1. Introduction
The Schrödinger equation is the fundamental equation in quantum mechanics, since it describes the evolution in time of the wave function’s state
$u(t,x)$ of a quantum particle with mass m:

where
$\hbar$ is the reduced Planck constant, V is a potential, and, as usual,
$\triangle:=\sum_{j=1}^n \partial_{x_j}^2$ is the Laplace operator. Literature on the topic is really wide, and various questions of existence of a unique solution in Sobolev spaces have been studied, see, for instance, [Reference Kenig, Ponce, Rolvung and Vega19] and the references therein. Also in the case of space and time depending coefficients, the literature is extensive, not only for the Schrödinger equation itself but also for the so-called class of Schrödinger type equations with lower order terms Su = f, where

with at least an assumption of continuity with respect to time and an assumption of smoothness in space for the coefficients.
It is indeed well known, see [Reference Ichinose16], that in the case
$c_j(t,x) = c_j(x) \in \mathcal{B}^\infty(\mathbb R^n)$ the Cauchy problem associated to the operator S may be well-posed in
$H^\infty(\mathbb R^n)= \bigcap_{m \in \mathbb R} H^m(\mathbb R^n)$, only if there exist
$M,N \gt 0$ such that

Here,
$\mathcal B^\infty(\mathbb R^n)$ denotes the space of all smooth functions on
$\mathbb R^n$ with uniformly bounded derivatives. Moreover, see [Reference Kajitani and Baba17], if

where
$\langle x\rangle:=(1+|x|^2)^{1/2}$ and
$\sigma \geq 1$, then the Cauchy problem associated to S is well-posed in
$H^m(\mathbb R^n)$ for every
$m \in \mathbb R$ when σ > 1, in
$H^\infty(\mathbb R^n)$ when σ = 1. In the latter case, a finite loss of derivatives appears in the solution. Results of well-posedness in Gevrey classes and Gelfand–Shilov spaces for operator S can be found respectively in [Reference Ascanelli and Cicognani5, Reference Cicognani and Reissig7, Reference Kajitani and Baba17] and [Reference Arias Junior1, Reference Ascanelli and Cappiello4]. A necessary condition for well-posedness in a Gevrey class of index θ > 1 can be found in [Reference Arias Junior, Ascanelli and Cappiello2].
We point out that a result of well-posedness in the Schwartz space of rapidly decreasing functions
$\mathscr S(\mathbb R^n)$ for the Cauchy problem associated to the operator S, which is interesting per se, appears as a byproduct in the present paper, see Theorem 7.1 at Section 7.
This paper concerns the following question: what happens when the coefficients are irregular, that is, if we lose the minimal regularity assumptions on the coefficients and they turn out to be discontinuous functions, or distributions? We focus on the Cauchy problem

for the Schrödinger-type operator given in (1.2) with the following assumption on the lower-order coefficients

and we take initial data satisfying

where, as usual,
$\mathscr{S}'(\mathbb R^n)$ stands for the space of tempered distributions.
In this case, the Cauchy problem (1.4) is ill-posed either in Sobolev-type spaces or in Gevrey-type spaces, see [Reference Arias Junior, Ascanelli and Cappiello2, Reference Ichinose16]. Moreover, due to the remarkable Schwartz impossibility result, the operator S might fail to be defined as an operator acting in the space
$C([0,T]; \mathscr{S}'(\mathbb R^n))$. Likewise, in this general setting, a non-trivial matter is to define what a solution to (1.4) should be.
The concept we are going to deal with is the notion of Schwartz very weak solution, see [Reference Arias Junior, Ascanelli, Cappiello and Garetto3, Reference Garetto13, Reference Garetto and Ruzhansky14]. Very weak solutions have been introduced in [Reference Garetto and Ruzhansky14] to provide a meaningful notion of solution for hyperbolic equations with highly irregular coefficients, namely distributions. See also [Reference Cardona, Chatzakou, Delgado and Ruzhansky6, Reference Discacciati, Garetto and Loizou9, Reference Discacciati, Garetto and Loizou10, Reference Garetto13, Reference Garetto and Sabitbek15, Reference Munoz, Ruzhansky and Tokmagambetov21]. The main idea is to replace the original equation with a family of regularized equations depending on the parameter
$\varepsilon$ and to investigate the
$\varepsilon$-behaviour of the net of corresponding solutions. While in [Reference Garetto and Ruzhansky14] very weak solutions were modelled on Gevrey classes, here, for the analysis of the Schrödinger type equations we are interested in, we will work with the space
$\mathscr{S}(\mathbb R^n)$ of smooth and rapidly decreasing functions (Schwartz functions).
To construct a
$\mathscr{S}$-very weak solution to (1.4), we take initial data
$f \in C([0,T]; \mathscr{S}'(\mathbb R^{n}))$ and
$g \in \mathscr{S}'(\mathbb R^{n})$, and we follow an approach similar to the one introduced in [Reference Garetto and Ruzhansky14] as follows:
(1) we regularize the tempered distributions
$c_j(t),f(t),g$ by a special net of Schwartz mollifiers parametrized by a positive scale
$\omega(\varepsilon)$ converging to 0 as
$\varepsilon\rightarrow0^+$; this regularisation produces a family of Schwartz functions indexed by
$\varepsilon$ > 0 and converging in
$\mathscr S'(\mathbb R^n)$ to the original tempered distributions as
$\varepsilon\rightarrow0^+$, see the Appendix (Section A) for the details;
(2) we then obtain a family of regularized Cauchy problems
(1.7)\begin{equation} \begin{cases} S_{\varepsilon} v(t,x) = f_\epsilon(t,x), \quad t \in [0,T],\, x \in \mathbb R^{n}, \\ v(0,x) = g_\epsilon(x), \quad \quad \,\, \, x \in \mathbb R^{n}, \end{cases} \end{equation}
where
(1.8)\begin{equation} S_{\varepsilon} = D_t + \sum_{j=1}^{n}D^{2}_{x_j} + \sum_{j=1}^{n} c_{j,\varepsilon}(t,x) D_{x_j} + c_{0,\varepsilon}(t,x). \end{equation}
These problems have Cauchy data and coefficients that are continuous functions in t and Schwartz functions in x, so they fulfil assumption (1.3) and we know that the Cauchy problems (1.7) admit unique solutions
$u_\epsilon\in C([0,T],H^\infty(\mathbb R^n)).$ But we search for well-posedness in
$\mathscr S$, so we prove in Theorem 3.2 of Section 3 that we have in fact
$u_\varepsilon\in C([0,T],\mathscr S(\mathbb{R}^n)).$ To conclude this, we study an auxiliary Cauchy problem, given by a suitable conjugation by powers of
$\langle x\rangle$, in order to conclude rapid decay for the solution
$u_\varepsilon$. A careful use of microlocal techniques and a sharp use of the energy method for evolution equations are needed in the proof. In addition, we provide a qualitative analysis of the solutions
$(u_\varepsilon)_\varepsilon$, by giving a suitable energy estimate.
(3) Using the energy estimate, we study the behaviour of the net
$(u_\varepsilon)_\varepsilon$ as
$\varepsilon\rightarrow0^+$, and we prove that the net
$(u_\varepsilon)_{\varepsilon\in(0,1\rbrack}$ defines a
$\mathscr S$-very weak solution to the Cauchy problem (1.4), that is the net is
$\mathscr S$-moderate and the solution is unique modulo negligible changes, see Section 2 for precise definitions.
(4) Finally, we show in Section 6 that the result obtained is consistent with the classical theory: in the case of regular coefficients and Schwartz Cauchy data, the net converges in
$\mathscr S$ to the unique classical solution, and the limit is independent of the regularization used.
This paper is a second step in the study of the Cauchy problem (1.4) with irregular coefficients. The first one was made in the recent paper [Reference Arias Junior, Ascanelli, Cappiello and Garetto3], where we dealt with a similar problem. There, we analysed the coefficients of the type

