1 Introduction

In this paper, we deal with a linear time-fractional diffusion equation in the form

$$\begin{aligned}{} & {} \partial _t^{\alpha }(u(x,t)-a(x)) = \sum _{i,j=1}^d \partial _i(a_{ij}(x)\partial _j u(x,t))\nonumber \\{} & {} \quad + \sum _{j=1}^d b_j(x,t)\partial _ju(x,t) + c(x,t)u(x,t) + F(x,t),\quad x \in \varOmega ,\, 0<t<T,\nonumber \\ \end{aligned}$$
(1.1)

where \(\partial _t^{\alpha }\) is the Caputo fractional derivative of order \(\alpha \in (0,1)\) defined on the fractional Sobolev spaces (see Section 2 for the details) and \(\varOmega \subset {\mathbb {R}}^d, \ d=1,2,3\) is a bounded domain with a smooth boundary \(\partial \varOmega \). All the functions under consideration are supposed to be real-valued.

In what follows, we always assume that the following conditions are satisfied:

$$\begin{aligned} \left\{ \begin{array}{rl} &{} a_{ij} = a_{ji} \in C^1(\overline{\varOmega }), \quad 1\le i,j \le d, \\ &{} b_j,\, c \in C^1([0,T]; C^1(\overline{\varOmega })) \cap C([0,T];C^2(\overline{\varOmega })), \quad 1\le j \le d, \\ &{} \text{ and } \text{ there } \text{ exists } \text{ a } \text{ constant }\ \kappa >0\ \text {such that}\\ &{} \sum _{i,j=1}^da_{ij}(x)\xi _i\xi _j \ge \kappa \sum _{j=1}^d \xi _j^2, \quad x\in \varOmega , \, \xi _1,\ldots , \xi _d \in {\mathbb {R}}. \end{array}\right. \end{aligned}$$
(1.2)

Using the notations \(\partial _j = \frac{\partial }{\partial x_j}\), \(j=1, 2,\ldots , d\), we define a conormal derivative \(\partial _{\nu _A}w\) with respect to the differential operator \(\sum _{i,j=1}^d\partial _j(a_{ij}\partial _i)\) by

$$\begin{aligned} \partial _{\nu _A}w(x) = \sum _{i,j=1}^da_{ij}(x)\partial _jw(x)\nu _i(x), \quad x\in \partial \varOmega , \end{aligned}$$
(1.3)

where \(\nu = \nu (x) =: (\nu _1(x),\ldots , \nu _d(x))\) is the unit outward normal vector to \(\partial \varOmega \) at the point \(x := (x_1,\ldots , x_d) \in \partial \varOmega \).

For the equation (1.1), we consider the initial-boundary value problems with the homogeneous Neumann boundary condition

$$\begin{aligned} \partial _{\nu _A}u = 0 \quad \text{ on }\ \partial \varOmega \times (0,T) \end{aligned}$$
(1.4)

or the more general homogeneous Robin boundary condition

$$\begin{aligned} \partial _{\nu _A}u + \sigma (x)u = 0 \quad \text{ on }\ \partial \varOmega \times (0,T), \end{aligned}$$
(1.5)

where \(\sigma \) is a sufficiently smooth function on \(\partial \varOmega \) that satisfies the condition \(\sigma (x) \ge 0,\ x\in \partial \varOmega \).

For partial differential equations of the parabolic type that correspond to the case \(\alpha =1\) in the equation (1.1), several important qualitative properties of solutions to the corresponding initial-boundary value problems are known. In particular, we mention a maximum principle and a comparison principle for the solutions to these problems ([28, 29]).

The main purpose of this paper is the comparison principles for the linear time-fractional diffusion equation (1.1) with the Neumann or the Robin boundary conditions.

For the equations of type (1.1) with the Dirichlet boundary conditions, the maximum principles in different formulations were derived and used in [3, 17,18,19,20,21,22,23, 33]. For a maximum principle for the time-fractional transport equations we refer to [24]. In [12], a maximum principle for more general space- and time-space-fractional partial differential equations has been derived.

Because any maximum principle involves the Dirichlet boundary values, its formulation in the case of the Neumann or Robin boundary conditions requires more cares. For this kind of the boundary conditions, both positivity of solutions and the comparison principles can be derived under some suitable restrictions on the problem data. One typical result of this sort says that the solution u to the equation (1.1) with the boundary condition (1.4) or (1.5) and an appropriately formulated initial condition is non-negative in \(\varOmega \times (0,T)\) if the initial value a and the non-homogeneous term F are non-negative in \(\varOmega \) and in \(\varOmega \times (0,T)\), respectively. Such positivity properties and their applications have been intensively discussed and used for the partial differential equations of parabolic type (\(\alpha =1\) in the equation (1.1)), see, e.g., [4, 5, 25], or [29].

However, to the best knowledge of the authors, no results of this kind have been published for the time-fractional diffusion equations in the case of the Neumann or Robin boundary conditions. The main subject of this paper is in derivation of a positivity property and the comparison principles for the linear equation (1.1) with the boundary condition (1.4) or (1.5) and an appropriately formulated initial condition. In the subsequent paper, these result will be extended to the case of the semilinear time-fractional diffusion equations. The arguments employed in these papers rely on an operator theoretical approach to the fractional integrals and derivatives in the fractional Sobolev spaces that is an extension of the theory well-known in the case \(\alpha =1\), see, e.g., [10, 26, 30]. We also refer to the recent publications [2] and [16] devoted to the comparison principles for solutions to the fractional differential inequalities with the general fractional derivatives and for solutions to the ordinary fractional differential equations, respectively.

The rest of this paper is organized as follows. In Section 2, some important results regarding the unique existence of solutions to the initial-boundary value problems for the linear time-fractional diffusion equations are presented. Section 3 is devoted to a proof of a key lemma that is a basis for the proofs of the comparison principles for the linear and semilinear time-fractional diffusion equations. The lemma asserts that each solution to (1.1) is non-negative in \(\varOmega \times (0,T)\) if \(a\ge 0\) and \(F \ge 0\), provided that u is assumed to satisfy some extra regularity. In Section 4, we prove a comparison principle that is our main result for the problem (1.1) for the linear time-fractional diffusion equation. Moreover, we establish the order-preserving properties for other problem data (the zeroth-order coefficient c of the equation and the coefficient \(\sigma \) of the Robin condition). Finally, a detailed proof of an important auxiliary statement is presented in an Appendix.

2 Well-posedness results

For \(x \in \varOmega , 0<t<T\), we define an operator

$$\begin{aligned} -Av(x,t) := \sum _{i,j=1}^d \partial _i(a_{ij}(x)\partial _j v(x,t) + \sum _{j=1}^d b_j(x,t)\partial _jv(x,t) + c(x,t)v(x,t)\nonumber \\ \end{aligned}$$
(2.1)

and assume that the conditions (1.2) for the coefficients \(a_{ij}, b_j, c\) are satisfied.

In this section, we deal with the following initial-boundary value problem for the linear time-fractional diffusion equation (1.1) with the time-fractional derivative of order \(\alpha \in (0,1)\)

$$\begin{aligned} \left\{ \begin{array}{rl} &{} \partial _t^{\alpha }(u(x,t)-a(x)) + Au(x,t) = F(x,t), \quad x \in \varOmega , 0<t<T, \\ &{} \partial _{\nu _A}u + \sigma (x)u(x,t) = 0, \quad x\in \partial \varOmega , \, 0<t<T, \end{array}\right. \end{aligned}$$
(2.2)

along with the initial condition (2.3) formulated below.

To appropriately define the Caputo fractional derivative \(d_t^{\alpha }w(t)\), \(0<\alpha <1\), we start with its definition on the space

$$\begin{aligned} {_{0}C^1[0,T]}:= \{ u \in C^1[0,T]; u(0) = 0\} \end{aligned}$$

that reads as follows:

$$\begin{aligned} d_t^{\alpha }w(t) = \frac{1}{\varGamma (1-\alpha )}\int ^t_0 (t-s)^{-\alpha }\frac{dw}{ds}(s) ds,\ w\in {_{0}C^1[0,T]}. \end{aligned}$$

Then we extend this operator from the domain \({\mathcal {D}}(d_t^{\alpha }) := {_{0}C^1[0,T]}\) to \(L^2(0,T)\) taking into account its closability ( [32]). As have been shown in [13], there exists a unique minimum closed extension of \(d_t^{\alpha }\) with the domain \({\mathcal {D}}(d_t^{\alpha }) = {_{0}C^1[0,T]}\). Moreover, the domain of this extension is the closure of \({_{0}C^1[0,T]}\) in the Sobolev-Slobodeckij space \(H^{\alpha }(0,T)\). Let us recall that the norm \(\Vert \cdot \Vert _{H^{\alpha } (0,T)}\) of the Sobolev-Slobodeckij space \(H^{\alpha }(0,T)\) is defined as follows ( [1]):

$$\begin{aligned} \Vert v\Vert _{H^{\alpha }(0,T)}:= \left( \Vert v\Vert ^2_{L^2(0,T)} + \int ^T_0\int ^T_0 \frac{\vert v(t)-v(s)\vert ^2}{\vert t-s\vert ^{1+2\alpha }} dtds \right) ^{\frac{1}{2}}. \end{aligned}$$

By setting

$$\begin{aligned} H_{\alpha }(0,T):= \overline{{_{0}C^1[0,T]}}^{H^{\alpha }(0,T)}, \end{aligned}$$

we obtain ( [13])

$$\begin{aligned} H_{\alpha }(0,T) = \left\{ \begin{array}{rl} &{}H^{\alpha }(0,T), \quad 0<\alpha<\frac{1}{2}, \\ &{}\left\{ v \in H^{\frac{1}{2}}(0,T);\, \int ^T_0 \frac{\vert v(t)\vert ^2}{t} dt< \infty \right\} , \quad \alpha =\frac{1}{2}, \\ &{} \{ v \in H^{\alpha }(0,T);\, v(0) = 0\}, \quad \frac{1}{2}< \alpha < 1, \end{array}\right. \end{aligned}$$

and

$$\begin{aligned} \Vert v\Vert _{H_{\alpha }(0,T)} = \left\{ \begin{array}{rl} &{}\Vert v\Vert _{H^{\alpha }(0,T)}, \quad \alpha \ne \frac{1}{2}, \\ &{}\left( \Vert v\Vert _{H^{\frac{1}{2}}(0,T)}^2 + \int ^T_0 \frac{\vert v(t)\vert ^2}{t}dt\right) ^{\frac{1}{2}}, \quad \alpha =\frac{1}{2}. \end{array}\right. \end{aligned}$$

In what follows, we also use the Riemann-Liouville fractional integral operator \(J^{\beta }\), \(\beta > 0\) defined by

$$\begin{aligned} (J^{\beta }f)(t) := \frac{1}{\varGamma (\beta )}\int ^t_0 (t-s)^{\beta -1}f(s) ds, \quad 0<t<T. \end{aligned}$$

Then, according to [9] and [13],

$$\begin{aligned} H_{\alpha }(0,T) = J^{\alpha }L^2(0,T),\quad 0<\alpha <1. \end{aligned}$$

Next we define

$$\begin{aligned} \partial _t^{\alpha }= (J^{\alpha })^{-1} \quad \text{ with }\ {\mathcal {D}}(\partial _t^{\alpha }) = H_{\alpha }(0,T). \end{aligned}$$

As have been shown in [9] and [13], there exists a constant \(C>0\) depending only on \(\alpha \) such that

$$\begin{aligned} C^{-1}\Vert v\Vert _{H_{\alpha }(0,T)} \le \Vert \partial _t^{\alpha }v\Vert _{L^2(0,T)} \le C\Vert v\Vert _{H_{\alpha }(0,T)} \quad \text{ for } \text{ all } \ v\in H_{\alpha }(0,T). \end{aligned}$$

Now we can introduce a suitable form of initial condition for the problem (2.2) as follows

$$\begin{aligned} u(x, \cdot ) - a(x) \in H_{\alpha }(0,T) \quad \text{ for } \text{ almost } \text{ all } x\in \varOmega \end{aligned}$$
(2.3)

and write down a complete formulation of an initial-boundary value problem for the linear time-fractional diffusion equation (1.1):

$$\begin{aligned} \left\{ \begin{array}{rl} &{} \partial _t^{\alpha }(u(x,t)-a(x)) + Au(x,t) = F(x,t), \quad x \in \varOmega , 0<t<T, \\ &{} \partial _{\nu _A}u(x,t) + \sigma (x)u(x,t) = 0, \quad x\in \partial \varOmega , \, 0<t<T,\\ &{} u(x, \cdot ) - a(x) \in H_{\alpha }(0,T) \quad \text{ for } \text{ almost } \text{ all } x\in \varOmega . \end{array}\right. \end{aligned}$$
(2.4)

It is worth mentioning that the term \(\partial _t^{\alpha }(u(x,t) - a(x))\) in the first line of (2.4) is well-defined due to inclusion formulated in the third line of (2.4). In particular, for \(\frac{1}{2}< \alpha < 1\), the Sobolev embedding leads to the inclusions \(H_{\alpha }(0,T) \subset H^{\alpha }(0,T)\subset C[0,T]\). This means that \(u\in H_{\alpha }(0,T;L^2(\varOmega ))\) implies \(u \in C([0,T];L^2(\varOmega ))\) and thus in this case the initial condition can be formulated as \(u(\cdot ,0) = a\) in \(L^2\)-sense. Moreover, for sufficiently smooth functions a and F, the solution to (2.4) can be proved to satisfy the initial condition in a usual sense: \(\lim _{t\rightarrow 0} u(\cdot ,t) = a\) in \(L^2(\varOmega )\) (see Lemma 4 in Section 4). Consequently, the third line of (2.4) can be interpreted as a generalized initial condition.

In the following theorem, a fundamental result regarding the unique existence of the solution to the initial-boundary value problem (2.4) is presented.

Theorem 1

For \(a\in H^1(\varOmega )\) and \(F \in L^2(0,T;L^2(\varOmega ))\), there exists a unique solution \(u(F,a) = u(F,a)(x,t) \in L^2(0,T;H^2(\varOmega ))\) to the initial-boundary value problem (2.4) such that \(u(F,a)-a \in H_{\alpha }(0,T;L^2(\varOmega ))\).

Moreover, there exists a constant \(C>0\) such that

$$\begin{aligned}&\Vert u(F,a)-a\Vert _{H_{\alpha }(0,T;L^2(\varOmega ))} + \Vert u(F,a)\Vert _{L^2(0,T;H^2(\varOmega ))} \\&\quad \le C(\Vert a\Vert _{H^1(\varOmega )} + \Vert F\Vert _{L^2(0,T;L^2(\varOmega ))}). \end{aligned}$$

Before starting with the proof of Theorem 1, we introduce some notations and derive several helpful results needed for the proof.

