1 Introduction

In this paper we consider the compressible Euler–Korteweg (EK) system

$$\begin{aligned} {\left\{ \begin{array}{ll} \partial _t \rho + {\text {div}}(\rho \vec {u})=0\\ \partial _t \vec {u} + \vec {u} \cdot \nabla \vec {u} + \nabla g(\rho )= \nabla \big ( K(\rho )\Delta \rho + \frac{1}{2} K'(\rho )|\nabla \rho |^2 \big ), \end{array}\right. } \end{aligned}$$
(1.1)

which is a modification of the Euler equations for compressible fluids to include capillary effects, under space periodic boundary conditions \( x \in {\mathbb {T}}^d {:}{=} ({\mathbb {R}}/2 \pi {\mathbb {Z}})^d \). The scalar variable \( \rho (t,x) > 0 \) is the density of the fluid and \( \vec {u} (t,x) \in {\mathbb {R}}^d \) is the time dependent velocity field.

The functions \( K (\rho ) \), \( g (\rho ) \) are defined on \( {\mathbb {R}}^+ \), smooth, and \( K (\rho ) \) is positive.

The quasi-linear equations (1.1) appear in a variety of physical contexts modeling phase transitions [17], water waves [14], quantum hydrodynamics where \( K(\rho ) = \kappa / \rho \) [4], see also [15].

Local well posedness results for the (EK)-system have been obtained in Benzoni-Gavage, Danchin and Descombes [8] for initial data sufficiently localized in the space variable \( x \in {\mathbb {R}}^d \). Then, for \( d \ge 3 \), thanks to dispersive estimates, global in time existence results have been obtained for small irrotational data by Audiard–Haspot [7], assuming the sign condition \(g'( \rho ) > 0\). The case of quantum hydrodynamics corresponds to \(K(\rho ) = \kappa /\rho \) and, in this case, the (EK)-system is formally equivalent, via Madelung transform, to a semilinear Schrödinder equation on \( {\mathbb {R}}^d \). Exploiting this fact, global in time weak solutions have been obtained by Antonelli–Marcati [4, 5] also allowing \(\rho (t,x) \) to become zero (see also the recent paper [6]).

In this paper we prove a local in time existence result for the solutions of (1.1), with space periodic boundary conditions, under natural minimal regularity assumptions on the initial datum in Sobolev spaces, see Theorem 1.1. Relying on this result, in a forthcoming paper [10], we shall prove a set of long time existence results for the (EK)-system in 1-space dimension, in the same spirit of [11, 12].

We consider irrotational velocity fields \( \vec {u} \), i.e. with vorticity \( \Omega := {\text {curl}}(\vec {u} ):= \nabla \vec {u} - (\nabla \vec {u} )^\top \) equal to zero. Note that, since \( \partial _t \Omega = -{\text {curl}}( \Omega \vec {u} ) \), if \( \Omega \) is initially zero, then \( \Omega \) remains zero under the evolution of (1.1). An irrotational vector field on \( {\mathbb {T}}^d \) reads (Helmotz decomposition)

$$\begin{aligned} \vec {u} = \vec {c} (t) + \nabla \phi , \quad \vec {c} (t) \in {\mathbb {R}}^d, \quad \vec {c} (t) = {\frac{1}{(2\pi )^d}} \int _{{\mathbb {T}}^d} \vec {u} \, {\text {d}}x, \end{aligned}$$
(1.2)

where \( \phi : {\mathbb {T}}^d \rightarrow {\mathbb {R}}\) is a scalar potential. By the second equation in (1.1) and \({\text {curl}}\, \vec {u} = 0 \), we get

$$\begin{aligned} \partial _t \vec {c} (t) = - {\frac{1}{(2\pi )^d}} \int _{{\mathbb {T}}^d} \vec {u} \cdot \nabla \vec {u} \, {\text {d}}x = {\frac{1}{(2\pi )^d}} \int _{{\mathbb {T}}^d} -\frac{1}{2} \nabla (|\vec {u}|^2) \, {\text {d}}x = 0 \quad \Longrightarrow \quad \vec {c}(t) = \vec {c}(0) \end{aligned}$$

is independent of time. Note that if the dimension \(d = 1\), the average \( {\frac{1}{2\pi }} \int _{{\mathbb {T}}} u(t,x) {\text {d}}x\) is an integral of motion for (1.1), and thus any solution u(tx), \( x \in {\mathbb {T}}\), of the (EK)-system (1.1) has the form (1.2) with \(c(t) = c(0) \) independent of time, that is \( u (t,x) = c(0) + \phi _x (t,x) \).

The (EK) system (1.1) is Galilean invariant: if \( (\rho (t,x), \vec {u} (t,x)) \) solves (1.1) then

$$\begin{aligned} \rho _{\vec {c}} (t,x) := \rho _{\vec {c}} (t,x + \vec {c} t), \quad \vec {u}_{\vec {c}} (t,x) := \vec {u} (t,x + \vec {c} t) - \vec {c} \end{aligned}$$

solve (1.1) as well. Thus, regarding the Euler–Korteweg system in a frame moving with a constant speed \( \vec {c} (0) \), we may always consider in (1.2) that

$$\begin{aligned} \vec {u} = \nabla \phi , \quad \phi : {\mathbb {T}}^d \rightarrow {\mathbb {R}}. \end{aligned}$$

The Euler–Korteweg equations (1.1) read, for irrotational fluids,

$$\begin{aligned} {\left\{ \begin{array}{ll} \partial _t \rho + {\text {div}}(\rho \nabla \phi )=0\\ \partial _t \phi +\frac{1}{2} |\nabla \phi |^2 + g(\rho )= K(\rho )\Delta \rho + \frac{1}{2} K'(\rho )|\nabla \rho |^2. \end{array}\right. } \end{aligned}$$
(1.3)

The main result of the paper proves local well posedness for the solutions of (1.3) with initial data \((\rho _0, \phi _0) \) in Sobolev spaces

$$\begin{aligned} H^s ({\mathbb {T}}^d) := \Big \{ u (x) = \sum _{j\in {\mathbb {Z}}^d} u_j e^{{\text {i}}j \cdot x} \ : \ \Vert u \Vert _s^2 := \sum _{j \in {\mathbb {Z}}^d} |u_j|^2 \langle j \rangle ^{2s} < + \infty \Big \} \end{aligned}$$

where \( \langle j \rangle := \max \{1,|j|\} \), under the natural mild regularity assumption \( s > 2 + (d/2)\). Along the paper, \( H^s ({\mathbb {T}}^d) \) may denote either the Sobolev space of real valued functions \( H^s ({\mathbb {T}}^d, {\mathbb {R}}) \) or the complex valued ones \( H^s ({\mathbb {T}}^d, {\mathbb {C}}) \).

Theorem 1.1

(Local existence on \({\mathbb {T}}^d\)) Let \(s > 2 + \frac{d}{2} \) and fix \(\frac{d}{2}<s_0\le s-2\). For any initial data

$$\begin{aligned} (\rho _0, \phi _0) \in H^s ({\mathbb {T}}^d,{\mathbb {R}}) \times H^s({\mathbb {T}}^d,{\mathbb {R}}) \quad {\text {with}} \quad \rho _0(x) > 0, \quad \forall x \in {\mathbb {T}}^d \,, \end{aligned}$$

there exists \(T:= T(\Vert (\rho _0, \phi _0)\Vert _{s_0+2}, \min _x \rho _0(x)) >0 \) and a unique solution \( (\rho , \phi )\) of (1.3) such that

$$\begin{aligned}&(\rho , \phi ) \in C^{0}\Big ([-T,T], H^{s}({\mathbb {T}}^d,{\mathbb {R}})\times H^{s}({\mathbb {T}}^d,{\mathbb {R}})\Big ) \\&\quad \cap C^{1}\Big ([-T,T], H^{s-2}({\mathbb {T}}^d,{\mathbb {R}})\times {H}^{s-2}({\mathbb {T}}^d,{\mathbb {R}}) \Big )\ \end{aligned}$$

and \(\rho (t,x) > 0 \) for any \(t \in [-T, T]\). Moreover, for \(|t| \le T\), the solution map \( (\rho _0, \phi _0) \mapsto (\rho (t, \cdot ), \phi (t, \cdot ) ) \) is locally defined and continuous in \( H^{s}({\mathbb {T}}^d,{\mathbb {R}})\times H^{s}({\mathbb {T}}^d,{\mathbb {R}}) \).

We remark that it is sufficient to prove the existence of a solution of (1.3) on [0, T] because system (1.3) is reversible: the Euler–Korteweg vector field X defined by (1.3) satisfies \( X\circ {\mathcal {S}}= - {\mathcal {S}}\circ X \), where \( {\mathcal {S}}\) is the involution

$$\begin{aligned} {\mathcal {S}}\left( \begin{matrix} \rho \\ \phi \end{matrix} \right) := \left( \begin{matrix} \rho ^\vee \\ - \phi ^\vee \end{matrix} \right) , \quad \rho ^\vee (x) := \rho (- x). \end{aligned}$$
(1.4)

Thus, denoting by \( (\rho , \phi )(t,x) = \Omega ^t (\rho _0, \phi _0) \) the solution of (1.3) with initial datum \((\rho _0, \phi _0)\) in the time interval [0, T], we have that \( {\mathcal {S}}\Omega ^{-t} ({\mathcal {S}}(\rho _0, \phi _0) ) \) solves (1.3) with the same initial datum but in the time interval \([-T,0]\).

Let us make some comments about the phase space of system (1.3). Note that the average \( {\frac{1}{(2\pi )^d}} \int _{{\mathbb {T}}^d} \rho (x) \, {\text {d}}x \) is a prime integral of (1.3) (conservation of the mass), namely

$$\begin{aligned} {\frac{1}{(2\pi )^d}} \int _{{\mathbb {T}}^d} \rho (x) \, {\text {d}}x = {{\mathtt {m}}}, \quad {{\mathtt {m}}}\in {\mathbb {R}}, \end{aligned}$$
(1.5)

remains constant along the solutions of (1.3). Note also that the vector field of (1.3) depends only on \( \phi - \frac{1}{(2 \pi )^d}\int _{{\mathbb {T}}^d} \phi \, {\text {d}}x \). As a consequence, the variables \( (\rho -{{\mathtt {m}}}, \phi ) \) belong naturally to some Sobolev space \( H^s_0({\mathbb {T}}^d) \times \dot{H}^s ({\mathbb {T}}^d) \), where \(H^s_0({\mathbb {T}}^d) \) denotes the Sobolev space of functions with zero average

$$\begin{aligned} H^s_0({\mathbb {T}}^d) := \Big \{ u \in H^s({\mathbb {T}}^d) \ :\ \int _{{\mathbb {T}}^d} u(x) {\text {d}}x = 0 \Big \} \end{aligned}$$

and \(\dot{H}^s({\mathbb {T}}^d)\), \(s \in {\mathbb {R}}\), the corresponding homogeneous Sobolev space, namely the quotient space obtained by identifying all the \(H^s({\mathbb {T}}^d)\) functions which differ only by a constant. For simplicity of notation we denote the equivalent class \( [u] := \{ u + c, c \in {\mathbb {R}}\} \), just by u. The homogeneous norm of \( u \in \dot{H}^s ({\mathbb {T}}^d) \) is \( \Vert u \Vert _s^2 := \sum _{j \in {\mathbb {Z}}^d {\setminus } \{0\}} |u_j|^2 |j |^{2s}\). We shall denote by \( \Vert \ \Vert _s \) either the Sobolev norm in \( H^s \) or that one in the homogenous space \( \dot{H}^s \), according to the context.

Let us make some comments about the proof. First, in view of (1.5), we rewrite system (1.3) in terms of \( \rho \leadsto {{\mathtt {m}}}+ \rho \) with \( \rho \in H^s_0 ({\mathbb {T}}^d) \), obtaining

$$\begin{aligned} {\left\{ \begin{array}{ll} \partial _t \rho = - {{\mathtt {m}}}\Delta \phi - {\text {div}}(\rho \nabla \phi )\\ \partial _t \phi = -\frac{1}{2} | \nabla \phi |^2 - g({{\mathtt {m}}}+\rho ) + K({{\mathtt {m}}}+\rho ) \Delta \rho + \frac{1}{2} K'({{\mathtt {m}}}+\rho )| \nabla \rho |^2. \end{array}\right. } \end{aligned}$$
(1.6)

Then Theorem 1.1 follows by the following result, that we are going to prove

Theorem 1.2

Let \(s > 2 + \frac{d}{2} \), \( \frac{d}{2}< s_0 < s- 2 \) and \( 0< {{\mathtt {m}}}_1 < {{\mathtt {m}}}_2 \). For any initial data of the form \(({{\mathtt {m}}}+ \rho _0, \phi _0)\) with \( (\rho _0, \phi _0) \in H^s_0({\mathbb {T}}^d) \times \dot{H}^s({\mathbb {T}}^d) \) and \( {{\mathtt {m}}}_1< {{\mathtt {m}}}+ \rho _0(x) < {{\mathtt {m}}}_2 \), \( \forall x \in {\mathbb {T}}^d \), there exist \( T= T\big (\Vert (\rho _0, \phi _0)\Vert _{s_0+2}, \min _x ({{\mathtt {m}}}+ \rho _0(x)) \big ) >0 \) and a unique solution \(({{\mathtt {m}}}+ \rho , \phi )\) of (1.6) such that

$$\begin{aligned}&(\rho , \phi ) \in C^{0}\Big ([0,T], H_0^{s}({\mathbb {T}}^d,{\mathbb {R}})\times {\dot{H}}^{s}({\mathbb {T}}^d,{\mathbb {R}})\Big ) \\&\quad \cap C^{1}\Big ([0,T], H_0^{s-2}({\mathbb {T}}^d,{\mathbb {R}})\times {\dot{H}}^{s-2}({\mathbb {T}}^d,{\mathbb {R}}) \Big )\ \end{aligned}$$

and \( {{\mathtt {m}}}_1< {{\mathtt {m}}}+ \rho (t,x) < {{\mathtt {m}}}_2 \) holds for any \(t \in [0, T]\). Moreover, for \(|t| \le T\), the solution map \( (\rho _0, \phi _0) \mapsto (\rho (t, \cdot ), \phi (t, \cdot ) ) \) is locally defined and continuous in \(H^s_0({\mathbb {T}}^d) \times \dot{H}^s({\mathbb {T}}^d) \).

We consider system (1.6) on the homogeneous space \( \dot{H}^s \times \dot{H}^s \), that is we study

$$\begin{aligned} {\left\{ \begin{array}{ll} \partial _t \rho = - {{\mathtt {m}}}\Delta \phi - {\text {div}}((\Pi _0^\bot \rho ) \nabla \phi )\\ \partial _t \phi = -\frac{1}{2} | \nabla \phi |^2 - g({{\mathtt {m}}}+ \Pi _0^\bot \rho ) + K({{\mathtt {m}}}+\Pi _0^\bot \rho ) \Delta \rho + \frac{1}{2} K'({{\mathtt {m}}}+ \Pi _0^\bot \rho )| \nabla \rho |^2 \, \end{array}\right. } \end{aligned}$$
(1.7)

where \(\Pi _0^\perp \) is the projector onto the Fourier modes of index \(\ne 0\). For simplicity of notation we shall not distinguish between systems (1.7) and (1.6), which are equivalent via the isomorphism \( \Pi _0^\bot : \dot{H}^s({\mathbb {T}}^d) \rightarrow H_0^s({\mathbb {T}}^d)\). In Sect. 3, we paralinearize (1.6), i.e. (1.7), up to bounded semilinear terms (for which we do not need Bony paralinearization formula). Then, introducing a suitable complex variable, we transform it into a quasi-linear type Schrödinger equation, see system (3.4), defined in the phase space

$$\begin{aligned} {\dot{{\mathbf {H}}}}^s := \Big \{ U = {\begin{pmatrix}u \\ {\overline{u}}\end{pmatrix}} :\quad u \in \dot{H}^s({\mathbb {T}}^d,{\mathbb {C}}) \Big \}, \quad \left\| U \right\| _{s}^2 := \left\| U \right\| _{{\dot{{\mathbf {H}}}}^s}^2 = \left\| u \right\| _s^2 + \left\| {\overline{u}} \right\| _s^2. \end{aligned}$$
(1.8)

We use paradifferential calculus in the Weyl quantization, because it is quite convenient to prove energy estimates for this system. Since (3.4) is a quasi-linear system, in order to prove local well posedness (Proposition 4.1) we follow the strategy, initiated by Kato [20], of constructing inductively a sequence of linear problems whose solutions converge to the solution of the quasilinear equation. Such a scheme has been widely used, see e.g. [1, 8, 18, 22] and reference therein.

The equation (1.3) is a Hamiltonian PDE. We do not exploit explicitly this fact, but it is indeed responsible for the energy estimate of Proposition 4.4. The method of proof of Theorem 1.1 is similar to the one in Feola–Iandoli [19] for Hamiltonian quasi-linear Schrödinger equations on \( {\mathbb {T}}^d \) (and Alazard–Burq–Zuily [1] in the case of gravity-capillary water waves in \( {\mathbb {R}}^d \)). The main difference is that we aim to obtain the minimal smoothness assumption \( s > 2 + (d/2) \). This requires to optimize several arguments, and, in particular, to develop a sharp para-differential calculus for periodic functions that we report in the Appendix in a self-contained way. Some other technical differences are in the use of the modified energy (Sect. 4.2), the mollifiers (4.17) which enables to prove energy estimates independent of \( \varepsilon \) for the regularized system, the argument for the continuity of the flow in \( H^s \). We expect that our approach would enable to extend the local existence result of [19] to initial data fulfilling the minimal smoothness assumptions \( s > 2 + (d/2) \).

We now set some notation that will be used throughout the paper. Since \( K : {\mathbb {R}}_+ \rightarrow {\mathbb {R}}\) is positive, given \( 0< {{\mathtt {m}}}_1 < {{\mathtt {m}}}_2 \), there exist constants \(c_K, C_K > 0 \) such that

$$\begin{aligned} c_K\le K(\rho ) \le C_K, \qquad \forall \rho \in ({{\mathtt {m}}}_1, {{\mathtt {m}}}_2). \end{aligned}$$
(1.9)

Since the velocity potential \( \phi \) is defined up to a constant, we may assume in (1.6) that

$$\begin{aligned} g({{\mathtt {m}}}) = 0 \,. \end{aligned}$$
(1.10)

From now on we fix \(s_0\) so that

$$\begin{aligned} \frac{d}{2}< s_0 < s-2. \end{aligned}$$
(1.11)

The initial datum \(\rho _0 (x) \) belongs to the open subset of \( H^{s_0}_0({\mathbb {T}}^d)\) defined by

$$\begin{aligned} {\mathcal {Q}} := \big \{ \rho \in H^{s_0}_0 ({\mathbb {T}}^d)\ \ : \ \ {{\mathtt {m}}}_1<{{\mathtt {m}}}+ \rho (x)< {{\mathtt {m}}}_2 \big \} \end{aligned}$$
(1.12)

and we shall prove that, locally in time, the solution of (1.6) stays in this set.

We write \(a\lesssim b\) with the meaning \(a \le C b\) for some constant \(C >0\) which does not depend on relevant quantities.

2 Functional Setting and Paradifferential Calculus

The Sobolev norms \( \Vert \ \Vert _s \) satisfy interpolation inequalities (see e.g. Sect. 3.5 in [9]):

  1. (i)

    for all \( s \ge s_0 > \frac{d}{2} \), \( u, v \in H^s \),

    $$\begin{aligned} \Vert u v \Vert _s \lesssim \Vert u \Vert _{s_0} \Vert v \Vert _s + \Vert u \Vert _{s} \Vert v \Vert _{s_0}. \end{aligned}$$
    (2.1)
  2. (ii)

    Let \(s_0 > \frac{d}{2} \). For all \( 0 \le s \le s_0 \), \( v \in H^s \), \( u \in H^{s_0} \),

    $$\begin{aligned} \Vert u v \Vert _s \lesssim \Vert u \Vert _{s_0} \Vert v \Vert _s. \end{aligned}$$
    (2.2)
  3. (iii)

    For all \( s_1 < s_2 \), \( \theta \in [0,1] \) and \( u \in H^{s_2} \),

    $$\begin{aligned} \Vert u \Vert _{\theta s_1 + (1- \theta ) s_2} \le \Vert u \Vert _{s_1}^\theta \Vert u \Vert _{s_2}^{1-\theta }. \end{aligned}$$
    (2.3)
  4. (iv)

    For all \(a\le \alpha \le \beta \le b\), with \(\alpha + \beta = a + b\), \(u,v\in H^b \),

    $$\begin{aligned} \Vert u \Vert _{\alpha } \Vert v \Vert _{\beta } \le \Vert u \Vert _{a} \Vert v \Vert _{b} + \Vert u \Vert _{b} \Vert v \Vert _{a}. \end{aligned}$$
    (2.4)

Paradifferential calculus. We now introduce the notions of paradifferential calculus that will be used in the proof of Theorem 1.1. We develop it in the Weyl quantization since it is more convenient to get the energy estimates of Sect. 4. The main results are the continuity Theorem 2.4 and the composition Theorem 2.5, which require mild regularity assumptions of the symbols in the space variable (they are deduced by the sharper results proved in Theorems A.7 and A.8 in the Appendix). This is needed in order to prove the local existence Theorem 1.1 with the natural minimal regularity on the initial datum \( (\rho _0, \phi _0) \in H^s \times H^s \) with \( s > 2 + \frac{d}{2} \).