where aj are distributions with compact support in
$\mathcal{E}'([0,T])$ and
$b_j$ belong to suitable subclasses
$H^{-\infty,j}(\mathbb R^n)$ of
$\mathscr S'(\mathbb R^n)$,
$j\geq 2$, where
$H^{-\infty,j}$ stands for the class of tempered distributions v such that all the derivatives
$\partial_\xi^\beta \hat v(\xi)$ have at most polynomial growth for
$|\beta|\leq j$; the standard regularization of these distributions produces uniformly bounded functions with decay
$\langle x\rangle^{-j}$,
$j\geq 2$, so, by [Reference Kajitani and Baba17], the regularized Cauchy problem turns out to be well posed in every Sobolev space and we prove that the Cauchy problem (1.4) admits a
$H^\infty$-very weak solution.
We remark that combining the ideas of the present manuscript with the ones developed in [Reference Arias Junior, Ascanelli, Cappiello and Garetto3], we might have considered coefficients of the type (1.9) with
$a_j \in \mathcal{E}'([0,T])$ and
$b_j \in \mathscr{S}'(\mathbb R^n)$ and still obtain an analogous of our main result (cf. Theorem 2.7). We decided to consider
$c_j \in C([0,T];\mathscr{S}'(\mathbb R^n))$ in order to obtain a less technical and more elegant result. In addition, the analysis of the regularizations in the space variable is the core of the problem and the most challenging part. Note that, in order to allow tempered distributions in the variable
$x$, we need to use a special kind of regularization (cf. Appendix A) and to change the strategy for solving the regularized problem, which becomes more involved with respect to the one used in [Reference Arias Junior, Ascanelli, Cappiello and Garetto3].
The manuscript is organized as follows. In Section 2, we introduce the notion of a very weak solution of Schwartz type and we state our main result that will be proven in Sections 3–5. Consistency with the classical theory in the case of regular coefficients is the topic of Section 6. The final Section 7 concerns a well-posedness result in the Schwartz space for (1.4) in a regular functional setting. We close the paper with two appendices: Appendix A concerns the regularization of tempered distributions by Schwartz functions, and Appendix B deals with some aspects of the theory of pseudo-differential operators relevant to this paper.
2. Schwartz very weak solutions and main result
The aim of this Section is to state the main result of this manuscript. As we already mentioned in the introduction, due to the Schwartz impossibility result about products of distributions, we need to define a suitable concept of solution. The approach that we shall adopt is the so-called very weak solutions (see [Reference Garetto13, Reference Garetto and Ruzhansky14]).
We consider the operator (1.2) under the hypothesis (1.5) on the coefficients. We are interested in the Cauchy problem (1.4), where the initial data satisfy
$g \in \mathscr{S}'(\mathbb R^{n})$ and
$f \in C([0,T]; \mathscr{S}'(\mathbb R^{n}))$. Once the main problem of the paper is posed, the next step is to define the notion of solution to (1.4) that we are going to deal with. To define it, we shall need some preliminaries.
Let ϕ and ψ be Schwartz functions such that
$\phi(0) = 1 = \widehat{\psi}(0)$. Given a positive scale
$\omega(\varepsilon)$,
$\varepsilon \in (0,1]$, i.e. ω is positive, bounded,
$\omega(\varepsilon) \to 0$ as
$\varepsilon \to 0^{+}$ and
$\omega(\varepsilon) \geq c \varepsilon^{r}$, for some
$c, r \gt 0$, we define

Now, given any
$u \in \mathscr{S}'(\mathbb R^n)$, the regularization

satisfies

See Appendix A at the end of the paper for more details on this kind of regularizations. The next proposition is standard, and for this reason, we are going to omit its proof.
(i) If
$u \in \mathscr{S}(\mathbb R^{n})$ then for any
$\beta \in \mathbb N_0^n$ and
$M \in \mathbb N_0$ there exists
$C_{\beta,M} \gt 0$ such that
\begin{equation*} |\langle x \rangle^{M} \partial^{\beta}_{x} \{\phi(\omega(\varepsilon)x)(\psi_{\omega(\varepsilon)} \ast u)(x)\}| \leq C_{\beta,M}; \end{equation*}
(ii) If
$u \in \mathscr{S}(\mathbb R^{n})$,
$(\partial^{\beta}\phi)(0) = 0$ for all β ≠ 0 and
$\int x^{\alpha}\psi(x) dx = 0$ for all α ≠ 0, then for any
$\beta \in \mathbb N_0^n$ and any
$q, M \in \mathbb N_0$ there exists C > 0 such that
\begin{equation*} |\langle x \rangle^{M} \partial^{\beta}_{x}\{\phi(\omega(\varepsilon) x ) (\psi_{\omega(\varepsilon)} \ast u)(x) - u(x)\}| \leq C (\omega(\varepsilon))^{q}. \end{equation*}
It is convenient to describe
$\mathscr{S}(\mathbb R^n)$ and
$\mathscr{S}'(\mathbb R^n)$ in terms of suitable weighted Sobolev spaces. For
$m, M \in \mathbb R$ we define

where
$\langle D \rangle^m$ stands for the Fourier multiplier given by the symbol
$\langle \xi \rangle^{m}$. It is well-known that
$H^{m,M}(\mathbb R^n)$ are Hilbert spaces and

