1 Introduction

1.1 The problem

The evolution of an inviscid perfect fluid that occupies a domain \(\Omega _t\) in \({\mathbb {R}}^n\) (\(n \ge 2\)) at time \(t\), is described by the free boundary incompressible Euler equations. If \(v\) and \(p\) denote respectively the velocity and the pressure of the fluid (which is assumed here to have constant density equal to \(1\)), these equations are:

$$\begin{aligned} \left\{ \begin{array}{l@{\quad }l} v_t + v \cdot \nabla v = - \nabla p - g e_n &{} x \in \Omega _t\\ \nabla \cdot v = 0 &{} x \in \Omega _t\\ v (0,x) = v_0 (x) &{} x \in \Omega _0 , \end{array} \right. \end{aligned}$$
(E)

where \(g\) is the gravitational constant, which we will assume to be \(1\) from now on. The free surface \(S_t := \partial \Omega _t\) moves with the normal component of the velocity, and, in absence of surface tension, the pressure vanishes on the boundary:

$$\begin{aligned} \left\{ \begin{array}{l} \partial _t + v \cdot \nabla \,\, \text{ is } \text{ tangent } \text{ to } \,\, \bigcup _t S_t \subset {\mathbb {R}}^{n+1}\\ p (t,x) = 0 , \,\,\, x \in S_t . \end{array} \right. \end{aligned}$$
(BC)

In the case of irrotational flows, i.e.

$$\begin{aligned} {\,\text{ curl }\,}v = 0 , \end{aligned}$$
(1.1)

one can reduce (E)–(BC) to a system on the boundary. Although this reduction can be performed identically regardless of the number of spatial dimensions, we only focus on the two dimensional case which is the one we are interested in. Assume that \(\Omega _t \subset {\mathbb {R}}^2\) is the region below the graph of a function \(h : {\mathbb {R}}_t \times {\mathbb {R}}_x \rightarrow {\mathbb {R}}\), that is \(\Omega _t = \{ (x,y) \in {\mathbb {R}}^2 \, : y \le h(t,x) \}\) and \(S_t = \{ (x,y) : y = h(t,x) \}\). Let us denote by \(\Phi \) the velocity potential: \(\nabla \Phi (t,x,y) = v (t,x,y)\), for \((x,y) \in \Omega _t\). If \(\phi (t,x) := \Phi (t, x, h(x,t))\) is the restriction of \(\Phi \) to the boundary \(S_t\), the equations of motion reduce to the following system for the unknowns \(h, \phi : {\mathbb {R}}_t \times {\mathbb {R}}_x \rightarrow {\mathbb {R}}\):

$$\begin{aligned} \left\{ \begin{array}{l} \partial _t h = G(h) \phi \\ \\ \partial _t \phi = -h - \frac{1}{2} {|\phi _x|}^2 + \frac{1}{2(1+{|h_x|}^2)} {\left( G(h)\phi + h_x \phi _x \right) }^2 \end{array} \right. \end{aligned}$$
(1.2)

with

$$\begin{aligned} G(h) := \sqrt{1+{|h_x|}^2} {\mathcal {N}}(h) \end{aligned}$$
(1.3)

where \({\mathcal {N}}(h)\) is the Dirichlet–Neumann operator associated to the domain \(\Omega _t\). We refer to [40, chap. 11] or [14] for the derivation of the water wave equations (1.2).

Another possible description for (E)–(BC) can be given in Lagrangian coordinates again by deriving, in the case of irrotational flows (1.1), a system of equations on the boundary \(S_t\). More precisely, following [45], let \(z(t,{\alpha })\), for \({\alpha }\in {\mathbb {R}}\), be the equation of the free interface \(S_t\) at time \(t\) in Lagrangian coordinates, i.e. \(z_t(t,{\alpha }) = v (t, z(t,{\alpha }))\). Identifying \({\mathbb {R}}^2\) with the complex plane we use the same notation for a point \(z=(x,y)\) and its complex form \(z = x + i y\). We will then denote \(\overline{z} = x - i y \sim (x,-y)\). The divergence and curl free condition on the velocity \(v\) imply that \(\overline{v}\) is holomorphic in \(\Omega _t\). Therefore \(\overline{z}_t = {\mathcal {H} } _z \overline{z}_t\), where \( {\mathcal {H} } _{\gamma }\) denotes the Hilbert transformFootnote 1 along a curve \({\gamma }\):

$$\begin{aligned} ( {\mathcal {H} } _{\gamma }f )(t,{\alpha }) := \frac{1}{i \pi } \text{ p.v. } \int _{\mathbb {R}}\frac{ f(t,{\beta }) }{{\gamma }(t,{\alpha }) - {\gamma }(t,{\beta })} \, {\gamma }_{\beta }(t,{\beta }) \, d{\beta }. \end{aligned}$$
(1.4)

The vanishing of the pressure in (BC) implies that \(\nabla p\) is perpendicular to \(S_t\) and therefore \(- \nabla p = i a z_{\alpha }\), with \(a := - \frac{\partial P}{\partial n} \frac{1}{|z_{\alpha }|}\). Since \(z_{tt}(t,{\alpha }) = \left( v_t + v\cdot \nabla v \right) (t,z(t,{\alpha }))\), one see that (E)–(BC)–(1.1) in two dimensions are equivalent to

$$\begin{aligned} \left\{ \begin{array}{l} z_{tt} + i = i a z_{\alpha }\\ \overline{z}_t = {\mathcal {H} } _z \overline{z}_t . \end{array} \right. \end{aligned}$$
(1.5)

In [43] Wu was able to reduce (1.5) to a quasilinear system, and to exploit the weakly hyperbolic structure of the new system to obtain local-in-time existence of solutions in Sobolev spaces by energy methods. Earlier results for small initial data in two dimensions were proven in [34, 42] and [13]. In [44] Wu was also able to prove local existence for the three dimensional problem (two dimensional interface) [44]. Following the breakthrough of [43, 44], there has been considerable amount of work on the local well-posedness of (E)–(BC), also including other effects on the wave motion, such as surface tension on the interface or a finite bottom. We refer the reader to [1, 2, 4, 5, 8, 9, 11, 31, 32, 39] for some of the works on the local well-posedness of (E). Recently, blow-up solutions in the form of turning waves [7] and “splash” solutions [6, 12] have been constructed.

The question of existence of global-in-time solutions for small, smooth, and suitably localized data, has also received attention in recent years. In the case of one dimensional interfaces, the only work investigating the long time behavior of small gravity waves is that of Wu [45], who was able to show almost global existence of solutions for (1.5). To do this, the author proposed some new unknowns, which we denote here by \(F\), and a fully nonlinear change of coordinates, reducing (1.5) to a system of the form

$$\begin{aligned} \partial _t^2 F + i\partial _{\alpha }F = G \end{aligned}$$
(1.6)

where \(G\) are quasilinear nonlinearities of cubic and higher order with suitable structure. Thanks to the cubic nature of this new system Wu was then able to perform (almost optimal) energy estimates and obtain existence of solutions up to times of order \(e^{c/{\varepsilon }}\), where \({\varepsilon }\) is the size of the initial data.

On the other hand, in the case of two dimensional interfaces, Germain, Masmoudi and Shatah [19] and Wu [46] obtained global solutions. The result of [19] relied on the energy method of [38, 39] and on the space-time resonance method introduced in [18]. In [46] the author used instead a three dimensional version of the arguments of [45] to derive a set of equations similar to (1.6), perform weighted energy estimates on them, and obtain decay via \(L^2 - L^\infty \) type estimates. Recently, Germain, Masmoudi and Shatah [20] obtained global solutions in three dimensions for capillary waves, i.e. with surface tension on the interface and no gravitational force.

Here we are interested in the gravity water waves system (E)–(BC)–(1.1) in the case of one dimensional interfaces which are a perturbation of the flat one, and initial velocity potentials which are suitably small in an appropriate norm. We aim to prove the existence of global-in-time and pointwise decaying solutions, and determine their asymptotic behavior as \(t \rightarrow \infty \).

1.2 The main theorem

We define first our main spaces of functions. FixFootnote 2 \(N_0 = 10^4\) and define \(N_1 := N_0/2 + 4\). Let \(S=\frac{1}{2} t \partial _t + {\alpha }\partial _{\alpha }\) be the scaling vector field. Given a time interval \(I\) and a function \(f:I\times \mathbb {R}\rightarrow \mathbb {C}\) we define the norm

$$\begin{aligned} {\Vert f (t)\Vert }_{X_k} := {\Vert f (t)\Vert }_{H^k} + {\Vert S f(t) \Vert }_{H^{k/2}}. \end{aligned}$$
(1.7)

\(X_{N_0}\) will be the weighted energy-space for the solution, expressed in some appropriate modified Lagrangian coordinates, as well as in Eulerian coordinates.

Given the height function \(h\) and the velocity potential \(\phi \) in Eulerian coordinates, we define the \(Z^{\prime }\) norm

$$\begin{aligned} {\Vert (h(t),\phi (t)) \Vert }_{Z^{\prime }} := {\Vert h(t) \Vert }_{W^{N_1+4,\infty }} + {\Vert \Lambda \phi (t) \Vert }_{W^{N_1+4,\infty }} ,\quad \text{ with } \quad \Lambda := {|\partial _x|}^{\frac{1}{2}}. \end{aligned}$$
(1.8)

This is the decaying norm that we will estimate. Decay of this norm at the rate of \(t^{-1/2}\) will give us a small global solution to the water wave problem.

Finally, we define the space \(Z\) by the norm

$$\begin{aligned} {\Vert f(t) \Vert }_{Z} := \sup _{\xi \in {\mathbb {R}}} \left| \left( {|\xi |}^{\beta }+ |\xi |^{N_1 + 15} \right) \widehat{f}(\xi ,t) \right| \end{aligned}$$
(1.9)

where \(\beta = 1/100\), and

$$\begin{aligned} \widehat{f}(\xi ,t) := \int _{\mathbb {R}}e^{-i x\xi } f(t,x) \, dx \end{aligned}$$

is the partial Fourier transform in the spatial variable. This space plays a key role in obtaining decay of the \(Z^{\prime }\) norm of solutions, see the linear estimate (1.18). The \(Z\) norm also plays an important role in proving modified scattering of solutions in Eulerian coordinates.

The paper is concerned with the proof of the following Main Theorem:

Theorem 1.1

Let \(h_0(x) = h(0,x)\) be the initial height of the surface \(S_0\), and let \(\phi _0 (x) = \phi (0,x)\) be the restriction to \(S_0\) of the initial velocity potential. Assume that at the initial time one has

$$\begin{aligned}&{\Vert (h_0, \Lambda \phi _0) \Vert }_{H^{N_0+2}} + {\left\| x \partial _x \left( h_0, \Lambda \phi _0 \right) \right\| }_{H^{N_0/2+1}}+ {\Vert h_0 + i \Lambda \phi _0 \Vert }_{Z} \le {\varepsilon }_0 , \end{aligned}$$
(1.10a)

where \(Z\) is defined in (1.9). Moreover, for \(x \in \Omega _0\) let \(v_0(x) = v(0,x)\), where \(v\) is the irrotational and divergence free velocity field of the fluid, and assume that

$$\begin{aligned} {\Vert |x|\nabla v_0 \Vert }_{H^{N_0/2} (\Omega _0)}\le {\varepsilon }_0\,. \end{aligned}$$
(1.10b)
  1. (i)

    (Global existence) Then there exists \(\overline{{\varepsilon }}_0\) small enough, such that for any \({\varepsilon }_0 \le \overline{{\varepsilon }}_0\), the initial value problem associated to (1.2) admits a unique global solution with

    $$\begin{aligned} \sup _{t} \left[ (1+t)^{-p_0} {\Vert (h(t), \phi _x (t)) \Vert }_{X_{N_0}} +{\Vert h(t) + i\Lambda \phi (t) \Vert }_{H^{N_1 + 10}}\right. \\ \left. +\sqrt{1+t} {\Vert (h(t), \phi (t)) \Vert }_{Z^{\prime }} \right] \lesssim {\varepsilon }_0 , \end{aligned}$$

    where \(p_0 = 10^{-4}\).

  2. (ii)

    (Modified scattering) Let \(u(t) := h(t) + i \Lambda \phi (t)\), with \(\Lambda := {|\partial _x|}^{1/2}\). Define

    $$\begin{aligned} G(\xi ,t):= \frac{{|\xi |}^{4}}{\pi } \int _0^t|\widehat{u}(\xi ,s)|^2 \frac{ds}{s+1} , \qquad t\in [0,T] . \end{aligned}$$

    Then there is \(p_1 > 0\) such that

    $$\begin{aligned}&(1+t_1)^{p_1}\left\| (1+|\xi |)^{N_1} \left[ e^{iG(\xi ,t_2)}e^{it_2\Lambda (\xi )}\widehat{u}(\xi ,t_2)\right. \right. \nonumber \\&\quad \left. \left. -e^{iG(\xi ,t_1)}e^{it_1\Lambda (\xi )} \widehat{u}(\xi ,t_1)\right] \right\| _{L^2_\xi }\lesssim \varepsilon _0 , \end{aligned}$$
    (1.11)

    for any \(t_1\le t_2\in [0,T]\). In particular, there is \(w_\infty \in L^2((1+|\xi |)^{2N_1}d\xi )\) with the property that

    $$\begin{aligned} \sup _{t\in [0,\infty )}(1+t)^{p_1}{ \left\| (1+|\xi |)^{N_1} \big ( e^{iG(\xi ,t)} e^{it\Lambda (\xi )} \widehat{u}(\xi ,t) - w_\infty (\xi ) \big ) \right\| }_{L^2_\xi }\lesssim \varepsilon _0\,.\nonumber \\ \end{aligned}$$
    (1.12)

Remark 1.2

The first norm in (1.10a) ensures that our initial data is small and smooth in Sobolev spaces of high regularity. Notice that we are assuming that the vertical variation from equilibrium of the interface, given by the graph of \(h\), as well as half derivative of the velocity potential are small. This is consistent with the conserved energy (and Hamiltonian)

$$\begin{aligned} E_0(h,\phi ) := \frac{1}{2} \int \phi G(h) \phi \, dx + \frac{1}{2} \int h^2 \, dx\approx \Vert h+i\Lambda \phi \Vert _{L^2}^2, \end{aligned}$$

for solutions of (1.2). The second norm in (1.10a), properly evolved in time, gives some control of certain weighted norms of the solution.

The key new ingredient in (1.10a) is the smallness of \(h_0\) and \(\Lambda \phi _0\) in the \(Z\)-norm defined in (1.9). The \(Z\)-norm is the key new component of our global argument: it is the only strong norm of the solution that we can control uniformly in time, while the other energy-type norms are allowed to increase slowly in time. Furthermore, the \(Z\)-norm allows us to describe properly the modified scattering of the solution.

Finally, (1.10b) is similar to a condition imposed by Wu [45]. We use it as in the cited paper to guarantee that the energy functional on which energy estimates are based is small at time \(t=0\).

Remark 1.3

The solutions can also be defined on the time interval \((-\infty ,0]\), since the equations are time-reversible. The global solutions we construct in Theorem 1.1 appear to be the first smooth nontrivial global solutions of the gravity water waves system (E)–(BC) in 2 dimensions.

Remark 1.4

A more precise statement of modified scattering can be found in Lemma 6.1, in terms of certain modified Eulerian variables. Also, more precise bounds on the solution, both in terms of the Eulerian variables \((h,\phi )\) and the Lagrangian variable \(z\) can be found in Sect. 3, see (3.1)–(3.3).

1.3 Main ideas in the proof

If one is interested in the long-time existence of small smooth solutions to quasilinear dispersive and wave equations, such as (1.2) or (1.5), there are two main aspects one needs to consider: controlling high frequencies and proving dispersion. The first aspect is generally connected to the construction of high order energies which control the Sobolev norm of a solution. The second aspect is related to \(L^p\) decay estimates, and to estimates of weighted norms. When dealing with the water waves system both of these aspects are extremely delicate.

1.3.1 Supercriticality, energy estimates and normal forms

The general strategy for obtaining a global small solution usually starts with local-in-time energy estimates. The aim of the energy method is to construct an energy functional \(E(t)\) such that

$$\begin{aligned} E(t) \sim {\Vert u(t) \Vert }_{H^N}^2 \qquad \text{ and } \qquad \frac{d}{dt} E(t) \lesssim E(t)^{3/2} . \end{aligned}$$
(1.13)

Here, the power \(3/2\) is dictated by the quadratic nature of the nonlinearities in (1.2) or (1.5). The estimates (1.13) are often the key ingredient in obtaining local solutions, and for initial data of size \({\varepsilon }\) they give existence for times of order \(1/{\varepsilon }\). We remark here that the construction of an energy satisfying (1.13) for the water waves system is particularly challenging. Nevertheless it has been done in several works, such as the already cited [1, 5, 9, 31, 32, 39, 43, 44], thanks to considerable insight into the structure of the equations.

To extend a local solution for longer times one needs to engage the dispersive effects of the equation. One possibility is to try to upgrade (1.13) to

$$\begin{aligned} {\Vert u(t) \Vert }^2_{H^N} + {\Vert u (t) \Vert }^2_W \lesssim E(t) \qquad \text{ and } \qquad \frac{d}{dt} E(t) \lesssim \frac{{\varepsilon }}{t^a} \, E(t) , \end{aligned}$$
(1.14)

provided the solution decays like \(t^{-a}\) in some \(L^\infty \)-based space. The W-norm in (1.14) is supposed to encode some information about the localization of the flow. In the best case scenario a bound of the form \({\Vert u (t) \Vert }_W \lesssim 1\) implies the desired \(t^{-a}\) decay. Thus, if one can prove (1.14) with \(a > 1\) small solutions will exist globally and scatter to a linear solution. If \(a = 1\) solutions will automatically exist almost globally and further analysisFootnote 3 is needed in this critical case to show global existence and determine the asymptotic behavior. If \(a < 1\) the problem of global existence and scattering is much more difficult. This case is referred to as scattering supercritical and it is the case of the \(2d\) water waves problem, since solutions of the linear equation \(i \partial _t u - \Lambda u = 0\), \(\Lambda = {|\partial _x|}^{1/2}\), decay at the rate \(t^{-1/2}\).

In the work of Wu [45] on the almost global existence in the two dimensional problem, the author relied on a nonlinear version of a normal form transformation. Starting from the Lagrangian formulation (1.5), Wu proposed some new quantities, and a diffeomorphism depending fully nonlinearly on the solution, such that the set of equations obtained in the new coordinates admit a certain type of energy estimates, consistent with cubic nonlinearities. More precisely, let us denote, schematically, by \(F\) the new transformed unknowns, after the nonlinear change of variables. The point of Wu’s remarkable construction is that the variables \(F\) satisfy nonlinear equations like (1.6) with cubic-type nonlinearities. As a consequence, Wu defines a suitable energy functional \(E(t)\) and proves the following type of energy estimates:

$$\begin{aligned}&{\Vert F(t) \Vert }^2_{H^N} + {\Vert F(t) \Vert }^2_{W} \lesssim E (t) \nonumber \\&\quad \text{ and } \qquad \frac{d}{dt} E (t) \lesssim {\Vert F(t) \Vert }^2_{W^{\frac{N}{2} ,\infty }} \, E(t) \log t + \frac{1}{t} E^2 (t). \end{aligned}$$
(1.15)

These estimates can then be combined with \(L^2-L^\infty \) estimates, which show that \(F(t)\) decays pointwise like \(t^{-1/2}\). Therefore \(E(t) \lesssim {\varepsilon }^2\) as long as \(\log t \lesssim {\varepsilon }^{-1}\), which concludes the proof of almost global existence.

1.3.2 Dispersion and asymptotic behavior

To pass to global existence, our first concern is to improve (1.15) by eliminating logarithmic losses. We will use the same approach of [45], and the same equations derived there, to show an estimate of the form

$$\begin{aligned} {\Vert F(t) \Vert }^2_{H^N} + {\Vert F(t) \Vert }^2_{W} \lesssim E (t) \qquad \text{ and } \qquad \frac{d}{dt} E (t) \lesssim {\Vert F (t) \Vert }^2_{Z_\infty } \, E(t) , \end{aligned}$$
(1.16)

for some \(L^\infty \) based space \(Z_\infty \) which is stronger than \(W^{\frac{N}{2},\infty }\). Such an estimate is achieved by carefully analyzing the singular integrals (Calderón commutators) appearing in the cubic nonlinearities, and exploiting special structures present in some of them.

Another important point is that in order to justify the existence of the Wu’s diffeomorphism \(k\) and of the new unknowns \(F\) for all times \(t\), we need appropriate a priori control on Wu’s change of coordinates. This is obtained by taking advantage of a certain null structure present in the transformation.

Thanks to (1.16) one can guarantee \(E(t) \lesssim _{p_0} {\varepsilon }^2 t^{2p_0}\), for any fixed \(p_0>0\), and for all \(t \in [0,\infty )\), provided \({\Vert F(t) \Vert }_{Z_\infty } \lesssim {\varepsilon }t^{-1/2}\). However, since \(E(t)\) is forced to grow in time, although just slightly, one cannot obtain the desired sharp \(Z_\infty \) decay through energy estimates like those in [45].

Our main idea in this paper is to use also the Eulerian formulation of the equations (1.2) for the purpose of proving decay and modified scattering. More precisely, we achieve this by bootstrapping at the same time several bounds, which control the Eulerian variables, the Lagrangian variables, and Wu’s diffeomorphism \(k\) (see (2.8)–(2.10)). This consists of several steps:

  • We show first that the Eulerian variables \(h\) and \(\phi _x\) in (1.2) are controlled in the energy norms \(H^{N_0}\) and in \(S^{-1}H^{N_0/2}\) by the energy \(E(t)\) of the modified Lagrangian variables. Here \(S=(1/2)t\partial _t+x\partial _x\) is the scaling vector field. In other words we transfer the energy and weighted energy bounds (which are expected to increase slowly in time as a result of the bootstrap argument) to the Eulerian variables \(h,\phi _x\).

  • The variables \(h\) and \(\phi \) satisfy the equations (1.2), which can be written schematically in the form

    $$\begin{aligned}&\partial _t h = |\partial _x| \phi - \partial _x( h \partial _x \phi ) - |\partial _x| (h |\partial _x| \phi )+\mathrm {Cubic}(h,\phi _x);\\&\partial _t \phi = - h - (1/2){|\phi _x|}^2 + (1/2){||\partial _x|\phi |}^2 + \mathrm {Cubic}(h,\phi _x), \end{aligned}$$

    where the remainders are cubic expression of \(h,\phi _x\). We use a normal form transformation \(H=h+A(h,h)\), \(\Psi =\phi +B(h,\phi )\), for suitable bilinear operators \(A\) and \(B\) to eliminate the quadratic nonlinearities and reduce this to an evolution equation of a complex variable with a cubic nonlinearity. More precisely, letting \(V=H+i\Lambda \Psi \), we show that \(V\) satisfies an equation of the form

    $$\begin{aligned} \partial _t V+i\Lambda V = C(V,\overline{V}),\qquad \Lambda =|\partial _x|^{1/2}, \end{aligned}$$
    (1.17)

    where \(C\) is a nonlocal cubic quasilinear nonlinearity depending on all possible combination of \(V\) and \(\overline{V}\), and some of their derivatives. Moreover, we show that the variable \(V\) satisfies similar energy and weighted energy estimates as the functions \(h\) and \(\phi _x\), i.e. with a slow increase in time.

  • Finally, to analyzeFootnote 4 the equation (1.17) we use the key \(Z\) norm defined in (1.9). Unlike all the other energy and weighted energy norms, which are allowed to increase slowly at \(t\rightarrow \infty \), the \(Z\) norm of the solution is not based on \(L^2\)-type of spaces and is the only strong norm we are able to control uniformly in time. Letting \(f(t)=e^{it\Lambda }V(t)\) and using the Fourier transform method, we identify an appropriate nonlinear correction \(L=L(\widehat{f})\) and show that the function \(t\rightarrow \mathcal {F}^{-1}[e^{iL(\xi ,t)}\widehat{f}(\xi ,t)]\) converges, at a polynomial rate, as \(t\rightarrow \infty \). This suffices to prove global existence and modified scattering.

The crucial \(t^{-1/2}\) pointwise decay of the solution, which is needed to close the energy estimates, is then a consequence of the linear bound in Lemma 6.2,

$$\begin{aligned} {\Vert e^{i t\Lambda }h\Vert }_{L^\infty }&\lesssim (1\!+\!|t|)^{-1/2} {\Vert \,|\xi |^{3/4}\widehat{h}(\xi )\Vert }_{L^\infty _\xi }\!+\!(1+|t|)^{-5/8}\nonumber \\&\left[ {\Vert x\cdot \partial _x h\Vert }_{L^2} \!+\! {\Vert h\Vert }_{H^2} \right] . \end{aligned}$$
(1.18)

Notice that this pointwise bound requires sharp control of the \(Z\) norm, but can tolerate slow increase of the energy norms in time. This is consistent with the information we have on our solutions.

1.3.3 Plan of the paper

In Sect. 2 we describe in detail the strategy of our proof through Propositions 2.1–2.5. In Sect. 3 we prove the main Theorem 1.1 assuming these propositions. In Sect. 4 we state Propositions 4.1–4.4 and show how they imply the decay and the control of lower Sobolev norms stated in Proposition 2.5. Propositions 4.1–4.4 are then proved in Sects. 5 and 6. In Sect. 7 we describe the change of coordinates used by Wu [45], the cubic equations obtained there, and the associated energy functional. In Sect. 8 we prove the energy estimates contained in Proposition 2.3 via Propositions 8.1–8.3. We then prove Proposition 2.2. i.e. that the change of coordinates used is a diffeomorphism, on any time interval where one has a small solution satisfying certain a priori bounds. The transition of energy norms to Eulerian coordinates is done in Sect. 10, where we prove Proposition 2.4. In Appendix A we first give some variants of the estimates used in [45] that are compatible with our energy estimates. Section A2 contains some estimates for singular integral operators of “Calderón commutators” type that are used in the course of the energy estimates. In appendix B we calculate the resonant contribution of the cubic nonlinearities in Eulerian coordinates, after the application of the normal form. Appendix C contains estimates for the quartic and higher order remainders in the equation (1.2).

2 Strategy of the proof

The proof of Theorem 1.1 relies on a set of different Propositions. We state these key Propositions below and make some comments.

2.1 Local existence

Our strategy for controlling high Sobolev norms of solutions relies on the energy method of Wu [45], which is developed starting from the Lagrangian formulation of the problem. Therefore we begin by describing the local existence theory in Lagrangian coordinates. Assume that at the initial time the interface \(S_0\) is given by the graph of a function \(h_0:{\mathbb {R}}\rightarrow {\mathbb {R}}\), with \(h_0({\alpha }) \rightarrow 0\) as \(|{\alpha }| \rightarrow \infty \). Let \(z_0({\alpha }) = {\alpha }+ i h_0({\alpha })\) be a parametrization of \(S_0 \subset {\mathbb {C}}\). Assume that for some \(\mu > 0\)

$$\begin{aligned} |z_0({\alpha }) - z_0({\beta })| \ge \mu |{\alpha }-{\beta }| \qquad \forall \, {\alpha }, \, {\beta }\in {\mathbb {R}}. \end{aligned}$$
(2.1)

Let \(z = z(t,{\alpha })\) be the equation of the free surface \(S_t\) at time \(t\), in the Lagrangian coordinate \({\alpha }\), with \(z(0,{\alpha }) = z_0({\alpha })\). The following local existence result holds:

Proposition 2.1

(Local existence in Lagrangian coordinates [45, Theorem 5.1], [43]) Let \(N \ge 4\) be an integer. Assume that

$$\begin{aligned} \sum _{ 0 \le j \le N} {\left\| \partial _{\alpha }^j\left( z_{\alpha }(0) \!-\! 1, z_t(0), \partial _{\alpha }z_t(0) \right) \right\| }_{H^{1/2}} \!+\!{\left\| \left( \partial _{\alpha }^j (z_{tt}(0), \partial _{\alpha }z_{tt}(0) \right) \right\| }_{L^2} \!\le \!{\varepsilon }_0 . \end{aligned}$$
(2.2)

Then there exists a time \(T>0\), depending only on the norm of the initial data, such that the initial value problem for (1.5)Footnote 5 has unique solution \(z = z(t,{\alpha })\) for \(t \in [0,T]\), satisfying for all \(j \le N\)

$$\begin{aligned} \partial _{\alpha }^j \left( z_{\alpha }- 1, z_t, \partial _{\alpha }z_t \right)&\in C \left( [0,T],H^{1/2}({\mathbb {R}}) \right) \\ \partial _{\alpha }^j \left( z_{tt}, \partial _{\alpha }z_{tt} \right)&\in C \left( [0,T],L^2({\mathbb {R}}) \right) \end{aligned}$$

and \(|z(t,{\alpha }) - z(t,{\beta })| \ge \nu |{\alpha }-{\beta }|\), for all \({\alpha },{\beta }\in {\mathbb {R}}\), \(t \in [0,T]\), and some \(\nu > 0\).

Moreover, one has the following continuation criterion: if \(T^*\) is the supremum of all such times \(T\), then either \(T^*= \infty \) or

$$\begin{aligned}&\sup _{t\in [0,T^*)} \left( \sum _{ 0 \le j \le \left[ \frac{N+1}{2}\right] + 2}{\Vert \partial ^j_{\alpha }z_{tt}(t) \Vert }_{L^2} + {\Vert \partial ^j_{\alpha }z_t(t) \Vert }_{H^{1/2}}\right. \nonumber \\&\qquad \qquad \left. + \sup _{{\alpha }\ne {\beta }} \left| \frac{{\alpha }-{\beta }}{z(t,{\alpha }) - z(t,{\beta })} \right| \right) = \infty . \end{aligned}$$
(2.3)

Given \(N = N_0\), and a local solution on \([0,T]\), with initial data as in the main Theorem, we assume that

$$\begin{aligned} \sup _{[0,T]} \left( \sum _{ 0 \le j \le \frac{N_0}{2} + 2}{\Vert \partial ^j_{\alpha }z_{tt} (t) \Vert }_{L^2} + {\Vert \partial ^j_{\alpha }z_t (t) \Vert }_{H^{1/2}}+ {\Vert z_{\alpha }(t) - 1 \Vert }_{L^\infty } \right) \le {\varepsilon }_0^{3/4} . \end{aligned}$$
(2.4)

To obtain a global solution it suffices to prove that the quantity in the left-hand side of the inequality above is bounded by \(C {\varepsilon }_0\).

2.2 The main a priori assumptions

Under the a priori assumption (2.4), we have that \(\text{ Re }z\) is a diffeomorphism on \([0,T]\). We can then relate the Eulerian variables \(h\) and \(\phi \) to the Lagrangian map \({\alpha }\rightarrow z(t,{\alpha })\), for \(t\in [0,T]\), via the following identities

$$\begin{aligned} h (t, \text{ Re }z(t,{\alpha }) ) = \text{ Im }z(t,{\alpha }),\qquad \phi (t, \text{ Re }z(t,{\alpha }) ) = \psi (t,{\alpha }) \,. \end{aligned}$$
(2.5)

Here \(\psi (t,\alpha )\) is the trace of the velocity potential in Lagrangian coordinates, which can be obtained from the map \(z\), for example, using the Bernoulli equation \(\psi _t=-\text{ Im }z+(1/2)|z_t|^2\), \(\psi (0,\alpha )=\phi _0(\alpha )\).

In addition, we also need Wu’s change of coordinates \(k\) as in [45], to obtain cubic equations amenable to energy estimates. The explicit form for \(k\) is given in (7.2), and is the same as the one used by Totz and Wu [41], see formula (2.3) there. As long as this transformation \(k\) is a well-defined diffeomorphism, one can associate to the Lagrangian map \(z\) a modified Lagrangian map \({\zeta }= z \circ k^{-1}\), and the following vector associated to the new coordinates:

$$\begin{aligned} \widetilde{L}(t,{\alpha }) := \left( {\zeta }_{\alpha }(t,{\alpha }) - 1, u(t,{\alpha }), w(t,{\alpha }), \text{ Im }{\zeta }(t,{\alpha }) \right) , \end{aligned}$$
(2.6)

with

$$\begin{aligned} {\zeta }:= z \circ k^{-1} , \quad u := z_t \circ k^{-1} , \quad w := z_{tt} \circ k^{-1} . \end{aligned}$$
(2.7)

Our main bootstrap assumptions on the solution are:

$$\begin{aligned}&\sup _{t\in [0,T]} \left[ (1+t)^{-p_0} {\Vert (h(t), \phi _x (t)) \Vert }_{X_{N_0}} +{\Vert h(t) + i\Lambda \phi (t) \Vert }_{H^{N_1 + 10}}\right. \nonumber \\&\quad \left. +\sqrt{1+t} {\Vert (h(t), \phi (t)) \Vert }_{Z^{\prime }} \right] \le {\varepsilon }_1,\end{aligned}$$
(2.8)
$$\begin{aligned}&\sup _{t \in [0,T]} \left[ (1+t)^{-p_0} {\Vert \widetilde{L}(t) \Vert }_{X_{N_0}}+{\Vert \widetilde{L}(t) \Vert }_{H^{N_1+5}} + \sqrt{1+t} {\Vert \widetilde{L}(t) \Vert }_{W^{N_1,\infty }} \right] \le {\varepsilon }_1 ,\nonumber \\ \end{aligned}$$
(2.9)

and

$$\begin{aligned} \sup _{t\in [0,T]} {\Vert k_{\alpha }(t) - 1 \Vert }_{W^{N_0/2+3,\infty }} \le {\varepsilon }_1 , \end{aligned}$$
(2.10)

where \(X_{N_0}\) is defined by (1.7) and \(Z^{\prime }\) is defined in (1.8), \({\varepsilon }_1\le {\varepsilon }_0^{3/4}\), and \(p_0=10^{-4}\). In other words we assume a priori control on the Eulerian variables \((h,\phi )\), on the modified Lagrangian variable \(\zeta \), and on the diffeomorphism \(k\). To close the bootstrap argument we need to prove improved control on these quantities; this is the content of Propositions 2.2–2.5 below.

2.3 The main Propositions

Our first proposition, which is proved in Sect. 9, provides improved control on the diffeomorphism \(k\).

Proposition 2.2

(Control on the diffeomorphism \(k\)) Assume that (2.8)–(2.10) hold and, in addition,

$$\begin{aligned} {\Vert k_{\alpha }(0) - 1 \Vert }_{W^{N_0/2+3,\infty }} \lesssim {\varepsilon }_0 \,. \end{aligned}$$
(2.11)

Then

$$\begin{aligned} \sup _{t\in [0,T]} {\Vert k_{\alpha }(t) - 1 \Vert }_{W^{N_0/2+3,\infty }} \lesssim {\varepsilon }_0 +{\varepsilon }_1^2. \end{aligned}$$
(2.12)

The proof of the above Proposition relies crucially on the exploitation of a special null structure present in the transformation \(k\).

Our second proposition concerns improved control of the highest energy norm in modified Lagrangian variables. More precisely:

Proposition 2.3

(Energy estimates for the modified Lagrangian variables) Assume that (2.8)–(2.10) hold. Then

$$\begin{aligned} \sup _{t\in [0,T]} (1+t)^{-p_0}{\Vert \widetilde{L}(t) \Vert }_{X_{N_0}} \lesssim {\varepsilon }_0 + {\varepsilon }_1^2. \end{aligned}$$
(2.13)

This is proved in Sect. 8. We follow, to a large extent, the proof of Wu [45], using cubic equations for the “good unknowns” related to \(\widetilde{L}\), and performing energy estimates. However, some arguments need to be adjusted in order to avoid the logarithmic losses in the energy bound (compare with (1.15)).

The next step consists in translating the bounds given by the energy estimates in terms of the modified Lagrangian coordinates, to bounds on the norms of the Eulerian variables \(h\) and \(\partial _x \phi \).

Proposition 2.4

(Transition to Eulerian coordinates) Assume that (2.8)–(2.10) hold. Then, for \(t\in [0,T]\),

$$\begin{aligned}&{\Vert (h(t) , \partial _x \phi (t)) \Vert }_{X_{N_0}} \lesssim {\Vert \widetilde{L}(t) \Vert }_{X_{N_0}} , \end{aligned}$$
(2.14)

and

$$\begin{aligned}&{\Vert \widetilde{L}(t) \Vert }_{H^{N_1+5}} \lesssim {\Vert (h(t) , \partial _x \phi (t)) \Vert }_{H^{N_1+7}} + {\varepsilon }_1^2,\end{aligned}$$
(2.15a)
$$\begin{aligned}&{\Vert \widetilde{L}(t) \Vert }_{W^{N_1,\infty }} \lesssim {\Vert (h(t), \phi (t)) \Vert }_{Z^{\prime }} . \end{aligned}$$
(2.15b)

The proof of Proposition 2.4 is given in Sect. 10.

Finally, we use the Eulerian formulation of the equations to show the decay of the \(Z^{\prime }\)-norm and bound lower Sobolev norms:

Proposition 2.5

(Control of dispersive and lower Sobolev norms) Assume that (2.8)–(2.10) hold. Then

$$\begin{aligned} \sup _{[0,T]} \sqrt{1+t} {\Vert (h(t),\phi (t)) \Vert }_{Z^{\prime }} \lesssim {\varepsilon }_0 + {\varepsilon }_1^2 \end{aligned}$$
(2.16)

and

$$\begin{aligned} {\Vert h(t) + i\Lambda \phi (t) \Vert }_{H^{N_1 + 10}} \lesssim {\varepsilon }_0 +{\varepsilon }_1^2. \end{aligned}$$
(2.17)

The detailed strategy for the proof of Proposition 2.5 can be found in Sect. 4, and relies on Propositions 4.1–4.4, which are proven in Sects. 5 and 6.

Proposition 2.5 is the main new ingredient in our proof of global regularity. We start from the equations (1.2), written schematically in (4.5) and perform first a normal form transformation to eliminate the quadratic terms in the nonlinearity. We then analyze the resulting cubic equation, given in Lemma 5.1. A crucial step in the analysis consists in applying a phase correction to the transformed solution, and estimating it in the auxiliary \(Z\)-norm defined in (1.9).

3 Proof of the main Theorem

Using Propositions 2.2–2.5 we can now complete the proof of the main theorem. Set \({\varepsilon }_1={\varepsilon }_0^{3/4}\) and assume that (2.8)–(2.10) hold. We apply Propositions 2.2–2.5 to conclude that

$$\begin{aligned}&\sup _{t\in [0,T]} \left[ (1+t)^{-p_0} {\Vert (h(t), \phi _x (t)) \Vert }_{X_{N_0}} + {\Vert h(t) + i\Lambda \phi (t) \Vert }_{H^{N_1 + 10}}\right. \nonumber \\&\quad \left. +\sqrt{1+t} {\Vert (h(t), \phi (t)) \Vert }_{Z^{\prime }} \right] \lesssim {\varepsilon }_0 ,\end{aligned}$$
(3.1)
$$\begin{aligned}&\sup _{t \in [0,T]} \left[ (1+t)^{-p_0} {\Vert \widetilde{L}(t) \Vert }_{X_{N_0}}+{\Vert \widetilde{L}(t) \Vert }_{H^{N_1+5}} + \sqrt{1+t} {\Vert \widetilde{L}(t) \Vert }_{W^{N_1,\infty }} \right] \lesssim {\varepsilon }_0,\nonumber \\ \end{aligned}$$
(3.2)

and

$$\begin{aligned} \sup _{t\in [0,T]} {\Vert k_{\alpha }(t) - 1 \Vert }_{W^{N_0/2+3,\infty }} \lesssim {\varepsilon }_0. \end{aligned}$$
(3.3)

This provides the desired improvement of the a priori assumptions (2.8)–(2.10). By continuity, the improved bounds (3.1)–(3.3) hold on the time interval \([0,T]\).

We consider now the a priori Assumption 2.4. To improve it, we need to show that

$$\begin{aligned} \sup _{[0,T]} \left( \sum _{ 0 \le j \le \frac{N_0}{2} + 2} {\Vert \partial ^j_{\alpha }z_{tt} (t) \Vert }_{L^2} + {\Vert \partial ^j_{\alpha }z_t (t) \Vert }_{H^{1/2}} + {\Vert z_{\alpha }(t) - 1 \Vert }_{L^\infty } \right) \lesssim {\varepsilon }_0. \end{aligned}$$
(3.4)

This follows using the chain rule from the identities \(z=\zeta \circ k\), \(z_t=u\circ k\), and \(z_{tt}=w\circ k\), see (2.7), and the bounds (3.2) and (3.3). As a consequence of Proposition 2.1, the solution extends globally, and part (i) of the Main Theorem follows.

To prove the modified scattering statement in part (ii) we may assume that \(t_1\le t_2\le 2t_1\) and use Lemma 6.1, more precisely the bound (6.13)

$$\begin{aligned} \Vert (|\xi |^\beta +|\xi |^{N_1+15})(g(\xi ,t_2)-g(\xi ,t_1))\Vert _{L^\infty _\xi }\lesssim \varepsilon _0(1+t_1)^{-p_1}, \end{aligned}$$
(3.5)

where, see (6.10) and Lemma 5.1,

$$\begin{aligned} g(\xi ,t)&=e^{iL(\xi ,t)}\widehat{f}(\xi ,t),\quad L(\xi ,t):=\frac{\xi ^4}{\pi }\int _0^t|\widehat{f}(\xi ,s)|^2 \frac{ds}{s+1},\nonumber \\ f(t)&=e^{it\Lambda }V(t),\quad V=H+i\Lambda \Psi =[h+A(h,h)]+i\Lambda [\phi +B(h,\phi )]. \end{aligned}$$
(3.6)

With the notation in Theorem 1.1, and using also (5.26), we notice that, for \(t\in \{t_1,t_2\}\),

$$\begin{aligned}&\left\| (1+|\xi |)^{N_1} \left[ e^{iG(\xi ,t)}e^{it\Lambda (\xi )}\widehat{u}(\xi ,t)-e^{iG(\xi ,t)}\widehat{f}(\xi ,t)\right] \right\| _{L^2_\xi }\\&\quad \lesssim \left\| (1+|\xi |)^{N_1} \left[ \widehat{u}(\xi ,t)-\widehat{V}(\xi ,t)\right] \right\| _{L^2_\xi }\lesssim {\varepsilon }_0(1+t)^{-1/4}. \end{aligned}$$

Therefore, for (1.11) it suffices to prove that

$$\begin{aligned} \left\| (1+|\xi |)^{N_1} \left[ e^{iG(\xi ,t_2)}\widehat{f}(\xi ,t_2) -e^{iG(\xi ,t_1)}\widehat{f}(\xi ,t_1)\right] \right\| _{L^2_\xi }\lesssim \varepsilon _0(1+t_1)^{-p_1}.\nonumber \\ \end{aligned}$$
(3.7)

Notice that \(e^{iG(\xi ,t)}\widehat{f}(\xi ,t)=g(\xi ,t)e^{i[G(\xi ,t)-L(\xi ,t)]}\), \(t\in \{t_1,t_2\}\). Moreover

$$\begin{aligned}&\big |e^{i[G(\xi ,t_2)-L(\xi ,t_2)]}-e^{i[G(\xi ,t_1)-L(\xi ,t_1)]}\big |\lesssim \! \big |[G(\xi ,t_2)\!-\!L(\xi ,t_2)]\nonumber \\&\quad -[G(\xi ,t_1)\!-\!L(\xi ,t_1)]\big | \lesssim \xi ^4\int _{t_1}^{t_2}\big |\widehat{u}(\xi ,s)-\widehat{V}(\xi ,s)\big |(|\widehat{u}(\xi ,s)|\\&\quad +|\widehat{V}(\xi ,s)|)\frac{ds}{s+1}. \end{aligned}$$

In view of Proposition 4.4, we have, for any \(\xi \in \mathbb {R}\) and \(s\in [t_1,t_2]\),

$$\begin{aligned} \big (|\xi |^\beta +|\xi |^{N_1+15}\big )\big (|\widehat{f}(\xi ,s)|+|\widehat{V}(\xi ,s)|+|\widehat{g}(\xi ,s)|\big )\lesssim {\varepsilon }_0. \end{aligned}$$

Also, using Proposition 5.1 and the bounds \(\Vert h(s)\Vert _{H^{N_0}}+\Vert \phi _x(s)\Vert _{H^{N_0}}\lesssim {\varepsilon }_0(1+s)^{p_0}\), see (2.8), it follows that, for any \(\xi \in \mathbb {R}\) and \(s\in [t_1,t_2]\),

$$\begin{aligned} \big (1+|\xi |^{N_1+15}\big )\big (|\widehat{u}(\xi ,s)-\widehat{V}(\xi ,s)|\big )\lesssim {\varepsilon }_0(1+s)^{2p_0}. \end{aligned}$$

Therefore, using these three bounds and (3.5), the left-hand side of (3.7) is dominated by

$$\begin{aligned}&\left\| (1+|\xi |)^{N_1} \left[ g(\xi ,t_2)-g(\xi ,t_1)\right] \right\| _{L^2_\xi }\\&\qquad +\left\| (1+|\xi |)^{N_1} g(\xi ,t_2)\big |e^{i[G(\xi ,t_2)-L(\xi ,t_2)]}-e^{i[G(\xi ,t_1)-L(\xi ,t_1)]}\big |\right\| _{L^2_\xi }\\&\qquad \quad \lesssim \varepsilon _0(1+t_1)^{-p_1}+{\varepsilon }_0\sup _{s\in [t_1,t_2]}(1+s)^{2p_0}\big \Vert \widehat{u}(\xi ,s)-\widehat{V}(\xi ,s)\big \Vert _{L^2_\xi }\\&\quad \qquad \lesssim {\varepsilon }_0(1+t)^{-p_1}. \end{aligned}$$

This completes the proof of the desired bound (3.7) and of the main theorem.