Along the paper \( {\mathscr {W}} \) may denote either the Banach space \( L^\infty ({\mathbb {T}}^d)\), or the Sobolev spaces \( H^s ({\mathbb {T}}^d)\), or the Hölder spaces \( W^{\varrho ,\infty } ({\mathbb {T}}^d)\), introduced in Definition A.3. Given a multi-index \( \beta \in {\mathbb {N}}_0^d \) we define \( |\beta | := \beta _1 + \cdots + \beta _d \).

Definition 2.1

(Symbols with finite regularity) Given \( m \in {\mathbb {R}}\) and a Banach space \({\mathscr {W}} \in \{ L^\infty ({\mathbb {T}}^d), H^s ({\mathbb {T}}^d), W^{\varrho ,\infty } ({\mathbb {T}}^d)\} \), we denote by \(\Gamma ^m_{\mathscr {W}}\) the space of functions \( a : {\mathbb {T}}^d \times {\mathbb {R}}^d \rightarrow {\mathbb {C}}\), \(a(x, \xi )\), which are \(C^\infty \) with respect to \(\xi \) and such that, for any \( \beta \in {\mathbb {N}}_0^d \), there exists a constant \(C_\beta >0\) such that

$$\begin{aligned} \big \Vert \partial _\xi ^\beta \, a(\cdot , \xi ) \big \Vert _{\mathscr {W}} \le C_\beta \, \left\langle \xi \right\rangle ^{m - |\beta |}, \quad \forall \xi \in {\mathbb {R}}^d . \end{aligned}$$
(2.5)

We denote by \(\Sigma ^m_{\mathscr {W}}\) the subclass of symbols \(a\in \Gamma ^m_{\mathscr {W}}\) which are spectrally localized, that is

$$\begin{aligned} \exists \, \delta \in (0, 1) \, :\qquad \quad {\widehat{a}}(j, \xi ) = 0, \quad \forall |j| \ge \delta \left\langle \xi \right\rangle , \end{aligned}$$
(2.6)

where \({\widehat{a}}(j, \xi ) := (2 \pi )^{-d} \int _{{\mathbb {T}}^d} a(x,\xi ) e^{- {{\text {i}}}j \cdot x} {\text {d}}x\), \( j \in {\mathbb {Z}}^d \), are the Fourier coefficients of the function \(x \mapsto a(x, \xi )\).

We endow \(\Gamma ^m_{\mathscr {W}}\) with the family of norms defined, for any \(n \in {\mathbb {N}}_0\), by

$$\begin{aligned} \left| a \right| _{m, {\mathscr {W}}, n}:= \max _{|\beta | \le n}\, \sup _{\xi \in {\mathbb {R}}^d} \, \big \Vert \left\langle \xi \right\rangle ^{-m+|\beta |} \, \partial _\xi ^{\beta } a(\cdot , \xi ) \big \Vert _{\mathscr {W}}. \end{aligned}$$
(2.7)

When \({\mathscr {W}} = H^s \), we also denote \(\Gamma ^m_s \equiv \Gamma ^m_{H^s}\) and \(\left| a \right| _{m, s, n} \equiv \left| a \right| _{m,H^s, n}\). We denote by \( \Gamma _s^m \otimes {{{\mathcal {M}}}}_2 ({\mathbb {C}}) \) the \( 2 \times 2 \) matrices \( A = \begin{pmatrix} a_{1} &{} \quad a_{2} \\ a_{3} &{} \quad a_{4} \end{pmatrix} \) of symbols in \( \Gamma _s^m \) and \( | A |_{m, {\mathscr {W}}, n} := \max _{i=1, \ldots , 4}\{ | a_{i} |_{m, {\mathscr {W}}, n}\} \). Similarly we denote by \( \Gamma _s^m \otimes {\mathbb {R}}^d \) the d-dimensional vectors of symbols in \( \Gamma _s^m \).

Let us make some simple remarks:

  • (\(\mathrm{i}\)) given a function \( a(x) \in {\mathscr {W}} \) then \(a(x) \in \Gamma ^0_{{\mathscr {W}}} \) and

    $$\begin{aligned} \left| u \right| _{0, {\mathscr {W}}, n} = \left\| u \right\| _{{\mathscr {W}}}, \forall n \in {\mathbb {N}}_0. \end{aligned}$$
    (2.8)
  • (\(\mathrm{ii}\)) For any \(s_0 > \frac{d}{2}\) and \(0 \le \varrho ' \le \varrho \), we have that

    $$\begin{aligned} \left| a \right| _{m, L^\infty , n} \lesssim \left| a \right| _{m, W^{\varrho ', \infty }, n} \lesssim \left| a \right| _{m, {W^{\varrho , \infty }}, n} \lesssim \left| a \right| _{m, H^{s_0+\varrho },n} , \quad \forall n \in {\mathbb {N}}_0. \end{aligned}$$
    (2.9)
  • (\(\mathrm{iii}\)) If \( a \in \Gamma ^m_{{\mathscr {W}}}\), then, for any \( \alpha \in {\mathbb {N}}_0^d \), we have \( \partial _\xi ^\alpha a \in \Gamma ^{m-|\alpha |}_{{\mathscr {W}}} \) and

    $$\begin{aligned} | \partial _\xi ^\alpha a |_{m-|\alpha |,{\mathscr {W}},n} \lesssim | a|_{m,{\mathscr {W}},n+|\alpha |}, \quad \forall n \in {\mathbb {N}}_0. \end{aligned}$$
    (2.10)
  • (\(\mathrm{iv}\)) If \( a \in \Gamma ^m_{H^{s}} \), resp. \( a \in \Gamma ^m_{W^{\varrho , \infty }} \), then \( \partial _x^\alpha a \in \Gamma ^m_{H^{s-|\alpha |}} \), resp. \( \partial _x^\alpha a \in \Gamma ^m_{W^{\varrho -|\alpha |,\infty }} \), and

    $$\begin{aligned} | \partial _x^\alpha a|_{m,s-|\alpha |,n} \lesssim | a|_{m,s,n}, \quad {\text {resp.}} \ |\partial _x^\alpha a|_{m,W^{\varrho -|\alpha |,\infty },n} \lesssim |a|_{m,W^{\varrho ,\infty },n}, \quad \forall n \in {\mathbb {N}}_0 \,.\nonumber \\ \end{aligned}$$
    (2.11)
  • (\(\mathrm{v}\)) If \( a,b \in \Gamma ^m_{{\mathscr {W}}} \) then \( ab \in \Gamma ^m_{{\mathscr {W}}} \) with \( |ab|_{m+m',{\mathscr {W}},n} \lesssim |a|_{m,{\mathscr {W}},N} |b|_{m',{\mathscr {W}},n}\) for any \( n \in {\mathbb {N}}_0 \). In particular, if \( a,b \in \Gamma ^m_{s} \) with \( s > d / 2 \) then \( ab \in \Gamma ^{m+m'}_{s} \) and

    $$\begin{aligned} | a b |_{m+m',s,n} \lesssim | a |_{m,s,n} | b |_{m',s_0,n} + | a |_{m,s_0,n} | b |_{m',s,n}, \quad \forall n \in {\mathbb {N}}_0. \end{aligned}$$
    (2.12)

    Let \(\epsilon \in (0,1)\) and consider a \( C^\infty \), even cut-off function \(\chi :{\mathbb {R}}^d \rightarrow [0,1]\) such that

    $$\begin{aligned} \chi (\xi ) = {\left\{ \begin{array}{ll} 1 &{} \quad \text{ if } |\xi | \le 1.1 \\ 0 &{} \quad \text{ if } |\xi | \ge 1.9, \end{array}\right. } \qquad \chi _\epsilon (\xi ) := \chi \left( \frac{\xi }{\epsilon }\right) . \end{aligned}$$
    (2.13)

    Given a symbol a in \( \Gamma ^m_{{\mathscr {W}}} \) we define the regularized symbol

    $$\begin{aligned} a_\chi (x, \xi ) := \chi _{\epsilon \left\langle \xi \right\rangle }(D) a(x, \xi ) = \sum _{j \in {\mathbb {Z}}^d} \chi _\epsilon \Big ( \frac{j}{\langle \xi \rangle } \Big ) \, {\widehat{a}} (j, \xi ) \, e^{{\text {i}}j \cdot x}. \end{aligned}$$
    (2.14)

Note that \(a_\chi \) is analytic in x (it is a trigonometric polynomial) and it is spectrally localized.

In order to define the Bony–Weyl quantization of a symbol \( a (x, \xi )\) we first remind the Weyl quantization formula

$$\begin{aligned} {\text {Op}}^W {(a)}[u] := \sum _{j \in {\mathbb {Z}}^d} \Big ( \sum _{k \in {\mathbb {Z}}^d} {\widehat{a}} \Big (j-k, \frac{k + j}{2}\Big ) \, u_k \Big ) e^{{\text {i}}j \cdot x }. \end{aligned}$$
(2.15)

Definition 2.2

(Bony–Weyl quantization) Given a symbol \(a \in \Gamma ^m_{{\mathscr {W}}}\), we define the Bony–Weyl paradifferential operator \( {\text {Op}}^{BW} {(a)} = {\text {Op}}^{W} (a_\chi ) \) that acts on a periodic function u as

$$\begin{aligned} \begin{aligned} \left( {\text {Op}}^{BW} {(a)}[u] \right) (x)&:= \sum _{j \in {\mathbb {Z}}^d} \Big ( \sum _{k \in {\mathbb {Z}}^d} {\widehat{a}}_\chi \Big (j-k, \frac{j+k}{2}\Big ) \, u_k \Big ) e^{{\text {i}}j \cdot x } \\&= \sum _{j \in {\mathbb {Z}}^d} \Big ( \sum _{k \in {\mathbb {Z}}^d} {\widehat{a}} \Big (j-k, \frac{j+k}{2}\Big ) \, \chi _\epsilon \Big ( \frac{j-k}{\langle j+k \rangle }\Big ) \, u_k \Big ) e^{{\text {i}}j \cdot x }. \end{aligned} \end{aligned}$$
(2.16)

If \( A = \begin{pmatrix} a_{1} &{} \quad a_{2} \\ a_{3} &{} \quad a_{4} \end{pmatrix} \) is a matrix of symbols in \( \Gamma _s^m \), then \( {\text {Op}}^{BW} (A) \) is defined as the matrix valued operator \( \begin{pmatrix} {\text {Op}}^{BW} (a_{1}) &{} \quad {\text {Op}}^{BW} (a_{2}) \\ {\text {Op}}^{BW} (a_{3}) &{} {\text {Op}}^{BW} (a_{4}) \end{pmatrix} \).

Given a symbol \( a(\xi ) \) independent of x, then \( {\text {Op}}^{BW} {(a)} \) is the Fourier multiplier operator

$$\begin{aligned} {\text {Op}}^{BW} {(a)} u = a(D) u = \sum _{j \in {\mathbb {Z}}^d} a(j) \, u_j \, e^{{\text {i}}j \cdot x }. \end{aligned}$$

Note that if \( \chi _\epsilon \Big ( \frac{k-j}{\langle k + j \rangle }\Big ) \ne 0 \) then \( |k-j| \le \epsilon \langle j + k \rangle \) and therefore, for \( \epsilon \in (0,1)\),

$$\begin{aligned} \frac{1-\epsilon }{1+\epsilon } |k| \le |j| \le \frac{1+\epsilon }{1-\epsilon }|k|, \quad \forall j, k \in {\mathbb {Z}}^d. \end{aligned}$$
(2.17)

This relation shows that the action of a para-differential operator does not spread much the Fourier support of functions. In particular \( {\text {Op}}^{\text {BW}}(a) \) sends a constant function into a constant function and therefore \( {\text {Op}}^{\text {BW}}(a) \) sends homogenous spaces into homogenous spaces.

Remark 2.3

Actually, if \( \chi _\epsilon \big ( \frac{k-j}{\langle k + j \rangle }\big ) \ne 0 \), \( \epsilon \in (0,1/4) \), then \( |j| \le |j+k| \le 3 |j| \), for all \( j,k \in {\mathbb {Z}}^d \).

Along the paper we shall use the following results concerning the action of a paradifferential operator in Sobolev spaces.

Theorem 2.4

(Continuity of Bony–Weyl operators) Let \( a \in \Gamma ^m_{s_0} \), resp. \( a \in \Gamma ^m_{L^\infty } \), with \( m \in {\mathbb {R}}\). Then \({{\text {Op}}}^{\text {BW}}\!\left( a\right) \) extends to a bounded operator \(\dot{H}^{s} \rightarrow \dot{H}^{s-m}\) for any \( s \in {\mathbb {R}}\) satisfying the estimate, for any \( u \in \dot{H}^s \),

$$\begin{aligned}&\left\| {{\text {Op}}}^{\text {BW}}\!\left( a\right) u \right\| _{{s-m}} \lesssim \, \left| a \right| _{m, s_0, 2(d+1)} \, \left\| u \right\| _{{s}} \end{aligned}$$
(2.18)

Moreover, for any \(\varrho \ge 0\), \( s \in {\mathbb {R}}\), \( u \in \dot{H}^s ({\mathbb {T}}^d)\),

$$\begin{aligned} \left\| {{\text {Op}}}^{\text {BW}}\!\left( a\right) u \right\| _{{s-m- \varrho }} \lesssim \, \left| a \right| _{m, {s_0-\varrho }, 2(d+1)} \, \left\| u \right\| _{{s}}. \end{aligned}$$
(2.19)

Proof

Since \( {{\text {Op}}}^{\text {BW}}\!\left( a\right) = {{\text {Op}}}^W\!\left( a_\chi \right) \), the estimate (2.18) follows by (A.35), (A.21) and \( \left| a \right| _{m, L^{\infty }, N} \lesssim \left| a \right| _{m, {s_0}, N} \). Note that the condition on the Fourier support of \(a_\chi \) in Theorem A.7 is automatically satisfied provided \(\epsilon \) in (2.13) is sufficiently small. To prove (2.19) we use also (A.22). \(\square \)

The second result of symbolic calculus that we shall use regards composition for Bony–Weyl paradifferential operators at the second order (as required in the paper) with mild smoothness assumptions for the symbols in the space variable x. Given symbols \(a \in \Gamma ^m_{s_0+\varrho }\), \(b \in \Gamma ^{m'}_{s_0+\varrho }\) with \(m, m' \in {\mathbb {R}}\) and \(\varrho \in (0,2]\) we define

$$\begin{aligned} a\#_\varrho b := {\left\{ \begin{array}{ll} ab, &{} \varrho \in (0,1] \\ ab + \frac{1}{2{\text {i}}}\{a, b\}, &{} \varrho \in (1,2], \quad {\text {where}} \quad \{a,b\} := \nabla _{\xi } a \cdot \nabla _{x} b - \nabla _x a \cdot \nabla _\xi b, \end{array}\right. } \end{aligned}$$
(2.20)

is the Poisson bracket between \( a (x, \xi )\) and \( b(x, \xi ) \). By (2.10) and (2.12) we have that ab is a symbol in \( \Gamma ^{m+m'}_{s_0+\varrho } \) and \( \{a, b\} \) is in \( \Gamma ^{m+m'-1}_{s_0+\varrho -1} \). The next result follows directly by Theorem A.8 and (2.9).

Theorem 2.5

(Composition) Let \(a \in \Gamma ^m_{s_0+\varrho }\), \(b \in \Gamma ^{m'}_{s_0+\varrho }\) with \(m, m' \in {\mathbb {R}}\) and \(\varrho \in (0,2]\). Then

$$\begin{aligned} {{\text {Op}}}^{\text {BW}}\!\left( a\right) {{\text {Op}}}^{\text {BW}}\!\left( b\right)&= {{\text {Op}}}^{\text {BW}}\!\left( a\#_\varrho b\right) + R^{-\varrho }(a,b) \end{aligned}$$
(2.21)

where the linear operator \(R^{-\varrho }(a,b):\dot{H}^s \rightarrow \dot{H}^{s-(m+m')+\varrho }\), \(\forall s \in {\mathbb {R}}\), satisfies, for any \( u \in \dot{H}^s \),

$$\begin{aligned}&\left\| R^{-\varrho }(a,b)u \right\| _{{s -(m+m') +\varrho }} \lesssim \left( \left| a \right| _{m, {s_0+\varrho }, N} \, \left| b \right| _{m', {s_0}, N} + \left| a \right| _{m, {s_0}, N} \, \left| b \right| _{m', {s_0+\varrho }, N} \right) \left\| u \right\| _{s} \end{aligned}$$
(2.22)

where \( N \ge 3d + 4 \).

A useful corollary of Theorems 2.5 and 2.4 [using also (2.10)–(2.12)] is the following:

Corollary 2.6

Let \(a\in \Gamma ^{m}_{s_0+2}\), \(b\in \Gamma ^{m'}_{s_0+2}\), \(c \in \Gamma ^{m''}_{s_0+2}\) with \(m, m', m'' \in {\mathbb {R}}\). Then

$$\begin{aligned} {{\text {Op}}}^{\text {BW}}\!\left( a\right) \circ {{\text {Op}}}^{\text {BW}}\!\left( b\right) \circ {{\text {Op}}}^{\text {BW}}\!\left( c\right) = {{\text {Op}}}^{\text {BW}}\!\left( abc\right) + R_1(a,b,c)+ R_0(a,b,c), \end{aligned}$$
(2.23)

where

$$\begin{aligned} R_1(a,b,c):= {\text {Op}}^{BW} \big ( \{ a,c\}b+ \{b,c\}a+\{ a,b\}c \big ) \end{aligned}$$
(2.24)

satisfies \( R_1(a,b,c)=-R_1(c,b,a) \) and \(R_0(a,b,c)\) is a bounded operator \(\dot{H}^s \rightarrow \dot{H}^{s-(m+m'+m'')+2}\), \(\, \forall s\in {\mathbb {R}}\), satisfying, for any \( u \in \dot{H}^s \),

$$\begin{aligned} \Vert R_0(a,b,c)\Vert _{s-(m+m'+m'') + 2} \lesssim |a|_{m,s_0+2,N} \ |b|_{m',s_0+2,N}\ |c|_{m'',s_0+2,N} \ \Vert u \Vert _s \end{aligned}$$
(2.25)

where \( N \ge 3d+ 5 \).

We now provide the Bony-paraproduct decomposition for the product of Sobolev functions in the Bony–Weyl quantization. Recall that \( \Pi _0^\bot \) denotes the projector on the subspace \( H^s_0 \).