with equivalent limit type topologies. For more details on weighted Sobolev spaces, see Proposition 2.3 and its corollaries in [Reference Parenti22], cf. also [Reference Cordes8].
Proposition 2.1 and (2.1) (cf. Theorem A.7) motivate the following Definition 2.2.
(i) Let
$(v_{\varepsilon})_{\varepsilon}$ be a net in
$\mathscr{S}(\mathbb R^{n})^{(0,1]}$. We say that the net
$(v_\varepsilon)_\varepsilon$ is
$\mathscr{S}$-moderate if for any
$M \in \mathbb N_0$ and any
$\beta \in \mathbb N_{0}^{n}$ there exists
$N(M,\beta) = N \in \mathbb N_0$ and
$C(M,\beta) = C \gt 0$ such that
\begin{equation*} \sup_{x\in\mathbb R^{n}} \langle x \rangle^{M} |\partial^{\beta}_{x} v_{\varepsilon}(x)| \leq C \varepsilon^{-N}, \quad \forall\, \epsilon\in (0,1]. \end{equation*}
This is equivalent to: for all
$M,m \in \mathbb N_0$ there exists
$N(M,m) = N \in \mathbb N_0$ and
$C(M,m) = C \gt 0$ such that
\begin{equation*} \|v_{\varepsilon}\|_{H^{m,M}(\mathbb R^{n})} \leq C \varepsilon^{-N}, \quad \forall\, \epsilon\in (0,1]. \end{equation*}
(ii) Let
$(v_{\varepsilon})_{\varepsilon}$ be a net in
$\mathscr{S}(\mathbb R^{n})^{(0,1]}$. We say that the net
$(v_{\varepsilon})_{\varepsilon}$ is
$\mathscr{S}$-negligible if for any
$M\in\mathbb N_0$, any
$\beta \in \mathbb N^{n}_{0}$ and any
$q\in\mathbb N_0$ there exist
$C(M,\beta,q) = C \gt 0$ such that
\begin{equation*} \sup_{x \in \mathbb R^{n}}|\langle x \rangle^{M} \partial^{\beta}_{x}\partial^{\beta}_{x} v_{\varepsilon}(x)| \leq C \varepsilon^{q}, \quad \forall\, \epsilon\in (0,1]. \end{equation*}
This is equivalent to: for any
$M,m \in \mathbb N_0$ and any
$q\in\mathbb N_0$ there exists
$C(M,m,q) = C \gt 0$ such that
\begin{equation*} \|v_{\varepsilon}\|_{H^{m,M}(\mathbb R^{n})} \leq C \varepsilon^{q},\quad \forall\, \epsilon\in (0,1]. \end{equation*}
Remark 2.3. We can extend Definition 2.2 to nets in
$\{C([0,T];\mathscr{S}(\mathbb R^{n}))\}^{(0,1]}$ just by asking uniform estimates with respect to the variable t.
Remark 2.4. The notions of Schwartz negligibility and Schwartz moderateness also appear in [Reference Garetto11] (cf. definition 2.8) and [Reference Garetto12] (cf. definition 3.1).
Remark 2.5. Let
$\mathcal{M}_{\mathscr{S}}$ be the set of all
$\mathscr{S}-$moderate nets and
$\mathcal{N}_{\mathscr{S}}$ be the set of all
$\mathscr{S}-$negligible nets. Then define the quotient space

The set
$\mathcal{G}_{\mathscr{S}}$ is known as the Colombeau algebra associated to the locally convex space
$\mathscr{S}(\mathbb R^n)$ (see definition 3.1 of [Reference Garetto12]). Next, for fixed mollifiers ϕ and ψ such that
$\phi(0) = 1 = \widehat{\psi}(0)$,
$(\partial^{\beta}\phi)(0) = 0$ for all β ≠ 0 and
$\int x^{\alpha}\psi(x) dx = 0$ for all α ≠ 0, the map

is a well-defined embedding (due to Theorem A.7 and Proposition 2.1). Then, the regularization
$(\phi^{\varepsilon} (\psi_{\varepsilon} \ast u))_{\varepsilon}$ gives a way to include tempered distributions in the Colombeau algebra
$\mathcal{G}_{\mathscr{S}}$. Notice that in
$\mathcal{G}_{\mathscr{S}}$ we have a well-defined multiplication operation which is consistent with the product of functions in
$\mathscr{S}(\mathbb R^n)$.
We are finally ready to introduce the concept of a very weak solution that we are interested in and the main result of this paper, namely Theorem 2.7.
Definition 2.6. The net
$(u_\varepsilon)_{\varepsilon} \in \{C([0,T];\mathscr{S}(\mathbb R^{n}))\}^{(0,1]}$ is a
$\mathscr{S}$-very weak solution for the Cauchy problem (1.4) if
$(u_\varepsilon)_\varepsilon$ is
$\mathscr{S}$-moderate and there exist
•
$\mathscr{S}$-moderate regularizations
$(c_{j,\varepsilon})_{\varepsilon}$ of the coefficients cj,
$j = 0, 1, \ldots, n$,
•
$\mathscr{S}$-moderate regularizations
$(f_{\varepsilon})_{\varepsilon}, (g_{\varepsilon})_{\varepsilon}$ of the Cauchy data
$f$ and
$g$,
such that, for every fixed
$\varepsilon$,
$u_\varepsilon$ solves the regularized Cauchy problem (1.7) for the operator (1.8)
Theorem 2.7 Under the assumptions (1.5) and (1.6), the Cauchy problem (1.4) is
$\mathscr{S}$-very weakly well-posed, i.e. (1.4) admits a
$\mathscr{S}$ very weak solution and the solution is unique in the following sense:
$\mathscr{S}$-negligible changes on the regularizations of the equation coefficients and
$\mathscr{S}$-negligible changes on the regularizations of the initial data lead to
$\mathscr{S}$-negligible changes in the corresponding
$\mathscr{S}$-very weak solution.
Remark 2.8. Theorem 2.7 implies that (1.4) is well-posed in
$\mathcal{G}_{\mathscr{S}}$. More precisely, we can embed the coefficients cj of (1.2) (via regularizations parametrized by a suitable positive scale
$\omega(\varepsilon)$, see (4.1)) and the Cauchy data of (1.4) in the Colombeau algebra
$\mathcal{G}_{\mathscr{S}}$ in such a way that the corresponding embedded Cauchy problem is well-posed in
$\mathcal{G}_{\mathscr{S}}$.
To prove Theorem 2.7, we need to obtain a
$\mathscr{S}$-very weak solution for (1.4), that is, we need to solve the regularized problem (1.7) in
$\mathscr{S}(\mathbb R^{n})$. In order to do it, we shall employ the classical techniques developed in [Reference Kajitani and Baba17]. In the sequel, we shall fix the considered regularizations for the coefficients of (1.2) and for the data in (1.4). In what follows, we shall use the standard positive scale
$\varepsilon$ for the sake of simplicity, and we shall replace it with a suitable positive scale
$\omega(\varepsilon)$ just when it is needed.
2.1. Regularization of
$c_j(t,x)$,
$j=0,1,\ldots,n$
Let
$\psi \in \mathscr{S}(\mathbb R^n)$ with
$\int \psi = 1$ and
$\phi \in \mathscr{S}(\mathbb R^{n})$ with
$\phi(0) = 1$. We then define