4 Eulerian formulation and proof of Proposition 2.5

In this section we first recall the water waves equations in Eulerian coordinates. We then explain our strategy for the proof of Proposition 2.5. This will be obtained as a consequence of Propositions 4.1, 4.2, 4.3 and 4.4 below.

4.1 The equations in Eulerian coordinates

The system of equations in Eulerian coordinates is

$$\begin{aligned} \left\{ \begin{array}{l} \partial _t h = G(h) \phi \\ \\ \partial _t \phi = -h - \frac{1}{2} {|\phi _x|}^2 + \frac{1}{2(1+{|h_x|}^2)} {\left( G(h)\phi + h_x \phi _x \right) }^2, \end{array} \right. \end{aligned}$$
(4.1)

where

$$\begin{aligned} G(h) := \sqrt{1+{|h_x|}^2} {\mathcal {N}}(h) \end{aligned}$$
(4.2)

and \({\mathcal {N}}\) denotes the Dirichlet–Neumann operator associated to \(\Omega _t\).

Given a multilinear expression of \(h\) and \(\phi \)

$$\begin{aligned} F = F(h,\phi ) = \sum _{j\ge 1} F_j(h,\phi ), \end{aligned}$$

where \(F_j\) is an homogeneous polynomial of order \(j\) in its arguments, we denote

$$\begin{aligned} {[F]}_k := F_k (h,\phi ) \end{aligned}$$
(4.3)

and

$$\begin{aligned} {[F]}_{\ge k} := \sum _{j \ge k} {[F]}_j, \qquad {[F]}_{\le k} := \sum _{1 \le j \le k} {[F]}_j . \end{aligned}$$
(4.4)

After expanding \({\mathcal {N}}\) for small displacements of the moving surface, see [19, 40], one obtains the equations

$$\begin{aligned} \left\{ \begin{array}{rcl} \partial _t h &{} =&{} |\partial _x| \phi - \partial _x( h \partial _x \phi ) - |\partial _x| (h |\partial _x| \phi )\\ &{}&{} - \frac{1}{2} |\partial _x| \left[ h^2 |\partial _x|^2 \phi + |\partial _x| (h^2 |\partial _x| \phi )- 2 h |\partial _x| (h |\partial _x| \phi ) \right] + R_1(h,\phi ),\\ \\ \partial _t \phi &{} =&{} - h - \frac{1}{2} {|\phi _x|}^2 + \frac{1}{2} {||\partial _x|\phi |}^2 +|\partial _x|\phi \left[ h |\partial _x|^2 \phi - |\partial _x| (h |\partial _x| \phi ) \right] \\ &{}&{}+ R_2(h,\phi ), \end{array} \right. \end{aligned}$$
(4.5)

where:

$$\begin{aligned} R_1 (h,\phi )&:= {[G(h)\phi ]}_{\ge 4}\end{aligned}$$
(4.6)
$$\begin{aligned} R_2 (h,\phi )&:= {\left[ \frac{(G(h)\phi + h_x\phi _x)^2}{2 (1+{|h_x|}^2)} \right] }_{\ge 4} . \end{aligned}$$
(4.7)

Let us denote

$$\begin{aligned} M_2(h,\phi )&:= -\partial _x( h \partial _x \phi ) - {|\partial _x|}(h {|\partial _x|}\phi ),\end{aligned}$$
(4.8)
$$\begin{aligned} M_3(h,h,\phi )&:= - \frac{1}{2} {|\partial _x|}\left[ h^2 {|\partial _x|}^2 \phi + {|\partial _x|}(h^2 {|\partial _x|}\phi )- 2 h {|\partial _x|}(h {|\partial _x|}\phi ) \right] ,\end{aligned}$$
(4.9)
$$\begin{aligned} Q_2(\phi ,\phi )&:= - \frac{1}{2} {|\phi _x|}^2 + \frac{1}{2} {|{|\partial _x|}\phi |}^2,\end{aligned}$$
(4.10)
$$\begin{aligned} Q_3(\phi ,h,\phi )&:= {|\partial _x|}\phi \left[ h {|\partial _x|}^2 \phi - {|\partial _x|}(h {|\partial _x|}\phi ) \right] , \end{aligned}$$
(4.11)

so that

$$\begin{aligned} \left\{ \begin{array}{l} \partial _t h = {|\partial _x|}\phi + M_2(h,\phi ) + M_3(h,h,\phi ) + R_1(h,\phi ), \\ \\ \partial _t \phi = - h + Q_2(\phi ,\phi ) + Q_3(\phi ,h,\phi ) + R_2(h,\phi ). \end{array} \right. \end{aligned}$$
(4.12)

4.2 Strategy for the proof of Proposition 2.5

Recall that in Proposition 2.5 we are making the following a priori assumptions:

$$\begin{aligned} \sup _{t\in [0,T]} \left[ (1+t)^{-p_0} {\Vert (h(t) , \partial _x \phi (t)) \Vert }_{X_{N_0}} +\sqrt{1+t} {\Vert (h(t) , \phi (t)) \Vert }_{Z^{\prime }}\right] \le {\varepsilon }_1 \end{aligned}$$
(4.13)

and

$$\begin{aligned} \Vert h_0+i\Lambda \phi _0\Vert _{H^{N_0+1}}+\Vert x\partial _x(h_0+i\Lambda \phi _0)\Vert _{H^{N_0/2}}+\Vert h_0+i\Lambda \phi _0\Vert _{Z}\le {\varepsilon }_0, \end{aligned}$$
(4.14)

for some \({\varepsilon }_1\in [{\varepsilon }_0,1]\). We then aim to prove

$$\begin{aligned} \sup _{t\in [0,T]} \sqrt{1+t} {\Vert (h(t), \phi (t)) \Vert }_{Z^{\prime }} \lesssim {\varepsilon }_0 +{\varepsilon }_1^2 \, \end{aligned}$$
(4.15)

and

$$\begin{aligned} \sup _{t\in [0,T]}{\Vert h(t) + i\Lambda \phi (t) \Vert }_{H^{N_1 + 10}} \lesssim {\varepsilon }_0 + {\varepsilon }_1^2. \end{aligned}$$
(4.16)

To prove (4.15) and (4.16) the idea to transform the quadratic equations into cubic ones, and then apply the strategy of our previous paper [27] to the cubic equations. We will proceed through several steps. We first perform a bilinear normal form transformation:

Proposition 4.1

(Poincaré-Shatah Normal Form) There exist bilinear operators \(A\) and \(B\) such that if

$$\begin{aligned} \left\{ \begin{array}{l} H \mathop {=}\limits ^{def} h + A (h,h), \\ \\ \Psi \mathop {=}\limits ^{def} \phi + B(h,\phi ), \end{array} \right. \end{aligned}$$
(4.17)

then the function \(V\) defined by

$$\begin{aligned} V \mathop {=}\limits ^{def} H + i \Lambda \Psi \end{aligned}$$
(4.18)

satisfies

$$\begin{aligned} \partial _t V + i \Lambda V = C \left( h, |\partial _x| \phi \right) \end{aligned}$$
(4.19)

where \(C\) is a nonlinearity consisting of cubic and higher order terms.

A more precise statement of this Proposition, with the explicit form of \(A,B\), and \(C\) is given in Sect. 5, followed by its proof. In Sect. 5.3 we show the following bounds on the transformation:

Proposition 4.2

(Bounds for the transformation) Under the a priori assumptions (4.13) on \(h = h(t)\) and \(\phi = \phi (t)\), we have for any \(t \in [0,T]\)

$$\begin{aligned} {\Vert A(h,h) \Vert }_{W^{N_1+4,\infty }}+ {\Vert \Lambda B(h,\phi ) \Vert }_{W^{N_1+4,\infty }}&\lesssim {\varepsilon }_1^2(1+t)^{-1/2},\end{aligned}$$
(4.20)
$$\begin{aligned} {\Vert A(h,h) \Vert }_{H^{N_0-5}} + {\Vert \Lambda B(h,\phi )\Vert }_{H^{N_0-5}}&\lesssim {\varepsilon }_1^2,\end{aligned}$$
(4.21)
$$\begin{aligned} {\Vert SA(h,h) \Vert }_{H^{\frac{N_0}{2}-5}}+ {\Vert S \Lambda B(h,\phi )\Vert }_{H^{\frac{N_0}{2}-5}}&\lesssim {\varepsilon }_1^2. \end{aligned}$$
(4.22)

In particular we have

$$\begin{aligned} {\Vert (h, \phi ) \Vert }_{Z^{\prime }}&\lesssim {\Vert V \Vert }_{W^{N_1+4,\infty }} + {\varepsilon }_1^2(1+t)^{-1/2} ,\end{aligned}$$
(4.23)
$$\begin{aligned} {\Vert h+i\Lambda \phi \Vert }_{H^{N_1+10}}&\lesssim {\Vert V \Vert }_{H^{N_1+11}} + {\varepsilon }_1^2 , \end{aligned}$$
(4.24)

and

$$\begin{aligned} {\Vert V \Vert }_{W^{N_1+4,\infty }}&\lesssim {\varepsilon }_1(1+t)^{-1/2},\end{aligned}$$
(4.25)
$$\begin{aligned} {\Vert H+i\partial _x\Psi \Vert }_{H^{N_0 -6}}&\lesssim \varepsilon _1(1+t)^{p_0},\end{aligned}$$
(4.26)
$$\begin{aligned} {\Vert S H \Vert }_{H^{\frac{N_0}{2}-6}} + {\Vert S \partial _x \Psi \Vert }_{H^{\frac{N_0}{2}-6}}&\lesssim {\varepsilon }_1(1+t)^{p_0}. \end{aligned}$$
(4.27)

The above Proposition shows that the a priori smallness assumption (4.13) can be suitably transferred to \(V\). The next step is to improve (4.27) by using the equation (4.19) and the specific properties of the nonlinearity:

Proposition 4.3

(Improvement of the weighted bound on \(V\)) Let \(V\) be the function defined by (4.18) and satisfying (4.19). Then

$$\begin{aligned}&\sup _{t\in [0,T]} {(1+t)}^{-5p_0} \left[ {\Vert S V(t) \Vert }_{H^{N_0/2-20}}+{\Vert V(t) \Vert }_{H^{N_0/2-20}}\right] \lesssim {\varepsilon }_0+{\varepsilon }_1^2 . \end{aligned}$$
(4.28)

Furthermore, if we define the profile of \(V\) as

$$\begin{aligned} f(t,x) := \left( e^{it\Lambda } V(t) \right) (x) . \end{aligned}$$
(4.29)

we have

$$\begin{aligned}&\sup _{t\in [0,T]} {(1+t)}^{-5p_0} \left[ {\Vert x\partial _x f(t) \Vert }_{H^{N_0/2-20}}+{\Vert f(t) \Vert }_{H^{N_0/2-20}}\right] \lesssim {\varepsilon }_0+{\varepsilon }_1^2. \end{aligned}$$
(4.30)

This is proved in Sect. 5.4. The bound (4.28) improves the bound (4.27) by gaining half derivative for low frequencies on the estimate for \(\Psi \), at the expense of losing a small amount of decay and some derivatives. This gives us (4.30) and allows us to exploit the bounds obtained in our previous paper [27].

Using the bounds given by Proposition 4.2 we will work on the scalar cubic equation (4.19) with the aim of showing:

Proposition 4.4

(Bound on the \(Z\)-norm and decay of the \(Z^{\prime }\)-norm of \(V\)) Let \(V\) be defined as above, and satisfying the bounds (4.25), (4.26), (4.30) . Assume further that

$$\begin{aligned} \sup _{t\in [0,T]} {\Vert V(t) \Vert }_{Z} \le {\varepsilon }_1 , \end{aligned}$$
(4.31)

where \(Z\) is the norm defined in (1.9). Then

$$\begin{aligned} \sup _{t\in [0,T]} {\Vert V(t) \Vert }_{Z} \lesssim {\varepsilon }_0 + {\varepsilon }_1^3 . \end{aligned}$$
(4.32)

As a consequence, using also Lemma 6.2,

$$\begin{aligned}&\sup _{t\in [0,T]} \sqrt{1+t} {\Vert V(t) \Vert }_{W^{N_1+4,\infty }} \lesssim {\varepsilon }_0 + {\varepsilon }_1^2, \end{aligned}$$
(4.33)

and

$$\begin{aligned} \sup _{t\in [0,T]} {\Vert V(t)\Vert }_{H^{N_1 + 11}} \lesssim {\varepsilon }_0 + {\varepsilon }_1^3. \end{aligned}$$
(4.34)

The proof of (4.32) constitutes the heart of the proof for the decay in Eulerian coordinates, and is performed in Sect. 6, using a construction similar to our paper [27].

Using (4.31) and an estimate similar to (4.20), we can also obtain the following:

Corollary 4.5

Under the a priori assumptions (4.13) and (4.31) we have

$$\begin{aligned} \sup _{t\in [0,T]} {(1+t)}^{1/8} {\Vert \phi (t) \Vert }_{L^\infty } \lesssim {\varepsilon }_0 + {\varepsilon }_1^2 . \end{aligned}$$
(4.35)

This shows in particular the validity of the assumption (8.128) in Lemma 8.10.

Proof of Corollary 4.5

For any \(k\in \mathbb {Z}\) let \(P_k\) be the Littlewood–Paley projector defined after (5.18). We estimate, using (4.33) and (4.23)

$$\begin{aligned} {\Vert P_k\phi (t)\Vert }_{L^\infty } \lesssim ({\varepsilon }_0 + {\varepsilon }_1^2)2^{-k/2}(1+t)^{-1/2}. \end{aligned}$$

At the same time, using (4.32), (4.20), and the bound

$$\begin{aligned}&{\Vert P_k B(h(t),\phi (t)) \Vert }_{L^\infty } \lesssim 2^k {\Vert {\mathcal {F}}P_k B(h(t),\phi (t)) \Vert }_{L^\infty } \\&\quad \lesssim 2^k {\Vert h(t) \Vert }_{L^2} {\Vert |\partial _x| \phi (t) \Vert }_{L^2} \lesssim 2^k {\varepsilon }_1^2 , \end{aligned}$$

see the explicit from of the symbol \(b\) in (5.3), we can obtain

$$\begin{aligned} {\Vert P_k\phi (t)\Vert }_{L^\infty } \lesssim {\Vert P_k\Psi (t)\Vert }_{L^\infty } \!+\! {\Vert P_kB(h(t),\phi (t))\Vert }_{L^\infty } \lesssim 2^{k/3}({\varepsilon }_0 \!+\! {\varepsilon }_1^2)+2^{k} {\varepsilon }_1^2. \end{aligned}$$

The desired conclusion follows from these estimates by considering the cases \(2^k \le {(1+t)}^{-3/8}\) and \(2^k \ge {(1+t)}^{-3/8}\). \(\square \)

Observe that (4.33) together with (4.23) imply (4.15). The bound (4.34) together with (4.24) implies (4.16), thereby concluding the proof of Proposition 2.5.

5 Proof of Propositions 4.1, 4.2, and 4.3: normal forms

In this section we aim to transform the quadratic equations (4.5) into cubic ones using a normal form transformation, as in [19, 37]. The possibility of doing this relies on the vanishing of the symbols of the quadratic interaction on the time resonant set. We remark that the structure of the transformation here is very important because we only have information on \(\partial _x \phi \) and not on \(\phi \) or \(\Lambda \phi \). Therefore we want to find \(H\) and \(\Psi \) as in (4.17), with \(A\) and \(B\) depending nicely on \(h\) and \(\partial _x \phi \), and such that \(H+i\Lambda \Psi \) satisfies a cubic equation.

5.1 Solving the homological equation

Given a suitable symbol \(m:\mathbb {R}\times \mathbb {R}\rightarrow \mathbb {C}\) we define the associated bilinear operator \(M(f,g)\) by the formula

$$\begin{aligned} \mathcal {F}\left[ M(f,g)\right] (\xi )=\frac{1}{2\pi }\int _{\mathbb {R}}m(\xi ,\eta )\widehat{f}(\xi -\eta )\widehat{g}(\eta )\,d\eta . \end{aligned}$$
(5.1)

The following lemma gives the explicit form for the transformation in Proposition 4.1:

Lemma 5.1

Let

$$\begin{aligned} a(\xi ,\eta )&:=-\frac{|\xi |}{2}\frac{\eta }{|\eta |}\frac{\xi -\eta }{|\xi -\eta |},\end{aligned}$$
(5.2)
$$\begin{aligned} b(\xi ,\eta )&:=-|\eta |\frac{\xi -\eta }{|\xi -\eta |}\frac{\xi }{|\xi |}, \end{aligned}$$
(5.3)

and

$$\begin{aligned} m_2(\xi ,\eta )&= \xi \eta - |\xi ||\eta | ,\end{aligned}$$
(5.4)
$$\begin{aligned} q_2(\xi ,\eta )&= \frac{1}{2}(\xi -\eta ) \eta + \frac{1}{2}|\xi -\eta ||\eta |. \end{aligned}$$
(5.5)

Then, the function \(V\) defined as

$$\begin{aligned} V \mathop {=}\limits ^{def} H + i \Lambda \Psi =[h+A(h,h)]+i\Lambda [\phi +B(h,\phi )] \end{aligned}$$
(5.6)

satisfies

$$\begin{aligned} \partial _t V&+ i \Lambda V = {\mathcal {N}}_3 + {\mathcal {N}}_4 \end{aligned}$$
(5.7)

where \({\mathcal {N}}_3\), respectively \({\mathcal {N}}_4\), are cubic, respectively quartic and higher, order terms explicitly given by

$$\begin{aligned} {\mathcal {N}}_3&\mathop {=}\limits ^{def} M_3(h,h,\phi )+2A(M_2(h,\phi ),h)+ i \Lambda \big [ Q_3(\phi ,h,\phi ) \nonumber \\&\qquad + B(M_2(h,\phi ),\phi ) + B(h, Q_2(\phi ,\phi ))\big ],\end{aligned}$$
(5.8)
$$\begin{aligned} {\mathcal {N}}_4&\mathop {=}\limits ^{def} R_1(h,\phi )+2A(M_3(h,h,\phi )+R_1(h,\phi ),h)+ i \Lambda \left[ R_2(h,\phi )\right. \nonumber \\&\qquad +B(h,Q_3(\phi ,h,\phi )+\left. R_2(h,\phi ))+B(M_3(h,h,\phi )+R_1(h,\phi ),\phi )\right] . \end{aligned}$$
(5.9)

Proof of Lemma 5.1

Given equation 4.12 we look for a transformation of the form \((h, \phi ) \rightarrow (H, \Psi )\), with

$$\begin{aligned} \left\{ \begin{array}{l} H = h + A_1 (\phi ,\phi ) + A_2 (h,h), \\ \\ \Psi = \phi + B(h,\phi ), \end{array} \right. \end{aligned}$$
(5.10)

where \(A_1,A_2\) are symmetric bilinear forms and \(B\) is a bilinear form. Our goal is to eliminate the quadratic nonlinear expressions, i. e.

$$\begin{aligned} \left\{ \begin{array}{l} \partial _t H = {|\partial _x|}\Psi + \text{ cubic } \text{ terms } \\ \\ \partial _t \Psi = - H + \text{ cubic } \text{ terms }. \end{array} \right. \end{aligned}$$
(5.11)

Indeed, using (5.10) and (4.12), we have

$$\begin{aligned} \partial _t H-|\partial _x|\Psi =&-|\partial _x|\phi -|\partial _x|B(h,\phi )+\partial _t h+2A_1(\partial _t\phi ,\phi )+2A_2(\partial _th,h)\\ =&-|\partial _x|B(h,\phi )+M_2(h,\phi )+M_3(h,h,\phi )+R_1(h,\phi )\\&-2A_1(h,\phi )+2A_1(Q_2(\phi ,\phi ),\phi ) + 2A_1(Q_3(\phi ,h,\phi ),\phi )\\&+ 2A_1(R_2(h,\phi ),\phi ) +2A_2({|\partial _x|}\phi ,h) + 2A_2(M_2(h,\phi ),h)\\&+ 2A_2(M_3(h,h,\phi ),h) + 2A_2(R_1(h,\phi ),h), \end{aligned}$$

and

$$\begin{aligned}&\partial _t \Psi +H = h + A_1 (\phi ,\phi ) + A_2 (h,h)+\partial _t\phi +B(h,\partial _t\phi )+B(\partial _th,\phi )\\&\quad =+A_1 (\phi ,\phi ) + A_2 (h,h)+Q_2(\phi ,\phi ) + Q_3(\phi ,h,\phi ) + R_2(h,\phi )\\&\qquad -B(h,h)+B(h,Q_2(\phi ,\phi )) + B(h,Q_3(\phi ,h,\phi )) + B(h,R_2(h,\phi ))\\&\qquad +B({|\partial _x|}\phi ,\phi ) \!+\! B(M_2(h,\phi ),\phi ) \!+\! B(M_3(h,h,\phi ),\phi ) \!+\! B(R_1(h,\phi ),\phi ). \end{aligned}$$

The condition (5.11) is equivalent to

$$\begin{aligned} -|\partial _x|B(h,\phi )+M_2(h,\phi )-2A_1(h,\phi )+2A_2({|\partial _x|}\phi ,h)&=0,\nonumber \\ A_1 (\phi ,\phi )+Q_2(\phi ,\phi )+B({|\partial _x|}\phi ,\phi )&=0,\nonumber \\ A_2 (h,h)-B(h,h)&=0. \end{aligned}$$
(5.12)

Therefore one can define

$$\begin{aligned}&a_2(\xi ,\eta )=\frac{b(\xi ,\eta )+b(\xi ,\xi -\eta )}{2},\nonumber \\&a_1(\xi ,\eta )=-q_2(\xi ,\eta )-\frac{b(\xi ,\eta )|\xi -\eta |+b(\xi ,\xi -\eta )|\eta |}{2},\nonumber \\&b(\xi ,\eta )\nonumber \\&=\frac{-2(|\xi |\!+\!|\eta |\!-\!|\xi \!-\!\eta |)q_2(\xi ,\eta )\!+\!(|\eta |\!+\!|\xi \!-\!\eta |\!-\!|\xi |)m_2(\xi ,\eta )\!-\!2|\eta |m_2(\xi ,\xi \!-\!\eta )}{D(\xi ,\eta )},\nonumber \\ \end{aligned}$$
(5.13)

where

$$\begin{aligned} D(\xi ,\eta ) = -{|\xi |}^2 - {|\xi -\eta |}^2 - {|\eta |}^2 +2|\xi ||\xi -\eta | +2|\xi ||\eta | +2|\eta ||\xi -\eta |\, \end{aligned}$$

and the identities (5.12) are easily seen to be verified.

The formulas can be simplified in the one-dimensional situation. Indeed, we notice first that \(m(\xi ,\eta )=m(-\xi ,-\eta )\) for all \(m\in \{q_2,m_2,a_1,a_2,b\}\). Moreover, for \(\xi > 0\), we calculateFootnote 6 explicitly:

if \(\eta < 0\) then

$$\begin{aligned}&q_2(\xi ,\eta )=0,\quad m_2(\xi ,\eta )=-2|\xi ||\eta |,\quad m_2(\xi ,\xi -\eta )=0,\quad \\&D(\xi ,\eta )=4|\xi ||\eta |, \quad b(\xi ,\eta )=-|\eta |; \end{aligned}$$

if \(\eta \in (0,\xi )\) then

$$\begin{aligned}&q_2(\xi ,\eta )=|\xi -\eta ||\eta |,\quad m_2(\xi ,\eta )=0,\quad m_2(\xi ,\xi -\eta )=0,\quad \\&D(\xi ,\eta )=4|\xi -\eta ||\eta |,\quad b(\xi ,\eta )=-|\eta |; \end{aligned}$$

if \(\eta > \xi \) then

$$\begin{aligned}&q_2(\xi ,\eta )=0,\quad m_2(\xi ,\eta )=0,\quad m_2(\xi ,\xi -\eta )=-2|\xi ||\xi -\eta |,\\&D(\xi ,\eta )=4|\xi ||\xi -\eta |,\quad b(\xi ,\eta )=|\eta |. \end{aligned}$$

Using also the formulas in the first line of (5.13), we calculate, for \(\xi > 0\),

$$\begin{aligned} \text { if } \quad \eta < 0 \quad&\text { then }\quad a_1(\xi ,\eta )=0,\quad a_2(\xi ,\eta )=\xi /2;\\ \text { if } \quad \eta \in (0,\xi ) \quad&\text { then }\quad a_1(\xi ,\eta )=0,\quad a_2(\xi ,\eta )=-\xi /2;\\ \text { if } \quad \eta > \xi \quad&\text { then }\quad a_1(\xi ,\eta )=0,\quad a_2(\xi ,\eta )=\xi /2. \end{aligned}$$

In particular \(A_1=0\) and the desired formulas in the lemma follow. \(\square \)

5.2 Analysis of the symbols

We now want to study the behavior of the symbols that appear in Lemma 5.1. We will describe our multipliers in terms of a simple class of symbols \(\mathcal {S}^{\infty }\), which is defined by

$$\begin{aligned} \mathcal {S}^\infty \mathop {=}\limits ^{def}\{m:\mathbb {R}^2\!\rightarrow \! \mathbb {C}:\,m\text { continuous and }\Vert m\Vert _{\mathcal {S}^\infty }\!:=\! \Vert \mathcal {F}^{-1}(m)\Vert _{L^1}<\infty \}.\nonumber \\ \end{aligned}$$
(5.14)

Clearly, \(\mathcal {S}^\infty \hookrightarrow L^\infty (\mathbb {R}\times \mathbb {R})\). Moreover, \(\mathcal {S}^\infty \) symbols are compatible with Hölder-type bounds on bilinear operators. Our first lemma summarizes some simple properties of the \(\mathcal {S}^{\infty }\) symbols.

Lemma 5.2

  1. (i)

    If \(m,m'\in \mathcal {S}^\infty \) then \(m\cdot m'\in \mathcal {S}^\infty \) and

    $$\begin{aligned} \Vert m\cdot m'\Vert _{\mathcal {S}^\infty }\lesssim \Vert m\Vert _{\mathcal {S}^\infty }\Vert m'\Vert _{\mathcal {S}^\infty }. \end{aligned}$$
    (5.15)

    Moreover, if \(m\in \mathcal {S}^\infty \), \(A:\mathbb {R}^2\rightarrow \mathbb {R}^2\) is a linear transformation, \(v\in \mathbb {R}^2\), and \(m_{A,v}(\xi ,\eta ):=m(A(\xi ,\eta )+v)\) then

    $$\begin{aligned} \Vert m_{A,v}\Vert _{\mathcal {S}^\infty }=\Vert m\Vert _{\mathcal {S}^\infty }. \end{aligned}$$
    (5.16)
  2. (ii)

    Assume \(p,q,r\in [1,\infty ]\) satisfy \(1/p+1/q=1/r\), and \(m\in \mathcal {S}^\infty \). Then, for any \(f,g\in L^2(\mathbb {R})\),

    $$\begin{aligned} \Vert M(f,g)\Vert _{L^r}\lesssim \Vert m\Vert _{S^\infty }\Vert f\Vert _{L^p}\Vert g\Vert _{L^q}, \end{aligned}$$
    (5.17)

    where \(M\) is defined as in (5.1).

Proof of Lemma 5.2

Part (i) follows directly from the definition. To prove (ii) let

$$\begin{aligned} K(x,y):=(\mathcal {F}^{-1}m)(x,y)=\int _{\mathbb {R}\times \mathbb {R}}m(\xi ,\eta )e^{ix\cdot \xi }e^{iy\cdot \eta }\,d\xi d\eta . \end{aligned}$$

Then

$$\begin{aligned} M(f,g)(x)&=C\int _{\mathbb {R}^2}e^{ix\xi }m(\xi ,\eta )\widehat{f}(\xi -\eta )\widehat{g}(\eta )\,d\eta d\xi \\&=C\int _{\mathbb {R}^2}K(u,v)f(x-u)g(x-u-v)\,dudv, \end{aligned}$$

and the desired bound (5.17) follows. \(\square \)

We fix \(\varphi :\mathbb {R}\rightarrow [0,1]\) an even smooth function supported in \([-8/5,8/5]\) and equal to \(1\) in \([-5/4,5/4]\). Let

$$\begin{aligned} \varphi _k(x):=\varphi (x/2^k)-\varphi (x/2^{k-1}),\qquad k\in \mathbb {Z},\,x\in \mathbb {R}. \end{aligned}$$
(5.18)

Let \(P_k\) denote the operator defined by the Fourier multiplier \(\xi \rightarrow \varphi _k(\xi )\). Given any multiplier \(m:\mathbb {R}^2\rightarrow \mathbb {C}\) and any \(k,k_1,k_2\in \mathbb {Z}\) we define

$$\begin{aligned} m^{k,k_1,k_2}(\xi ,\eta ):=m(\xi ,\eta )\cdot \varphi _k(\xi )\varphi _{k_1}(\xi -\eta )\varphi _{k_2}(\eta ). \end{aligned}$$
(5.19)

Our next lemma, which is an easy consequence of the explicit formulas (5.2)–(5.5), describes our main multipliers \(m_2,q_2,a,b\) in terms of \(\mathcal {S}^\infty \) symbols.

Lemma 5.3

For any \(k,k_1,k_2\in \mathbb {Z}\) we have

$$\begin{aligned} \Vert m_2^{k,k_1,k_2}\Vert _{\mathcal {S}^\infty }+\Vert q_2^{k,k_1,k_2}\Vert _{\mathcal {S}^\infty }&\lesssim 2^k2^{\min (k_1,k_2)},\end{aligned}$$
(5.20)
$$\begin{aligned} \Vert a^{k,k_1,k_2}\Vert _{\mathcal {S}^\infty }&\lesssim 2^k, \end{aligned}$$
(5.21)

and

$$\begin{aligned} \Vert b^{k,k_1,k_2}\Vert _{\mathcal {S}^\infty }\lesssim 2^{k_2}. \end{aligned}$$
(5.22)

5.3 Proof of Proposition 4.2: bounds on the normal form

Recall that we are assuming (4.13) and we want to show the three estimates (4.20), (4.21) and (4.22) for the bilinear operators \(A,B\) defined through their symbols \(a,b\) in Lemma 5.1.

As a consequence of (4.13), we have the following bounds on \(h=h(t)\) and \(\phi =\phi (t)\), for any \(k\in \mathbb {Z}\):

$$\begin{aligned} \Vert P_k h\Vert _{L^2}+2^k\Vert P_k\phi \Vert _{L^2}&\lesssim {\varepsilon }_1(1+t)^{p_0}2^{-N_0k_+},\nonumber \\ \Vert P_k h\Vert _{L^\infty }+2^{k/2}\Vert P_k\phi \Vert _{L^\infty }&\lesssim {\varepsilon }_1(1+t)^{-1/2}2^{-(N_1+4)k_+},\\ \Vert P_k Sh\Vert _{L^2}+2^k\Vert P_k S\phi \Vert _{L^2}&\lesssim {\varepsilon }_1(1+t)^{p_0}2^{-N_0k_+/2},\nonumber \end{aligned}$$
(5.23)

where here, and from now on, we denote \(k_+ = \max (k,0)\).

For any \(k\in \mathbb {Z}\) let

$$\begin{aligned}&\mathcal {X}_k:=\mathcal {X}_k^1\cup \mathcal {X}_k^2,\nonumber \\&\mathcal {X}_k^1:=\{(k_1,k_2)\in \mathbb {Z}\times \mathbb {Z}:\min (k_1,k_2)\le k+4,\,|\max (k_1,k_2)-k|\le 4\},\\&\mathcal {X}_k^2:=\{(k_1,k_2)\in \mathbb {Z}\times \mathbb {Z}:\min (k_1,k_2)\ge k-4,\,|k_1-k_2|\le 4\}.\nonumber \end{aligned}$$
(5.24)

Also let

$$\begin{aligned} \begin{aligned}&\mathcal {X}_{k,s}:=\{(k_1,k_2)\in \mathcal {X}_k:2^{\min (k_1,k_2)}\le \min (2^{k-10},(1+t)^{-10})\},\\&\mathcal {X}_{k,l}:=\{(k_1,k_2)\in \mathcal {X}_k:2^{\min (k_1,k_2)}\ge \min (2^{k-10},(1+t)^{-10})\}. \end{aligned} \end{aligned}$$
(5.25)

To prove (4.20)–(4.22) we estimate for any \(k\in \mathbb {Z}\), using (5.23), (5.21)–(5.22), and Lemma 5.2 (ii),

$$\begin{aligned} \Vert P_kA(h,h)\Vert _{L^2}&\lesssim \sum _{(k_1,k_2)\in \mathcal {X}_k}\Vert P_kA(P_{k_1}h,P_{k_2}h)\Vert _{L^2}\\&\lesssim \sum _{k_1\le k+4,\,|k_2-k|\le 4}2^k\Vert P_{k_1}h\Vert _{L^\infty }\Vert P_{k_2}h\Vert _{L^2}\\&+\sum _{(k_1,k_2)\in \mathcal {X}_k^2}2^k\Vert P_{k_1}h\Vert _{L^\infty }\Vert P_{k_2}h\Vert _{L^2}\\&\lesssim {\varepsilon }_1^2(1+t)^{2p_0-1/2}2^{k/4}2^{-(N_0-3)k_+} \end{aligned}$$

and

$$\begin{aligned} \Vert P_k\Lambda B(h,\phi )\Vert _{L^2}&\lesssim \sum _{(k_1,k_2)\in \mathcal {X}_k}2^{k/2}\Vert P_kB(P_{k_1}h,P_{k_2}\phi )\Vert _{L^2}\\&\lesssim \sum _{k_1\le k+4,\,|k_2-k|\le 4}2^{k/2}2^{k_2}\Vert P_{k_1}h\Vert _{L^\infty }\Vert P_{k_2}\phi \Vert _{L^2}\\&\quad +\sum _{k_2\le k+4,\,|k_1-k|\le 4}2^{k/2}2^{k_2}\Vert P_{k_1}h\Vert _{L^2}\Vert P_{k_2}\phi \Vert _{L^\infty }\\&\quad +\sum _{(k_1,k_2)\in \mathcal {X}_k^2}2^k2^{k_1}\Vert P_{k_1}h\Vert _{L^\infty }\Vert P_{k_2}\phi \Vert _{L^2}\\&\lesssim {\varepsilon }_1^2(1+t)^{2p_0-1/2}2^{k/4}2^{-(N_0-3)k_+}. \end{aligned}$$

Therefore, for any \(k\in \mathbb {Z}\),

$$\begin{aligned} \Vert P_kA(h,h)\Vert _{L^2}+\Vert P_k\Lambda B(h,\phi )\Vert _{L^2}\lesssim {\varepsilon }_1^2(1+t)^{2p_0-1/2}2^{k/4}2^{-(N_0-3)k_+},\nonumber \\ \end{aligned}$$
(5.26)

and the desired bound (4.21) follows.

Similarly, we also have the \(L^\infty \) bounds,

$$\begin{aligned} \Vert P_kA(h,h)\Vert _{L^\infty }&\lesssim \sum _{(k_1,k_2)\in \mathcal {X}_k}\Vert P_kA(P_{k_1}h,P_{k_2}h)\Vert _{L^\infty }\\&\lesssim \sum _{(k_1,k_2)\in \mathcal {X}_k}2^k\Vert P_{k_1}h\Vert _{L^\infty }\Vert P_{k_2}h\Vert _{L^\infty }\\&\lesssim {\varepsilon }_1^2(1+t)^{p_0-1}2^{k/4}2^{-N_0k_+/2} \end{aligned}$$

and

$$\begin{aligned} \Vert P_k\Lambda B(h,\phi )\Vert _{L^\infty }&\lesssim \sum _{(k_1,k_2)\in \mathcal {X}_k}2^{k/2}\Vert P_kB(P_{k_1}h,P_{k_2}\phi )\Vert _{L^\infty }\\&\lesssim \sum _{(k_1,k_2)\in \mathcal {X}_k}2^{k/2}2^{k_2}\Vert P_{k_1}h\Vert _{L^\infty }\Vert P_{k_2}\phi \Vert _{L^\infty }\\&\lesssim {\varepsilon }_1^2(1+t)^{p_0-1}2^{k/4}2^{-N_0k_+/2}. \end{aligned}$$

Therefore, for any \(k\in \mathbb {Z}\),

$$\begin{aligned} \Vert P_kA(h,h)\Vert _{L^\infty }+\Vert P_k\Lambda B(h,\phi )\Vert _{L^\infty }\lesssim {\varepsilon }_1^2(1+t)^{p_0-1}2^{k/4}2^{-N_0k_+/2},\nonumber \\ \end{aligned}$$
(5.27)

and the desired bound (4.20) follows.

To prove (4.22) we notice first that the symbol \(a\) is homogeneous of degree \(1\), i. e.

$$\begin{aligned} a(\lambda \xi ,\lambda \eta )=\lambda a(\xi ,\eta )\qquad \text { for any }\xi ,\eta \in \mathbb {R},\lambda \in (0,\infty ). \end{aligned}$$

Differentiating this identity with respect to \(\lambda \) and then setting \(\lambda =1\), we have

$$\begin{aligned} (\xi \partial _\xi a)(\xi ,\eta )+(\eta \partial _\eta a)(\xi ,\eta )=a(\xi ,\eta ). \end{aligned}$$

The symbol to \((\xi ,\eta )\rightarrow b(\xi ,\eta )\) is homogeneous of degree \(1\). As a consequence, we have the identities

$$\begin{aligned}&(\xi \partial _\xi a)(\xi ,\eta )+(\eta \partial _\eta a)(\xi ,\eta )=a(\xi ,\eta ),\nonumber \\&(\xi \partial _\xi b)(\xi ,\eta )+(\eta \partial _\eta b)(\xi ,\eta )=b(\xi ,\eta ). \end{aligned}$$
(5.28)

Using the first formula in (5.28) we calculate

$$\begin{aligned}&\mathcal {F}\left[ SA(h,h)\right] (\xi )\\&\quad =\Big [\frac{1}{2}t\partial _t-\xi \partial _\xi -I\Big ]\Big [\int _{\mathbb {R}}a(\xi ,\eta )\widehat{h}(\xi -\eta ,t)\widehat{h}(\eta ,t)\,d\eta \Big ]\\&\quad =\int _{\mathbb {R}}a(\xi ,\eta )\Big [\frac{1}{2}t(\partial _t\widehat{h})-\widehat{h}\Big ](\xi -\eta ,t)\widehat{h}(\eta ,t)\,d\eta \\&\qquad +\int _{\mathbb {R}}a(\xi ,\eta )\widehat{h}(\xi -\eta ,t)\frac{1}{2}t(\partial _t\widehat{h})(\eta ,t)\,d\eta \\&\quad \quad -\int _{\mathbb {R}}(\xi \partial _\xi a)(\xi ,\eta )\widehat{h}(\xi -\eta ,t)\widehat{h}(\eta ,t)\,d\eta \\&\qquad -\int _{\mathbb {R}}a(\xi ,\eta )\xi (\partial \widehat{h})(\xi -\eta ,t)\widehat{h}(\eta ,t)\,d\eta \\&\quad =\int _{\mathbb {R}}a(\xi ,\eta )\widehat{Sh}(\xi -\eta ,t)\widehat{h}(\eta ,t)\,d\eta -\int _{\mathbb {R}}a(\xi ,\eta )\eta (\partial \widehat{h})(\xi -\eta ,t)\widehat{h}(\eta ,t)\,d\eta \\&\quad \quad +\int _{\mathbb {R}}a(\xi ,\eta )\widehat{h}(\xi -\eta ,t)\widehat{Sh}(\eta ,t)\,d\eta \\&\qquad +\int _{\mathbb {R}}a(\xi ,\eta )\widehat{h}(\xi -\eta ,t)[\eta (\partial \widehat{h})(\eta ,t)+\widehat{h}(\eta ,t)]\,d\eta \\&\quad \quad +\int _{\mathbb {R}}[(\eta \partial _\eta a)(\xi ,\eta )-a(\xi ,\eta )]\widehat{h}(\xi -\eta ,t)\widehat{h}(\eta ,t)\,d\eta \\&\quad =\mathcal {F}\left[ A(Sh,h)\right] (\xi )+\mathcal {F}\left[ A(h,Sh)\right] (\xi )-\mathcal {F}\left[ A(h,h)\right] (\xi ). \end{aligned}$$

A similar calculation can be applied to the operator \(B\), using also (5.28). Therefore

$$\begin{aligned} \begin{aligned} SA(h,h)&=A(Sh,h)+A(h,Sh)-A(h,h),\\ SB(h,\phi )&=B(Sh,\phi )+B(h,S\phi )-B(h,\phi ). \end{aligned} \end{aligned}$$
(5.29)

For any \(k\in \mathbb {Z}\) we estimate, using (5.23), Lemma 5.2 (ii), and (5.21)–(5.22), and recalling (5.25),

$$\begin{aligned} \Vert P_kA(Sh,h)\Vert _{L^2}&\lesssim \sum _{(k_1,k_2)\in \mathcal {X}_k}\Vert P_kA_2(P_{k_1}Sh,P_{k_2}h)\Vert _{L^2}\\&\lesssim \sum _{(k_1,k_2)\in \mathcal {X}_{k,l}}2^k\Vert P_{k_1}Sh\Vert _{L^2}\Vert P_{k_2}h\Vert _{L^\infty }\\&\quad +\sum _{(k_1,k_2)\in \mathcal {X}_{k,s}}2^k2^{\min (k_1,k_2)/2}\Vert P_{k_1}Sh\Vert _{L^2}\Vert P_{k_2}h\Vert _{L^2}\\&\lesssim {\varepsilon }_1^2(1+t)^{2p_0-1/2}2^{k/4}2^{-(N_0/2-3)k_+}, \end{aligned}$$

and

$$\begin{aligned}&\Vert P_k\Lambda B(Sh,\phi )\Vert _{L^2}+\Vert P_k\Lambda B(h,S\phi )\Vert _{L^2}\\&\quad \lesssim 2^{k/2}\sum _{(k_1,k_2)\in \mathcal {X}_k}\left[ \Vert P_kB(P_{k_1}Sh,P_{k_2}\phi )\Vert _{L^2}+\Vert P_kB(P_{k_1}h,P_{k_2}S\phi )\Vert _{L^2}\right] \\&\quad \lesssim 2^{k/2}\sum _{(k_1,k_2)\in \mathcal {X}_{k,l}}\left[ 2^{k_2}\Vert P_{k_1}Sh\Vert _{L^2}\Vert P_{k_2}\phi \Vert _{L^\infty }+2^{k_2}\Vert P_{k_1}h\Vert _{L^\infty }\Vert P_{k_2}S\phi \Vert _{L^2}\right] \\&\qquad +2^{k/2}\sum _{(k_1,k_2)\in \mathcal {X}_{k,s}}2^{\min (k_1,k_2)/2}\\&\qquad \times \left[ 2^{k_2}\Vert P_{k_1}Sh\Vert _{L^2}\Vert P_{k_2}\phi \Vert _{L^2}+2^{k_2}\Vert P_{k_1}h\Vert _{L^2}\Vert P_{k_2}S\phi \Vert _{L^2}\right] \\&\quad \lesssim {\varepsilon }_1^2(1+t)^{2p_0-1/2}2^{k/4}2^{-(N_0/2-3)k_+}. \end{aligned}$$

Therefore, using also (5.29) and (5.26), for any \(k\in \mathbb {Z}\),

$$\begin{aligned} \Vert P_kSA(h,h)\Vert _{L^2}\!+\! \Vert P_kS\Lambda B(h,\phi )\Vert _{L^2}\!\lesssim \!{\varepsilon }_1^2(1+t)^{2p_0-1/2}2^{k/4}2^{-(N_0/2-3)k_+},\nonumber \\ \end{aligned}$$
(5.30)

and the desired bound (4.22) follows.