Lemma 2.7

(Bony paraproduct decomposition) Let \(u \in H^s\), \(v \in H^r\) with \(s + r \ge 0\). Then

$$\begin{aligned} uv = {{\text {Op}}}^{\text {BW}}\!\left( u\right) v + {{\text {Op}}}^{\text {BW}}\!\left( v\right) u + R(u,v) \end{aligned}$$
(2.26)

where the bilinear operator \(R:H^s \times H^r \rightarrow H^{s+r-s_0}\) is symmetric and satisfies the estimate

$$\begin{aligned} \left\| R(u,v) \right\| _{s+ r - s_0} \lesssim \left\| u \right\| _s \, \left\| v \right\| _{r}. \end{aligned}$$
(2.27)

Moreover \( R(u,v) = R(\Pi _0^\bot u, \Pi _0^\bot v) - u_0 v_0 \) and then

$$\begin{aligned} \Vert \Pi _0^\bot R(u,v) \Vert _{s+ r - s_0} \lesssim \Vert \Pi _0^\bot u \Vert _s \, \Vert \Pi _0^\bot v \Vert _{r}. \end{aligned}$$
(2.28)

Proof

Introduce the function \(\theta _\epsilon (j,k)\) by

$$\begin{aligned} 1 =\chi _\epsilon \Big ( \frac{j-k}{\langle j+k \rangle }\Big ) + \chi _\epsilon \Big ( \frac{k}{\langle 2j-k \rangle }\Big ) + \theta _\epsilon (j,k). \end{aligned}$$
(2.29)

Note that \(\left| \theta _\epsilon (j, k ) \right| \le 1\). Let \( \Sigma := \{ (j, k) \in {\mathbb {Z}}^d \times {\mathbb {Z}}^d \, : \, \theta _\epsilon (j,k) \ne 0 \} \) denote the support of \(\theta _\epsilon \). We claim that

$$\begin{aligned} (j,k) \in \Sigma \qquad \Longrightarrow \qquad |j| \le C_\epsilon \min ( |j-k|, \ |k |). \end{aligned}$$
(2.30)

Indeed, recalling the definition of the cut-off function \( \chi \) in (2.13), we first note thatFootnote 1

$$\begin{aligned} \Sigma= & {} \{(0,0)\} \cup \Big \{ |j-k| \ge \epsilon \left\langle j + k \right\rangle ,\, |k| \ge \epsilon \left\langle 2j - k \right\rangle \Big \}. \end{aligned}$$

Thus, for any \( (j,k) \in \Sigma \),

$$\begin{aligned} |j|&\le \frac{1}{2} |j-k| + \frac{1}{2} |j+k| \le \left( \frac{1}{2} + \frac{1}{2\epsilon } \right) \, |j-k|,\\ |j|&\le \frac{1}{2} |2j - k| + \frac{1}{2} |k| \le \left( \frac{1}{2} + \frac{1}{2\epsilon } \right) \, |k| \end{aligned}$$

proving (2.30). Using (2.29) we decompose

$$\begin{aligned} uv&= \sum _{j,k } {\widehat{u}}_{j-k} \, \chi _\epsilon \Big ( \frac{j-k}{\langle j+k \rangle }\Big ) \, {\widehat{v}}_{k} \, e^{{\text {i}}j \cdot x}\\&\quad + \sum _{j,k} {\widehat{v}}_{k} \, \chi _\epsilon \Big ( \frac{k}{\langle 2j-k \rangle }\Big )\, {\widehat{u}}_{j-k} \, e^{{\text {i}}j \cdot x} + \sum _{j, k} \theta _\epsilon (j, k){\widehat{u}}_{j-k}\, {\widehat{v}}_{k} \, e^{{\text {i}}j \cdot x} \\&= {{\text {Op}}}^{\text {BW}}\!\left( u\right) v + {{\text {Op}}}^{\text {BW}}\!\left( v\right) u + R(u,v). \end{aligned}$$

By (2.30), \( s+ r \ge 0 \), and the Cauchy–Schwartz inequality, we get

$$\begin{aligned} \left\| R(u,v) \right\| _{s+r-s_0}^2&\le \sum _j \left\langle j \right\rangle ^{2(s+r-s_0)} \Big | \sum _{k} \theta _\epsilon (j, k){\widehat{u}}_{j-k}\, {\widehat{v}}_{k} \Big |^2 \\&\lesssim \sum _j \left\langle j \right\rangle ^{-2s_0} \Big | \sum _{k} \left\langle j-k \right\rangle ^{s} \, |{\widehat{u}}_{j-k}|\, \left\langle k \right\rangle ^{r }\, |{\widehat{v}}_{k}| \Big |^2 \lesssim \left\| u \right\| _s^2 \, \left\| v \right\| _r^2 \end{aligned}$$

proving (2.27). Finally, since on the support of \( \theta _\epsilon \) we have or \( (j,k) = (0,0)\) or \( j - k \ne 0 \) and \( k \ne 0 \), we deduce that

$$\begin{aligned} R(u,v) = \theta _\epsilon (0,0){\widehat{u}}_{0}\, {\widehat{v}}_{0} + \sum _{j-k \ne 0, k \ne 0} \theta _\epsilon (j, k){\widehat{u}}_{j-k}\, {\widehat{v}}_{k} \, e^{{\text {i}}j \cdot x} = - {\widehat{u}}_{0}\, {\widehat{v}}_{0} + R(\Pi _0^\bot u,\Pi _0^\bot v) \end{aligned}$$

and we deduce (2.28). \(\square \)

Composition estimates.

We will use the following Moser estimates for composition of functions in Sobolev spaces.

Theorem 2.8

Let \(I \subseteq {\mathbb {R}}\) be an open interval and \(F\in C^\infty (I; {\mathbb {C}})\) a smooth function. Let \(J \subset I\) be a compact interval. For any function \(u, v \in H^s({\mathbb {T}}^d, {\mathbb {R}})\), \(s>\frac{d}{2} \), with values in J, we have

$$\begin{aligned}&\Vert F(u)\Vert _{s} \le C({s,F,J}) \left( 1+ \Vert u \Vert _{s}\right) , \nonumber \\&\Vert F(u)- F(v)\Vert _{s} \le C({s,F,J}) \left( \Vert u-v \Vert _{s} + (\Vert u \Vert _s+ \Vert v \Vert _s)\Vert u - v \Vert _{L^\infty }\right) \, \nonumber \\&\Vert F(u)\Vert _{s} \le C({s,F,J}) \Vert u \Vert _{s} \quad {\text {if}} \quad F(0) = 0. \end{aligned}$$
(2.31)

Proof

Take an extension \({\tilde{F}}\in C^\infty ({\mathbb {R}};{\mathbb {C}})\) such that \({\tilde{F}}_{| I}= F \). Then \( F(u) = {\tilde{F}}(u)\) for any \(u \in H^s({\mathbb {T}}^d;{\mathbb {R}})\) with values in J, and apply the usual Moser estimate, see e.g. [3], replacing the Littlewood–Paley decomposition on \({\mathbb {R}}^d\) with the one on \({\mathbb {T}}^d\) in (A.12). \(\square \)

3 Paralinearization of (EK)-System and Complex Form

In this section we paralinearize the Euler–Korteweg system (1.6) and write it in terms of the complex variable

$$\begin{aligned} u := \frac{1}{\sqrt{2}} \left( \frac{{{\mathtt {m}}}}{K({{\mathtt {m}}})}\right) ^{-1/4} \rho + \frac{{\text {i}}}{\sqrt{2}}\, \left( \frac{{{\mathtt {m}}}}{K({{\mathtt {m}}})}\right) ^{1/4} \phi , \quad \rho \in \dot{H}^s, \ \phi \in \dot{H}^s. \end{aligned}$$
(3.1)

The variable \( u \in \dot{H}^s \). We denote this change of coordinates in \( \dot{H}^s \times \dot{H}^s \) by

$$\begin{aligned}&{\begin{pmatrix}u \\ {\overline{u}}\end{pmatrix}}= {\textit{C}}^{-1} {\begin{pmatrix}\rho \\ \phi \end{pmatrix}},\nonumber \\&\quad {\textit{C}}:=\frac{1}{\sqrt{2}} \begin{pmatrix} \left( \frac{{{\mathtt {m}}}}{K({{\mathtt {m}}})}\right) ^{\frac{1}{4}}&{}\quad \left( \frac{{{\mathtt {m}}}}{K({{\mathtt {m}}})}\right) ^{\frac{1}{4}}\\ -{\text {i}}\left( \frac{{{\mathtt {m}}}}{K({{\mathtt {m}}})}\right) ^{-\frac{1}{4}}&{}\quad {\text {i}}\left( \frac{{{\mathtt {m}}}}{K({{\mathtt {m}}})}\right) ^{-\frac{1}{4}}\end{pmatrix},\quad {\textit{C}}^{-1}=\frac{1}{\sqrt{2}} \begin{pmatrix} \left( \frac{{{\mathtt {m}}}}{K({{\mathtt {m}}})}\right) ^{-\frac{1}{4}}&{}\quad {\text {i}}\left( \frac{{{\mathtt {m}}}}{K({{\mathtt {m}}})}\right) ^{\frac{1}{4}}\\ \left( \frac{{{\mathtt {m}}}}{K({{\mathtt {m}}})}\right) ^{-\frac{1}{4}}&{}\quad -{\text {i}}\left( \frac{{{\mathtt {m}}}}{K({{\mathtt {m}}})}\right) ^{\frac{1}{4}}\end{pmatrix}.\nonumber \\ \end{aligned}$$
(3.2)

We also define the matrices

(3.3)

Proposition 3.1

(Paralinearized Euler–Korteweg equations in complex coordinates) The (EK)-system (1.6) can be written in terms of the complex variable \( U := {\begin{pmatrix}u \\ {\overline{u}}\end{pmatrix}} \) with u defined in (3.1), in the paralinearized form

$$\begin{aligned} \begin{aligned} \partial _t U&= \ {\mathbb {J}}\, \Big [ {{\text {Op}}}^{\text {BW}}\!\left( A_2(U; x, \xi ) + A_1(U; x, \xi ) \right) \Big ] U+ R(U) \end{aligned} \end{aligned}$$
(3.4)

where, for any function \(U\in \dot{\mathbf{H}}^{s_0+2} \) such that

$$\begin{aligned} \rho (U) := \frac{1}{\sqrt{2}} \left( \frac{{{\mathtt {m}}}}{K({{\mathtt {m}}})}\right) ^{1/4} \Pi _0^\bot (u + {\overline{u}}) \in {\mathcal {Q}}\ \ {\text {(see}}\,(1.12)) , \end{aligned}$$
(3.5)

one has

  1. (i)

    \(A_2(U;x,\xi ) \in \Gamma _{s_0+2}^2 \otimes {\mathcal {M}}_2({\mathbb {C}}) \) is the matrix of symbols

    $$\begin{aligned} A_2(U; x, \xi ) := \sqrt{{{\mathtt {m}}}K({{\mathtt {m}}})} |\xi |^2 \begin{bmatrix} 1 + {\mathtt {a}}_+(U; x) &{} {\mathtt {a}}_-(U; x) \\ {\mathtt {a}}_-(U; x) &{} 1+ {\mathtt {a}}_+(U; x) \end{bmatrix} \end{aligned}$$
    (3.6)

    where \({\mathtt {a}}_\pm (U; x) \in \Gamma _{s_0+2}^0\) are the \(\xi \)-independent functions

    $$\begin{aligned} \begin{aligned}&{\mathtt {a}}_\pm (U; x) := \frac{1}{2} \left( \frac{K(\rho + {{\mathtt {m}}})- K({{\mathtt {m}}}) }{K({{\mathtt {m}}})} \pm \frac{\rho }{{{\mathtt {m}}}} \right) . \end{aligned} \end{aligned}$$
    (3.7)
  2. (ii)

    \(A_1(U; x, \xi ) \in \Gamma _{s_0+1}^1 \otimes {\mathcal {M}}_2({\mathbb {C}}) \) is the diagonal matrix of symbols

    $$\begin{aligned} A_1(U; x, \xi ) := \begin{bmatrix} {\mathtt {b}}(U; x) \cdot \xi &{} 0 \\ 0 &{} -{\mathtt {b}}(U; x) \cdot \xi \end{bmatrix},\qquad {\mathtt {b}}(U; x) := \nabla \phi \in \Gamma _{s_0+1}^0\otimes {\mathbb {R}}^d.\nonumber \\ \end{aligned}$$
    (3.8)

    Moreover for any \(\sigma \ge 0 \) there exists a non decreasing function \({\mathtt {C}}( \ ) : {\mathbb {R}}_+ \rightarrow {\mathbb {R}}_+ \) (depending on K) such that, for any \(U,V\in \dot{\mathbf{H}}^{s_0} \) with \(\rho (U), \rho (V) \in {\mathcal {Q}}\), \(W \in {\dot{{\mathbf {H}}}}^{{\sigma }+2}\) and \(j=1,2\), we have

    $$\begin{aligned}&\Vert {{\text {Op}}}^{\text {BW}}\!\left( A_j(U)\right) W\Vert _{{\sigma }} \le {\mathtt {C}}\left( \Vert U\Vert _{s_0}\right) \Vert W\Vert _{{\sigma }+2} \end{aligned}$$
    (3.9)
    $$\begin{aligned}&\Vert {{\text {Op}}}^{\text {BW}}\!\left( A_j(U)-A_j(V)\right) W\Vert _{{\sigma }}\le {\mathtt {C}}\left( \Vert U\Vert _{s_0}, \Vert V\Vert _{s_0}\right) \Vert W\Vert _{{\sigma }+2} \Vert U-V\Vert _{s_0} \end{aligned}$$
    (3.10)

    where in (3.10) we denoted by \({\mathtt {C}}(\cdot , \cdot ):= {\mathtt {C}}\left( \max \{\cdot ,\cdot \}\right) \).

  3. (iii)

    The vector field R(U) satisfies the following “semilinear” estimates: for any \(\sigma \ge s_0 > d / 2 \) there exists a non decreasing function \({\mathtt {C}}( \ ) : {\mathbb {R}}_+ \rightarrow {\mathbb {R}}_+ \) (depending also on gK) such that, for any \(U,V\in \dot{\mathbf{H}}^{\sigma +2} \) such that \(\rho (U), \rho (V) \in {\mathcal {Q}}\), we have

    $$\begin{aligned}&\Vert R(U) \Vert _{ \sigma } \le {\mathtt {C}}\left( \Vert U\Vert _{ {s_0+2}} \right) \Vert U\Vert _{ \sigma }, \qquad \Vert R(U)\Vert _{ {\sigma }} \le {\mathtt {C}}\left( \Vert U\Vert _{ {s_0}}\right) \Vert U\Vert _{ {\sigma +2}}, \end{aligned}$$
    (3.11)
    $$\begin{aligned}&\Vert R(U)-R(V) \Vert _{ \sigma } \le {\mathtt {C}}\left( \Vert U\Vert _{ {s_0+2}}, \, \Vert V\Vert _{ {s_0+2}}\right) \Vert U-V\Vert _{{\sigma }}\nonumber \\&\qquad \qquad \qquad \qquad \qquad \quad \ + {\mathtt {C}}\left( \Vert U\Vert _{ {\sigma }}, \, \Vert V\Vert _{ {\sigma }}\right) \Vert U-V\Vert _{{s_0+2}} \end{aligned}$$
    (3.12)
    $$\begin{aligned}&\Vert R(U)-R(V) \Vert _{ s_0} \le {\mathtt {C}}\left( \Vert U\Vert _{ {s_0+2}}, \, \Vert V\Vert _{ {s_0+2}}\right) \Vert U-V\Vert _{{s_0}} \,, \end{aligned}$$
    (3.13)

    where in (3.12) and (3.13) we denoted again by \({\mathtt {C}}(\cdot , \cdot ):= {\mathtt {C}}\left( \max \{\cdot ,\cdot \}\right) \).

Proof

We first paralinearize the original equations (1.6), then we switch to complex coordinates.

Step 1: paralinearization of (1.6). We apply several times the paraproduct Lemma 2.7 and the composition Theorem 2.5. In the following we denote by \(R^p\) the remainder that comes from Lemma 2.7, and by \(R^{-\varrho }\), \(\varrho =1,2\), the remainder that comes from Theorem 2.5. We shall adopt the following convention: given \({\mathbb {R}}^d\)-valued symbols \( a = (a_j)_{j=1,\ldots , d} \), \( b = (b_j)_{j=1,\ldots , d} \) in some class \(\Gamma ^m_s\otimes {\mathbb {R}}^d\), we denote \( R^p(a, b) := \sum _{j=1}^d R^p(a_j, b_j) \),

$$\begin{aligned} R^{-\varrho }(a, b) := \sum _{j=1}^d R^{-\varrho }(a_j, b_j) \quad {\text {and}} \quad {{\text {Op}}}^{\text {BW}}\!\left( a\right) \cdot {{\text {Op}}}^{\text {BW}}\!\left( b\right) := \sum _{j=1}^d {{\text {Op}}}^{\text {BW}}\!\left( a_j\right) {{\text {Op}}}^{\text {BW}}\!\left( b_j\right) . \end{aligned}$$

We paralinearize the terms in the first line of (1.6). We have \( \Delta \phi = - {{\text {Op}}}^{\text {BW}}\!\left( |\xi |^2\right) \phi \) and \( {\text {div}}(\rho \nabla \phi ) = \nabla \rho \cdot \nabla \phi + \rho \Delta \phi \) can be written as

$$\begin{aligned} \rho \Delta \phi&= -{{\text {Op}}}^{\text {BW}}\!\left( \rho |\xi |^2 + \nabla \rho \cdot {\text {i}}\xi \right) \phi \nonumber \\&\quad + {{\text {Op}}}^{\text {BW}}\!\left( \Delta \phi \right) \rho + R^p(\rho , \Delta \phi )+ R^{-2}(\rho , |\xi |^2)\phi , \end{aligned}$$
(3.14)
$$\begin{aligned} \nabla \rho \cdot \nabla \phi&= {{\text {Op}}}^{\text {BW}}\!\left( \nabla \rho \cdot {\text {i}}\xi \right) \phi + {{\text {Op}}}^{\text {BW}}\!\left( \nabla \phi \cdot {\text {i}}\xi \right) \rho \nonumber \\&\quad + R^p(\nabla \rho , \nabla \phi ) + R^{-1}(\nabla \rho , {\text {i}}\xi )\phi +R^{-1}(\nabla \phi , {\text {i}}\xi )\rho . \end{aligned}$$
(3.15)

Then we paralinearize the terms in the second line of (1.6). We have

$$\begin{aligned} \frac{1}{2} | \nabla \phi |^2&= {{\text {Op}}}^{\text {BW}}\!\left( \nabla \phi \cdot {\text {i}}\xi \right) \phi \nonumber \\&\quad + \frac{1}{2} R^p(\nabla \phi ,\nabla \phi )+R^{-1}(\nabla \phi , {\text {i}}\xi )\phi . \end{aligned}$$
(3.16)

Using (1.10) we regard the semilinear term

$$\begin{aligned} g({{\mathtt {m}}}+ \rho ) = g({{\mathtt {m}}}+ \rho ) - g({{\mathtt {m}}}) =: R(\rho ) \end{aligned}$$
(3.17)

directly as a remainder. Moreover, writing \( \Delta \rho = - {{\text {Op}}}^{\text {BW}}\!\left( |\xi |^2\right) \rho \), we get

$$\begin{aligned} K({{\mathtt {m}}}+ \rho )\Delta \rho&= {{\text {Op}}}^{\text {BW}}\!\left( K({{\mathtt {m}}}+\rho )\right) \Delta \rho +{{\text {Op}}}^{\text {BW}}\!\left( \Delta \rho \right) K({{\mathtt {m}}}+\rho )+R^p(\Delta \rho , K({{\mathtt {m}}}+\rho )) \nonumber \\&= - {{\text {Op}}}^{\text {BW}}\!\left( K({{\mathtt {m}}}+\rho )|\xi |^2 + K'({{\mathtt {m}}}+\rho ) \nabla \rho \cdot {\text {i}}\xi \right) \rho \nonumber \\&\quad +{{\text {Op}}}^{\text {BW}}\!\left( \Delta \rho \right) K({{\mathtt {m}}}+\rho )+ R^p(\Delta \rho , K({{\mathtt {m}}}+\rho )) - R^{-2}(K({{\mathtt {m}}}+\rho ),|\xi |^2)\rho . \end{aligned}$$
(3.18)

Finally, using for \( \frac{1}{2} | \nabla \rho |^2 \) the expansion (3.16) for \(\rho \) instead of \( \phi \), we obtain

$$\begin{aligned} \frac{1}{2} K'({{\mathtt {m}}}+ \rho )| \nabla \rho | ^2&= \frac{1}{2} {{\text {Op}}}^{\text {BW}}\!\left( K'({{\mathtt {m}}}+\rho )\right) |\nabla \rho |^2+ \frac{1}{2} {{\text {Op}}}^{\text {BW}}\!\left( |\nabla \rho |^2\right) K'({{\mathtt {m}}}+\rho ) \\&\quad + \frac{1}{2} R^p(|\nabla \rho |^2, K'({{\mathtt {m}}}+\rho )) = {{\text {Op}}}^{\text {BW}}\!\left( K'({{\mathtt {m}}}+ \rho ) \nabla \rho \cdot \, {\text {i}}\xi \right) \rho + {\mathtt R}(\rho ) \end{aligned}$$

where

$$\begin{aligned} {\mathtt R} (\rho )&:= \frac{1}{2}{{\text {Op}}}^{\text {BW}}\!\left( |\nabla \rho |^2\right) K'({{\mathtt {m}}}+\rho )+ \frac{1}{2} R^p(|\nabla \rho |^2, K'({{\mathtt {m}}}+\rho ))\end{aligned}$$
(3.19)
$$\begin{aligned}&\quad + \frac{1}{2} {{\text {Op}}}^{\text {BW}}\!\left( K'({{\mathtt {m}}}+\rho )\right) R^p(\nabla \rho , \nabla \rho )\end{aligned}$$
(3.20)
$$\begin{aligned}&\quad +{{\text {Op}}}^{\text {BW}}\!\left( K'({{\mathtt {m}}}+\rho )\right) R^{-1}(\nabla \rho , {\text {i}}\xi ) \rho + R^{-1}(K'({{\mathtt {m}}}+\rho ), {\text {i}}\nabla \rho \cdot \xi )\rho . \end{aligned}$$
(3.21)

Collecting all the above expansions and recalling the definition of the symplectic matrix J in (3.3), the system (1.6) can be written in the paralinearized form

$$\begin{aligned} \begin{aligned} \partial _t {\begin{pmatrix}\rho \\ \phi \end{pmatrix}}&= J {{\text {Op}}}^{\text {BW}}\!\left( \begin{bmatrix} K({{\mathtt {m}}}+\rho )|\xi |^2 &{} \nabla \phi \cdot \,{\text {i}}\xi \\ -\nabla \phi \cdot \,{\text {i}}\xi &{} ({{\mathtt {m}}}+\rho )|\xi |^2 \end{bmatrix}\right) {\begin{pmatrix}\rho \\ \phi \end{pmatrix}} + R(\rho ,\phi ) \end{aligned} \end{aligned}$$
(3.22)

where we collected in \(R(\rho , \phi )\) all the terms in lines (3.14)–(3.21).