Then we have
$c_{j,\varepsilon} \in C([0,T];\mathscr{S}(\mathbb R^{n}))$ and the following estimate holds

where N > 0 is a number depending on the coefficients cj,
$j = 0, 1, \ldots, n$, and on the dimension.
2.2. Regularization of the Cauchy data
Let
$\mu, \nu \in \mathscr{S}(\mathbb R^{n})$ with
$\int \mu = 1$ and
$\nu(0) = 1$. Let
$f_\epsilon(t,x) = \nu(\varepsilon x) (\mu_{\varepsilon} \ast_x f)(t,x)$ and
$g_\epsilon(x) = \nu(\varepsilon x) \mu_{\varepsilon}(x) \ast g(x)$. Then we have
$f_{\varepsilon} \in C([0,T]; \mathscr{S}(\mathbb R^{n}))$ and the following estimates

where
$\tilde{N}_f \gt 0$ is a number depending on f and on the dimension n,
$\tilde{N}_g \gt 0$ is a number depending on g and on the dimension n. By these estimates we immediately get that for all
$m,M \in \mathbb N_0$ there exist C > 0,
$N_f\in\mathbb N_0$ and
$N_g\in\mathbb N_0$ such that


We are finally ready to define the family of regularized Cauchy problems that we shall study in the subsequent sections. We consider the family of regularized operators (1.8) for
$\varepsilon \in (0, 1]$ and then the family of regularized Cauchy problems

where
$f \in C([0,T];\mathscr{S}(\mathbb R^{n}))$,
$g \in \mathscr{S}(\mathbb R^{n})$ and the coefficients of Sɛ defined in (1.8) are given in (2.2). In the next sections, we will obtain a net of solutions
$(u_{\varepsilon})_{\varepsilon \in (0,1]}$ where for every
$\varepsilon$ the function
$u_{\varepsilon} \in C([0,T]; \mathscr{S}(\mathbb R^{n}))$ is the unique solution of the Cauchy problem (2.7). Moreover, we will also derive energy estimates for the solutions uɛ expliciting how the constants depend on the parameter ɛ.
3. Solving the regularized problem (2.7)
We want to prove that there exists a unique solution in
$\mathscr{S}(\mathbb R^{n})$ for the problem (2.7). Since
$c_{j,\varepsilon} \in C([0,T];\mathscr{S}(\mathbb R^{n}))$ we know that (2.7) is well-posed in
$H^{\infty}(\mathbb R^{n})$ without loss of derivatives, i.e., there exists a unique solution
$u \in C([0,T];H^{\infty}(\mathbb R^{n}))$ satisfying (cf. Proposition 5.1 in [Reference Arias Junior, Ascanelli, Cappiello and Garetto3])

where N is a natural number depending on the coefficients cj and
$\theta_{n,m}$ is a natural number depending on the dimension
$n$ and on the Sobolev index m. We want to conclude that the solution u is also rapidly decreasing at infinity. This motivates the following conjugation: for a fixed
$s \in \mathbb N_0$, we define

where

Observe that in
$S_{\varepsilon, s}$ the new purely imaginary first-order term

appears. This term decays exactly like
$\langle x \rangle^{-1}$ for
$|x| \to \infty$. From the classical theory presented in [Reference Kajitani and Baba17] we know that the Cauchy problem for the operator
$S_{\varepsilon, s}$ is well-posed in
$H^{\infty}(\mathbb R^{n})$ and the decay
$\langle x \rangle^{-1}$ on the first-order coefficients produces a finite loss of Sobolev regularity with respect to the initial data. Since (3.2) does not depend on ɛ, the loss of regularity produced by this new term will not depend on ɛ, but will depend on s. This intuitive idea leads to the following result.
Proposition 3.1. Let
$\tilde{f} \in C([0,T]; \mathscr{S}(\mathbb R^{n}))$ and
$\tilde{g} \in \mathscr{S}(\mathbb R^{n})$. There exists a unique solution
$u$ in
$C([0,T];H^{\infty}(\mathbb R^{n}))$ to the Cauchy problem

and
$u$ satisfies: for each
$m \in \mathbb N_0$

where
$c(s) = Cs$ for some constant
$C \gt 0$ and
$J(N, s, m,n)$ is a natural number depending on
$N,s,m$ and
$n$.
Proposition 3.1 is crucial to prove the main result of the paper. Its proof is the most challenging part of the paper. However, since it is long and technical, we postpone it to Section 5 in order to address the reader as fast as possible to the proof of the main result.
The next theorem is a direct consequence of Proposition 3.1.
Theorem 3.2 For every
$\varepsilon \in (0,1]$ consider the regularized Cauchy problem (2.7) with initial data
$f \in C([0,T];\mathscr{S}(\mathbb R^{n}))$ and
$g \in \mathscr{S}(\mathbb R^{n})$. Then there exists a unique solution
$u_{\varepsilon} \in C([0,T]; \mathscr{S}(\mathbb R^{n}))$ for the problem (2.7). Moreover, the solution
$u_{\varepsilon}$ satisfies for every
$m,s\in\mathbb N_0$:

where
(i) The constants
$C_{m,s,n,T}$ and
$\tilde{C}_{m,s,n,T}$ depend on the coefficients
$c_j$,
$j=0,1,\ldots,n$, and on the mollifiers
$\psi, \phi$;
(ii)
$N$ is a positive integer depending on the coefficients
$c_0, c_1, \ldots, c_n$ and on the dimension;
(iii)
$J(N,s,m,n)$ is a natural number depending on
$N,s,m$ and
$n$.
4. From the regularized problem to the original problem
In this section, we shall apply Theorem 3.2 to prove Theorem 2.7.
4.1. Existence
The regularized Cauchy data
$f_\epsilon (t), g_\epsilon$ fulfil estimates (2.5), (2.6); moreover, from Theorem 3.2, the regularized problem (2.7) with data
$f_\varepsilon(t)$ and gɛ has a unique solution
$u_\varepsilon\in C([0,T];\mathscr S(\mathbb R^{n}))$ satisfying for every
$m, s\in\mathbb N_0$

where
$A_{m,s}, B_{m,s} \gt 0$ are independent from ɛ and
$N_{m,s}\in \mathbb N_0$ is independent from ɛ but depends on the coefficients cj,
$j = 0, 1, \ldots, n$, on the dimension n and on Sobolev indices
$m,s$. We thus obtain a
$\mathscr{S}$-moderate net of solutions for the problem (1.4) by regularizing the coefficients cj via mollifiers indexed by the positive scale

Indeed, by inequality
$\log(y) \leq C_{r} y^{\frac{1}{r}}$, for all
$y \geq 2$ and all
$r \geq 1$, we get

and so we have

where
$D_{N_{m,s}} \in \mathbb N_0$ is a large number depending on
$N_{m,s}$ and on
$B_{m,s}$. In this way, we obtain that for any
$m,s \in \mathbb N_0$ we find constants
$A_{m,s}, D_{N_{m,s}}$ such that