5.4 Proof of Proposition 4.3

We start from the formula (5.7),

$$\begin{aligned} \partial _tV+i\Lambda V=\mathcal {N}_3+\mathcal {N}_4, \end{aligned}$$

where \(\mathcal {N}_3\) and \(\mathcal {N}_4\) are given in (5.8) and (5.9). Applying \(S\) and commuting we derive the equation

$$\begin{aligned} (\partial _t+i\Lambda ) SV=S\mathcal {N}_3+S\mathcal {N}_4+(1/2)(\mathcal {N}_3+\mathcal {N}_4). \end{aligned}$$

Moreover, with \(f(t)=e^{it\Lambda }V(t)\), we have

$$\begin{aligned} (x\partial _xf)(t)=e^{it\Lambda }\left[ SV-(t/2)(\mathcal {N}_3+\mathcal {N}_4)\right] . \end{aligned}$$

It follows from the assumption (4.14) and Proposition 4.2 that

$$\begin{aligned} \Vert V(0)\Vert _{H^{N_0/2-5}}+\Vert SV(0)\Vert _{H^{N_0/2-5}}\lesssim {\varepsilon }_0+{\varepsilon }_1^2. \end{aligned}$$

Therefore, it suffices to prove the following:

Lemma 5.4

For any \(t\),

$$\begin{aligned} \Vert \mathcal {N}_3\Vert _{H^{N_0-20}}+\Vert S\mathcal {N}_3\Vert _{H^{N_0/2-20}}\lesssim {\varepsilon }_1^3(1+t)^{5p_0-1}, \end{aligned}$$
(5.31)

and

$$\begin{aligned} \Vert \mathcal {N}_4\Vert _{H^{N_0-20}}+\Vert S\mathcal {N}_4\Vert _{H^{N_0/2-20}}\lesssim {\varepsilon }_1^3(1+t)^{5p_0-1}. \end{aligned}$$
(5.32)

Proof of Lemma 5.4

As in the proof of Proposition 4.2, see (5.26), (5.27), and (5.30), for any \(k\in \mathbb {Z}\) we have

$$\begin{aligned} \Vert P_kQ_2(\phi ,\phi )\Vert _{L^2}+\Vert P_kM_2(h,\phi )\Vert _{L^2}&\lesssim {\varepsilon }_1^2(1+t)^{2p_0-1/2}2^{k/4}2^{-(N_0-3)k_+},\nonumber \\ \Vert P_kSQ_2(\phi ,\phi )\Vert _{L^2}\!+\!\Vert P_kSM_2(h,\phi )\Vert _{L^2}&\lesssim {\varepsilon }_1^2(1+t)^{2p_0-1/2} 2^{k/4}2^{-(N_0/2-3)k_+},\\ \Vert P_kQ_2(\phi ,\phi )\Vert _{L^\infty }+\Vert P_kM_2(h,\phi )\Vert _{L^\infty }&\lesssim {\varepsilon }_1^2(1+t)^{p_0-1}2^{k/4}2^{-(N_0/2-3)k_+}.\nonumber \end{aligned}$$
(5.33)

We examine now the trilinear expressions \(M_3(h,h,\phi )\) and \(Q_3(\phi ,h,\phi )\) in (4.9) and (4.11). These expressions appear in both nonlinearities \(\mathcal {N}_3\) and \(\mathcal {N}_4\). To estimate them we start by estimating \(h|\partial _x|\phi \) and \(h|\partial _x|^2\phi \): using (5.23), for any \(k\in \mathbb {Z}\), we obtain as before

$$\begin{aligned} \Vert P_k(h|\partial _x|\phi )\Vert _{L^2}+\Vert P_k(h|\partial _x|^2\phi )\Vert _{L^2}&\lesssim {\varepsilon }_1^2(1+t)^{2p_0-1/2}2^{-(N_0-3)k_+},\nonumber \\ \Vert P_kS(h|\partial _x|\phi )\Vert _{L^2}+\Vert P_kS(h|\partial _x|^2\phi )\Vert _{L^2}&\lesssim {\varepsilon }_1^2(1+t)^{2p_0-1/2}2^{-(N_0/2-3)k_+},\\ \Vert P_k(h|\partial _x|\phi )\Vert _{L^\infty }+\Vert P_k(h|\partial _x|^2\phi )\Vert _{L^\infty }&\lesssim {\varepsilon }_1^2(1+t)^{p_0-1}2^{-(N_0/2-3)k_+}.\nonumber \end{aligned}$$
(5.34)

We examine the formulas (4.9) and (4.11). For any \(k\in \mathbb {Z}\) we use (5.23) and (5.34) and estimate as before, for any \(k\in \mathbb {Z}\),

$$\begin{aligned} 2^{-k/4}\Vert P_kM_3(h,h,\phi )\Vert _{L^2}+\Vert P_kQ_3(\phi ,h,\phi )\Vert _{L^2}&\lesssim {\varepsilon }_1^3(1+t)^{3p_0-1}2^{-(N_0-6)k_+},\nonumber \\ 2^{-k/4}\Vert P_kSM_3(h,h,\phi )\Vert _{L^2}+\Vert P_kSQ_3(\phi ,h,\phi )\Vert _{L^2}&\lesssim {\varepsilon }_1^3(1+t)^{3p_0-1}2^{-(N_0/2-6)k_+},\\ 2^{-k/4}\Vert P_kM_3(h,h,\phi )\Vert _{L^\infty }+\Vert P_kQ_3(\phi ,h,\phi )\Vert _{L^\infty }&\lesssim {\varepsilon }_1^3(1+t)^{3p_0-3/2}2^{-(N_0/2-6)k_+}.\nonumber \end{aligned}$$
(5.35)

Recall the formulas (5.8) and (5.9),

$$\begin{aligned} {\mathcal {N}}_3&= M_3(h,h,\phi )+2A(M_2(h,\phi ),h)+ i \Lambda \left[ Q_3(\phi ,h,\phi )\right. \\&\quad \left. + B(M_2(h,\phi ),\phi ) + B(h, Q_2(\phi ,\phi ))\right] ,\\ {\mathcal {N}}_4&= R_1(h,\phi ) +2A(M_3(h,h,\phi )+R_1(h,\phi ),h)+ i \Lambda \left[ R_2(h,\phi )\right. \\&\quad +B(h,Q_3(\phi ,h,\phi )+\left. R_2(h,\phi ))+B(M_3(h,h,\phi )+R_1(h,\phi ),\phi )\right] , \end{aligned}$$

and the bounds (13.10)

$$\begin{aligned}&\Vert R_1(h,\phi )+i\Lambda R_2(h,\phi )\Vert _{H^{N_0-10}}+\Vert S\big (R_1(h,\phi )+i\Lambda R_2(h,\phi )\big )\Vert _{H^{N_0/2-10}}\\&\quad \lesssim {\varepsilon }_1^4(1+t)^{-5/4}. \end{aligned}$$

The desired bounds (5.31) and (5.32) follow using (5.33), (5.35), and Lemma 5.5 below, with \(G=R_2(h,\phi )\) and

$$\begin{aligned} F\in \{Q_2(\phi ,\phi ),M_2(h,\phi ),M_3(h,h,\phi ),Q_3(\phi ,h,\phi ),R_1(h,\phi )\}. \end{aligned}$$

\(\square \)

Lemma 5.5

Assume \(F\) and \(G\) satisfy the bounds, for any \(k\in \mathbb {Z}\),

$$\begin{aligned} \Vert P_kF\Vert _{L^2}+2^{k/2}\Vert P_kG\Vert _{L^2}&\lesssim {\varepsilon }_1^2(1+t)^{3p_0-1/2}2^{-(N_0-12)k_+},\nonumber \\ \Vert P_kSF\Vert _{L^2}+2^{k/2}\Vert P_kSG\Vert _{L^2}&\lesssim {\varepsilon }_1^2(1+t)^{3p_0-1/2}2^{-(N_0/2-12)k_+},\\ \Vert P_kF\Vert _{L^\infty }+2^{k/2}\Vert P_kG\Vert _{L^\infty }&\lesssim {\varepsilon }_1^2(1+t)^{3p_0-1}2^{-(N_0/2-12)k_+}.\nonumber \end{aligned}$$
(5.36)

Then, for any \(k\in \mathbb {Z}\),

$$\begin{aligned} \begin{aligned}&\Vert P_kA(F,h)\Vert _{L^2}+2^{k/2}\Vert P_kB(F,\phi )\Vert _{L^2}+2^{k/2}\Vert P_kB(h,G)\Vert _{L^2}\\&\quad \lesssim {\varepsilon }_1^3(1+t)^{5p_0-1}2^{k/4}2^{-(N_0-16)k_+},\\&\Vert P_kSA(F,h)\Vert _{L^2}+2^{k/2}\Vert P_kSB(F,\phi )\Vert _{L^2}+2^{k/2}\Vert P_kSB(h,G)\Vert _{L^2}\\&\quad \lesssim {\varepsilon }_1^3(1+t)^{5p_0-1}2^{k/4}2^{-(N_0/2-16)k_+}. \end{aligned} \end{aligned}$$
(5.37)

Proof of Lemma 5.5

We estimate, using (5.23), (5.36), Lemma 5.3, and Lemma 5.2 (ii),

$$\begin{aligned} \Vert P_kA(F,h)\Vert _{L^2}&\lesssim \sum _{k_1\le k+4,\,|k_2-k|\le 4}2^k\Vert P_{k_1}F\Vert _{L^\infty }\Vert P_{k_2}h\Vert _{L^2}\\&\quad +\sum _{k_2\le k+4,\,|k_1-k|\le 4}2^k\Vert P_{k_1}F\Vert _{L^2}\Vert P_{k_2}h\Vert _{L^\infty }\\&\quad +\sum _{(k_1,k_2)\in \mathcal {X}_k^2}2^k\Vert P_{k_1}F\Vert _{L^\infty }\Vert P_{k_2}h\Vert _{L^2}\\&\lesssim {\varepsilon }_1^3(1+t)^{5p_0-1}2^{k/4}2^{-(N_0-16)k_+} \end{aligned}$$

and

$$\begin{aligned}&\Vert P_kA(F,Sh)\Vert _{L^2}+\Vert P_kA(SF,h)\Vert _{L^2}\\&\quad \lesssim \sum _{(k_1,k_2)\in \mathcal {X}_{k,l}}2^k\left[ \Vert P_{k_1}F\Vert _{L^\infty }\Vert P_{k_2}Sh\Vert _{L^2}+\Vert P_{k_1}SF\Vert _{L^2}\Vert P_{k_2}h\Vert _{L^\infty }\right] \\&\quad \quad +\sum _{(k_1,k_2)\in \mathcal {X}_{k,s}}2^k2^{\min (k_1,k_2)/2}\left[ \Vert P_{k_1}F\Vert _{L^2}\Vert P_{k_2}Sh\Vert _{L^2}+\Vert P_{k_1}SF\Vert _{L^2}\Vert P_{k_2}h\Vert _{L^2}\right] \\&\quad \lesssim {\varepsilon }_1^3(1+t)^{5p_0-1}2^{k/4}2^{-(N_0/2-16)k_+}. \end{aligned}$$

This proves the desired bounds for \(\Vert P_kA(F,h)\Vert _{L^2}\) and \(\Vert P_kSA(F,h)\Vert _{L^2}\). The other bounds in (5.37) are similar. \(\square \)

6 Proof of Proposition 4.4: uniform control of the \(Z\) norm

In this section we prove Proposition 4.4, which is our main bootstrap estimate. We start by rewriting the equation (5.7) in the form

$$\begin{aligned} \partial _t V + i\Lambda V = \widetilde{{\mathcal {N}}}_3 + R, \end{aligned}$$
(6.1)

where

$$\begin{aligned}&\widetilde{{\mathcal {N}}}_3:=M_3(H,H,\Psi ) + 2A_2(M_2(H,\Psi ),H)\nonumber \\&\quad + i \Lambda \left[ Q_3(\Psi ,H,\Psi ) + B(M_2(H,\Psi ),\Psi ) + B(H, Q_2(\Psi ,\Psi ))\right] ,\qquad \end{aligned}$$
(6.2)

and

$$\begin{aligned} R:=\mathcal {N}_3+\mathcal {N}_4-\widetilde{\mathcal {N}}_3. \end{aligned}$$
(6.3)

The point of the above decomposition is to express the cubic part of the nonlinearity in terms of \(H\) and \(\Psi \), hence of the solution \(V\).

Letting \(f(t)=e^{it\Lambda }V(t)\) as in (4.29), we have

$$\begin{aligned} \partial _tf=e^{it\Lambda }(\widetilde{{\mathcal {N}}}_3(t)+R(t)). \end{aligned}$$
(6.4)

Notice that

$$\begin{aligned} \begin{aligned}&H(t)=\frac{V(t)+\overline{V}(t)}{2}=\frac{e^{-it\Lambda }f(t)+e^{it\Lambda }\overline{f}(t)}{2},\\&\Psi (t)=\frac{i\Lambda ^{-1}(\overline{V}(t)-V(t))}{2}=i\frac{e^{it\Lambda }(\Lambda ^{-1}\overline{f})(t)-e^{-it\Lambda }(\Lambda ^{-1}f)(t)}{2}. \end{aligned} \end{aligned}$$
(6.5)

Therefore, the trilinear expression \(\widetilde{{\mathcal {N}}}_3\) can be written in terms of \(f(t)\) and \(\overline{f}(t)\), in the form

$$\begin{aligned}&\mathcal {F}(e^{it\Lambda }\widetilde{{\mathcal {N}}}_3(t))(\xi )\!=\!\frac{i}{(2\pi )^2} \big [I^{++-}(\xi ,t)\!+\!I^{--+}(\xi ,t)\nonumber \\&\qquad \qquad \qquad \quad +I^{+++}(\xi ,t)\!+\!I^{---}(\xi ,t)\big ],\nonumber \\&I^{++-}(\xi ,t):=\int _{\mathbb {R}\times \mathbb {R}}e^{it[\Lambda (\xi )-\Lambda (\xi -\eta )-\Lambda (\eta -\sigma )+\Lambda (\sigma )]}c^{++-}(\xi ,\eta ,\sigma )\nonumber \\&\qquad \qquad \qquad \quad \times \widehat{f}(\xi -\eta ,t)\widehat{f}(\eta -\sigma ,t)\widehat{\overline{f}}(\sigma ,t)\,d\eta d\sigma ,\nonumber \\&I^{--+}(\xi ,t):=\int _{\mathbb {R}\times \mathbb {R}}e^{it[\Lambda (\xi )-\Lambda (\xi -\eta )+\Lambda (\eta -\sigma )+\Lambda (\sigma )]}c^{--+}(\xi ,\eta ,\sigma )\widehat{\overline{f}}(\xi -\eta ,t)\nonumber \\&\qquad \qquad \qquad \quad \times \widehat{\overline{f}}(\eta -\sigma ,t)\widehat{f}(\sigma ,t)\,d\eta d\sigma ,\\&I^{+++}(\xi ,t):=\int _{\mathbb {R}\times \mathbb {R}}e^{it[\Lambda (\xi )-\Lambda (\xi -\eta )-\Lambda (\eta -\sigma )-\Lambda (\sigma )]}c^{+++}(\xi ,\eta ,\sigma )\widehat{f}(\xi -\eta ,t)\nonumber \\&\qquad \qquad \qquad \quad \times \widehat{f}(\eta -\sigma ,t)\widehat{f}(\sigma ,t)\,d\eta d\sigma ,\nonumber \\&I^{---}(\xi ,t):=\int _{\mathbb {R}\times \mathbb {R}}e^{it[\Lambda (\xi )+\Lambda (\xi -\eta )+\Lambda (\eta -\sigma )+\Lambda (\sigma )]}c^{---}(\xi ,\eta ,\sigma )\widehat{\overline{f}}(\xi -\eta ,t)\nonumber \\&\qquad \qquad \qquad \quad \times \widehat{\overline{f}}(\eta -\sigma ,t)\widehat{\overline{f}}(\sigma ,t)\,d\eta d\sigma .\nonumber \end{aligned}$$
(6.6)

The symbols \(c^{++-}, c^{--+}, c^{+++}, c^{---}\) can be calculated explicitly, using the formulas (4.9), (4.11), and (5.2)–(5.5), see Appendix B. For us it is important to notice that these symbols are real-valued, and satisfy the uniform bounds

$$\begin{aligned}&\big \Vert \mathcal {F}^{-1}[c^{\iota _1\iota _2\iota _3}(\xi ,\eta ,\sigma )\cdot \varphi _l(\xi )\varphi _{k_1}(\xi -\eta )\varphi _{k_2}(\eta -\sigma )\varphi _{k_3}(\sigma )]\big \Vert _{L^1(\mathbb {R}^3)}\nonumber \\&\quad \lesssim 2^{l/2}2^{2\max (k_1,k_2,k_3)}, \end{aligned}$$
(6.7)

for any \((\iota _1\iota _2\iota _3)\in \{(++-),(--+),(+++),(---)\}\) and \(l,k_1,k_2,k_3\in \mathbb {Z}\). As a consequence,

$$\begin{aligned} \big \Vert c^{\iota _1\iota _2\iota _3}(\xi ,\eta ,\sigma )\!\cdot \! \varphi _l(\xi )\varphi _{k_1}(\xi \!-\!\eta )\varphi _{k_2}(\eta \!-\!\sigma ) \varphi _{k_3}(\sigma ) \big \Vert _{\mathcal {S}^\infty _{\eta ,\sigma }} \!\lesssim \!2^{l/2}2^{2\max (k_1,k_2,k_3)},\nonumber \\ \end{aligned}$$
(6.8)

for any \((\iota _1\iota _2\iota _3)\in \{(++-),(--+),(+++),(---)\}\), \(\xi \in \mathbb {R}\), and \(l,k_1,k_2,k_3\in \mathbb {Z}\). Moreover, for any \(\mathbf {k}=(k_1,k_2,k_3),\mathbf {l}=(l_1,l_2,l_3)\in \mathbb {Z}^3\) let

$$\begin{aligned} c^*_\xi (x,y)&:=c^{++-}(\xi ,-x,-\xi -x-y),\\ (\partial _xc^*_\xi )_{\mathbf {k},\mathbf {l}}(x,y)&:=(\partial _xc^*_\xi )(x,y)\cdot \varphi _{k_1}(\xi +x)\varphi _{k_2}(\xi +y)\\&\quad \times \varphi _{k_3}(\xi +x+y)\varphi _{l_1}(x)\varphi _{l_2}(y)\varphi _{l_3}(2\xi +x+y),\\ (\partial _yc^*_\xi )_{\mathbf {k},\mathbf {l}}(x,y)&:=(\partial _yc^*_\xi )(x,y)\cdot \varphi _{k_1}(\xi +x)\varphi _{k_2}(\xi +y)\\&\quad \times \varphi _{k_3}(\xi +x+y)\varphi _{l_1}(x)\varphi _{l_2}(y)\varphi _{l_3}(2\xi +x+y). \end{aligned}$$

Then, for any \(\mathbf {k},\mathbf {l}\in \mathbb {Z}^3\), and \(\xi \in \mathbb {R}\),

$$\begin{aligned}&\Vert (\partial _xc^*_\xi )_{\mathbf {k},\mathbf {l}}\Vert _{\mathcal {S}^\infty }\lesssim 2^{-\min (k_1,k_3)}2^{5\max (k_1,k_2,k_3)/2},\nonumber \\&\Vert (\partial _yc^*_\xi )_{\mathbf {k},\mathbf {l}}\Vert _{\mathcal {S}^\infty }\lesssim 2^{-\min (k_2,k_3)}2^{5\max (k_1,k_2,k_3)/2}. \end{aligned}$$
(6.9)

These bounds are proved in Lemma 12.1.

Let (compare with (12.9)),

$$\begin{aligned} \widetilde{c}(\xi )&:=-8\pi |\xi |^{3/2}c^*_\xi (0,0)=4\pi |\xi |^4,\nonumber \\ L(\xi ,t)&:=\frac{\widetilde{c}(\xi )}{4\pi ^2}\int _0^t|\widehat{f}(\xi ,s)|^2 \frac{ds}{s+1},\nonumber \\ g(\xi ,t)&:=e^{iL(\xi ,t)}\widehat{f}(\xi ,t). \end{aligned}$$
(6.10)

It follows from (6.4) that

$$\begin{aligned} (\partial _tg)(\xi ,t)&=\frac{i}{(2\pi )^2}e^{iL(\xi ,t)}\Big [I^{++-}(\xi ,t)+\widetilde{c}(\xi )\frac{|\widehat{f}(\xi ,t)|^2}{t+1}\widehat{f}(\xi ,t)\Big ]\nonumber \\&\quad +\frac{i}{(2\pi )^2}e^{iL(\xi ,t)} [I^{--+}(\xi ,t)+I^{+++}(\xi ,t)+I^{---}(\xi ,t)]\nonumber \\&\quad +e^{iL(\xi ,t)}e^{it\Lambda (\xi )}\widehat{R}(\xi ,t). \end{aligned}$$
(6.11)

Proposition 4.4 clearly follows from Lemma 6.1 below.

Lemma 6.1

With the same notation as before, recall that \(f\) satisfies the bounds

$$\begin{aligned}&\sup _{t\in [0,T]}\Big [(1\!+\!t)^{-p_0}\Vert f(t)\Vert _{H^{N_0-10}}\nonumber \\&\quad +\,(1+t)^{-5p_0}\Vert (x\partial _xf)(t)\Vert _{H^{N_0/2-20}}\!+\!\Vert f(t)\Vert _{Z}\Big ]\!\le \!\varepsilon _1. \end{aligned}$$
(6.12)

Then there is \(p_1>0\) such that, for any \(m\in \{1,2,\ldots \}\) and any \(t_1\le t_2\in [2^{m}-2, 2^{m+1}]\),

$$\begin{aligned} \Vert (|\xi |^\beta +|\xi |^{N_1+15})(g(\xi ,t_2)-g(\xi ,t_1))\Vert _{L^\infty _\xi }\lesssim \varepsilon _1^32^{-p_1m}. \end{aligned}$$
(6.13)

The rest of the section is concerned with the proof of Lemma 6.1. We will use the following dispersive linear estimate from [27, Lemma 2.3]:

Lemma 6.2

For any \(t\in \mathbb {R}\) we have

$$\begin{aligned}&\Vert e^{i t\Lambda }h\Vert _{L^\infty }\!\lesssim \! (1+|t|)^{-1/2}\Vert \,|\xi |^{3/4} \widehat{h}(\xi )\Vert _{L^\infty _\xi }\nonumber \\&\quad +\,(1+|t|)^{-5/8}\left[ \Vert x\cdot \partial _x h\Vert _{L^2}\!+\!\Vert h\Vert _{H^2}\right] . \end{aligned}$$
(6.14)

For any \(k\in \mathbb {Z}\) let \(f_k^+:=P_kf\), \(f_k^-:=P_k\overline{f}\), and decompose,

$$\begin{aligned} I^{\iota _1\iota _2\iota _3}=\sum _{k_1,k_2,k_3\in \mathbb {Z}}I_{k_1,k_2,k_3}^{\iota _1\iota _2\iota _3}, \end{aligned}$$

for \((\iota _1\iota _2\iota _3)\in \{(++-),(--+),(+++),(---)\}\), where

$$\begin{aligned} I_{k_1,k_2,k_3}^{\iota _1\iota _2\iota _3}(\xi ,t)&:=\int _{\mathbb {R}\times \mathbb {R}}e^{it[\Lambda (\xi )-\iota _1\Lambda (\xi -\eta )-\iota _2\Lambda (\eta -\sigma )-\iota _3\Lambda (\sigma )]}\nonumber \\&\quad \times c^{\iota _1\iota _2\iota _3}(\xi ,\eta ,\sigma )\widehat{f_{k_1}^{\iota _1}}(\xi -\eta ,t)\widehat{f_{k_2}^{\iota _2}}(\eta -\sigma ,t)\widehat{f_{k_3}^{\iota _3}}(\sigma ,t)\,d\eta d\sigma . \end{aligned}$$
(6.15)

Using (6.11), for (6.13) it suffices to prove that if \(k\in \mathbb {Z}\), \(m\in \{1,2,\ldots \}\), \(|\xi |\in [2^k,2^{k+1}]\), and \(t_1\le t_2\in [2^m-2,2^{m+1}]\cap [0,T]\) then

$$\begin{aligned}&\sum _{k_1,k_2,k_3\in \mathbb {Z}}\left| \int _{t_1}^{t_2} e^{iL(\xi ,s)}\Big [I_{k_1,k_2,k_3}^{++-}(\xi ,s)+\widetilde{c}(\xi )\frac{\widehat{f_{k_1}^+}(\xi ,s)\widehat{f_{k_2}^+}(\xi ,s)\widehat{f_{k_3}^-}(-\xi ,s)}{s+1}\Big ]\,ds\right| \nonumber \\&\qquad \lesssim \varepsilon _1^32^{-p_1m}(2^{\beta k}+2^{(N_1+15)k})^{-1}, \end{aligned}$$
(6.16)
$$\begin{aligned} \sum _{k_1,k_2,k_3\in \mathbb {Z}}\left| \int _{t_1}^{t_2} e^{iL(\xi ,s)}I_{k_1,k_2,k_3}^{\iota _1\iota _2\iota _3}(\xi ,s)\,ds\right| \lesssim \varepsilon _1^32^{-p_1m}(2^{\beta k}+2^{(N_1+15)k})^{-1}\nonumber \\ \end{aligned}$$
(6.17)

for any \((\iota _1,\iota _2,\iota _3)\in \{(--+),(+++),(---)\}\), and

$$\begin{aligned} \left| \int _{t_1}^{t_2} e^{iL(\xi ,s)}e^{is\Lambda (\xi )}\widehat{R}(\xi ,s)\,ds\right| \lesssim \varepsilon _1^32^{-p_1m}(2^{\beta k}+2^{(N_1+15)k})^{-1}. \end{aligned}$$
(6.18)

In view of (6.12) and Lemma 6.14, we have

$$\begin{aligned} \Vert \widehat{f_l^{\pm }}(s)\Vert _{L^2}&\lesssim \varepsilon _12^{p_0 m}2^{-(N_0-10) l_+},\nonumber \\ \Vert (\partial \widehat{f_l^{\pm }})(s)\Vert _{L^2}&\lesssim \varepsilon _12^{5p_0 m}2^{-l}2^{-(N_0/2-20) l_+},\nonumber \\ \Vert \widehat{f_l^{\pm }}(s)\Vert _{L^\infty }&\lesssim \varepsilon _1(2^{\beta l}+2^{(N_1+15)l})^{-1},\nonumber \\ \Vert e^{\mp is\Lambda }f_l^{\pm }(s)\Vert _{L^\infty }&\lesssim \varepsilon _12^{-m/2}2^{-(N_0/2-20) l_+}, \end{aligned}$$
(6.19)

for any \(l\in \mathbb {Z}\) and \(s\in [2^m-2,2^{m+1}]\cap [0,T]\). Using only the \(L^2\) bounds in the first line of (6.19) it is easy to see that

$$\begin{aligned} |I_{k_1,k_2,k_3}^{\iota _1\iota _2\iota _3}(\xi ,s)|\lesssim \varepsilon _1^32^{3p_0 m}2^{\min (k_1,k_2,k_3)/2}(1+2^{\max (k_1,k_2,k_3)})^{-(N_0-20)},\nonumber \\ \end{aligned}$$
(6.20)

for any \((\iota _1\iota _2\iota _3)\in \{(++-),(--+),(+++),(---)\}\), \(k_1,k_2,k_3\in \mathbb {Z}\). Moreover, using the \(L^\infty \) bounds in (6.19),

$$\begin{aligned}&\left| \widetilde{c}(\xi )\frac{\widehat{f_{k_1}^+}(\xi ,s)\widehat{f_{k_2}^+}(\xi ,s)\widehat{f_{k_3}^-}(\xi ,s)}{s+1}\right| \lesssim 2^{-m}\varepsilon _1^32^{k}2^{-3N_1k_+}\mathbf {1}_{[0,4]}\\&\quad (\max (|k_1-k|,|k_2-k|,|k_3-k|)). \end{aligned}$$

Using these two bounds it is easy to see that the sums in (6.16) and (6.17) over those \((k_1,k_2,k_3)\) for which \(\max (k_1,k_2,k_3)\ge 3m/N_0-1000\) or \(\min (k_1,k_2,k_3)\le -4m\) are bounded by \(C\varepsilon _1^32^{-p_1m}2^{-(N_1+15)k_+}\), as desired. The remaining sums have only \(Cm^3\) terms. Therefore it suffices to prove the desired estimates for each \((k_1,k_2,k_3)\) fixed satisfying \(k_1,k_2,k_3\in [-4m,3m/N_0-1000]\).

At the same time, using (6.19), together with the symbol estimates (6.8), it follows that

$$\begin{aligned}&|I_{k_1,k_2,k_3}^{\iota _1\iota _2\iota _3}(\xi ,s)|\lesssim \varepsilon _1^3|\xi |^{1/2}2^{\min (k_1,k_2,k_3)(1-\beta )}\nonumber \\&\quad 2^{\mathrm {med}(k_1,k_2,k_3)(1-\beta )}2^{-(N_1+10)\max (k_1,k_2,k_3,0)}, \end{aligned}$$
(6.21)

for any \((\iota _1\iota _2\iota _3)\in \{(++-),(--+),(+++),(---)\}\), \(k_1,k_2,k_3\in \mathbb {Z}\), and \(\xi \in \mathbb {R}\), where \(\mathrm {med}(k_1,k_2,k_3)\) denotes the second largest frequency among \(k_1,k_2,k_3\).

After these reductions, it suffices to prove the following lemma:

Lemma 6.3

Assume that \(m\in \{1,2,\ldots \}\), \(k\in \mathbb {Z}\), \(|\xi |\in [2^k,2^{k+1}]\) and \(t_1\le t_2\in [2^{m}-2,2^{m+1}]\cap [0,T]\). Then, for any \(k_1,k_2,k_3\) satisfying

$$\begin{aligned}&k_1,k_2,k_3\in [-4m,3m/N_0-1000]\cap \mathbb {Z},\nonumber \\&\quad \min (k_1,k_2,k_3)+\,\mathrm {med}(k_1,k_2,k_3)\ge -m(1+3\beta ), \end{aligned}$$
(6.22)

we have

$$\begin{aligned}&\left| \int _{t_1}^{t_2} e^{iL(\xi ,s)}\Big [I_{k_1,k_2,k_3}^{++-}(\xi ,s)+\widetilde{c}(\xi )\frac{\widehat{f_{k_1}^+}(\xi ,s)\widehat{f_{k_2}^+}(\xi ,s)\widehat{f_{k_3}^-}(-\xi ,s)}{s+1}\Big ]\,ds\right| \nonumber \\&\quad \lesssim \varepsilon _1^32^{-2p_1m}(2^{\beta k}+2^{(N_1+15)k})^{-1}, \end{aligned}$$
(6.23)

and, for any \((\iota _1\iota _2\iota _3)\in \{(--+),(+++),(---)\}\),

$$\begin{aligned} \left| \int _{t_1}^{t_2} e^{iL(\xi ,s)}I_{k_1,k_2,k_3}^{\iota _1\iota _2\iota _3}(\xi ,s)\,ds\right| \lesssim \varepsilon _1^32^{-2p_1m}(2^{\beta k}+2^{(N_1+15)k})^{-1}. \end{aligned}$$
(6.24)

Moreover

$$\begin{aligned} \left| \int _{t_1}^{t_2} e^{iL(\xi ,s)}e^{is\Lambda (\xi )}\widehat{R}(\xi ,s)\,ds\right| \lesssim \varepsilon _1^32^{-p_1m}(2^{\beta k}+2^{(N_1+15)k})^{-1}. \end{aligned}$$
(6.25)

We will prove this main lemma in several steps. The bounds (6.23) and (6.24) are proved in the remaining part of this section. The main ingredients are the bounds (6.19). We will also use the following consequence of Lemma 5.2 (ii): if \((p,q,r)\in \{(2,2,\infty ),(2,\infty ,2),(\infty ,2,2)\}\) then

$$\begin{aligned} \left| \int _{\mathbb {R}\times \mathbb {R}}\widehat{f}(\eta )\widehat{g}(\sigma )\widehat{h}(-\eta -\sigma )m(\eta ,\sigma )\,d\eta d\sigma \right| \lesssim \Vert m\Vert _{\mathcal {S}^\infty }\Vert f\Vert _{L^p}\Vert g\Vert _{L^q}\Vert h\Vert _{L^r}.\nonumber \\ \end{aligned}$$
(6.26)

We also need suitable \(L^2\) bounds on the derivatives \((\partial _sf^{\pm }_l)\), in order to be able to integrate by parts in time. More precisely, we have

$$\begin{aligned} \Vert (\partial _s\widehat{f^{\pm }_l})(s)\Vert _{L^2}\lesssim \varepsilon _12^{5p_0m-m}2^{-(N_0-20)l_+}, \end{aligned}$$
(6.27)

which is a consequence of the (5.31)–(5.32) and the formula \(\partial _t f=e^{it\Lambda }[\mathcal {N}_3(t)+\mathcal {N}_4(t)]\), see (6.4).

The bound (6.25) is proved in appendix C.

6.1 Proof of (6.23)

We divide the proof in several cases.

Lemma 6.4

The bounds (6.23) hold provided that (6.22) holds and, in addition,

$$\begin{aligned} k_1,k_2,k_3\in [k-20,k+20]\cap \mathbb {Z}. \end{aligned}$$
(6.28)

Proof of Lemma 6.4

This is the main case, when the specific correction in the left-hand side of (6.23) is important. We will prove that

$$\begin{aligned}&\left| I_{k_1,k_2,k_3}^{++-}(\xi ,s)+\widetilde{c}(\xi )\frac{\widehat{f_{k_1}^+}(\xi ,s)\widehat{f_{k_2}^+} (\xi ,s)\widehat{f_{k_3}^-}(-\xi ,s)}{s+1}\right| \nonumber \\&\quad \lesssim 2^{-m}\varepsilon _1^32^{-2p_1m}(2^{\beta k}+2^{(N_1+15)k})^{-1}, \end{aligned}$$
(6.29)

for any \(s\in [t_1,t_2]\), which is clearly stronger than the desired bound (6.23).

The bound (6.29) follows easily from the bound (6.21) if \(k\le -m/2\). Therefore, in the rest of the proof of (6.29) we may assume that

$$\begin{aligned} k\ge -m/2. \end{aligned}$$
(6.30)

After changes of variables we rewriteFootnote 7

$$\begin{aligned} I_{k_1,k_2,k_3}^{++-}(\xi ,s)&= \int _{\mathbb {R}^2}e^{is\Phi (\xi ,\eta ,\sigma )} \widehat{f_{k_1}^+}(\xi +\eta ,s)\nonumber \\&\times \widehat{f_{k_2}^+}(\xi +\sigma ,s)\widehat{f_{k_3}^-}(-\xi -\eta -\sigma ,s)c^*_\xi (\eta ,\sigma )\,d\eta d\sigma , \end{aligned}$$

where

$$\begin{aligned} \Phi (\xi ,\eta ,\sigma ):=\Lambda (\xi )-\Lambda (\xi +\eta )-\Lambda (\xi +\sigma )+\Lambda (\xi +\eta +\sigma ). \end{aligned}$$
(6.31)

Let \(\overline{l}\) denote the smallest integer with the property that \(2^{\overline{l}}\ge 2^{3k/4}2^{-49m/100}\) (in view of (6.30) \(\overline{l}\le k-10\)). For any \(m,k\in \mathbb {Z}\), \(m\le k\), we define

$$\begin{aligned} \varphi ^{(m)}_k(x):= {\left\{ \begin{array}{ll} \varphi (x/2^k)-\varphi (x/2^{k-1}),\qquad &{}\text { if }\quad k\ge m+1,\\ \varphi (x/2^k),\qquad &{}\text { if }\quad k=m. \end{array}\right. } \end{aligned}$$
(6.32)

We decompose

$$\begin{aligned} I_{k_1,k_2,k_3}^{++-}(\xi ,s)=\sum _{l_1,l_2=\overline{l}}^{k+20}J_{l_1,l_2}(\xi ,s), \end{aligned}$$
(6.33)

where, for any \(l_1,l_2\ge \overline{l}\),

$$\begin{aligned} J_{l_1,l_2}(\xi ,s)&:= \int _{\mathbb {R}^2}e^{is\Phi (\xi ,\eta ,\sigma )}\widehat{f_{k_1}^+}(\xi +\eta ,s)\widehat{f_{k_2}^+}(\xi +\sigma ,s) \widehat{f_{k_3}^-}(-\xi -\eta -\sigma ,s)\nonumber \\&\times \,\varphi ^{(\overline{l})}_{l_1}(\eta )\varphi ^{(\overline{l})}_{l_2}(\sigma )c^*_\xi (\eta ,\sigma )\,d\eta d\sigma . \end{aligned}$$
(6.34)

Step 1. We show first that

$$\begin{aligned} |J_{l_1,l_2}(\xi ,s)|\lesssim 2^{-m}\varepsilon _1^32^{-3p_1m}(2^{\beta k}+2^{(N_1+15)k})^{-1},\quad \text { if }l_2\ge \max (l_1,\overline{l}+1).\nonumber \\ \end{aligned}$$
(6.35)

For this we integrate by parts in \(\eta \) in the formula (6.34). Recalling that \(\Lambda (\theta )=\sqrt{|\theta |}\), we observe that

$$\begin{aligned} \big |(\partial _\eta \Phi )(\xi ,\eta ,\sigma )\big |\!=\!\big |\Lambda '(\xi +\eta +\sigma )-\Lambda '(\xi +\eta )\big |\gtrsim 2^{l_2}2^{-3k/2}, \end{aligned}$$
(6.36)

provided that \(|\xi +\eta |\approx 2^k, |\xi +\eta +\sigma |\approx 2^k, |\sigma |\approx 2^{l_2}\). After integration by parts in \(\eta \) we see that

$$\begin{aligned} |J_{l_1,l_2}(\xi ,s)|\!\le \! |J^1_{l_1,l_2,1}(\xi ,s)|+|J^2_{l_1,l_2,1}(\xi ,s)|+|F_{l_1,l_2,1}(\xi ,s)|\!+\!|G_{l_1,l_2,1}(\xi ,s)|, \end{aligned}$$

where

$$\begin{aligned} J^1_{l_1,l_2,1}(\xi ,s)&:=\int _{\mathbb {R}^2}e^{is\Phi (\xi ,\eta ,\sigma )}\widehat{f_{k_1}^+}(\xi +\eta ,s)\widehat{f_{k_2}^+}(\xi +\sigma ,s)\nonumber \\&\quad \times \widehat{f_{k_3}^-}(-\xi -\eta -\sigma ,s)(\partial _\eta r_1)(\eta ,\sigma )c^*_\xi (\eta ,\sigma )\,d\eta d\sigma ,\nonumber \\ J^2_{l_1,l_2,1}(\xi ,s)&:=\int _{\mathbb {R}^2}e^{is\Phi (\xi ,\eta ,\sigma )}\widehat{f_{k_1}^+}(\xi +\eta ,s)\widehat{f_{k_2}^+}(\xi +\sigma ,s)\nonumber \\&\quad \times \widehat{f_{k_3}^-}(-\xi -\eta -\sigma ,s) r_1(\eta ,\sigma )(\partial _\eta c^*_\xi )(\eta ,\sigma )\,d\eta d\sigma ,\nonumber \\ F_{l_1,l_2,1}(\xi ,s)&:=\int _{\mathbb {R}^2}e^{is\Phi (\xi ,\eta ,\sigma )}(\partial \widehat{f_{k_1}^+})(\xi +\eta ,s)\widehat{f_{k_2}^+}(\xi +\sigma ,s)\nonumber \\&\quad \times \widehat{f_{k_3}^-}(-\xi -\eta -\sigma ,s)r_1(\eta ,\sigma )c^*_\xi (\eta ,\sigma )\,d\eta d\sigma ,\nonumber \\ G_{l_1,l_2,1}(\xi ,s)&:=\int _{\mathbb {R}^2}e^{is\Phi (\xi ,\eta ,\sigma )}\widehat{f_{k_1}^+}(\xi +\eta ,s)\widehat{f_{k_2}^+}(\xi +\sigma ,s)\nonumber \\&\quad \times (\partial \widehat{f_{k_3}^-})(-\xi -\eta -\sigma ,s)r_1(\eta ,\sigma )c^*_\xi (\eta ,\sigma )\,d\eta d\sigma , \end{aligned}$$
(6.37)

and

$$\begin{aligned} r_1(\eta ,\sigma ):=\frac{\varphi ^{(\overline{l})}_{l_1}(\eta )\varphi _{l_2}(\sigma )}{s(\partial _\eta \Phi )(\xi ,\eta ,\sigma )}\cdot \varphi _{[k_1-2,k_1+2]}(\xi +\eta )\varphi _{[k_3-2,k_3+2]}(\xi +\eta +\sigma ). \end{aligned}$$

To estimate \(|F_{l_1,l_2,1}(\xi ,s)|\) we recall that \(\xi \) and \(s\) are fixed and use (6.26) with

$$\begin{aligned} \widehat{f}(\eta )&:=e^{-is\Lambda (\xi +\eta )}(\partial \widehat{f_{k_1}^+})(\xi +\eta ,s),\\ \widehat{g}(\sigma )&:=e^{-is\Lambda (\xi +\sigma )}\widehat{f_{k_2}^+}(\xi +\sigma ,s)\cdot \varphi (\sigma /2^{l_2+4}),\\ \widehat{h}(\theta )&:=e^{is\Lambda (\xi -\theta )}\widehat{f_{k_3}^-}(-\xi +\theta ,s)\cdot \varphi (\theta /2^{l_2+4}). \end{aligned}$$

It is easy to see, compare with (6.36), that \(r_1\) satisfies the symbol-type estimates

$$\begin{aligned}&|(\partial _\eta ^a\partial _\sigma ^br_1)(\eta ,\sigma )|\lesssim (2^{-m}2^{-l_2}2^{3k/2})(2^{-al_1}2^{-bl_2})\nonumber \\&\quad \times \,\mathbf {1}_{[0,2^{l_1+4}]}(|\eta |)\mathbf {1}_{[2^{l_2-4},2^{l_2+4}]}(|\sigma |), \end{aligned}$$
(6.38)

for any \(a,b\in [0,20]\cap \mathbb {Z}\). It follows from (6.19) that

$$\begin{aligned} \Vert f\Vert _{L^2}&\lesssim \varepsilon _12^{-k}2^{5p_0m}2^{-(N_0/2-20)k_+},\quad \Vert g\Vert _{L^\infty }\lesssim \varepsilon _12^{-m/2}2^{-(N_0/2-20)k_+},\\ \Vert h\Vert _{L^2}&\lesssim \varepsilon _12^{l_2/2}2^{-\beta k}. \end{aligned}$$