Step 2: complex coordinates. We now write system (3.22) in the complex coordinates \(U = {\textit{C}}^{-1}{\begin{pmatrix}\rho \\ \phi \end{pmatrix}}\). Note that \( {\textit{C}}^{-1}\) conjugates the Poisson tensor J to \({\mathbb {J}}\) defined in (3.3), i.e. \( {\textit{C}}^{-1} \, J = {\mathbb {J}}\, {\textit{C}}^* \) and therefore system (3.22) is conjugated to

$$\begin{aligned} \partial _t U&= {\mathbb {J}}{\textit{C}}^{*} {{\text {Op}}}^{\text {BW}}\!\left( \begin{bmatrix} K({{\mathtt {m}}}+\rho )|\xi |^2 &{} \nabla \phi \cdot \,{\text {i}}\xi \\ - \nabla \phi \cdot \,{\text {i}}\xi &{} \rho \, |\xi |^2 \end{bmatrix}\right) {\textit{C}}U + {\varvec{C}}^{-1} R({\textit{C}}U). \end{aligned}$$
(3.23)

Using (3.2), system (3.23) reads as system (3.4)–(3.8) with \( R(U) := {\textit{C}}^{-1} R({\textit{C}}U) \).

We note also that estimates (3.9) and (3.10) for \(j = 2\) follow by (2.18) and (2.31), whereas in case \(j = 1\) follow by (2.19) applied with \(m=1\), \(\varrho =1\).

Step 3: Estimate of the remainder R(U) . We now prove (3.11)–(3.13). Since \( \Vert \rho \Vert _\sigma , \Vert \phi \Vert _\sigma \sim \Vert U \Vert _\sigma \) for any \( \sigma \in {\mathbb {R}}\) by (3.2), the estimates (3.11)–(3.13) directly follow from those of \( R(\rho , \phi ) \) in (3.22). We now estimate each term in (3.14)–(3.21). In the sequel \(\sigma \ge s_0 > d / 2 \).

Estimate of the term in line (3.14). Applying first (2.18) with \( m = 0 \), and then (2.19) with \(\varrho =2\), we have

$$\begin{aligned} \Vert {{\text {Op}}}^{\text {BW}}\!\left( \Delta \phi \right) \rho \Vert _{ \sigma }\lesssim \Vert \phi \Vert _{ {s_0+2}} \Vert \rho \Vert _{ \sigma }, \quad \Vert {{\text {Op}}}^{\text {BW}}\!\left( \Delta \phi \right) \rho \Vert _{ \sigma }\lesssim \Vert \phi \Vert _{ {s_0}} \Vert \rho \Vert _{ {\sigma +2}}. \end{aligned}$$
(3.24)

By (2.27), the smoothing remainder in line (3.14) satisfies the estimates

$$\begin{aligned} \Vert R^p(\rho ,\Delta \phi )\Vert _{ \sigma } \lesssim \Vert \phi \Vert _{ {s_0+2}}\Vert \rho \Vert _{ \sigma }, \quad \Vert R^p(\rho ,\Delta \phi )\Vert _{ \sigma } \lesssim \Vert \phi \Vert _{ {s_0}}\Vert \rho \Vert _{ {\sigma +2}}, \end{aligned}$$
(3.25)

and, by (2.22) with \( \varrho = 2 \), and the interpolation estimate (2.4),

$$\begin{aligned} \Vert R^{-2}(\rho , |\xi |^2)\phi \Vert _{\sigma }\lesssim \Vert \rho \Vert _{{s_0+2}} \Vert \phi \Vert _{\sigma } \lesssim \Vert \phi \Vert _{{s_0}} \Vert \rho \Vert _{{\sigma +2}}+\Vert \rho \Vert _{{s_0}} \Vert \phi \Vert _{{\sigma +2}}. \end{aligned}$$
(3.26)

By (3.24)–(3.26) and \( \Vert \rho \Vert _\sigma , \Vert \phi \Vert _\sigma \sim \Vert U \Vert _\sigma \) we deduce that the terms in line (3.14), written in function of U, satisfy (3.11). Next we write

$$\begin{aligned} {{\text {Op}}}^{\text {BW}}\!\left( \Delta \phi _1\right) \rho _1- {{\text {Op}}}^{\text {BW}}\!\left( \Delta \phi _2\right) \rho _2 = {{\text {Op}}}^{\text {BW}}\!\left( \Delta \phi _1\right) [\rho _1-\rho _2] + {{\text {Op}}}^{\text {BW}}\!\left( \Delta \phi _1-\Delta \phi _2\right) \rho _2 \end{aligned}$$

and, applying (2.18) with \( m = 0\), and (2.19) with \(\varrho = 2\) to \({{\text {Op}}}^{\text {BW}}\!\left( \Delta \phi _1-\Delta \phi _2\right) \rho _2\), we get

$$\begin{aligned} \Vert {{\text {Op}}}^{\text {BW}}\!\left( \Delta \phi _1\right) \rho _1- {{\text {Op}}}^{\text {BW}}\!\left( \Delta \phi _2\right) \rho _2\Vert _{ \sigma }&\lesssim \Vert \phi _1\Vert _{ {s_0+2}}\Vert \rho _1-\rho _2\Vert _{ \sigma }+ \Vert \phi _1-\phi _2\Vert _{ {s_0+2}}\Vert \rho _2\Vert _{ \sigma } \nonumber \\ \Vert {{\text {Op}}}^{\text {BW}}\!\left( \Delta \phi _1\right) \rho _1- {{\text {Op}}}^{\text {BW}}\!\left( \Delta \phi _2\right) \rho _2\Vert _{ \sigma }&\lesssim \Vert \phi _1\Vert _{ {s_0+2}}\Vert \rho _1-\rho _2\Vert _{ \sigma }+ \Vert \phi _1-\phi _2\Vert _{ {s_0}}\Vert \rho _2\Vert _{ {\sigma +2}}.\nonumber \\ \end{aligned}$$
(3.27)

Concerning the remainder \(R^p(\rho , \Delta \phi )\), we write \(R^p(\rho _1,\Delta \phi _1)-R^p(\rho _2,\Delta \phi _2) = R^p(\rho _1-\rho _2,\Delta \phi _1) +R^p(\rho _2,\Delta \phi _2- \Delta \phi _1)\) and, applying (2.27), we get

$$\begin{aligned} \Vert R^p(\rho _1,\Delta \phi _1)-R^p(\rho _2,\Delta \phi _2)\Vert _{ \sigma }&\lesssim \Vert \phi _1\Vert _{ {s_0+2}} \Vert \rho _1-\rho _2\Vert _{ \sigma } + \Vert \rho _2\Vert _{ \sigma } \Vert \phi _1-\phi _2\Vert _{ {s_0+2}} \nonumber \\ \Vert R^p(\rho _1,\Delta \phi _1)-R^p(\rho _2,\Delta \phi _2)\Vert _{ \sigma }&\lesssim \Vert \phi _1\Vert _{ {s_0+2}} \Vert \rho _1-\rho _2\Vert _{ \sigma } + \Vert \rho _2\Vert _{ {\sigma +2}} \Vert \phi _1-\phi _2\Vert _{ {s_0}}.\nonumber \\ \end{aligned}$$
(3.28)

Finally we write \( R^{-2}( \rho _1, |\xi |^2)\phi _1- R^{-2}( \rho _2, |\xi |^2)\phi _2 = R^{-2}( \rho _1-\rho _2, |\xi |^2)\phi _1 + R^{-2}( \rho _2, |\xi |^2)[\phi _1- \phi _2] \). Using (2.22) we get

$$\begin{aligned}&\Vert R^{-2}( \rho _1, |\xi |^2)\phi _1- R^{-2}( \rho _2, |\xi |^2)\phi _2\Vert _{\sigma } \lesssim \Vert \phi _1\Vert _{\sigma }\Vert \rho _1-\rho _2\Vert _{s_0+2}\nonumber \\&\quad + \Vert \phi _1-\phi _2\Vert _{\sigma }\Vert \rho _2\Vert _{s_0+2}. \end{aligned}$$
(3.29)

We also claim that

$$\begin{aligned}&\Vert R^{-2}( \rho _1, |\xi |^2)\phi _1- R^{-2}( \rho _2, |\xi |^2)\phi _2\Vert _{\sigma }\nonumber \\&\quad \lesssim \Vert \rho _1-\rho _2\Vert _{s_0} \Vert \phi _1\Vert _{\sigma +2}+ \Vert \phi _1-\phi _2\Vert _{\sigma }\Vert \rho _2\Vert _{s_0+2}. \end{aligned}$$
(3.30)

Indeed, we bound

$$\begin{aligned}&\Vert R^{-2}( \rho _1, |\xi |^2)\phi _1- R^{-2}( \rho _2, |\xi |^2)\phi _2\Vert _{\sigma }\\&\quad \lesssim \Vert R^{-2}( \rho _1-\rho _2, |\xi |^2)\phi _1\Vert _{\sigma }+ \Vert \phi _1-\phi _2\Vert _{\sigma }\Vert \rho _2\Vert _{s_0+2} \, \end{aligned}$$

and, to control \(R^{-2}( \rho _1-\rho _2, |\xi |^2)\phi _1\), we use that, by definition, it equals

$$\begin{aligned} {{\text {Op}}}^{\text {BW}}\!\left( \rho _1-\rho _2\right) {{\text {Op}}}^{\text {BW}}\!\left( |\xi |^2\right) \phi _1- {{\text {Op}}}^{\text {BW}}\!\left( (\rho _1-\rho _2)|\xi |^2\right) \phi _1 -{{\text {Op}}}^{\text {BW}}\!\left( \nabla (\rho _1-\rho _2)\cdot {\text {i}}\xi \right) \phi _1 \end{aligned}$$

and we estimate the first two terms using (2.19) with \(\varrho = 0\) and the last term with \(\varrho = 1\), by \(\Vert R^{-2}( \rho _1-\rho _2, |\xi |^2)\phi _1\Vert _{\sigma }\lesssim \Vert \rho _1-\rho _2\Vert _{s_0} \Vert \phi _1\Vert _{\sigma +2} \), proving (3.30). By (3.27)–(3.30) and \( \Vert \rho \Vert _\sigma , \Vert \phi \Vert _\sigma \sim \Vert U \Vert _\sigma \) we deduce that the terms in line (3.14), written in function of U, satisfy (3.12)–(3.13).

The estimates (3.11)–(3.13) for the terms in lines (3.15), (3.16), (3.18) and (3.17), follow by similar arguments, using also (2.31).

Estimates of \( {\mathtt R}(\rho ) \) defined in (3.19)–(3.21).

Writing \( {{\text {Op}}}^{\text {BW}}\!\left( |\nabla \rho |^2\right) K'({{\mathtt {m}}}+\rho ) = {{\text {Op}}}^{\text {BW}}\!\left( |\nabla \rho |^2\right) (K'({{\mathtt {m}}}+\rho ) - K'({{\mathtt {m}}}))\) (in the homogeneous spaces \( \dot{H}^s \)), we have, by (2.18), the fact that \( \rho \in {\mathcal {Q}}\), Theorem 2.8, (2.27), (2.2), (2.22) with \(\varrho =1\),

$$\begin{aligned} \Vert {\mathtt R} (\rho )\Vert _{ \sigma } \le {\mathtt {C}}\big (\Vert \rho \Vert _{ {s_0+2}}\big ) \Vert \rho \Vert _{ \sigma }. \end{aligned}$$

Thus \( {\mathtt R} (\rho ) \), written as a function of U, satisfies (3.11). The estimates (3.12)–(3.13) follow by

$$\begin{aligned}&\Vert {\mathtt R}(\rho _1)-{\mathtt R}(\rho _2)\Vert _{ \sigma } \le {\mathtt {C}}\big (\Vert \rho _1\Vert _{ {s_0+2}},\Vert \rho _2\Vert _{ {s_0+2}}\big )\Vert \rho _1-\rho _2\Vert _{ \sigma }+ {\mathtt {C}}\big (\Vert \rho _1\Vert _{ {\sigma }},\Vert \rho _2\Vert _{ {\sigma }}\big )\Vert \rho _1-\rho _2\Vert _{ {s_0+2}} \end{aligned}$$
(3.31)
$$\begin{aligned}&\Vert {\mathtt R}(\rho _1)-{\mathtt R}(\rho _2)\Vert _{ {s_0}} \le {\mathtt {C}}\big (\Vert \rho _1\Vert _{ {s_0+2}},\Vert \rho _2\Vert _{ {s_0+2}}\big )\Vert \rho _1-\rho _2\Vert _{ {s_0}}. \end{aligned}$$
(3.32)

Proof of (3.31). Defining \( w := \nabla ( \rho _1+\rho _2)\), \(v:= \nabla (\rho _1-\rho _2)\), then we have, by (2.1),

$$\begin{aligned}&\Vert |\nabla \rho _1|^2- |\nabla \rho _2|^2\Vert _{ {s_0}} = \Vert w \cdot v\Vert _{ {s_0}} \lesssim \big ( \Vert \rho _1\Vert _{ {s_0+1}} +\Vert \rho _2\Vert _{ {s_0+1}}\big ) \Vert \rho _1-\rho _2\Vert _{ {s_0+1}}\end{aligned}$$
(3.33)
$$\begin{aligned}&\Vert |\nabla \rho _1|^2- |\nabla \rho _2|^2\Vert _{ {s_0-1}} = \Vert w \cdot v\Vert _{ {s_0-1}} {\mathop {\lesssim }\limits ^{(2.2)}}\big ( \Vert \rho _1\Vert _{ {s_0+1}} +\Vert \rho _2\Vert _{ {s_0+1}}\big ) \Vert \rho _1-\rho _2\Vert _{ {s_0}}. \end{aligned}$$
(3.34)

Let us prove (3.31) for the first term in (3.19). Remind that \( \rho _1, \rho _2 \) are in \( {\mathcal {Q}}\). We have

$$\begin{aligned} \Vert&{{\text {Op}}}^{\text {BW}}\!\left( |\nabla \rho _1|^2\right) K'({{\mathtt {m}}}+\rho _1)-{{\text {Op}}}^{\text {BW}}\!\left( |\nabla \rho _2|^2\right) K'({{\mathtt {m}}}+\rho _2)\Vert _{ \sigma } \nonumber \\&\quad \le \Vert {{\text {Op}}}^{\text {BW}}\!\left( w \cdot v\right) \big ( K'({{\mathtt {m}}}+ \rho _1)\Vert _{ \sigma } + \Vert {{\text {Op}}}^{\text {BW}}\!\left( |\nabla \rho _2|^2\right) \big [K'({{\mathtt {m}}}+\rho _1)-K'({{\mathtt {m}}}+\rho _2)\big ]\Vert _{ \sigma } \nonumber \\&\quad {\mathop {\lesssim }\limits ^{(2.18)}}\Vert w\cdot v\Vert _{ {s_0}} \Vert K'({{\mathtt {m}}}+ \rho _1)-K'({{\mathtt {m}}})\Vert _{ \sigma }+ \Vert \rho _2\Vert _{ {s_0+1}}^2 \Vert K'({{\mathtt {m}}}+\rho _1)-K'({{\mathtt {m}}}+\rho _2)\Vert _{ \sigma } \nonumber \\&\quad {\mathop {\lesssim }\limits ^{(2.4),(2.31),(3.33)}}\Vert \rho _1\Vert _\sigma (\Vert \rho _1\Vert _{s_0+1}+ \Vert \rho _2\Vert _{s_0+1}) \Vert \rho _1-\rho _2\Vert _{s_0+1} \nonumber \\&\quad \quad + {\mathtt {C}}\big ( \Vert \rho _1\Vert _{s_0+1}, \Vert \rho _2\Vert _{ {s_0+1}} \big ) \Vert \rho _1-\rho _2\Vert _{ {\sigma }} \nonumber \\&\quad \quad + \Vert \rho _2\Vert _{ {s_0}}\Vert \rho _2\Vert _{s_0+2}( \Vert \rho _1\Vert _\sigma + \Vert \rho _2\Vert _\sigma ) \Vert \rho _1-\rho _2\Vert _{s_0} \nonumber \\&\quad {\mathop {\le }\limits ^{(2.4)}}{\mathtt {C}}\big ( \Vert \rho _1\Vert _\sigma , \Vert \rho _2\Vert _{ {\sigma }} \big ) \Vert \rho _1-\rho _2\Vert _{ {s_0+2}} + {\mathtt {C}}\big ( \Vert \rho _1\Vert _{s_0+2}, \Vert \rho _2\Vert _{ {s_0+2}} \big ) \Vert \rho _1-\rho _2\Vert _{ {\sigma }}. \end{aligned}$$
(3.35)

In the same way the second term in (3.19) is bounded by (3.35). Regarding the term in (3.20), using that \(R^p(\cdot ,\cdot )\) is bilinear and symmetric, we have

$$\begin{aligned} \Vert&{{\text {Op}}}^{\text {BW}}\!\left( K'({{\mathtt {m}}}+\rho _1)\right) R^p(\nabla \rho _1, \nabla \rho _1)-{{\text {Op}}}^{\text {BW}}\!\left( K'({{\mathtt {m}}}+\rho _2)\right) R^p(\nabla \rho _2, \nabla \rho _2)\Vert _{ \sigma }\nonumber \\&\quad \le \Vert {{\text {Op}}}^{\text {BW}}\!\left( K'({{\mathtt {m}}}+\rho _1)-K'({{\mathtt {m}}}+\rho _2)\right) R^p(\nabla \rho _1, \nabla \rho _1)\Vert _{ \sigma } \nonumber \\&\quad \quad + \Vert {{\text {Op}}}^{\text {BW}}\!\left( K'({{\mathtt {m}}}+\rho _2)\right) R^p(w, v)\Vert _{ \sigma } \nonumber \\&\quad {\mathop {\lesssim }\limits ^{(2.18),(2.27)}}\Vert \rho _1\Vert _{ {\sigma }}\Vert \rho _1\Vert _{s_0+2} \Vert K'({{\mathtt {m}}}+ \rho _1)-K'({{\mathtt {m}}}+\rho _2)\Vert _{ {s_0}}\nonumber \\&\quad \quad + \Vert w\Vert _{ {s_0+1}}\Vert v\Vert _{\sigma -1} \Vert K'({{\mathtt {m}}}+\rho _2)\Vert _{ {s_0}} \nonumber \\&\quad {\mathop {\le }\limits ^{(2.31),(3.33)}}{\mathtt {C}}\big ( \Vert \rho _1\Vert _{\sigma }, \Vert \rho _2\Vert _{\sigma } \big ) \Vert \rho _1-\rho _2\Vert _{s_0}+ {\mathtt {C}}\big ( \Vert \rho _1\Vert _{s_0+2}, \Vert \rho _2\Vert _{s_0+2} \big ) \Vert \rho _1-\rho _2\Vert _{\sigma } . \end{aligned}$$
(3.36)

Also the terms in (3.21) are bounded by (3.35), proving that \({\mathtt R}(\rho )\) satisfies (3.31).