Therefore, the Cauchy problem (1.4) admits a
$\mathscr S'$-very weak solution.
4.2. Uniqueness
Suppose that we perturb the coefficients of the regularized operator Sɛ in the following way

where
$(c'_{j,\varepsilon})_{\varepsilon}$ are
$\mathscr{S}$-negligible. Suppose moreover that the net
$(u'_{\varepsilon})_\varepsilon$ solves the Cauchy problem

where
$(p_\varepsilon)_{\varepsilon}$ and
$(q_\varepsilon)_{\varepsilon}$ are
$\mathscr{S}$-negligible nets. We now want to compare the two very weak solutions uɛ and
$u'_\varepsilon$. Since

and

it follows that

By applying estimate (3.5), using the scale (4.1) to regularize the coefficients cj, we have that for all
$m, s \in \mathbb N_0$ there exist
$\tilde{C} \gt 0$ and
$\tilde{N}\in\mathbb N_0$ such that

Since
$(p_\varepsilon)$ and
$(q_\varepsilon)$ are both
$\mathscr{S}$-negligible, the coefficients of the operator
$(S_\varepsilon-S'_\varepsilon)_\varepsilon$ are
$\mathscr{S}$-negligible and
$(u'_{\varepsilon})_{\varepsilon}$ is
$\mathscr{S}$-moderate, we conclude that the right-hand side of (4.2) can be estimated by any positive power of ɛ. Hence,
$(u_{\varepsilon}-u'_{\varepsilon})_{\varepsilon}$ is
$\mathscr{S}$-negligible.
5. Proof of Proposition 3.1
To solve (3.3), we take inspiration from the approach used in [Reference Kajitani and Baba17], which is based on a suitable change of variable that turns (3.3) into an auxiliary Cauchy problem well-posed in
$L^2(\mathbb R^n)$. The change of variable we use is the composition of two pseudodifferential operators. In the next lines, we introduce this change of variable. A survival toolkit of what we need about the theory of pseudodifferential operators is reported in the Appendix B at the end of the paper.
In [Reference Kajitani and Baba17] the authors proved that for every
$M_1, M_2 \gt 0, $ there exist real-valued functions
$\tilde\lambda_1, \tilde\lambda_2$ satisfying for
$x \in \mathbb R^n, |\xi|\geq 1, \alpha, \beta \in \mathbb N_0^n$, β ≠ 0:

and

where
$\chi \in \mathcal{C}^{\infty}_{c}(\mathbb R)$ such that
$\chi(t) = 0$ for
$|t| \gt 1$,
$\chi(t) = 1$ for
$|t| \lt \frac{1}{2}$,
$t \chi'(t) \leq 0$ and
$0 \leq \chi \leq 1$.
For a large parameter
$h \geq 1$ to be chosen later on, we consider

Then, since
$\langle \xi \rangle_{h} = \sqrt{h^2 + |\xi|^2 } \leq \sqrt{5}h$ on the support of
$(1-\chi)(h^{-1}|\xi|)$, we have

where the constants Cα and
$C_{\alpha,\beta}$ do not depend on
$M_{\ell}$ and on h. To define the change of variable, we shall use the symbols
$e^{\lambda_\ell(x,\xi)}, \ell =1,2.$ The following estimates are followed by a standard application of Faà di Bruno formula:

On the other hand,

Hence, it follows that there exists a number
$c \gt 0$ such that

and

Concerning
$e^{\lambda_2(x,\xi)}$ we have that it belongs to
$S^0(\mathbb R^{2n}).$
We shall use the operators
$e^{\lambda_\ell}(x,D)$ to handle the first-order terms of the operator
$S_{\varepsilon,s}$ defined in (3.1). These terms are given by

and

Since
$c_{j,\varepsilon}(t,x)$ decays faster than
$\langle x \rangle^{-1}$, we have that (5.2) does not produce any loss of Sobolev regularity. On the other hand, the term (5.3) decays exactly as
$\langle x \rangle^{-1}$, so it produces a loss of regularity, but this loss will be independent on the parameter ɛ, since (5.3) does not depend on ɛ. Our idea is then to use
$e^{\lambda_{2}}(x,D)$ to handle the term (5.2) and
$e^{\lambda_{1}}(x,D)$ to control the new first-order term (5.3) produced by the conjugation by
$\langle x \rangle^{s}$. Namely, we shall consider a change of variable of the form

Remark 5.1. Notice that in the conjugation, the leading term is
$e^{\lambda_1}(x,D)$. However, we need to perform two conjugations in order to avoid a loss of Sobolev regularity depending on ɛ, which would appear if we consider the sole transformation
$e^{\lambda_1}(x,D)$.
In the remaining part of this section, we shall derive a priori energy estimates to the auxiliary problem (5.4), see Proposition 5.4. Then the proof of Proposition 3.1 will be a consequence of the derived energy estimates and the mapping properties of
$e^{\lambda_1}(x,D)$ and
$e^{\lambda_2}(x,D)$.
5.1. Invertibility of
$e^{\lambda_1}(x,D)$ and conjugation theorem
The reverse operator
$^{R}\{e^{\lambda_1}(x,D)\}$ was introduced in [Reference Kajitani and Wakabayashi18] (see Proposition 2.13) as the transpose operator of
$e^{\lambda_1}(x,-D)$. Since λ 1 is real-valued, in this particular case, the reverse operator coincides with the L 2 adjoint of
$e^{\lambda_1}(x,D)$. We have that

where

So, Taylor’s formula and usual computations give

where

The advantage of using the reverse operator is that in the above asymptotic formula, the terms

appear, and they are of order exactly
$-|\alpha|$ (i.e. the x-derivatives in the above asymptotic formula allow us to avoid the log growth appearing in the symbol estimates of λ 1). The symbols
$e^{\pm \lambda_1}$ are of finite order, so we can estimate rN in the following way