For an arbitrary constant \(c_0>0\), we define an elliptic operator \(A_0\) as follows:

$$\begin{aligned} \left\{ \begin{array}{rl} &{} (-A_0v)(x) := \sum _{i,j=1}^d\partial _i(a_{ij}(x)\partial _jv(x)) - c_0v(x), \quad x\in \varOmega , \\ &{} {\mathcal {D}}(A_0) = \left\{ v\in H^2(\varOmega );\, \partial _{\nu _A}v + \sigma v = 0 \quad \text{ on } \partial \varOmega \right\} . \end{array}\right. \end{aligned}$$
(2.5)

We recall that in the definition (2.5), \(\sigma \) is a smooth function, the inequality \(\sigma (x)\ge 0,\ x\in \partial \varOmega \) holds true, and the coefficients \(a_{ij}\) satisfy the conditions (1.2).

Henceforth, by \(\Vert \cdot \Vert \) and \((\cdot ,\cdot )\) we denote the standard norm and the scalar product in \(L^2(\varOmega )\), respectively. It is well-known that the operator \(A_0\) is self-adjoint and its resolvent is a compact operator. Moreover, for a sufficiently large constant \(c_0>0\), by Lemma 6 in Section 1, we can verify that \(A_0\) is positive definite. Therefore, by choosing the constant \(c_0>0\) large enough, the spectrum of \(A_0\) consists entirely of discrete positive eigenvalues \(0 < \lambda _1 \le \lambda _2 \le \cdots \), which are numbered according to their multiplicities and \(\lambda _n \rightarrow \infty \) as \(n\rightarrow \infty \). Let \(\varphi _n\) be an eigenvector corresponding to the eigenvalue \(\lambda _n\) such that \(A\varphi _n = \lambda _n\varphi _n\) and \((\varphi _n, \varphi _m) = 0\) if \(n \ne m\) and \((\varphi _n,\varphi _n) = 1\). Then the system \(\{ \varphi _n\}_{n\in {\mathbb {N}}}\) of the eigenvectors forms an orthonormal basis in \(L^2(\varOmega )\) and for any \(\gamma \ge 0\) we can define the fractional powers \(A_0^{\gamma }\) of the operator \(A_0\) by the following relation (see, e.g., [26]):

$$\begin{aligned} A_0^{\gamma }v = \sum _{n=1}^{\infty } \lambda _n^{\gamma } (v,\varphi _n)\varphi _n, \end{aligned}$$

where

$$\begin{aligned} v \in {\mathcal {D}}(A_0^{\gamma }) := \left\{ v\in L^2(\varOmega ): \sum _{n=1}^{\infty } \lambda _n^{2\gamma } (v,\varphi _n)^2 < \infty \right\} \end{aligned}$$

and

$$\begin{aligned} \Vert A_0^{\gamma }v\Vert = \left( \sum _{n=1}^{\infty } \lambda _n^{2\gamma } (v,\varphi _n)^2 \right) ^{\frac{1}{2}}. \end{aligned}$$

We note that \({\mathcal {D}}(A_0^{\gamma }) \subset H^{2\gamma }(\varOmega )\).

Our proof of Theorem 1 is similar to the one presented in [9, 13] for the case of the homogeneous Dirichlet boundary condition. In particular, we employ the operators S(t) and K(t) defined by ( [9, 13])

$$\begin{aligned} S(t)a = \sum _{n=1}^{\infty } E_{\alpha ,1}(-\lambda _n t^{\alpha }) (a,\varphi _n)\varphi _n, \quad a\in L^2(\varOmega ), t>0 \end{aligned}$$
(2.6)

and

$$\begin{aligned} K(t)a = -A_0^{-1}S'(t)a = \sum _{n=1}^{\infty } t^{\alpha -1}E_{\alpha ,\alpha }(-\lambda _n t^{\alpha }) (a,\varphi _n)\varphi _n, \quad a\in L^2(\varOmega ), t>0.\nonumber \\ \end{aligned}$$
(2.7)

In the above formulas, \(E_{\alpha ,\beta }(z)\) denotes the Mittag-Leffler function defined by a convergent series as follows:

$$\begin{aligned} E_{\alpha ,\beta }(z) = \sum _{k=0}^\infty \frac{z^k}{\varGamma (\alpha \, k + \beta )}, \ \alpha >0,\ \beta \in {\mathbb {C}},\ z \in {\mathbb {C}}. \end{aligned}$$

It follows directly from the definitions given above that \(A_0^{\gamma }K(t)a = K(t)A_0^{\gamma }a\) and \(A_0^{\gamma }S(t)a = S(t)A_0^{\gamma }a\) for \(a \in {\mathcal {D}} (A_0^{\gamma })\). Moreover, the inequality (see, e.g., Theorem 1.6 (p. 35) in [27])

$$\begin{aligned} \max \{ \vert E_{\alpha ,1}(-\lambda _nt^{\alpha })\vert , \, \vert E_{\alpha ,\alpha }(-\lambda _nt^{\alpha })\vert \} \le \frac{C}{1+\lambda _nt^{\alpha }} \quad \hbox { for all}\ t>0 \end{aligned}$$

implicates the estimations ( [9])

$$\begin{aligned} \left\{ \begin{array}{l} \Vert A_0^{\gamma }S(t)a\Vert \le Ct^{-\alpha \gamma }\Vert a\Vert , \\ \Vert A_0^{\gamma }K(t)a\Vert \le Ct^{\alpha (1-\gamma )-1} \Vert a\Vert , \quad a \in L^2(\varOmega ), t > 0, 0 \le \gamma \le 1. \end{array}\right. \end{aligned}$$
(2.8)

In order to shorten the notations and to focus on the dependence on the time variable t, henceforth we sometimes omit the variable x in the functions of two variables x and t and write, say, u(t) instead of \(u(\cdot ,t)\).

Due to the inequalities (2.8), the estimations provided in the formulation of Theorem 1 can be derived as in the case of the fractional powers of generators of the analytic semigroups ( [10]). To do this, we first formulate and prove the following lemma:

Lemma 1

Under the conditions formulated above, the following estimates hold true for \(F\in L^2(0,T;L^2(\varOmega ))\) and \(a \in L^2(\varOmega )\):

  1. (i)
    $$\begin{aligned} \left\| \int ^t_0 A_0K(t-s)F(s) ds \right\| _{L^2(0,T;L^2(\varOmega ))} \le C\Vert F\Vert _{L^2(0,T;L^2(\varOmega ))}, \end{aligned}$$
  2. (ii)
    $$\begin{aligned} \left\| \int ^t_0 K(t-s)F(s) ds \right\| _{H_{\alpha }(0,T;L^2(\varOmega ))} \le C\Vert F\Vert _{L^2(0,T;L^2(\varOmega ))}, \end{aligned}$$
  3. (iii)
    $$\begin{aligned} \Vert S(t)a - a\Vert _{H_{\alpha }(0,T;L^2(\varOmega ))} + \Vert S(t)a\Vert _{L^2(0,T;H^2(\varOmega ))} \le C\Vert a\Vert . \end{aligned}$$

Proof

We start with proving the estimate (i). By (2.7), we have

$$\begin{aligned}&\left\| \int ^t_0 A_0 K(t-s)F(s) ds \right\| ^2\\&\quad = \left\| \sum _{n=1}^{\infty }\left( \int ^t_0 \lambda _n(t-s)^{\alpha -1} E_{\alpha ,\alpha }(-\lambda _n(t-s)^{\alpha }) (F(s), \varphi _n) ds\right) \varphi _n\right\| ^2\\&\quad = \sum _{n=1}^{\infty }\left| \int ^t_0 \lambda _n(t-s)^{\alpha -1}E_{\alpha ,\alpha }(-\lambda _n(t-s)^{\alpha }) (F(s), \varphi _n) ds \right| ^2. \end{aligned}$$

Therefore, using the Parseval equality and the Young inequality for the convolution, we obtain

$$\begin{aligned}&\left\| \int ^t_0 A_0K(t-s)F(s) ds \right\| ^2_{L^2(0,T;L^2(\varOmega ))}\\&\quad = \sum _{n=1}^{\infty }\int ^T_0 \vert (\lambda _ns^{\alpha -1}E_{\alpha ,\alpha }(-\lambda _ns^{\alpha }) \, *\, (F(s), \varphi _n) \vert ^2 ds\\&\quad = \sum _{n=1}^{\infty }\Vert \lambda _ns^{\alpha -1}E_{\alpha ,\alpha }(-\lambda _ns^{\alpha }) \, * \, (F(s), \varphi _n) \Vert ^2_{L^2(0,T)}\\&\quad \le \sum _{n=1}^{\infty }\left( \lambda _n\int ^t_0 \vert t^{\alpha -1}E_{\alpha ,\alpha }(-\lambda _nt^{\alpha }) \vert dt \right) ^2 \Vert (F(t),\varphi _n)\Vert ^2_{L^2(0,T)}. \end{aligned}$$

Then we employ the representation

$$\begin{aligned} \frac{d}{dt}E_{\alpha ,1}(-\lambda _nt^{\alpha }) = -\lambda _nt^{\alpha -1}E_{\alpha ,\alpha }(-\lambda _nt^{\alpha }), \end{aligned}$$
(2.9)

and the complete monotonicity of the Mittag-Leffler function ( [8])

$$\begin{aligned} E_{\alpha ,1}(-\lambda _nt^{\alpha }) > 0, \quad \frac{d}{dt}E_{\alpha ,1}(-\lambda _nt^{\alpha }) \le 0, \quad t\ge 0, \quad 0<\alpha \le 1 \end{aligned}$$

to get the inequality

$$\begin{aligned}{} & {} \int ^T_0 \vert \lambda _nt^{\alpha -1}E_{\alpha ,\alpha }(-\lambda _nt^{\alpha })\vert dt = -\int ^T_0 \frac{d}{dt}E_{\alpha ,1}(-\lambda _nt^{\alpha })dt\nonumber \\{} & {} \quad = 1 - E_{\alpha ,1}(-\lambda _nT^{\alpha }) \le 1 \quad \hbox { for all}\ n\in {\mathbb {N}}. \end{aligned}$$
(2.10)

Hence,

$$\begin{aligned}&\left\| \int ^t_0 A_0K(t-s)F(s) ds \right\| ^2_{L^2(0,T;L^2(\varOmega ))} \le \sum _{n=1}^{\infty }\Vert (F(t), \varphi _n)\Vert ^2_{L^2(0,T)}\\&\quad = \int ^T_0 \sum _{n=1}^{\infty }\vert (F(t), \varphi _n) \vert ^2 dt = \int ^T_0 \Vert F(\cdot ,t)\Vert ^2 dt = \Vert F\Vert _{L^2(0,T;L^2(\varOmega ))}^2. \end{aligned}$$

Now we proceed with proving the estimate (ii). For \(0<t<T, \, n\in {\mathbb {N}}\) and \(f\in L^2(0,T)\), we set

$$\begin{aligned} (L_nf)(t) := \int ^t_0 (t-s)^{\alpha -1}E_{\alpha ,\alpha }(-\lambda _n(t-s)^{\alpha }) f(s) ds. \end{aligned}$$

Then

$$\begin{aligned} \int ^t_0 K(t-s)F(s) ds = \sum _{n=1}^{\infty }(L_nf)(t)\varphi _n \end{aligned}$$

in \(L^2(\varOmega )\) for any fixed \(t \in [0,T]\).

First we prove that

$$\begin{aligned} \left\{ \begin{array}{rl} &{} L_nf \in H_{\alpha }(0,T), \\ &{} \partial _t^{\alpha }(L_nf)(t) = -\lambda _nL_nf(t) + f(t), \quad 0<t<T, \\ &{} \Vert L_nf\Vert _{H_{\alpha }(0,T)} \le C\Vert f\Vert _{L^2(0,T)}, \quad n\in {\mathbb {N}}\quad \text{ for } \text{ each } f\in L^2(0,T). \end{array}\right. \end{aligned}$$
(2.11)

In order to prove this, we apply the Riemann-Liouville fractional integral \(J^{\alpha }\) to \(L_nf\) and get the representation

$$\begin{aligned}&J^{\alpha }(L_nf)(t) = \frac{1}{\varGamma (\alpha )}\int ^t_0 (t-s)^{\alpha -1}(L_nf)(s) ds\\&\quad = \frac{1}{\varGamma (\alpha )} \int ^t_0 (t-s)^{\alpha -1} \left( \int ^s_0 (s-\xi )^{\alpha -1}E_{\alpha ,\alpha }(-\lambda _n (s-\xi )^{\alpha })f(\xi ) d\xi \right) ds\\&\quad = \frac{1}{\varGamma (\alpha )}\int ^t_0 f(\xi ) \left( \int ^t_{\xi } (t-s)^{\alpha -1}(s-\xi )^{\alpha -1}E_{\alpha ,\alpha }(-\lambda _n (s-\xi )^{\alpha }) ds \right) d\xi . \end{aligned}$$

By direct calculations, using (2.9), we obtain the formula

$$\begin{aligned}&\frac{1}{\varGamma (\alpha )}\int ^t_{\xi } (t-s)^{\alpha -1}(s-\xi )^{\alpha -1} E_{\alpha ,\alpha }(-\lambda _n (s-\xi )^{\alpha }) ds\\&\quad = -\frac{1}{\lambda _n}(t-\xi )^{\alpha -1}\left( E_{\alpha ,\alpha }(-\lambda _n t^{\alpha }) - \frac{1}{\varGamma (\alpha )}\right) . \end{aligned}$$

Therefore, we have the relation

$$\begin{aligned}&J^{\alpha }(L_nf)(t) = -\frac{1}{\lambda _n}(L_nf)(t) + \frac{1}{\lambda _n}\int ^t_0 (t-\xi )^{\alpha -1}\frac{1}{\varGamma (\alpha )} f(\xi ) d\xi \\&\quad = -\frac{1}{\lambda _n}(L_nf)(t) + \frac{1}{\lambda _n}(J^{\alpha }f)(t), \quad n\in {\mathbb {N}}, \end{aligned}$$

that is,

$$\begin{aligned} (L_nf)(t) = -\lambda _n J^{\alpha }(L_nf)(t) + (J^{\alpha }f)(t), \quad 0<t<T. \end{aligned}$$

Hence, \(L_nf \in H_{\alpha }(0,T) = J^{\alpha }L^2(0,T)\). By definition, \(\partial _t^{\alpha }= (J^{\alpha })^{-1}\) ( [13]) and thus the last formula can be rewritten in the form

$$\begin{aligned} \partial _t^{\alpha }(L_nf) = -\lambda _n L_nf + f \quad \text{ in } (0,T). \end{aligned}$$

Using the inequality (2.10), we obtain

$$\begin{aligned} \lambda _n\Vert L_nf\Vert _{L^2(0,T)} \le \lambda _n\Vert s^{\alpha -1} E_{\alpha ,\alpha }(-\lambda _ns^{\alpha })\Vert _{L^1(0,T)}\Vert f\Vert _{L^2(0,T)} \le \Vert f\Vert _{L^2(0,T)}. \end{aligned}$$

Therefore,

$$\begin{aligned}&\Vert L_nf\Vert _{H_{\alpha }(0,T)} \le C\Vert \partial _t^{\alpha }(L_nf)\Vert _{L^2(0,T)} \le C(\Vert -\lambda _nL_nf\Vert _{L^2(0,T)} + \Vert f\Vert _{L^2(0,T)})\\&\quad \le C\Vert f\Vert _{L^2(0,T)}, \quad n\in {\mathbb {N}}, \quad f\in L^2(0,T). \end{aligned}$$

Thus, the estimate (2.11) is proved.