Proof of (3.32). Regarding the first term (3.19), we have

$$\begin{aligned}&\Vert {{\text {Op}}}^{\text {BW}}\!\left( |\nabla \rho _1|^2\right) K'({{\mathtt {m}}}+\rho _1)-{{\text {Op}}}^{\text {BW}}\!\left( |\nabla \rho _2|^2\right) K'({{\mathtt {m}}}+\rho _2)\Vert _{ s_0} \nonumber \\&\le \Vert {{\text {Op}}}^{\text {BW}}\!\left( w \cdot v\right) K'({{\mathtt {m}}}+ \rho _1)\Vert _{ s_0} + \Vert {{\text {Op}}}^{\text {BW}}\!\left( |\nabla \rho _2|^2\right) \big [K'({{\mathtt {m}}}+\rho _1)-K'({{\mathtt {m}}}+\rho _2)\big ]\Vert _{ s_0} \nonumber \\&{\mathop {\lesssim }\limits ^{(2.18),(2.19)}}\Vert w \cdot v \Vert _{ s_0-1}\Vert K'({{\mathtt {m}}}+ \rho _1)\Vert _{s_0+1}+ \Vert \rho _2\Vert _{ {s_0+1}}^2 \Vert K'({{\mathtt {m}}}+\rho _1)-K'({{\mathtt {m}}}+\rho _2)\Vert _{ s_0} \nonumber \\&{\mathop {\le }\limits ^{(2.31),(3.34)}}{\mathtt {C}}\big ( \Vert \rho _1\Vert _{s_0+1}, \Vert \rho _2\Vert _{ {s_0+1}} \big ) \Vert \rho _1-\rho _2\Vert _{ {s_0}}. \end{aligned}$$
(3.37)

Similarly we deduce that the second term in (3.19) is bounded as in (3.37). Regarding the term in (3.20), note that the bound (3.32) follows from (3.36) applied for \(\sigma = s_0\). The estimate for last two terms in (3.21) follows in the same way so we analyze the last one. First we have

$$\begin{aligned}&\ \Vert R^{-1} (K'({{\mathtt {m}}}+\rho _1), {\text {i}}\nabla \rho _1\cdot \xi )\rho _1- R^{-1}(K'({{\mathtt {m}}}+\rho _2), {\text {i}}\nabla \rho _2\cdot \xi )\rho _2\Vert _{ s_0}\\&\quad \le \Vert \big [ R^{-1}(K'({{\mathtt {m}}}+\rho _1),\nabla \rho _1\cdot {\text {i}}\xi )- R^{-1}(K'({{\mathtt {m}}}+\rho _2), {\text {i}}\nabla \rho _2\cdot \xi )\big ] \rho _1\Vert _{ s_0} \\&\quad \quad +\Vert R^{-1}(K'({{\mathtt {m}}}+\rho _2),\nabla \rho _2\cdot {\text {i}}\xi ) (\rho _1-\rho _2)\Vert _{ s_0}\\&\quad {\mathop {\le }\limits ^{(2.22),(2.31)}}\Vert \big [ R^{-1}(K'({{\mathtt {m}}}+\rho _1),\nabla \rho _1\cdot {\text {i}}\xi )- R^{-1}(K'({{\mathtt {m}}}+\rho _2), {\text {i}}\nabla \rho _2\cdot \xi )\big ] \rho _1\Vert _{ s_0}\\&\quad \quad + {\mathtt {C}}\big (\Vert \rho _2\Vert _{s_0+2}\big ) \Vert \rho _1-\rho _2\Vert _{s_0}. \end{aligned}$$

On the other hand, by definition, we have

$$\begin{aligned}&\big [ R^{-1}(K'({{\mathtt {m}}}+\rho _1),\nabla \rho _1\cdot {\text {i}}\xi )- R^{-1}(K'({{\mathtt {m}}}+\rho _2), {\text {i}}\nabla \rho _2\cdot \xi )\big ] \rho _1 \nonumber \\&\quad = \big [{{\text {Op}}}^{\text {BW}}\!\left( K'({{\mathtt {m}}}+\rho _1)\right) {{\text {Op}}}^{\text {BW}}\!\left( \nabla \rho _1\cdot {\text {i}}\xi \right) -{{\text {Op}}}^{\text {BW}}\!\left( K'({{\mathtt {m}}}+\rho _2)\right) {{\text {Op}}}^{\text {BW}}\!\left( \nabla \rho _2\cdot {\text {i}}\xi \right) \big ]\rho _1 \nonumber \\&\quad \quad + {{\text {Op}}}^{\text {BW}}\!\left( K'({{\mathtt {m}}}+\rho _1)\nabla \rho _1\cdot {\text {i}}\xi -K'({{\mathtt {m}}}+\rho _2)\nabla \rho _2\cdot {\text {i}}\xi \right) \rho _1 \nonumber \\&\quad = {{\text {Op}}}^{\text {BW}}\!\left( K'({{\mathtt {m}}}+\rho _1)-K'({{\mathtt {m}}}+\rho _2)\right) {{\text {Op}}}^{\text {BW}}\!\left( \nabla \rho _1\cdot {\text {i}}\xi \right) \rho _1 \nonumber \\&\quad \quad +{{\text {Op}}}^{\text {BW}}\!\left( K'({{\mathtt {m}}}+\rho _2)\right) {{\text {Op}}}^{\text {BW}}\!\left( \nabla ( \rho _1-\rho _2)\cdot {\text {i}}\xi \right) \rho _1 \nonumber \\&\quad \quad +{{\text {Op}}}^{\text {BW}}\!\left( \nabla ( K({{\mathtt {m}}}+\rho _1)-K({{\mathtt {m}}}+\rho _2))\cdot {\text {i}}\xi \right) \rho _1. \end{aligned}$$
(3.38)

Then, applying first (2.18) to the first term and then (2.19) with \(\varrho =1\), \(m=1\) and (2.31) to each term, we deduce that the \( \Vert \ \Vert _{s_0} \)-norm of (3.38) is bounded by \( {\mathtt {C}}\big (\Vert \rho _1\Vert _{s_0+2}, \Vert \rho _2\Vert _{s_0+2}\big ) \Vert \rho _1-\rho _2\Vert _{s_0}\). Thus (3.32) is proved. \(\square \)

4 Local Existence

In this section we prove the existence of a local in time solution of system (3.4). For any \(s \in {\mathbb {R}}\) and \(T >0\), we denote \(L^\infty _T {\dot{{\mathbf {H}}}}^s := L^\infty ([0,T], {\dot{{\mathbf {H}}}}^s) \). For \( \delta > 0 \) we also introduce

$$\begin{aligned} {\mathcal {Q}}_\delta := \big \{ \rho \in H^{s_0}_0 \ :\ {{\mathtt {m}}}_1 + \delta \le {{\mathtt {m}}}+ \rho (x) \le {{\mathtt {m}}}_2 - \delta \big \} \subset {\mathcal {Q}}\end{aligned}$$
(4.1)

where \( {\mathcal {Q}}\) is defined in (1.12).

Proposition 4.1

(Local well-posedness in \({\mathbb {T}}^d\)) For any \(s > \frac{d}{2}+2\), any initial datum \(U_0 \in {\dot{{\mathbf {H}}}}^s \) with \( \rho (U_0) \in {\mathcal {Q}}_\delta \) for some \( \delta > 0 \), there exist \(T := T(\Vert U \Vert _{s_0+2}, \delta ) > 0 \) and a unique solution \(U \in C^0\big ([0, T], {\dot{{\mathbf {H}}}}^s \big ) \cap C^1\big ([0, T], {\dot{{\mathbf {H}}}}^{s-2} \big )\) of (3.4) satisfying \( \rho (U) \in {\mathcal {Q}}\), for any \( t \in [0,T] \). Moreover the solution depends continuously with respect to the initial datum in \( {\dot{{\mathbf {H}}}}^s \).

Proposition 4.1 proves Theorem 1.2 and thus Theorem 1.1.

The first step is to prove the local well-posedness result of a linear inhomogeneous problem.

Proposition 4.2

(Linear local well-posedness) Let \(\Theta \ge r > 0 \) and U be a function in \( C^0([0,T],{\dot{{\mathbf {H}}}}^{s_0+2}) \cap C^1([0,T],{\dot{{\mathbf {H}}}}^{s_0}) \) satisfying

$$\begin{aligned} \left\| U \right\| _{L^\infty _T {\dot{{\mathbf {H}}}}^{s_0+2} } + \left\| \partial _t U \right\| _{L^\infty _T {\dot{{\mathbf {H}}}}^{s_0} } \le \Theta , \quad \Vert U\Vert _{L^\infty _T {\dot{{\mathbf {H}}}}^{s_0}}\le r, \quad \rho (U(t)) \in {\mathcal {Q}}, \ \forall t \in [0, T].\nonumber \\ \end{aligned}$$
(4.2)

Let \( \sigma \ge 0\) and \( t \mapsto R(t)\) be a function in \( C^0([0, T], {\dot{{\mathbf {H}}}}^\sigma ) \). Then there exists a unique solution \(V\in C^0([0,T], {\dot{{\mathbf {H}}}}^\sigma )\cap C^1([0,T], {\dot{{\mathbf {H}}}}^{\sigma -2})\) of the linear inhomogeneous system

$$\begin{aligned} \partial _t V = {\mathbb {J}}\, {{\text {Op}}}^{\text {BW}}\!\left( A_2(U(t); x, \xi ) + A_1(U(t); x, \xi ) \right) V + R(t), \quad V(0,x) = V_0(x) \in {\dot{{\mathbf {H}}}}^{\sigma },\nonumber \\ \end{aligned}$$
(4.3)

satisfying, for some \( C_\Theta := C_{\Theta ,\sigma } > 0 \) and \(C_r := C_{r,\sigma } > 0 \), the estimate

$$\begin{aligned} \Vert V\Vert _{L^\infty _T {\dot{{\mathbf {H}}}}^\sigma } \le C_r e^{C_\Theta T} \Vert V_0\Vert _{\sigma }+ C_\Theta e^{C_\Theta T } T \Vert R\Vert _{L^\infty _T {\dot{{\mathbf {H}}}}^\sigma }. \end{aligned}$$
(4.4)

The following two sections are devoted to the proof of Proposition 4.2. The key step is the construction of a modified energy which is controlled by the \({\dot{{\mathbf {H}}}}^{\sigma }\)-norm, and whose time variation is bounded by the \({\dot{{\mathbf {H}}}}^{\sigma }\) norm of the solution, as done e.g. in [1, 21] for linear systems. In order to construct such modified energy, the first step is to diagonalize the matrix \( {\mathbb {J}}A_2\) in (4.3).

4.1 Diagonalization at Highest Order

We diagonalize the matrix of symbols \({\mathbb {J}}A_2(U; x, \xi )\). The eigenvalues of the matrix

$$\begin{aligned} {\mathbb {J}}\begin{bmatrix} 1 + {\mathtt {a}}_+(U; x) &{} {\mathtt {a}}_-(U; x) \\ {\mathtt {a}}_-(U; x) &{} 1+ {\mathtt {a}}_+(U; x) \end{bmatrix} \end{aligned}$$
(4.5)

with \({\mathtt {a}}_\pm (U;x)\) defined in (3.7) are given by \(\pm {\text {i}}\lambda (U; x)\) with

$$\begin{aligned} \lambda (U; x) := \sqrt{(1+{\mathtt {a}}_+(U;x))^2 - {\mathtt {a}}_-(U;x)^2} = \sqrt{\frac{({{\mathtt {m}}}+\rho (U)) \, K({{\mathtt {m}}}+\rho (U))}{{{\mathtt {m}}}\, K({{\mathtt {m}}})}}. \end{aligned}$$
(4.6)

These eigenvalues are purely imaginary because \(\rho (U) \in {\mathcal {Q}}\) [see (1.12)] and (1.9), which guarantees that \(\lambda (U;x) \) is real valued and fulfills

$$\begin{aligned} 0< \lambda _{\min }:= \sqrt{\frac{{{\mathtt {m}}}_1 c_K}{{{\mathtt {m}}}K({{\mathtt {m}}})}}\le \lambda (U;x) \le \sqrt{\frac{{{\mathtt {m}}}_2 C_K}{{{\mathtt {m}}}K({{\mathtt {m}}})}}=: \lambda _{\max }. \end{aligned}$$
(4.7)

A matrix which diagonalizes (4.5) is

$$\begin{aligned} \begin{aligned} F&: = \begin{pmatrix} f (U; x )&{}\quad g(U; x ) \\ g(U; x ) &{} \quad f(U; x ) \end{pmatrix} , \\ f&:= \displaystyle { \frac{1+{\mathtt {a}}_+ + \lambda }{\sqrt{(1+{\mathtt {a}}_+ + \lambda )^2 - {\mathtt {a}}_-^2}} }, \qquad g:= \displaystyle { \frac{ -{\mathtt {a}}_-}{\sqrt{(1+{\mathtt {a}}_+ + \lambda )^2 - {\mathtt {a}}_-^2}} }. \end{aligned} \end{aligned}$$
(4.8)

Note that F(Ux) is well defined because

$$\begin{aligned} (1+{\mathtt {a}}_+ + \lambda )^2 - {\mathtt {a}}_-^2&= \Big (\frac{K({{\mathtt {m}}}+\rho (U))}{K({{\mathtt {m}}})}+ \lambda \Big ) \Big (\frac{{{\mathtt {m}}}+\rho (U)}{{{\mathtt {m}}}}+ \lambda \Big ) \nonumber \\&> \frac{({{\mathtt {m}}}+\rho (U) )K({{\mathtt {m}}}+\rho (U))}{{{\mathtt {m}}}K({{\mathtt {m}}})} \ge \frac{{{\mathtt {m}}}_1 c_K}{{{\mathtt {m}}}K({{\mathtt {m}}})} \end{aligned}$$
(4.9)

by (1.12) and (1.9). The matrix F(Ux) has \( \det F(U;x) = {f^2 - g^2} = 1 \) and its inverse is

$$\begin{aligned} F(U; x )^{-1} : = \begin{pmatrix} f (U; x )&{}\quad - g(U; x ) \\ -g(U; x ) &{}\quad f(U; x ) \end{pmatrix}. \end{aligned}$$
(4.10)

We have that

$$\begin{aligned} F(U;x)^{-1} \, {\mathbb {J}}\begin{bmatrix} 1 + {\mathtt {a}}_+(U; x) &{} {\mathtt {a}}_-(U; x) \\ {\mathtt {a}}_-(U; x) &{} 1+ {\mathtt {a}}_+(U; x) \end{bmatrix}\, F(U;x) = {\mathbb {J}}\lambda (U; x). \end{aligned}$$
(4.11)

By (2.31) and (4.9) we deduce the following estimates: for any \( N \in {\mathbb {N}}_0 \), \( s \ge 0 \) and \({\sigma }>\frac{d}{2}\),

$$\begin{aligned} \begin{aligned}&\left\| {\mathtt {a}}_\pm (U) \right\| _{\sigma }, \, \left\| f(U) \right\| _{{\sigma }}, \, \left\| g(U) \right\| _{{\sigma }}\le {\mathtt {C}}\big ( \Vert U\Vert _{{\sigma }}\big ), \\&|\lambda (U;x)|\xi |^{2s}|_{2s,{\sigma },N}\le {\mathtt {C}}_N \big (\Vert U\Vert _{\sigma }\big ), \quad | {\mathtt {b}}(U)\cdot \xi |_{1,{\sigma },N}\le {\mathtt {C}}_N (\Vert U\Vert _{{\sigma }+1}\big ). \end{aligned} \end{aligned}$$
(4.12)

For any \(\varepsilon > 0 \), consider the regularized matrix symbol

$$\begin{aligned} A^\varepsilon (U;x,\xi ) := \left( A_2(U;x,\xi )+A_1(U;x,\xi ) \right) \chi ( \varepsilon \lambda (U;x) |\xi |^2 ), \end{aligned}$$
(4.13)

where \( \chi \) is the cut-off function in (2.13) and \(\lambda (U;x)\) is the function defined in (4.6). In what follows we will denote by \(\chi _\varepsilon := \chi ( \varepsilon \lambda (U; x) |\xi |^2)\). Note that, by (2.31), (4.7) and by the fact that the function \(y \mapsto \langle \xi \rangle ^{|\alpha |}\partial _\xi ^\alpha \chi ( \varepsilon y |\xi |^2)\) is bounded together with its derivatives uniformly in \(\varepsilon \in (0,1)\), \( \xi \in {\mathbb {R}}^d\) and \(y\in [\lambda _{\min }, \lambda _{\max }]\), the symbol \( \chi _\varepsilon \) satisfies, for any \( N\in {\mathbb {N}}_0 \), \( \sigma > d/2 \)

$$\begin{aligned} | \chi _\varepsilon |_{0, \sigma , N} \le {\mathtt {C}}\left( \Vert U\Vert _{\sigma }\right) , \quad {\text {uniformly}} \ \text {in} \ \varepsilon . \end{aligned}$$
(4.14)

The diagonalization (4.11) has the following operatorial consequence.