It follows from (6.26), (6.38), and (6.8) that

$$\begin{aligned} \Vert \mathcal {F}^{-1}(r_1\cdot c^*_\xi )\Vert _{L^1}\lesssim 2^{-m}2^{-l_2}2^{4k}. \end{aligned}$$

Therefore, using (6.26) and recalling that \(2^{-l_2/2}\lesssim 2^{m/4}2^{-3k/8}\) and that \(k\le m/10\),

$$\begin{aligned} |F_{l_1,l_2,1}(\xi ,s)|&\lesssim \varepsilon _1^32^{-k}2^{5p_0m}2^{-(N_0/2-20)k_+}\cdot 2^{-m/2}2^{-(N_0/2-20)k_+}\cdot 2^{l_2/2}2^{-\beta k}\\&\cdot 2^{-m}2^{-l_2}2^{4k}\lesssim \varepsilon _1^32^{-(N_0/2+40)k_+}2^{-m}\cdot 2^{-m/8}. \end{aligned}$$

Similar arguments show that \(|G_{l_1,l_2,1}(\xi ,s)|\lesssim \varepsilon _1^32^{-(N_0/2+40)k_+}2^{-9m/8}\) and, using also the bound (6.9)Footnote 8, \(|J^2_{l_1,l_2,1}(\xi ,s)|\lesssim \varepsilon _1^32^{-(N_0/2+40)k_+}2^{-9m/8}\). Therefore, for (6.35) it suffices to prove that

$$\begin{aligned} |J^1_{l_1,l_2,1}(\xi ,s)|\lesssim 2^{-m}\varepsilon _1^32^{-3p_1m}(2^{\beta k}+2^{(N_1+15)k})^{-1}. \end{aligned}$$
(6.39)

For this we integrate by parts again in \(\eta \) and estimate

$$\begin{aligned} |J^1_{l_1,l_2,1}(\xi ,s)|&\le |J^1_{l_1,l_2,2}(\xi ,s)|+|J^2_{l_1,l_2,2}(\xi ,s)|+|F_{l_1,l_2,2}(\xi ,s)|\\&+|G_{l_1,l_2,2}(\xi ,s)|, \end{aligned}$$

where

$$\begin{aligned} J^1_{l_1,l_2,2}(\xi ,s)&:=\int _{\mathbb {R}^2}e^{is\Phi (\xi ,\eta ,\sigma )}\widehat{f_{k_1}^+}(\xi +\eta ,s)\widehat{f_{k_2}^+}(\xi +\sigma ,s)\\&\quad \times \widehat{f_{k_3}^-}(-\xi -\eta -\sigma ,s)(\partial _\eta r_2)(\eta ,\sigma )c^*_\xi (\eta ,\sigma )\,d\eta d\sigma ,\\ J^2_{l_1,l_2,2}(\xi ,s)&:=\int _{\mathbb {R}^2}e^{is\Phi (\xi ,\eta ,\sigma )}\widehat{f_{k_1}^+}(\xi +\eta ,s)\widehat{f_{k_2}^+}(\xi +\sigma ,s)\\&\quad \times \widehat{f_{k_3}^-}(-\xi -\eta -\sigma ,s)r_2(\eta ,\sigma )(\partial _\eta c^*_\xi )(\eta ,\sigma )\,d\eta d\sigma ,\\ F_{l_1,l_2,2}(\xi ,s)&:=\int _{\mathbb {R}^2}e^{is\Phi (\xi ,\eta ,\sigma )}(\partial \widehat{f_{k_1}^+})(\xi +\eta ,s)\widehat{f_{k_2}^+}(\xi +\sigma ,s)\\&\quad \times \widehat{f_{k_3}^-}(-\xi -\eta -\sigma ,s)r_2(\eta ,\sigma )c^*_\xi (\eta ,\sigma )\,d\eta d\sigma ,\\ G_{l_1,l_2,2}(\xi ,s)&:=\int _{\mathbb {R}^2}e^{is\Phi (\xi ,\eta ,\sigma )}\widehat{f_{k_1}^+}(\xi +\eta ,s)\widehat{f_{k_2}^+}(\xi +\sigma ,s)\\&\quad \times (\partial \widehat{f_{k_3}^-})(-\xi -\eta -\sigma ,s)r_2(\eta ,\sigma )c^*_\xi (\eta ,\sigma )\,d\eta d\sigma . \end{aligned}$$

and

$$\begin{aligned} r_2(\eta ,\sigma ):=\frac{(\partial _\eta r_1)(\eta ,\sigma )}{s(\partial _\eta \Phi )(\xi ,\eta ,\sigma )} . \end{aligned}$$

It follows from (6.38) that \(r_2\) satisfies the stronger symbol-type bounds

$$\begin{aligned}&|(\partial _\eta ^a\partial _\sigma ^br_2)(\eta ,\sigma )|\lesssim (2^{-m}2^{-l_1-l_2}2^{3k/2})(2^{-m}2^{-l_2}2^{3k/2})(2^{-al_1}2^{-bl_2})\nonumber \\&\quad \times \,\mathbf {1}_{[0,2^{l_1+4}]}(|\eta |)\mathbf {1}_{[2^{l_2-4},2^{l_2+4}]}(|\sigma |), \end{aligned}$$
(6.40)

for \(a,b\in [0,19]\cap \mathbb {Z}\). Therefore, using Lemma 5.2 as before,

$$\begin{aligned}&|F_{l_1,l_2,2}(\xi ,s)|+|G_{l_1,l_2,2}(\xi ,s)|+|J^2_{l_1,l_2,2}(\xi ,s)|\nonumber \\&\quad \lesssim \varepsilon _1^32^{-(N_0/2+40)k_+}2^{-m}\cdot 2^{-m/8}. \end{aligned}$$

Moreover, we can now estimate \(|J^1_{l_1,l_2,2}(\xi ,s)|\) using only (6.40) and the first \(L^\infty \) bounds in (6.19),

$$\begin{aligned} |J_{l_1,l_2,2}(\xi ,s)|&\lesssim 2^{l_1+l_2}\cdot \varepsilon _1^3(2^{\beta k}+2^{(N_1+15)k})^{-3}\cdot (2^{-m}2^{-l_1-l_2}2^{3k/2})^22^{5k/2}\\&\lesssim \varepsilon _1^32^{-(N_0/2+40)k_+}2^{-m}\cdot 2^{-m/50}. \end{aligned}$$

This completes the proof of (6.39) and (6.35).

A similar argument shows that

$$\begin{aligned} |J_{l_1,l_2}(\xi ,s)|\lesssim 2^{-m}\varepsilon _1^32^{-3p_1m}(2^{\beta k}+2^{(N_1+15)k})^{-1},\qquad \text { if }l_1\ge \max (l_2,\overline{l}+1). \end{aligned}$$

Step 2. Using the decomposition (6.33), for (6.29) it suffices to prove that

$$\begin{aligned}&\left| J_{\overline{l},\overline{l}}(\xi ,s)+\widetilde{c}(\xi )\frac{\widehat{f_{k_1}^+}(\xi ,s)\widehat{f_{k_2}^+}(\xi ,s)\widehat{f_{k_3}^-}(-\xi ,s)}{s+1}\right| \nonumber \\&\quad \lesssim 2^{-m}\varepsilon _1^32^{-2p_1m}(2^{\beta k}+2^{(N_1+15)k})^{-1}. \end{aligned}$$
(6.41)

To prove (6.41) we notice that

$$\begin{aligned} \left| \Phi (\xi ,\eta ,\sigma )+\frac{\eta \sigma }{4|\xi |^{3/2}}\right| \lesssim 2^{-5k/2}(|\eta |+|\sigma |)^3, \end{aligned}$$

as long as \(|\eta |+|\sigma |\le 2^{k-5}\). Therefore, using the \(L^\infty \) bounds in (6.19)

$$\begin{aligned}&\left| J_{\overline{l},\overline{l}}(\xi ,s)-J'_{\overline{l},\overline{l}}(\xi ,s)\right| \lesssim \varepsilon _1^3(2^{\beta k}+2^{(N_1+15)k})^{-3}\cdot 2^m2^{-5k/2}2^{5\overline{l}}\nonumber \\&\quad \lesssim \varepsilon _1^3(2^{\beta k}+2^{(N_1+15)k})^{-1}2^{-5m/4}, \end{aligned}$$
(6.42)

where

$$\begin{aligned} J'_{\overline{l},\overline{l}}(\xi ,s)&:=\int _{\mathbb {R}^2}e^{-is\eta \sigma /(4|\xi |^{3/2})}\widehat{f_{k_1}^+}(\xi +\eta ,s)\widehat{f_{k_2}^+}(\xi +\sigma ,s)\nonumber \\&\quad \times \widehat{f_{k_3}^-}(-\xi -\eta -\sigma ,s)\varphi (2^{-\overline{l}}\eta )\varphi (2^{-\overline{l}}\sigma )c^*_\xi (\eta ,\sigma )\,d\eta d\sigma . \end{aligned}$$
(6.43)

Using the second bound in (6.19), we see that

$$\begin{aligned}&| \widehat{f_l} (\xi +\rho ,s) - \widehat{f_l} (\xi ,s)| \lesssim \int _0^\rho | \partial \widehat{f_l}(\xi +\mu ,s)| \,d\mu \\&\quad \lesssim {|\rho |}^{1/2} {\Vert \partial \widehat{f_l}(s) \Vert }_{L^2} \lesssim {|\rho |}^{1/2} {\varepsilon }_1 2^{5p_0m} 2^{-l} \end{aligned}$$

whenever \(|\rho | \lesssim 2^{l-10}\). Then, using the third bound in (6.19) it follows that

$$\begin{aligned}&|\widehat{f_{k_1}^+}(\xi +\eta ,s)\widehat{f_{k_2}^+}(\xi +\sigma ,s)\widehat{f_{k_3}^-}(-\xi -\eta -\sigma ,s)\\&\quad -\widehat{f_{k_1}^+}(\xi ,s) \widehat{f_{k_2}^+}(\xi ,s)\widehat{f_{k_3}^-}(-\xi ,s)|\lesssim \varepsilon _1^32^{\overline{l}/2}\cdot 2^{-N_0k_+}2^{5p_0m}2^{-k(1+2\beta )}, \end{aligned}$$

whenever \(|\eta |+|\sigma |\le 2^{\overline{l}+4}\). In addition, using (6.9),

$$\begin{aligned} |c^*_\xi (\eta ,\sigma )-c^*_\xi (0,0)|\lesssim 2^{3k/2}2^{\overline{l}}, \end{aligned}$$

provided that \(|\eta |+|\sigma |\le 2^{\overline{l}+4}\). Therefore

$$\begin{aligned}&\left| J'_{\overline{l},\overline{l}}(\xi ,s)-\int _{\mathbb {R}^2}e^{-is\eta \sigma /(4|\xi |^{3/2})}\widehat{f_{k_1}^+}(\xi ,s)\widehat{f_{k_2}^+}(\xi ,s)\widehat{f_{k_3}^-}(-\xi ,s)\varphi (2^{-\overline{l}}\eta )\right. \nonumber \\&\qquad \times \left. \varphi (2^{-\overline{l}}\sigma )c^*_\xi (0,0)\,d\eta d\sigma \right| \lesssim 2^{2\overline{l}}2^{3k/2}\cdot \varepsilon _1^32^{\overline{l}/2}2^{-N_0k_+}2^{5p_0m}2^{-k(1+2\beta )}\nonumber \\&\qquad \lesssim \varepsilon _1^3(2^{\beta k}+2^{(N_1+15)k})^{-1}2^{-9m/8}. \end{aligned}$$
(6.44)

Starting from the general formula

$$\begin{aligned} \int _{\mathbb {R}}e^{-ax^2-bx}\,dx=e^{b^2/(4a)}\sqrt{\pi }/\sqrt{a},\qquad a,b\in \mathbb {C},\,\,\text{ Re }\,a>0, \end{aligned}$$

we calculate, for any \(N\ge 1\),

$$\begin{aligned}&\int _{\mathbb {R}\times \mathbb {R}}e^{-ixy}e^{-x^2/N^2}e^{-y^2/N^2}\,dxdy=\sqrt{\pi }N\int _{\mathbb {R}}e^{-y^2/N^2}e^{-N^2y^2/4}\,dy\\&\quad =2\pi +O(N^{-1}). \end{aligned}$$

Therefore, for \(N\ge 1\),

$$\begin{aligned} \int _{\mathbb {R}\times \mathbb {R}}e^{-ixy}\varphi (x/N)\varphi (y/N)\,dxdy=2\pi +O(N^{-1/2}). \end{aligned}$$

Recalling also that \(2^{\overline{l}}\approx |\xi |^{3/4}2^{-49m/100}\), it follows that

$$\begin{aligned} \left| \int _{\mathbb {R}^2}e^{-is\eta \sigma /(4|\xi |^{3/2})}\varphi (2^{-\overline{l}}\eta )\varphi (2^{-\overline{l}}\sigma )\,d\eta d\sigma -\frac{4|\xi |^{3/2}}{s}(2\pi )\right| \lesssim 2^{3k/2}2^{-(1+4p_1)m}. \end{aligned}$$

Therefore, using also (6.19),

$$\begin{aligned}&\left| \int _{\mathbb {R}^2}e^{-is\eta \sigma /(4|\xi |^{3/2})}\widehat{f_{k_1}^+}(\xi ,s)\widehat{f_{k_2}^+}(\xi ,s)\widehat{f_{k_3}^-}(-\xi ,s)\varphi (2^{-\overline{l}}\eta )\varphi (2^{-\overline{l}}\sigma )c^*_\xi (0,0)\,d\eta d\sigma \nonumber \right. \\&\left. \quad -\frac{8\pi |\xi |^{3/2}c^*_\xi (0,0)\cdot \widehat{f_{k_1}^+}(\xi ,s)\widehat{f_{k_2}^+}(\xi ,s)\widehat{f_{k_3}^-}(-\xi ,s)}{s}\right| \lesssim \varepsilon _1^32^{-(1+4p_1)m}2^{-N_0k_+}. \end{aligned}$$
(6.45)

and the bound (6.41) follows from (6.42), (6.44), and (6.45). This completes the proof of the lemma. \(\square \)

Lemma 6.5

The bounds (6.23) hold provided that (6.22) holds and, in addition,

$$\begin{aligned}&\max (|k_1-k|,|k_2-k|,|k_3-k|)\ge 21,\nonumber \\&\min (k_1,k_2,k_3)\ge -19m/20,\qquad \max (|k_1-k_3|,|k_2-k_3|)\ge 5. \end{aligned}$$
(6.46)

Proof of Lemma 6.5

Recall the definition

$$\begin{aligned} I_{k_1,k_2,k_3}^{++-}(\xi ,s)&= \int _{\mathbb {R}^2}e^{is\Phi (\xi ,\eta ,\sigma )}\widehat{f_{k_1}^+}(\xi +\eta ,s)\widehat{f_{k_2}^+}(\xi +\sigma ,s)\nonumber \\&\times \widehat{f_{k_3}^-}(-\xi -\eta -\sigma ,s)c^*_\xi (\eta ,\sigma )\,d\eta d\sigma , \end{aligned}$$
(6.47)

where

$$\begin{aligned} \Phi (\xi ,\eta ,\sigma )=\Lambda (\xi )-\Lambda (\xi +\eta )-\Lambda (\xi +\sigma )+\Lambda (\xi +\eta +\sigma ). \end{aligned}$$

It suffices to prove that, for any \(s\in [2^m-1,2^{m+1}]\),

$$\begin{aligned} \big |I_{k_1,k_2,k_3}^{++-}(\xi ,s)\big |\lesssim 2^{-m}\varepsilon _1^32^{-2p_1m}2^{-(N_1+15)k_+}. \end{aligned}$$
(6.48)

By symmetry, we may assume that \(|k_1-k_3|\ge 5\) and notice that

$$\begin{aligned} |(\partial _\eta \Phi )(\xi ,\eta ,\sigma )|=|-\Lambda '(\xi +\eta )+\Lambda '(\xi +\eta +\sigma )|\gtrsim 2^{-\min (k_1,k_3)/2},\nonumber \\ \end{aligned}$$
(6.49)

provided that \(|\xi +\eta |\in [2^{k_1-2},2^{k_1+2}]\), \(|\xi +\eta +\sigma |\in [2^{k_3-2},2^{k_3+2}]\). As in the proof of Lemma 6.4, we integrate by parts in \(\eta \) to estimate

$$\begin{aligned} |I_{k_1,k_2,k_3}^{++-}(\xi ,s)|\le |J^1_{1}(\xi ,s)|+|J^2_{1}(\xi ,s)|+|F_{1}(\xi ,s)|+|G_{1}(\xi ,s)|, \end{aligned}$$

where

$$\begin{aligned} J_{1}^1(\xi ,s)&:=\int _{\mathbb {R}^2}e^{is\Phi (\xi ,\eta ,\sigma )}\widehat{f_{k_1}^+}(\xi +\eta ,s)\widehat{f_{k_2}^+}(\xi +\sigma ,s) \widehat{f_{k_3}^-}(-\xi -\eta -\sigma ,s)\\&\quad \times (\partial _\eta r_3)(\eta ,\sigma )c^*_\xi (\eta ,\sigma )\,d\eta d\sigma ,\\ J_{1}^2(\xi ,s)&:=\int _{\mathbb {R}^2}e^{is\Phi (\xi ,\eta ,\sigma )}\widehat{f_{k_1}^+}(\xi +\eta ,s)\widehat{f_{k_2}^+}(\xi +\sigma ,s) \widehat{f_{k_3}^-}(-\xi -\eta -\sigma ,s) \\&\quad \times r_3(\eta ,\sigma )(\partial _\eta c^*_\xi )(\eta ,\sigma )\,d\eta d\sigma ,\\ F_{1}(\xi ,s)&:=\int _{\mathbb {R}^2}e^{is\Phi (\xi ,\eta ,\sigma )}(\partial \widehat{f_{k_1}^+})(\xi +\eta ,s)\widehat{f_{k_2}^+}(\xi +\sigma ,s) \widehat{f_{k_3}^-}(-\xi -\eta -\sigma ,s)\\&\quad \times r_3(\eta ,\sigma )c^*_\xi (\eta ,\sigma )\,d\eta d\sigma ,\\ G_{1}(\xi ,s)&=\int _{\mathbb {R}^2}e^{is\Phi (\xi ,\eta ,\sigma )}\widehat{f_{k_1}^+}(\xi +\eta ,s)\widehat{f_{k_2}^+}(\xi +\sigma ,s)\\&\quad \times (\partial \widehat{f_{k_3}^-})(-\xi -\eta -\sigma ,s)r_3(\eta ,\sigma )c^*_\xi (\eta ,\sigma )\,d\eta d\sigma , \end{aligned}$$

and

$$\begin{aligned} r_3(\eta ,\sigma ):=\frac{1}{s(\partial _\eta \Phi )(\xi ,\eta ,\sigma )}\cdot \varphi _{[k_1-1,k_1+1]}(\xi +\eta )\varphi _{[k_3-1,k_3+1]}(\xi +\eta +\sigma ). \end{aligned}$$

Using also (6.49), it follows easily that

$$\begin{aligned} \Vert \mathcal {F}^{-1}(r_3)\Vert _{L^1}\lesssim 2^{-m}2^{\min (k_1,k_3)/2}, \qquad \Vert \mathcal {F}^{-1}(\partial _\eta r_3)\Vert _{L^1}\lesssim 2^{-m}2^{-\min (k_1,k_3)/2}. \end{aligned}$$

We apply first (6.26) with

$$\begin{aligned} \widehat{f}(\eta )&:=e^{-is\Lambda (\xi +\eta )}\widehat{f_{k_1}^+}(\xi +\eta ,s),\, \widehat{g}(\sigma ):=e^{-is\Lambda (\xi +\sigma )}\widehat{f_{k_2}^+}(\xi +\sigma ,s),\, \\ \widehat{h}(\theta )&:=e^{is\Lambda (\xi -\theta )}\widehat{f_{k_3}^-}(-\xi +\theta ,s). \end{aligned}$$

Using also (6.19), (6.8), and (6.9), we conclude that

$$\begin{aligned}&|J^1_{1}(\xi ,s)| +|J^2_{1}(\xi ,s)|\lesssim \varepsilon _1^32^{-m/2}2^{3p_0m}2^{-(N_0-20)\max (k_1,k_2,k_3)_+}\\&\quad \cdot 2^{-m}2^{-\min (k_1,k_3)/2}. \end{aligned}$$

Similarly, we apply (6.26) with

$$\begin{aligned} \widehat{f}(\eta )&:=e^{-is\Lambda (\xi +\eta )}(\partial \widehat{f_{k_1}^+})(\xi +\eta ,s),\, \widehat{g}(\sigma ):=e^{-is\Lambda (\xi +\sigma )}\widehat{f_{k_2}^+}(\xi +\sigma ,s),\\ \widehat{h}(\theta )&:=e^{is\Lambda (\xi -\theta )}\widehat{f_{k_3}^-}(-\xi +\theta ,s), \end{aligned}$$

and use (6.19) to conclude that

$$\begin{aligned} |F_{1}(\xi ,s)|\lesssim \varepsilon _1^32^{-m/2}2^{8p_0m}2^{-k_1}2^{-(N_0/2-30)\max (k_1,k_2,k_3)_+}\cdot 2^{-m}2^{\min (k_1,k_3)/2}. \end{aligned}$$

Finally, we apply (6.26) with

$$\begin{aligned} \widehat{f}(\eta )&:=e^{-is\Lambda (\xi +\eta )}\widehat{f_{k_1}^+}(\xi +\eta ,s),\, \widehat{g}(\sigma ):=e^{-is\Lambda (\xi +\sigma )}\widehat{f_{k_2}^+}(\xi +\sigma ,s), \\ \widehat{h}(\theta )&:=e^{is\Lambda (\xi -\theta )}(\partial \widehat{f_{k_3}^-})(-\xi +\theta ,s), \end{aligned}$$

and use (6.19) to conclude that

$$\begin{aligned} |G_{1}(\xi ,s)|\lesssim \varepsilon _1^32^{-m/2}2^{8p_0m}2^{-k_3}2^{-(N_0/2-30)\max (k_1,k_2,k_3)_+}\cdot 2^{-m}2^{\min (k_1,k_3)/2}. \end{aligned}$$

Therefore

$$\begin{aligned}&|J_{1}(\xi ,s)|+|F_{1}(\xi ,s)|+|G_{1}(\xi ,s)|\lesssim \varepsilon _1^32^{-m}2^{-(N_0/2-30)k_+}\nonumber \\&\quad 2^{-m/2+8p_0m}2^{-\min (k_1,k_3)/2}, \end{aligned}$$

and the desired bound (6.48) follows from the assumptions \(-\min (k_1,k_3)/2\le 19m/40\), see (6.46), and \(\max (k_1,k_2,k_3)_+\le 3m/N_0\) (see the hypothesis of Lemma 6.3). \(\square \)

Lemma 6.6

The bounds (6.23) hold provided that (6.22) holds and, in addition,

$$\begin{aligned}&\max (|k_1-k|,|k_2-k|,|k_3-k|)\ge 21,\nonumber \\&\min (k_1,k_2,k_3)\ge -19m/20,\qquad \max (|k_1-k_3|,|k_2-k_3|)\le 4. \end{aligned}$$
(6.50)

Proof of Lemma 6.6

We may assume that

$$\begin{aligned} \min (k_1,k_2,k_3)\ge k+10, \end{aligned}$$
(6.51)

and rewrite

$$\begin{aligned} I_{k_1,k_2,k_3}^{++-}(\xi ,s)&=\int _{\mathbb {R}^2}e^{is\Phi (\xi ,\eta ,\sigma )}\widehat{f_{k_1}^+}(\xi +\eta ,s)\widehat{f_{k_2}^+}(\xi +\sigma ,s)\nonumber \\&\quad \times \widehat{f_{k_3}^-}(-\xi -\eta -\sigma ,s)\varphi _{[k_2-4,k_2+4]}(\sigma )c_\xi ^*(\eta ,\sigma )\,d\eta d\sigma , \end{aligned}$$
(6.52)

where, as before,

$$\begin{aligned} \Phi (\xi ,\eta ,\sigma )=\Lambda (\xi )-\Lambda (\xi +\eta )-\Lambda (\xi +\sigma )+\Lambda (\xi +\eta +\sigma ). \end{aligned}$$

It suffices to prove that, for any \(s\in [t_1,t_2]\),

$$\begin{aligned} \big |I_{k_1,k_2,k_3}^{++-}(\xi ,s)\big |\lesssim 2^{-m}\varepsilon _1^32^{-2p_1m}2^{-(N_1+15)k_+}. \end{aligned}$$
(6.53)

Notice that

$$\begin{aligned} |(\partial _\eta \Phi )(\xi ,\eta ,\sigma )|=|-\Lambda '(\xi +\eta )+\Lambda '(\xi +\eta +\sigma )|\gtrsim 2^{-k_2/2}, \end{aligned}$$
(6.54)

provided that \(|\xi +\eta |\in [2^{k_1-2},2^{k_1+2}]\), \(|\xi +\eta +\sigma |\in [2^{k_3-2},2^{k_3+2}]\), and \(|\sigma |\approx 2^{k_2}\) (recall also that \(2^{k_1}\approx 2^{k_2}\approx 2^{k_3}\)). The bound (6.53) follows by integration by parts in \(\eta \), as in the proof of Lemma 6.5. \(\square \)

Lemma 6.7

The bounds (6.23) hold provided that (6.22) holds and, in addition,

$$\begin{aligned}&\max (|k_1-k|,|k_2-k|,|k_3-k|)\ge 21,\nonumber \\&\min (k_1,k_2,k_3)\le -19m/20,\qquad k\le -m/5. \end{aligned}$$
(6.55)

Proof of Lemma 6.7

It follows from the definition and the bounds (6.8) and (6.19) that, for any \(s\in [t_1,t_2]\),

$$\begin{aligned} |I_{k_1,k_2,k_3}^{++-}(\xi ,s)|&\lesssim \varepsilon _1^3|\xi |^{1/2}2^{(1-\beta )\min (k_1,k_2,k_3)}2^{2p_0m}2^{-(N_0-20)\max (k_1,k_2,k_3,0)}\\&\lesssim 2^{-m}\varepsilon _1^32^{-2p_1m}2^{-(N_1+15)k_+}. \end{aligned}$$

The desired estimate (6.23) follows in this case. \(\square \)

Lemma 6.8

The bounds (6.23) hold provided that (6.22) holds and, in addition,

$$\begin{aligned}&\max (|k_1-k|,|k_2-k|,|k_3-k|)\ge 21,\nonumber \\&\min (k_1,k_2,k_3)\le -19m/20,\qquad k\ge -m/5. \end{aligned}$$
(6.56)

Proof of Lemma 6.8

In this case we cannot prove pointwise bounds on \(\big |I_{k_1,k_2,k_3}^{+,+,-}(\xi ,s)|\) and we need to integrate by parts in \(s\). For (6.23) it suffices to prove that

$$\begin{aligned}&\left| \int _{\mathbb {R}^2\times [t_1,t_2]}e^{iL(\xi ,s)}e^{is\Phi (\xi ,\eta ,\sigma )}\widehat{f_{k_1}^+}(\xi +\eta ,s)\widehat{f_{k_2}^+}(\xi +\sigma ,s)\right. \nonumber \\&\quad \left. \times \widehat{f_{k_3}^-}(-\xi -\eta -\sigma ,s)c^*_\xi (\eta ,\sigma )\,d\eta d\sigma ds \right| \lesssim \varepsilon _1^32^{-2p_1m}2^{-(N_1+15)k_+},\qquad \end{aligned}$$
(6.57)

where

$$\begin{aligned} \Phi (\xi ,\eta ,\sigma )&=\Lambda (\xi )-\Lambda (\xi +\eta )-\Lambda (\xi +\sigma )+\Lambda (\xi +\eta +\sigma ),\\ L(\xi ,s)&=\frac{\widetilde{c}(\xi )}{4\pi ^2}\int _0^s|\widehat{f}(\xi ,r)|^2 \frac{dr}{r+1}. \end{aligned}$$

The assumptions (6.56) and (6.22) show that

$$\begin{aligned} \widetilde{k}:=\min (k,\mathrm {med}(k_1,k_2,k_3))\ge -m/5. \end{aligned}$$
(6.58)

Then we make the simple observation that

$$\begin{aligned} \Lambda (a)+\Lambda (b)-\Lambda (a+b)\ge \Lambda (a)/2\qquad \text { if }\qquad 0\le a\le b \end{aligned}$$

to conclude that \(\Phi \) satisfies the weakly elliptic bound

$$\begin{aligned} |\Phi (\xi ,\eta ,\sigma )|\ge 2^{\widetilde{k}/2-10} \end{aligned}$$
(6.59)

provided that \(|\xi +\eta |\in [2^{k_1-2},2^{k_1+2}]\), \(|\xi +\sigma |\in [2^{k_2-2},2^{k_2+2}]\), \(|\xi +\eta +\sigma |\in [2^{k_3-2},2^{k_3+2}]\). Letting \(\dot{L}(\xi ,s):=(\partial _sL)(\xi ,s)\), where \(L\) is defined in (6.10), and using (6.19), we notice that, for any \(s\in [2^{m}-1,2^{m+1}]\),

$$\begin{aligned} |\dot{L}(\xi ,s)|\lesssim \varepsilon _1^22^{2k}2^{-N_0k_+}2^{-m}. \end{aligned}$$
(6.60)

We integrate by parts in \(s\) to conclude that the integral in the left-hand side of (6.57) is dominated by

$$\begin{aligned} B^0(\xi )+\sum _{j=1}^2B_j(\xi ), \end{aligned}$$

where

$$\begin{aligned} B^0(\xi )&:=\int _{t_1}^{t_2}\left| \int _{\mathbb {R}^2} e^{is\Phi (\xi ,\eta ,\sigma )} \frac{1}{\Phi (\xi ,\eta ,\sigma )} \frac{d}{ds} \left[ e^{i L(\xi ,s)}\widehat{f_{k_1}^+}(\xi +\eta ,s)\widehat{f_{k_2}^+}(\xi +\sigma ,s) \right. \right. \\&\quad \left. \left. \times \widehat{f_{k_3}^-}(-\xi -\eta -\sigma ,s)c^*_\xi (\eta ,\sigma )\right] \,d\eta d\sigma \right| ds \end{aligned}$$

and, for \(j=1,2\),

$$\begin{aligned} B_j(\xi )&:=\left| \int _{\mathbb {R}^2}e^{it_j\Phi (\xi ,\eta ,\sigma )}\frac{1}{\Phi (\xi ,\eta ,\sigma )}\widehat{f_{k_1}^+}(\xi +\eta ,t_j)\widehat{f_{k_2}^+}(\xi +\sigma ,t_j)\right. \\&\left. \quad \times \widehat{f_{k_3}^-}(-\xi -\eta -\sigma ,t_j)c^*_\xi (\eta ,\sigma )\,d\eta d\sigma \right| \end{aligned}$$

Let

$$\begin{aligned} r_4(\eta ,\sigma )&:= \frac{1}{\Phi (\xi ,\eta ,\sigma )}\cdot \varphi _{[k_1-1,k_1+1]}(\xi +\eta )\varphi _{[k_2-1,k_2+1]}(\xi +\sigma )\\&\times \varphi _{[k_3-1,k_3+1]}(\xi +\eta +\sigma ). \end{aligned}$$

Using (6.59)–(6.60) and integration by parts it is easy to see that, for any \(s\in [t_1,t_2]\),

$$\begin{aligned} \Vert \mathcal {F}^{-1}(r_4)\Vert _{L^1}\lesssim 2^{-\widetilde{k}/2}. \end{aligned}$$
(6.61)

Using the first \(L^\infty \) bound in (6.19) we estimate, for \(j\in \{1,2\}\),

$$\begin{aligned} B_j(\xi )\lesssim \varepsilon _1^32^{(1-\beta )\min (k_1,k_2,k_3)}2^{-(N_0/2-20)k_+}\lesssim \varepsilon _1^32^{-3m/4}2^{-(N_0/2-20)k_+}.\nonumber \\ \end{aligned}$$
(6.62)

Expanding the \(d/ds\) derivative we estimate

$$\begin{aligned} B^0(\xi )&lesssim 2^m\sup _{s\in [t_1,t_2]}[B^0_0(\xi ,s)+B^0_1(\xi ,s)+B^0_2(\xi ,s)+B^0_3(\xi ,s)],\\ B^0_0(\xi ,s)&=\int _{\mathbb {R}^2}\left| \frac{\partial _s L(\xi ,s)}{\Phi (\xi ,\eta ,\sigma )}\widehat{f_{k_1}^+}(\xi +\eta ,s)\widehat{f_{k_2}^+}(\xi +\sigma ,s)\right. \\&\quad \times \widehat{f_{k_3}^-} \left. (-\xi -\eta -\sigma ,s)c^*_\xi (\eta ,\sigma )\right| \,d\eta d\sigma ,\\ B^0_1(\xi ,s)&=\left| \,\int _{\mathbb {R}^2}e^{is\Phi (\xi ,\eta ,\sigma )}r_4(\eta ,\sigma )(\partial _s\widehat{f_{k_1}^+})(\xi +\eta ,s)\widehat{f_{k_2}^+}(\xi +\sigma ,s)\right. \\&\quad \times \,\widehat{f_{k_3}^-}\left. (-\xi -\eta -\sigma ,s)c^*_\xi (\eta ,\sigma )\,d\eta d\sigma \right| ,\\ B^0_2(\xi ,s)&=\left| \,\int _{\mathbb {R}^2}e^{is\Phi (\xi ,\eta ,\sigma )}r_4(\eta ,\sigma )\widehat{f_{k_1}^+}(\xi +\eta ,s)(\partial _s\widehat{f_{k_2}^+})(\xi +\sigma ,s)\right. \\&\quad \times \,\widehat{f_{k_3}^-}\left. (-\xi -\eta -\sigma ,s)c^*_\xi (\eta ,\sigma )\,d\eta d\sigma \right| ,\\ B^0_3(\xi ,s)&=\left| \,\int _{\mathbb {R}^2}e^{is\Phi (\xi ,\eta ,\sigma )}r_4(\eta ,\sigma )\widehat{f_{k_1}^+}(\xi +\eta ,s)\widehat{f_{k_2}^+}(\xi +\sigma ,s)\right. \\&\quad \times \,(\partial _s\widehat{f_{k_3}^-})\left. (-\xi -\eta -\sigma ,s)c^*_\xi (\eta ,\sigma )\,d\eta d\sigma \right| . \end{aligned}$$

As before, we combine (6.26), (6.61), and the bounds (6.19) and (6.27) to conclude that

$$\begin{aligned} \sup _{s\in [t_1,t_2]}[B^0_1(\xi ,s)+B^0_2(\xi ,s)+B^0_3(\xi ,s)]\lesssim \varepsilon _1^3 2^{-5m/4} 2^{-(N_0/2-20)k_+}.\nonumber \\ \end{aligned}$$
(6.63)

In addition, using the definition of the function \(L\), we have

$$\begin{aligned} \sup _{s\in [t_1,t_2]}|\partial _s L(\xi ,s)|\lesssim \varepsilon _1^2 2^{-m}. \end{aligned}$$
(6.64)

Therefore

$$\begin{aligned} \sup _{s\in [t_1,t_2]}B^0_0(\xi ,s)\lesssim \varepsilon _1^3 2^{-5m/4} 2^{-(N_0/2-20)k_+}. \end{aligned}$$
(6.65)

The desired bound (6.57) follows from (6.62), (6.63), and (6.65). \(\square \)

6.2 Proof of (6.24)

After changes of variables, it suffices to prove that

$$\begin{aligned}&\left| \,\int _{\mathbb {R}^2\times [t_1,t_2]}e^{iL(\xi ,s)}e^{is\Phi ^{\iota _1,\iota _2,\iota _3}(\xi ,\eta ,\sigma )}\widehat{f_{k_1}^{\iota _1}}(\xi +\eta ,s)\widehat{f_{k_2}^{\iota _2}}(\xi +\sigma ,s) \right. \nonumber \\&\quad \left. \times \widehat{f_{k_3}^{\iota _3}}(-\xi -\eta -\sigma ,s)c^{\iota _1\iota _2\iota _3}(\xi ,-\eta ,-\xi -\eta -\sigma )\,d\eta d\sigma ds\right| \nonumber \\&\quad \quad \lesssim \varepsilon _1^32^{-2p_1m}2^{-(N_1+15)k_+}, \end{aligned}$$
(6.66)

where \((\iota _1,\iota _2,\iota _3)\in \{(+,+,+),(-,-,+),(-,-,-)\}\) and

$$\begin{aligned} \Phi ^{\iota _1,\iota _2,\iota _3}(\xi ,\eta ,\sigma )=\Lambda (\xi )-\iota _1\Lambda (\xi +\eta )-\iota _2\Lambda (\xi +\sigma )-\iota _3\Lambda (\xi +\eta +\sigma ). \end{aligned}$$

The main observation is that the phases \(\Phi ^{\iota _1,\iota _2,\iota _3}\) are weakly elliptic, i.e.

$$\begin{aligned} |\Phi ^{\iota _1,\iota _2,\iota _3}(\xi ,\eta ,\sigma )|\ge 2^{\mathrm {med}(k_1,k_2,k_3)/2-100}, \end{aligned}$$

provided that \(|\xi +\eta |\in [2^{k_1-2},2^{k_1+2}]\), \(|\xi +\sigma |\in [2^{k_2-2},2^{k_2+2}]\), \(|\xi +\eta +\sigma |\in [2^{k_3-2},2^{k_3+2}]\), and \((\iota _1,\iota _2,\iota _3)\in \{(+,+,+),(-,-,+),(-,-,-)\}\). The proof then proceeds as in the proof of Lemma 6.8, using integration by parts in \(s\).

7 Lagrangian formulation and “Wu’s good coordinates”

In this section we review the formulation of the problem used by Wu [45]. We use the same notations as in the cited paper to summarize the construction of Wu’s modified Lagrangian variables. We recall the cubic evolution equations associated to these good unknowns, some other (elliptic) relations between these variables, and the relative cubic energies.

Let \(v\) be the fluid’s velocity field, recall that we denote by \(z: (t,{\alpha }) \in [0,T] \times {\mathbb {R}}\rightarrow {\mathbb {C}}\) the Lagrangian map (restricted to the surface parametrized by \(\alpha \)), that is the solutions of

$$\begin{aligned} z_t (t, {\alpha }) = v (t, z({\alpha },t)) , \quad z(0,{\alpha }) = {\alpha }+ i y_0 ({\alpha }) . \end{aligned}$$
(7.1)

\(z(t,{\alpha }) = x(t,{\alpha }) + i y(t,{\alpha })\) is the equation of the free interface. The imaginary part of \(z\), \(\text{ Im }z\), measures the height of the interface.

