{d+1})} \right) , \end{aligned}$$
(3.8)

where \(N=N(d,p,q,\lambda ,\Lambda ,\delta ,R_{0})\). Now, let us first take \(\kappa >0\) being big number, and then take small \(\delta >0\) such that

$$\begin{aligned} N \left( \kappa ^{\frac{d+2}{\gamma }} \delta ^{\frac{1}{(d+1)\, \beta '}} + \kappa ^{- \alpha } \right) \le \frac{1}{2}. \end{aligned}$$

Therefore, it follows from the Gagliardo-Nirenberg interpolation inequality (2.3) that

$$\begin{aligned} \left\| D^{2}u \right\| _{L^{(p,q)}(\mathbb {R}^{d+1})}\le & {} N \left( \left\| h \right\| _{L^{(p,q)}(\mathbb {R}^{d+1})} \right. \nonumber \\&\left. + \left\| u\right\| _{L^{(p,q)}(\mathbb {R}^{d+1})} \right) . \end{aligned}$$
(3.9)

On the other hand, we notice that

$$\begin{aligned} \partial _{t} u(x,t) = - F(D^{2}u, x,t) + h(x,t) \qquad \text {in} \quad \mathbb {R}^{d+1}. \end{aligned}$$

By (H1) we have

$$\begin{aligned} \left\| \partial _{t} u \right\| _{L^{(p,q)}(\mathbb {R}^{d+1})}\le & {} N \left\| D^{2}u \right\| _{L^{(p,q)}(\mathbb {R}^{d+1})} + \left\| h \right\| _{L^{(p,q)}(\mathbb {R}^{d+1})}\nonumber \\\le & {} N \left( \left\| h \right\| _{L^{(p,q)}(\mathbb {R}^{d+1})} + \left\| u \right\| _{L^{(p,q)}(\mathbb {R}^{d+1})} \right) . \end{aligned}$$
(3.10)

Combining (3.9) with (3.10) yields (3.7). \(\square\)

In the following we consider the initial-boundary problem of (3.1) over a half space. We let

$$\begin{aligned} \partial \, \mathbb {R}^{d+1}_{+} = \left\{ \, (x,t) = (x,t^{1}, \cdots ,x^{d}) \in \mathbb {R}^{d+1}: \,\, x^{1}=0 \, \right\} \end{aligned}$$

and consider

$$\begin{aligned} \left\{ \begin{array}{ll} \partial _{t} u(x,t) + F(D^{2}u, x,t) = h(x,t) &{} \quad \text {in} \quad \mathbb {R}^{d+1}_{+}, \\ u=0 &{} \quad \text {on} \quad \partial \, \mathbb {R}^{d+1}_{+}. \end{array} \right. \end{aligned}$$
(3.11)

Here, we suppose that there exists a small constant \(\delta >0\) such that Assumption 1.1 holds with \(\overline{\mathbb {R}^{d+1}_{+}}\) instead of \(\Omega _{T}\). For any \(r>0\) and \((x,t)\in \overline{\mathbb {R}^{d+1}_{+}}\), we denote

$$\begin{aligned} Q_{r}^{+}(x,t) = Q_{r}(x,t) \cap \mathbb {R}^{d+1}_{+}, \qquad Q_{r}^{+} = Q_{r}^{+}(0,0), \qquad Q_{r}^{+}(x^{1}) = Q_{r}^{+}(0,x^{1},0). \end{aligned}$$

Let us fix \((t_{0}, x_{0})\in \overline{\mathbb {R}^{d+1}_{+}}\). To attain the mean oscillation estimate on \(Q^{+}_{r}(t_{0},x_{0})\) for any \(r>0\) we mainly make use of Lemma 7.4 in [8] in terms of the odd extension technique.