Now we set \(f_n(s) := (F(s), \, \varphi _n)\) for \(0<s<T\) and \(n\in {\mathbb {N}}\). Since

$$\begin{aligned} \partial _t^{\alpha }\int ^t_0 K(t-s)F(s) ds = \sum _{n=1}^{\infty }\partial _t^{\alpha }(L_nf_n)(t)\varphi _n, \end{aligned}$$

we obtain

$$\begin{aligned} \left\| \partial _t^{\alpha }\int ^t_0 K(t-s)F(s) ds\right\| ^2_{L^2(\varOmega )} = \sum _{n=1}^{\infty }\vert \partial _t^{\alpha }(L_nf_n)(t)\vert ^2. \end{aligned}$$

By applying (2.11), we get the following chain of inequalities and equations:

$$\begin{aligned}&\left\| \partial _t^{\alpha }\int ^t_0 K(t-s)F(s) ds\right\| ^2 _{H_{\alpha }(0,T;L^2(\varOmega ))} \le C\left\| \partial _t^{\alpha }\int ^t_0 K(t-s)F(s) ds\right\| ^2 _{L^2(0,T;L^2(\varOmega ))}\\&\quad = C\sum _{n=1}^{\infty }\Vert \partial _t^{\alpha }(L_nf_n)\Vert ^2_{L^2(0,T)} \le C\sum _{n=1}^{\infty }\Vert L_nf_n\Vert ^2_{H_{\alpha }(0,T)} \le C\sum _{n=1}^{\infty }\Vert f_n\Vert ^2_{L^2(0,T)}\\&\quad = C\int ^T_0 \sum _{n=1}^{\infty }\vert (F(s),\varphi _n) \vert ^2 ds = C\int ^T_0 \Vert F(s)\Vert _{L^2(\varOmega )}^2 ds = C\Vert F\Vert ^2_{L^2(0,T;L^2(\varOmega ))}. \end{aligned}$$

Thus, the proof of the estimate (ii) is completed.

The estimate (iii) from Lemma 1 follows from the standard estimates of the operator S(t). It can be derived by the same arguments as those that were employed in Section 6 of Chapter 4 in [13] for the case of the homogeneous Dirichlet boundary condition and we omit here the technical details. \(\square \)

Now we proceed to the proof of Theorem 1.

Proof

In the first line of the problem (2.4), we regard the expressions \(\sum _{j=1}^d b_j(x,t)\partial _ju\) and c(xt)u as some non-homogeneous terms. Then this problem can be rewritten in terms of the operator \(A_0\) as follows

$$\begin{aligned} \left\{ \begin{array}{rl} &{} \partial _t^{\alpha }(u-a) + A_0u(x,t) = F(x,t)\\ +&{} \sum _{j=1}^d b_j(x,t)\partial _ju + (c_0+c(x,t))u, \quad x\in \varOmega ,\, 0<t<T,\\ &{} \partial _{\nu _A}u + \sigma (x) u = 0 \quad \text{ on } \partial \varOmega \times (0,T),\\ &{} u(x,\cdot ) - a(x) \in H_{\alpha }(0,T) \quad \text{ for } \text{ almost } \text{ all } x\in \varOmega . \end{array}\right. \end{aligned}$$
(2.12)

In its turn, the first line of (2.12) can be represented in the form ( [9, 13])

$$\begin{aligned}{} & {} u(t) = S(t)a + \int ^t_0 K(t-s)F(s) ds\nonumber \\{} & {} \quad + \int ^t_0 K(t-s) \left( \sum _{j=1}^d b_j(s)\partial _ju(s) + (c_0+c(s))u(s) \right) ds, \quad 0<t<T.\nonumber \\ \end{aligned}$$
(2.13)

Moreover, it is known that if \(u\in L^2(0,T;H^2(\varOmega ))\) satisfies the initial condition \(u-a \in H_{\alpha }(0,T;L^2(\varOmega ))\) and the equation (2.13), then u is a solution to the problem (2.12). With the notations

$$\begin{aligned} \left\{ \begin{array}{rl} &{} G(t):= \int ^t_0 K(t-s)F(s) ds + S(t)a, \\ &{} Qv(t) = Q(t)v(t) := \sum _{j=1}^d b_j(\cdot ,t)\partial _jv(t) + (c_0+c(\cdot ,t))v(t), \\ &{} Rv(t):= \int ^t_0 K(t-s) \left( \sum _{j=1}^d b_j(\cdot ,s)\partial _jv(s) + (c_0+c(\cdot ,s))v(s) \right) ds,\\ &{}\qquad \qquad \qquad \hbox { for}\ 0<t<T, \end{array}\right. \end{aligned}$$
(2.14)

the equation (2.13) can be represented in form of a fixed point equation \(u = Ru + G\) on the space \(L^2(0,T;H^2(\varOmega ))\).

Lemma 1 yields the inclusion \(G \in L^2(0,T;H^2(\varOmega ))\). Moreover, since \(\Vert A_0^{\frac{1}{2}}a\Vert \le C\Vert a\Vert _{H^1(\varOmega )}\) and \({\mathcal {D}}(A_0^{\frac{1}{2}}) = H^1(\varOmega )\) (see, e.g., [6]), the estimate (2.8) implies

$$\begin{aligned} \Vert S(t)a\Vert _{H^2(\varOmega )} \le C\Vert A_0S(t)a\Vert = C\Vert A_0^{\frac{1}{2}}S(t)A_0^{\frac{1}{2}}a\Vert \le Ct^{-\frac{1}{2}\alpha }\Vert a\Vert _{H^1(\varOmega )} \end{aligned}$$

and thus

$$\begin{aligned} \Vert S(t)a\Vert ^2_{L^2(0,T;H^2(\varOmega ))} \le C\left( \int ^T_0 t^{-\alpha }dt \right) \Vert a\Vert ^2_{H^1(\varOmega )} \le \frac{CT^{1-\alpha }}{1-\alpha }\Vert a\Vert ^2_{H^1(\varOmega )}. \end{aligned}$$

Consequently, the inclusion \(S(t)a \in L^2(0,T;H^2(\varOmega ))\) holds valid.

For \(0<t<T\), we next estimate \(\Vert Rv(\cdot ,t)\Vert _{H^2(\varOmega )}\) for \(v(\cdot ,t) \in {\mathcal {D}}(A_0)\) as follows:

$$\begin{aligned}&\Vert Rv(\cdot ,t)\Vert _{H^2(\varOmega )} \le C\Vert A_0Rv(\cdot ,t)\Vert _{L^2(\varOmega )}\\&\quad \le \int ^t_0 \left\| A_0^{\frac{1}{2}}K(t-s)A_0^{\frac{1}{2}} \left( \sum _{j=1}^d b_j(s)\partial _jv(s) + (c_0+c(s))v(s) \right) \right\| ds\\&\quad \le C\int ^t_0 \Vert A_0^{\frac{1}{2}}K(t-s)\Vert \left\| A_0^{\frac{1}{2}} \left( \sum _{j=1}^d b_j(s)\partial _jv(s) + (c_0+c(s))v(s) \right) \right\| ds\\&\quad \le C\int ^t_0 (t-s)^{\frac{1}{2}\alpha -1}\Vert v(s)\Vert _{H^2(\varOmega )}ds = C\left( \varGamma \left( \frac{1}{2}\alpha \right) J^{\frac{1}{2}\alpha }\Vert v\Vert _{H^2(\varOmega )}\right) (t). \end{aligned}$$

For derivation of this estimate, we employed the inequalities

$$\begin{aligned} \Vert A_0^{\frac{1}{2}}b_j(s)\partial _jv(t)\Vert \le C\Vert b_j(s)\partial _jv(s)\Vert _{H^1(\varOmega )} \le C\Vert v(s)\Vert _{H^2(\varOmega )} \end{aligned}$$

and \(\Vert (c(s)+c_0)v(s)\Vert _{H^1(\varOmega )} \le C\Vert v(s)\Vert _{H^2(\varOmega )}\) that are valid because of the inclusions \(b_j \in C^1(\overline{\varOmega }\times [0,T])\)) and \(c+c_0\in C([0,T];C^1(\overline{\varOmega }))\).

Since \((J^{\frac{1}{2}\alpha }w_1)(t) \ge (J^{\frac{1}{2}\alpha }w_2)(t)\) if \(w_1(t) \ge w_2(t)\) for \(0\le t\le T\), and \(J^{\frac{1}{2}\alpha }J^{\frac{1}{2}\alpha }w = J^{\alpha }w\) for \(w_1, w_2, w \in L^2(0,T)\), we have

$$\begin{aligned}&\Vert R^2v(t)\Vert _{H^2(\varOmega )} = \Vert R(Rv)(t)\Vert _{H^2(\varOmega )}\\&\quad \le C\left( \varGamma \left( \frac{1}{2}\alpha \right) J^{\frac{1}{2}\alpha } \left( C\varGamma \left( \frac{1}{2}\alpha \right) J^{\frac{1}{2}\alpha } \Vert v\Vert _{H^2(\varOmega )}\right) \right) (t)\\&\quad = \left( C\varGamma \left( \frac{1}{2}\alpha \right) \right) ^2 (J^{\alpha }\Vert v\Vert _{H^2(\varOmega )})(t). \end{aligned}$$

Repeating this argumentation m-times, we obtain

$$\begin{aligned}&\Vert R^mv(t)\Vert _{H^2(\varOmega )} \le \left( C\varGamma \left( \frac{1}{2}\alpha \right) \right) ^m \left( J^{\frac{1}{2}\alpha m}\Vert v\Vert _{H^2(\varOmega )}\right) (t)\\&\quad \le \frac{\left( C\varGamma \left( \frac{1}{2}\alpha \right) \right) ^m}{\varGamma \left( \frac{1}{2}\alpha m\right) } \int ^t_0 (t-s)^{\frac{m}{2}\alpha -1} \Vert v(\xi )\Vert _{H^2(\varOmega )}ds, \quad 0<t<T. \end{aligned}$$

Applying the Young inequality to the integral at the right-hand side of the last estimate, we arrive to the inequality

$$\begin{aligned}&\Vert R^mv(t)\Vert _{L^2(0,T; H^2(\varOmega ))}^2 \le \left( \frac{\left( C\varGamma \left( \frac{1}{2}\alpha \right) \right) ^m}{\varGamma \left( \frac{1}{2}\alpha m\right) }\right) ^2 \Vert t^{\frac{\alpha m}{2}-1}\Vert _{L^1(0,T)}^2 \Vert v\Vert _{L^2(0,T;H^2(\varOmega ))}^2\\&\quad = \frac{\left( CT^{\frac{\alpha }{2}} \varGamma \left( \frac{1}{2}\alpha \right) \right) ^{2m}}{\varGamma \left( \frac{1}{2}\alpha m +1\right) ^2} \Vert v\Vert _{L^2(0,T;H^2(\varOmega ))}^2. \end{aligned}$$

Employing the known asymptotic behavior of the gamma function, we obtain the relation

$$\begin{aligned} \lim _{m\rightarrow \infty } \frac{\left( CT^{\frac{\alpha }{2}} \varGamma \left( \frac{1}{2}\alpha \right) \right) ^m}{\varGamma \left( \frac{1}{2}\alpha m +1\right) } = 0 \end{aligned}$$

that means that for sufficiently large \(m\in {\mathbb {N}}\), the map**

$$\begin{aligned} R^m: L^2(0,T;H^2(\varOmega ))\, \longrightarrow \, L^2(0,T;H^2(\varOmega )) \end{aligned}$$

is a contraction. Hence, by the Banach fixed point theorem, the equation (2.13) possesses a unique fixed point. Therefore, by the first equation in (2.4), we obtain the inclusion \(\partial _t^{\alpha }(u-a) \in L^2(0,T;L^2(\varOmega ))\). Since \(\Vert \eta \Vert _{H_{\alpha }(0,T)} \sim \Vert \partial _t^{\alpha }\eta \Vert _{L^2(0,T)}\) for \(\eta \in H_{\alpha }(0,T)\) ( [13]), we finally obtain the estimate

$$\begin{aligned} \Vert u-a\Vert _{H_{\alpha }(0,T;L^2(\varOmega ))} + \Vert u\Vert _{L^2(0,T;H^2(\varOmega ))} \le C(\Vert a\Vert _{H^1(\varOmega )} + \Vert F\Vert _{L^2(0,T;L^2(\varOmega ))}). \end{aligned}$$

The proof of Theorem 1 is completed. \(\square \)

3 Key lemma

For derivation of the comparison principles for solutions to the initial-boundary value problems for the linear and semilinear time-fractional diffusion equations, we need some auxiliary results that are formulated and proved in this section.

In addition to the operator \(-A_0\) defined by (2.5), we define an elliptic operator \(-A_1\) with a positive zeroth-order coefficient:

$$\begin{aligned}{} & {} (-A_1(t)v)(x):= (-A_1v)(x)\nonumber \\{} & {} := \sum _{i,j=1}^d\partial _i(a_{ij}(x)\partial _jv(x)) + \sum _{j=1}^d b_j(x,t)\partial _jv - b_0(x,t)v, \end{aligned}$$
(3.1)

where \(b_0 \in C^1([0,T];C^1(\overline{\varOmega })) \cap C([0,T];C^2(\overline{\varOmega }))\), \(b_0(x,t) > 0,\ (x,t)\in \overline{\varOmega }\times [0,T]\), and \(\min _{(x,t)\in \overline{\varOmega }\times [0,T]} b_0(x,t)\) is sufficiently large.

We also recall that for \(y\in W^{1,1}(0,T)\), the pointwise Caputo derivative \(d_t^{\alpha }\) is defined by

$$\begin{aligned} d_t^{\alpha }y(t) = \frac{1}{\varGamma (1-\alpha )} \int ^t_0 (t-s)^{-\alpha }\frac{dy}{ds}(s) ds. \end{aligned}$$
(3.2)

In what follows, we employ an extremum principle for the Caputo fractional derivative formulated below.

Lemma 2

([17]) Let the inclusions \(y\in C[0,T]\) and \(t^{1-\alpha }y' \in C[0,T]\) hold true.