7.1 The transformation \(k\) and modified Lagrangian coordinates

Define the change of coordinates \(k\)

$$\begin{aligned} k (t,{\alpha }) := \overline{z} (t,{\alpha }) + \frac{1}{2} (I + {\mathcal {H} } _z) {(I + {\mathcal {K} } _z)}^{-1} (z (t,{\alpha }) - \overline{z} (t,{\alpha }) ) \end{aligned}$$
(7.2)

where \( {\mathcal {H} } _{\gamma }\) denotes the Hilbert transform along the curve \({\gamma }\), see (1.4), and \( {\mathcal {K} } _z = \text{ Re } {\mathcal {H} } _z\). The explicit change of coordinates (7.2) is the same as the one given by Totz and Wu in [41, formula (2.3)]. \(k\) will be shown to be a diffeomorphism on \({\mathbb {R}}\). Given the change of coordinates \(k\) we can define the transformed Lagrangian unknowns as in [45]

$$\begin{aligned} \widetilde{L} (t,{\alpha }) := ( {\zeta }_{\alpha }(t,{\alpha }) - 1, u(t,{\alpha }) , w(t,{\alpha }), \text{ Im }{\zeta }(t,{\alpha }) ) \end{aligned}$$

with

$$\begin{aligned} {\zeta }(t,{\alpha })&:= z (t,k^{-1}(t,{\alpha })) , \ u(t,{\alpha }) := z_t (t,k^{-1}(t,{\alpha })), \nonumber \\ \ w(t,{\alpha })&:= z_{tt} (t,k^{-1}(t,{\alpha })) . \end{aligned}$$
(7.3)

Note that

$$\begin{aligned} w(t,{\alpha })&:= z_{tt} (t,k^{-1}(t,{\alpha })) = (\partial _t + b (t,{\alpha }) \partial _{\alpha }) u(t,{\alpha }) , \end{aligned}$$

where \(b (t,{\alpha }) = k_t (t, k^{-1}(t,{\alpha }))\). Also, following [45] we define the “good quantities” in terms of the transformed Lagrangian unknowns

$$\begin{aligned} \chi&:= 2i(I- {\mathcal {H} } _{\zeta }) \text{ Im }{\zeta }= (I- {\mathcal {H} } _z)(z-\overline{z}) \circ k^{-1}\end{aligned}$$
(7.4)
$$\begin{aligned} {\lambda }&:= (I- {\mathcal {H} } _{\zeta }) ( \psi \circ k^{-1} ) = (I- {\mathcal {H} } _z)\psi \circ k^{-1}\end{aligned}$$
(7.5)
$$\begin{aligned} v&:= \partial _t (I- {\mathcal {H} } _z)(z-\overline{z}) \circ k^{-1} = (\partial _t + b \partial _{\alpha })\chi , \quad v_1 := (I- {\mathcal {H} } _{\zeta }) v . \end{aligned}$$
(7.6)

In what follows we will denote \( {\mathcal {H} } \) for \( {\mathcal {H} } _{\zeta }\), where this creates no confusion. We recall that \(\psi \) is the trace of the velocity potential in Lagrangian coordinates: \(\psi (t,{\alpha }) = \Phi (t , z(t,{\alpha }))\). The quantities \(\chi ,{\lambda }\) and \(v\) are those for which cubic equations are derived and the Energy argument is performed. The relation between the Eulerian trace of the velocity potential \(\phi \) and the surface elevation \(h\) with the Lagrangian quantities is given by the identities:

$$\begin{aligned} h(t, \text{ Re }z(t,{\alpha }))&= \text{ Im }z (t,{\alpha }) \end{aligned}$$
(7.7)
$$\begin{aligned} \phi (t, \text{ Re }z(t,{\alpha }))&= \psi (t,{\alpha }) . \end{aligned}$$
(7.8)

Precomposing with \(k^{-1}\) these become

$$\begin{aligned} h(t, \text{ Re }{\zeta }(t,{\alpha })) = \text{ Im }{\zeta }(t,{\alpha }) , \quad \phi (t, \text{ Re }{\zeta }(t,{\alpha }))&= \psi (t, k^{-1}(t,{\alpha })) . \end{aligned}$$

7.2 The cubic equations

In [45] cubic equations are derived for the “good quantities” \(\chi , {\lambda }\) and \(v\) given by (7.4)–(7.6). We will not discuss the derivation of these equations here, but refer the reader to section 2 in [45] for the details. In the cited work it is shown that for \(F = \chi , {\lambda }\) or \(v\)

$$\begin{aligned} {(\partial _t + b \partial _{\alpha })}^2 F - i A \partial _{\alpha }F = G(\widetilde{L}) \end{aligned}$$
(7.9)

where the right-hand side \(G(\widetilde{L})\) can be thought of as a cubic expression in the variables

$$\begin{aligned} \widetilde{L} = (u,w, \text{ Im }{\zeta }, {\zeta }_{\alpha }- 1) , \end{aligned}$$
(7.10)

which involves singular integrals related to the Cauchy integral and to Calderón commutators. The functions \(b\) and \(A\) are defined by

$$\begin{aligned} b (t,{\alpha })&= k_t \left( t , k^{-1}(t,{\alpha }) \right) \end{aligned}$$
(7.11)
$$\begin{aligned} A (t, {\alpha })&= (a(t, \cdot ) k_{\alpha }(t, \cdot ) ) \circ k^{-1}(t,{\alpha }) \end{aligned}$$
(7.12)

where \(a(t,{\alpha })\) is the Rayleigh-Taylor coefficient appearing in (1.5). As shown in [45, Proposition 2.1], both \(b\) and \(A-1\) are real valued and quadratically small if \(\widetilde{L}\) is small. More precisely:

Lemma 7.1

Let \(b\) and \(A\) be given by (7.11) and (7.12). Under the a priori assumptions (2.9) we have

$$\begin{aligned} {\Vert A(t)-1 \Vert }_{H^{N_1}} + {\Vert b(t) \Vert }_{H^{N_1}}&\le {\varepsilon }_1 , \end{aligned}$$
(7.13)

and for \(0\le k\le N_0\)

$$\begin{aligned} {\Vert A(t)-1 \Vert }_{X_k}&\lesssim {\varepsilon }_1 {\Vert \widetilde{L}(t) \Vert }_{X_k} ,\end{aligned}$$
(7.14)
$$\begin{aligned} {\Vert b(t) \Vert }_{X_k}&\lesssim {\varepsilon }_1 {\Vert \widetilde{L}(t) \Vert }_{X_k} . \end{aligned}$$
(7.15)

Proof

The identities in (2.30) of [45] (see Proposition 2.4 in [45] for their derivation) read

$$\begin{aligned} (I- {\mathcal {H} } )(A-1)&= i [u, {\mathcal {H} } ] \frac{\overline{u}_{\alpha }}{{\zeta }_{\alpha }} + i [w, {\mathcal {H} } ] \frac{\overline{{\zeta }}_{\alpha }- 1}{{\zeta }_{\alpha }}\end{aligned}$$
(7.16)
$$\begin{aligned} (I- {\mathcal {H} } )b&= - [u, {\mathcal {H} } ] \frac{\overline{{\zeta }}_{\alpha }- 1}{{\zeta }_{\alpha }} . \end{aligned}$$
(7.17)

The above right-hand sides are all of the form \(Q_0(\widetilde{L}, \widetilde{L})\) or \(Q_0(\widetilde{L}, \partial _{\alpha }\widetilde{L})\), where \(Q_0\) is defined in (11.18). (7.16) and (7.17) then follow from the bounds for operators of the type \(Q_0\) given in (11.30), Sobolev’s embedding, the estimate (11.3) for \( {\mathcal {H} } \), and the a priori smallness assumptions on \(\widetilde{L}\). The bounds (7.14) and (7.15) follow by combining (7.16) and (7.17) with Lemma 11.2. \(\square \)

The cubic equations verified by \(\chi , {\lambda }\) and \(v\) are (see formulas (2.27)-(2.29) and (4.33) in [45])

$$\begin{aligned} {(\partial _t + b \partial _{\alpha })}^2 \chi - i A \partial _\alpha \chi&= G^\chi (u,\text{ Im }{\zeta })\end{aligned}$$
(7.18)
$$\begin{aligned} {(\partial _t + b \partial _{\alpha })}^2 v - i A \partial _\alpha v&= G^v (u,w,{\zeta }_{\alpha }- 1,\text{ Im }{\zeta },\chi _{\alpha })\end{aligned}$$
(7.19)
$$\begin{aligned} {(\partial _t + b \partial _{\alpha })}^2 {\lambda }- i A \partial _\alpha {\lambda }&= G^{\lambda }(u,w, \text{ Im }{\zeta }) \end{aligned}$$
(7.20)

where

$$\begin{aligned} G^\chi (u,\text{ Im }{\zeta })&:= \frac{4}{\pi } \int \frac{ (u({\alpha }) - u({\beta })) (\text{ Im }{\zeta }({\alpha }) - \text{ Im }{\zeta }({\beta })) }{ {|{\zeta }({\alpha })-{\zeta }({\beta })|}^2 } u_{\beta }({\beta }) \, d{\beta }\nonumber \\&\quad + \frac{2}{\pi } \int { \left( \frac{ u({\alpha }) - u({\beta }) }{ {\zeta }({\alpha })-{\zeta }({\beta }) } \right) }^2 \text{ Im }{\zeta }_{\beta }({\beta }) \, d{\beta }, \end{aligned}$$
(7.21)
$$\begin{aligned}&G^v (u,w,{\zeta }_{\alpha }- 1,\text{ Im }{\zeta }, \chi _{\alpha }) \nonumber \\&\quad :=\frac{4}{\pi } \int \frac{ (w({\alpha }) - w({\beta })) (\text{ Im }{\zeta }({\alpha }) - \text{ Im }{\zeta }({\beta })) }{ {|{\zeta }({\alpha })-{\zeta }({\beta })|}^2 } u_{\beta }({\beta }) \, d{\beta }\nonumber \\&\quad \quad + \frac{4}{\pi } \int \frac{ (u({\alpha }) - u({\beta })) (\text{ Im }{\zeta }({\alpha }) - \text{ Im }{\zeta }({\beta })) }{ {|{\zeta }({\alpha })-{\zeta }({\beta })|}^2 } w_{\beta }({\beta }) \, d{\beta }\nonumber \\&\quad \quad + \frac{2}{i\pi } \int { \left( \frac{ u({\alpha }) - u({\beta }) }{ {\zeta }({\alpha })-{\zeta }({\beta }) } \right) }^2 u_{\beta }({\beta }) \, d{\beta }- \frac{2}{i\pi } \int \frac{ {|u({\alpha }) - u({\beta })|}^2 }{ {(\overline{{\zeta }}({\alpha })- \overline{{\zeta }}({\beta }))}^2 } u_{\beta }({\beta }) \, d{\beta }\nonumber \\&\quad \quad + \frac{2}{i\pi } \int \frac{ (u({\alpha }) - u({\beta })) (w({\alpha }) - w({\beta })) }{ {({\zeta }({\alpha })-{\zeta }({\beta }))}^2 } ({\zeta }_{\beta }({\beta }) - \overline{{\zeta }}_{\beta }({\beta })) \, d{\beta }\nonumber \\&\quad \quad + \frac{1}{i\pi } \int { \left( \frac{ u({\alpha }) - u({\beta }) }{ {\zeta }({\alpha })-{\zeta }({\beta }) } \right) }^2 (u_{\beta }({\beta }) - \overline{u}_{\beta }({\beta })) \, d{\beta }\nonumber \\&\quad \quad - \frac{2}{i\pi } \int { \left( \frac{ u({\alpha }) - u({\beta }) }{ {\zeta }({\alpha })-{\zeta }({\beta }) } \right) }^3 ({\zeta }_{\beta }({\beta }) - \overline{{\zeta }}_{\beta }({\beta })) \, d{\beta }+ i \frac{a_t}{a} \circ k^{-1} A \partial _{\alpha }\chi , \end{aligned}$$
(7.22)
$$\begin{aligned} G^{\lambda }(u,w,\text{ Im }{\zeta })&:= \frac{2}{\pi } \int \frac{ (u({\alpha }) - u({\beta })) (\text{ Im }{\zeta }({\alpha }) - \text{ Im }{\zeta }({\beta })) }{ {|{\zeta }({\alpha })-{\zeta }({\beta })|}^2 } \overline{{\zeta }}_{\beta }({\beta }) w({\beta }) \, d{\beta }\nonumber \\&\quad + [u,\overline{ {\mathcal {H} } }] \left( \overline{u} \frac{u_{\alpha }}{\overline{{\zeta }}_{\alpha }} \right) + u [u, {\mathcal {H} } ] \frac{\overline{u}_{\alpha }}{{\zeta }_{\alpha }} - 2 [u, {\mathcal {H} } ] \frac{u \cdot u_{\alpha }}{{\zeta }_{\alpha }} \nonumber \\&\quad + \frac{1}{i\pi } \int { \left( \frac{ u({\alpha }) - u({\beta }) }{ {\zeta }({\alpha })-{\zeta }({\beta }) } \right) }^2 u ({\beta }) \cdot {\zeta }_{\beta }({\beta }) \, d{\beta }. \end{aligned}$$
(7.23)

A fourth additional equation is also derived in [45, formula (4.32)] for the quantity

$$\begin{aligned} v_1 := (I- {\mathcal {H} } _{\zeta }) v \end{aligned}$$
(7.24)

and has the same form as above

$$\begin{aligned}&{(\partial _t + b \partial _{\alpha })}^2 v_1 - i A \partial _\alpha v_1 = G^{v_1} (u,w,{\zeta }_{\alpha }- 1, \text{ Im }{\zeta }, v, \chi ) \end{aligned}$$
(7.25)

with

$$\begin{aligned}&G^{v_1} \left( u, w, {\zeta }_{\alpha }- 1, \text{ Im }{\zeta }, v, \chi \right) \nonumber \\&\quad : = (I- {\mathcal {H} } ) {\mathcal {P}}v - 2 [u, {\mathcal {H} } ] \frac{\partial _{\alpha }}{{\zeta }_{\alpha }} {\mathcal {P}}\chi - 2[u, {\mathcal {H} } ] \frac{\partial _{\alpha }}{{\zeta }_{\alpha }} \left( w \frac{\partial _{\alpha }}{{\zeta }_{\alpha }} \chi \right) \nonumber \\&\quad \quad \,-i [( {\mathcal {H} } + \overline{ {\mathcal {H} } })u, {\mathcal {H} } ] {\left( \frac{\partial _{\alpha }}{{\zeta }_{\alpha }} \right) }^2 \chi + \frac{1}{i\pi } \int { \left( \frac{ u({\alpha }) - u({\beta }) }{ {\zeta }({\alpha })-{\zeta }({\beta }) } \right) }^2 v_{\beta }({\beta }) \, d{\beta }, \end{aligned}$$
(7.26)

where \({\mathcal {P}}:= {(\partial _t + b \partial _{\alpha })}^2 - i A \partial _\alpha \). The above cubic equations allow the construction of energy functionals, see Sect. 7.3 below, which suitably control the variables \(\chi ,v\) and \({\lambda }\). The quantities appearing on the right-hand sides of these cubic equations, such as \(u,w\) and \({\zeta }_{\alpha }-1\), are in turned controlled by the energies, see Proposition 8.1, using the identities (8.24)–(8.27).

In order to simplify our presentation and the estimates performed on the above equations, we define the following types of trilinear operators:

$$\begin{aligned} T^1 (f,g,h)&:= \int \frac{ (f({\alpha }) - f({\beta })) (g({\alpha }) - g({\beta })) }{ {|{\zeta }({\alpha })-{\zeta }({\beta })|}^2 } h ({\beta }) \, d{\beta }\end{aligned}$$
(7.27a)
$$\begin{aligned} T^2 (f,g,h)&:= \int \frac{ (f({\alpha }) - f({\beta })) (g({\alpha }) - g({\beta })) }{ {({\zeta }({\alpha })-{\zeta }({\beta }))}^2 } h ({\beta }) \, d{\beta }\end{aligned}$$
(7.27b)
$$\begin{aligned} T^3 (f,g,h)&:= \int \frac{ (f({\alpha }) - f({\beta })) (g({\alpha }) - g({\beta })) }{ {(\overline{{\zeta }}({\alpha })-\overline{{\zeta }}({\beta }))}^2 } h ({\beta }) \, d{\beta }, \end{aligned}$$
(7.27c)

and denote by \(\mathbf{T}\) any scalar multiple of them:

$$\begin{aligned} \mathbf{T}(f,g,h) = c_i T_i (f,g,h) \, \end{aligned}$$
(7.28)

for some constant \(c_i \in {\mathbb {C}}\), \(i=1,2,3\). We can then write the nonlinearity (7.21) appearing in the equation (7.18) as:

$$\begin{aligned} G^\chi (u,\text{ Im }{\zeta }) = \mathbf{T}(u,\text{ Im }{\zeta }, u_{\alpha }) + \mathbf{T}(u,u,\text{ Im }{\zeta }_{\alpha }) . \end{aligned}$$
(7.29)

Similarly we have

$$\begin{aligned}&G^v (u,w,{\zeta }_{\alpha }- 1,\text{ Im }{\zeta }) \nonumber \\&\quad = \mathbf{T}(w,\text{ Im }{\zeta }, u_{\alpha }) + \mathbf{T}(u,\text{ Im }{\zeta }, w_{\alpha }) + \mathbf{T}(u,u, u_{\alpha }) + \mathbf{T}(u,w,\text{ Im }{\zeta }_{\alpha }) \nonumber \\&\quad \quad \,+ \mathbf{T}(u,u, \overline{u}_{\alpha }) + u \mathbf{T}(u,u, \text{ Im }{\zeta }_{\alpha }) + \mathbf{T}(u,u, u \text{ Im }{\zeta }_{\alpha }) + i \frac{a_t}{a} \circ k^{-1} A \partial _{\alpha }\chi , \end{aligned}$$
(7.30)

and

$$\begin{aligned} G^{\lambda }(u,w, \text{ Im }{\zeta })&= \mathbf{T}(u,\text{ Im }{\zeta }, \overline{{\zeta }}_{\alpha }w) + \mathbf{T}(u,u, u \cdot {\zeta }_{\alpha }) \nonumber \\&\quad + [u,\overline{ {\mathcal {H} } }] \left( \overline{u} \frac{u_{\alpha }}{\overline{{\zeta }}_{\alpha }} \right) + u [u, {\mathcal {H} } ] \frac{\overline{u}_{\alpha }}{{\zeta }_{\alpha }} - 2 [u, {\mathcal {H} } ] \frac{u \cdot u_{\alpha }}{{\zeta }_{\alpha }} , \end{aligned}$$
(7.31)

and

$$\begin{aligned}&G^{v_1} \left( u, w, {\zeta }_{\alpha }- 1, \text{ Im }{\zeta }, v, \chi \right) \nonumber \\&\quad : = (I- {\mathcal {H} } ) {\mathcal {P}}v - 2 [u, {\mathcal {H} } ] \frac{\partial _{\alpha }}{{\zeta }_{\alpha }} {\mathcal {P}}\chi - 2[u, {\mathcal {H} } ] \frac{\partial _{\alpha }}{{\zeta }_{\alpha }} \left( w \frac{\partial _{\alpha }}{{\zeta }_{\alpha }} \chi \right) \nonumber \\&\quad \quad \,-i [( {\mathcal {H} } + \overline{ {\mathcal {H} } })u, {\mathcal {H} } ] {\left( \frac{\partial _{\alpha }}{{\zeta }_{\alpha }} \right) }^2 \chi + \mathbf{T}(u,u, v_{\alpha }) . \end{aligned}$$
(7.32)

By writing the nonlinearities (7.21), (7.22), (7.23) and (7.26) in terms of operators of the type (7.27) above, we will be able to efficiently estimate them by making use of a general Proposition giving \(L^2\)-type bounds of such operators. These estimates are given in Proposition 11.7, and are obtained by improving two statements contained in [45] (which in turn rely in part on the work of Coifman, McIntosh and Meyer [10]). All the terms that cannot be written as a trilinear operators of the form (7.27), acting on the components of \(\widetilde{L}\), need to be treated separately. To bound these remaining nonlinear terms, namely

$$\begin{aligned}&i \frac{a_t}{a} \circ k^{-1} A \partial _{\alpha }\chi , \quad [u,\overline{ {\mathcal {H} } }] \left( \overline{u} \frac{u_{\alpha }}{\overline{{\zeta }}_{\alpha }} \right) + u [u, {\mathcal {H} } ] \frac{\overline{u}_{\alpha }}{{\zeta }_{\alpha }} , \quad [u, {\mathcal {H} } ] \frac{u \cdot u_{\alpha }}{{\zeta }_{\alpha }} \\&[( {\mathcal {H} } + \overline{ {\mathcal {H} } })u, {\mathcal {H} } ] {\left( \frac{\partial _{\alpha }}{{\zeta }_{\alpha }} \right) }^2 \chi , \quad [u, {\mathcal {H} } ] \frac{\partial _{\alpha }}{{\zeta }_{\alpha }} \left( w \frac{\partial _{\alpha }}{{\zeta }_{\alpha }} \chi \right) \end{aligned}$$

we will make use of some additional special structure present in them.

7.3 The energy

The total energy for the system is given by the sum of three energies naturally associated to the equations (7.18), (7.20) and (7.25). Let us define

$$\begin{aligned} S_k := D^{k} S . \end{aligned}$$
(7.33)

The first term in the energy is given by

$$\begin{aligned} E^\chi (t)&= \sum _{k=0}^{N_0} \int _{\mathbb {R}}\frac{1}{A} {\left| (\partial _t + b \partial _{\alpha }) D^k \chi \right| }^2 + i {\left( D^k \chi \right) }^h \partial _\alpha \overline{{\left( D^k \chi \right) }^h} \, d{\alpha }\nonumber \\&\quad \,+ \sum _{k=0}^{N_0/2} \int _{\mathbb {R}}\frac{1}{A} {\left| (\partial _t + b \partial _{\alpha }) S_k \chi \right| }^2 + i {\left( S_k \chi \right) }^h \partial _\alpha \overline{{\left( S_k \chi \right) }^h} \, d{\alpha }, \end{aligned}$$
(7.34)

where \(f^h\) denotes the anti-holomorphic part of a function \(f\):

$$\begin{aligned} f^h := \frac{I - {\mathcal {H} } }{2} f . \end{aligned}$$
(7.35)

By considering only the anti-holomorphic parts of \(D^k \chi \) and \(S_k \chi \) in (7.34), one obtains that all summands in \(E^\chi \) are non-negative, see Lemma (8.7). Similarly one constructs the energy associated to (7.20):

$$\begin{aligned} E^{\lambda }(t)&= \sum _{k=0}^{N_0-2} \int _{\mathbb {R}}\frac{1}{A} {\left| (\partial _t + b \partial _{\alpha }) D^k {\lambda }\right| }^2 + i {\left( D^k {\lambda }\right) }^h \partial _\alpha \overline{{\left( D^k{\lambda }\right) }^h} \, d{\alpha }\nonumber \\&\quad \,+ \sum _{k=0}^{N_0/2} \int _{\mathbb {R}}\frac{1}{A} {\left| (\partial _t + b \partial _{\alpha }) S_k {\lambda }\right| }^2 + i {\left( S_k {\lambda }\right) }^h \partial _\alpha \overline{{\left( S_k {\lambda }\right) }^h} \, d{\alpha }, \end{aligned}$$
(7.36)

The energy controlling \(v\) instead is given in terms of \(v_1\), and based on the equation (7.25):

$$\begin{aligned} E^{v} (t)&= \sum _{k=0}^{N_0} \int _{\mathbb {R}}\frac{1}{A} {\left| (\partial _t + b \partial _{\alpha }) D^k v_1 \right| }^2 + i D^k v_1 \partial _\alpha D^k \overline{v}_1 \, d{\alpha }\nonumber \\&\quad \,+ \sum _{k=0}^{N_0/2} \int _{\mathbb {R}}\frac{1}{A} {\left| (\partial _t + b \partial _{\alpha }) S_k v_1 \right| }^2 + i S_k v_1 \partial _\alpha S_k \overline{v}_1 \, d{\alpha }. \end{aligned}$$
(7.37)

Here in the second summand there is no restriction to \((D^k v_1)^h\) or \((S v_1)^h\). Therefore the Energy \(E^v\) has no definite sign. Nevertheless, it can be shown that this Energy controls the norms of \(v\) and \(v_1\) up to cubic lower order contributions. The total energy is then given by

$$\begin{aligned} E(t) = E^{\chi }(t) + E^{v}(t) + E^{{\lambda }}(t) . \end{aligned}$$
(7.38)

8 Proof of Proposition 2.3: energy estimates

Once the energy \(E(t)\) has been defined, we can proceed with the proof of Proposition 2.3. The main ideas are essentially the same as those used in Wu’s paper [45]. In fact, we use a similar procedure and borrow several identities and estimates from this paper. However, some arguments there need to be adjusted in order to make the energy estimates valid for all times, and compatible with the growth of the highest Sobolev and weighted norms. In particular, we need to pay special attention to certain dangerous nonlinear terms that can potentially create logarithmic losses, compare with Wu’s energy estimate (1.15), and show how these losses can be avoided.

Recall that we denoted \(\widetilde{L} = (u,w, \text{ Im }{\zeta }, {\zeta }_{\alpha }- 1)\), and define the vectors

$$\begin{aligned} L^- := ({\zeta }_{\alpha }- 1, u, w, \partial _{\alpha }\chi , v) , \quad L := (L^-, \text{ Im }{\zeta }) . \end{aligned}$$

We separate the proof of Proposition 2.3 into three main steps. We first show how \(E(t)\) controls the \(X_{N_0}\)-norm of \(\widetilde{L}(t)\):

Proposition 8.1

Under the a priori assumptions (2.9), that is

$$\begin{aligned}&\sup _{t \in [0,T]} \left( (1+t)^{-p_0} {\Vert \widetilde{L}(t) \Vert }_{X_{N_0}} + {\Vert \widetilde{L}(t) \Vert }_{H^{N_1+5}} + {\Vert \widetilde{L}(t) \Vert }_{W^{N_1,\infty }}\sqrt{1+t} \right) \nonumber \\&\quad \le {\varepsilon }_1 \ll 1 , \end{aligned}$$
(8.1)

we have

$$\begin{aligned} {\Vert \widetilde{L} (t) \Vert }_{X_{N_0}} \lesssim \sqrt{E(t)} , \end{aligned}$$
(8.2)

for any \(t\in [0,T]\) and \({\varepsilon }_1\) sufficiently small.

This is proven in Sect. 8.1. Our next proposition shows how to estimate the energy increment.

Proposition 8.2

Assume again that (8.1) holds for \({\varepsilon }_1\) small enough. Then

$$\begin{aligned} \frac{d}{dt} \sqrt{E(t)} \lesssim \left( {\Vert L (t) \Vert }_{W^{N_1,\infty }} + {\Vert {\mathcal {H} } L^- (t) \Vert }_{W^{N_1,\infty }} \right) ^2 \sqrt{E(t)} , \end{aligned}$$
(8.3)

for any \(t\in [0,T]\).

The above Proposition will be proven in Sect. 8.2. We eventually need to bound the \(L^\infty \)-norms on the right hand side of (8.3) in terms of the \(Z^{\prime }\)-norm of \(h\) and \(\phi \):

Proposition 8.3

Under the a priori assumption (8.1) and (2.8) we have

$$\begin{aligned} {\Vert L (t) \Vert }_{W^{N_1,\infty }} \!+\! {\Vert {\mathcal {H} } L^- (t) \Vert }_{W^{N_1,\infty }} \lesssim {\Vert (h,\phi ) \Vert }_{Z^{\prime }} \!=\! {\Vert (h, \Lambda \phi ) \Vert }_{W^{N_1+4,\infty }} . \end{aligned}$$
(8.4)

The proof of this Proposition is in Sect. 8.3.

In view of the initial assumptions (1.10a)–(1.10b) and the discussion in [45, sec 5.1], one has \(E(0) \lesssim {\varepsilon }_0^2\). It is then clear that Propositions 8.1–8.3 imply Proposition 2.3.

8.1 Proof of Proposition 8.1: energy bounds

The estimate (8.2) will be achieved through a sequence of Lemmas which we state below and prove in the remaining of this section. We start by using some formulae derived in [45] to bound \({\Vert \widetilde{L}(t) \Vert }_{X_{N_0}}\) by the \(X_{N_0}\)-norms of \((\partial _t + b \partial _{\alpha }) \chi \), \((\partial _t + b \partial _{\alpha }) v\) and \((\partial _t + b \partial _{\alpha }) {\lambda }\):

Lemma 8.4

Under the assumption (8.1) it is possible to write

$$\begin{aligned} 2 u&= (\partial _t + b \partial _{\alpha }) \chi + Q \end{aligned}$$
(8.5)
$$\begin{aligned} 2 w&= (\partial _t + b \partial _{\alpha }) v + Q \end{aligned}$$
(8.6)
$$\begin{aligned} 2 ({\zeta }_{\alpha }- 1)&= -i (\partial _t + b \partial _{\alpha }) v + Q \end{aligned}$$
(8.7)
$$\begin{aligned} (I- {\mathcal {H} } ) \text{ Im }{\zeta }&= - (\partial _t + b \partial _{\alpha }) {\lambda }+ Q \end{aligned}$$
(8.8)

where \(Q\) denotes a generic term which is at least quadratic in \(\widetilde{L}\) and satisfies

$$\begin{aligned} {\Vert Q(t) \Vert }_{X_{N_0}}&\lesssim {\varepsilon }_1 {\Vert \widetilde{L}(t) \Vert }_{X_{N_0}} \end{aligned}$$
(8.9)
$$\begin{aligned} {\Vert Q(t) \Vert }_{W^{N_1,\infty }}&\lesssim {\varepsilon }_1 {\Vert \widetilde{L}(t) \Vert }_{W^{N_1,\infty }} \end{aligned}$$
(8.10)

for \({\varepsilon }_1\) small enough. In particular we see that

$$\begin{aligned} {\Vert \widetilde{L}(t) \Vert }_{X_{N_0}} \sim {\Vert (\partial _t + b\partial _{\alpha }) \chi \Vert }_{X_{N_0}} + {\Vert (\partial _t + b\partial _{\alpha }) v \Vert }_{X_{N_0}} + {\Vert (\partial _t + b\partial _{\alpha }) {\lambda }\Vert }_{X_{N_0}} \end{aligned}$$
(8.11)

and

$$\begin{aligned}&{\Vert \widetilde{L}(t) \Vert }_{W^{N_1,\infty }} \lesssim {\Vert (\partial _t + b\partial _{\alpha }) \chi \Vert }_{W^{N_1,\infty }} + {\Vert (\partial _t + b\partial _{\alpha }) v \Vert }_{W^{N_1,\infty }}\nonumber \\&\quad + {\Vert (\partial _t + b\partial _{\alpha }) {\lambda }\Vert }_{W^{N_1,\infty }} . \end{aligned}$$
(8.12)

The proof of this Lemma is given in Sect. 8.1.1. We then establish the following commutators estimates:

Lemma 8.5

Assume again (8.1) holds for \({\varepsilon }_1\) small enough. Then, for \(Q\) as in (8.9) above, we have

$$\begin{aligned} \partial _{\alpha }\chi&= (I- {\mathcal {H} } ) ({\zeta }_{\alpha }- \overline{{\zeta }}_{\alpha }) + Q \end{aligned}$$
(8.13)
$$\begin{aligned} v&= 2u + Q \end{aligned}$$
(8.14)
$$\begin{aligned} \partial _{\alpha }{\lambda }&= u + Q . \end{aligned}$$
(8.15)

Moreover, denoting \(D\) for \(\partial _{\alpha }\), we have for all \(0\le k \le N_0\) and \(f = \chi ,v\) or \({\lambda }\)

$$\begin{aligned}&{\Vert (\partial _t + b\partial _{\alpha }) D^k f - D^k (\partial _t + b\partial _{\alpha }) f \Vert }_{L^2} \lesssim {\varepsilon }_1 {\Vert \widetilde{L}(t) \Vert }_{X_{N_0}} , \end{aligned}$$
(8.16)

and for all \(0\le k \le \frac{N_0}{2}\)

$$\begin{aligned}&{\Vert (\partial _t \!+\! b\partial _{\alpha }) D^k S f \!-\! D^k S (\partial _t \!+\! b\partial _{\alpha }) f \Vert }_{L^2} \lesssim {\varepsilon }_1 {\Vert \widetilde{L}(t) \Vert }_{X_{N_0}} \!+\! {\Vert (\partial _t + b\partial _{\alpha }) f \Vert }_{H^k} . \end{aligned}$$
(8.17)

The above Lemma is proven in Sect. 8.1.2. Eventually we show how \(E(t)\) controls \(\chi \) and \(v\):

Lemma 8.6

Under the assumption (8.1), we have

$$\begin{aligned} {\Vert (\partial _t + b\partial _{\alpha }) D^k \chi \Vert }_{L^2}^2&\lesssim E(t) , \end{aligned}$$
(8.18)
$$\begin{aligned} {\Vert (\partial _t + b\partial _{\alpha }) D^k v \Vert }_{L^2}^2&\lesssim E(t) + {\varepsilon }_1 {\Vert \widetilde{L}(t) \Vert }_{H^{N_0}}^2 \end{aligned}$$
(8.19)

for all \(0 \le k \le N_0\). Also

$$\begin{aligned}&{\Vert (\partial _t + b \partial _{\alpha }) D^k {\lambda }\Vert }^2_{L^2} \lesssim E(t) , \end{aligned}$$
(8.20)

for all \(0 \le k \le N_0-2\). Moreover, for \(0 \le k \le \frac{N_0}{2}\)

$$\begin{aligned} {\Vert (\partial _t + b\partial _{\alpha }) D^k S \chi \Vert }_{L^2}^2&\lesssim E(t) , \end{aligned}$$
(8.21)
$$\begin{aligned} {\Vert (\partial _t + b\partial _{\alpha }) D^k S v \Vert }_{L^2}^2&\lesssim E(t) + {\varepsilon }_1 {\Vert \widetilde{L}(t) \Vert }_{X_{N_0}}^2 , \end{aligned}$$
(8.22)
$$\begin{aligned} {\Vert (\partial _t + b \partial _{\alpha }) D^k S {\lambda }\Vert }^2_{L^2}&\lesssim E(t) . \end{aligned}$$
(8.23)

This Lemma is proven in 8.1.3. One can actually obtain the stronger bound

$$\begin{aligned} {\Vert (\partial _t + b\partial _{\alpha }) {\Gamma }v \Vert }_{L^2}^2 + {\Vert (\partial _t + b\partial _{\alpha }) {\Gamma }v_1 \Vert }_{L^2}^2 \lesssim E(t) + {\varepsilon }_1 {\Vert \widetilde{L}(t) \Vert }_{X_{N_0}}^2 . \end{aligned}$$

The bound (8.2), and hence Proposition 8.1, follow easily from the above three lemmas.

8.1.1 Proof of Lemma 8.4

Equations (2.35), (2.43), (3.38) and (2.51) in [45] respectively read

$$\begin{aligned}&2 u = (\partial _t + b \partial _{\alpha }) \chi + ( {\mathcal {H} } + \overline{ {\mathcal {H} } }) u + [u, {\mathcal {H} } ] \frac{{\zeta }_{\alpha }- \overline{{\zeta }}_{\alpha }}{{\zeta }_{\alpha }}\end{aligned}$$
(8.24)
$$\begin{aligned}&2 w = (\partial _t + b \partial _{\alpha }) v + [\overline{u}, \overline{ {\mathcal {H} } }] \frac{u_{\alpha }}{\overline{{\zeta }}_{\alpha }} +[u, {\mathcal {H} } ] \frac{2 u_{\alpha }- \overline{u}_{\alpha }}{{\zeta }_{\alpha }} + ( {\mathcal {H} } + \overline{ {\mathcal {H} } }) w\nonumber \\&\quad \,+ [w, {\mathcal {H} } ] \frac{{\zeta }_{\alpha }- \overline{{\zeta }}_{\alpha }}{{\zeta }_{\alpha }}- \frac{1}{i \pi } \int {\left( \frac{u({\alpha }) - u({\beta })}{{\zeta }({\alpha }) - {\zeta }(b)} \right) }^2 ({\zeta }_{\beta }({\beta }) - \overline{{\zeta }}_{\beta }({\beta })) \, d{\beta }\,\end{aligned}$$
(8.25)
$$\begin{aligned}&{\zeta }_{\alpha }- 1 = \frac{w}{i A} - \frac{A - 1}{A}\end{aligned}$$
(8.26)
$$\begin{aligned}&(I- {\mathcal {H} } ) \text{ Im }{\zeta }= - (\partial _t + b \partial _{\alpha }) {\lambda }-\frac{1}{2} [u, {\mathcal {H} } ] \frac{\overline{{\zeta }}_{\alpha }u}{{\zeta }_{\alpha }} . \end{aligned}$$
(8.27)

Proof of (8.5). To show (8.5) it is enough to prove that \(( {\mathcal {H} } + \overline{ {\mathcal {H} } }) u\) and \([u, {\mathcal {H} } ] \frac{{\zeta }_{\alpha }- \overline{{\zeta }}_{\alpha }}{{\zeta }_{\alpha }}\) are quadratic terms satisfying (8.9). Estimate (11.4) and the a priori assumption (8.1) give

$$\begin{aligned}&{\Vert ( {\mathcal {H} } + \overline{ {\mathcal {H} } }) u \Vert }_{X_{N_0}} \lesssim {\Vert {\zeta }_{\alpha }- 1 \Vert }_{H^{\frac{N_0}{2} + 1}} {\Vert u \Vert }_{X_{N_0}}\\&\quad + {\Vert {\zeta }_{\alpha }- 1 \Vert }_{X_{N_0}} {\Vert u \Vert }_{H^{\frac{N_0}{2} + 1}} \lesssim {\varepsilon }_1 {\Vert \widetilde{L} \Vert }_{X_{N_0}} . \end{aligned}$$

Similarly, under the a priori assumptions (8.1), estimate (11.31) implies

$$\begin{aligned} {\left\| [u, {\mathcal {H} } ] \frac{{\zeta }_{\alpha }- \overline{{\zeta }}_{\alpha }}{{\zeta }_{\alpha }} \right\| }_{X_{N_0}}&\le {\varepsilon }_1 {\Vert u \Vert }_{X_{N_0}} + {\varepsilon }_1 {\Vert \text{ Im }{\zeta }\Vert }_{X_{N_0}} \\ + {\varepsilon }_1 {\Vert {\zeta }_{\alpha }- 1 \Vert }_{X_{N_0}}&\lesssim {\varepsilon }_1 {\Vert \widetilde{L} \Vert }_{X_{N_0}} . \end{aligned}$$

Proof of (8.6). We need to estimate all of the terms in the difference \(2w - (\partial _t + b \partial _{\alpha }) v\) from (8.25). The terms

$$\begin{aligned}{}[w, {\mathcal {H} } ] \frac{{\zeta }_{\alpha }- \overline{{\zeta }}_{\alpha }}{{\zeta }_{\alpha }} \qquad \text{ and } \qquad ( {\mathcal {H} } + \overline{ {\mathcal {H} } } )w \end{aligned}$$

can be estimated as above, the only difference being the appearance of \(w\) instead of \(u\). The last term in (8.25) is of the form \(\mathbf{T}(u,u, \text{ Im }{\zeta }_{\alpha })\), where \(\mathbf{T}\) is defined in (7.27). This can be directly estimated using (11.35). The remaining terms are

$$\begin{aligned} I_1 := [\overline{u}, \overline{ {\mathcal {H} } }] \frac{u_{\alpha }}{\overline{{\zeta }}_{\alpha }} \qquad \text{ and } \qquad I_2 := [u, {\mathcal {H} } ] \frac{2 u_{\alpha }- \overline{u}_{\alpha }}{{\zeta }_{\alpha }} . \end{aligned}$$

These are terms of the form \(Q_0 (\widetilde{L}, \partial _{\alpha }\widetilde{L})\) and can be bounded by making use of (11.30):

$$\begin{aligned} {\left\| I_1 \right\| }_{X_{N_0}} + {\left\| I_2 \right\| }_{X_{N_0}}&\lesssim {\left\| Q_0 (u,\partial _{\alpha }u) \right\| }_{X_{N_0}} \lesssim {\varepsilon }_1 {\Vert \widetilde{L} \Vert }_{X_{N_0}} . \end{aligned}$$

Proof of (8.7). We start from (8.26) and rewrite as:

$$\begin{aligned} {\zeta }_{\alpha }- 1&= -iw + i w \left( 1 - \frac{1}{A} \right) - \frac{A - 1}{A} . \end{aligned}$$
(8.28)

Since \(w\) satisfies (8.6), in order to show (8.7) is suffices to verify that

$$\begin{aligned} {\left\| w \left( 1 - \frac{1}{A} \right) \right\| }_{X_{N_0}} + {\left\| \frac{A - 1}{A} \right\| }_{X_{N_0}} \lesssim {\varepsilon }_1 {\Vert \widetilde{L} \Vert }_{X_{N_0}} . \end{aligned}$$

This follows from the bounds on \(A-1\) (7.13) and (7.14), and the a priori assumptions (8.1).

Proof of (8.8). This follows directly from (8.27) and arguments similar to the ones above.

Proof of (8.11) and (8.12). Using (11.9) in combination with (8.8) one can deduce that

$$\begin{aligned} {\Vert \text{ Im }{\zeta }\Vert }_{X_{N_0}}&\lesssim {\Vert (\partial _t + b\partial _{\alpha }) {\lambda }\Vert }_{X_{N_0}} + {\varepsilon }_1 {\Vert \widetilde{L} \Vert }_{X_{N_0}} . \end{aligned}$$

In view of (8.5)–(8.9) we have then obtained (8.11).

From the a priori bound on the \(H^{N_1 + 5}\)-norm of \(\widetilde{L}\) it is not hard to see that (8.10) holds true by using similar arguments as above. (8.12) immediately follows, concluding the proof of Lemma 8.4. \(\square \)

8.1.2 Proof of Lemma 8.5

Identities (2.50), (2.35) and (2.36) in [45] respectively read:

$$\begin{aligned} \frac{i}{2} \partial _{\alpha }\chi&= w \overline{{\zeta }}_{\alpha }- \frac{1}{2} {\mathcal {H} } \left( u_{\alpha }\frac{u \overline{{\zeta }}_{\alpha }}{{\zeta }_{\alpha }} \right) +\frac{1}{2} [u, {\mathcal {H} } ] \left( \frac{\partial _{\alpha }(u \overline{{\zeta }}_{\alpha })}{{\zeta }_{\alpha }} \right) - \frac{1}{2\pi i} \nonumber \\&\quad \,\int \frac{ (u({\alpha }) - u({\beta })) ({\zeta }_{\alpha }({\alpha }) - {\zeta }_{\beta }({\beta })) }{ {({\zeta }({\alpha }) - {\zeta }({\beta }))}^2 } u({\beta }) \overline{{\zeta }}_{\beta }({\beta }) \, d{\beta }+ \frac{1}{2}({\zeta }_{\alpha }- \overline{{\zeta }}_{\alpha }) w\nonumber \\&\quad \, +\frac{1}{2} \overline{u}_{\alpha }u -\frac{1}{2} {\zeta }_{\alpha }\left( {\mathcal {H} } \frac{1}{{\zeta }_{\alpha }} + \overline{ {\mathcal {H} } } \frac{1}{\overline{{\zeta }}_{\alpha }} \right) (w\overline{{\zeta }}_{\alpha }+ u\overline{u}_{\alpha }) \nonumber \\&\quad +\, \frac{{\zeta }_{\alpha }}{\pi } \int \text{ Im }\left( \frac{ u({\alpha }) - u({\beta }) }{ {({\zeta }({\alpha }) - {\zeta }({\beta }))}^2 } \right) u({\beta }) \overline{{\zeta }}_{\beta }({\beta }) \, d{\beta },\end{aligned}$$
(8.29)
$$\begin{aligned} v&= 2u - ( {\mathcal {H} } + \overline{ {\mathcal {H} } }) u - [u, {\mathcal {H} } ] \frac{{\zeta }_{\alpha }- \overline{{\zeta }}_{\alpha }}{{\zeta }_{\alpha }} ,\end{aligned}$$
(8.30)
$$\begin{aligned} \partial _{\alpha }{\lambda }&= u \overline{{\zeta }}_{\alpha }- \frac{1}{2} \left( {\zeta }_{\alpha } {\mathcal {H} } \frac{1}{{\zeta }_{\alpha }} + \overline{{\zeta }}_{\alpha }\overline{ {\mathcal {H} } } \frac{1}{\overline{{\zeta }}_{\alpha }} \right) (u \overline{{\zeta }}_{\alpha }) . \end{aligned}$$
(8.31)

From these, Proposition 11.7 and the a priori assumption (8.1), it is not hard to see that

$$\begin{aligned} {\Vert \partial _{\alpha }\chi \Vert }_{H^{N_1}} + {\Vert v \Vert }_{H^{N_1}} + {\Vert \partial _{\alpha }{\lambda }\Vert }_{H^{N_1}}&\lesssim {\varepsilon }_1 ,\end{aligned}$$
(8.32)
$$\begin{aligned} {\Vert \partial _{\alpha }\chi \Vert }_{X_{N_0}} + {\Vert v \Vert }_{X_{N_0}} + {\Vert \partial _{\alpha }{\lambda }\Vert }_{X_{N_0}}&\lesssim {\Vert \widetilde{L} \Vert }_{X_{N_0}} . \end{aligned}$$
(8.33)

We are now going to use these bounds to control the commutators \([\partial _t + b\partial _{\alpha }, D^k] f\) and \([\partial _t + b\partial _{\alpha }, S] f\) for \(f = \chi ,v,{\lambda }\). By direct computation one sees that

$$\begin{aligned} \left[ D^k, \partial _t + b\partial _{\alpha }\right]&= \left[ D^k, b\partial _{\alpha }\right] = \sum _{j=1}^k \partial ^j b \partial ^{k-j} \partial _{\alpha }\end{aligned}$$
(8.34)
$$\begin{aligned} \left[ S, \partial _t + b\partial _{\alpha }\right]&= \left( Sb -\frac{1}{2}b \right) \partial _{\alpha }- \frac{1}{2}(\partial _t + b\partial _{\alpha }) . \end{aligned}$$
(8.35)

Moreover, for \(S_k = D^k S\), we have

$$\begin{aligned}&[S_k, \partial _t + b\partial _{\alpha }] f = D^k [S, \partial _t + b\partial _{\alpha }] f + [D^k, \partial _t + b\partial _{\alpha }] S f . \end{aligned}$$
(8.36)

From (8.34) it follows that for any \(0\le k \le N_0\)

$$\begin{aligned}&{\left\| D^k (\partial _t + b\partial _{\alpha }) f - (\partial _t + b\partial _{\alpha }) D^k f \right\| }_{L^2} \\&\quad \lesssim \sum _{j=1}^{N_0/2} {\Vert \partial ^j b \Vert }_{L^\infty } {\Vert \partial _{\alpha }f \Vert }_{H^{k-1}} + {\Vert \partial _{\alpha }b \Vert }_{H^{k-1}} \sum _{j=1}^{N_0/2} {\Vert \partial _{\alpha }f \Vert }_{L^\infty } \\&\quad \lesssim {\Vert \partial _{\alpha }b \Vert }_{H^{\frac{N_0}{2}+1}} {\Vert \partial _{\alpha }f \Vert }_{H^{k-2}} + {\Vert \partial _{\alpha }b \Vert }_{H^{k-1}} {\Vert \partial _{\alpha }f \Vert }_{H^{\frac{N_0}{2}+1}} . \end{aligned}$$

Using (8.32) and (8.33) we see that for \(f = \chi ,v\) and \(0\le k \le N_0\), or \(f={\lambda }\) and \(0\le k \le N_0 - 2\), we have

$$\begin{aligned}&{\left\| D^k (\partial _t + b\partial _{\alpha }) f - (\partial _t + b\partial _{\alpha }) D^k f \right\| }_{L^2} \lesssim {\Vert \partial _{\alpha }b \Vert }_{H^{\frac{N_0}{2}+1}} {\Vert \widetilde{L} \Vert }_{X_{N_0}} \\&\quad + {\varepsilon }_1 {\Vert \partial _{\alpha }b \Vert }_{H^{N_0-1}} . \end{aligned}$$

This would conclude the proof of (8.16) provided one has

$$\begin{aligned}&{\Vert \partial _{\alpha }b \Vert }_{H^{\frac{N_0}{2}+1}} \lesssim {\varepsilon }_1 \qquad \text{ and } \qquad {\Vert \partial _{\alpha }b \Vert }_{H^{N_0-1}} \lesssim {\Vert \widetilde{L} \Vert }_{X_{N_0}} . \end{aligned}$$

These two estimates follow from Lemma 7.1.