Lemma 3.3

For \(\kappa \ge 2\), \(r>0\), \(\beta \in (1, \infty )\) and \(\alpha \in (0,1)\), let \(u \in \mathring{W}^{1,\, 2}_{d+1}(\mathbb {R}_{+}^{d+1})\) be the solution of (3.11) and \((t_{0}, x_{0})\in \overline{\mathbb {R}_{+}^{d+1}}\). Then there exists a small constant \(\delta _0=\delta _0(d,\lambda ,\Lambda ,\beta )>0\) such that the following holds: if Assumption 1.1holds for any \(\delta \in (0,\delta _0)\), then we have

$$\begin{aligned}&\fint _{Q^{+}_{r}(t_{0}, x_{0})} \fint _{Q^{+}_{r}(t_{0}, x_{0})} \left| D^{2}u(x,t) - D^{2}u(s,y)\right| ^{\gamma } \, dxdtdyds \\&\quad \le N \kappa ^{d+2} \left( \fint _{Q^{+}_{\kappa r}(t_{0}, x_{0})} |h|^{d+1} \, dxdt \right) ^{\frac{\gamma }{d+1}}\\&\qquad + N \kappa ^{d+2} \delta ^{\frac{\gamma }{(d+1)\, \beta '}} \left( \fint _{Q^{+}_{\kappa r}(t_{0}, x_{0})} |D^{2}u|^{(d+1) \, \beta } \, dxdt \right) ^{\frac{\gamma }{(d+1)\, \beta }} \\&\qquad + N \kappa ^{- \alpha \gamma } \left( \fint _{Q^{+}_{\kappa r}(t_{0}, x_{0})} |D^{2}u|^{d+1} \, dxdt \right) ^{\frac{\gamma }{d+1}} \\&\qquad + N \kappa ^{d+2} R_{0}^{-2\gamma } \left( \fint _{Q^{+}_{\kappa r}(t_{0}, x_{0})} |u|^{d+1} \, dxdt \right) ^{\frac{\gamma }{d+1}} \\&\qquad + N \kappa ^{d+2} R_{0}^{-\gamma } \left( \fint _{Q^{+}_{\kappa r}(t_{0}, x_{0})} |Du|^{d+1} \, dxdt \right) ^{\frac{\gamma }{d+1}} \end{aligned}$$

with some \(\gamma \in (0,1]\), where \(\beta ' = {\beta }/{(\beta -1)}\) and \(N=N(d,\delta ,\lambda ,\Lambda ,\beta )\).

If f is measurable function defined in \(\mathbb {R}^{d+1}_{+}\), then for any \((x,t) \in \overline{\mathbb {R}^{d+1}_{+}}\) and \(r>0\),

$$\begin{aligned} \fint _{Q^{+}_{r}(x,t)} |f(s,y)| \, dyds \le 2 \fint _{Q_{r}(x,t)} |f(s,y) I_{\mathbb {R}^{d+1}_{+}}| \, dyds \le N(d)\, \mathcal {M} \, (f \, I_{\mathbb {R}^{d+1}_{+}} )\,(x,t), \end{aligned}$$

where \(I_{\mathbb {R}^{d+1}_{+}}\) is the indicator function of \(\mathbb {R}^{d+1}_{+}\). On the other hand, for the smallest \(r>0\) such that \(\mathcal {C} \subset Q^{+}_{r}\), we have

$$\begin{aligned}&\left( \fint _{\mathcal {C}} \fint _{\mathcal {C}} \left| f(x,t) - f(s,y)\right| ^{\gamma } \, dxdtdyds \right) ^{\frac{1}{\gamma }} \\&\quad \le N(d,\gamma ) \left( \fint _{Q^{+}_{r}} \fint _{Q^{+}_{r}} \left| f(x,t) - f(s,y)\right| ^{\gamma } \, dxdtdyds \right) ^{\frac{1}{\gamma }}. \end{aligned}$$

With Lemma 3.3 we obtain the following result.