If the function \(y=y(t)\) attains its minimum over the interval [0, T] at the point \(t_0 \in (0, \,T]\), then

$$\begin{aligned} d_t^{\alpha }y(t_0) \le 0. \end{aligned}$$

In Lemma 2, the assumption \(t_0>0\) is essential. This lemma was formulated and proved in [17] under a weaker regularity condition posed on the function y, but for our arguments we can assume that \(y\in C[0,T]\) and \(t^{1-\alpha }y' \in C[0,T]\).

Employing Lemma 2, we now formulate and prove our key lemma that is a basis for further derivations in this paper.

Lemma 3

(Positivity of a smooth solution) For \(F\in L^2(0,T;L^2(\varOmega ))\) and \(a\in H^1(\varOmega )\), let \(F(x,t) \ge 0,\ (x,t)\in \varOmega \times (0,T)\), \(a(x)\ge 0,\ x\in \varOmega \), and \(\min _{(x,t)\in \overline{\varOmega }\times [0,T]} b_0(x,t)\) be a sufficiently large positive constant. Furthermore, we assume that there exists a solution \(u\in C([0,T];C^2(\overline{\varOmega }))\) to the initial-boundary value problem

$$\begin{aligned} \left\{ \begin{array}{rl} &{} \partial _t^{\alpha }(u-a) + A_1u = F(x,t), \quad x\in \varOmega ,\, 0<t<T, \\ &{} \partial _{\nu _A}u + \sigma (x)u = 0 \quad \text{ on } \partial \varOmega \times (0,T),\\ &{} u(x,\cdot ) - a(x) \in H_{\alpha }(0,T) \quad \text{ for } \text{ almost } \text{ all } x\in \varOmega \end{array} \right. \end{aligned}$$
(3.3)

and u satisfies the condition \(t^{1-\alpha }\partial _tu \in C([0,T];C(\overline{\varOmega }))\).

Then the solution u is non-negative:

$$\begin{aligned} u (x,t)\ge 0,\ (x,t)\in \varOmega \times (0,T). \end{aligned}$$

For the partial differential equations of parabolic type with the Robin boundary condition (\(\alpha =1\) in (3.3)), a similar positivity property is well-known (see, e.g., [11]). However, it is worth mentioning that the regularity of the solution to the problem (3.3) at the point \(t=0\) is a more delicate question compared to the one in the case \(\alpha =1\). In particular, we cannot expect the inclusion \(u(x,\cdot ) \in C^1[0,T]\). This can be illustrated by a simple example of the equation \(\partial _t^{\alpha }y(t) = y(t)\) with \(y(t)-1 \in H_{\alpha }(0,T)\) whose unique solution \(y(t) = E_{\alpha ,1}(t^{\alpha })\) does not belong to the space \(C^1[0,T]\).

Proof

First we introduce an auxiliary function \(\psi \in C^1([0,T];C^2(\overline{\varOmega }))\) that satisfies the conditions

$$\begin{aligned} \left\{ \begin{array}{rl} &{} A_1\psi (x,t) = 1, \quad (x,t) \in \varOmega \times [0,T], \\ &{} \partial _{\nu _A}\psi + \sigma \psi = 1 \quad \text{ on } \partial \varOmega \times [0,T]. \end{array}\right. \end{aligned}$$
(3.4)

Proving existence of such function \(\psi \) is non-trivial. In this section, we focus on the proof of the lemma and then come back to the problem (3.4) in Appendix.

Now, choosing \(M>0\) sufficiently large and \(\varepsilon >0\) sufficiently small, we set

$$\begin{aligned} w(x,t) := u(x,t) + \varepsilon (M + \psi (x,t) + t^{\alpha }), \quad x\in \varOmega ,\, 0<t<T. \end{aligned}$$
(3.5)

For a fixed \(x\in \varOmega \), by the assumption on the regularity of u, we have the inclusion

$$\begin{aligned} t^{1-\alpha }\partial _tu(x,\cdot ) \in C[0,T]. \end{aligned}$$
(3.6)

Then, \(\partial _tu(x,\cdot ) \in L^1(0,T)\), that is, \(u(x,\cdot ) \in W^{1,1}(0,T)\). Moreover,

$$\begin{aligned} u(x,0) - a(x) = 0, \quad x\in \varOmega . \end{aligned}$$
(3.7)

On the other hand, for \(w\in H_{\alpha }(0,T) \cap W^{1,1}(0,T)\) and \(w(0) = 0\), the equality

$$\begin{aligned} \partial _t^{\alpha }w = d_t^{\alpha }w = d_t^{\alpha }(w+c) \end{aligned}$$

holds true with any constant c (see, e.g., Theorem 2.4 of Chapter 2 in [13]).

Since \(u(x,\cdot ) - a \in H_{\alpha }(0,T)\) and \(u(x,\cdot ) \in W^{1,1}(0,T)\), by (3.7), the relations \(\partial _t^{\alpha }(u-a) = d_t^{\alpha }(u-a) = d_t^{\alpha }u\) hold true for almost all \(x\in \varOmega \).

Furthermore, since \(\varepsilon (M+\psi (\cdot ,t)+t^{\alpha }) \in W^{1,1}(0,T)\), we obtain

$$\begin{aligned}&d_t^{\alpha }w = d_t^{\alpha }(u + \varepsilon (M+\psi (x,t)+t^{\alpha }))) = d_t^{\alpha }u + \varepsilon d_t^{\alpha }(M+\psi (x,t)+t^{\alpha })\\&\quad = \partial _t^{\alpha }(u-a) + \varepsilon (d_t^{\alpha }(\psi + t^{\alpha })) = \partial _t^{\alpha }(u-a) + \varepsilon (d_t^{\alpha }\psi + \varGamma (\alpha +1)) \end{aligned}$$

and

$$\begin{aligned}&A_1w = A_1u + \varepsilon A_1\psi + \varepsilon A_1t^{\alpha } + \varepsilon A_1M\\&\quad = A_1u + \varepsilon + \varepsilon b_0(x,t)t^{\alpha } + b_0(x,t)\varepsilon M. \end{aligned}$$

Now we choose a constant \(M>0\) such that \(M + \psi (x,t) \ge 0\) and \(d_t^{\alpha }\psi (x,t) + b_0(x,t)M > 0\) for \((x,t) \in \overline{\varOmega } \times [0,T]\), so that

$$\begin{aligned}{} & {} d_t^{\alpha }w + A_1w\nonumber \\{} & {} \quad = F + \varepsilon (\varGamma (\alpha +1) + d_t^{\alpha }\psi + 1 + b_0(x,t)t^{\alpha } + b_0(x,t) M) > 0 \quad \text{ in } \varOmega \times (0,T).\nonumber \\ \end{aligned}$$
(3.8)

Moreover, because of the relation \(\partial _{\nu _A}w = \partial _{\nu _A}u + \varepsilon \partial _{\nu _A}\psi \), we obtain the following estimate:

$$\begin{aligned}{} & {} \partial _{\nu _A}w + \sigma w = \partial _{\nu _A}u + \sigma u + \varepsilon + \sigma \varepsilon t^{\alpha } + \varepsilon \sigma M\nonumber \\{} & {} \quad \ge \varepsilon + \sigma \varepsilon t^{\alpha } + \varepsilon \sigma M \ge \varepsilon \quad \text{ on } \partial \varOmega \times (0,T). \end{aligned}$$
(3.9)

Evaluation of the representation (3.5) at the point \(t=0\) immediately leads to the formula

$$\begin{aligned} w(x,0) = u(x,0) + \varepsilon (\psi (x,0) + M), \quad x\in \varOmega . \end{aligned}$$

Let us assume that the inequality

$$\begin{aligned} \min _{(x,t)\in \overline{\varOmega }\times [0,T]} w(x,t) \ge 0 \end{aligned}$$

does not hold valid, that is, there exists a point \((x_0,t_0) \in \overline{\varOmega }\times [0,T]\) such that

$$\begin{aligned} w(x_0,t_0):= \min _{(x,t)\in \overline{\varOmega }\times [0,T]} w(x,t) < 0. \end{aligned}$$
(3.10)

Since \(M>0\) is sufficiently large and u(x, 0) is non-negative, we obtain the inequality

$$\begin{aligned} w(x,0) = u(x,0) + \varepsilon (\psi (x,0) + M) \ge u(x,0) \ge 0, \quad x\in \overline{\varOmega }, \end{aligned}$$

and thus \(t_0\) cannot be zero.

Next, we show that \(x_0 \not \in \partial \varOmega \). Indeed, let us assume that \(x_0 \in \partial \varOmega \). Then the estimate (3.9) yields that \(\partial _{\nu _A}w(x_0,t_0) + \sigma (x_0)w(x_0,t_0) \ge \varepsilon \). By (3.10) and \(\sigma (x_0)\ge 0\), we obtain

$$\begin{aligned} \partial _{\nu _A}w(x_0,t_0) \ge -\sigma (x_0)w(x_0,t_0) + \varepsilon \ge \varepsilon > 0, \end{aligned}$$

which implies

$$\begin{aligned}{} & {} \partial _{\nu _A}w(x_0,t_0) = \sum _{i,j=1}^da_{ij}(x_0)\nu _j(x_0)\partial _iw(x_0,t_0) = \nabla w(x_0,t_0) \cdot {\mathcal {A}}(x_0)\nu (x_0)\nonumber \\{} & {} \quad = \sum _{i=1}^d (\partial _iw)(x_0,t_0)[{\mathcal {A}}(x_0)\nu (x_0)]_i > 0. \end{aligned}$$
(3.11)

Here \({\mathcal {A}}(x) = (a_{ij}(x))_{1\le i,j\le d}\) and \([b]_i\) means the i-th element of a vector b.

For sufficiently small \(\varepsilon _0>0\) and \(x_0\in \partial \varOmega \), we now verify the inclusion

$$\begin{aligned} x_0 - \varepsilon _0{\mathcal {A}}(x_0)\nu (x_0) \in \varOmega . \end{aligned}$$
(3.12)

Indeed, since the matrix \({\mathcal {A}}(x_0)\) is positive-definite, the inequality

$$\begin{aligned} (\nu (x_0)\, \cdot \, -\varepsilon _0{\mathcal {A}}(x_0)\nu (x_0)) = -\varepsilon _0({\mathcal {A}}(x_0)\nu (x_0)\, \cdot \, \nu (x_0)) < 0 \end{aligned}$$

holds true. In other words, the inequality

$$\begin{aligned} \angle (\nu (x_0), \, (x_0 - \varepsilon _0{\mathcal {A}}(x_0)\nu (x_0)) - x_0) > \frac{\pi }{2} \end{aligned}$$

is satisfied. Because the boundary \(\partial \varOmega \) is smooth, the domain \(\varOmega \) is locally located on one side of \(\partial \varOmega \). In a small neighborhood of the point \(x_0\in \partial \varOmega \), the boundary \(\partial \varOmega \) can be described in the local coordinates composed of its tangential component in \({\mathbb {R}}^{d-1}\) and the normal component along \(\nu (x_0)\). Consequently, if \(y \in {\mathbb {R}}^d\) satisfies the inequality \(\angle (\nu (x_0), y-x_0) > \frac{\pi }{2}\), then \(y\in \varOmega \). Therefore, for a sufficiently small \(\varepsilon _0>0\), the point \(x_0-\varepsilon _0{\mathcal {A}}(x_0)\nu (x_0)\) is located in \(\varOmega \) and we have proved the inclusion (3.12).

Moreover, for sufficiently small \(\varepsilon _0>0\), we can prove that

$$\begin{aligned} w(x_0 - \varepsilon _0{\mathcal {A}}(x_0)\nu (x_0),\,t_0) < w(x_0,t_0). \end{aligned}$$
(3.13)

Indeed, the inequality (3.10) yields

$$\begin{aligned} \sum _{i=1}^d (\partial _iw)(x_0-\eta {\mathcal {A}}(x_0)\nu (x_0),\,t_0) [{\mathcal {A}}(x_0)\nu (x_0)]_i > 0 \quad \text{ if } \vert \eta \vert < \varepsilon _0. \end{aligned}$$

Then, by the mean value theorem, we obtain the inequality

$$\begin{aligned}&w(x_0 - \xi {\mathcal {A}}(x_0)\nu (x_0),\,t_0) - w(x_0,t_0)\\&\quad = \xi \sum _{i=1}^d \partial _iw(x_0 - \theta {\mathcal {A}}(x_0)\nu (x_0),\,t_0) (-[{\mathcal {A}}(x_0)\nu (x_0)]_i) < 0, \end{aligned}$$

where \(\theta \) is a number between 0 and \(\xi \in (0,\varepsilon _0)\). Thus, the inequality (3.13) is verified.

By combining (3.13) with (3.12), we conclude that there exists a point \(\widetilde{x_0} \in \varOmega \) such that the inequality \(w(\widetilde{x_0},t_0) < w(x_0,t_0)\) holds true, which contradicts the assumption (3.10). Thus, we have proved that \(x_0 \not \in \partial \varOmega \).

According to (3.10), the function w attains its minimum at the point \((x_0,t_0)\). Because \(0 < t_0 \le T\), Lemma 2 yields the inequality

$$\begin{aligned} d_t^{\alpha }w(x_0,t_0) \le 0. \end{aligned}$$
(3.14)

Since \(x_0 \in \varOmega \), the necessary condition for an extremum point leads to the equality

$$\begin{aligned} \nabla w(x_0,t_0) = 0. \end{aligned}$$
(3.15)

Moreover, because the function w attains its minimum at the point \(x_0 \in \varOmega \), in view of the sign of the Hessian, the inequality

$$\begin{aligned} \sum _{i,j=1}^da_{ij}(x_0)\partial _i\partial _j w(x_0,t_0) \ge 0 \end{aligned}$$
(3.16)

holds true (see, e.g., the proof of Lemma 1 in Section 1 of Chapter 2 in [5]).

The inequalities \(b(x_0,t_0)>0\), \(w(x_0,t_0) < 0\), and (3.14)-(3.16) lead to the estimate

$$\begin{aligned}&d_t^{\alpha }w(x_0,t_0) + A_1w(x_0,t_0)\\&\quad = d_t^{\alpha }w (x_0,t_0) - \sum _{i,j=1}^da_{ij}(x_0)\partial _i\partial _jw(x_0,t_0) - \sum _{i=1}^d (\partial _ia_{ij})(x_0)\partial _jw(x_0,t_0)\\&\qquad - \sum _{i=1}^d b_i(x_0,t_0)\partial _iw(x_0,t_0) + b(x_0,t_0)w(x_0,t_0) < 0, \end{aligned}$$

which contradicts the inequality (3.8).

Thus, we have proved that

$$\begin{aligned} u(x,t) + \varepsilon (M+\psi (x,t)+t^{\alpha }) = w(x,t) \ge 0, \quad (x,t) \in \varOmega \times (0,T). \end{aligned}$$

Since \(\varepsilon >0\) is arbitrary, we let \(\varepsilon \downarrow 0\) to obtain the inequality \(u(x,t) \ge 0\) for \((x,t) \in \varOmega \times (0,T)\) and the proof of Lemma 3 is completed. \(\square \)

Let us finally mention that the positivity of the function \(b_0\) from the definition of the operator \(-A_1\) is an essential condition for validity of our proof of Lemma 3. However, in the next section, we remove this condition while deriving the comparison principles for the solutions to the initial-boundary value problem (2.4).