Since (5.5) holds for any
$N \in \mathbb N$, it follows that

where r −2 has order −2. Hence, if we assume that
$h \geq h_0(M_1)$ we obtain that

where
$\sum_{j \geq 0} (-r)^j$ has symbol of order zero and
$r_{-2}$ has symbol of order −2 satisfying

for some
$C \gt 0$ independent on M 1 and h (cf. theorem
$I.1$ at page 372 of [Reference Kumano-Go20]).
For a
$p \in S^{m}(\mathbb R^{2n})$ such that

we consider

Thus

Now we need to understand the composition
$
e^{\lambda_1}(x,D) \circ p(x,D) \circ\, \{e^{\lambda_1}(x,D)\}^{-1} .
$ This is done in two steps in view of (5.6). The next theorem is a direct consequence of Theorem B.3.
Theorem 5.2 For any
$N \in \mathbb N$ the symbol of
$e^{\lambda_1}(x,D)\circ p(x,D) \circ\, ^{R}\{e^{-\lambda_1}(x,D)\}$ is given by

where

$J$ being a natural number depending on
$M_1, m$ and on the dimension
$n$.
Theorem 5.2 implies the following: if we know that pɛ satisfies

then taking
$N(M_1) = \lfloor 2cM_1 + 1 \rfloor + 1$ we get that rN is a symbol of order zero and satisfies

where
$J$ is a natural number depending on
$M_1$ and
$n$. Thus, (5.6) becomes

where r 0 satisfies (5.7). Now note that we can write qɛ in the following way:

for a new zero-order term r 0 satisfying an estimate like (5.7). Hence

Since
$
\sum\limits_{j=1}^{n} \partial_{\xi_j} \lambda_1 D_{x_j} p_{\varepsilon}
$ behaves like
$\varepsilon^{-N_1-1} \log\langle \xi \rangle_{h}$, it cannot be regarded as a zero-order term.
5.2. Invertibility of
$e^{\lambda_2}(x,D)$
Here, we just report results derived in [Reference Arias Junior, Ascanelli, Cappiello and Garetto3] at Section 6.1.
Lemma 5.3. For all
$h \geq h_0(M_2,n) := A e^{(2C+1)M_2}$, for some
$A, C \gt 0$ independent on M 2, the operator
$e^{\lambda_2}(x,D)$ is invertible and we have

where

and

Moreover,

where
$s(x,\xi)$ is a zero-order symbol and its seminorms do not depend on
$h$ and on
$M_2$.
We need to write
$s(x,D)$ in the following convenient way:

So,

where
$s_{-2}(x,\xi)$ has order −2 and satisfies

where the constants
$C_{\alpha,\beta,n}$ and
$C$ do not depend on
$M_2$ and on
$h$.
5.3. Conjugation of
$S_{\varepsilon,s}$ by
$e^{\lambda_1}(x,D)$
We recall that

where

Let us set

Applying Theorem 5.2 to
$\sum\limits_{j=1}^{n} D^{2}_{x_j}$ and formula (5.8) to the first-order terms of
$S_{\varepsilon, s}$ we get

where

and
$d_{0,\varepsilon, s}$ is a zero-order term satisfying

In order to simplify the expression for
$S_{\varepsilon,s,\lambda_1}$ we set

We note that

and we may write

5.4. Conjugation of
$S_{\varepsilon, s, \lambda_1}$ by
$e^{\lambda_2}(x,D)$
We shall denote

We recall that
$e^{\lambda_2}$ has order zero. So, using Lemma 5.3, Theorem B.3, and repeating similar ideas of the previous subsection, we end up with

for a new zero-order term
$d_{0,\varepsilon,s}(t,x,\xi)$ satisfying

where
$J(N, M_1, n)$ is a natural number depending on
$N, M_1$ and on the dimension and the constants
$C_{\alpha,\beta,M_1,n,s}$,
$C$ do not depend on
$\varepsilon, M_2$ and
$h$.
5.5. Choices of M 1, M 2, h and L 2 energy estimate
Inserting (5.12) in
$p_{\varepsilon,\log}(t,x,\xi)$ we obtain

Since
$c_{j,\varepsilon}(t,x)$ are Schwartz functions in x we obtain

Analogously

So, we can consider
$q_{1,\varepsilon}$ and
$q_{2,\varepsilon}$ as symbols of order 1 with decay
$\langle x \rangle^{-2}$ and seminorms bounded by
$M_1C_{\alpha, \beta, s, n}\varepsilon^{-N - |\beta| - 3}h^{-\frac{1}{2}}$.
In order to derive an L 2 energy estimate for
$S_{\varepsilon,\lambda}$ we write

We also recall that

Note that if a symbol
$c(x,\xi)$ of order 1 decays like
$\langle x \rangle^{-2}$, then we can split it as follows

So, outside of the support of
$\chi\left( \frac{\langle x \rangle}{|\xi|} \right) (1-\chi)\left( \frac{|\xi| }{h } \right)$ we have that

where

On the other hand, on the support of
$\chi\left( \frac{\langle x \rangle}{|\xi|} \right) (1-\chi)\left( \frac{|\xi| }{h } \right)$ we can take
$M_1(s)$ and
$M_2(\varepsilon)$ large in order to get that the symbols of the first-order part are non-negative. Indeed, we have

and

Hence, choosing

and taking h large so that

we obtain that the real part of the first-order symbols in (5.18) are all non-negative.
Hence, we may apply the sharp Gårding inequality to get the following energy estimate:

To obtain (5.19) with general
$m-$norms we apply the same ideas to the operator

which is almost equal to
$S_{\varepsilon,s,\lambda_1,\lambda_2}$ except for a zero-order term. So,

Gronwall’s inequality then gives (recalling that
$M_1$ depends only on
$s$ and
$M_2 = C\varepsilon^{-N-2}$)

The next proposition is a consequence of (5.20).
Proposition 5.4. Let
$f \in C([0,T]; H^m(\mathbb R^{n}))$ and
$g \in H^{m}(\mathbb R^{n})$. There exists a unique solution
$u$ in
$C([0,T];H^{m}(\mathbb R^{n}))$ to the Cauchy problem

and
$u$ satisfies (5.20).
Proof of Proposition 3.1
We observe that
$e^{\lambda_2}(x,D): H^{m} \to H^{m}$ and
$e^{\lambda_1}(x,D): H^m \to H^{m-\frac{c(s)}{2}}$, for some c(s) depending on s. Hence
$e^{\lambda_2}(x,D)$ does not produce any loss of regularity and
$e^{\lambda_1}(x,D)$ produces a loss depending on s. Then, Proposition 3.1 is a consequence of Proposition 5.4 and the mapping properties of
$e^{\lambda_1}(x,D)$ and
$e^{\lambda_2}(x,D)$.
6. Consistency with the classical theory in the case of regular coefficients
Suppose that the coefficients of the operator
$S$ are Schwartz functions, more precisely:

In this situation the Cauchy problem (1.4) with data
$f \in C([0,T];\mathscr{S}(\mathbb R^{n}))$ and
$g \in \mathscr{S}(\mathbb R^{n})$ admits a unique solution
$u \in C([0,T]; \mathscr{S}(\mathbb R^{n}))$ satisfying an energy estimate like (3.5) (with a constant which does not depend on
$\varepsilon$). The goal of this section is to verify that the net obtained in Theorem 3.2 converges to u in the
$\mathscr{S}(\mathbb R^{n})$ topology.
Let uɛ be the solution of the Cauchy problem (2.7). Then
$u - u_{\varepsilon}$ solves the following Cauchy problem

where

So, we have the following estimate: for any
$m, s \in \mathbb N_0$

Now we observe the following:
(i) Since, in this case, we are regularizing regular functions, it is easy to conclude that the regularizations
$c_{j,\varepsilon}$ satisfy uniform estimates with respect to
$\varepsilon$. So, in this particular case, we obtain estimate (3.5) uniformly with respect to
$\varepsilon$. Hence, for any
$m,s \in \mathbb N_0$,
\begin{equation*} \|u_{\varepsilon}(\tau)\|^{2}_{H^{m,s}} \leq C_{m,s}(f,g), \quad \forall \varepsilon \in (0,\varepsilon_0). \end{equation*}
(ii) We have the following convergence: for any
$\beta \in \mathbb N^n_0$ and
$j = 0, 1, \ldots, n$ it holds
\begin{equation*} \sup_{t \in [0,T],\, x \in \mathbb R^{n}} |\partial^{\beta}_{x}c_{j,\varepsilon}(t,x) - \partial^{\beta}_{x}c_{j}(t,x)| \to 0, \quad \text{as} \, \varepsilon \to 0. \end{equation*}
(iii) The regularized data satisfy
\begin{equation*} \|g-g_\varepsilon\|_{H^{m,s}} \to 0, \quad \sup_{t \in [0,T]} \|f(t)-f_{\varepsilon}(t)\|_{H^{m,s}} \to 0, \quad \text{when}\,\, \varepsilon \to 0, \end{equation*}
for all
$m, s \in \mathbb N_0$.
We therefore conclude
$u_{\varepsilon} \to u$ in
$C([0,T];\mathscr{S}(\mathbb R^{n}))$. Summing up, we have proven the following result.
Proposition 6.1. Assume that
$c_j \in C([0,T]; \mathscr{S}(\mathbb R^{n}))$ for
$j = 0, 1, \ldots, n$. Assume moreover
$g \in \mathscr{S}(\mathbb R^n)$ and
$f \in C([0,T];\mathscr{S}(\mathbb R^n))$. Then, any
$\mathscr{S}$-very weak solution converges to the unique classical solution in the space
$C([0,T];\mathscr{S}(\mathbb R^{n}))$. In particular, in the classical case, the limit of very weak solutions always exists and does not depend on the regularization.
7. Classical Schwartz well-posedness result
As we mentioned in the introduction, from the theory developed in this paper, we obtain a result concerning the well-posedness in Schwartz spaces of the Cauchy problem (1.4), which is new and interesting per se. In this section, we shall state it and explain how to prove it from the ideas used throughout the manuscript.
Let
$S$ be the Schrödinger type operator given in (1.2) with the following hypotheses on the coefficients:

Then the following result holds.
Theorem 7.1 Suppose the coefficients of
$S$ satisfy (7.1). Let
$f \in C([0,T]; \mathscr{S}(\mathbb R^n))$ and
$g \in \mathscr{S}(\mathbb R^n)$ and consider the Cauchy problem (1.4). Then there exists a unique solution
$u \in C([0,T]; \mathscr{S}(\mathbb R^n))$. Moreover, the solution
$u$ satisfies the following energy inequality

for some constant
$c(s)$ depending only on
$s$. In particular, the Cauchy problem (1.4) is well-posed in
$\mathscr{S}(\mathbb R^n)$.
To prove the above theorem, the idea is to consider the operator

and to repeat the steps of Section 5. We just remark that since we do not have to worry with a parameter ɛ and the coefficients have rapid decay, the sole conjugation by
$e^{\lambda_1}(x,D)$ is enough to prove the above result via an auxiliary Cauchy problem (cf. again Section 5).
Acknowledgements
A. Arias Junior was supported by the grants
$2022/01712-3$ from São Paulo Research Foundation (FAPESP) and
$306699/2025-7$ from Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq). A. Ascanelli and M. Cappiello were supported by the INdAM-GNAMPA projects CUP E53C22001930001 and CUP E53C23001670001. C. Garetto was supported by the EPSRC Grant EP/V005529/2.
Appendix A. Regularization of tempered distributions by Schwartz functions
This Appendix is devoted to the study of a special kind of regularizations of tempered distributions. Namely, we want to regularize distributions in
$\mathscr{S}'(\mathbb R^n)$ by functions in
$\mathscr{S}(\mathbb R^n)$. Despite the results reported here are quite natural and straightforward, we decided to present them in detail for the sake of completeness.
We recall that if
$u \in \mathscr{S}'(\mathbb R^n)$ then there exist C > 0 and
$N \in \mathbb N_0$ such that

Let
$\phi, \psi \in \mathscr{S}(\mathbb R^{n})$ such that
$\phi(0) = 1$ and
$\widehat{\psi}(0) = 1$. We shall use the following notations

For any given
$u \in \mathscr{S}'(\mathbb R^n)$, we consider the following type of regularization

Lemma A.1. Let
$u \in \mathscr{S}'(\mathbb R^{n})$. Then

where

More precisely, the following estimate holds:

where
$N$ depends on
$u$ and
$C_{N,\alpha}$ depends on
$u$,
$\alpha$ and on seminorms of
$\psi$ with order up to
$N+|\alpha|$.
Proof. The proof follows from the following estimate

From Leibniz formula and the above lemma, we easily conclude that
$u_\varepsilon$ given in (A.1) belongs to
$\mathscr{S}(\mathbb R^{n})$ for all
$\varepsilon \in (0,1)$ and the following estimate holds:

Now we want to check that the regularization uɛ converges to u in
$\mathscr{S}'(\mathbb R^{n})$. This is done in Theorem A.7 below; to prove it, we shall need the following lemmas.
Lemma A.2. Let
$\{u_j\}_{j \in \mathbb N_0} \subset \mathscr{S}'(\mathbb R^{n})$,
$u \in \mathscr{S}'(\mathbb R^{n})$,
$\{h_j\}_{j \in \mathbb N_0} \subset \mathscr{S}(\mathbb R^{n})$ and
$h \in \mathscr{S}(\mathbb R^{n})$. If

then
$u_j(h_j) \to u(h)$ in
$\mathbb C$.
Proof. Since
$u_j \to u$ in
$\mathscr{S}'(\mathbb R^{n})$, we have that the family
$\{u_j\}$ is pointwise bounded. Therefore, Banach–Steinhaus theorem gives that the sequence
$\{u_j\}$ is equicontinuous. In this way, for any
$\delta_1 \gt 0$ there exists
$\delta_2 \gt 0$ such that

where d denotes the usual distance in
$\mathscr{S}(\mathbb R^{n})$. In particular, we conclude that

To finish the proof, it suffices to observe that

Lemma A.3. For any
$h \in \mathscr{S}(\mathbb R^{n})$ we have

Proof. Since
$\partial^{\alpha}_{x} \{\psi_{\varepsilon} \ast h \} = \psi_{\varepsilon} \ast \{\partial^{\alpha}_{x} h\}$ we only need to prove that