Corollary 3.4

Under the same hypotheses as in Lemma 3.3, we have

$$\begin{aligned}&\left( D^{2}u \right) ^{\sharp }_{\gamma }(t_{0}, x_{0}) \\&\quad \le N \kappa ^{\frac{d+2}{\gamma }} \mathcal {M}^{\frac{1}{d+1}}\left( |h|^{d+1} \, I_{\mathbb {R}^{d+1}_{+}} \right) (t_{0}, x_{0})\\&\qquad + N \kappa ^{\frac{d+2}{\gamma }} \delta ^{\frac{1}{(d+1)\, \beta '}} \mathcal {M}^{\frac{1}{(d+1)\, \beta }}\left( |D^{2}u|^{(d+1) \, \beta } \, I_{\mathbb {R}^{d+1}_{+}} \right) (t_{0}, x_{0}) \\&\qquad + N \kappa ^{- \alpha } \mathcal {M}^{\frac{1}{d+1}}\left( |D^{2}u|^{d+1} \, I_{\mathbb {R}^{d+1}_{+}} \right) (t_{0}, x_{0})\nonumber \\&\qquad + N \kappa ^{\frac{d+2}{\gamma }}R_{0}^{-2} \mathcal {M}^{\frac{1}{d+1}}\left( |u|^{d+1} I_{\mathbb {R}^{d+1}_{+}} \right) (t_{0}, x_{0}) \\&\qquad + N \kappa ^{\frac{d+2}{\gamma }}R_{0}^{-1} \mathcal {M}^{\frac{1}{d+1}}\left( |Du|^{d+1} I_{\mathbb {R}^{d+1}_{+}} \right) (t_{0}, x_{0}) \end{aligned}$$

for any \((t_{0}, x_{0})\in \overline{\mathbb {R}_{+}^{d+1}}\), where \(N=N(d,\delta ,\lambda ,\Lambda ,\beta )\).

Hence, the following regularity theorem in \(\mathring{W}^{1,\, 2}_{(p,q)}(\mathbb {R}^{d+1}_{+})\) can be deduced from Corollary 3.4 and Lemma 2.3 with the same way as the proof of Theorem 3.2.

Theorem 3.5

For \(p\in (d+1, \, +\infty )\) and \(q\in (0, \, +\infty ]\), let \(u \in \mathring{W}^{1,\, 2}_{(p,q)}(\mathbb {R}^{d+1}_{+})\) satisfying (3.11) with \(h\in L^{(p,q)}(\mathbb {R}^{d+1}_{+})\). Then there exists a small constant \(\delta _0=\delta _0(d,\lambda ,\Lambda ,R_{0})>0\) for some \(R_{0}>0\) such that if Assumption 1.1holds for any \(\delta \in (0,\delta _0)\) in \(\mathbb {R}^{d+1}_{+}\), then we have

$$\begin{aligned}&\left\| D^{2}u \right\| _{L^{(p,q)}(\mathbb {R}^{d+1}_{+})} + \left\| \partial _{t} u \right\| _{L^{(p,q)}(\mathbb {R}^{d+1}_{+})}\nonumber \\&\quad \le N \left( \left\| h \right\| _{L^{(p,q)}(\mathbb {R}^{d+1}_{+})} + \left\| u \right\| _{L^{(p,q)}(\mathbb {R}^{d+1}_{+})} \right) , \end{aligned}$$
(3.12)

where \(N=N(d,p,q,\lambda ,\Lambda ,\delta ,R_{0})\).

On the basis of Theorems 3.2 and 3.5, we now prove the main Theorem 1.3 to the Cauchy-Dirichlet problems (1.1) in \(\Omega _{T}\). Here \(\Omega\) is a bounded \(C^{1,1}\) domain and there exists a small constant \(\delta >0\) such that Assumption 1.1 holds in \(\Omega _{T}\).