Lemma 4.3

We have

$$\begin{aligned} {{\text {Op}}}^{\text {BW}}\!\left( F^{-1}\right) \, {\mathbb {J}}{{\text {Op}}}^{\text {BW}}\!\left( A^\varepsilon \right) \, {{\text {Op}}}^{\text {BW}}\!\left( F\right) ={\mathbb {J}}{{\text {Op}}}^{\text {BW}}\!\left( (\sqrt{{{\mathtt {m}}}K({{\mathtt {m}}})}\lambda |\xi |^2+ {\mathtt {b}}\cdot \xi )\chi _\varepsilon \right) + {\mathcal {F}}(U)\nonumber \\ \end{aligned}$$
(4.15)

where \({\mathcal {F}}(U):= {\mathcal {F}}_\varepsilon (U):{\dot{{\mathbf {H}}}}^\sigma \rightarrow {\dot{{\mathbf {H}}}}^\sigma \), \(\forall \sigma \ge 0\), satisfies, uniformly in \( \varepsilon \),

$$\begin{aligned} \Vert {\mathcal {F}}(U) W\Vert _{\sigma }\le {{\mathtt {C}}( \Vert U \Vert _{s_0+2})} \Vert W\Vert _{\sigma }, \quad \forall W \in {\dot{{\mathbf {H}}}}^\sigma . \end{aligned}$$
(4.16)

Proof

We have that

$$\begin{aligned} {{\text {Op}}}^{\text {BW}}\!\left( F^{-1}\right) {\mathbb {J}}{{\text {Op}}}^{\text {BW}}\!\left( A_2\chi _\varepsilon \right) {{\text {Op}}}^{\text {BW}}\!\left( F\right) = {\mathbb {J}}\sqrt{{{\mathtt {m}}}K({{\mathtt {m}}})} \begin{bmatrix} D_2 &{} B_2 \\ B_2 &{} D_2 \end{bmatrix}, \end{aligned}$$

where

$$\begin{aligned} D_2&= {{\text {Op}}}^{\text {BW}}\!\left( f\right) {{\text {Op}}}^{\text {BW}}\!\left( |\xi |^2(1+{\mathtt {a}}_+) \chi _\varepsilon \right) {{\text {Op}}}^{\text {BW}}\!\left( f\right) \\&\quad + {{\text {Op}}}^{\text {BW}}\!\left( g\right) {{\text {Op}}}^{\text {BW}}\!\left( |\xi |^2(1+{\mathtt {a}}_+)\chi _\varepsilon \right) {{\text {Op}}}^{\text {BW}}\!\left( g\right) \\&\quad + {{\text {Op}}}^{\text {BW}}\!\left( f\right) {{\text {Op}}}^{\text {BW}}\!\left( |\xi |^2 {\mathtt {a}}_- \chi _\varepsilon \right) {{\text {Op}}}^{\text {BW}}\!\left( g\right) \\&\quad + {{\text {Op}}}^{\text {BW}}\!\left( g\right) {{\text {Op}}}^{\text {BW}}\!\left( |\xi |^2 {\mathtt {a}}_-\chi _\varepsilon \right) {{\text {Op}}}^{\text {BW}}\!\left( f\right) \\ B_2&= {{\text {Op}}}^{\text {BW}}\!\left( f\right) {{\text {Op}}}^{\text {BW}}\!\left( |\xi |^2(1+{\mathtt {a}}_+)\chi _\varepsilon \right) {{\text {Op}}}^{\text {BW}}\!\left( g\right) \\&\quad + {{\text {Op}}}^{\text {BW}}\!\left( g\right) {{\text {Op}}}^{\text {BW}}\!\left( |\xi |^2(1+{\mathtt {a}}_+)\chi _\varepsilon \right) {{\text {Op}}}^{\text {BW}}\!\left( f\right) \\&\quad +{{\text {Op}}}^{\text {BW}}\!\left( f \right) {{\text {Op}}}^{\text {BW}}\!\left( |\xi |^2 {\mathtt {a}}_-\chi _\varepsilon \right) {{\text {Op}}}^{\text {BW}}\!\left( f\right) \\&\quad + {{\text {Op}}}^{\text {BW}}\!\left( g\right) {{\text {Op}}}^{\text {BW}}\!\left( |\xi |^2 {\mathtt {a}}_-\chi _\varepsilon \right) {{\text {Op}}}^{\text {BW}}\!\left( g\right) . \end{aligned}$$

By Corollary 2.6 we obtain

$$\begin{aligned} D_2&= {{\text {Op}}}^{\text {BW}}\!\left( \left[ (f^2+g^2)(1+{\mathtt {a}}_+) +2fg {\mathtt {a}}_-\right] |\xi |^2\chi _\varepsilon \right) +{\mathcal {F}}_1(U)\\&={{\text {Op}}}^{\text {BW}}\!\left( \lambda (U)|\xi |^2\chi _\varepsilon \right) + {\mathcal {F}}_1(U), \\ B_2&= {{\text {Op}}}^{\text {BW}}\!\left( \left[ (f^2+g^2){\mathtt {a}}_-+2fg(1+{\mathtt {a}}_+)\right] |\xi |^2\chi _\varepsilon \right) + {\mathcal {F}}_2(U)={\mathcal {F}}_2(U), \end{aligned}$$

where \( {\mathcal {F}}_1,{\mathcal {F}}_2 \) satisfy (4.16) by (2.25), (4.12), and (4.14) and since, by the definition of f and g in (4.8) and \( \lambda \) in (4.6), we have \( (f^2+g^2)(1+{\mathtt {a}}_+) + 2fg{\mathtt {a}}_- = \lambda \) and \( (f^2+g^2){\mathtt {a}}_-+2fg(1+{\mathtt {a}}_+)=0 \). Moreover

$$\begin{aligned} {{\text {Op}}}^{\text {BW}}\!\left( F^{-1}\right) {\mathbb {J}}{{\text {Op}}}^{\text {BW}}\!\left( A_1\chi _\varepsilon \right) {{\text {Op}}}^{\text {BW}}\!\left( F\right) = {\mathbb {J}}\begin{bmatrix} D_1 &{} B_1\\ -B_1&{} -D_1 \end{bmatrix}, \end{aligned}$$

where

$$\begin{aligned} D_1&= {{\text {Op}}}^{\text {BW}}\!\left( f\right) {{\text {Op}}}^{\text {BW}}\!\left( {\mathtt {b}}\cdot \xi \chi _\varepsilon \right) {{\text {Op}}}^{\text {BW}}\!\left( f\right) - {{\text {Op}}}^{\text {BW}}\!\left( g\right) {{\text {Op}}}^{\text {BW}}\!\left( {\mathtt {b}}\cdot \xi \chi _\varepsilon \right) {{\text {Op}}}^{\text {BW}}\!\left( g\right) \\ B_1&= {{\text {Op}}}^{\text {BW}}\!\left( f\right) {{\text {Op}}}^{\text {BW}}\!\left( {\mathtt {b}}\cdot \xi \chi _\varepsilon \right) {{\text {Op}}}^{\text {BW}}\!\left( g\right) - {{\text {Op}}}^{\text {BW}}\!\left( g\right) {{\text {Op}}}^{\text {BW}}\!\left( {\mathtt {b}}\cdot \xi \chi _\varepsilon \right) {{\text {Op}}}^{\text {BW}}\!\left( f\right) . \end{aligned}$$

Applying Theorem 2.5, (4.12), (4.14), using that \(f^2-g^2=1\) we obtain \( D_1= {{\text {Op}}}^{\text {BW}}\!\left( {\mathtt {b}}\cdot \xi \, \chi _\varepsilon \right) + {\mathcal {F}}_1(U) \) and \( B_1= {\mathcal {F}}_2(U) \) with \({\mathcal {F}}_1,{\mathcal {F}}_2\) satisfying (4.16). \(\square \)

4.2 Energy Estimate for Smoothed System

We first solve (4.3) in the case \(R(t)=0\) and \(V_0 \in \dot{C}^\infty := \cap _{{\sigma }\in {\mathbb {R}}} {\dot{{\mathbf {H}}}}^{\sigma }\). Consider the regularized Cauchy problem

$$\begin{aligned} \partial _t V^\varepsilon = {\mathbb {J}}{{\text {Op}}}^{\text {BW}}\!\left( A^\varepsilon (U(t);x,\xi )\right) V^\varepsilon , \quad V^\varepsilon (0)= V_0 \in \dot{C}^\infty , \end{aligned}$$
(4.17)

where \(A^\varepsilon (U; x, \xi ) \) is defined in (4.13). As the operator \( {{\text {Op}}}^{\text {BW}}\!\left( A^\varepsilon (U;x,\xi )\right) \) is bounded for any \( \varepsilon >0\), and U(t) satisfies (4.2), the differential equation (4.17) has a unique solution \(V^\varepsilon (t)\) which belongs to \( C^2 ([0,T],{\dot{{\mathbf {H}}}}^\sigma )\) for any \( \sigma \ge 0\). The important fact is that it admits the following \(\varepsilon \)-independent energy estimate.

Proposition 4.4

(Energy estimate) Let U satisfy (4.2). For any \(\sigma \ge 0\), there exist constants \( C_r, C_{\Theta } > 0 \) (depending also on \(\sigma \)), such that for any \( \varepsilon > 0 \), the unique solution of (4.17)

$$\begin{aligned} \Vert V^\varepsilon (t)\Vert _{\sigma }^2\le C_{r}\Vert V_0\Vert _{\sigma }^2+ C_{\Theta } \int _0^t \Vert V^\varepsilon (\tau )\Vert ^2_{\sigma } \, {\text {d}}\tau , \quad \forall t\in [0,T] \,. \end{aligned}$$
(4.18)

As a consequence, there are constants \( C_r, C_\Theta \) independent of \(\varepsilon \), such that

$$\begin{aligned} {\Vert V^\varepsilon (t)\Vert _{{{\sigma }}}\le C_r e^{C_\Theta t} \Vert V_0\Vert _{{{\sigma }}}, \quad \forall t\in [0,T].} \end{aligned}$$
(4.19)

In order to prove Proposition 4.4, we define, for any \(\sigma \ge 0\), the modified energy

$$\begin{aligned} \Vert V\Vert _{{\sigma },U}^2 := \langle {{\text {Op}}}^{\text {BW}}\!\left( \lambda ^{\sigma }(U;x) |\xi |^{2{\sigma }}\right) {{\text {Op}}}^{\text {BW}}\!\left( F^{-1}(U;x)\right) V, {{\text {Op}}}^{\text {BW}}\!\left( F^{-1}(U;x)\right) V\rangle ,\nonumber \\ \end{aligned}$$
(4.20)

where we introduce the real scalar product

$$\begin{aligned} \langle V,W\rangle :=2 {{\text {Re }}}\, \int _{{\mathbb {T}}^d} v(x) {\overline{w}} (x)\, {\text {d}}x, \qquad V=\begin{bmatrix} v \\ {\overline{v}}\end{bmatrix}, \ \ \ W=\begin{bmatrix} w \\ {\overline{w}}\end{bmatrix}. \end{aligned}$$

Lemma 4.5

Fix \(\sigma \ge 0\), \(r >0\). There exists a constant \(C_{r}>0\) (depending also on \({\sigma }\)) such that for any \(U \in {\dot{{\mathbf {H}}}}^{s_0}\) with \(\left\| U \right\| _{s_0} \le r\) and \(\rho (U) \in {\mathcal {Q}}\) we have

$$\begin{aligned} C_{r}^{-1} \Vert V\Vert _{{\sigma }}^2 -\Vert V\Vert _{-2}^2 \le \Vert V\Vert _{{{\sigma }},U}^2\le C_{r} \Vert V\Vert _{{\sigma }}^2, \quad \forall V\in {\dot{{\mathbf {H}}}}^{{\sigma }}. \end{aligned}$$
(4.21)

Proof

We first prove the upper bound in (4.21). We note that, by (4.12), \(\lambda ^{\sigma }(U;x) |\xi |^{2{\sigma }}\in \Gamma _{s_0}^{2{\sigma }}\) and \(F^{-1}(U;x) \in \Gamma _{s_0}^0\otimes {\mathcal {M}}_2({\mathbb {C}})\) and, by Theorem 2.4 and (4.12) we have

$$\begin{aligned} \Vert V\Vert _{{\sigma },U}^2&\le \Vert {{\text {Op}}}^{\text {BW}}\!\left( \lambda ^{\sigma }(U;x) |\xi |^{2{\sigma }}\right) {{\text {Op}}}^{\text {BW}}\!\left( F^{-1}(U;x)\right) V\Vert _{{-{\sigma }}}\Vert {{\text {Op}}}^{\text {BW}}\!\left( F^{-1}(U;x)\right) V\Vert _{{\sigma }}\\&\le C_{r} \Vert V\Vert _{{\sigma }}^2. \end{aligned}$$

In order to prove the lower bound, we fix \(\delta \in (0,1)\) such that \(s_0-\delta >\frac{d}{2}\) and, due to (4.7), we have \( \lambda ^{-\frac{{\sigma }}{2}}\in \Gamma _{s_0-\delta }^{0} \). So, applying Theorem 2.5 and (4.12) with \(s_0-\delta \) instead of \(s_0\) and with \(\varrho =\delta \), we have

(4.22)

where for any \(\sigma ' \in {\mathbb {R}}\) there exists a constant \(C_{r, \sigma '}>0\) such that

$$\begin{aligned} \Vert {\mathcal {F}}^{-\delta }(U)f\Vert _{\sigma '}\le C_{r,\sigma '}\Vert f\Vert _{{\sigma '-\delta }}, \quad \forall f\in {\dot{{\mathbf {H}}}}^{\sigma '-\delta }. \end{aligned}$$
(4.23)

Again, applying Theorem 2.5 with \(s_0-\delta \) instead of \(s_0\) and with \(\varrho =\delta \), we have also

$$\begin{aligned} {{\text {Op}}}^{\text {BW}}\!\left( \lambda ^{\frac{{\sigma }}{2}}\right) {{\text {Op}}}^{\text {BW}}\!\left( |\xi |^{2{\sigma }}\right) {{\text {Op}}}^{\text {BW}}\!\left( \lambda ^{\frac{{\sigma }}{2}}\right) = {{\text {Op}}}^{\text {BW}}\!\left( \lambda ^{\sigma }|\xi |^{2{\sigma }}\right) + {\mathcal {F}}^{2{\sigma }-\delta }(U), \end{aligned}$$
(4.24)

where for any \(\sigma ' \in {\mathbb {R}}\) there exists a constant \(C_{r,\sigma '}>0\) such that

$$\begin{aligned} \Vert {\mathcal {F}}^{2{\sigma }-\delta }(U)f\Vert _{{\sigma '-2{\sigma }+\delta }}\le C_{r,\sigma '}\Vert f\Vert _{{\sigma '}}, \quad \forall f\in {\dot{{\mathbf {H}}}}^{\sigma '}. \end{aligned}$$
(4.25)

By (4.22)–(4.25), Theorem 2.4 and (4.12) and using also that \({{\text {Op}}}^{\text {BW}}\!\left( \lambda ^{\frac{{\sigma }}{2}}\right) \) is symmetric with respect to \(\langle \cdot , \cdot \rangle \), we have

$$\begin{aligned} \Vert V\Vert _{{\sigma }}^2&\le 2 \Vert {{\text {Op}}}^{\text {BW}}\!\left( \lambda ^{-\frac{{\sigma }}{2}}\right) {{\text {Op}}}^{\text {BW}}\!\left( F\right) {{\text {Op}}}^{\text {BW}}\!\left( \lambda ^{\frac{{\sigma }}{2}}\right) {{\text {Op}}}^{\text {BW}}\!\left( F^{-1}\right) V\Vert _{{\sigma }}^2+ 2\Vert {\mathcal {F}}^{-\delta }(U)V\Vert _{{\sigma }}^2\\&\le C_{r} \left( \Vert {{\text {Op}}}^{\text {BW}}\!\left( \lambda ^{\frac{{\sigma }}{2}}\right) {{\text {Op}}}^{\text {BW}}\!\left( F^{-1}\right) V\Vert _{{\sigma }}^2+ \Vert V\Vert _{{{\sigma }-\delta }}^2\right) \\&=C_{r}\left( \langle {{\text {Op}}}^{\text {BW}}\!\left( \lambda ^{\frac{{\sigma }}{2}}\right) {{\text {Op}}}^{\text {BW}}\!\left( |\xi |^{2{\sigma }}\right) {{\text {Op}}}^{\text {BW}}\!\left( \lambda ^{\frac{{\sigma }}{2}}\right) {{\text {Op}}}^{\text {BW}}\!\left( F^{-1}\right) V,{{\text {Op}}}^{\text {BW}}\!\left( F^{-1}\right) V\rangle \right. \\&\quad \left. +\Vert V\Vert _{{{\sigma }-\delta }}^2\right) \\&= C_{r}\left( \Vert V\Vert _{{\sigma },U}^2+ \langle {\mathcal {F}}^{2{\sigma }-\delta }(U) {{\text {Op}}}^{\text {BW}}\!\left( F^{-1}\right) V, {{\text {Op}}}^{\text {BW}}\!\left( F^{-1}\right) V\rangle +\Vert V\Vert _{{{\sigma }-\delta }}^2\right) \\&\le C_{r}\big ( \Vert V\Vert _{{\sigma },U}^2+\Vert V\Vert _{{{\sigma }-\frac{\delta }{2}}}^2\big ). \end{aligned}$$

Now we use (2.3) and the asymmetric Young inequality to get, for any \(\epsilon > 0\),

$$\begin{aligned} \left\| V \right\| _{{{\sigma }-\frac{\delta }{2}}}^2 \le \left\| V \right\| _{{-2}}^{\frac{\delta }{{\sigma }+2}} \ \left\| V \right\| _{{\sigma }}^{\frac{2({\sigma }+2)-\delta }{{\sigma }+2}} \le \epsilon ^{-\frac{2({\sigma }+2)}{\delta }} \left\| V \right\| _{{-2}}^2 + \epsilon ^{\frac{2({\sigma }+2)}{2({\sigma }+2)-\delta }}\left\| V \right\| _{{\sigma }}^2 ; \end{aligned}$$

we choose \(\epsilon \) so small so that \( \epsilon ^{\frac{2({\sigma }+2)}{2({\sigma }+2)-\delta }}C_{r}=\frac{1}{2}\) and we get \( \Vert V\Vert _{{\sigma }}^2\le 2C_{r} \big ( \Vert V\Vert _{{\sigma },U}^2+\Vert V\Vert _{{-2}}^2 \big ) \). This proves the lower bound in (4.21). \(\square \)

Proof of Proposition 4.4

The time derivative of the modified energy (4.20) along a solution \( V^\varepsilon (t) \) of (4.17) is

$$\begin{aligned} \frac{d}{dt} \Vert V^\varepsilon \Vert _{{\sigma },U(t)}^2&= \langle {{\text {Op}}}^{\text {BW}}\!\left( \partial _t(\lambda ^{\sigma }) |\xi |^{2{\sigma }}\right) {{\text {Op}}}^{\text {BW}}\!\left( F^{-1}\right) V^\varepsilon , {{\text {Op}}}^{\text {BW}}\!\left( F^{-1}\right) V^\varepsilon \rangle \end{aligned}$$
(4.26)
$$\begin{aligned}&\quad + 2\langle {{\text {Op}}}^{\text {BW}}\!\left( \lambda ^{\sigma }|\xi |^{2{\sigma }}\right) {{\text {Op}}}^{\text {BW}}\!\left( \partial _t F^{-1}\right) V^\varepsilon , {{\text {Op}}}^{\text {BW}}\!\left( F^{-1}\right) V^\varepsilon \rangle \end{aligned}$$
(4.27)
$$\begin{aligned}&\quad +2\langle {{\text {Op}}}^{\text {BW}}\!\left( \lambda ^{\sigma }|\xi |^{2{\sigma }}\right) {{\text {Op}}}^{\text {BW}}\!\left( F^{-1}\right) \partial _t V^\varepsilon , {{\text {Op}}}^{\text {BW}}\!\left( F^{-1}\right) V^\varepsilon \rangle . \end{aligned}$$
(4.28)

By Theorem 2.4 and using that \(\forall \sigma \ge 0 \), \( N \in {\mathbb {N}}_0 \),

$$\begin{aligned} \left| \partial _t \lambda ^\sigma (U) |\xi |^{2{\sigma }} \right| _{2{\sigma }, s_0, N}, \quad \left| \partial _t F^{-1}(U) \right| _{0, s_0, N} \le {\mathtt {C}}_N(\left\| U \right\| _{s_0}, \left\| \partial _t U \right\| _{s_0}) \end{aligned}$$

and the assumption (4.2), there exists a constant \(C_{\Theta }>0\) (depending also on \({\sigma }\)) such that

$$\begin{aligned} (4.26)+(4.27)\le C_{\Theta } \Vert V^\varepsilon \Vert _{{\sigma }}^2. \end{aligned}$$
(4.29)

We now estimate (4.28). By Theorem 2.5 with \( \varrho = 2 \) and (4.2) we have

(4.30)

where \( {\mathcal {F}}_\pm ^{-2}(U)\) are bounded operators from \({\dot{{\mathbf {H}}}}^{\sigma '}\) to \({\dot{{\mathbf {H}}}}^{\sigma '+2}\), \(\forall \sigma ' \in {\mathbb {R}}\), satisfying

$$\begin{aligned} \Vert {\mathcal {F}}_\pm ^{-2}(U) W\Vert _{{\sigma '+2}}\le C_{\Theta ,\sigma '} \left\| W \right\| _{\sigma '}, \quad \forall \, W \in {\dot{{\mathbf {H}}}}^{\sigma '}. \end{aligned}$$
(4.31)

Thus, denoting \( {\widetilde{V}}^\varepsilon := {{\text {Op}}}^{\text {BW}}\!\left( F^{-1}\right) V^\varepsilon \), by (4.30), we have

$$\begin{aligned} {{\text {Op}}}^{\text {BW}}\!\left( F\right) {\widetilde{V}}^\varepsilon = V^\varepsilon + {\mathcal {F}}_+^{-2}(U)V^\varepsilon . \end{aligned}$$
(4.32)