To show the commutator estimate (8.17) we start by using (8.36):

$$\begin{aligned}&{\Vert [S_k, \partial _t + b\partial _{\alpha }] f \Vert }_{L^2} \lesssim {\Vert [D^k, \partial _t + b\partial _{\alpha }] S f \Vert }_{L^2} + {\Vert [S, \partial _t + b\partial _{\alpha }] f \Vert }_{H^k} . \end{aligned}$$
(8.37)

From (8.34), and for \(0\le k \le \frac{N_0}{2}\), we get

$$\begin{aligned} {\Vert [D^k, \partial _t + b\partial _{\alpha }] S f \Vert }_{L^2} \lesssim {\Vert \partial _{\alpha }b \Vert }_{H^{\frac{k}{2}+1}} {\Vert \partial _{\alpha }S f \Vert }_{H^{k-1}} + {\Vert \partial _{\alpha }b \Vert }_{H^{k-1}} {\Vert \partial _{\alpha }S f \Vert }_{H^{\frac{k}{2}+1}} . \end{aligned}$$

Using (8.33) we can bound the above right-hand side to obtain

$$\begin{aligned}&{\Vert [D^k, \partial _t + b\partial _{\alpha }] S f \Vert }_{L^2} \lesssim {\Vert \partial _{\alpha }b \Vert }_{H^{\frac{N_0}{2}}} {\Vert \widetilde{L} \Vert }_{X_{N_0}}, \end{aligned}$$
(8.38)

for \(f = \chi \), \(v\) or \({\lambda }\). Moreover, from (8.35) we see that

$$\begin{aligned} {\Vert [S, \partial _t + b\partial _{\alpha }] f \Vert }_{H^k} \lesssim {\Vert S b \partial _{\alpha }f \Vert }_{H^k} + {\Vert b \partial _{\alpha }f \Vert }_{H^k} + {\Vert (\partial _t + b \partial _{\alpha })f \Vert }_{H^k} . \end{aligned}$$

With \(f = \chi \), \(v\) or \({\lambda }\), and using (8.32) and (8.33) we deduce

$$\begin{aligned} {\Vert [S, \partial _t + b\partial _{\alpha }] f \Vert }_{H^k} \lesssim {\varepsilon }_1 \left( {\Vert S b \Vert }_{H^k} + {\Vert b \Vert }_{H^k} \right) + {\Vert (\partial _t + b \partial _{\alpha })f \Vert }_{H^k} . \end{aligned}$$
(8.39)

Putting together (8.37) with (8.38) and (8.39) we get

$$\begin{aligned}&{\Vert [S_k, \partial _t + b\partial _{\alpha }] f \Vert }_{L^2} \lesssim {\Vert \partial _{\alpha }b \Vert }_{H^{\frac{N_0}{2}}} {\Vert \widetilde{L} \Vert }_{X_{N_0}} + {\varepsilon }_1 \left( {\Vert S b \Vert }_{H^{\frac{N_0}{2}}} + {\Vert b \Vert }_{H^{\frac{N_0}{2}}} \right) \\&\quad + {\Vert (\partial _t + b \partial _{\alpha })f \Vert }_{H^k} . \end{aligned}$$

To obtain (8.17) it then suffices to have

$$\begin{aligned}&{\Vert S b \Vert }_{H^{\frac{N_0}{2}}} \lesssim {\Vert \widetilde{L} \Vert }_{X_{N_0}} \qquad \text{ and } \qquad {\Vert b \Vert }_{H^{\frac{N_0}{2}+1}} \lesssim {\varepsilon }_1 . \end{aligned}$$

These two estimates are again direct consequence of Lemma 7.1. \(\square \)

8.1.3 Proof of Lemma 8.6

Recall the definition of \(E^\chi \) given in (7.34):

$$\begin{aligned} E^\chi (t)&= \sum _{k=0}^{N_0} \int _{\mathbb {R}}\frac{1}{A} {\left| (\partial _t + b \partial _{\alpha }) D^k \chi \right| }^2 + i {\left( D^k \chi \right) }^h \partial _\alpha \overline{{\left( D^k\chi \right) }^h} \, d{\alpha }\end{aligned}$$
(8.40)
$$\begin{aligned}&\quad + \sum _{k=0}^{N_0/2} \int _{\mathbb {R}}\frac{1}{A} {\left| (\partial _t + b \partial _{\alpha }) S_k \chi \right| }^2 + i {\left( S_k \chi \right) }^h \partial _\alpha \overline{ {\left( S_k \chi \right) }^h} \, d{\alpha }\, \, \end{aligned}$$
(8.41)

where \( f^h := (I- {\mathcal {H} } )f/2\). As in Lemma 4.1 of [45] (see Lemma 8.7 below) we know that if \(\Theta \) is the boundary value of an holomorphic function in \(\Omega (t)^c\), such as \(f^h\), then

$$\begin{aligned} i \int \Theta \partial _{\alpha }\overline{\Theta } \, d{\alpha }\ge 0 . \end{aligned}$$

Therefore both summands in (8.40) and (8.41) are nonnegative. In particular

$$\begin{aligned} E^\chi (t) \gtrsim \sum _{k=0}^{N_0} \int _{\mathbb {R}}\frac{1}{A} {\left| (\partial _t + b \partial _{\alpha }) D^k \chi \right| }^2 + \, \sum _{k=0}^{N_0/2} \int _{\mathbb {R}}\frac{1}{A} {\left| (\partial _t + b \partial _{\alpha }) S_k \chi \right| }^2 . \end{aligned}$$
(8.42)

From this, and since \(| A-1 | \le 1/2\), see (7.14), we get the desired bounds (8.18) and (8.21). The exact same argument can be used to show (8.20) and (8.23).

To prove (8.19) and (8.22) the argument is more complicated since \(E^v\) is not nonnegative, but we just need to adapt the proof of Lemma 4.2 in [45, pp. 89-92] to see that

$$\begin{aligned} E^v (t) \ge \frac{1}{4} {\Vert (\partial _t + b \partial _{\alpha }) {\Gamma }v \Vert }^2_{L^2} + \frac{1}{8} {\Vert (\partial _t + b \partial _{\alpha }) {\Gamma }v_1 \Vert }^2_{L^2} - {\varepsilon }_1 E^\chi - {\varepsilon }_1 {\Vert \widetilde{L} \Vert }^2_{X_{N_0}} \end{aligned}$$

where \({\Gamma }= D^k\) for \(0 \le k \le N_0\), or \({\Gamma }= D^k S\) for \(0\le k \le \frac{N_0}{2}\). \(\Box \)

8.2 Proof of Proposition 8.2: evolution of the energy

We want to show, under the a priori assumptions (8.1), the following bound for the evolution of the Energy

$$\begin{aligned} \frac{d}{dt} \sqrt{E(t)} \lesssim {\left( {\Vert L(t) \Vert }_{W^{N_1,\infty }} + {\Vert {\mathcal {H} } L^-(t) \Vert }_{W^{N_1,\infty }} \right) }^2 \sqrt{E(t)} \, \end{aligned}$$
(8.43)

for any \(t\in [0,T]\). From the definition of \(E\) in (7.38), and the bound (8.2), we see that it suffices to prove

$$\begin{aligned} \frac{d}{dt} E^f(t) \lesssim {\left( {\Vert L(t) \Vert }_{W^{N_1,\infty }} + {\Vert {\mathcal {H} } L^-(t)\Vert }_{W^{N_1,\infty }} \right) }^2 \left( {\Vert \widetilde{L}(t) \Vert }_{X_{N_0}}^2 + E(t) \right) \,\nonumber \\ \end{aligned}$$
(8.44)

for any \(t\in [0,T]\) and \(f = \chi ,{\lambda }\) and \(v\).

8.2.1 Basic energy equality

Assume that \(F\) is a smooth function vanishing sufficiently fast at infinity and satisfying the equation

$$\begin{aligned} {\mathcal {P}}F := {(\partial _t + b(t,{\alpha }) \partial _{\alpha })}^2 F(t,{\alpha }) - i A(t,{\alpha }) \partial _{\alpha }F(t,{\alpha }) = G(t,{\alpha }) . \end{aligned}$$
(8.45)

Define the zero-th energy associated to (8.45) by

$$\begin{aligned} E_0^F (t)&= \int _{\mathbb {R}}\frac{1}{A(t,{\alpha })} {\left| (\partial _t + b(t,{\alpha }) \partial _{\alpha }) F(t,{\alpha }) \right| }^2 + i F(t,{\alpha }) \partial _\alpha \overline{F}(t,{\alpha }) \, d{\alpha }. \end{aligned}$$
(8.46)

Also define higher order energies

$$\begin{aligned} E_j^F (t)&\! = \!\int _{\mathbb {R}}\frac{1}{A(t,{\alpha })} {\left| (\partial _t \!+\! b(t,{\alpha }) \partial _{\alpha }) {\Gamma }^j F(t,{\alpha }) \right| }^2\nonumber \\ {}&\quad + i {({\Gamma }^j F)}^h (t,{\alpha }) \partial _\alpha \overline{{({\Gamma }^j F)}^h} (t,{\alpha }) \, d{\alpha }. \end{aligned}$$
(8.47)

where \(j\ge 1\), \({\Gamma }^j=D^j\) for \(j \le N_0\), or \({\Gamma }^j=S_j\) for \(j \le N_0/2\), and \(f^h\) is defined in (7.35). Then the following holds:

Lemma 8.7

([45]) Assume that \(F\) and \(E_0\) are as above, then

$$\begin{aligned} \frac{d}{dt} E_0^F (t)&= \int _{\mathbb {R}}\frac{2}{A} \text{ Re }\left( (\partial _t + b \partial _{\alpha }) F \, \overline{G} \right) - \frac{1}{A} \frac{a_t}{a} \circ k^{-1} {\left| (\partial _t + b \partial _{\alpha }) F \right| }^2 \, d{\alpha }. \end{aligned}$$
(8.48)

Furthermore, if \(\Theta \) is the boundary value of an holomorphic function in \(\Omega _t^c\), that is \(\Theta =\Theta ^h\), then

$$\begin{aligned} i \int _{\mathbb {R}}\Theta (t,{\alpha }) \partial _\alpha \overline{\Theta }(t,{\alpha }) \, d{\alpha }\ge 0 . \end{aligned}$$

Let \(E_j^F (t)\) be as in (8.47), then

$$\begin{aligned}&\frac{d}{dt} E_j^F (t) = \int _{\mathbb {R}}\frac{2}{A} \text{ Re }\left( (\partial _t + b \partial _{\alpha }) {\Gamma }^j F \overline{G_j} \right) - \frac{1}{A} \frac{a_t}{a} \circ k^{-1} {\left| (\partial _t + b \partial _{\alpha }) {\Gamma }^j F \right| }^2 \, d{\alpha }\end{aligned}$$
(8.49)
$$\begin{aligned}&\quad - 2\text{ Re }\int _{\mathbb {R}}i\partial _t {({\Gamma }^j F)}^h \partial _{\alpha }\overline{{({\Gamma }^j F)}^r} + i \partial _t {({\Gamma }^j F)}^r \partial _{\alpha }\overline{{({\Gamma }^j F)}^h}\nonumber \\&\quad + \partial _t {({\Gamma }^j F)}^r \partial _{\alpha }\overline{{({\Gamma }^j F)}^r} \, d{\alpha }, \end{aligned}$$
(8.50)

where

$$\begin{aligned} G_j := {\Gamma }^j G + [{\mathcal {P}},{\Gamma }^j]F \end{aligned}$$

and \(f^r\) is defined by \(f^r := (1/2)(I+ {\mathcal {H} } )f\).

The proof of the above Lemma can be found in [45, pp. 83–85].

8.2.2 Evolution of \(E^\chi \)

We want to show

$$\begin{aligned} \frac{d}{dt} E^\chi (t) \lesssim {\left( {\Vert L \Vert }_{W^{N_1,\infty }} + {\Vert {\mathcal {H} } L^- \Vert }_{W^{N_1,\infty }} \right) }^2 E(t) . \end{aligned}$$
(8.51)

We recall that the energy \(E^\chi \) is given by

$$\begin{aligned} E^\chi (t)&= \sum _{k=0}^{N_0} \int _{\mathbb {R}}\frac{1}{A} {\left| (\partial _t + b \partial _{\alpha }) D^k \chi \right| }^2 + i {\left( D^k \chi \right) }^h \partial _\alpha \overline{{\left( D^k \chi \right) }^h} \, d{\alpha }\\&\quad \, + \sum _{k=0}^{N_0/2} \int _{\mathbb {R}}\frac{1}{A} {\left| (\partial _t + b \partial _{\alpha }) S_k \chi \right| }^2 + i {\left( S_k \chi \right) }^h \partial _\alpha \overline{{\left( S_k \chi \right) }^h} \, d{\alpha }. \end{aligned}$$

From (7.18) and (7.29) we know that \(\chi \) satisfies an equation of the form \({\mathcal {P}}\chi = G^\chi \) with a cubic nonlinearity \(G^\chi = \mathbf{T}(u,\text{ Im }{\zeta }, u_{\alpha }) + \mathbf{T}(u,u,\text{ Im }{\zeta }_{\alpha })\), where \(\mathbf{T}\) are operators of the type defined in (7.28). This nonlinearity can be schematically rewritten as

$$\begin{aligned} G^\chi = \mathbf{T}(\widetilde{L}, \widetilde{L}, \widetilde{L}_{\alpha }) . \end{aligned}$$
(8.52)

By using Lemma 8.7 one obtains

$$\begin{aligned}&\frac{d}{dt} E^\chi (t)= \sum _{k=0}^{N_0} \int \frac{2}{A} \text{ Re }\left( (\partial _t + b \partial _{\alpha }) D^k \chi \, {\mathcal {P}}D^k \chi \right) \nonumber \\&\quad - \frac{1}{A} \frac{a_t}{a} \circ k^{-1} {\left| (\partial _t + b \partial _{\alpha }) D^k \chi \right| }^2 \, d{\alpha }\end{aligned}$$
(8.53)
$$\begin{aligned}&\quad - 2 \sum _{k=0}^{N_0} \text{ Re }\int i\partial _t (D^k \chi )^h \partial _{\alpha }\overline{(D^k \chi )^r} + i\partial _t (D^k \chi )^r \partial _{\alpha }\overline{(D^k \chi )^h}\nonumber \\&\quad + i \partial _t (D^k \chi )^r \partial _{\alpha }\overline{(D^k \chi )^r} \, d{\alpha }\end{aligned}$$
(8.54)
$$\begin{aligned}&\quad + \sum _{k=0}^{N_0/2} \int \frac{2}{A} \text{ Re }\left( (\partial _t + b \partial _{\alpha }) S_k \chi \, {\mathcal {P}}S_k \chi \right) - \frac{1}{A} \frac{a_t}{a} \circ k^{-1} {\left| (\partial _t + b \partial _{\alpha }) S_k \chi \right| }^2 \, d{\alpha }\end{aligned}$$
(8.55)
$$\begin{aligned}&\quad - 2 \sum _{k=0}^{N_0/2} \text{ Re }\int i\partial _t (S_k \chi )^h \partial _{\alpha }\overline{(S_k \chi )^r} + i\partial _t (S_k \chi )^r \partial _{\alpha }\overline{(S_k \chi )^h}\nonumber \\&\quad + i \partial _t (S_k \chi )^r \partial _{\alpha }\overline{(S_k \chi )^r} \, d{\alpha }. \end{aligned}$$
(8.56)

Since \({\Vert A -1 \Vert }_{L^\infty } \le \frac{1}{2}\), we have

$$\begin{aligned} (8.53) \!+\! (8.55)&\lesssim \sqrt{E^\chi (t)} \sum _{k=0}^{N_0} \left( {\Vert D^k G^\chi \Vert }_{L^2} \!+\! {\Vert [{\mathcal {P}}, D^k] \chi \Vert }_{L^2} \right) + \sqrt{E^\chi (t)}\\&\quad \times \sum _{k=0}^{N_0/2} \left( {\Vert S_k G^\chi \Vert }_{L^2} \!+\! {\Vert [{\mathcal {P}}, S_k] \chi \Vert }_{L^2} \right) \!+\! E^\chi (t) { \left\| \frac{a_t}{a} \circ k^{-1} \right\| }_{L^\infty } . \end{aligned}$$

The terms in (8.54) and (8.56) are remainder terms. Since there is no logarithmic loss in the estimates for those terms in [45, p. 94-98], they can be estimated exactly as in the cited paper, and therefore we skip them. Thus, to obtain the desired bound (8.51) on the evolution of \(E^\chi \) it suffices to show

$$\begin{aligned} {\Vert G^\chi \Vert }_{X_{N_0}}&\lesssim {\Vert L \Vert }^2_{W^{N_1+3,\infty }} \sqrt{E}\end{aligned}$$
(8.57)
$$\begin{aligned} \sum _{k=0}^{N_0} {\Vert [{\mathcal {P}}, D^k] \chi \Vert }_{L^2}&\lesssim {\left( {\Vert L \Vert }_{W^{N_1,\infty }} + {\Vert {\mathcal {H} } L^- \Vert }_{W^{N_1,\infty }} \right) }^2 \sqrt{E}\end{aligned}$$
(8.58)
$$\begin{aligned} \sum _{k=0}^{N_0/2} {\Vert [{\mathcal {P}}, S_k] \chi \Vert }_{L^2}&\lesssim {\left( {\Vert L \Vert }_{W^{N_1,\infty }} + {\Vert {\mathcal {H} } L^- \Vert }_{W^{N_1,\infty }} \right) }^2 \sqrt{E}\end{aligned}$$
(8.59)
$$\begin{aligned} {\left\| \frac{a_t}{a} \circ k^{-1} \right\| }_{L^\infty }&\lesssim {\Vert \widetilde{L} \Vert }^2_{W^{N_1,\infty }}. \end{aligned}$$
(8.60)

These estimates are performed in the next four subsections.

Proof of (8.57). The bound (8.57) follows directly from (8.52) and Proposition 11.7. Indeed applying (11.35) we see that

$$\begin{aligned} {\Vert G^\chi \Vert }_{X_{N_0}} = {\Vert \mathbf{T}(\widetilde{L}, \widetilde{L}, \widetilde{L}_{\alpha }) \Vert }_{X_{N_0}}\lesssim {\Vert \widetilde{L} \Vert }^2_{W^{N_1,\infty }} {\Vert \widetilde{L} \Vert }_{X_{N_0}} \lesssim {\Vert L \Vert }^2_{W^{N_1,\infty }} \sqrt{E} , \end{aligned}$$

having used (8.2) in the last inequality.

Proof of (8.58). Recall the definition \({\mathcal {P}}= {(\partial _t + b \partial _{\alpha })}^2 - i A \partial _{\alpha }\). By direct computation we see that

$$\begin{aligned}&[D^k, {\mathcal {P}}] f =\! [D^k, \partial _t \!+\! b\partial _{\alpha }] (\partial _t \!+\! b\partial _{\alpha })f \!+\! (\partial _t + b\partial _{\alpha }) [D^k, \partial _t + b\partial _{\alpha }] f \nonumber \\&\quad - i [D^k, A \partial _{\alpha }] f= [D^k, b\partial _{\alpha }] (\partial _t + b\partial _{\alpha })f + (\partial _t + b\partial _{\alpha }) [D^k, b\partial _{\alpha }] f \nonumber \\&\quad - i [D^k, A \partial _{\alpha }] f= \sum _{j=1}^k c_{k,j} D^j b D^{k-j} \partial _{\alpha }(\partial _t + b\partial _{\alpha })f \nonumber \\&\quad + (\partial _t + b\partial _{\alpha }) \sum _{j=1}^k c_{k,j} D^j b D^{k-j} \partial _{\alpha }f- i \sum _{j=1}^k c_{k,j} D^j A D^{k-j} \partial _{\alpha }f , \end{aligned}$$
(8.61)

for some coefficients \(c_{k,j}\). It follows that for any \(0 \le k \le N_0\)

$$\begin{aligned}&{\Vert [{\mathcal {P}}, D^k] \chi \Vert }_{L^2} \nonumber \lesssim {\Vert (\partial _t + b\partial _{\alpha }) \chi \Vert }_{H^{N_0}} {\Vert \partial _{\alpha }b \Vert }_{W^{N_0/2,\infty }}+ {\Vert \partial _{\alpha }b \Vert }_{H^{N_0-1}} {\Vert (\partial _t + b\partial _{\alpha }) \chi \Vert }_{W^{N_0/2,\infty }}\end{aligned}$$
(8.62a)
$$\begin{aligned}&\quad \quad + \sum _{j=0}^{N_0-1} {\Vert (\partial _t + b\partial _{\alpha }) D^j \partial _{\alpha }b \Vert }_{L^2} {\Vert \partial _{\alpha }\chi \Vert }_{W^{N_0/2,\infty }}+ {\Vert \partial _{\alpha }b \Vert }_{H^{N_0-1}} \times \sum _{j=0}^{N_0/2} {\Vert (\partial _t + b\partial _{\alpha }) D^j \partial _{\alpha }\chi \Vert }_{L^\infty }\end{aligned}$$
(8.62b)
$$\begin{aligned}&\quad \quad + \sum _{j=0}^{N_0/2} {\Vert (\partial _t + b\partial _{\alpha }) D^j \partial _{\alpha }b \Vert }_{L^\infty } {\Vert \partial _{\alpha }\chi \Vert }_{H^{N_0-1}}+ {\Vert \partial _{\alpha }b \Vert }_{W^{N_0/2,\infty }} \times \sum _{j=0}^{N_0-1} {\Vert (\partial _t + b\partial _{\alpha }) D^j \partial _{\alpha }\chi \Vert }_{L^2}\end{aligned}$$
(8.62c)
$$\begin{aligned}&\quad \quad + {\Vert \partial _{\alpha }\chi \Vert }_{H^{N_0}} {\Vert \partial _{\alpha }A \Vert }_{W^{N_0/2,\infty }}+ {\Vert \partial _{\alpha }A \Vert }_{H^{N_0}} {\Vert \partial _{\alpha }\chi \Vert }_{W^{N_0/2,\infty }} . \end{aligned}$$
(8.62d)

Combining (8.11), commutator estimates for \([\partial _t + b\partial _{\alpha }, D^j]\), (8.13) and (8.2), it follows that

$$\begin{aligned} {\Vert (\partial _t + b \partial _{\alpha }) \chi \Vert }_{H^{N_0}} +\sum _{j=0}^{N_0-1} {\Vert (\partial _t + b \partial _{\alpha }) D^j \partial _{\alpha }\chi \Vert }_{L^2} +{\Vert \partial _{\alpha }\chi \Vert }_{H^{N_0}} \lesssim \sqrt{E(t)} . \end{aligned}$$
(8.63)

From Lemma 8.4, and commutation estimates for \([\partial _t + b\partial _{\alpha }, D^j]\), we also control the following \(L^\infty \) norms:

$$\begin{aligned}&{\Vert (\partial _t + b \partial _{\alpha }) \chi \Vert }_{W^{N_0/2,\infty }} \!+\! {\Vert \partial _{\alpha }\chi \Vert }_{W^{N_0/2,\infty }} +\sum _{j=0}^{N_0/2} {\Vert (\partial _t + b \partial _{\alpha }) D^j \partial _{\alpha }\chi \Vert }_{L^\infty }\nonumber \\&\quad \lesssim {\Vert L \Vert }_{W^{N_1,\infty }} . \end{aligned}$$
(8.64)

Estimate of (8.62a). Using (8.63) and (8.64) we can bound

$$\begin{aligned} (8.62a) \lesssim \sqrt{E(t)} {\Vert \partial _{\alpha }b \Vert }_{W^{N_0/2,\infty }}+ {\Vert \partial _{\alpha }b \Vert }_{H^{N_0-1}} {\Vert L \Vert }_{W^{N_1,\infty }} . \end{aligned}$$

To obtain the desired bound in then suffices to show

$$\begin{aligned} {\Vert \partial _{\alpha }b \Vert }_{W^{N_0/2,\infty }}&\lesssim \left( {\Vert {\mathcal {H} } L^- \Vert }_{W^{N_1,\infty }} + {\Vert L \Vert }_{W^{N_1,\infty }} \right) ^2 \end{aligned}$$
(8.65)
$$\begin{aligned} {\Vert \partial _{\alpha }b \Vert }_{H^{N_0-1}}&\lesssim {\Vert L \Vert }_{W^{N_1,\infty }} \sqrt{E} . \end{aligned}$$
(8.66)

From formula (7.17) we see that \((I- {\mathcal {H} } )b = g\) with

$$\begin{aligned} g = - [u, {\mathcal {H} } ] \frac{\overline{{\zeta }}_{\alpha }- 1}{{\zeta }_{\alpha }} . \end{aligned}$$
(8.67)

Using (11.11) in Lemma 11.2 we get

$$\begin{aligned} {\Vert \partial _{\alpha }b \Vert }_{W^{N_0/2,\infty }}&\lesssim {\Vert \partial _{\alpha }g \Vert }_{W^{N_0/2,\infty }}+ {\Vert \widetilde{L} \Vert }_{W^{N_0/2+1,\infty }} {\Vert \partial _{\alpha }g \Vert }_{H^{N_0/2+1}} . \end{aligned}$$

Since \(g\) above is an operator of the form \(Q_0(\widetilde{L} ,\widetilde{L})\), with \(Q_0\) defined in (11.18), we can use the \(L^\infty \) bound provided by (11.33) to deduce

$$\begin{aligned} {\Vert \partial _{\alpha }b \Vert }_{W^{N_0/2,\infty }}&\lesssim {\Vert L \Vert }_{W^{N_0/2+2,\infty }} {\left( {\Vert {\mathcal {H} } L^- \Vert }_{W^{N_0/2+2,\infty }} + {\Vert L \Vert }_{W^{N_0/2+2,\infty }} \right) } . \end{aligned}$$
(8.68)

Here we have also used the \(L^2\)-bounds from 11.4 to estimate \({\Vert \partial _{\alpha }g \Vert }_{H^{N_0/2+1}}\), and the a priori assumption \({\Vert \widetilde{L} \Vert }_{H^{N_1+5}} \lesssim {\varepsilon }_1\). (8.65) is proven. Using again Lemma 11.2 we can estimate

$$\begin{aligned}&{\Vert \partial _{\alpha }b \Vert }_{H^{N_0-1}} \lesssim {\Vert \partial _{\alpha }g \Vert }_{H^{N_0-1}} \lesssim {\Vert \widetilde{L} \Vert }_{H^{N_0}} {\Vert \widetilde{L} \Vert }_{W^{N_0/2+2,\infty }} , \end{aligned}$$
(8.69)

which in light of (8.2) suffices to obtain (8.66).

Estimate of (8.62b). Using (8.63) and (8.64) we see that

$$\begin{aligned} (8.62b)&\lesssim \sum _{j=0}^{N_0-1} {\Vert (\partial _t + b\partial _{\alpha }) D^j \partial _{\alpha }b \Vert }_{L^2} {\Vert L \Vert }_{W^{N_0/2+2,\infty }} \\&\quad + {\Vert \partial _{\alpha }b \Vert }_{H^{N_0-1}} {\Vert L \Vert }_{W^{N_0/2+2,\infty }} . \end{aligned}$$

Since we already have the bound (8.66) for \({\Vert \partial _{\alpha }b \Vert }_{H^{N_0-1}}\), in order to estimate (8.62b) it suffices to show

$$\begin{aligned} \sum _{j=0}^{N_0-1} {\Vert (\partial _t + b\partial _{\alpha }) D^j \partial _{\alpha }b \Vert }_{L^2} \lesssim {\Vert \widetilde{L} \Vert }_{W^{N_0/2+2,\infty }} {\Vert \widetilde{L} \Vert }_{X_{N_0}} . \end{aligned}$$
(8.70)

To establish this estimate we use the following identity derived in [45, formula (2.52)]:

$$\begin{aligned}&(I - {\mathcal {H} } ) (\partial _t + b\partial _{\alpha })b = B_3 \nonumber \\&B_3 \!:=\! [u, {\mathcal {H} } ] \frac{\partial _{\alpha }( 2b - \overline{u}) }{{\zeta }_{\alpha }} \!-\! [w, {\mathcal {H} } ] \frac{\overline{{\zeta }}_{\alpha }- 1}{{\zeta }_{\alpha }} \!+\! \frac{1}{i\pi } \nonumber \\&\quad \int {\left( \frac{u({\alpha }) - u({\beta })}{{\zeta }({\alpha }) - {\zeta }({\beta })} \right) }^2 (\overline{{\zeta }}_{\beta }({\beta }) - 1) \, d{\beta }. \end{aligned}$$
(8.71)

Using this formula, (11.8) in Lemma 11.2, and the fact that \([\partial _{\alpha }, \partial _t + b\partial _{\alpha }]b\) gives higher order quartic terms, it is not hard to see that (8.70) holds

Estimate of (8.62c). Using again (8.63) and (8.64) we can bound

$$\begin{aligned} (8.62c) \lesssim \sum _{j=0}^{N_0/2} {\Vert (\partial _t + b\partial _{\alpha }) D^j \partial _{\alpha }b \Vert }_{L^\infty } \sqrt{E(t)} + {\Vert \partial _{\alpha }b \Vert }_{W^{N_0/2,\infty }} \sqrt{E(t)} . \end{aligned}$$

By virtue of (8.65) it suffices to establish the bound

$$\begin{aligned} \sum _{j=0}^{N_0/2} {\Vert (\partial _t + b\partial _{\alpha }) D^j \partial _{\alpha }b \Vert }_{L^\infty } \lesssim {\left( {\Vert {\mathcal {H} } L^- \Vert }_{W^{N_0/2+2,\infty }} + {\Vert L \Vert }_{W^{N_0/2+2,\infty }} \right) }^2 . \end{aligned}$$
(8.72)

Commuting \(\partial _t + b\partial _{\alpha }\) and \(\partial _{\alpha }^j\), using (8.71) and (11.11) in Lemma 11.2, and (8.68), one can see that for any \(0 \le j \le N_0/2\)

$$\begin{aligned}&{\Vert (\partial _t + b\partial _{\alpha }) D^j \partial _{\alpha }b \Vert }_{L^\infty } \lesssim {\Vert \partial _{\alpha }(\partial _t + b\partial _{\alpha }) b \Vert }_{W^{j,\infty }} + {\Vert L \Vert }_{W^{N_0/2+2,\infty }}^2 \\&\quad \lesssim {\Vert \partial _{\alpha }B_3 \Vert }_{W^{N_0/2,\infty }} + {\Vert L \Vert }_{W^{N_0/2+2,\infty }} {\Vert \partial _{\alpha }B_3 \Vert }_{H^{N_0/2+1}} + {\Vert L \Vert }_{W^{N_0/2+2,\infty }}^2. \end{aligned}$$

Since \({\Vert \partial _{\alpha }B_3 \Vert }_{H^{N_0/2+1}}\) can be estimated by means of Corollary 11.4, we only need to bound \({\Vert \partial _{\alpha }B_3 \Vert }_{W^{N_0/2,\infty }}\). \(B_3\) contains potentially dangerous terms, but since they are of the form \(Q_0(L^-, L^-)\) we can use again (11.33) in Proposition 11.7 to obtain

$$\begin{aligned} {\Vert \partial _{\alpha }B_3 \Vert }_{W^{N_0/2,\infty }}&\lesssim {\left( {\Vert {\mathcal {H} } L^- \Vert }_{W^{N_0/2+2,\infty }} + {\Vert L \Vert }_{W^{N_0/2+2,\infty }} \right) }^2 . \end{aligned}$$

This gives (8.72), which in turn allows to bound (8.62c) by the right-hand side of (8.58) as desired.

Estimate of (8.62d). From (8.63) and (8.64) we see that

$$\begin{aligned} (8.62d)&\lesssim \sqrt{E(t)} {\Vert \partial _{\alpha }A \Vert }_{W^{N_0/2,\infty }} + {\Vert \partial _{\alpha }A \Vert }_{H^{N_0}} {\Vert L \Vert }_{W^{N_0/2+2,\infty }} . \end{aligned}$$

To obtain the desired bound it is sufficient to show the following two estimates:

$$\begin{aligned} {\Vert \partial _{\alpha }A \Vert }_{H^{N_0}}&\lesssim \sqrt{E(t)} {\Vert L \Vert }_{W^{N_0/2+2,\infty }} \end{aligned}$$
(8.73)
$$\begin{aligned} {\Vert \partial _{\alpha }(A-1) \Vert }_{W^{N_0/2,\infty }}&\lesssim {\left( {\Vert {\mathcal {H} } L^- \Vert }_{W^{N_0/2+2,\infty }} + {\Vert L \Vert }_{W^{N_0/2+2,\infty }} \right) }^2 . \end{aligned}$$
(8.74)

Recalling the identity (7.16) for \((I- {\mathcal {H} } )(A-1)\) we see that the two terms in the right hand side of that formula are of the same type of the one appearing in the formula (8.67) for \((I- {\mathcal {H} } )b\). Therefore, in order to show (8.73) and (8.74), one can proceed in the exact same fashion as was done before to obtain (8.65) and (8.66). Also in this case the presence of a derivative acting on \(A-1\) in (8.74) plays a crucial role, allowing us to use the bound (11.33) on operators of the type \(\partial _{\alpha }Q_0\).

Proof of (8.59). Since \(S_k = D^k S\) for any \(0 \le k \le N_0/2\), we can write

$$\begin{aligned} {[{\mathcal {P}}, S_k] f = [{\mathcal {P}}, D^k] S + D^k [{\mathcal {P}}, S]}\,. \end{aligned}$$

Thus, to prove (8.59) it is enough to show

$$\begin{aligned} {\Vert [{\mathcal {P}}, D^k] S \chi \Vert }_{L^2}&\lesssim {\left( {\Vert L \Vert }_{W^{N_1,\infty }} + {\Vert {\mathcal {H} } L^- \Vert }_{W^{N_1,\infty }} \right) }^2 \sqrt{E} \end{aligned}$$
(8.75)
$$\begin{aligned} {\Vert [{\mathcal {P}}, S] \chi \Vert }_{H^k}&\lesssim {\left( {\Vert L \Vert }_{W^{N_1,\infty }} + {\Vert {\mathcal {H} } L^- \Vert }_{W^{N_1,\infty }} \right) }^2 \sqrt{E} \end{aligned}$$
(8.76)

for any \(0 \le k \le N_0/2\). Recall that the commutation \([{\mathcal {P}}, D^k]\) is explicitly given in (8.61), whereas a direct computation shows that

$$\begin{aligned}{}[{\mathcal {P}},S]&= {\mathcal {P}}+ \left\{ \left( S b - \frac{1}{2}b \right) b_{\alpha }- S(\partial _t+b\partial _{\alpha })b \right\} \partial _{\alpha }\nonumber \\&\quad - \left( S b - \frac{1}{2}b \right) \left\{ (\partial _t+b\partial _{\alpha })\partial _{\alpha }+ \partial _{\alpha }(\partial _t+b\partial _{\alpha }) \right\} + i S A \partial _{\alpha }. \end{aligned}$$
(8.77)

Proof of (8.75). For \(0 \le k \le N_0/2\), we first use (8.61) to obtain

$$\begin{aligned}&{\Vert [{\mathcal {P}}, D^k] S \chi \Vert }_{L^2}\lesssim {\Vert \partial _{\alpha }b \Vert }_{W^{k-1,\infty }} {\Vert \partial _{\alpha }(\partial _t + b\partial _{\alpha }) S \chi \Vert }_{H^{k-1}}\nonumber \\&\quad + \sum _{j=1}^{k} {\Vert (\partial _t + b\partial _{\alpha }) D^j b \Vert }_{L^\infty } {\Vert \partial _{\alpha }S \chi \Vert }_{H^{k-1}} + {\Vert \partial _{\alpha }b \Vert }_{W^{k-1,\infty }} \nonumber \\&\quad \times \sum _{j=1}^{k} {\Vert (\partial _t + b\partial _{\alpha }) D^j S \chi \Vert }_{L^2} + {\Vert \partial _{\alpha }A \Vert }_{W^{k-1,\infty }} {\Vert \partial _{\alpha }S \chi \Vert }_{H^{k-1}} . \end{aligned}$$
(8.78)

Commuting \(\partial _t + b\partial _{\alpha }\) with \(\partial _{\alpha }\) and \(S\) in the appropriate fashion, using (8.21) to control the \(L^2\) norm of \(\partial _{\alpha }(\partial _t + b\partial _{\alpha }) S_k \chi \), and (8.13) to control \({\Vert \partial _{\alpha }S \chi \Vert }_{H^{k-1}}\), we see that all of the \(L^2\)-based norms in (8.78) are controlled by \(\sqrt{E}\). To obtain (8.75) We are then left with proving that for \(0 \le k \le N_0/2\)

$$\begin{aligned} {\Vert \partial _{\alpha }b \Vert }_{W^{k-1,\infty }}&\lesssim {\left( {\Vert L \Vert }_{W^{N_1,\infty }} + {\Vert {\mathcal {H} } L^- \Vert }_{W^{N_1,\infty }} \right) }^2 \end{aligned}$$
(8.79)
$$\begin{aligned} {\Vert (\partial _t + b\partial _{\alpha }) \partial _{\alpha }b \Vert }_{W^{k-1,\infty }}&\lesssim {\left( {\Vert L \Vert }_{W^{N_1,\infty }} + {\Vert {\mathcal {H} } L^- \Vert }_{W^{N_1,\infty }} \right) }^2 \end{aligned}$$
(8.80)
$$\begin{aligned} {\Vert \partial _{\alpha }A \Vert }_{W^{k-1,\infty }}&\lesssim {\left( {\Vert L \Vert }_{W^{N_1,\infty }} + {\Vert {\mathcal {H} } L^- \Vert }_{W^{N_1,\infty }} \right) }^2 . \end{aligned}$$
(8.81)

The bound (8.79) is implied by (8.68) which has been already proven. Up to commuting \(\partial _t + b\partial _{\alpha }\) and \(\partial _{\alpha }^j\), for \(0\le j \le k-1\), we see that (8.80) would follow from obtaining the same bound for \((\partial _t + b\partial _{\alpha }) \partial _{\alpha }^j \partial _{\alpha }b\). Such an estimate has been already obtained in (8.72). Since also (8.81) has been shown to hold true before, see (8.74), we have completed the proof of (8.75).

Proof of (8.76). Using (8.77), for any \(0 \le k \le N_0/2\), we can estimate

$$\begin{aligned} {\Vert [{\mathcal {P}},S] \chi \Vert }_{H^k}&\lesssim {\Vert {\mathcal {P}}\chi \Vert }_{H^k} + \left( {\left\| S b - \frac{1}{2}b \right\| }_{H^k} {\Vert b_{\alpha }\Vert }_{H^{k+1}} + {\Vert S(\partial _t+b\partial _{\alpha })b \Vert }_{H^k} \right) \\&\quad \times {\Vert \partial _{\alpha }\chi \Vert }_{W^{\frac{N_0}{2}, \infty }}+ {\left\| S b - \frac{1}{2}b \right\| }_{H^k} {\Vert \partial _{\alpha }(\partial _t + b\partial _{\alpha }) \chi \Vert }_{W^{\frac{N_0}{2}, \infty }}\nonumber \\&\quad + {\Vert S (A-1) \Vert }_{H^k} {\Vert \partial _{\alpha }\chi \Vert }_{W^{\frac{N_0}{2}, \infty }} . \end{aligned}$$

From (8.12) we know that

$$\begin{aligned} {\Vert \partial _{\alpha }\chi \Vert }_{W^{\frac{N_0}{2}, \infty }} + {\Vert \partial _{\alpha }(\partial _t+b\partial _{\alpha }) \chi \Vert }_{W^{\frac{N_0}{2}, \infty }} \lesssim {\Vert L \Vert }_{W^{\frac{N_0}{2}, \infty }} . \end{aligned}$$

From Lemma 7.1 we see that \({\Vert b \Vert }_{H^{k+1}} \lesssim 1\). To conclude the desired bound it then suffices to show the following \(L^2\)-estimates:

$$\begin{aligned} {\Vert b \Vert }_{H^k} + {\Vert S b \Vert }_{H^k}&\lesssim \left( {\Vert L \Vert }_{W^{N_1,\infty }} + {\Vert {\mathcal {H} } L^- \Vert }_{W^{N_1,\infty }} \right) \sqrt{E} \end{aligned}$$
(8.82)
$$\begin{aligned} {\Vert S(\partial _t+b\partial _{\alpha }) b \Vert }_{H^k}&\lesssim \left( {\Vert L \Vert }_{W^{N_1,\infty }} + {\Vert {\mathcal {H} } L^- \Vert }_{W^{N_1,\infty }} \right) \sqrt{E} \end{aligned}$$
(8.83)
$$\begin{aligned} {\Vert S (A-1) \Vert }_{H^k}&\lesssim \left( {\Vert L \Vert }_{W^{N_1,\infty }} + {\Vert {\mathcal {H} } L^- \Vert }_{W^{N_1,\infty }} \right) \sqrt{E} \end{aligned}$$
(8.84)

for any \(0 \le k \le N_0/2\).

The \(L^2\) estimates above can all be proven in the same fashion, so we just give details for the first one. As before, from (7.17) we know that \((I- {\mathcal {H} } ) b = Q_0(\widetilde{L},\widetilde{L})\), where \(Q_0\) is as in(11.18). From (11.9) we then have

$$\begin{aligned}&{\Vert b \Vert }_{H^k} + {\Vert S b \Vert }_{H^k} \lesssim {\Vert Q_0(\widetilde{L},\widetilde{L}) \Vert }_{X_k} + {\Vert {\zeta }_{\alpha }-1 \Vert }_{X_k} \big ( {\Vert Q_0(\widetilde{L},\widetilde{L}) \Vert }_{W^{\frac{k}{2},\infty }}\\&\quad + {\Vert \text{ Im }{\zeta }_{\alpha }\Vert }_{W^{k,\infty }} {\Vert Q_0(\widetilde{L},\widetilde{L}) \Vert }_{H^{\frac{k}{2}+1}} \big ) . \end{aligned}$$

Using the estimate (11.30), and the energy bounds (8.2), we see that \({\Vert Q_0(\widetilde{L},\widetilde{L}) \Vert }_{X_k}\) is bounded by the right-hand side of (8.82). Moreover we can use (11.3), and the a priori assumptions, to deduce that

$$\begin{aligned} {\Vert Q_0(\widetilde{L},\widetilde{L}) \Vert }_{H^{\frac{k}{2}+1}} \lesssim {\Vert L \Vert }_{W^{\frac{N_0}{2} + 2,\infty }} {\Vert L \Vert }_{H^{\frac{N_0}{2} + 3}} \lesssim {\Vert L \Vert }_{W^{\frac{N_0}{2} + 2,\infty }} . \end{aligned}$$

This, and \({\Vert {\zeta }_{\alpha }-1 \Vert }_{X_k} \lesssim \sqrt{E}\), suffice to obtain (8.82). One can easily see that (8.83) and (8.84) follow analogously, by using respectively the identities (8.71) and (7.16).