Proof of Theorem 1.3

(i)  Let us define

$$\begin{aligned} h(x,t) = - G(D^{2}u, Du, u, x,t). \end{aligned}$$
(3.13)

We employ the technique of flattening the boundary and Theorems 3.2 and 3.5 which is a similar way shown as in [17]. Then, we obtain

$$\begin{aligned} \left\| D^{2}u \right\| _{L^{(p,q)}(\Omega _{T})} + \left\| \partial _{t} u \right\| _{L^{(p,q)}(\Omega _{T})} \le N \left( \left\| h \right\| _{L^{(p,q)}(\Omega _{T})} + \left\| u \right\| _{L^{(p,q)}(\Omega _{T})} \right) , \end{aligned}$$
(3.14)

where \(N=N(d,p,q,\lambda ,\Lambda ,\delta ,R_{0},T, C^{1,1} \,\text {norm of}\,\partial \Omega )\). The definition of h together with Assumption (H2) leads to that

$$\begin{aligned} \left\| h \right\| _{L^{(p,q)}(\Omega _{T})}= & {} \left\| G(D^{2}u, Du, u, x,t) \right\| _{L^{(p,q)}(\Omega _{T})} \nonumber \\\le & {} \left\| \psi \left( |D^{2}u| \right) \, |D^{2}u| \right\| _{L^{(p,q)}(\Omega _{T})} + K \, \left( \left\| Du \right\| _{L^{(p,q)}(\Omega _{T})} \right. \nonumber \\&\left. + \left\| u \right\| _{L^{(p,q)}(\Omega _{T})} \right) + \left\| g \right\| _{L^{(p,q)}(\Omega _{T})}. \end{aligned}$$
(3.15)

Recall that the function \(\psi (\rho ): \bar{\mathbb {R}}^{+} \rightarrow \bar{\mathbb {R}}^{+}\) is bounded, monotonically decreasing and \(\psi (\rho ) \rightarrow 0\) as \(\rho \rightarrow 0\). Therefore, for some \(\bar{\rho }>0\) we obtain

$$\begin{aligned}&\left\| \psi \left( |D^{2}u| \right) \, |D^{2}u| \right\| _{L^{(p,q)}(\Omega _{T})}\nonumber \\&\quad \le \psi \left( \bar{\rho } \right) \, \left\| D^{2}u \right\| _{L^{(p,q)}(\Omega _{T})} + N(p,q) \, \bar{\rho } \left\| \psi \right\| _{L^{\infty }} |\Omega _{T}|^{\frac{1}{p}}. \end{aligned}$$
(3.16)

This together with (3.15) and (3.14) yield

$$\begin{aligned}&\left\| D^{2}u \right\| _{L^{(p,q)}(\Omega _{T})} + \left\| \partial _{t} u \right\| _{L^{(p,q)}(\Omega _{T})}\\&\quad \le N \, \psi \left( \bar{\rho } \right) \, \left\| D^{2}u \right\| _{L^{(p,q)}(\Omega _{T})} + N \, \bar{\rho } \left\| \psi \right\| _{L^{\infty }} |\Omega _{T}|^{\frac{1}{p}} \\&\qquad + N \, \left( \left\| Du \right\| _{L^{(p,q)}(\Omega _{T})} + \left\| u \right\| _{L^{(p,q)}(\Omega _{T})} + \left\| g \right\| _{L^{(p,q)}(\Omega _{T})} \right) , \end{aligned}$$

where \(N=N(d,p,q,\lambda ,\Lambda ,\delta ,R_{0},T, K, C^{1,1} \,\text {norm of}\,\partial \Omega )\). Taking \(\bar{\rho }\) large enough that \(\displaystyle { N \, \psi \left( \bar{\rho } \right) \le \frac{1}{2}}\) and using again the Gagliardo-Nirenberg interpolation inequality (2.3) in the Lorentz spaces, we obtain

$$\begin{aligned}&\left\| D^{2}u \right\| _{L^{(p,q)}(\Omega _{T})} + \left\| \partial _{t} u \right\| _{L^{(p,q)}(\Omega _{T})}\nonumber \\&\quad \le N \, \left( \left\| u \right\| _{L^{(p,q)}(\Omega _{T})} + \left\| g \right\| _{L^{(p,q)}(\Omega _{T})} + \bar{\rho } \left\| \psi \right\| _{L^{\infty }} \right) , \end{aligned}$$
(3.17)

where \(N=N(d,p,q,\lambda ,\Lambda ,\delta ,R_{0},T, K, |\Omega |, C^{1,1} \,\text {norm of}\,\partial \Omega )\). Finally, to estimate the Lorentz norm of u, we rewrite the Cauchy-Dirichlet problem (1.1) as

$$\begin{aligned} \left\{ \begin{array}{ll} \partial _{t} u(x,t) + F(D^{2}u, x,t) + G(D^{2}u, Du, u, x,t) - G(D^{2}u, 0, 0, x,t) = - G(D^{2}u, 0, 0, x,t) &{} \quad \text {in} \quad \Omega _{T}, \\ u=0 &{} \quad \text {on} \quad \partial \Omega _{T}. \end{array} \right. \end{aligned}$$