We have

Let
$\delta \gt 0$. When
$|y| \leq \delta$ we have

Since

if
$|y| \geq \delta$ we get

From the above inequalities, we can find
$\varepsilon_0 \gt 0$ such that

Remark A.4. From the previous lemma we conclude that for any
$u \in \mathscr{S}'(\mathbb R^{n})$

Indeed, it suffices to observe

where
$\tilde{\psi}(\xi) = \psi(-\xi)$, for any
$\psi \in \mathscr{S}(\mathbb R^n)$ and
$u \in \mathscr{S}'(\mathbb R^n)$.
Lemma A.5. For any
$h \in \mathscr{S}(\mathbb R^{n})$ we have

Proof. We have

So we only need to prove that
$ \langle x \rangle^{M}\{\phi(\varepsilon x) - 1\}\partial^{\alpha}_{x} h(x)$ converges to zero uniformly in
$x$. We first write

Let then δ > 0. There is
$R_{\delta} \gt 0$ such that

On the other hand, since
$\phi_{\varepsilon}(x)$ converges uniformly to 1 on compact sets, we may find
$\varepsilon_0 \gt 0$ such that

Hence, if
$\varepsilon \leq \varepsilon_0$ we have

Remark A.6. From the previous two lemmas we conclude
$u_\varepsilon \to u$ in
$\mathscr{S}(\mathbb R^{n})$ provided that
$u \in \mathscr{S}(\mathbb R^{n})$.
We are finally ready to prove that the regularization
$u_\varepsilon$ defined by (A.1) converges to u in
$\mathscr{S}'(\mathbb R^{n})$. Indeed, this follows immediately from the previous three lemmas. As a matter of fact

We summarize what we have proven so far in this Appendix in the following result.
Theorem A.7 Let
$\phi, \psi \in \mathscr{S}(\mathbb R^{n})$ such that
$\phi(0) = 1$ and
$\widehat{\psi}(0) = 1$ and define

For any given
$u \in \mathscr{S}'(\mathbb R^n)$ consider

Then
$u_{\varepsilon} \in \mathscr{S}(\mathbb R^{n})$ for any ɛ,
$u_{\varepsilon} \to u$ in
$\mathscr{S}'(\mathbb R^{n})$ and

Remark A.8. Of course, the above theorem still holds if we replace
$\varepsilon$ by another positive scale
$\omega(\varepsilon)$.
It is also in our interest to regularize curves of temperate distributions.
Proposition A.9. Let
$u \in C([0,T]; \mathscr{S}'(\mathbb R^{n}))$ and set

Then
$u_{\varepsilon} \in C([0,T];\mathscr{S}(\mathbb R^{n}))$ and the following estimate holds

where N > 0 and C > 0 do not depend on
$t$.
Proof. First of all we notice that for every fixed
$\psi, \phi \in \mathscr{S}(\mathbb R^n)$ the map
$v \mapsto \phi( \psi \ast v)$ is continuous from
$\mathscr{S}'(\mathbb R^n)$ to
$\mathscr{S}(\mathbb R^n)$. Therefore the map
$t \mapsto \psi(\varepsilon \cdot) \{\psi \ast u(t)\}(\cdot)$ is continuous from
$[0,T]$ to
$\mathscr{S}(\mathbb R^n)$ for every fixed
$\varepsilon$, since it is a composition of continuous maps. Hence
$u_{\varepsilon} \in C([0,T];\mathscr{S}(\mathbb R^{n}))$.
In order to obtain (A.3), we observe that Banach–Steinhaus theorem implies that
$\{u(t)\}_{t \in [0,T]}$ is equicontinuous. Hence, there exists an open neighbourhood
$V \subset \mathscr{S}(\mathbb R^{n})$ of 0 such that

where
$B(0,1) = \{z \in \mathbb C: |z| \lt 1\}$. Since
$V$ is open, from the topology of
$\mathscr{S}(\mathbb R^{n})$, we find
$N \gt 0$ and
$r \gt 0$ such that

where

Hence for any
$t \in [0,T]$ and any
$h \in \mathscr{S}(\mathbb R^{n})$ we get

This means that we can find
$C, N \gt 0$ independent of
$t \in [0,T]$ such that

Appendix B. Pseudodifferential operators
This Appendix is devoted to reporting some results and definitions concerning pseudodifferential operators that were employed throughout the paper. For the proofs, we refer the reader to [Reference Kumano-Go20].
Definition B.1. For a given
$m\in\mathbb R$,
$S^{m}(\mathbb R^{2n})$ denotes the space of all smooth functions
$p(x,\xi) \in C^{\infty}(\mathbb R^{2n})$ such that for any
$\alpha, \beta \in \mathbb N^{n}_{0}$ there exists a positive constant
$C_{\alpha,\beta}$ for which

The Fréchet topology of the space
$S^{m}(\mathbb R^{2n})$ is induced by the following family of seminorms

Given a symbol
$p(x,\xi)$ we associate to it the operator

which maps continuously
$\mathscr{S}(\mathbb R^{n})$ into itself.
We will sometimes also use the notation
$p(x,D) = {\rm op}(p(x,\xi))$. The next result concerns the action of such operator on Sobolev spaces.
Theorem B.2 (Calderón-Vaillancourt)
Let
$p \in S^{m}(\mathbb R^{2n})$. Then for any real number
$s \in \mathbb R$ there exist
$\ell := \ell(s,m) \in \mathbb N_0$ and
$C:= C_{s,m} \gt 0$ such that

Moreover, when
$m = s = 0$ we can replace
$|p|^{(m)}_{\ell}$ by

where

In the following, we consider the algebra properties of
$S^{m}(\mathbb R^{2n})$ with respect to the composition of operators. In the sequel,
$Os-$ in front of the integral sign stands for oscillatory integral. Let
$p_j \in S^{m_j}(\mathbb R^{n})$,
$j = 1, 2$, and define

Theorem B.3 Let
$p_j \in S^{m_j}(\mathbb R^{2n})$,
$j = 1, 2$, and consider
$q$ defined by (B.1). Then
$q \in S^{m_1+m_2}(\mathbb R^{2n})$ and
$q(x,D) = p_1(x,D) p_2(x,D)$. Moreover, the symbol
$q$ has the following asymptotic expansion

where

and the seminorms of
$r_N$ may be estimated in the following way: for any
$\ell_{0} \in \mathbb N_0$ there exists
$C_{\ell, N, n} \gt 0$ such that

The last theorem that we recall is the celebrated sharp Gårding inequality (see, for instance, Theorem 2.1.3 in [Reference Kenig, Ponce, Rolvung and Vega19]).
Theorem B.4 Let
$p \in S^{1}(\mathbb R^{2n})$ and suppose
$Re\,p(x,\xi) \geq 0$ for all
$x \in \mathbb R^n$ and
$|\xi| \geq R$ for some
$R$ > 0. Then there exist
$k = k(n) \in \mathbb N_0$ and
$C = C(n,R)$ such that