Proof of (8.60). In order to complete the Energy estimate for \(E^\chi \) we want to prove the \(L^\infty \) bound

$$\begin{aligned} { \left\| \frac{a_t}{a} \circ k^{-1} \right\| }_{L^\infty } \lesssim {\Vert \widetilde{L} \Vert }_{W^{N_1,\infty }}^2 . \end{aligned}$$

In what follows we are going to establish the stronger bound

$$\begin{aligned} { \left\| \frac{a_t}{a} \circ k^{-1} \right\| }_{W^{\frac{N_0}{2},\infty }} \lesssim {\Vert \widetilde{L} \Vert }_{W^{N_1,\infty }}^2 \end{aligned}$$
(8.85)

and the additional estimate

$$\begin{aligned} { \left\| \frac{a_t}{a} \circ k^{-1} \right\| }_{X_{N_0}} \lesssim {\Vert \widetilde{L} \Vert }_{W^{N_1,\infty }} {\Vert \widetilde{L} \Vert }_{X_{N_0}} . \end{aligned}$$
(8.86)

The above bounds will also be useful later on. To prove (8.85)–(8.86) we will use the formula (2.32) from [45], which reads

$$\begin{aligned} (I - {\mathcal {H} } )&\left( \frac{a_t}{a} \circ k^{-1} A \overline{{\zeta }}_{\alpha }\right) = I_1 + I_2 \nonumber \\ I_1&:= 2i [w, {\mathcal {H} } ] \frac{\overline{u}_{\alpha }}{{\zeta }_{\alpha }} + 2i [u, {\mathcal {H} } ] \frac{\overline{w}_{\alpha }}{{\zeta }_{\alpha }} \nonumber \\ I_2&:= - \frac{1}{\pi } \int { \left( \frac{ u({\alpha }) - u({\beta }) }{ {\zeta }({\alpha })-{\zeta }({\beta }) } \right) }^2 \overline{u}_{\beta }({\beta }) \, d{\beta }, \end{aligned}$$
(8.87)

in combination with the following Lemma:

Lemma 8.8

Let \(f \in X_k\), \(0 \le k \le N_0\), let \(A\) be as in (7.16), and \(w\) as defined in (7.3). Assume that \(f\) and \(g\) are related by

$$\begin{aligned} (I - {\mathcal {H} } ) (f A \overline{{\zeta }}_{\alpha }) = g . \end{aligned}$$

Then, for any \(0 \le k \le N_0\)

$$\begin{aligned} {\Vert f \Vert }_{X_k}&\lesssim {\Vert g \Vert }_{X_k} + \left( {\Vert w \Vert }_{X_k} + {\Vert {\zeta }_{\alpha }- 1 \Vert }_{X_k} \right) \nonumber \\&\quad \times \left( {\Vert g \Vert }_{W^{\frac{k}{2},\infty }} + {\Vert \text{ Im }{\zeta }_{\alpha }\Vert }_{W^{\frac{k}{2}+1,\infty }} {\Vert (g,w) \Vert }_{H^{\frac{k}{2}+1,\infty }} \right) , \end{aligned}$$
(8.88)

and for \(0 \le k \le \frac{N_0}{2} + 2\)

$$\begin{aligned} {\Vert f \Vert }_{W^{k,\infty }} \lesssim {\Vert \text{ Re }\, g \Vert }_{W^{k,\infty }} + {\Vert \text{ Im }{\zeta }_{\alpha }\Vert }_{W^{k+1,\infty }} {\Vert g \Vert }_{H^{k + 1}} + {\Vert g \Vert }_{H^{k+1}} {\Vert w \Vert }_{W^{k+1,\infty }} . \end{aligned}$$
(8.89)

Proof

From the identity (8.135) we see that \(A \overline{{\zeta }_{\alpha }} = 1 - i w\), and therefore

$$\begin{aligned} (I - {\mathcal {H} } )f = g + (I- {\mathcal {H} } ) (i f w) . \end{aligned}$$

(8.88) can then be derived by applying (11.9) to the above identity, using (11.2) and the a priori bounds on \(w\). The estimate (8.89) follows similarly from (11.10), together with (11.7) and (11.3). \(\square \)

Notice that \(I_2\) in (8.87) is of the form \(\mathbf{T}(u,u,\overline{u}_{\alpha })\), and therefore is easier to estimate, so that we can skip its treatment and focus on \(I_1\). From Lemma 8.8 we see that in order to prove (8.85) it suffices to obtain the bounds

$$\begin{aligned} {\Vert \text{ Re }I_1 \Vert }_{W^{\frac{N_0}{2}+1,\infty }}&\lesssim {\Vert \widetilde{L} \Vert }_{W^{N_1,\infty }}^2 \end{aligned}$$
(8.90)
$$\begin{aligned} {\Vert I_1 \Vert }_{H^{\frac{N_0}{2}+1}}&\lesssim {\Vert \widetilde{L} \Vert }_{W^{N_1,\infty }} . \end{aligned}$$
(8.91)

For (8.86) it is enough to prove

$$\begin{aligned} {\Vert I_1 \Vert }_{X_{N_0}}&\lesssim {\Vert \widetilde{L} \Vert }_{W^{N_1,\infty }} {\Vert \widetilde{L} \Vert }_{X_{N_0}} . \end{aligned}$$
(8.92)

To show (8.90) we write explicitly \(\text{ Re }I_1\) as follows:

$$\begin{aligned} \text{ Re }I_1&= \text{ Re }\left( \frac{2}{\pi } \int \frac{ w({\alpha }) - w({\beta }) }{ {\zeta }({\alpha })-{\zeta }({\beta }) } \overline{u}_{\beta }({\beta }) \, d{\beta }+ \frac{2}{\pi } \int \frac{ u({\alpha }) - u({\beta }) }{ {\zeta }({\alpha })-{\zeta }({\beta }) } \overline{w}_{\beta }({\beta }) \, d{\beta }\right) \nonumber \\&= \text{ Re }\left( w 2i {\mathcal {H} } \frac{\overline{u}_{\alpha }}{{\zeta }_{\alpha }} + u 2i {\mathcal {H} } \frac{\overline{w}_{\alpha }}{{\zeta }_{\alpha }} \right) \nonumber \\&\quad - \text{ Re }\left( \frac{2}{\pi } \int \frac{w({\beta }) \overline{u}_{\beta }({\beta }) + u({\beta }) \overline{w}_{\beta }({\beta })}{ {\zeta }({\alpha })-{\zeta }({\beta }) } \, d{\beta }\right) . \end{aligned}$$
(8.93)

The first contribution above is estimated using (11.36):

$$\begin{aligned} {\left\| w 2i {\mathcal {H} } \frac{\overline{u}_{\alpha }}{{\zeta }_{\alpha }} + u 2i {\mathcal {H} } \frac{\overline{w}_{\alpha }}{{\zeta }_{\alpha }} \right\| }_{W^{\frac{N_0}{2}+1,\infty }} \lesssim {\Vert w \Vert }_{W^{N_1,\infty }} {\Vert u \Vert }_{W^{N_1,\infty }} . \end{aligned}$$
(8.94)

To bound the second summand in (8.93) we use the identity

$$\begin{aligned}&2 \text{ Re }\int \frac{f({\beta })}{{\zeta }({\alpha }) - {\zeta }({\beta }) }\, d{\beta }= \int \frac{2 \text{ Re }f({\beta })}{{\zeta }({\alpha }) - {\zeta }({\beta }) }\, d{\beta }\\&\quad + 2i \int \frac{\overline{f}({\beta }) (\text{ Im }{\zeta }({\alpha }) - \text{ Im }{\zeta }({\beta })) }{{|{\zeta }({\alpha }) - {\zeta }({\beta })|}^2}\, d{\beta }\end{aligned}$$

and notice that

$$\begin{aligned} 2 \text{ Re }\left( w \overline{u}_{\beta }+ u \overline{w}_{\beta }\right) = \partial _{\beta }(w \overline{u} + u \overline{w} ) . \end{aligned}$$

It follows that

$$\begin{aligned} \text{ Re }\left( \frac{2}{\pi } \int \frac{w({\beta }) \overline{u}_{\beta }({\beta }) + u({\beta }) \overline{w}_{\beta }({\beta })}{ {\zeta }({\alpha })-{\zeta }({\beta }) } \, d{\beta }\right) = 2 i {\mathcal {H} } \frac{ \partial _{\alpha }(w \overline{u} + u \overline{w} ) }{{\zeta }_{\alpha }} + I_r \end{aligned}$$
(8.95)

where \(I_r\) is cubic remainder which can be easily estimated. Since

$$\begin{aligned} {\left\| {\mathcal {H} } \frac{ \partial _{\alpha }(w \overline{u} + \overline{u} w ) }{{\zeta }_{\alpha }} \right\| }_{W^{\frac{N_0}{2}+1,\infty }} \lesssim {\Vert w \Vert }_{W^{\frac{N_0}{2}+3,\infty }} {\Vert u \Vert }_{W^{\frac{N_0}{2}+3,\infty }} \end{aligned}$$
(8.96)

we have concluded the proof of (8.90). The estimates (8.91) and (8.92) follow respectively from (11.3), and the bound (11.31) provided by Proposition 11.7.

8.2.3 Evolution of \(E^{\lambda }\)

The energy associated with \({\lambda }\) is given by

$$\begin{aligned} E^{{\lambda }} (t)&= \sum _{k=0}^{N_0-2} \int \frac{1}{A} {\left| (\partial _t + b \partial _{\alpha }) D^k {\lambda }\right| }^2 + i {\left( D^k {\lambda }\right) }^h \partial _\alpha \overline{{\left( D^k {\lambda }\right) }^h} \, d{\alpha }\\&\quad + \sum _{k=0}^{N_0/2} \int \frac{1}{A} {\left| (\partial _t + b \partial _{\alpha }) S_k {\lambda }\right| }^2 + i{\left( S_k {\lambda }\right) }^h \partial _\alpha \overline{{\left( S_k {\lambda }\right) }^h} \, d{\alpha }. \end{aligned}$$

From (7.20) and (7.23) we have \({(\partial _t + b \partial _{\alpha })}^2 {\lambda }+ i A \partial _\alpha {\lambda }= \sum _{j=1}^4 G^{{\lambda }}_j\), where

$$\begin{aligned} G^{{\lambda }}_1&= - \left[ u, {\mathcal {H} } \frac{1}{{\zeta }_{\alpha }} + \overline{ {\mathcal {H} } }\frac{1}{\overline{{\zeta }}_{\alpha }} \right] (\overline{{\zeta }}_{\alpha }w ) \end{aligned}$$
(8.97)
$$\begin{aligned} G^{{\lambda }}_2&= [u,\overline{ {\mathcal {H} } }] \left( \overline{u} \frac{u_{\alpha }}{\overline{{\zeta }}_{\alpha }} \right) + u [u, {\mathcal {H} } ] \frac{\overline{u}_{\alpha }}{{\zeta }_{\alpha }} \end{aligned}$$
(8.98)
$$\begin{aligned} G^{{\lambda }}_3&= - 2 [u, {\mathcal {H} } ] \frac{u \cdot u_{\alpha }}{{\zeta }_{\alpha }} \end{aligned}$$
(8.99)
$$\begin{aligned} G^{{\lambda }}_4&= \frac{1}{i\pi } \int { \left( \frac{ u({\alpha }) - u({\beta }) }{ {\zeta }({\alpha })-{\zeta }({\beta }) } \right) }^2 u ({\beta }) \cdot {\zeta }_{\beta }({\beta }) \, d{\beta }. \end{aligned}$$
(8.100)

As already done for \(\chi \) before, we use Proposition 8.7 to compute

$$\begin{aligned} \frac{d}{dt} E^{\lambda }(t)&= \sum _{k=0}^{N_0-2} \int \frac{2}{A} \text{ Re }\left( (\partial _t + b \partial _{\alpha }) D^k {\lambda }\, {\mathcal {P}}D^k {\lambda }\right) \nonumber \\&\quad - \frac{1}{A} \frac{a_t}{a} \circ k^{-1} {\left| (\partial _t + b \partial _{\alpha }) D^k {\lambda }\right| }^2 \, d{\alpha }\end{aligned}$$
(8.101)
$$\begin{aligned}&\quad - 2 \sum _{k=0}^{N_0-2} \text{ Re }\int i\partial _t (D^k {\lambda })^h \partial _{\alpha }\overline{(D^k {\lambda })^r}\nonumber \\&\quad + i\partial _t (D^k {\lambda })^r \partial _{\alpha }\overline{(D^k {\lambda })^h} + i \partial _t (D^k {\lambda })^r \partial _{\alpha }\overline{(D^k {\lambda })^r} \, d{\alpha }\end{aligned}$$
(8.102)
$$\begin{aligned}&\quad + \sum _{k=0}^{N_0/2} \int \frac{2}{A} \text{ Re }\left( (\partial _t + b \partial _{\alpha }) S_k {\lambda }\, {\mathcal {P}}S_k {\lambda }\right) \nonumber \\&\quad - \frac{1}{A} \frac{a_t}{a} \circ k^{-1} {\left| (\partial _t + b \partial _{\alpha }) S_k {\lambda }\right| }^2 \, d{\alpha }\end{aligned}$$
(8.103)
$$\begin{aligned}&\quad - 2 \sum _{k=0}^{N_0/2} \text{ Re }\int i\partial _t (S_k {\lambda })^h \partial _{\alpha }\overline{(S_k {\lambda })^r} + i\partial _t (S_k {\lambda })^r \partial _{\alpha }\overline{(S_k {\lambda })^h}\nonumber \\&\quad + i \partial _t (S_k {\lambda })^r \partial _{\alpha }\overline{(S_k {\lambda })^r} \, d{\alpha }. \end{aligned}$$
(8.104)

Since \({\Vert A-1 \Vert }_{L^\infty } \le \frac{1}{2}\) we see that

$$\begin{aligned} (8.101) + (8.103)&\lesssim \sqrt{E^{\lambda }} (t) \sum _{k=0}^{N_0-2} \left( {\Vert D^k G^{{\lambda }} \Vert }_{L^2} + {\Vert [{\mathcal {P}},D^k] {\lambda }\Vert }_{L^2} \right) \end{aligned}$$
(8.105)
$$\begin{aligned}&\quad + \sqrt{E^{\lambda }} (t) \sum _{k=0}^{N_0/2} \left( {\Vert S_k G^{{\lambda }} \Vert }_{L^2} + {\Vert [{\mathcal {P}},S_k] {\lambda }\Vert }_{L^2} \right) \end{aligned}$$
(8.106)
$$\begin{aligned}&\quad + E^{\lambda }(t) { \left\| \frac{a_t}{a} \circ k^{-1} \right\| }_{L^\infty } . \end{aligned}$$
(8.107)

As before, the terms in (8.102) and (8.104) are remainder terms: there is no logarithmic loss in estimating them already in [45], so we can disregard them. Moreover \({ \Vert a_t/a \circ k^{-1} \Vert }_{L^\infty }\) has been already estimated in Sect. 8.2.2, see (8.85). Also, the terms \({\Vert [{\mathcal {P}},D^k] {\lambda }\Vert }_{L^2}\) and \({\Vert [{\mathcal {P}},S_k] {\lambda }\Vert }_{L^2}\) can be treated exactly as done in Sect. 8.2.2. Therefore, to control the time evolution of \(E^{\lambda }\) by the right-hand side of (8.44) it suffices to show

$$\begin{aligned} {\Vert G^{\lambda }\Vert }_{X_{N_0}} \lesssim {\Vert L(t) \Vert }_{W^{N_1,\infty }}^2 {\Vert \widetilde{L} \Vert }_{X_{N_0}} . \end{aligned}$$
(8.108)

This is done in the following sections by estimating each of the terms in (8.97)–(8.100).

Estimate of (8.97). Observe that

$$\begin{aligned} G_1^{\lambda }= - \frac{2}{\pi } \int \frac{ (u({\alpha }) - u({\beta })) (\text{ Im }{\zeta }({\alpha }) - \text{ Im }{\zeta }({\beta })) }{ {|{\zeta }({\alpha })-{\zeta }({\beta })|}^2 } w({\beta }) \overline{{\zeta }}_{\beta }({\beta }) \, d{\beta }\end{aligned}$$

and therefore it is an operator of the form \(\mathbf{T}(u,\text{ Im }{\zeta }, w {\zeta }_{\alpha })\) which can be estimated by means of Proposition 11.7.

Estimate of (8.98). The term \(G_2^{\lambda }\) is more delicate. In order to estimate it we need to exploit its special structure, which allows the appearance of the Hilbert transform acting on products only when the arguments are perfect derivative of functions that we can control. Let us start by explicitly rewriting \(G_2^{\lambda }\) as

$$\begin{aligned} G_2^{\lambda }&= u \overline{ {\mathcal {H} } } \left( \overline{u} \frac{u_{\alpha }}{\overline{{\zeta }}_{\alpha }} \right) - \overline{ {\mathcal {H} } } \left( u \overline{u} \frac{u_{\alpha }}{\overline{{\zeta }}_{\alpha }} \right) + u^2 {\mathcal {H} } \frac{\overline{u}_{\alpha }}{{\zeta }_{\alpha }} - u {\mathcal {H} } \left( u \frac{\overline{u}_{\alpha }}{{\zeta }_{\alpha }} \right) . \end{aligned}$$
(8.109)

We can then apply \(D^k\) for \(k=1,\dots ,N_0\), and \(S_k\), for \(k=0,\dots ,N_0/2\), to the above expression and use the commutation identities to distribute them. This procedure will give many terms, most of which can be estimated directly using the \(L^2\)-bounds given by Proposition 11.7. There will only two be types of dangerous terms:

  1. 1.

    Terms for which \(D^{N_0}\) or \(S_{N_0/2}\) fall on \(u_{\alpha }\);

  2. 2.

    Terms that require the estimate in \(L^\infty \) of an Hilbert transform of a product, such as \( {\mathcal {H} } \frac{1}{{\zeta }_{\alpha }} u \overline{u}_{\alpha }\).

More precisely, denoting by \({\Gamma }^{N_0}\) either \(D^{N_0}\) or \(S_{N_0/2}\), all the dangerous terms are:

$$\begin{aligned} A_1&= {\Gamma }^{N_0} u \overline{ {\mathcal {H} } } \left( \overline{u} \frac{u_{\alpha }}{\overline{{\zeta }}_{\alpha }} \right) , \,\, A_2 = u \overline{ {\mathcal {H} } } \left( \overline{u} \frac{{\Gamma }^{N_0} u_{\alpha }}{\overline{{\zeta }}_{\alpha }} \right) , \,\, A_3 = - \overline{ {\mathcal {H} } } \left( u \overline{u} \frac{{\Gamma }^{N_0} u_{\alpha }}{\overline{{\zeta }}_{\alpha }} \right) \\ A_4&= 2 u {\Gamma }^{N_0} u \overline{ {\mathcal {H} } } \frac{u_{\alpha }}{\overline{{\zeta }}_{\alpha }} ,\,\, A_5 = u^2 {\mathcal {H} } \frac{{\Gamma }^{N_0} \overline{u}_{\alpha }}{{\zeta }_{\alpha }} , \,\, A_6 = - {\Gamma }^{N_0} u {\mathcal {H} } \left( u \frac{\overline{u}_{\alpha }}{{\zeta }_{\alpha }} \right) , \,\, \\ A_7&= - u {\mathcal {H} } \left( u \frac{{\Gamma }^{N_0} \overline{u}_{\alpha }}{{\zeta }_{\alpha }} \right) . \end{aligned}$$

In particular, using Propositions 11.7 and 11.4, one can verify that

$$\begin{aligned}&{\left\| {\Gamma }^{N_0} \left( u \overline{ {\mathcal {H} } } \left( \overline{u} \frac{u_{\alpha }}{\overline{{\zeta }}_{\alpha }} \right) \right) - (A_1 + A_2) \right\| }_{L^2} + {\left\| - {\Gamma }^{N_0} \overline{ {\mathcal {H} } } \left( u \overline{u} \frac{u_{\alpha }}{\overline{{\zeta }}_{\alpha }} \right) - A_3 \right\| }_{L^2} \\&\qquad + {\left\| {\Gamma }^{N_0} \left( u^2 {\mathcal {H} } \frac{\overline{u}_{\alpha }}{{\zeta }_{\alpha }} \right) - (A_4 + A_5) \right\| }_{L^2}\\&\qquad + {\left\| - {\Gamma }^{N_0} \left( u {\mathcal {H} } \left( u \frac{\overline{u}_{\alpha }}{{\zeta }_{\alpha }} \right) \right) - (A_6 + A_7) \right\| }_{L^2} \lesssim {\Vert u \Vert }_{W^{N_1,\infty }}^2 {\Vert u \Vert }_{X_{N_0}}. \end{aligned}$$

It follows that

$$\begin{aligned} {\left\| {\Gamma }^{N_0} G_2^{\lambda }\right\| }_{L^2} \lesssim \left\| \sum _{j=1}^7 A_j \right\| _{L^2} + {\Vert u \Vert }_{W^{N_1,\infty }}^2 {\Vert u \Vert }_{X_{N_0}} . \end{aligned}$$

To estimate the terms \(A_j\) we need to combine them appropriately. More precisely we look at the combinations

$$\begin{aligned} A_2 + A_3&= -\frac{1}{i\pi } \int \frac{(u({\alpha }) - u({\beta })) \overline{u}({\beta })}{ \overline{{\zeta }}({\alpha })-\overline{{\zeta }}({\beta }) } {\Gamma }^{N_0} u_{\beta }({\beta }) \, d{\beta }\end{aligned}$$
(8.110)
$$\begin{aligned} A_5 + A_7&= \frac{1}{i\pi } u({\alpha }) \int \frac{u({\alpha }) - u({\beta })}{ {\zeta }({\alpha })-{\zeta }({\beta }) } {\Gamma }^{N_0} \overline{u}_{\beta }({\beta }) \, d{\beta }\end{aligned}$$
(8.111)
$$\begin{aligned} A_1 + A_6&= {\Gamma }^{N_0} u \left( \overline{ {\mathcal {H} } } \left( \overline{u} \frac{u_{\alpha }}{\overline{{\zeta }}_{\alpha }} \right) - {\mathcal {H} } \left( u \frac{\overline{u}_{\alpha }}{{\zeta }_{\alpha }} \right) \right) . \end{aligned}$$
(8.112)

The remaining term \(A_4\) can be directly bounded using Lemma 11.8

$$\begin{aligned} {\left\| A_4 \right\| }_{L^2} \lesssim {\left\| u {\Gamma }^{N_0} u \overline{ {\mathcal {H} } } \frac{u_{\alpha }}{\overline{{\zeta }}_{\alpha }} \right\| }_{L^2} \lesssim {\Vert u \Vert }_{L^\infty } {\Vert {\Gamma }^{N_0} u \Vert }_{L^2} {\left\| \overline{ {\mathcal {H} } } \frac{u_{\alpha }}{\overline{{\zeta }}_{\alpha }} \right\| }_{L^\infty } \lesssim {\Vert u \Vert }_{W^{N_1,\infty }}^2 {\Vert u \Vert }_{X_{N_0}} . \end{aligned}$$

The terms (8.110) and (8.111) can be estimated by means of (11.31) (case \(k=0\)) in Proposition 11.7 since, upon commuting \({\Gamma }^{N_0}\) and \(\partial _{\alpha }\), they are linear combinations of terms of the form

$$\begin{aligned} Q_0 (u, \partial _{\alpha }\overline{u}{\Gamma }^{N_0} u) , \quad Q_0 \left( u, \partial _{\alpha }(\overline{u}{\Gamma }^{N_0}u) \right) , \quad Q_0 (u,\overline{u} {\Gamma }^{N_0} u) \quad \text{ or } \quad u Q_0 (u, \partial _{\alpha }{\Gamma }^{N_0} u) . \end{aligned}$$

Notice that there are no singular integrals that need to be estimated in \(L^\infty \) here.

To bound (8.112) we rewrite it as:

$$\begin{aligned} (8.112)&= {\Gamma }^{N_0} u\left( \overline{ {\mathcal {H} } } \frac{1}{\overline{{\zeta }}_{\alpha }} (\overline{u} u_{\alpha }) + {\mathcal {H} } \frac{1}{{\zeta }_{\alpha }} ( \overline{u} u_{\alpha }) \right) - {\Gamma }^{N_0} u\left( {\mathcal {H} } \frac{1}{{\zeta }_{\alpha }} (\overline{u} u_{\alpha }) + {\mathcal {H} } \frac{1}{{\zeta }_{\alpha }} ( u \overline{u}_{\alpha }) \right) \nonumber \\&= {\Gamma }^{N_0} u \left( \overline{ {\mathcal {H} } } \frac{1}{\overline{{\zeta }}_{\alpha }} + {\mathcal {H} } \frac{1}{{\zeta }_{\alpha }} \right) ( \overline{u} u_{\alpha })- {\Gamma }^{N_0} u \, {\mathcal {H} } \frac{1}{{\zeta }_{\alpha }} \partial _{\alpha }( \overline{u} u ) =: B_1 + B_2. \end{aligned}$$
(8.113)

The first term \(B_1\) in (8.113) is quartic and can therefore be easily estimated. The second term \(B_2\) can instead be bounded by means of Lemma 11.8 using the fact that the argument of the Hilbert transform is a perfect derivative:

$$\begin{aligned} {\Vert B_2 \Vert }&\lesssim {\Vert {\Gamma }^{N_0} u \Vert }_{L^2} {\left\| {\mathcal {H} } \frac{1}{{\zeta }_{\alpha }} \partial _{\alpha }( \overline{u} u ) \right\| }_{L^\infty } \lesssim {\Vert u \Vert }_{X_{N_0}} {\Vert u^2 \Vert }_{W^{N_1,\infty }} . \end{aligned}$$

Estimate of (8.99). The term \(G_3^{\lambda }\) can be treated similarly to \(G_2^{\lambda }\). Using the same notation as above, and indicating with “\(+ \cdots \)” harmless terms that can be controlled directly by means of Propositions 11.7 or 11.4, we can write

$$\begin{aligned}&{\Gamma }^{N_0} G_3^{\lambda }= - 2 {\Gamma }^{N_0} \left( [u, {\mathcal {H} } ] \frac{u \cdot u_{\alpha }}{{\zeta }_{\alpha }} \right) = - 2 {\Gamma }^{N_0} u {\mathcal {H} } \frac{u \cdot u_{\alpha }}{{\zeta }_{\alpha }} - 2 [u, {\mathcal {H} } ] \frac{u \cdot {\Gamma }^{N_0} u_{\alpha }}{{\zeta }_{\alpha }} + \cdots . \end{aligned}$$
(8.114)

Since \(u \cdot u_\alpha = \text{ Re }\,\,(u \bar{u}_\alpha )\) the first summand in (8.114) is like \(B_2\) above. The second summand in (8.114) is of the same form as (8.110) above, and therefore can be estimated similarly.

Estimate of (8.100). The term \(G_4^{\lambda }\) is of the type \(\mathbf{T}(\widetilde{L},\widetilde{L},\widetilde{L})\) and can therefore be treated directly using (11.34) in Proposition 11.7. This concludes the proof of (8.108), hence of (8.44) for \(f = {\lambda }\).

8.2.4 Evolution of \(E^v\)

The energy associated with \(v\) is given by

$$\begin{aligned} E^{v} (t)&=\sum _{k=0}^{N_0} \int \frac{1}{A} {\left| (\partial _t + b \partial _{\alpha }) D^k v_1 \right| }^2 + i D^k v_1 \partial _\alpha \overline{D^k v_1} \, d{\alpha }\\&\quad + \sum _{k=0}^{N_0/2} \int \frac{1}{A} {\left| (\partial _t + b \partial _{\alpha }) S_k v_1 \right| }^2+ i S_k v_1 \, \partial _\alpha \overline{S_k v_1} \, d{\alpha }\,. \end{aligned}$$

Here \(v_1 = (I- {\mathcal {H} } ) v\) and from (7.19), (7.22), (7.26) we have \({(\partial _t + b \partial _{\alpha })}^2 v_1 + i A \partial _\alpha v_1 = \sum _{j=1}^5 G^{v_1}_j\), with

$$\begin{aligned} G^{v_1}_1&= (I- {\mathcal {H} } ) {\mathcal {P}}v \end{aligned}$$
(8.115)
$$\begin{aligned} G^{v_1}_2&= - 2 [u, {\mathcal {H} } ] \frac{\partial _{\alpha }}{{\zeta }_{\alpha }} {\mathcal {P}}\chi \end{aligned}$$
(8.116)
$$\begin{aligned} G^{v_1}_3&= - 2[u, {\mathcal {H} } ] \frac{\partial _{\alpha }}{{\zeta }_{\alpha }} \left( w \frac{\partial _{\alpha }}{{\zeta }_{\alpha }} \chi \right) \end{aligned}$$
(8.117)
$$\begin{aligned} G^{v_1}_4&= -i [( {\mathcal {H} } + \overline{ {\mathcal {H} } })u, {\mathcal {H} } ] {\left( \frac{\partial _{\alpha }}{{\zeta }_{\alpha }} \right) }^2 \chi \end{aligned}$$
(8.118)
$$\begin{aligned} G^{v_1}_5&= \frac{1}{i\pi } \int { \left( \frac{ u({\alpha }) - u({\beta }) }{ {\zeta }({\alpha })-{\zeta }({\beta }) } \right) }^2 v_{\beta }({\beta }) \, d{\beta }. \end{aligned}$$
(8.119)

Using (8.48) we compute

$$\begin{aligned} \frac{d}{dt} E^v (t)&= \sum _{k=0}^{N_0} \int \frac{2}{A} \text{ Re }\left( (\partial _t \!+\! b \partial _{\alpha }) D^k v_1 \, {\mathcal {P}}D^k v_1 \right) - \frac{1}{A} \frac{a_t}{a} \circ k^{-1} {\left| (\partial _t + b \partial _{\alpha }) D^k v_1 \right| }^2 \, d{\alpha }\\&\quad + \sum _{k=0}^{N_0/2} \int \frac{2}{A} \text{ Re }\left( (\partial _t + b \partial _{\alpha }) S_k v_1 \, {\mathcal {P}}S_k v_1 \right) - \frac{1}{A} \frac{a_t}{a} \circ k^{-1} {\left| (\partial _t + b \partial _{\alpha }) S_k v_1 \right| }^2 \, d{\alpha }. \end{aligned}$$

Since \({\Vert A-1 \Vert }_{L^\infty } \le \frac{1}{2}\) we see that

$$\begin{aligned} \frac{d}{dt} E^v (t)&\lesssim \sqrt{E^{v_1}} (t) \sum _{k=0}^{N_0} \left( {\Vert D^k G^{v_1} \Vert }_{L^2} + {\Vert [{\mathcal {P}},D^k] v_1 \Vert }_{L^2} \right) \\&\quad + \sqrt{E^{v_1}} (t) \sum _{k=0}^{N_0/2} \left( {\Vert S_k G^{v_1} \Vert }_{L^2} + {\Vert [{\mathcal {P}},S_k] v_1 \Vert }_{L^2} \right) + E^{v_1} (t) { \left\| \frac{a_t}{a} \circ k^{-1} \right\| }_{L^\infty } . \end{aligned}$$

Since \({ \Vert a_t/a \circ k^{-1} \Vert }_{L^\infty }\) has been already estimated in (8.85), and \({\Vert [{\mathcal {P}},D^k] v_1 \Vert }_{L^2}\) and \({\Vert [{\mathcal {P}},S] {v_1} \Vert }_{L^2}\) can be bounded as in Sect. 8.2.2, we only need to suitably control \({\Vert G^{v_1} \Vert }_{X_{N_0}}\) by showing

$$\begin{aligned} {\Vert G^{v_1} \Vert }_{X_{N_0}} \lesssim {\left( {\Vert L(t) \Vert }_{W^{N_1,\infty }} + {\Vert {\mathcal {H} } L^-(t) \Vert }_{W^{N_1,\infty }} \right) }^2 {\Vert L \Vert }_{X_{N_0}} . \end{aligned}$$
(8.120)

This bound is proven below by separately estimating each of the terms (8.115)–(8.119).

Estimate of (8.115). From (11.2) we see that

$$\begin{aligned} {\Vert G^{v_1}_1 \Vert }_{X_{N_0}} \lesssim {\Vert {\mathcal {P}}v \Vert }_{X_{N_0}} + {\Vert \widetilde{L} \Vert }_{X_{N_0}} {\Vert {\mathcal {P}}v\Vert }_{H^{{N_0/2}+1}} . \end{aligned}$$
(8.121)

Since the second summand above is easier to estimate (it is a quartic expression) we just show how to control the \(X_{N_0}\)-norm of \({\mathcal {P}}v\) in (7.22). With the exception of the last term, \(i (a_t/a) \circ k^{-1} \, A \partial _{\alpha }\chi \), these are all terms of the form \(\mathbf{T}(L,L,L_{\alpha })\) for which (11.35) applies directly. To estimate the remaining term we first notice that we can essentially replace \(A\) with \(1\) in view of Lemma 7.1 and the estimates (8.85) and (8.86). We then have

$$\begin{aligned} {\left\| \frac{a_t}{a} \circ k^{-1} \, \partial _{\alpha }\chi \right\| }_{X_{N_0}} \lesssim {\left\| \frac{a_t}{a} \circ k^{-1} \right\| }_{X_{N_0}} {\Vert \partial _{\alpha }\chi \Vert }_{W^{\frac{N_0}{2},\infty }} + {\left\| \frac{a_t}{a} \circ k^{-1} \right\| }_{W^{\frac{N_0}{2},\infty }} {\Vert \partial _{\alpha }\chi \Vert }_{X_{N_0}} . \end{aligned}$$

Using (8.85)–(8.86) to bound the norms of \((a_t/a) \circ k^{-1}\), and (8.32)–(8.33) to control the norms of \(\partial _{\alpha }\chi \) we see

$$\begin{aligned} {\left\| \frac{a_t}{a} \circ k^{-1} \, A \partial _{\alpha }\chi \right\| }_{X_{N_0}} \lesssim {\Vert L(t) \Vert }_{W^{N_1,\infty }}^2 {\Vert L \Vert }_{X_{N_0}} \end{aligned}$$

as desired.

Estimate of (8.116). Since \(P \chi \) is a cubic term, (8.116) is a quartic term. Moreover it is of the form \(Q_0(u, \partial _{\alpha }{\mathcal {P}}\chi )\). Thus we can use (11.31) to obtain

$$\begin{aligned} {\left\| (8.116) \right\| }_{X_{N_0}}&\lesssim {\Vert u \Vert }_{W^{\frac{N_0}{2}+2,\infty }} {\Vert {\mathcal {P}}\chi \Vert }_{X_{N_0}} + {\Vert {\mathcal {P}}\chi \Vert }_{W^{\frac{N_0}{2}+2,\infty }} {\Vert u \Vert }_{X_{N_0}} \\&\lesssim {\varepsilon }_1 {\Vert {\mathcal {P}}\chi \Vert }_{X_{N_0}} + {\Vert {\mathcal {P}}\chi \Vert }_{W^{\frac{N_0}{2}+3,\infty }} {\Vert u \Vert }_{X_{N_0}} . \end{aligned}$$

One can then use bounds obtained previously on \({\Vert {\mathcal {P}}\chi \Vert }_{X_{N_0}}\) and the estimate

$$\begin{aligned} {\Vert {\mathcal {P}}\chi \Vert }_{W^{\frac{N_0}{2}+3,\infty }} \lesssim {\varepsilon }_1 {\Vert L \Vert }_{W^{\frac{N_0}{2},\infty }}^2 , \end{aligned}$$

which is easy to derive, to deduce the desired bound for (8.116).

Estimate of (8.117). This term is of the form

$$\begin{aligned} (8.117) = Q_0 \left( u, \partial _{\alpha }\left( w \frac{\partial _{\alpha }}{{\zeta }_{\alpha }} \chi \right) \right) . \end{aligned}$$

Applying (11.31), followed by product estimates, (8.32)–(8.33), and the use of (8.13), we see that

$$\begin{aligned} {\left\| (8.117) \right\| }_{X_{N_0}}&\lesssim {\Vert u \Vert }_{W^{\frac{N_0}{2}+2,\infty }} {\left\| w \frac{\partial _{\alpha }}{{\zeta }_{\alpha }} \chi \right\| }_{X_{N_0}} + {\left\| w \frac{\partial _{\alpha }}{{\zeta }_{\alpha }} \chi \right\| }_{W^{\frac{N_0}{2}+2,\infty }} {\Vert u \Vert }_{X_{N_0}} \\&\quad + {\left\| u \right\| }_{W^{\frac{N_0}{2}+1,\infty }} {\left\| w \frac{\partial _{\alpha }}{{\zeta }_{\alpha }} \chi \right\| }_{W^{\frac{N_0}{2}+1,\infty }} {\Vert {\zeta }_{\alpha }-1 \Vert }_{X_{N_0}} \lesssim {\Vert L \Vert }_{X_{N_0}} {\Vert L \Vert }_{W^{N_1,\infty }}^2 . \end{aligned}$$

Estimate of (8.118). This contribution can also be written in terms of the operator \(Q_0\) as

$$\begin{aligned} (8.118) = Q_0 \left( ( {\mathcal {H} } + \overline{ {\mathcal {H} } }) u, \partial _{\alpha }\left( \frac{\partial _{\alpha }}{{\zeta }_{\alpha }} \chi \right) \right) . \end{aligned}$$

We can then use (11.31) to obtain

$$\begin{aligned}&{\left\| (8.118) \right\| }_{X_{N_0}} \lesssim {\Vert \partial _{\alpha }( {\mathcal {H} } + \overline{ {\mathcal {H} } }) u \Vert }_{W^{\frac{N_0}{2}+1,\infty }} {\left\| \frac{\partial _{\alpha }}{{\zeta }_{\alpha }} \chi \right\| }_{X_{N_0}}\\&\quad + {\left\| \frac{\partial _{\alpha }}{{\zeta }_{\alpha }} \chi \right\| }_{W^{\frac{N_0}{2}+1,\infty }} {\Vert ( {\mathcal {H} } + \overline{ {\mathcal {H} } }) u \Vert }_{X_{N_0}} + {\left\| ( {\mathcal {H} } + \overline{ {\mathcal {H} } }) u \right\| }_{H^{\frac{N_0}{2}+2,\infty }} \\&\quad \times {\left\| \frac{\partial _{\alpha }}{{\zeta }_{\alpha }} \chi \right\| }_{W^{\frac{N_0}{2}+2,\infty }} {\Vert {\zeta }_{\alpha }-1 \Vert }_{X_{N_0}}. \end{aligned}$$

From (11.6), (11.28) and (11.36), and (11.4), we deduce

$$\begin{aligned} {\Vert \partial _{\alpha }( {\mathcal {H} } + \overline{ {\mathcal {H} } }) u \Vert }_{W^{\frac{N_0}{2}+1,\infty }}&\lesssim {\Vert L \Vert }^2_{W^{\frac{N_0}{2}+3,\infty }} \\ {\Vert ( {\mathcal {H} } + \overline{ {\mathcal {H} } }) u \Vert }_{X_{N_0}}&\lesssim {\Vert L \Vert }_{W^{\frac{N_0}{2}+1,\infty }} {\Vert L \Vert }_{X_{N_0}} . \end{aligned}$$

Using these bounds we eventually see that

$$\begin{aligned} {\left\| (8.118) \right\| }_{X_{N_0}}&\lesssim {\Vert L \Vert }_{X_{N_0}} {\Vert L \Vert }_{W^{\frac{N_0}{2}+3,\infty }}^2 . \end{aligned}$$

Estimate of (8.119). This term is of the form \(\mathbf{T}(L,L,L_{\alpha })\) and therefore (11.35) can be applied directly.