Conditions (H1) and (H2) imply that there exists an operator \(L \, \cdot = \,a^{ij}\, D_{ij} \, \cdot + \,b^{i}\, D_{i} \, \cdot - \, c \, \cdot\) with (2.6) and (2.7), such that

$$\begin{aligned} \left\{ \begin{array}{ll} \partial _{t} u(x,t) + Lu = - G(D^{2}u, 0, 0, x,t) &{} \quad \text {in} \quad \Omega _{T}, \\ u=0 &{} \quad \text {on} \quad \partial \Omega _{T}. \end{array} \right. \end{aligned}$$

Since G is a monotonically nonincreasing function in u, by the Alexandrov estimate in the Lorentz spaces (cf. Lemma 2.7) we obtain

$$\begin{aligned} \left\| \,u\,\right\| _{L^{(p,q)}(\Omega _{T})} \le N \left\| \,G(D^{2}u, 0, 0, x,t)\,\right\| _{L^{(p,q)}(\Omega _{T})}, \end{aligned}$$
(3.18)

where \(N=N(d,p,q,\lambda ,\Lambda ,T,|\Omega |)\). This combined with Assumption (H2) yields

$$\begin{aligned} \left\| \,u\,\right\| _{L^{(p,q)}(\Omega _{T})} \le N \, \left( \, \left\| \psi \left( |D^{2}u| \right) \, |D^{2}u| \right\| _{L^{(p,q)}(\Omega _{T})} + \left\| g \right\| _{L^{(p,q)}(\Omega _{T})} \right) . \end{aligned}$$

Therefore, with the above estimate and (3.17), we use the same argument as in the proof of (3.16) to obtain

$$\begin{aligned} \left\| D^{2}u \right\| _{L^{(p,q)}(\Omega _{T})} + \left\| \partial _{t} u \right\| _{L^{(p,q)}(\Omega _{T})} \le N \, \left( \left\| g \right\| _{L^{(p,q)}(\Omega _{T})} + \bar{\rho } \left\| \psi \right\| _{L^{\infty }} \right) \end{aligned}$$
(3.19)

for some \(\bar{\rho }>0\), where \(N=N(d,p,q,\lambda ,\Lambda ,\delta ,R_{0},T, K, |\Omega |, \Vert \partial \Omega \Vert _{C^{1,1}})\).

(ii)   With the a prior estimate (3.19) in hand, we make use of an approximating argument to the nonlinear function H and the domain \(\Omega\), which is just as the same proof of [8, Theorem 1.2], to lead to the existence and uniqueness claims of (1.1) in the Sobolev-Lorentz spaces \(W^{1,\, 2}_{(p,q)}(\Omega _{T})\). It completes the proof. \(\square\)

Notice that parabolic Bellman’s equation is a typical prototype as fully nonlinear parabolic equations. Therefore, as a consequence of Theorem 1.3, it yields the following Sobolev-Lorentz estimate of strong solutions to parabolic Bellman’s equations under the assumption of small BMO coefficients in a bounded \(C^{1,1}\) domain. More precisely, we consider the initial-boundary problem of parabolic Bellman’s equation

$$\begin{aligned} \left\{ \begin{array}{ll} \partial _{t} u(x,t) + \underset{\omega \in \mathcal {A}}{\sup } \,\, \Big [ a^{ij}(\omega ,x,t) D_{ij}u(x,t) + b^{i}(\omega ,x,t) D_{i}u(x,t) \\ \qquad \qquad \qquad \quad - c(\omega , x,t) u(x,t) + g(\omega ,x,t) \Big ] = 0 &{} \,\,\text {in} \,\, \Omega _{T}, \\ u(x,t)=0 &{} \,\, \text {on} \,\, \partial \Omega _{T}, \end{array} \right. \end{aligned}$$
(3.20)