4 Comparison principles

According to the results formulated in Theorem 1, in this section, we consider the solutions to the initial-boundary value problem (2.4) that belong to the following space of functions:

$$\begin{aligned} {\mathcal {Y}}_\alpha := \{ u; \, u-a\in H_{\alpha }(0,T;L^2(\varOmega )), \, u\in L^2(0,T;H^2(\varOmega ))\}. \end{aligned}$$
(4.1)

In what follows, by u(Fa) we denote the solution to the problem (2.4) with the initial data a and the source function F.

Our first result concerning the comparison principles for the solutions to the initial-boundary value problems for the linear time-fractional diffusion equation is presented in the next theorem.

Theorem 2

Let the functions \(a \in H^1(\varOmega )\) and \(F \in L^2(\varOmega \times (0,T))\) satisfy the inequalities \(F(x,t) \ge 0,\ (x,t)\in \varOmega \times (0,T)\) and \(a(x) \ge 0,\ x\in \varOmega \), respectively.

Then the solution \(u(F,a) \in {\mathcal {Y}}_\alpha \) to the initial-boundary value problem (2.4) is non-negative, e.g., the inequality

$$\begin{aligned} u(F,a)(x,t) \ge 0,\ (x,t)\in \varOmega \times (0,T) \end{aligned}$$

holds true.

Let us emphasize that the non-negativity of the solution u to the problem (2.4) holds true for the space \( {\mathcal {Y}}_\alpha \) and thus u does not necessarily satisfy the inclusions \(u \in C([0,T];C^2(\overline{\varOmega }))\) and \(t^{1-\alpha }\partial _tu \in C([0,T]; C(\overline{\varOmega }))\). Therefore, Theorem 2 is widely applicable. Before presenting its proof, let us discuss one of its corollaries in form of a comparison property:

Corollary 1

Let \(a_1, a_2 \in H^1(\varOmega )\) and \(F_1, F_2 \in L^2(\varOmega \times (0,T))\) satisfy the inequalities \(a_1(x) \ge a_2(x),\ x\in \varOmega \) and \(F_1(x,t) \ge F_2(x,t), \ (x,t)\in \varOmega \times (0,T)\), respectively.

Then the inequality

$$\begin{aligned} u(F_1, a_1)(x,t) \ge u(F_2,a_2)(x,t), \ (x,t)\in \varOmega \times (0,T) \end{aligned}$$

holds true.

Proof

Setting \(a:= a_1-a_2\), \(F:= F_1 - F_2\) and \(u:= u(F_1,a_1) - u(F_2,a_2)\), we immediately obtain the inequalities \(a(x)\ge 0,\ x\in \varOmega \) and \(F(x,t)\ge 0, \ (x,t)\in \varOmega \times (0,T)\) and

$$\begin{aligned} \left\{ \begin{array}{rl} &{} \partial _t^{\alpha }(u-a) + Au = F \ge 0 \quad \text{ in } \varOmega \times (0,T), \\ &{} \partial _{\nu _A}u + \sigma u = 0 \quad \text{ on } \partial \varOmega . \end{array}\right. \end{aligned}$$

Therefore, Theorem 2 implies that \(u(x,t)\ge 0, \ (x,t)\in \varOmega \times (0,T)\), that is, \(u(F_1, a_1)(x,t) \ge u(F_2,a_2)(x,t), \ (x,t)\in \varOmega \times (0,T)\). \(\square \)

In its turn, Corollary 1 can be applied for derivation of the lower and upper bounds for the solutions to the initial-boundary value problem (2.4) by suitably choosing the initial values and the source functions. Let us demonstrate this technique on an example.

Example 1

Let the coefficients \(a_{ij}, b_j\), \(1\le i,j\le d\) of the operator

$$\begin{aligned} -Av(x) = \sum _{i,j=1}^d \partial _i(a_{ij}(x)\partial _jv(x)) + \sum _{j=1}^d b_j(x,t)\partial _jv(x) \end{aligned}$$

from the initial-boundary value problem (2.4) satisfy the conditions (1.2). Now we consider the homogeneous initial condition \(a(x)=0,\ x\in \varOmega \) and assume that the source function \(F \in L^2(0,T;L^2(\varOmega ))\) satisfies the inequality

$$\begin{aligned} F(x,t) \ge \delta t^{\beta }, \quad x\in \varOmega ,\, 0<t<T \end{aligned}$$

with certain constants \(\beta \ge 0\) and \(\delta >0\).

Then the solution u(F, 0) can be estimated from below as follows:

$$\begin{aligned} u(F,0)(x,t) \ge \frac{\delta \varGamma (\beta +1)}{\varGamma (\alpha +\beta +1)} t^{\alpha +\beta }, \quad x\in \varOmega , \, 0\le t \le T. \end{aligned}$$
(4.2)

Indeed, it is easy to verify that the function

$$\begin{aligned} {\underline{u}}(x,t):= \frac{\delta \varGamma (\beta +1)}{\varGamma (\alpha +\beta +1)} t^{\alpha +\beta }, \quad x\in \varOmega , \, t>0 \end{aligned}$$

is a solution to the following problem:

$$\begin{aligned} \left\{ \begin{array}{rl} &{} \partial _t^{\alpha }{\underline{u}} + A{\underline{u}} = \delta t^{\beta } \quad \text{ in }\ \varOmega \times (0,T), \\ &{} \partial _{\nu _A}{\underline{u}} = 0 \quad \text{ on } \partial \varOmega \times (0,T), \\ &{} {\underline{u}}(x,\cdot ) \in H_{\alpha }(0,T). \end{array}\right. \end{aligned}$$

Due to the inequality \(F(x,t) \ge \delta t^{\beta },\ (x,t) \in \varOmega \times (0,T)\), we can apply Corollary 1 to the solutions u and \({\underline{u}}\) and the inequality (4.2) immediately follows.

In particular, for the spatial dimensions \(d \le 3\), the Sobolev embedding theorem leads to the inclusion \(u \in L^2(0,T;H^2(\varOmega )) \subset L^2(0,T;C(\overline{\varOmega }))\) and thus the strict inequality \(u(F,0)(x,t) > 0\) holds true for almost all \(t>0\) and all \(x\in \overline{\varOmega }\).

Now we proceed to the proof of Theorem 2.

Proof

In the proof, we employ the operators Qv(t) and G(t) defined by (2.14). In terms of these operators, the solution \(u(t):= u(F,a)(t)\) to the initial-boundary problem (2.4) satisfies the integral equation

$$\begin{aligned} u(F,a)(t) = G(t) + \int ^t_0 K(t-s)Qu(s) ds, \quad 0<t<T. \end{aligned}$$
(4.3)

For readers’ convenience, we split the proof into three parts.

I. First part of the proof: existence of a smoother solution

In the formulation of Lemma 3, we assumed existence of a solution \(u\in C([0,T];C^2(\overline{\varOmega }))\) to the initial-boundary value problem (3.3) satisfying the inclusion \(t^{1-\alpha }\partial _tu \in C([0,T];C(\overline{\varOmega }))\). On the other hand, Theorem 1 asserts the unique existence of solution u to the initial-boundary value problem (2.4) from the space \({\mathcal {Y}}_\alpha \), i.e., of the solution u that satisfies the inclusions \(u\in L^2(0,T;H^2(\varOmega ))\) and \(u - a \in H_{\alpha }(0,T;L^2(\varOmega ))\).

In this part of the proof, we show that for \(a \in C^{\infty }_0(\varOmega )\) and \(F \in C^{\infty }_0(\varOmega \times (0,T))\), the solution to the problem (2.4) satisfies the regularity assumptions formulated in Lemma 3.

More precisely, we first prove the following lemma:

Lemma 4

Let \(a_{ij}\), \(b_j\), c satisfy the conditions (1.2) and the inclusions \(a\in C^{\infty }_0(\varOmega )\), \(F \in C^{\infty }_0(\varOmega \times (0,T))\) hold true.

Then the solution \(u=u(F,a)\) to the problem (2.4) satisfies the inclusions

$$\begin{aligned} u \in C([0,T];C^2(\overline{\varOmega })), \quad t^{1-\alpha }\partial _tu \in C([0,T];C(\overline{\varOmega })) \end{aligned}$$

and \(\lim _{t\rightarrow 0} \Vert u(t) - a\Vert _{L^2(\varOmega )} = 0\).

Proof

We recall that \(c_0>0\) is a positive fixed constant and

$$\begin{aligned} -A_0v = \sum _{i,j=1}^d\partial _i(a_{ij}(x)\partial _jv) - c_0v,\ {\mathcal {D}}(A_0) = \{ v \in H^2(\varOmega );\, \partial _{\nu _A}v + \sigma v = 0 \,\, \text{ on } \partial \varOmega \}. \end{aligned}$$

Then \({\mathcal {D}}(A_0^{\frac{1}{2}}) = H^1(\varOmega )\) and \(\Vert A_0^{\frac{1}{2}}v\Vert \sim \Vert v\Vert _{H^1(\varOmega )}\) ( [6]). Moreover, for the operators S(t) and K(t) defined by (2.6) and (2.7), the estimates (2.8) hold true.

In what follows, we denote \( \frac{\partial u}{\partial t}(\cdot ,t)\) by \(u'(t) = \frac{du}{dt}(t)\) if there is no fear of confusion.

The solution u to the integral equation (4.3) can be constructed as a fixed point of the equation

$$\begin{aligned} A_0u(t) = A_0G(t) + \int ^t_0 A_0^{\frac{1}{2}}K(t-s)A_0^{\frac{1}{2}}Qu(s) ds, \quad 0<t<T. \end{aligned}$$
(4.4)

As already proved, this fixed point satisfies the inclusion \(u\in L^2(0,T;H^2(\varOmega ))\) \( \cap \) \( (H_{\alpha }(0,T;L^2(\varOmega )) + \{ a\}).\)

Now we derive some estimates for the norms \(\Vert A_0^{\kappa }u(t)\Vert \), \(\kappa =1,2\) and \(\Vert A_0u'(t)\Vert \) for \(0<t<T\). First we set

$$\begin{aligned} D:= \sup _{0<t<T} (\Vert A_0F(t)\Vert + \Vert A_0F'(t)\Vert + \Vert A_0^2F(t)\Vert ) + \Vert a\Vert _{H^4(\varOmega )}. \end{aligned}$$

Since \(F\in C^{\infty }_0(\varOmega \times (0,T))\), we obtain the inclusion \(F\in L^{\infty }(0,T;{\mathcal {D}}(A_0^2))\) and the inequality \(D < +\infty \). Moreover, in view of (2.8), for \(\kappa =1,2\), we get the estimates

$$\begin{aligned}&\left\| A_0^{\kappa }\int ^t_0 K(t-s)F(s) ds \right\| \le C\int ^t_0 \Vert K(t-s)\Vert \Vert A_0^{\kappa }F(s)\Vert ds\\&\quad \le C\left( \int ^t_0 (t-s)^{\alpha -1}ds \right) \sup _{0<s<T} \Vert A_0^{\kappa }F(s)\Vert \le CD,\\&\left\| A_0\frac{d}{dt}\int ^t_0 K(t-s)F(s) ds \right\| = \left\| A_0\frac{d}{dt}\int ^t_0 K(s)F(t-s) ds \right\| \\&\quad = \left\| A_0K(t)F(0) + A_0\int ^t_0 K(s)F'(t-s) ds \right\| \\&\quad \le C\left\| A_0\int ^t_0 K(s)F'(t-s) ds \right\| \le C \int ^t_0 s^{\alpha -1} \Vert A_0F'(t-s) \Vert ds < CD. \end{aligned}$$

The regularity conditions (1.2) lead to the estimates

$$\begin{aligned}{} & {} \Vert A_0^{\frac{1}{2}}Q(s)u(s)\Vert \le C\Vert Q(s)u(s)\Vert _{H^1(\varOmega )}\nonumber \\{} & {} \quad = C\left\| \sum _{j=1}^d b_j(s)\partial _ju(s) + (c_0+c(s))u(s) \right\| _{H^1(\varOmega )}\nonumber \\{} & {} \quad \le C\Vert u(s)\Vert _{H^2(\varOmega )} \le C\Vert A_0u(s)\Vert ,\quad 0<s<T. \end{aligned}$$
(4.5)

Moreover,

$$\begin{aligned} \Vert A_0S(t)a\Vert = \Vert S(t)A_0a\Vert \le C\Vert a\Vert _{H^2(\varOmega )} \le CD \end{aligned}$$

by using the inequalities (2.8). Then

$$\begin{aligned}&\Vert A_0u(t)\Vert \le CD + \int ^t_0 \Vert A_0^{\frac{1}{2}}K(t-s)\Vert \Vert A_0^{\frac{1}{2}}Q(s)u(s)\Vert ds\\&\quad \le CD + C\int ^t_0 (t-s)^{\frac{1}{2}\alpha -1}\Vert A_0u(s)\Vert ds, \quad 0<s<T. \end{aligned}$$

The generalized Gronwall inequality yields the estimate

$$\begin{aligned} \Vert A_0u(t)\Vert \le CD + C\int ^t_0 (t-s)^{\frac{1}{2}\alpha -1}D ds \le CD, \quad 0<t<T, \end{aligned}$$

which implies the inequality

$$\begin{aligned} \Vert A_0u\Vert _{L^{\infty }(0,T;H^2(\varOmega ))} \le CD. \end{aligned}$$

Next, for the space \(C([0,T]; L^2(\varOmega ))\), we can repeat the same arguments as the ones employed for the iterations \(R^n\) of the operator R in the proof of Theorem 1 and apply the fixed point theorem to the equation (4.3) that leads to the inclusion \(A_0u \in C([0,T];L^2(\varOmega ))\). The obtained results implicate

$$\begin{aligned} u\in C([0,T];H^2(\varOmega )), \quad \Vert u\Vert _{C([0,T];H^2(\varOmega ))} \le CD. \end{aligned}$$
(4.6)

Choosing \(\varepsilon _0 > 0\) sufficiently small, we have the equation

$$\begin{aligned} A_0^{\frac{3}{2}}u(t) = A_0^{\frac{3}{2}}G(t) + \int ^t_0 A_0^{\frac{3}{4}+\varepsilon _0} K(t-s)A_0^{\frac{3}{4}-\varepsilon _0}Q(s)u(s)ds, \quad 0<t<T. \end{aligned}$$
(4.7)

Next, according to [6], the inclusion

$$\begin{aligned} {\mathcal {D}}(A_0^{\frac{3}{4}-\varepsilon _0}) \subset H^{\frac{3}{2}-2\varepsilon _0}(\varOmega ) \end{aligned}$$

holds true. Now we proceed to the proof of the inclusion \(Q(s)u(s) \in {\mathcal {D}}(A_0^{\frac{3}{4}-\varepsilon _0})\). By (2.8), we obtain the inequality

$$\begin{aligned} \Vert A_0^{\frac{3}{2}}u(t)\Vert \le CD + \int ^t_0 (t-s)^{(\frac{1}{4}-\varepsilon _0)\alpha -1} \Vert A_0^{\frac{3}{4}-\varepsilon _0}Q(s)u(s)\Vert ds, \end{aligned}$$

which leads to the estimate

$$\begin{aligned} \Vert u(t)\Vert _{H^3(\varOmega )} \le CD + \int ^t_0 (t-s)^{(\frac{1}{4}-\varepsilon _0)\alpha -1} \Vert u(s)\Vert _{H^3(\varOmega )} ds, \quad 0<t<T \end{aligned}$$
(4.8)

because of the inequality

$$\begin{aligned} \Vert A_0^{\frac{3}{4}-\varepsilon _0}Q(s)u(s)\Vert \le C\Vert Q(s)u(s)\Vert _{H^{\frac{3}{2}}(\varOmega )} \le C\Vert Q(s)u(s)\Vert _{H^2(\varOmega )} \le C\Vert u(s)\Vert _{H^3(\varOmega )}, \end{aligned}$$

which follows from the regularity conditions (1.2) posed on the coefficients \(b_j, c\).