Recalling (4.17) we have

$$\begin{aligned} (4.28)&= 2 \langle {{\text {Op}}}^{\text {BW}}\!\left( \lambda ^{\sigma }|\xi |^{2{\sigma }}\right) {{\text {Op}}}^{\text {BW}}\!\left( F^{-1}\right) {\mathbb {J}}{{\text {Op}}}^{\text {BW}}\!\left( A^\varepsilon \right) V^\varepsilon , {\widetilde{V}}^\varepsilon \rangle \\&\quad {\mathop {=}\limits ^{(4.32)}} 2 \langle {{\text {Op}}}^{\text {BW}}\!\left( \lambda ^{\sigma }|\xi |^{2{\sigma }}\right) {{\text {Op}}}^{\text {BW}}\!\left( F^{-1}\right) {\mathbb {J}}{{\text {Op}}}^{\text {BW}}\!\left( A^\varepsilon \right) {{\text {Op}}}^{\text {BW}}\!\left( F\right) {\widetilde{V}}^\varepsilon , {\widetilde{V}}^\varepsilon \rangle \\&\quad \quad \ - 2 \langle {{\text {Op}}}^{\text {BW}}\!\left( \lambda ^{\sigma }|\xi |^{2{\sigma }}\right) {{\text {Op}}}^{\text {BW}}\!\left( F^{-1}\right) {\mathbb {J}}{{\text {Op}}}^{\text {BW}}\!\left( A^\varepsilon \right) {\mathcal {F}}_+^{-2} V^\varepsilon , {\widetilde{V}}^\varepsilon \rangle \end{aligned}$$

and by Lemma 4.3 we get

$$\begin{aligned} (4.28){\mathop { =}\limits ^{(4.15)}}&\langle {\mathbb {J}}\big [ {{\text {Op}}}^{\text {BW}}\!\left( \lambda ^{\sigma }|\xi |^{2{\sigma }}\right) , {\text {Op}}^{BW} \big ( {\sqrt{{{\mathtt {m}}}K({{\mathtt {m}}})}\lambda |\xi |^2 \chi _\varepsilon } \big ) \big ] {\widetilde{V}}^\varepsilon , {\widetilde{V}}^\varepsilon \rangle \end{aligned}$$
(4.33)
$$\begin{aligned}&+ \langle {\mathbb {J}}\big [ {{\text {Op}}}^{\text {BW}}\!\left( \lambda ^{\sigma }|\xi |^{2{\sigma }}\right) , {{\text {Op}}}^{\text {BW}}\!\left( {\mathtt {b}}\cdot \xi \, \chi _\varepsilon \right) \big ]{\widetilde{V}}^\varepsilon , {\widetilde{V}}^\varepsilon \rangle \end{aligned}$$
(4.34)
$$\begin{aligned}&+2 \langle {{\text {Op}}}^{\text {BW}}\!\left( \lambda ^{\sigma }|\xi |^{2{\sigma }}\right) {\mathcal {F}} {\widetilde{V}}^\varepsilon , {\widetilde{V}}^\varepsilon \rangle \end{aligned}$$
(4.35)
$$\begin{aligned}&-2\langle {{\text {Op}}}^{\text {BW}}\!\left( \lambda ^{\sigma }|\xi |^{2{\sigma }}\right) {{\text {Op}}}^{\text {BW}}\!\left( F^{-1}\right) {\mathbb {J}}{{\text {Op}}}^{\text {BW}}\!\left( A^\varepsilon \right) {\mathcal {F}}_+^{-2} V^\varepsilon , {\widetilde{V}}^\varepsilon \rangle \end{aligned}$$
(4.36)

where in line (4.35) the operator \({\mathcal {F}}(U)\) is the bounded remainder of Lemma 4.3. We estimate each contribution. First we consider line (4.33). Using Theorem 2.5 with \( \varrho = 2 \), the principal symbol of the commutator is

$$\begin{aligned} {{\text {i}}^{-1}} \big \{ \lambda ^{\sigma }|\xi |^{2{\sigma }},\sqrt{{{\mathtt {m}}}K({{\mathtt {m}}})}\lambda |\xi |^2\chi \big (\varepsilon \lambda |\xi |^2\big ) \big \} = 0, \end{aligned}$$

and, using (4.14), (4.12) and assumption (4.2), we get

$$\begin{aligned} | (4.33)| \le {C_{\Theta }' \Vert {\widetilde{V}}^\varepsilon \Vert _{{\sigma }}^2} \le C_{\Theta } \Vert V^\varepsilon \Vert _{{\sigma }}^2. \end{aligned}$$
(4.37)

Similarly, using Theorem 2.5 with \(\varrho =1\), Theorem 2.4, (4.12) and estimates (4.31) and (4.16), we obtain

$$\begin{aligned} |(4.34)|+ | (4.35)|+ |(4.36)| \le C_{\Theta }\Vert V^\varepsilon \Vert _{{\sigma }}^2. \end{aligned}$$
(4.38)

In conclusion, by (4.29), (4.37), (4.38), we deduce the bound \(\frac{d}{dt} \Vert V^\varepsilon (t)\Vert _{{\sigma },U(t)}^2 \le C_{\Theta } \left\| V^\varepsilon (t) \right\| _{\sigma }^2 \), that gives, for any \( t \in [0,T] \)

$$\begin{aligned} \Vert V^\varepsilon (t)\Vert _{{\sigma },U(t)}^2&\le \Vert V^\varepsilon (0)\Vert _{{\sigma },U(0)}^2+ C_{\Theta } \int _0^t \Vert V^\varepsilon (\tau )\Vert _{{\sigma }}^2 \, {\text {d}}\tau \nonumber \\&{\mathop {\le }\limits ^{(4.21)}} C_{r} \Vert V^\varepsilon (0)\Vert _{{\sigma }}^2+ C_{\Theta } \int _0^t \Vert V^\varepsilon (\tau )\Vert _{{\sigma }}^2 \, {\text {d}}\tau . \end{aligned}$$
(4.39)

Since \(V^\varepsilon (t)\) solves (4.17), by Theorem 2.4, (4.12), (4.14) there exists a constant \(C_{\Theta }>0\) (independent on \(\varepsilon \)) such that \( \Vert \partial _t V^\varepsilon (t)\Vert _{{-2}}^2\le C_{\Theta }\Vert V^\varepsilon (t)\Vert _{0}^2\le C_{\Theta }\Vert V^\varepsilon (t)\Vert _{{\sigma }}^2 \) and therefore

$$\begin{aligned} \Vert V^\varepsilon (t)\Vert _{{-2}}^2\le \Vert V^\varepsilon (0)\Vert _{{-2}}^2+C_{\Theta } \int _0^t \Vert V^\varepsilon (\tau )\Vert _{{\sigma }}^2 \, {\text {d}}\tau , \quad \forall t \in [0,T]. \end{aligned}$$
(4.40)

We finally deduce (4.18) by (4.39), the lower bound in (4.21) and (4.40). The estimate (4.19) follows by Gronwall inequality. \(\square \)

Proof of Proposition 4.2. By Proposition 4.4, Ascoli–Arzelá theorem ensures that, for any \( \sigma \ge 0 \), \(V^\varepsilon \) converges up to subsequence to a limit V in \( C^1([0,T],{\dot{{\mathbf {H}}}}^{\sigma })\), as \(\varepsilon \rightarrow 0\) that solves (4.3) with \(R(t)=0 \), initial datum \(V_0\in \dot{C}^\infty \), and satisfies \( \Vert V(t)\Vert _{{{\sigma }}}\le C_r e^{C_\Theta t} \Vert V_0\Vert _{{{\sigma }}} \), for any \( \sigma \ge 0 \). The case \(V_0\in {\dot{{\mathbf {H}}}}^{\sigma }\) follows by a classical approximation argument with smooth initial data. This shows that the propagator of \({\mathbb {J}}{{\text {Op}}}^{\text {BW}}\!\left( A_2(U(t);x, \xi )+A_1(U(t);x, \xi )\right) \) is, for any \( \sigma \ge 0 \), a well defined bounded linear operator

$$\begin{aligned}&\Phi (t):{\dot{{\mathbf {H}}}}^{\sigma }\mapsto {\dot{{\mathbf {H}}}}^{\sigma }, \ V_0\mapsto \Phi (t)V_0:= V(t), \ \forall t\in [0,T], \\&\quad {\text { satisfying}} \ \ \Vert \Phi (t) V_0 \Vert _\sigma \le C_r e^{C_\Theta t} \Vert V_0 \Vert _\sigma . \end{aligned}$$

In the inhomogeneous case \(R\not =0 \), the solutions of (4.3) is given by the Duhamel formula \( V(t)= \Phi (t) V_0 + \Phi (t)\int _0^t \Phi ^{-1}(\tau ) R(\tau )\, {\text {d}}\tau , \) and the estimate (4.4) follows.

4.3 Iterative Scheme

In order to prove that the nonlinear system (3.4) has a local in time solution we consider the sequence of linear Cauchy problems

$$\begin{aligned} {\mathcal {P}}_1:= & {} {\left\{ \begin{array}{ll} \partial _t U_1 = - {\mathbb {J}}\sqrt{{{\mathtt {m}}}K({{\mathtt {m}}})} \, \Delta U_1 \\ U_1(0) = U_0, \end{array}\right. }\nonumber \\ {\mathcal {P}}_n:= & {} {\left\{ \begin{array}{ll} \partial _t U_n = {\mathbb {J}}\, {{\text {Op}}}^{\text {BW}}\!\left( A(U_{n-1}; x, \xi ) \right) U_n + R(U_{n-1})\\ U_n(0) = U_0, \end{array}\right. } \end{aligned}$$

for \( n \ge 2 \), where \( A := A_2 + A_1 \), cfr. (3.6), (3.8). The strategy is to prove that the sequence of solutions \(U_n\) of the approximated problems \({\mathcal {P}}_n \) converges to a solution U of system (3.4).

Lemma 4.6

Let \( U_0\in {\dot{{\mathbf {H}}}}^s \), \( s > 2 + \frac{d}{2}\), such that \( \rho (U_0) \in {\mathcal {Q}}_\delta \) for some \(\delta >0\) (recall (3.5) and (4.1)) and define \(r:= 2\Vert U_0\Vert _{{s_0}} \). Then there exists a time \(T:=T(\Vert U_0\Vert _{{s_0+2}},\delta )>0\) such that, for any \(n\in {\mathbb {N}}\):

\((S0)_n\)::

The problem \({\mathcal {P}}_n\) admits a unique solution \(U_n\in C^0([0,T], {\dot{{\mathbf {H}}}}^s)\cap C^1 ([0,T], {\dot{{\mathbf {H}}}}^{s-2})\).

\((S1)_n\)::

For any \( t\in [0,T]\), \( \rho (U_n(t)) \) belongs to \({\mathcal {Q}}_{\frac{\delta }{2}}\).

\((S2)_n\)::

There exists a constant \({C_r\ge 1}\) (depending also on s) such that, defining \(\Theta :=4C_r \Vert U_0 \Vert _{{s_0+2}}\) and \( M:= 4C_r \Vert U_0\Vert _{{s}}\), for any \(1\le m\le n\) one has

$$\begin{aligned}&\Vert U_m\Vert _{L^\infty _T {\dot{{\mathbf {H}}}}^{s_0}} \le r \, ;\end{aligned}$$
(4.41)
$$\begin{aligned}&\Vert U_m\Vert _{L^{\infty }_T {\dot{{\mathbf {H}}}}^{s_0+2}}\le \Theta , \quad \Vert \partial _t U_m\Vert _{L^{\infty }_T {\dot{{\mathbf {H}}}}^{s_0}}\le C_r \Theta \, ;\end{aligned}$$
(4.42)
$$\begin{aligned}&\Vert U_m\Vert _{L^{\infty }_T {\dot{{\mathbf {H}}}}^{s}}\le M, \quad \Vert \partial _t U_m\Vert _{L^{\infty }_T {\dot{{\mathbf {H}}}}^{s-2}}\le C_r M . \end{aligned}$$
(4.43)
\((S3)_n\)::

For \(1\le m\le n\) one has

$$\begin{aligned} \Vert U_1\Vert _{L^{\infty }_T{\dot{{\mathbf {H}}}}^{s_0}} = r / 2, \qquad \Vert U_m -U_{m-1}\Vert _{L^{\infty }_T{\dot{{\mathbf {H}}}}^{s_0}}\le 2^{-m} {r}, \ \ m \ge 2. \end{aligned}$$

Proof

We prove the statement by induction on \(n\in {\mathbb {N}}\). Given \(r >0\), we define

$$\begin{aligned} C_r := \max \left\{ 1, C_{r, s_0}, \ C_{r, s_0+2}, \ C_{r,s}, \ 2 \, {\mathtt {C}}(r) \right\} , \end{aligned}$$

where \(C_{r, \sigma }\) is the constant in Proposition 4.2 (where we stress that it depends also on \({\sigma }\)) and \({\mathtt {C}}(\cdot )\) is the function in (3.9) and (3.11). In the following we shall denote by \( C_\Theta \) all the constants depending on \(\Theta \), which can vary from line to line.

Proof of \((S0)_1\): The problem \({\mathcal {P}}_1\) admits a unique global solution which preserves Sobolev norms.

Proof of \((S1)_1\): We have \(\rho (U_0) \in {\mathcal {Q}}_{\delta }\). In addition

$$\begin{aligned} \Vert \rho (U_1(t) - U_0) \Vert _{L^\infty ({\mathbb {T}}^d)} \lesssim \Vert U_1(t)- U_0\Vert _{{s_0}}\lesssim T \Vert U_0\Vert _{s_0+2} \le \delta /2 \end{aligned}$$

for \(T := T(\Vert U_0\Vert _{s_0+2}, \delta )>0\) sufficiently small, which implies \(\rho (U_1(t)) \in {\mathcal {Q}}_{\frac{\delta }{2}}\), for any \( t \in [0, T]\).

Proof of \((S2)_1\) and \((S3)_1\): The flow of \({\mathcal {P}}_1\) is an isometry and \(M\ge \Vert U_0\Vert _{ s}\), \( \Theta \ge \Vert U_0\Vert _{{s_0+2}}\).

Suppose that \((S0)_{n-1}\)\((S3)_{n-1}\) hold true. We prove \((S0)_{n}\)\((S3)_{n}\).

Proof of \((S0)_n\): We apply Proposition 4.2 with \(\sigma = s \), \( U \leadsto U_{n-1} \) and \(R(t) := R(U_{n-1}(t))\). By \((S1)_{n-1}\) and \((S2)_{n-1}\), the function \(U_{n-1}\) satisfies assumption (4.2) with \(\Theta \leadsto (1+C_r)\Theta \). In addition \( R(U_{n-1}(t))\) belongs to \(C^0([0, T], {\dot{{\mathbf {H}}}}^{s})\) thanks to (3.12) and \( U_{n-1} \in C^0 ([0,T]; {\dot{{\mathbf {H}}}}^s)\). Thus Proposition 4.2 with \(\sigma = s\) implies \((S0)_n \). In particular \( U_n \) satisfies the estimate (4.4).

Proof of \((S2)_n\): We first prove (4.42). The estimate (4.4) with \({\sigma }= s_0+2\), the bound (3.11) of \(R(U_{n-1}(t))\) and (4.42) at the step \( n - 1 \), imply

$$\begin{aligned} \Vert U_n\Vert _{L^\infty _T {\dot{{\mathbf {H}}}}^{s_0+2}}\le C_{r} e^{C_{\Theta }T} \Vert U_0\Vert _{{s_0+2}} + T C_{\Theta } e^{C_\Theta T} \Theta . \end{aligned}$$
(4.44)

As \(\Theta = 4C_r \Vert U_0\Vert _{{s_0+2}} \), we take \(T>0\) small such that

$$\begin{aligned} C_{\Theta }T\le 1, \quad T C_{\Theta } e^{C_\Theta T} \le 1 / 4 , \end{aligned}$$
(4.45)

which, by (4.44), gives \(\Vert U_n\Vert _{L^\infty _T{\dot{{\mathbf {H}}}}^{s_0+2}}\le \Theta \). This proves the first estimate of (4.42). Regarding the control of \(\partial _t U_n\), we use the equation \({\mathcal {P}}_n\), the second estimate in (3.11) and (3.9) with \({\sigma }=s_0\) to obtain

$$\begin{aligned} \Vert \partial _t U_n(t)\Vert _{{s_0}}\le {\mathtt {C}}\left( \Vert U_{n-1}(t)\Vert _{{s_0}}\right) \Vert U_n(t)\Vert _{{s_0+2}} + {\mathtt {C}}\left( \Vert U_{n-1}(t)\Vert _{{s_0}}\right) \Vert U_{n-1}(t)\Vert _{{s_0+2}} \le C_r\Theta \nonumber \\ \end{aligned}$$
(4.46)

which proves the second estimate of (4.42).

Next we prove (4.43). Applying estimate (4.4) with \({\sigma }= s\), we have

$$\begin{aligned} \Vert U_n\Vert _{L^\infty _T {\dot{{\mathbf {H}}}}^{s}}\le C_{r} e^{C_{\Theta }T} \Vert U_0\Vert _{s} + T C_{\Theta } e^{C_\Theta T} M \le M \end{aligned}$$

for \( M= 4C_r \Vert U_0\Vert _{s} \) and since \(T>0\) is chosen as in (4.45). The estimate for \(\left\| \partial _t U_n \right\| _{s-2}\) is similar to (4.46), and we omit it. Estimate (4.41) is a consequence of \((S3)_n\), which we prove below.

Proof of \((S1)_n\): We use estimate (4.46) to get

$$\begin{aligned} \Vert \rho (U_n(t) - U_0) \Vert _{L^\infty ({\mathbb {T}}^d)} \le C \Vert U_n(t)- U_0\Vert _{{s_0}}&\le C \int _0^T \Vert \partial _t U_n(t)\Vert _{{s_0}}\, {\text {d}}t \le C\, C_r\, T\, \Theta \le \delta / 2 \end{aligned}$$

provided that \(T< \delta /{(2 C C_r \Theta )}\). This shows that \(\rho (U_n(t)) \in {\mathcal {Q}}_{\frac{\delta }{2}}\).

Proof of \((S3)_n\): Define \(V_n:= U_n-U_{n-1}\) if \( n \ge 2 \) and \( V_1 := U_1 \). Note that \( V_n\), \( n \ge 2 \), solves

$$\begin{aligned} \partial _t V_n= {\mathbb {J}}{{\text {Op}}}^{\text {BW}}\!\left( A(U_{n-1})\right) V_n + f_n, \quad V_n(0)=0, \end{aligned}$$
(4.47)

where \(A:= A_2+A_1\) and

$$\begin{aligned} \begin{aligned} f_n&:= {\mathbb {J}}{{\text {Op}}}^{\text {BW}}\!\left( A(U_{n-1})- A(U_{n-2})\right) U_{n-1}+ R(U_{n-1}) - R(U_{n-2}), \ {\text {for}} \ n > 2, \\ f_2&:= {\mathbb {J}}{\text {Op}}^{BW}\big ({A(U_{1})- \sqrt{{{\mathtt {m}}}K({{\mathtt {m}}})}|\xi |^2} \big ) U_{1}+ R(U_{1}). \end{aligned} \end{aligned}$$

Applying estimates (3.13), (3.10), (3.11) and (4.42) we obtain, for \( n \ge 2 \),

$$\begin{aligned} \Vert f_n \Vert _{{s_0}} \le C_\Theta \Vert V_{n-1}\Vert _{s_0}, \quad \forall t \in [0,T]. \end{aligned}$$
(4.48)

We apply Proposition 4.2 to (4.47) with \({\sigma }=s_0\). Thus by (4.4) and (4.48) we get

$$\begin{aligned} \begin{aligned} \Vert V_n\Vert _{L^\infty _T {\dot{{\mathbf {H}}}}^{s_0}}&\le C_\Theta e^{C_\Theta T } T \Vert f_n\Vert _{L^\infty _T {\dot{{\mathbf {H}}}}^{s_0}} \le C_\Theta e^{C_\Theta T } T \Vert V_{n-1}\Vert _{L^\infty _T {\dot{{\mathbf {H}}}}^{s_0}} \le \frac{1}{2} \Vert V_{n-1}\Vert _{L^\infty _T {\dot{{\mathbf {H}}}}^{s_0}} \end{aligned} \end{aligned}$$

provided \( C_\Theta e^{C_\Theta T} T\le \frac{1}{2}\). The proof of Lemma 4.6 is complete. \(\square \)

Corollary 4.7

With the same assumptions of Lemma 4.6, for any \(s_0+2 \le s'<s\):

  1. (i)

    \((U_n)_{n \ge 1}\) is a Cauchy sequence in \(C^0([0,T],{\dot{{\mathbf {H}}}}^{s'})\cap C^1([0,T],{\dot{{\mathbf {H}}}}^{s'-2})\) with \( T = T(\Vert U_0\Vert _{{s_0+2}},\delta ) \) given by Lemma 4.6. It converges to the unique solution U(t) of (3.4) with initial datum \(U_0\), U(t) is in \( C^0([0,T],{\dot{{\mathbf {H}}}}^{s'})\cap C^1([0,T],{\dot{{\mathbf {H}}}}^{s'-2}) \). Moreover \(\rho (U(t)) \in {\mathcal {Q}}\), \( \forall t \in [0, T]\).