where the set \(\mathcal {A}\) is a separable metric space, \(a(\omega ,x,t) = \left( a^{ij}(\omega ,x,t)\right)\), \(b(\omega ,x,t)=\left( b^{i}(\omega ,x,t)\right)\), \(c(\omega , x,t)\ge 0\) and \(g(\omega ,x,t)\) are given functions. Here, we assume that these functions are measurable in (xt) for each \(\omega\), continuous in \(\omega\) for each (xt) and there exist positive constants \(0<{\lambda }\le {\Lambda }< \infty\) such that

$$\begin{aligned} |b(\omega ,x,t)| \le {\Lambda }, \qquad c(\omega , x,t)\le {\Lambda }, \qquad \hat{g}(x,t):= \underset{\omega \in \mathcal {A}}{\sup } \left[ g(\omega ,x,t)\right] < \infty \end{aligned}$$
(3.21)

and

$$\begin{aligned} {\lambda } |\xi |^{2} \le a^{ij}(\omega ,x,t) \, \xi _{i} \xi _{j} \le {\Lambda } |\xi |^{2} \end{aligned}$$
(3.22)

for any \(\xi \in \mathbb {R}^{d}\), \(\omega \in \mathcal {A}\) and almostly all \((x,t)\in \Omega _{T}\). Moreover, \(a^{ij}(\omega ,x,t)\) is \((\delta , R_{0})\)-vanishing of (xt) which means that there exists \(R_{0}>0\) such that for any \(r\in (0, R_{0}]\) and \((s,y) \in \overline{\Omega }_{T}\),

$$\begin{aligned} \sup _{i,j} \fint _{Q_{r}(s,y)} \underset{\omega \in \mathcal {A}}{\sup } \, \left| a^{ij} (\omega ,x,t) - \bar{a}^{ij}_{Q_{r}(s,y)}(\omega ) \right| \, dxdt \le \delta , \end{aligned}$$
(3.23)

where the parameter \(\delta >0\) will be specified later.

One can check that H1,H2 and Assumptions 1.1 hold in Theorem 1.3 by taking

$$\begin{aligned} F= & {} \underset{\omega \in \mathcal {A}}{\sup } \,\, a^{ij}(\omega ,x,t) D_{ij}u(x,t) \quad \text {and} \\ G= & {} \underset{\omega \in \mathcal {A}}{\sup } \,\, \Big [ b^{i}(\omega ,x,t) D_{i}u(x,t) - c(\omega , x,t) u(x,t) + g(\omega ,x,t) \Big ], \end{aligned}$$

which immediately leads to the following conclusion.

Corollary 3.6

For \(p\in (d+1, \, +\infty )\), \(q\in (0, \, +\infty ]\), \(\Omega \in C^{1,1}\), let \(u \in \mathring{W}^{1,\,2}_{(p,q)}(\Omega _{T})\) with \(T>0\) be the solution of the Cauchy-Dirichlet problems (3.20) with assumptions (3.21, 3.22) and \(\hat{g} \in L^{(p,q)}(\Omega _{T})\). Then, there exists a small constant \(\delta =\delta (d,p,q,\lambda ,\Lambda ,R_{0})>0\) for some \(R_{0}>0\) such that if the condition (3.23) holds, then we have

$$\begin{aligned} \left\| u \right\| _{W^{1,2}_{(p,q)}(\Omega _{T})} \le N \left\| \hat{g} \right\| _{L^{(p,q)}(\Omega _{T})}. \end{aligned}$$
(3.24)

Here, N is a positive constant depending only on \(d,p,q,\lambda ,\Lambda ,\delta ,R_{0},T,|\Omega |\) and \(\Vert \partial \Omega \Vert _{C^{1,1}}\).