For \(0<t<T\), the generalized Gronwall inequality applied to the integral inequality (4.8) yields the estimate

$$\begin{aligned} \Vert u(t)\Vert _{H^3(\varOmega )} \le C\left( 1 + t^{\alpha \left( \frac{1}{4}-\varepsilon _0\right) } \right) D. \end{aligned}$$

For the relation (4.7), we repeat the same arguments as the ones employed in the proof of Theorem 1 to estimate \(A_0^{\frac{3}{2}}u(t)\) in the norm \(C([0,T];L^2(\varOmega ))\) by the fixed point theorem arguments and thus we obtain the inclusion \(A_0^{\frac{3}{2}}u \in C([0,T];L^2(\varOmega ))\).

Summarising the estimates derived above, we have shown that

$$\begin{aligned} \left\{ \begin{array}{rl} &{} u \in C([0,T];{\mathcal {D}}(A_0^{\frac{3}{2}})) \subset C([0,T];H^3(\varOmega )), \\ &{} \Vert u(t)\Vert _{H^3(\varOmega )} \le C\left( 1 + t^{\alpha \left( \frac{1}{4}-\varepsilon _0\right) } \right) D, \quad 0<t<T. \end{array}\right. \end{aligned}$$
(4.9)

Next we estimate the norm \(\Vert Au'(t)\Vert \). First, \(u'(t)\) is represented in the form

$$\begin{aligned}&u'(t) = G'(t) + \frac{d}{dt}\int ^t_0 K(t-s)Q(s)u(s) ds\\&\quad = G'(t) + \frac{d}{dt}\int ^t_0 K(s)Q(t-s)u(t-s) ds \, = \, G'(t) + K(t)Q(0)u(0) \\&\qquad + \int ^t_0 K(s) (Q(t-s)u'(t-s) + Q'(t-s)u(t-s)) ds, \quad 0<t<T, \end{aligned}$$

so that

$$\begin{aligned}{} & {} A_0u'(t) = A_0G'(t) + A_0K(t)Q(0)u(0)\nonumber \\{} & {} \quad + \int ^t_0 A_0^{\frac{1}{2}}K(s) A_0^{\frac{1}{2}}(Q(t-s)u'(t-s) + Q'(t-s)u(t-s)) ds, \quad 0<t<T.\nonumber \\ \end{aligned}$$
(4.10)

Similarly to the arguments applied for derivation of (4.5), we obtain the inequality

$$\begin{aligned} \Vert A_0^{\frac{1}{2}}(Q(t-s)u'(t-s) + Q'(t-s)u(t-s))\Vert \le C\Vert A_0u'(t-s)\Vert , \quad 0<t<T. \end{aligned}$$

The inclusion \(Q(0)u(0) = Q(0)a \in C^2_0(\varOmega ) \subset {\mathcal {D}}(A_0)\) follows from the regularity conditions (1.2) and the inclusion \(a \in C^{\infty }_0(\varOmega )\). Furthermore, by (2.7) and (2.8), we obtain

$$\begin{aligned} \Vert A_0S'(t)a\Vert = \Vert A_0^2K(t)a\Vert = \Vert K(t)A_0^2a\Vert \le Ct^{\alpha -1}\Vert A_0^2a\Vert \le Ct^{\alpha -1}\Vert a\Vert _{H^4(\varOmega )} \end{aligned}$$

and

$$\begin{aligned} \Vert K(t)A_0(Q(0)a)\Vert \le Ct^{1-\alpha }\Vert A_0(Q(0)a)\Vert \le Ct^{\alpha -1}\Vert a\Vert _{H^3(\varOmega )}. \end{aligned}$$

Hence, the representation (4.10) leads to the estimate

$$\begin{aligned} \Vert A_0u'(t)\Vert \le Ct^{\alpha -1}D + C\int ^t_0 s^{\frac{1}{2}\alpha -1}\Vert A_0u'(t-s)\Vert ds, \quad 0<t<T. \end{aligned}$$

Now we consider a vector space

$$\begin{aligned} \widetilde{X}:= \{v\in C([0,T];L^2(\varOmega )) \cap C^1((0,T];L^2(\varOmega ));\, t^{1-\alpha }\partial _tv \in C([0,T];L^2(\varOmega ))\} \end{aligned}$$

with the norm

$$\begin{aligned} \Vert v\Vert _{\widetilde{X}}:= \max _{0\le t\le T} \Vert t^{1-\alpha }\partial _tv (\cdot ,t)\Vert _{L^2(\varOmega )} + \max _{0\le t\le T} \Vert v(\cdot ,t)\Vert _{L^2(\varOmega )}. \end{aligned}$$

It is easy to verify that \(\widetilde{X}\) with the norm \(\Vert v\Vert _{\widetilde{X}}\) defined above is a Banach space.

Arguing similarly to the proof of Theorem 1 and applying the fixed point theorem in the Banach space \(\widetilde{X}\), we conclude that \(A_0u \in \widetilde{X}\), that is, \(t^{1-\alpha }A_0u' \in C([0,T];L^2(\varOmega ))\). Using the inclusion \({\mathcal {D}}(A_0) \subset C(\overline{\varOmega })\) in the spatial dimensions \(d=1,2,3\), the Sobolev embedding theorem yields

$$\begin{aligned} u' \in C(\overline{\varOmega } \times (0,T]), \quad \Vert A_0u'(t)\Vert \le CDt^{\alpha -1}, \quad 0\le t\le T. \end{aligned}$$
(4.11)

Now we proceed to the estimation of \(A_0^2u(t)\). Since \(\frac{d}{ds}(-A_0^{-1}S(s)) = K(s)\) for \(0<s<T\) by (2.7), the integration by parts yields

$$\begin{aligned}&\int ^t_0 K(t-s)Q(s)u(s) ds = \int ^t_0 K(s)Q(t-s)u(t-s) ds\nonumber \\&\quad = \left[ -A_0^{-1}S(s)Q(t-s)u(t-s)\right] ^{s=t}_{s=0}\nonumber \\&\qquad - \int ^t_0 A_0^{-1}S(s)(Q'(t-s)u(t-s)+Q(t-s)u'(t-s)) ds\nonumber \\&\quad = A_0^{-1}Q(t)u(t) - A_0^{-1}S(t)Q(0)u(0)\nonumber \\&\qquad - \int ^t_0 A_0^{-1}S(s)(Q'(t-s)u(t-s)+Q(t-s)u'(t-s)) ds, \quad 0<t<T. \end{aligned}$$
(4.12)

Applying the Lebesgue convergence theorem and the estimate \(\vert E_{\alpha ,1}(\eta )\vert \le \frac{C}{1+\eta },\ \eta >0\) (Theorem 1.6 in [27]), we readily reach

$$\begin{aligned} \Vert S(t)a - a\Vert ^2 = \sum _{n=1}^{\infty }\vert (a,\varphi _n)\vert ^2 (E_{\alpha ,1}(-\lambda _nt^{\alpha }) - 1)^2 \, \longrightarrow \, 0 \end{aligned}$$

as \(t \rightarrow \infty \) for \(a \in L^2(\varOmega )\).

Hence, \(u \in C([0,T];L^2(\varOmega ))\) and \(\lim _{t\downarrow 0} \Vert (S(t)-1)a\Vert = 0\) and thus

$$\begin{aligned} \lim _{s\downarrow 0} S(s)Q(t-s)u(t-s) = S(0)Q(t)u(t) \quad \text{ in } L^2(\varOmega ) \end{aligned}$$

and

$$\begin{aligned} \lim _{s\uparrow t} S(s)Q(t-s)u(t-s) = S(t)Q(0)u(0) \quad \text{ in } L^2(\varOmega ), \end{aligned}$$

which justify the last equality in the formula (4.12).

Thus, in terms of (4.12), the representation (2.13) can be rewritten in the form

$$\begin{aligned}{} & {} A_0^2(u(t) - A_0^{-1}Q(t)u(t)) = A_0^2G(t) -A_0S(t)Q(0)u(0)\nonumber \\{} & {} - \int ^t_0 A_0^{\frac{1}{2}}S(s)A_0^{\frac{1}{2}} (Q'(t-s)u(t-s) + Q(t-s)u'(t-s)) ds, \quad 0<t<T.\nonumber \\ \end{aligned}$$
(4.13)

Since \(u(0) = a \in C^{\infty }_0(\varOmega )\) and \(F \in C^{\infty }_0(\varOmega \times (0,T))\), in view of (1.2) we have the inclusions

$$\begin{aligned}{} & {} A_0^2G(\cdot ) \in C([0,T];L^2(\varOmega )), \ A_0S(t)Q(0)u(0)\\{} & {} \quad = S(t)(A_0Q(0)a) \in C([0,T];L^2(\varOmega )). \end{aligned}$$

Now we use the conditions (1.2) and (2.8) and repeat the arguments employed for derivation of (4.5) by means of (4.6) and (4.11) to obtain the estimates

$$\begin{aligned}&\left\| \int ^t_0 A_0^{\frac{1}{2}}S(s)A_0^{\frac{1}{2}} (Q'(t-s)u(t-s) + Q(t-s)u'(t-s)) ds \right\| \\&\quad \le C\int ^t_0 s^{-\frac{1}{2}\alpha }\Vert Q'(t-s)u(t-s) + Q(t-s)u'(t-s) \Vert _{H^1(\varOmega )} ds\\&\quad \le C\int ^t_0 s^{-\frac{1}{2}\alpha }(\Vert A_0u'(t-s)\Vert + \Vert A_0u(t-s)\Vert ) ds \le Ct^{\frac{1}{2}\alpha }D \end{aligned}$$

and the inclusion

$$\begin{aligned} -\int ^t_0 A_0^{\frac{1}{2}}S(s)A_0^{\frac{1}{2}}(Q'(t-s)u(t-s) + Q(t-s)u'(t-s)) ds \in C([0,T];L^2(\varOmega )). \end{aligned}$$

Therefore,

$$\begin{aligned} A_0^2(u(t) - A_0^{-1}Q(t)u(t)) = A_0(A_0u(t) - Q(t)u(t)) \in C([0,T];L^2(\varOmega )), \end{aligned}$$

that is,

$$\begin{aligned} A_0u(t) - Q(t)u(t) \in C([0,T]; {\mathcal {D}}(A_0)) \subset C([0,T];H^2(\varOmega )). \end{aligned}$$

On the other hand, the estimate (4.9) implies \(Q(t)u(t) \in C([0,T];H^2(\varOmega ))\) and we obtain

$$\begin{aligned} A_0u(t) \in C([0,T];H^2(\varOmega )). \end{aligned}$$
(4.14)

For further arguments, we define the Schauder spaces \(C^{\theta }(\overline{\varOmega })\) and \(C^{2+\theta }(\overline{\varOmega })\) with \(0<\theta <1\) (see, e.g., [7, 14]) as follows: A function w is said to belong to the space \(C^{\theta }(\overline{\varOmega })\) if

$$\begin{aligned} \sup _{x, x'\in \varOmega , \, x \ne x'} \frac{\vert w(x) - w(x')\vert }{\vert x-x'\vert ^{\theta }} < \infty . \end{aligned}$$

For \(w \in C^{\theta }(\overline{\varOmega })\), we define the norm

$$\begin{aligned} \Vert w\Vert _{C^{\theta }(\overline{\varOmega })} := \Vert w\Vert _{C(\overline{\varOmega })} + \sup _{x, x'\in \varOmega , \, x \ne x'} \frac{\vert w(x) - w(x')\vert }{\vert x-x'\vert ^{\theta }} \end{aligned}$$

and for \(w\in C^{2+\theta }(\overline{\varOmega })\), the norm is given by

$$\begin{aligned} \Vert w\Vert _{C^{2+\theta }(\overline{\varOmega })} := \Vert w\Vert _{C^2(\overline{\varOmega })} + \sum _{\vert \tau \vert =2} \sup _{x, x'\in \varOmega , \, x \ne x'} \frac{\vert \partial _x^{\tau }w(x) - \partial _x^{\tau }w(x')\vert }{\vert x-x'\vert ^{\theta }}. \end{aligned}$$

In the last formula, the notations \(\tau := (\tau _1,\ldots , \tau _d) \in ({\mathbb {N}}\cup \{0\})^d\), \(\partial _x^{\tau }:= \partial _1^{\tau _1}\cdots \partial _d^{\tau _d}\), and \(\vert \tau \vert := \tau _1 + \cdots + \tau _d\) are employed.

For \(d=1,2,3\), the Sobolev embedding theorem says that \(H^2(\varOmega ) \subset C^{\theta }(\overline{\varOmega })\) with some \(\theta \in (0,1)\) ( [1]).

Therefore, in view of (4.14), we obtain the inclusion \(h:= A_0u(\cdot ,t) \in C^{\theta }(\overline{\varOmega })\) for each \(t \in [0,T]\). Now we apply the Schauder estimate (see, e.g., [7] or [14]) for solutions to the elliptic boundary value problem

$$\begin{aligned} A_0u(\cdot ,t) = h\in C^{\theta }(\overline{\varOmega }) \quad \text{ in } \varOmega \end{aligned}$$

with the boundary condition \(\partial _{\nu _A}u(\cdot ,t) + \sigma (\cdot )u(\cdot ,t) = 0\) on \(\partial \varOmega \) to reach the inclusion

$$\begin{aligned} u \in C([0,T]; C^{2+\theta }(\overline{\varOmega })). \end{aligned}$$

This inclusion and (4.11) yield the conclusion \(u \in C([0,T];C^2(\overline{\varOmega }))\) and \(t^{1-\alpha }\partial _tu \in C([0,T];C(\overline{\varOmega }))\) of the lemma.