  2. (ii)

    For any \(t\in [0,T]\), \(U(t) \in {\dot{{\mathbf {H}}}}^s \) and \(\left\| U(t) \right\| _s \le 4 C_r \left\| U_0 \right\| _s\) where \(C_r\) is the constant of \((S2)_n\).

Proof

\(\mathrm{(i)}\) :

If \(s' = s_0\) it is the content of \((S3)_n\). For \( s' \in (s_0, s)\), we use interpolation estimate (2.3), (4.43) and \((S3)_n\) to get, for \( n \ge 2 \),

$$\begin{aligned} \Vert U_n-U_{n-1}\Vert _{L^\infty _T {\dot{{\mathbf {H}}}}^{s'}} \le \Vert U_n-U_{n-1}\Vert _{L^\infty _T {\dot{{\mathbf {H}}}}^{s_0}}^{\theta } \Vert U_n-U_{n-1}\Vert _{L^\infty _T {\dot{{\mathbf {H}}}}^{s}}^{1-\theta } \le 2^{-n\theta } C_M, \end{aligned}$$

where \(\theta \in (0,1) \) is chosen so that \(s' = \theta s_0 + (1-\theta )s\). Thus \((U_n)_{n \ge 1}\) is a Cauchy sequence in \(C^0([0,T], {\dot{{\mathbf {H}}}}^{s'})\); we denote by \(U(t)\in C^0([0,T],{\dot{{\mathbf {H}}}}^{s'})\) its limit. Similarly using that \( \partial _t U_n \) solves \( {\mathcal {P}}_n \), one proves that \(\partial _tU_n\) is a Cauchy sequence in \(C^0([0,T],{\dot{{\mathbf {H}}}}^{s'-2})\) that converges to \(\partial _t U\) in \(C^0([0,T],{\dot{{\mathbf {H}}}}^{s'-2})\). In order to prove that U(t) solves (3.4), it is enough to show that

$$\begin{aligned}&{\mathcal {R}}(U, U_{n-1}, U_n) := {\mathbb {J}}{{\text {Op}}}^{\text {BW}}\!\left( A(U_{n-1})\right) U_n - {\mathbb {J}}{{\text {Op}}}^{\text {BW}}\!\left( A(U)\right) U + R(U_{n-1}) - R(U) \end{aligned}$$

converges to 0 in \( L^\infty _T {\dot{{\mathbf {H}}}}^{s'-2} \). This holds true because by estimates (2.18), (3.10), (3.12), \( (S2)_n \), and the fact that \(U(t) \in C^0([0,T],{\dot{{\mathbf {H}}}}^{s'})\), we have

$$\begin{aligned}&\Vert {\mathcal {R}}(U, U_{n-1}, U_n) \Vert _{L^\infty _T {\dot{{\mathbf {H}}}}^{s'-2}}\le C_M\left( \Vert U - U_{n-1}\Vert _{L^\infty _T {\dot{{\mathbf {H}}}}^{s'}} + \left\| U - U_{n-1} \right\| _{L^\infty _T {\dot{{\mathbf {H}}}}^{s_0+2}} + \Vert U - U_n\Vert _{L^\infty _T {\dot{{\mathbf {H}}}}^{s'}} \right) \end{aligned}$$

which converges to 0 as \(n \rightarrow \infty \). Let us now prove the uniqueness. Suppose that \(V_1,V_2\in C^0([0,T],{\dot{{\mathbf {H}}}}^{s'})\cap C^1([0,T],{\dot{{\mathbf {H}}}}^{s'-2})\) are solutions of (3.4) with initial datum \(U_0\). Then \(W:= V_1-V_2\) solves

$$\begin{aligned} \partial _tW = {\mathbb {J}}{{\text {Op}}}^{\text {BW}}\!\left( A(V_1)\right) W + \varvec{R}(t), \quad W(0)=0, \end{aligned}$$

where \({\varvec{R}}(t):={\mathbb {J}}{{\text {Op}}}^{\text {BW}}\!\left( A(V_1)-A(V_2)\right) V_2+ R(V_1)-R(V_2). \) Applying Proposition 4.2 with \(\sigma = s_0\) and \(\Theta , r\) defined by

$$\begin{aligned} \Theta := \max _{j=1,2}\big ( \Vert V_j\Vert _{L^\infty _T {\dot{{\mathbf {H}}}}^{s_0+2}} + \Vert \partial _t V_j\Vert _{L^\infty _T {\dot{{\mathbf {H}}}}^{s_0}} \big ), \quad r := \max _{j = 1,2} \Vert V_j\Vert _{L^\infty _T {\dot{{\mathbf {H}}}}^{s_0}}, \end{aligned}$$

together with estimates (3.13) and (3.10) we have, for any \( t \in [0,T] \),

$$\begin{aligned} \Vert W\Vert _{{L^\infty _t {\dot{{\mathbf {H}}}}^{s_0}}}&\le C_\Theta e^{C_\Theta t }t \Vert {\varvec{R}}\Vert _{L^\infty _t {\dot{{\mathbf {H}}}}^{s_0}} \le C_\Theta e^{C_\Theta t }t \Vert W\Vert _{L^\infty _t {\dot{{\mathbf {H}}}}^{s_0}}. \end{aligned}$$

Therefore, provided t is so small that \(C_\Theta e^{C_\Theta t }t< 1\), we get \(V_1(\tau )=V_2(\tau )\) \(\, \forall \tau \in [0,t]\). As (3.4) is autonomous, actually one has \(V_1(t)=V_2(t)\) for all \(t\in [0,T]\). This proves the uniqueness. Finally, as \(\rho (U_n(t)) \in {\mathcal {Q}}_{\frac{\delta }{2}}\) and \(U_n(t) \rightarrow U(t)\) in \({\dot{{\mathbf {H}}}}^{s_0}\), then \(\rho (U(t)) \in {\mathcal {Q}}_{\frac{\delta }{2}} \subset {\mathcal {Q}}\).

\(\mathrm{(ii)}\) :

Since \( \Vert U_n(t)\Vert _s \le 4 C_r \left\| U_0 \right\| _s \) and \(U_n(t) \rightarrow U(t)\) in \({\dot{{\mathbf {H}}}}^{s'}\) then \(\Vert U (t)\Vert _s \le 4 C_r \left\| U_0 \right\| _s \). \(\square \)

Let \(\displaystyle { \Pi _{N} U := \Big ({\sum _{1\le |j|\le N} {u}_j e^{{\text {i}}j\cdot x}},{ \sum _{1\le |j|\le N} \overline{{ u}_{j}} e^{-{\text {i}}j \cdot x}}\Big )} \). We need below the following technical lemma.

Lemma 4.8

Let \( U_0\in {\dot{{\mathbf {H}}}}^s\), \( s > 2 + \frac{d}{2} \), with \(\rho (U_0) \in {\mathcal {Q}}_\delta \) for some \(\delta >0\). Then there exists a time \({\tilde{T}}:= {\tilde{T}}( \Vert U_0\Vert _{s_0+2}, \delta )>0\) and \(N_0 >0\) such that for any \(N > N_0\):

  1. (i)

    system (3.4) with initial datum \( \Pi _{N} U_0\) has a unique solution \(U_N \in C^0([0, {\tilde{T}}], {\dot{{\mathbf {H}}}}^{s+2})\).

  2. (ii)

    Let U be the unique solution of (3.4) with initial datum \(U_{0}\) defined in the time interval [0, T] (which exists by Corollary 4.7). Then there is \( \breve{T} < \min \{T, {\tilde{T}} \} \), depending on \( \Vert U_0 \Vert _s \), independent of N, such that

    $$\begin{aligned} \left\| U - U_N \right\| _{ L^\infty _{\breve{T}} {\dot{{\mathbf {H}}}}^s} \le {\mathtt {C}}(\left\| U_0 \right\| _s) \, \left( \left\| U_0 - \Pi _{N} U_{0} \right\| _s + N^{s_0 +2 -s} \right) . \end{aligned}$$
    (4.49)

    In particular \(U_N \rightarrow U\) in \(C^0([0, \breve{T}], {\dot{{\mathbf {H}}}}^{s})\) when \(N \rightarrow \infty \).

Proof

Clearly \( \Pi _{N} U_0 \in \dot{C}^\infty \). Moreover, as \( \Vert \rho (U_0 - \Pi _{N} U_0) \Vert _{L^\infty ({\mathbb {T}}^d)} \rightarrow 0 \) when \(N \rightarrow \infty \), one has \(\rho (\Pi _{N} U_0) \in {\mathcal {Q}}_{\frac{\delta }{2}}\) provided \(N \ge N_0 \) is sufficiently large. So we can apply Corollary 4.7 and obtain a time \({\tilde{T}} > 0 \), independent on N, and a unique solution \(U_N \in C^0([0, {\tilde{T}}], {\dot{{\mathbf {H}}}}^{s+2}) \) of (3.4) with initial datum \( \Pi _{N} U_0 \). Moreover, by item \(\mathrm{(ii)}\) of that corollary, setting \( r = 2 \left\| \Pi _{N} U_0 \right\| _{s_0} \),

$$\begin{aligned}&\Vert U_N\Vert _{{L^\infty _{{\tilde{T}}}} {\dot{{\mathbf {H}}}}^{s}}\le 4 {C_r \Vert \Pi _{N} U_0 \Vert _{{s}}} \le {\mathtt {C}}(\left\| U_{0} \right\| _{s_0}) \, \Vert U_0\Vert _{{s}},\end{aligned}$$
(4.50)
$$\begin{aligned}&\Vert U_N\Vert _{{L^\infty _{{\tilde{T}}}} {\dot{{\mathbf {H}}}}^{s+2}} \le 4 {C_r \Vert \Pi _{N} U_0 \Vert _{{s+2}}} \le {\mathtt {C}}(\left\| U_{0} \right\| _{s_0}) \, N^2 \, \Vert U_0\Vert _{{s}}. \end{aligned}$$
(4.51)

This proves item \(\mathrm{(i)}\). In the following let \( {\breve{T}}\le \min \{ {\tilde{T}}, T \} \).

Let us prove \(\mathrm{(ii)}\). Let \( \Theta := \left\| U \right\| _{L^\infty _{\breve{T}}{\dot{{\mathbf {H}}}}^{s_0+2} } + \left\| \partial _t U \right\| _{L^\infty _{\breve{T}}{\dot{{\mathbf {H}}}}^{s_0} }\) and \( r := \Vert U\Vert _{L^\infty _{\breve{T}}{\dot{{\mathbf {H}}}}^{s_0}} \). The function \(W_N(t):=U(t)-U_N(t)\) satisfies \( \left\| W_N(t) \right\| _s \le \left\| U(t) \right\| _s + \left\| U_N(t) \right\| _s \le {\mathtt {C}}(\left\| U_{0} \right\| _{s}) \), \(\, \forall t \in [0, {\breve{T}}]\), by Corollary 4.7-(ii). Moreover, \(W_N\) solves

$$\begin{aligned} \partial _t W_N&= {\mathbb {J}}{{\text {Op}}}^{\text {BW}}\!\left( A(U)\right) W_N +{\varvec{R}}(t), \quad W_N(0) = U_0- \Pi _{N} U_0 \end{aligned}$$

where \( {\varvec{R}}(t):= {\mathbb {J}}{{\text {Op}}}^{\text {BW}}\!\left( A(U)-A(U_N)\right) U_N + R(U)-R(U_N). \) Applying Proposition 4.2 with \( \sigma = s_0 \) and estimates (3.10), (3.13), (4.50) one obtains

$$\begin{aligned} \Vert W_N\Vert _{L^\infty _{ {\breve{T}}} {\dot{{\mathbf {H}}}}^{s_0 }}&\le C_r e^{C_\Theta {\breve{T}}} \Vert U_0 - \Pi _{N} U_0 \Vert _{{s_0}} + {\breve{T}}C_\Theta e^{C_\Theta {\breve{T}}}\, {\mathtt {C}}(\Vert {U_0} \Vert _{s_0+2}) \left\| W_N \right\| _{L^\infty _{\breve{T}}{\dot{{\mathbf {H}}}}^{s_0}}, \end{aligned}$$

which, provided \({\breve{T}}\) is so small that \( {\breve{T}}C_\Theta e^{C_\Theta {\breve{T}}}\, {\mathtt {C}}(\Vert U_0\Vert _{s_0+2}) \le \frac{1}{2}\) (eventually shrinking it), gives

$$\begin{aligned} \Vert W_N\Vert _{{L^\infty _{\breve{T}}{\dot{{\mathbf {H}}}}^{s_0}}} \le C_r \Vert U_0- \Pi _{N} U_0 \Vert _{{s_0}} \le C_r \, N^{s_0 - s} \left\| U_0 \right\| _s. \end{aligned}$$
(4.52)

Similarly one estimates \(\left\| W_N(t) \right\| _{{\dot{{\mathbf {H}}}}^s}\), getting

$$\begin{aligned}&\Vert W_N\Vert _{{L^\infty _{\breve{T}}{\dot{{\mathbf {H}}}}^s }} \\&\quad \le C_r e^{C_\Theta {\breve{T}}}\Vert U_0- \Pi _{N} U_0 \Vert _s + C_\Theta e^{C_\Theta {\breve{T}}} {\breve{T}}{\mathtt {C}}\big ( \Vert U_0\Vert _{s} \big ) \left( \Vert W_N\Vert _{L^\infty _{\breve{T}}{\dot{{\mathbf {H}}}}^{s_0}} \Vert {U_N}\Vert _{L^\infty _{\breve{T}}{\dot{{\mathbf {H}}}}^{s+2}} + \Vert W_N\Vert _{L^\infty _{\breve{T}}{\dot{{\mathbf {H}}}}^s}\right) \\&\quad {\mathop {\le }\limits ^{(4.51),(4.52)}} C_r e^{C_\Theta {\breve{T}}} \Vert U_0- \Pi _{N} U_0\Vert _{{s}} + C_\Theta e^{C_\Theta {\breve{T}}} {\breve{T}}{\mathtt {C}}(\Vert U_0\Vert _{s}) \big ( N^{s_0-s +2} + \Vert W_N\Vert _{L^\infty _{\breve{T}}{\dot{{\mathbf {H}}}}^s}\big ) \end{aligned}$$

from which (4.49) follows provided \( {\breve{T}}\) (depending on \(\Vert U_0\Vert _{s}\)) is sufficiently small. \(\square \)

Proof of Proposition 4.1

Given an initial datum \( U_0\in {\dot{{\mathbf {H}}}}^s\) with \(\rho (U_0) \in {\mathcal {Q}}\), choose \(\delta >0\) so small that \( \rho (U_0) \in {\mathcal {Q}}_\delta \). Then Corollary 4.7 gives us a time \(T = T(\left\| U_0 \right\| _{s_0+2}, \delta ) > 0 \) and a unique solution \(U \in C^0([0,T],{\dot{{\mathbf {H}}}}^{s'})\cap C^1([0,T],{\dot{{\mathbf {H}}}}^{s'-2})\), \(\forall s_0+2 \le s' < s\), of (3.4) with initial datum \(U_0\). Now take an open neighborhood \({\mathcal {U}}\subset {\dot{{\mathbf {H}}}}^{s}\) of \(U_0\) such that \(\forall V \in {\mathcal {U}}\) one has \(\rho (V) \in {\mathcal {Q}}_{\frac{\delta }{2}} \) and \(\left\| V \right\| _s \le 2 \left\| U_0 \right\| _s\). Then there exists \({\tilde{T}} \in (0, T] \) such that the flow map of (3.4),

$$\begin{aligned} \Omega ^t: {\mathcal {U}}\rightarrow {\dot{{\mathbf {H}}}}^s \cap \big \{ U \in {\dot{{\mathbf {H}}}}^s: \, \rho (U)\in {\mathcal {Q}}_{\frac{\delta }{4}} \big \}, \quad U_0 \mapsto \Omega ^t(U_0):= U(t), \end{aligned}$$

is well defined for any \(t \in [0, {\tilde{T}}]\), it satisfies the group property

$$\begin{aligned} \Omega ^{t+\tau }=\Omega ^t\circ \Omega ^\tau , \quad \forall t,\tau , t+\tau \in [0,{\tilde{T}}], \end{aligned}$$
(4.53)

and \( \left\| \Omega ^t(U_0) \right\| _s \le {\mathtt {C}}(\left\| U_0 \right\| _s) \) for all \( U_0 \in {\mathcal {U}}\), \( t\in [0, {\tilde{T}}] \). For simplicity of notation in the sequel we denote by T a time, independent of N, smaller than \( {\tilde{T}} \).

Continuity of \(t \mapsto U(t)\): We show that \( U \in C^0([0,T], {\dot{{\mathbf {H}}}}^s )\). By (4.53), it is enough to prove that \(t\mapsto U(t)\) is continuous in a neighborhood of \(t=0\). This follows by Lemma 4.8, as U is the uniform limit of continuous functions.

Continuity of the flow map: We shall follow the method by [8, 13]. Let \( U^n_0 \rightarrow U_0\in {\dot{{\mathbf {H}}}}^s\) and pick \(\delta >0\) such that \(\rho (U^n_0)\), \(\rho (U_0) \), \(\rho (\Pi _{N} U^n_0)\), \(\rho (\Pi _{N} U_0) \in {\mathcal {Q}}_{\delta }\), for any \( n \ge n_0 \), \( N \ge N_0 \) sufficiently large. Denote by \(U^n,U \in C^0([0,T], {\dot{{\mathbf {H}}}}^s)\) the solutions of (3.4) with initial data \( U^n_0 \), respectively \( U_0 \), and \( U_N(t):= \Omega ^t(\Pi _{ N} U_0)\), \( U^n_N(t):= \Omega ^t(\Pi _{ N} U^n_0)\). Note that these solutions are well defined in \({\dot{{\mathbf {H}}}}^s\) up to a common time \(T' \in (0,T]\), depending on \( \Vert U_0 \Vert _s \), thanks to Lemma 4.8. By triangular inequality we have, by (4.49), for any \( n \ge n_0 \), \( N \ge N_0 \),

(4.54)

For any \(\varepsilon > 0 \), since \(s>s_0+2\), there exists \(N_\varepsilon \in {\mathbb {N}}\) (independent of n) such that

(4.55)

Consider now the term \( \Vert U^n_N - U_N\Vert _{L^\infty _T {\dot{{\mathbf {H}}}}^s} \). As \(\Pi _{ N} U_0, \Pi _{ N} U_0^n \in \dot{C}^\infty \), the solutions \(U_N(t)\), \(U_N^n(t) \) belong actually to \( {\dot{{\mathbf {H}}}}^{s+2}\). By interpolation and by item \(\mathrm{(ii)}\) of Corollary 4.7 applied with \( s \leadsto s+2\) one has, for \( s+ 2 = \theta s_0 + (1-\theta ) (s+2) \),

$$\begin{aligned} \Vert U^n_{N_\varepsilon } - U_{N_\varepsilon }\Vert _{L^\infty _{T} {\dot{{\mathbf {H}}}}^s}&\le {\mathtt {C}}\big ( \Vert \Pi _{N_\varepsilon }U_{0}\Vert _{s+2},\Vert \Pi _{N_\varepsilon } U^n_{0} \Vert _{s+2}\big ) \Vert U^n_{N_\varepsilon } - U_{N_\varepsilon }\Vert _{L^\infty _T \dot{\mathbf {H}}^{s_0}}^{\theta } \nonumber \\&\le {\mathtt {C}}\big ( N_\varepsilon ^2\Vert U_{0}\Vert _{s}\big ) \Vert U^n_{N_\varepsilon } - U_{N_\varepsilon }\Vert _{L^\infty _T {\dot{{\mathbf {H}}}}^{s_0}}^{\theta }. \end{aligned}$$
(4.56)

Arguing in the same way of the proof of (4.52) one obtains

$$\begin{aligned} \Vert U^n_{N_\varepsilon } - U_{N_\varepsilon }\Vert _{L^\infty _T {\dot{{\mathbf {H}}}}^{s_0}}\le {\mathtt {C}}\big ( \Vert U_0\Vert _{s_0+2}\big ) \Vert \Pi _{ N_\varepsilon } (U^n_{ 0}- U_{ 0})\Vert _{s_0}. \end{aligned}$$
(4.57)

By (4.54)–(4.57), we have \( \limsup _{n\rightarrow \infty } \Vert U^n-U\Vert _{L^\infty _T {\dot{{\mathbf {H}}}}^s}\le \varepsilon \), \( \forall \varepsilon \in (0,1).\) \(\square \)