Finally we prove that \(\lim _{t\rightarrow 0} \Vert u(t) - a \Vert = 0\). By (2.8), we have

$$\begin{aligned}&\left\| \int ^t_0 K(t-s)h(s) ds\right\| \le \int ^t_0 \Vert K(t-s)h(s) \Vert ds \le C\int ^t_0 (t-s)^{\alpha -1} \Vert h(s)\Vert ds\\&\quad \le \frac{Ct^{\alpha }}{\alpha }\Vert h\Vert _{L^{\infty }(0,T;L^2(\varOmega ))}, \end{aligned}$$

and so

$$\begin{aligned} \lim _{t\rightarrow 0} \int ^t_0 K(t-s)h(s) ds = 0 \quad \hbox { in}\ L^2(\varOmega ) \end{aligned}$$
(4.15)

for each \(h \in L^{\infty }(0,T;L^2(\varOmega ))\). Therefore by the regularity \(u \in C([0,T];C^2(\overline{\varOmega }))\), we see that

$$\begin{aligned} \lim _{t\rightarrow 0} \left( \int ^t_0 K(t-s)F(s) ds + Ru(t) \right) = 0 \quad \hbox { in}\ L^2(\varOmega ), \end{aligned}$$

where R is defined in (2.14). Moreover, for justifying (4.12), we have already proved \(\lim _{t\rightarrow 0} \Vert S(t)a - a\Vert = 0\) for \(a \in L^2(\varOmega )\). Thus the proof of Lemma 4 is complete. \(\square \)

II. Second part of the proof.

In this part, we weaken the regularity conditions posed on the solution u to (3.3) in Lemma 3 and prove the same results provided that \(u\in L^2(0,T;H^2(\varOmega ))\) and \(u-a \in H_{\alpha }(0,T;L^2(\varOmega ))\), under the assumption that \(\min \limits _{(x,t)\in \overline{\varOmega }\times [0,T]} b_0(x,t) > 0\) is sufficiently large.

Let \(F \in L^2(0,T;L^2(\varOmega ))\) and \(a\in H^1(\varOmega )\) satisfy the inequalities \(F(x,t)\ge 0,\ (x,t)\in \varOmega \times (0,T)\) and \(a(x)\ge 0,\ x\in \varOmega \).

Now we apply the standard mollification procedure (see, e.g., [1]) and construct the sequences \(F_n \in C^{\infty }_0(\varOmega \times (0,T))\) and \(a_n \in C^{\infty }_0(\varOmega )\), \(n\in {\mathbb {N}}\) such that \(F_n(x,t)\ge 0,\ (x,t)\in \varOmega \times (0,T)\) and \(a_n(x)\ge 0,\ x\in \varOmega \), \(n\in {\mathbb {N}}\) and \(\lim _{n\rightarrow \infty } \Vert F_n-F\Vert _{L^2(0,T;L^2(\varOmega ))} = 0\) and \(\lim _{n\rightarrow \infty }\Vert a_n-a\Vert _{H^1(\varOmega )} = 0\). Then Lemma 4 yields the inclusion

$$\begin{aligned} u(F_n,a_n) \in C([0,T];C^2(\overline{\varOmega })), \quad t^{1-\alpha }\partial _tu(F_n,a_n) \in C([0,T];C(\overline{\varOmega })), \quad n\in {\mathbb {N}}\end{aligned}$$

and thus Lemma 3 ensures the inequalities

$$\begin{aligned} u(F_n,a_n)(x,t) \ge 0 ,\ \ (x,t)\in \varOmega \times (0,T), \, n\in {\mathbb {N}}. \end{aligned}$$
(4.16)

Since Theorem 1 holds true for the initial-boundary value problem (3.3) with F and a replaced by \(F-F_n\) and \(a-a_n\), respectively, we have

$$\begin{aligned}{} & {} \Vert u(F,a) - u(F_n,a_n)\Vert _{L^2(0,T; H^2(\varOmega ))}\\{} & {} \quad \le C(\Vert a-a_n\Vert _{H^1(\varOmega )} + \Vert F-F_n\Vert _{L^2(0,T;L^2(\varOmega ))}) \, \rightarrow \, 0 \end{aligned}$$

as \(n\rightarrow \infty \). Therefore, we can choose a subsequence \(m(n)\in {\mathbb {N}}\) such that \(u(F,a)(x,t) = \lim _{m(n)\rightarrow \infty } u(F_{m(n)},a_{m(n)})(x,t)\) for almost all \((x,t) \in \varOmega \times (0,T)\). Then the inequality (4.16) leads to the desired result, namely, to the inequality \(u(F,a)(x,t) \ge 0\) for almost all \((x,t) \in \varOmega \times (0,T)\).

III. Third part of the proof.

Let the inequalities \(a(x)\ge 0,\ x\in \varOmega \) and \(F(x,t)\ge 0,\ (x,t)\in \varOmega \times (0,T)\) hold true for \(a\in H^1(\varOmega )\) and \(F \in L^2(0,T;L^2(\varOmega ))\) and let \(u=u(F,a) \in L^2(0,T;H^2(\varOmega ))\) is a solution to the problem (2.4). In order to complete the proof of Theorem 2, we have to demonstrate the non-negativity of the solution without any assumptions on the sign of the zeroth-order coefficient.

First, the zeroth-order coefficient \(b_0(x,t)\) in the definition (3.1) of the operator \(-A_1\) is set to a constant \(b_0>0\) that is assumed to be sufficiently large. In this case, the initial-boundary value problem (2.4) can be rewritten as follows:

$$\begin{aligned} \left\{ \begin{array}{rl} &{} \partial _t^{\alpha }(u-a) + A_1u = (b_0+c(x,t))u + F(x,t), \quad (x,t) \in \varOmega \times (0,T), \\ &{} \partial _{\nu _A}u + \sigma u = 0 \quad \hbox { on}\ \partial \varOmega \times (0,T). \end{array}\right. \end{aligned}$$
(4.17)

In what follows, we choose sufficiently large \(b_0>0\) such that \(b_0 \ge \Vert c\Vert _{C(\overline{\varOmega } \times [0,T])}\).

In the previous parts of the proof, we already interpreted the solution u as a unique fixed point for the equation (4.3). Now let us construct an appropriate approximating sequence \(u_n\), \(n\in {\mathbb {N}}\) for the fixed point u. First we set \(u_0(x,t) := 0\) for \((x,t) \in \varOmega \times (0,T)\) and \(u_1(x,t) = a(x) \ge 0, \ (x,t) \in \varOmega \times (0,T)\). Then we define a sequence \(u_{n+1},\ n\in {\mathbb {N}}\) of solutions to the following initial-boundary value problems with the given \(u_n\):

$$\begin{aligned} \left\{ \begin{array}{rl} &{}\partial _t^{\alpha }(u_{n+1}-a) + A_1u_{n+1} = (b_0+c(x,t))u_n + F(x,t) \quad \text{ in } \varOmega \times (0,T),\\ &{} \partial _{\nu _A}u_{n+1} + \sigma u_{n+1} = 0 \quad \text{ on } \partial \varOmega \times (0,T),\\ &{} u_{n+1} - a \in H_{\alpha }(0,T;L^2(\varOmega )), \quad n\in {\mathbb {N}}. \end{array}\right. \end{aligned}$$
(4.18)

First we show that

$$\begin{aligned} u_n(x,t) \ge 0, \quad (x,t) \in \varOmega \times (0,T), \quad n\in {\mathbb {N}}. \end{aligned}$$
(4.19)

Indeed, the inequality (4.19) holds for \(n=1\). Now we assume that \(u_n(x,t) \ge 0,\ (x,t)\in \varOmega \times (0,T)\). Then \((b_0+c(x,t))u_n(x,t) + F(x,t) \ge 0,\ (x,t)\in \varOmega \times (0,T)\), and thus by the results established in the second part of the proof of Theorem 2, we obtain the inequality \(u_{n+1}(x,t) \ge 0,\ (x,t)\in \varOmega \times (0,T)\). By the principle of mathematical induction, the inequality (4.19) holds true for all \(n\in {\mathbb {N}}\).

Now we rewrite the problem (4.18) as

$$\begin{aligned} \partial _t^{\alpha }(u_{n+1}(t) - a) + A_0u_{n+1}(t)= & {} (Q(t)u_{n+1}-(c(t)+b_0)u_{n+1}) \\{} & {} + (b_0+c(t))u_n + F, \end{aligned}$$

where \(A_0\) and Q(t) are defined by (2.5) and (2.14), respectively. Next we estimate \(w_{n+1}:= u_{n+1} - u_n\). By the relation (4.18), \(w_{n+1}\) is a solution to the problem

$$\begin{aligned} \left\{ \begin{array}{rl} &{}\partial _t^{\alpha }w_{n+1} + A_0w_{n+1} = (Q(t)w_{n+1} - (c(t)+b_0)w_{n+1}) + (b_0+c(x,t))w_n \\ &{} \qquad \qquad \quad \text{ in } \varOmega \times (0,T),\\ &{} \partial _{\nu _A}w_{n+1} + \sigma w_{n+1} = 0 \quad \text{ on } \partial \varOmega \times (0,T),\\ &{} w_{n+1} \in H_{\alpha }(0,T;L^2(\varOmega )), \quad n\in {\mathbb {N}}. \end{array}\right. \end{aligned}$$

In terms of the operator K(t) defined by (2.7), acting similarly to our analysis of the fixed point equation (4.3), we obtain the integral equation

$$\begin{aligned} w_{n+1}(t)&= \int ^t_0 K(t-s)(Qw_{n+1})(s) ds - \int ^t_0 K(t-s)(c(s)+b_0)w_{n+1}(s) ds\\&\quad + \int ^t_0 K(t-s)(b_0+c(s))w_n(s) ds, \quad 0<t<T, \end{aligned}$$

which leads to the inequalities

$$\begin{aligned}&\Vert A_0^{\frac{1}{2}}w_{n+1}(t)\Vert \le \int ^t_0 \Vert A_0^{\frac{1}{2}}K(t-s)\Vert \Vert Q(s)w_{n+1}(s)\Vert ds\\&\qquad + \int ^t_0 \Vert A_0^{\frac{1}{2}}K(t-s)\Vert \Vert (c(s)+b_0)w_{n+1}(s)\Vert ds\\&\qquad + \int ^t_0 \Vert A_0^{\frac{1}{2}}K(t-s)\Vert \Vert (b_0+c(s))w_n(s)\Vert ds\\&\quad \le C\int ^t_0 (t-s)^{\frac{1}{2}\alpha -1} \Vert A_0^{\frac{1}{2}}w_{n+1}(s)\Vert ds\\&\quad + C\int ^t_0 (t-s)^{\frac{1}{2}\alpha -1}\Vert A_0^{\frac{1}{2}}w_n(s)\Vert ds \quad \text{ for } 0<t<T\text{. } \end{aligned}$$

For their derivation, we used the norm estimates

$$\begin{aligned} \Vert Q(s)w_{n+1}(s)\Vert \le C\Vert w_{n+1}(s)\Vert _{H^1(\varOmega )} \le C\Vert A_0^{\frac{1}{2}}w_{n+1}(s)\Vert \end{aligned}$$

and

$$\begin{aligned} \Vert (c(s)+b_0)w_\ell (s)\Vert \le C\Vert w_\ell (s)\Vert _{H^1(\varOmega )} \le C\Vert A_0^{\frac{1}{2}}w_\ell (s)\Vert , \quad \ell =n, n+1 \end{aligned}$$

that hold true under the conditions (1.2). Thus we arrive at the integral inequality

$$\begin{aligned}&\Vert A_0^{\frac{1}{2}}w_{n+1}(t)\Vert \le C\int ^t_0 (t-s)^{\frac{1}{2}\alpha -1} \Vert A_0^{\frac{1}{2}}w_{n+1}(s)\Vert ds\\&\quad + C\int ^t_0 (t-s)^{\frac{1}{2}\alpha -1} \Vert A_0^{\frac{1}{2}}w_n(s)\Vert ds, \quad 0<t<T. \end{aligned}$$

The generalized Gronwall inequality yields now the estimate

$$\begin{aligned}&\Vert A_0^{\frac{1}{2}}w_{n+1}(t)\Vert \le C\int ^t_0 (t-s)^{\frac{1}{2}\alpha -1}\Vert A_0^{\frac{1}{2}}w_n(s)\Vert ds\\&\quad + C\int ^t_0 (t-s)^{\frac{1}{2}\alpha -1} \left( \int ^s_0 (s-\xi )^{\frac{1}{2}\alpha -1} \Vert A_0^{\frac{1}{2}}w_n(\xi )\Vert d\xi \right) ds. \end{aligned}$$

The second term at the right-hand side of the last inequality can be represented as follows:

$$\begin{aligned}&\int ^t_0 (t-s)^{\frac{1}{2}\alpha -1} \left( \int ^s_0 (s-\xi )^{\frac{1}{2}\alpha -1}\Vert A_0^{\frac{1}{2}}w_n(\xi )\Vert d\xi \right) ds\\&\quad = \int ^t_0 \Vert A_0^{\frac{1}{2}}w_n(\xi )\Vert \left( \int ^t_{\xi } (t-s)^{\frac{1}{2}\alpha -1} (s-\xi )^{\frac{1}{2}\alpha -1} ds \right) d\xi \\&\quad = \frac{\varGamma \left( \frac{1}{2}\alpha \right) \varGamma \left( \frac{1}{2}\alpha \right) }{\varGamma (\alpha )}\int ^t_0 (t-\xi )^{\alpha -1}\Vert A_0^{\frac{1}{2}}w_n(\xi )\Vert d\xi \\&\quad = \frac{\varGamma \left( \frac{1}{2}\alpha \right) ^2}{\varGamma (\alpha )}T^{\frac{1}{2}\alpha } \int ^t_0 (t-s)^{\frac{1}{2}\alpha -1}\Vert A_0^{\frac{1}{2}}w_n(s)\Vert ds. \end{aligned}$$

Thus, we can choose a constant \(C>0\) depending on \(\alpha \) and T, such that

$$\begin{aligned} \Vert A_0^{\frac{1}{2}}w_{n+1}(t)\Vert \le C\int ^t_0 (t-\xi )^{\frac{1}{2}\alpha -1}\Vert A_0^{\frac{1}{2}}w_n(s)\Vert ds, \quad 0<t<T,\, n\in {\mathbb {N}}.\qquad \end{aligned}$$
(4.20)

Recalling that

$$\begin{aligned} \int ^t_0 (t-s)^{\frac{1}{2}\alpha -1}\eta (s)ds = \varGamma \left( \frac{1}{2}\alpha \right) (J^{\frac{1}{2}\alpha }\eta )(t), \quad t>0, \end{aligned}$$

and setting \(\eta _n(t):= \Vert A_0^{\frac{1}{2}}w_n(t)\Vert \), we can rewrite (4.20) in the form

$$\begin{aligned} \eta _{n+1}(t) \le C\varGamma \left( \frac{1}{2}\alpha \right) (J^{\frac{1}{2}\alpha }\eta _n)(t), \quad 0<t<T, \, n\in {\mathbb {N}}. \end{aligned}$$
(4.21)

Since the Riemann-Liouville integral \(J^{\frac{1}{2}\alpha }\) preserves the sign and the semi-group property \(J^{\beta _1}(J^{\beta _2}\eta )(t) = J^{\beta _1+\beta _2}\eta (t)\) is valid for any \(\beta _1, \beta _2 > 0\), applying the inequality (4.21) repeatedly, we obtain the estimates

$$\begin{aligned}&\eta _n(t) \le \left( C\varGamma \left( \frac{1}{2}\alpha \right) \right) ^{n-1} (J^{(n-1)\frac{\alpha }{2}}\eta _1)(t) \\&\quad = \frac{\left( C\varGamma \left( \frac{1}{2}\alpha \right) \right) ^{n-1}}{\varGamma \left( \frac{\alpha }{2}(n-1)\right) } \left( \int ^t_0 (t-s)^{(n-1)\frac{1}{2}\alpha -1} ds\right) \Vert A_0^{\frac{1}{2}}a\Vert \\&\quad = \frac{\left( C\varGamma \left( \frac{1}{2}\alpha \right) \right) ^{n-1}}{\varGamma \left( \frac{\alpha }{2}(n-1)\right) } \frac{t^{(n-1)\frac{1}{2}\alpha }}{(n-1)\frac{1}{2}\alpha } \Vert A_0^{\frac{1}{2}}a\Vert \le C_1\frac{\left( C\varGamma \left( \frac{1}{2}\alpha \right) T^{\frac{\alpha }{2}} \right) ^{n-1}}{\varGamma \left( \frac{\alpha }{2}(n-1)\right) }. \end{aligned}$$

The known asymptotic behavior of the gamma function justifies the relation

$$\begin{aligned} \lim _{n\rightarrow \infty } \frac{\left( C\varGamma \left( \frac{1}{2}\alpha \right) T^{\frac{\alpha }{2}} \right) ^{n-1}}{\varGamma \left( \frac{\alpha }{2}(n-1)\right) } = 0. \end{aligned}$$

Thus we have proved that the sequence \(u_N = w_0 + \cdots + w_N\) converges to the solution u in \(L^{\infty }(0,T;H^1(\varOmega ))\) as \(N \rightarrow \infty \). Therefore, we can choose a subsequence \(m(n)\in {\mathbb {N}}\) such that \(\lim _{m(n)\rightarrow \infty } u_{m(n)}(x,t) = u(x,t)\) for almost all \((x,t) \in \varOmega \times (0,T)\). This statement in combination with the inequality (4.19) means that \(u(x,t) \ge 0\) for almost all \((x,t) \in \varOmega \times (0,T)\). The proof of Theorem 2 is completed. \(\square \)

Now let us fix a source function \(F = F(x,t) \ge 0,\ (x,t)\in \varOmega \times (0,T)\) and an initial value \(a \in H^1(\varOmega )\) in the initial-boundary value problem (2.4) and denote by \(u(c,\sigma ) = u(c,\sigma )(x,t)\) the solution to the problem (2.4) with the functions \(c=c(x,t)\) and \(\sigma = \sigma (x)\). Then the following comparison property regarding the coefficients c and \(\sigma \) is valid:

Theorem 3

Let \(a\in H^1(\varOmega )\) and \(F \in L^2(\varOmega \times (0,T))\) and the inequalities \(a(x)\ge 0,\ x\in \varOmega \) and \(F(x,t)\ge 0,\ (x,t)\in \varOmega \times (0,T)\) hold true.

  1. (i)

    Let \(c_1, c_2 \in C^1([0,T]; C^1(\overline{\varOmega })) \cap C([0,T];C^2(\overline{\varOmega }))\) and \(c_1(x,t) \ge c_2(x,t)\) for \((x,t)\in \varOmega \). Then \(u(c_1,\sigma )(x,t) \ge u(c_2,\sigma )(x,t)\) in \(\varOmega \times (0,T)\).

  2. (ii)

    Let \(c(x,t) < 0,\ (x,t) \in \varOmega \times (0,T)\) and a constant \(\sigma _0>0\) be arbitrary and fixed. If the smooth functions \(\sigma _1, \sigma _2\) on \(\partial \varOmega \) satisfy the conditions

    $$\begin{aligned} \sigma _2(x) \ge \sigma _1(x) \ge \sigma _0,\ x\in \partial \varOmega , \end{aligned}$$

    then the inequality \(u(c,\sigma _1) \ge u(c, \sigma _2),\ x\in \varOmega \times (0,T)\) holds true.

Proof

We start with a proof of the statement (i). Because \(a(x)\ge 0,\ x\in \varOmega \) and \(F(x,t)\ge 0,\ (x,t)\in \varOmega \times (0,T)\), Theorem 2 yields the inequality \(u(c_2,\sigma )(x,t)\ge 0,\ (x,t)\in \varOmega \times (0,T)\). Setting \(u(x,t):= u(c_1,\sigma )(x,t) - u(c_2,\sigma )(x,t)\) for \((x,t) \in \varOmega \times (0,T)\), we obtain

$$\begin{aligned} \left\{ \begin{array}{rl} &{} \partial _t^{\alpha }u - \sum _{i,j=1}^d\partial _i(a_{ij}\partial _ju) - \sum _{j=1}^d b_j \partial _ju\\ -&{} c_1(x,t)u = (c_1-c_2)u(c_2,\sigma )(x,t) \quad \text{ in } \varOmega \times (0,T), \\ &{} \partial _{\nu _A}u + \sigma u = 0 \quad \text{ on } \partial \varOmega ,\\ &{} u \in H_{\alpha }(0,T;L^2(\varOmega )). \end{array}\right. \end{aligned}$$

Since \(u(c_2,\sigma )(x,t) \ge 0\) and \((c_1-c_2)(x,t) \ge 0\) for \((x,t) \in \varOmega \times (0,T)\), Theorem 2 leads to the estimate \(u(x,t)\ge 0\) for \((x,t) \in \varOmega \times (0,T)\), which is equivalent to the inequality \(u(c_1,\sigma )(x,t) \ge u(c_2,\sigma )(x,t)\) for \((x,t) \in \varOmega \times (0,T)\) and the statement (i) is proved.

Now we proceed to the proof of the statement (ii). Similarly to the procedure applied for the second part of the proof of Theorem 2, we choose the sequences \(F_n \ge 0\), \(F_n \in C^{\infty }_0(\varOmega \times (0,T))\) and \(a_n \ge 0\), \(a_n \in C^{\infty }_0(\varOmega )\), \(n\in {\mathbb {N}}\) such that \(F_n \rightarrow F\) in \(L^2(\varOmega \times (0,T))\) and \(a_n \rightarrow a\) in \(H^1(\varOmega )\). Let \(u_n\), \(v_n\) be the solutions to the initial-boundary value problem (2.4) with \(F=F_n\), \(a=a_n\) and the coefficients \(\sigma _1\) and \(\sigma _2\) in the boundary condition, respectively. According to Lemma 4, the inclusions \(v_n, u_n \in C(\overline{\varOmega } \times [0,T])\) and \(t^{1-\alpha }\partial _tv_n, \, t^{1-\alpha }\partial _tu_n \in C([0,T];C(\overline{\varOmega }))\), \(n\in {\mathbb {N}}\) hold true and thus Theorem 2 yields

$$\begin{aligned} v_n(x,t) \ge 0, \quad (x,t) \in \partial \varOmega \times (0,T). \end{aligned}$$
(4.22)

Moreover, the relation

$$\begin{aligned} \lim _{n\rightarrow \infty }\Vert u_n - u(c,\sigma _1)\Vert _{L^2(0,T;L^2(\varOmega ))} = \lim _{n\rightarrow \infty }\Vert v_n - u(c,\sigma _2)\Vert _{L^2(0,T;L^2(\varOmega ))} = 0\nonumber \\ \end{aligned}$$
(4.23)

follows from Theorem 1. Let us now define an auxiliary function \(w_n:= u_n - v_n\). For this function, the inclusions

$$\begin{aligned} t^{1-\alpha }\partial _tw_n \in C([0,T];C(\overline{\varOmega })), \quad w_n \in C([0,T];C^2(\overline{\varOmega })), \quad n\in {\mathbb {N}}\end{aligned}$$
(4.24)

hold true. Furthermore, it is a solution to the initial-boundary value problem

$$\begin{aligned} \left\{ \begin{array}{rl} &{} \partial _t^{\alpha }w_n + Aw_n = 0 \quad \text{ in } \varOmega \times (0,T),\\ &{} \partial _{\nu _A}w_n + \sigma _1w_n = (\sigma _2-\sigma _1)v_n \quad \text{ on } \partial \varOmega \times (0,T),\\ &{} w_n(x,\cdot ) \in H_{\alpha }(0,T) \quad \text{ for } \text{ almost } \text{ all } x\in \varOmega . \end{array}\right. \end{aligned}$$
(4.25)

The inequalities (4.22) and \(\sigma _2(x) \ge \sigma _1(x), \ x\in \partial \varOmega \) lead to the estimate

$$\begin{aligned} \partial _{\nu _A}w_n + \sigma _1w_n \ge 0 \quad \text{ on } \partial \varOmega \times (0,T). \end{aligned}$$
(4.26)

To finalize the proof of the theorem, a variant of Lemma 3 formulated below will be employed.

Lemma 5

Let the elliptic operator \(-A\) be defined by (2.1) and the conditions (1.2) be satisfied. Moreover, let the inequality \(c(x,t) < 0\) for \(x \in \overline{\varOmega }\) and \(0\le t \le T\) hold true and there exist a constant \(\sigma _0>0\) such that

$$\begin{aligned} \sigma (x) \ge \sigma _0 \quad \hbox { for all}\ x\in \partial \varOmega . \end{aligned}$$

For \(a \in H^1(\varOmega )\) and \(F\in L^2(\varOmega \times (0,T))\), we further assume that there exists a solution \(u\in C([0,T];C^2(\overline{\varOmega }))\) to the initial-boundary value problem

$$\begin{aligned} \left\{ \begin{array}{rl} &{} \partial _t^{\alpha }(u-a) + Au = F \quad \hbox { in}\ \varOmega \times (0,T), \\ &{} \partial _{\nu _A}u + \sigma (x)u \ge 0 \quad \text{ on } \partial \varOmega \times (0,T), \\ &{} u(x,\cdot ) - a\in H_{\alpha }(0,T) \quad \text{ for } \text{ almost } \text{ all } x\in \varOmega \end{array}\right. \end{aligned}$$

that satisfies the inclusion \(t^{1-\alpha }\partial _tu \in C([0,T];C(\overline{\varOmega }))\). Then the inequalities \(F(x,t) \ge 0,\ (x,t)\in \varOmega \times (0,T)\) and \(a(x)\ge 0, \ \varOmega \) implicate the inequality \(u(x,t) \ge 0,\ (x,t)\in \varOmega \times (0,T)\).

In the formulation of this lemma, at the expense of the extra condition \(\sigma (x) > 0\) on \(\partial \varOmega \), we do not assume that \(\min \limits _{(x,t)\in \overline{\varOmega }\times [0,T]} (-c(x,t))\) is sufficiently large. This is the main difference between the conditions supposed in Lemma 5 and in Lemma 3. The proof of Lemma 5 is much simpler compared to the one of Lemma 3; it will be presented at the end of this section.

Now we complete the proof of Theorem 3. Since \(c(x,t) < 0\) for \((x,t)\in \varOmega \times (0,T)\) and \(\sigma _1(x) \ge \sigma _0 > 0\) on \(\partial \varOmega \) and taking into account the conditions (4.24) and (4.26), we can apply Lemma 5 to the initial-boundary value problem (4.25) and deduce the inequality \(w_n(x,t) \ge 0,\ (x,t)\in \varOmega \times (0,T)\), that is, \(u_n(x,t) \ge v_n(x,t),\ (x,t)\in \varOmega \times (0,T)\) for \(n\in {\mathbb {N}}\). Due to the relation (4.23), we can choose a suitable subsequence of \(w_n,\ n\in {\mathbb {N}}\) and pass to the limit as n tends to infinity thus arriving at the inequality \(u(c,\sigma _1)(x,t) \ge u(c,\sigma _2)(x,t)\) in \(\varOmega \times (0,T)\). The proof of Theorem 3 is completed. \(\square \)

At this point, let us mention a direction for further research in connection with the results formulated and proved in this sections. In order to remove the negativity condition posed on the coefficient \(c=c(x,t)\) in Theorem 3 (ii), one needs a unique existence result for solutions to the initial-boundary value problems of type (2.4) with non-zero Robin boundary condition similar to the one formulated in Theorem 1. There are several works that treat the case of the initial-boundary value problems with non-homogeneous Dirichlet boundary conditions (see, e.g., [31] and the references therein). However, to the best of the authors’ knowledge, analogous results are not available for the initial-boundary value problems with the non-homogeneous Neumann or Robin boundary conditions. Thus, in Theorem 3 (ii), we assumed the condition \(c(x,t)<0,\ (x,t)\in \varOmega \times (0,T)\), although our conjecture is that this result holds true for an arbitrary coefficient \(c=c(x,t)\).

We conclude this section with a proof of Lemma 5 that is simple because in this case we do not need the function \(\psi \) defined as in (3.4).

Proof

First we introduce an auxiliary function as follows:

$$\begin{aligned} \widetilde{w}(x,t):= u(x,t) + \varepsilon (1+t^{\alpha }), \quad x\in \varOmega ,\, 0<t<T. \end{aligned}$$

The inequalities \(c(x,t)<0,\ (x,t)\in \overline{\varOmega } \times [0,T]\) and \(\sigma (x) \ge \sigma _0>0,\ x\in \partial \varOmega \) and the calculations similar to the ones done in the proof of Lemma 3 implicate the inequalities

$$\begin{aligned}{} & {} d_t^{\alpha }\widetilde{w} + A\widetilde{w} = F + \varepsilon \varGamma (\alpha +1) - c(x,t)\varepsilon (1+t^{\alpha }) > 0 \quad \hbox { in}\ \varOmega \times (0,T),\\{} & {} \partial _{\nu _A}\widetilde{w} + \sigma \widetilde{w} = \partial _{\nu _A}u + \sigma u + \sigma \varepsilon (1+t^{\alpha }) \ge \sigma _0\varepsilon \quad \text{ on }\ \partial \varOmega \times (0,T) \end{aligned}$$

and

$$\begin{aligned} \widetilde{w}(x,0) = a(x) + \varepsilon \ge \varepsilon \quad \hbox { in}\ \varOmega . \end{aligned}$$

Based on these inequalities, the same arguments that were employed after the formula (3.10) in the proof of Lemma 3 readily complete the proof of Lemma 5. \(\square \)