8.3 Proof of Proposition 8.3: control in terms of the \(Z^{\prime }\) norm

Recall the definitions

$$\begin{aligned} \widetilde{L} := ({\zeta }_{\alpha }- 1, u, \text{ Im }{\zeta }, w) , \quad L := ({\zeta }_{\alpha }- 1, u, w, \text{ Im }{\zeta }, \partial _{\alpha }\chi , v) \end{aligned}$$
(8.122)

and

$$\begin{aligned} L^- := ({\zeta }_{\alpha }- 1, u, w, \partial _{\alpha }\chi , v) . \end{aligned}$$
(8.123)

We want to show

$$\begin{aligned} {\Vert L (t) \Vert }_{W^{N_1,\infty }} \lesssim {\Vert (h(t),\phi (t)) \Vert }_{Z^{\prime }} \end{aligned}$$
(8.124)

and

$$\begin{aligned} {\left\| {\mathcal {H} } L^- (t) \right\| }_{W^{N_1,\infty }} \lesssim {\Vert (h(t),\phi (t)) \Vert }_{Z^{\prime }} \end{aligned}$$
(8.125)

where the \(Z^{\prime }\) norm is defined in (1.8). These estimates rely on the following Lemmas:

Lemma 8.9

Assume that the a priori estimate (2.9) holds. Then there exists constants \(c_i,d_i\), for \(i=1,\dots ,5\) such that

$$\begin{aligned} L_i^-(t,{\alpha }) - c_i \partial _{\alpha }\chi (t,{\alpha }) - d_i \partial _{\alpha }{\lambda }(t,{\alpha }) = Q (t,{\alpha }) \end{aligned}$$
(8.126)

where \(Q\) denotes a quadratic expression in \(\widetilde{L}\) satisfying

$$\begin{aligned} {\Vert Q(t) \Vert }_{H^{k}} \lesssim {\Vert \widetilde{L}(t) \Vert }_{H^{k+1}} {\Vert \widetilde{L}(t) \Vert }_{W^{\frac{k}{2}+1,\infty }} . \end{aligned}$$
(8.127)

Lemma 8.10

Assume that the a priori estimates (2.8) and (2.9) hold, and that

$$\begin{aligned} \sup _{[0,T]} {(1+t)}^\frac{1}{8} {\Vert \phi (t) \Vert }_{L^\infty } \le {\varepsilon }_1 , \end{aligned}$$
(8.128)

as guaranteed by (4.35). Then for any \(0 \le k \le N_1\) and any \(t \in [0,T]\) we have,

$$\begin{aligned} {\Vert \partial _{\alpha }\chi (t) \Vert }_{W^{k,\infty }}&\lesssim {\Vert (h(t),\phi (t)) \Vert }_{Z^{\prime }} \end{aligned}$$
(8.129)
$$\begin{aligned} {\Vert \partial _{\alpha }{\lambda }(t) \Vert }_{W^{k,\infty }}&\lesssim {\Vert (h(t),\phi (t)) \Vert }_{Z^{\prime }} \end{aligned}$$
(8.130)

and

$$\begin{aligned} {\Vert {\mathcal {H} } \partial _{\alpha }\chi (t) \Vert }_{W^{k,\infty }}&\lesssim {\Vert (h(t),\phi (t)) \Vert }_{Z^{\prime }} \end{aligned}$$
(8.131)
$$\begin{aligned} {\Vert {\mathcal {H} } \partial _{\alpha }{\lambda }(t) \Vert }_{W^{k,\infty }}&\lesssim {\Vert (h(t),\phi (t)) \Vert }_{Z^{\prime }} . \end{aligned}$$
(8.132)

Proof of Proposition 8.3

The estimate

$$\begin{aligned} {\Vert L^- \Vert }_{W^{N_1,\infty }} + {\Vert {\mathcal {H} } L^- \Vert }_{W^{N_1,\infty }} \lesssim {\Vert (h,\phi ) \Vert }_{Z^{\prime }} \end{aligned}$$

clearly follows by combining the above Lemmas. To obtain (8.124) we need to use in addition the identity

$$\begin{aligned} \text{ Im }{\zeta }(t,{\alpha })= h(t ,\text{ Re }{\zeta }(t,{\alpha })) \end{aligned}$$

to estimate \(\text{ Im }{\zeta }\) in \(W^{N_1,\infty }\):

$$\begin{aligned}&{\Vert \text{ Im }{\zeta }\Vert }_{W^{N_1,\infty }} \lesssim {\Vert h \Vert }_{W^{N_1,\infty }} {\Vert \text{ Re }{\zeta }_{\alpha }\Vert }_{W^{N_1,\infty }} \lesssim {\Vert h \Vert }_{W^{N_1,\infty }} \\&\quad \left( 1 + {\Vert \text{ Re }{\zeta }_{\alpha }- 1\Vert }_{H^{N_1+1}} \right) \lesssim {\Vert h \Vert }_{W^{N_1,\infty }} . \end{aligned}$$

\(\square \)

8.3.1 Proof of Lemma 8.9

To prove Lemma 8.9 we use the identities (2.44), (2.50), (3.38) and (2.35) derived by Wu [45], which relate the components of \(L^-\), \({\zeta }_{\alpha }-1,u,w\) and \(v\), to \(\partial _{\alpha }\chi \) and \(\partial _{\alpha }{\lambda }\). These are given respectively by

$$\begin{aligned} \overline{u}&= \partial _{\alpha }\overline{{\lambda }} + \overline{u} (1 - {\zeta }_{\alpha }) + \frac{1}{2} ({\zeta }_{\alpha }- \overline{{\zeta }}_{\alpha }) \overline{u}+ \frac{1}{2} \overline{{\zeta }}_{\alpha }\left( {\mathcal {H} } \frac{1}{{\zeta }_{\alpha }} + \overline{ {\mathcal {H} } } \frac{1}{\overline{{\zeta }}_{\alpha }} \right) (\overline{u} {\zeta }_{\alpha })\qquad \end{aligned}$$
(8.133)
$$\begin{aligned} \overline{w} {\zeta }_{\alpha }&= -\frac{i}{2} \partial _{\alpha }\overline{\chi } + \frac{1}{2} \overline{ {\mathcal {H} } } \left( \overline{u}_{\alpha }\frac{\overline{u}{\zeta }_{\alpha }}{\overline{{\zeta }_{\alpha }}} \right) - \frac{1}{2} [\overline{u}, \overline{ {\mathcal {H} } }] \frac{\partial _{\alpha }( \overline{u}{\zeta }_{\alpha })}{\overline{{\zeta }_{\alpha }}}\nonumber \\&\quad - \frac{1}{2\pi i} \int \frac{ (\overline{u}({\alpha }) - \overline{u}({\beta }) ) ({\zeta }_{\alpha }({\alpha }) - {\zeta }_{\beta }({\beta })) }{{ (\overline{{\zeta }}({\alpha }) - \overline{{\zeta }}({\beta })) }^2} \overline{u}({\beta }) {\zeta }_{\beta }({\beta }) \,d{\beta }\nonumber \\&\quad + \frac{1}{2} ({\zeta }_{\alpha }- \overline{{\zeta }}_{\alpha }) \overline{w}\nonumber - \frac{1}{2} u_{\alpha }\overline{u}+ \frac{1}{2} \overline{{\zeta }}_{\alpha }\left( {\mathcal {H} } \frac{1}{{\zeta }_{\alpha }} + \overline{ {\mathcal {H} } } \frac{1}{\overline{{\zeta }}_{\alpha }} \right) (\overline{w} {\zeta }_{\alpha }+ \overline{u} u_{\alpha })\nonumber \\&\quad -\frac{\overline{{\zeta }}_{\alpha }}{\pi } \int \text{ Im }{\left( \frac{u({\alpha }) - u({\beta })}{ \overline{{\zeta }}({\alpha }) - \overline{{\zeta }}({\beta }) } \right) }^2 \overline{u}({\beta }) {\zeta }_{\beta }({\beta }) \, d{\beta }\end{aligned}$$
(8.134)
$$\begin{aligned} {\zeta }_{\alpha }- 1&= \frac{w}{i A} - \frac{A-1}{A}\end{aligned}$$
(8.135)
$$\begin{aligned} v&= 2u - ( {\mathcal {H} } + \overline{ {\mathcal {H} } }) u - [u, {\mathcal {H} } ] \frac{{\zeta }_{\alpha }- \overline{{\zeta }}_{\alpha }}{{\zeta }_{\alpha }} . \end{aligned}$$
(8.136)

From (8.133) we can schematically write

$$\begin{aligned} u - \partial _{\alpha }{\lambda }= \widetilde{L} \cdot \widetilde{L} + \mathbf{Q}( \widetilde{L}, \widetilde{L} ) \end{aligned}$$

up to cubic and higher order terms whose arguments have the same regularity of \(\widetilde{L}\). Commuting derivatives via (11.39b) and using (11.48), it is then easy to verify that \(u - \partial _{\alpha }{\lambda }= Q\), where \(Q\) is a quadratic term satisfying the estimate (8.127) in the statement. Similarly, from (8.134) we deduce that up to cubic terms

$$\begin{aligned} w - \frac{i}{2} \partial _{\alpha }\chi = {\mathcal {H} } ( \widetilde{L} \cdot \widetilde{L}_{\alpha }) + Q_0 (\widetilde{L}, \widetilde{L}_{\alpha }) + \mathbf{Q}(\widetilde{L}, \widetilde{L}) . \end{aligned}$$

Arguing as above using the bounds (11.3), (11.30) and (11.48), it follows that \(w - \frac{i}{2} \partial _{\alpha }\chi = Q\), for some \(Q\) satisfying (8.127). Using (8.135), the last equality above, and the quadratic bounds on \(A-1\) given by (7.16), we can write

$$\begin{aligned} {\zeta }_{\alpha }- 1&= \frac{w}{i A} - \frac{A-1}{A} = - i w + i w \frac{A-1}{A} - \frac{A-1}{A} = \frac{1}{2} \partial _{\alpha }\chi + Q , \end{aligned}$$

so that (8.126) is verified also for the component \({\zeta }_{\alpha }- 1\). Combining (8.133) with the identity \(u - \partial _{\alpha }{\lambda }= Q\), and the quadratic bounds on \( {\mathcal {H} } + \overline{ {\mathcal {H} } }\) given in (11.6), we see that \(v - 2 \partial _{\alpha }{\lambda }= Q\), for \(Q\) as above. Thus we have checked that (8.126) holds true for all \(i=1,\dots ,5\). \(\square \)

8.3.2 Proof of Lemma 8.10

Let \( {\mathcal {H} } = {\mathcal {H} } _{\zeta }\) and let \(H_0\) be the flat Hilbert transform. We start by establishing the following estimate:

$$\begin{aligned}&{\left\| (I- {\mathcal {H} } ) (f \circ \text{ Re }{\zeta }) - \left[ (I-H_0) f(1+h^{\prime }) \right] \circ \text{ Re }{\zeta }\right\| }_{W^{k,\infty }}\nonumber \\&\quad \lesssim {\Vert f \Vert }_{W^{k+1,\infty }} {\Vert h \Vert }_{W^{k+2,\infty }} . \end{aligned}$$
(8.137)

for any \(0 \le k \le N_1 + 4\). Notice that since \(\text{ Im }{\zeta }(t,{\alpha }) = h (t, \text{ Re }{\zeta }(t,{\alpha }))\) we have \({\zeta }= ( x + i h(x) ) \circ \text{ Re }{\zeta }\). Thus by a change of variables one has

$$\begin{aligned} (I- {\mathcal {H} } ) (f \circ \text{ Re }{\zeta }) = (I- {\mathcal {H} } _{x+ih(x)}) f \circ \text{ Re }{\zeta }. \end{aligned}$$

Expanding out the denominator in the expression for \( {\mathcal {H} } _{x+ih(x)} f\) we see that

$$\begin{aligned}&(I- {\mathcal {H} } _{x+ih(t,x)}) f = (I-H_0) \left[ f(t,\cdot ) (1+ h^{\prime }(t,\cdot )) \right] (x) + R_f (t,x) \,\nonumber \\&\text{ with } \quad R_f (t,x) := \frac{1}{i\pi } \int H\left( \frac{h(x) - h(y)}{x-y} \right) \frac{h(x) - h(y)}{ {(x-y)}^2 } f (y)(1+ h^{\prime }(y)) \, dy \end{aligned}$$
(8.138)

for some smooth function \(H\). To prove (8.137) it then suffices to show that \(R_f\) in \(W^{k,\infty }\) is bounded by the right-hand side of (8.137). Applying the commutation identity (11.39b) in order to distribute derivatives, and the \(L^\infty \) estimate (11.40), it is not hard to see that

$$\begin{aligned} {\Vert R_f \Vert }_{W^{k,\infty }}&\lesssim {\Vert h \Vert }_{W^{k+2,\infty }} {\Vert f(1+h^{\prime }) \Vert }_{W^{k+1,\infty }}\lesssim {\Vert f \Vert }_{W^{k+1,\infty }} {\Vert h \Vert }_{W^{k+2,\infty }} . \end{aligned}$$

We will now show the estimates (8.130) and (8.132). The estimates (8.129) and (8.131) can be proven in a similar fashion, so we will not detail the proof here.

Proof of (8.130). Recall the definition of \({\lambda }\)

$$\begin{aligned} {\lambda }= (I - {\mathcal {H} } )\psi = (I - {\mathcal {H} } ) ( \phi \circ \text{ Re }{\zeta }) . \end{aligned}$$

Applying (8.137) with \(f = \phi \), and using the assumption (8.128), we see that

$$\begin{aligned}&{\left\| \partial _{\alpha }{\lambda }- \partial _{\alpha }\left[ (I-H_0) \phi (1+h^{\prime }) \circ \text{ Re }{\zeta }\right] \right\| }_{W^{k,\infty }}\nonumber \\&\quad \lesssim {\Vert \phi \Vert }_{W^{k+1,\infty }} {\Vert h \Vert }_{W^{k+2,\infty }} \lesssim {\Vert (h,\phi ) \Vert }_{Z^{\prime }} \end{aligned}$$
(8.139)

for any \(0 \le k \le N_1 + 2\). Moreover, using \({\Vert \partial _x H_0 f \Vert }_{L^\infty } \lesssim {\Vert \Lambda f \Vert }_{W^{1,\infty }}\), standard product estimates, and the hypothesis (8.128), one can estimate

$$\begin{aligned}&{\left\| \partial _{\alpha }\left[ (I-H_0) \phi (1+h^{\prime }) \right] \circ \text{ Re }{\zeta }\right\| }_{W^{k,\infty }}\nonumber \\&\quad \lesssim {\left\| \partial _x \left[ (I-H_0) \phi (1+h^{\prime }) \right] \right\| }_{W^{k,\infty }} \nonumber \\&\quad \lesssim {\Vert \partial _x (I-H_0) \phi \Vert }_{W^{k+1,\infty }} + {\Vert \phi h^{\prime }\Vert }_{W^{k+2,\infty }} \lesssim {\Vert (h,\phi ) \Vert }_{Z^{\prime }} . \end{aligned}$$
(8.140)

This and (8.139) give us for all \(0 \le k \le N_1 + 2\)

$$\begin{aligned} {\left\| \partial _{\alpha }{\lambda }\right\| }_{W^{k,\infty }}&\lesssim {\Vert (h,\phi ) \Vert }_{Z^{\prime }} , \end{aligned}$$
(8.141)

which in particular implies (8.130).

Proof of (8.132). Let us write

$$\begin{aligned} {\lambda }= F \circ \text{ Re }{\zeta }\qquad&\text{ with } \qquad F(t,x) = \left( I - {\mathcal {H} } _{x+ih(t,x)} \right) \phi (t,\cdot ) . \end{aligned}$$
(8.142)

Then \(\partial _{\alpha }{\lambda }= \partial _x F \circ \text{ Re }{\zeta }\, \partial _{\alpha }\text{ Re }{\zeta }\), and since we know that \({\zeta }_{\alpha }-1\) has uniformly bounded \(H^{N_1+5}\) norm, using also (8.141), we can easily deduce that

$$\begin{aligned} {\Vert \partial _x F\Vert }_{W^{k,\infty }} \lesssim {\Vert \partial _{\alpha }{\lambda }\Vert }_{W^{k,\infty }} \lesssim {\Vert (h,\phi ) \Vert }_{Z^{\prime }} , \end{aligned}$$
(8.143)

for any \(0 \le k \le N_1 + 2\). Using the definition of \( {\mathcal {H} } \), and making a change of variables, we see that

$$\begin{aligned} {\mathcal {H} } \frac{1}{{\zeta }_{\alpha }} \partial _{\alpha }{\lambda }(t,{\alpha })&= \frac{1}{i\pi } \int \frac{\partial _x F(t, \text{ Re }{\zeta }(t,{\beta }))}{{\zeta }({\alpha })-{\zeta }({\beta })} \partial _{\beta }\text{ Re }{\zeta }(t,{\beta }) \, d{\beta }\\&= \frac{1}{i\pi } \left. \int \frac{\partial _y F(t,y)}{ x + ih(x) -(y + ih(y)) } \, dy \right| _{x = \text{ Re }{\zeta }(t,{\alpha })} . \end{aligned}$$

Using also (8.141), it follows that

$$\begin{aligned} {\left\| {\mathcal {H} } \partial _{\alpha }{\lambda }(t,{\alpha }) \right\| }_{W^{N_1,\infty }}&\lesssim {\left\| {\mathcal {H} } \frac{1}{{\zeta }_{\alpha }} \partial _{\alpha }{\lambda }(t,{\alpha }) \right\| }_{W^{N_1,\infty }} + {\Vert \partial _{\alpha }{\lambda }\Vert }_{W^{N_1+1,\infty }} \\&\lesssim {\left\| \int \frac{\partial _y F(t,y)}{ x + ih(x) -(y + ih(y)) } \, dy \right\| }_{W^{N_1,\infty }} \!+\! {\Vert (h,\phi ) \Vert }_{Z^{\prime }}. \end{aligned}$$

Setting

$$\begin{aligned} G(t,x) = \int \frac{\partial _y F(t,y)}{ x + ih(x) -(y + ih(y)) } \, dy , \end{aligned}$$
(8.144)

we see that in order to obtain (8.132) it suffices to show

$$\begin{aligned} {\Vert G \Vert }_{W^{N_1,\infty }} \lesssim {\Vert (h,\phi ) \Vert }_{Z^{\prime }} . \end{aligned}$$
(8.145)

Expanding the denominator in (8.144), we can write

$$\begin{aligned}&G(t,x) = (I - H_0) \partial _x F (t,\cdot ) (x) + G_1 (t,x) \,\nonumber \\&\text{ with } \qquad G_1(t,x) := \frac{1}{i\pi } \int H\left( \frac{h(x) - h(y)}{x-y} \right) \frac{h(x) - h(y)}{ {(x-y)}^2 } \partial _y F (t,y) \, dy \end{aligned}$$
(8.146)

for some smooth function \(H\). Expanding the denominator in the expression for \(F\) we can write:

$$\begin{aligned}&F(t,x) = (I - H_0) [(1+h^{\prime }(t,\cdot ))\phi (t,\cdot )] (x) + R_\phi (t,x) \, \end{aligned}$$

where \(R_\phi \) is given by (8.138). It follows that

$$\begin{aligned} G&= (I - H_0) \partial _x (I-H_0) [\phi (1+h^{\prime })] + (I - H_0) \partial _x R_\phi + G_1 . \end{aligned}$$
(8.147)

To obtain (8.145) it is then enough to have

$$\begin{aligned} {\Vert H_0 \partial _x [ \phi (1+h^{\prime }) ] \Vert }_{W^{N_1,\infty }}&\lesssim {\Vert (h,\phi ) \Vert }_{Z^{\prime }} , \end{aligned}$$
(8.148)
$$\begin{aligned} {\Vert (I-H_0) \partial _x R_\phi \Vert }_{W^{N_1,\infty }}&\lesssim {\Vert (h,\phi ) \Vert }_{Z^{\prime }} , \end{aligned}$$
(8.149)
$$\begin{aligned} {\Vert G_1 \Vert }_{W^{N_1,\infty }}&\lesssim {\Vert (h,\phi ) \Vert }_{Z^{\prime }} . \end{aligned}$$
(8.150)

The bound (8.148) has been already shown to hold true in the above paragraph, see (8.140). We also have

$$\begin{aligned} {\Vert R_\phi \Vert }_{W^{N_1+2,\infty }} \lesssim {\Vert \phi \Vert }_{W^{N_1+3,\infty }} {\Vert h \Vert }_{W^{N_1+4,\infty }} \lesssim {\Vert (h,\phi ) \Vert }_{Z^{\prime }} \end{aligned}$$

which is stronger than (8.149). Using again commutation identities and (11.40), together with (8.143), we get

$$\begin{aligned} {\Vert G_1 \Vert }_{W^{N_1,\infty }} \lesssim {\Vert h \Vert }_{W^{N_1+2,\infty }} {\Vert \partial _x F \Vert }_{W^{N_1+1,\infty }} \lesssim {\Vert h \Vert }_{W^{N_1+2,\infty }} {\Vert (h,\phi ) \Vert }_{Z^{\prime }} \end{aligned}$$

which is enough for (8.150). \(\square \)

9 Proof of Proposition 2.2: the diffeomorphism \(k\)

This section contains the proof of Proposition 2.2. The main issue is to show that the change of coordinates \(k\) is a uniformly controlled diffeomorphism for all times. This is a substantial improvement of the analogous analysis performed by Wu [45, p. 124-127], and relies on a special null structure present in the transport equation (9.5) for \(k\).

We start by assuming a priori that

$$\begin{aligned} \sup _{t\in [0,T]} {\Vert k_{\alpha }(t) - 1 \Vert }_{W^{N_0/2+3,\infty }} \le 1/2 . \end{aligned}$$
(9.1)

Furthermore we assume, see (2.8),

$$\begin{aligned} \sup _{t\in [0,T]} \left[ (1+t)^{-p_0} {\Vert (h(t), \partial _x \phi (t)) \Vert }_{X_{N_0}} + \sqrt{1+t} {\Vert (h(t), \phi (t)) \Vert }_{Z^{\prime }}\right] \le {\varepsilon }_1 , \end{aligned}$$
(9.2)

and (2.9), that is

$$\begin{aligned}&\sup _{t \in [0,T]} \left[ (1+t)^{-p_0} {\Vert \widetilde{L}(t) \Vert }_{X_{N_0}} + {\Vert \widetilde{L}(t) \Vert }_{H^{N_1+5}} + \sqrt{1+t} {\Vert \widetilde{L}(t) \Vert }_{W^{N_1,\infty }} \right] \le {\varepsilon }_1 . \end{aligned}$$
(9.3)

We then aim to conclude

$$\begin{aligned} \sup _{[0,T]} {\Vert k_{\alpha }(t) - 1 \Vert }_{W^{N_0/2+3,\infty }} \lesssim {\varepsilon }_0 + {\varepsilon }_1^2 \end{aligned}$$
(9.4)

as a consequence of the following Lemmas:

Lemma 9.1

(Approximation of \(k_t\)) Let \(k\) be defined as in (7.2) then the following formula holds:

$$\begin{aligned} (I- {\mathcal {H} } _z) k_t = - [z_t, {\mathcal {H} } _z] \frac{\overline{z}_{\alpha }- k_{\alpha }}{z_{\alpha }} . \end{aligned}$$
(9.5)

Under the assumptions (9.2) and (9.3) there exists \({\gamma }>0\), such that for any \(t\in [0,T]\)

$$\begin{aligned} {\left\| \partial _{\alpha }[u, {\mathcal {H} } _{\zeta }] \frac{\overline{{\zeta }}_{\alpha }- 1}{{\zeta }_{\alpha }} - T_0 (h,\phi ) \circ \text{ Re }{\zeta }\right\| }_{W^{N_0/2+3,\infty }}&\lesssim {\varepsilon }_1^2 {(1+t)}^{-1-{\gamma }} \end{aligned}$$
(9.6)

and

$$\begin{aligned} {\left\| \partial _{\alpha }[u, {\mathcal {H} } _{\zeta }] \frac{\overline{{\zeta }}_{\alpha }- 1}{{\zeta }_{\alpha }} - T_0 (h,\phi ) \circ \text{ Re }{\zeta }\right\| }_{H^{N_0/2+4}}&\lesssim {\varepsilon }_1^2 {(1+t)}^{-1/2-{\gamma }} , \end{aligned}$$
(9.7)

with \(T_0\) given by

$$\begin{aligned} T_0 (f,g) := \partial _x [(I-H_0) g_x, H_0] (I-H_0) f_x , \end{aligned}$$
(9.8)

where \(H_0\) is the flat Hilbert transform, \(H_0 = {\mathcal {H} } _\mathrm {id}\) according to (1.4).

Lemma 9.2

(Estimate for \(T_0\)) Under the a priori assumptions (9.2), there exists \({\gamma }>0\) such that

$$\begin{aligned} {\left\| T_0 (h,\phi ) \right\| }_{W^{N_0/2+3,\infty }}&\lesssim {\varepsilon }_1^2 {(1+t)}^{-1-{\gamma }} , \end{aligned}$$
(9.9)
$$\begin{aligned} {\left\| T_0 (h,\phi ) \right\| }_{H^{N_0/2+4}}&\lesssim {\varepsilon }_1^2 {(1+t)}^{-1/2-{\gamma }} . \end{aligned}$$
(9.10)

The proofs of Lemma 9.1 and 9.2 are in Sect. 9.1 and 9.2 respectively. We now show how Proposition 2.2 follows from them.

Proof of Proposition 2.2

Since we know by our a priori assumption that \(k\) is diffeomorphism, we can define

$$\begin{aligned} K(t,{\alpha }) := [z_t, {\mathcal {H} } _z] \frac{\overline{z}_{\alpha }- k_{\alpha }}{z_{\alpha }} \circ k^{-1} . \end{aligned}$$
(9.11)

From the properties of the Hilbert transform and the definition of \({\zeta }\) and \(u\) in (7.3), we see that

$$\begin{aligned} K(t,{\alpha }) = [u, {\mathcal {H} } _{\zeta }] \frac{\overline{{\zeta }}_{\alpha }- 1}{{\zeta }_{\alpha }} \, \end{aligned}$$
(9.12)

and

$$\begin{aligned} (I- {\mathcal {H} } _{\zeta }) (k_t \circ k^{-1}) = - K(t,{\alpha }) . \end{aligned}$$
(9.13)

Applying the estimate (11.11) for the inversion of \(I- {\mathcal {H} } \), with \(f = k_t \circ k^{-1}\) and \(g = - K\), we see that

$$\begin{aligned}&{\Vert \partial _{\alpha }(k_t \circ k^{-1}) \Vert }_{W^{N_0/2+3,\infty }} \lesssim {\Vert \partial _{\alpha }K(t,{\alpha }) \Vert }_{W^{N_0/2+3,\infty }} + {\Vert {\zeta }_{\alpha }- 1 \Vert }_{W^{N_0/2+4,\infty }}\\&\quad {\Vert \partial _{\alpha }K(t,{\alpha }) \Vert }_{H^{N_0/2+4}} . \end{aligned}$$

From the a priori assumption (9.3) and (9.1) it follows that

$$\begin{aligned} {\Vert \partial _{\alpha }k_t \Vert }_{W^{N_0/2+3,\infty }} \!\lesssim \! {\Vert \partial _{\alpha }K(t,{\alpha }) \Vert }_{W^{N_0/2+3,\infty }} \!+\! {\varepsilon }_1 {(1\!+\!t)}^{-1/2} {\Vert \partial _{\alpha }K(t,{\alpha }) \Vert }_{H^{N_0/2+4}} . \end{aligned}$$
(9.14)

Applying successively (9.6) and (9.9) we see that

$$\begin{aligned} {\Vert \partial _{\alpha }K(t,{\alpha }) \Vert }_{W^{N_0/2+3,\infty }} \lesssim {\varepsilon }_1^2 {(1+t)}^{-1-{\gamma }} + {\Vert T_0 (h,\phi ) \Vert }_{W^{N_0/2+3,\infty }} \lesssim {\varepsilon }_1^2 {(1+t)}^{-1-{\gamma }} . \end{aligned}$$
(9.15)

Similarly, from (9.7) and (9.10) we have

$$\begin{aligned} {\Vert \partial _{\alpha }K(t,{\alpha }) \Vert }_{H^{N_0/2+4}} \lesssim {\varepsilon }_1^2 {(1+t)}^{-1/2-{\gamma }} + {\Vert T_0 (h,\phi ) \Vert }_{H^{N_0/2+4}} \lesssim {\varepsilon }_1^2 {(1+t)}^{-1/2-{\gamma }} . \end{aligned}$$
(9.16)

Plugging (9.15) and (9.16) into (9.14) gives

$$\begin{aligned} {\Vert \partial _{\alpha }k_t (t) \Vert }_{W^{N_0/2+3,\infty }} \lesssim {\varepsilon }_1^2 {(1+t)}^{-1-{\gamma }} \end{aligned}$$

whence

$$\begin{aligned} {\Vert k_{\alpha }(t) \!-\! 1 \Vert }_{W^{N_0/2+3,\infty }}&\lesssim {\Vert k_{\alpha }(0) \!-\! 1 \Vert }_{W^{N_0/2+3,\infty }}\\&\quad + \int _0^t {\Vert \partial _s k_{\alpha }(s) \Vert }_{W^{N_0/2+3,\infty }} \, ds \!\lesssim \! {\varepsilon }_0 \!+\! C_{\gamma }{\varepsilon }_1^2 . \end{aligned}$$

\(\square \)

9.1 Proof of Lemma 9.1

The identity (9.5) is proven by Wu in Proposition 2.4 of [45]. Let \(K\) be given by (9.11):

$$\begin{aligned} K(t,{\alpha }) = [u, {\mathcal {H} } _{\zeta }] \frac{\overline{{\zeta }}_{\alpha }- 1}{{\zeta }_{\alpha }} = Q_0 (u, \overline{{\zeta }}_{\alpha }- 1) , \end{aligned}$$
(9.17)

where \(Q_0\) is the bilinear operator defined in (11.18). We aim to approximate \(\partial _{\alpha }K\) by \(T_0(h,\phi )\) showing

$$\begin{aligned}&{\left\| \partial _{\alpha }K - T_0 (h,\phi ) \circ \text{ Re }{\zeta }\right\| }_{W^{N_0/2+3,\infty }} \lesssim {\varepsilon }_1^2 {(1+t)}^{-1-{\gamma }} \end{aligned}$$
(9.18)
$$\begin{aligned}&\quad {\left\| \partial _{\alpha }K - T_0 (h,\phi ) \circ \text{ Re }{\zeta }\right\| }_{H^{N_0/2+4}} \lesssim {\varepsilon }_1^2 {(1+t)}^{-1/2-{\gamma }} . \end{aligned}$$
(9.19)

Step 1: Approximation of \(u\). Let \(H_0\) denote the flat Hilbert transform, \(H_0 = {\mathcal {H} } _\mathrm {id}\) according to (1.4). We start by showing

$$\begin{aligned} {\left\| u - (I-H_0) \phi _x \circ \text{ Re }{\zeta }\right\| }_{W^{N_0/2+5,\infty }}&\lesssim {\varepsilon }_1^2 {(1+t)}^{-1/2-{\gamma }} . \end{aligned}$$
(9.20)

Using the identity (8.133) as in the proof of Lemma 8.9, we can schematically write

$$\begin{aligned} u - \partial _{\alpha }{\lambda }= \widetilde{L} \cdot \widetilde{L} + \mathbf{Q}( \widetilde{L}, \widetilde{L} ) , \end{aligned}$$

so that using (11.41) to estimate \(\mathbf{Q}\), and interpolating between the a priori decay assumption and the Sobolev bounds, we get

$$\begin{aligned} {\Vert u - \partial _{\alpha }{\lambda }\Vert }_{W^{N_0/2+5,\infty }} \lesssim {\Vert \widetilde{L} \Vert }^2_{W^{N_0/2+7,\infty }} \lesssim {\varepsilon }_1^2 {(1+t)}^{-3/4} . \end{aligned}$$

To obtain (9.20) it then suffices to show

$$\begin{aligned} {\left\| \partial _{\alpha }{\lambda }- (I-H_0) \phi _x \circ \text{ Re }{\zeta }\right\| }_{W^{N_0/2+5,\infty }}&\lesssim {\varepsilon }_1^2 {(1+t)}^{-1/2-{\gamma }} . \end{aligned}$$
(9.21)

Looking at (8.139) and (8.140) in the proof of Lemma 8.10 one can see that

$$\begin{aligned}&{\left\| \partial _{\alpha }{\lambda }- \partial _{\alpha }\left[ (I-H_0) \phi \circ \text{ Re }{\zeta }\right] \right\| }_{W^{N_0/2+5,\infty }} \lesssim {\left\| \phi \right\| }_{W^{N_0/2+7,\infty }} {\left\| h \right\| }_{W^{N_0/2+8,\infty }} \\&\quad \lesssim {\varepsilon }_1^2 {(1+t)}^{-5/8} , \end{aligned}$$

having used (4.35) in the last inequality. Since we also have

$$\begin{aligned}&{\left\| \partial _{\alpha }\left[ (I-H_0) \phi \circ \text{ Re }{\zeta }\right] - (I-H_0) \phi _x \circ \text{ Re }{\zeta }\right\| }_{W^{N_0/2+5,\infty }} \\&\quad = {\left\| (I-H_0) \phi _x \circ \text{ Re }{\zeta }\, (\text{ Re }{\zeta }_{\alpha }- 1) \right\| }_{W^{N_0/2+5,\infty }}\\&\quad \lesssim {\left\| \Lambda \phi \right\| }_{W^{N_0/2+6,\infty }} {\Vert {\zeta }_{\alpha }- 1 \Vert }_{W^{N_0/2+5,\infty }} \lesssim {\varepsilon }_1^2 {(1+t)}^{-1} , \end{aligned}$$

we have verified (9.21), hence (9.20) with \({\beta }= 1/8\).

Step 2: Approximation of \({\zeta }_{\alpha }- 1\). We want to show that \({\zeta }_{\alpha }-1\) can be approximated as follows:

$$\begin{aligned} {\left\| {\zeta }_{\alpha }- 1 - i(I-H_0) h_x \circ \text{ Re }{\zeta }\, \partial _{\alpha }\text{ Re }{\zeta }\right\| }_{W^{N_0/2+5,\infty }}&\lesssim {\varepsilon }_1^2 {(1+t)}^{-1/2-{\gamma }} \end{aligned}$$
(9.22)
$$\begin{aligned} {\left\| {\mathcal {H} } _{\zeta }\left( {\zeta }_{\alpha }- 1 - i(I-H_0) h_x \circ \text{ Re }{\zeta }\, \partial _{\alpha }\text{ Re }{\zeta }\right) \right\| }_{W^{N_0/2+3,\infty }}&\lesssim {\varepsilon }_1^2 {(1+t)}^{-1/2-{\gamma }} , \end{aligned}$$
(9.23)

for some \({\gamma }>0\). Putting together the identities (8.135) and (8.134), as in the proof of Lemma 8.9, we can write

$$\begin{aligned} {\zeta }_{\alpha }- 1 - \frac{1}{2} \partial _{\alpha }\chi = {\mathcal {H} } (\widetilde{L} \cdot \widetilde{L}) + \widetilde{L} \cdot {\mathcal {H} } \widetilde{L}_{\alpha }+ \widetilde{L} \cdot \widetilde{L}_{\alpha }+ \mathbf{Q}( \widetilde{L}, \widetilde{L} ) + A-1 + \cdots \, \end{aligned}$$
(9.24)

where operators of the type \(\mathbf{Q}\) are defined by (11.19)–(11.21), \(A\) is defined in (7.12), and “\(\cdots \)” denotes cubic or higher order terms which are more easily estimated, and we will therefore disregard. Notice that interpolating between the bounds provided by the a priori assumptions (9.3), for large enough \(p\) one has

$$\begin{aligned} {\Vert \widetilde{L}(t) \Vert }_{W^{N_0/2+7,p}} \lesssim {\varepsilon }_1 {(1+t)}^{-2/5} . \end{aligned}$$
(9.25)

Combining this with the estimates (11.37) and (11.36) for \( {\mathcal {H} } \) in \(W^{k,\infty }\), we see that

$$\begin{aligned}&{\left\| {\mathcal {H} } (\widetilde{L} \cdot \widetilde{L}) + \widetilde{L} \cdot {\mathcal {H} } \widetilde{L}_{\alpha }+ \widetilde{L} \cdot \widetilde{L}_{\alpha }\right\| }_{W^{N_0/2+5,\infty }} \nonumber \\&\quad \lesssim \left( {\Vert \widetilde{L} \Vert }_{W^{N_0/2+6,p}} + {\Vert \widetilde{L} \Vert }_{W^{N_0/2+6,\infty }} \right) {\Vert \widetilde{L} \Vert }_{W^{N_0/2+6,\infty }} \lesssim {\varepsilon }_1^2 {(1+t)}^{-4/5} \end{aligned}$$
(9.26)

and

$$\begin{aligned}&{\left\| {\mathcal {H} } \left( {\mathcal {H} } (\widetilde{L} \cdot \widetilde{L}) + \widetilde{L} \cdot {\mathcal {H} } \widetilde{L}_{\alpha }+ \widetilde{L} \cdot \widetilde{L}_{\alpha }\right) \right\| }_{W^{N_0/2+3,\infty }} \nonumber \\&\quad \lesssim \left( {\Vert \widetilde{L} \Vert }_{W^{N_0/2+6,p}} + {\Vert \widetilde{L} \Vert }_{W^{N_0/2+5,\infty }} \right) {\Vert \widetilde{L} \Vert }_{W^{N_0/2+6,\infty }} \lesssim {\varepsilon }_1^2 {(1+t)}^{-4/5} \end{aligned}$$
(9.27)

having chosen \(p\) large enough so that (9.25) holds.

We now want to obtain similar bounds for \(\mathbf{Q}( \widetilde{L}, \widetilde{L} )\) and \(A-1\), and more precisely show

$$\begin{aligned} {\big \Vert \mathbf{Q}( \widetilde{L}, \widetilde{L} ) \big \Vert }_{W^{N_0/2+5,\infty }}&\lesssim {\varepsilon }_1^2 {(1+t)}^{-4/5} \end{aligned}$$
(9.28)
$$\begin{aligned} {\big \Vert {\mathcal {H} } \mathbf{Q}( \widetilde{L}, \widetilde{L} ) \big \Vert }_{W^{N_0/2+3,\infty }}&\lesssim {\varepsilon }_1^2 {(1+t)}^{-3/5} \end{aligned}$$
(9.29)

and

$$\begin{aligned} {\left\| A-1 \right\| }_{W^{N_0/2+5,\infty }}&\lesssim {\varepsilon }_1^2 {(1+t)}^{-4/5} \end{aligned}$$
(9.30)
$$\begin{aligned} {\left\| {\mathcal {H} } (A-1) \right\| }_{W^{N_0/2+3,\infty }}&\lesssim {\varepsilon }_1^2 {(1+t)}^{-3/5} . \end{aligned}$$
(9.31)

The first bound (9.28) follows directly from (11.29) and (9.25). To obtain (9.29) first notice that the inequality in (11.48), which is an application of the \(L^2\) estimates in Corollary 11.4, gives

$$\begin{aligned} {\big \Vert \mathbf{Q}( \widetilde{L}, \widetilde{L} ) \big \Vert }_{H^{N_0/2+5}}&\lesssim {\Vert \widetilde{L} \Vert }_{H^{N_0/2+6}} {\Vert \widetilde{L} \Vert }_{W^{N_0/2+5,\infty }} \lesssim {\varepsilon }_1^2 {(1+t)}^{-2/5} . \end{aligned}$$

Interpolating this and the \(L^\infty \) bound (9.28) gives

$$\begin{aligned} {\big \Vert \mathbf{Q}( \widetilde{L}, \widetilde{L} ) \big \Vert }_{W^{N_0/2+3,p}}&\lesssim {\varepsilon }_1^2 {(1+t)}^{-3/5} , \end{aligned}$$

provided \(p\) is large enough. (9.29) then follows by applying (11.37).

Both (9.30) and (9.31) rely on the identity (7.16), which we can schematically write as

$$\begin{aligned} (I - {\mathcal {H} } )(A-1) = [\widetilde{L}, {\mathcal {H} } ] \widetilde{L}_{\alpha }+ [\widetilde{L}, {\mathcal {H} } ] \widetilde{L} + \cdots \end{aligned}$$
(9.32)

where once again “\(\cdots \)”stands for cubic order terms which we are going to disregard. Applying (11.10) we get

$$\begin{aligned} {\Vert A-1 \Vert }_{W^{N_0/2+5,\infty }}&\lesssim {\big \Vert [\widetilde{L}, {\mathcal {H} } ] \widetilde{L}_{\alpha }+ [\widetilde{L}, {\mathcal {H} } ] \widetilde{L} \big \Vert }_{W^{N_0/2+5,\infty }}\\&\quad + {\varepsilon }_1 {(1+t)}^{-2/5} {\big \Vert [\widetilde{L}, {\mathcal {H} } ] \widetilde{L}_{\alpha }+ [\widetilde{L}, {\mathcal {H} } ] \widetilde{L} \big \Vert }_{H^{N_0/2+6}} . \end{aligned}$$

We can then bound the above right-hand side by using (11.37), (9.25), and the boundedness of \( {\mathcal {H} } \) in (11.3):

$$\begin{aligned} {\Vert A-1 \Vert }_{W^{N_0/2+5,\infty }}&\lesssim {\big ( {\Vert \widetilde{L} \Vert }_{W^{N_0/2+7,p}} + {\Vert \widetilde{L} \Vert }_{W^{N_0/2+7,\infty }} \big )}^2\nonumber \\&\quad + {\varepsilon }_1^2 {(1+t)}^{-1+p_0} \lesssim {\varepsilon }_1^2 {(1+t)}^{-4/5} . \end{aligned}$$

We have therefore obtained (9.30). Since (9.32) holds true also for \( {\mathcal {H} } (A-1)\), up to a sign, the estimate (9.31) follows as above.

Putting together (9.24)–(9.31) we have

$$\begin{aligned} {\left\| {\zeta }_{\alpha }- 1 - \frac{1}{2} \partial _{\alpha }\chi \right\| }_{W^{N_0/2+5,\infty }}&\lesssim {\varepsilon }_1^2 {(1+t)}^{-1/2-{\gamma }} \\ {\left\| {\mathcal {H} } _{\zeta }\left( {\zeta }_{\alpha }- 1 - \frac{1}{2} \partial _{\alpha }\chi \right) \right\| }_{W^{N_0/2+3,\infty }}&\lesssim {\varepsilon }_1^2 {(1+t)}^{-1/2-{\gamma }} , \end{aligned}$$

for some \({\gamma }>0\). To obtain (9.22) and (9.23) it is then enough to show

$$\begin{aligned} {\left\| \partial _{\alpha }\chi - 2 i (I-H_0) h_x \circ \text{ Re }{\zeta }\, \partial _{\alpha }\text{ Re }{\zeta }\right\| }_{W^{N_0/2+5,\infty }}&\lesssim {\varepsilon }_1^2 {(1+t)}^{-1/2-{\gamma }} \\ {\left\| {\mathcal {H} } _{\zeta }\left( \partial _{\alpha }\chi - 2 i (I-H_0) h_x \circ \text{ Re }{\zeta }\, \partial _{\alpha }\text{ Re }{\zeta }\right) \right\| }_{W^{N_0/2+3,\infty }}&\lesssim {\varepsilon }_1^2 {(1+t)}^{-1/2-{\gamma }} . \end{aligned}$$

In light of (11.36) both bounds would follow from

$$\begin{aligned}&{\left\| \chi - 2 i (I-H_0) h \circ \text{ Re }{\zeta }\right\| }_{W^{N_0/2+6,\infty }} \lesssim {\varepsilon }_1^2 {(1+t)}^{-1/2-{\gamma }} . \end{aligned}$$
(9.33)

From the definition of \(\chi \) in (7.4) and (7.7) we see that

$$\begin{aligned} \chi = 2i(I- {\mathcal {H} } ) (h \circ \text{ Re }{\zeta }) . \end{aligned}$$

Applying the inequality (8.137) with \(f= 2i h\), one gets

$$\begin{aligned}&{\left\| \chi - 2i (I-H_0) h \circ \text{ Re }{\zeta }\right\| }_{W^{N_0/2+6,\infty }}\\&\quad \lesssim {\Vert h \Vert }_{W^{N_0/2+7,\infty }} {\Vert h \Vert }_{W^{N_0/2+8,\infty }} + {\Vert (I-H_0) (h h^{\prime }) \Vert }_{W^{N_0/2+8,\infty }} \\&\quad \lesssim {\Vert h \Vert }^2_{W^{N_0/2+9,\infty }} + {\Vert \partial _x H_0 h^2 \Vert }_{W^{N_0/2+8,\infty }}\\&\quad \lesssim {\Vert h \Vert }^2_{W^{N_0/2+10,\infty }} \lesssim {\varepsilon }_1^2 {(1+t)}^{-4/5} . \end{aligned}$$

This gives us (9.33) and concludes the proof of (9.22)–(9.23).

Step 3: First approximation of \(\partial _{\alpha }K\). We now want to show

$$\begin{aligned}&{\left\| \partial _{\alpha }K - \partial _{\alpha }Q_0 \left( (I-H_0)\phi _x \circ \text{ Re }{\zeta }, (I-H_0) h_x \circ \text{ Re }{\zeta }\, \text{ Re }{\zeta }_{\alpha }\right) \right\| }_{W^{N_0/2+3,\infty }} \nonumber \\&\quad \lesssim {\varepsilon }_1^2 {(1+t)}^{-1-{\gamma }} \end{aligned}$$
(9.34)
$$\begin{aligned}&{\left\| \partial _{\alpha }K - \partial _{\alpha }Q_0 \left( (I-H_0)\phi _x \circ \text{ Re }{\zeta }, (I-H_0) h_x \circ \text{ Re }{\zeta }\, \text{ Re }{\zeta }_{\alpha }\right) \right\| }_{H^{N_0/2+4}}\nonumber \\&\quad \lesssim {\varepsilon }_1^2 {(1+t)}^{-1/2-{\gamma }} . \end{aligned}$$
(9.35)

Let us denote

$$\begin{aligned} A_1&:= u - (I-H_0)\phi _x \circ \text{ Re }{\zeta }\\ A_2&:= {\zeta }_{\alpha }- 1 - i(I-H_0) h_x \circ \text{ Re }{\zeta }\, \partial _{\alpha }\text{ Re }{\zeta }. \end{aligned}$$

Using (9.17) we can write

$$\begin{aligned} \partial _{\alpha }K - \partial _{\alpha }Q_0 \left( (I-H_0)\phi _x \circ \text{ Re }{\zeta }, (I-H_0) h_x \circ \text{ Re }{\zeta }\, \partial _{\alpha }\text{ Re }{\zeta }\right) = K_1 + K_2 \end{aligned}$$

where

$$\begin{aligned} K_1&:= \partial _{\alpha }Q_0 \left( A_1, {\zeta }_{\alpha }-1 \right) \\ K_2&:= \partial _{\alpha }Q_0 \left( (I-H_0)\phi _x \circ \text{ Re }{\zeta }, A_2 \right) . \end{aligned}$$

From (9.20), (9.22) and (9.23) we know that there exists \({\gamma }>0\) such that

$$\begin{aligned} {\left\| A_1 \right\| }_{W^{N_0/2+5,\infty }}&\lesssim {\varepsilon }_1^2 {(1+t)}^{-1/2-{\gamma }} , \end{aligned}$$
(9.36)
$$\begin{aligned} {\left\| A_2 \right\| }_{W^{N_0/2+4,\infty }}&\lesssim {\varepsilon }_1^2 {(1+t)}^{-1/2-{\gamma }} , \end{aligned}$$
(9.37)
$$\begin{aligned} {\left\| {\mathcal {H} } A_2 \right\| }_{W^{N_0/2+3,\infty }}&\lesssim {\varepsilon }_1^2 {(1+t)}^{-1/2-{\gamma }} . \end{aligned}$$
(9.38)

Using the \(L^\infty \) type bound (11.32) for operators of the type \(\partial _{\alpha }Q_0\), (9.36) above, and the bound in Proposition 8.3 together with the a priori decay assumption in (9.2), we see that

$$\begin{aligned} {\left\| K_1 \right\| }_{W^{N_0/2+3,\infty }}&\!\lesssim \! {\left\| A_1 \right\| }_{W^{N_0/2+5,\infty }} \left( {\Vert {\mathcal {H} } ({\zeta }_{\alpha }\!-\! 1) \Vert }_{W^{N_0/2+3,\infty }} \!+\! {\Vert {\zeta }_{\alpha }\!-\! 1 \Vert }_{W^{N_0/2+4,\infty }} \right) \\&\lesssim {\varepsilon }_1^3 {(1+t)}^{-1-{\gamma }} . \end{aligned}$$

Similarly, in view of (9.37) and (9.38) above, one has

$$\begin{aligned} {\left\| K_2 \right\| }_{W^{N_0/2+3,\infty }}&\lesssim {\left\| (I-H_0)\phi _x \circ \text{ Re }{\zeta }\right\| }_{W^{N_0/2+5,\infty }} \\&\quad \times \left( {\Vert {\mathcal {H} } A_2 \Vert }_{W^{N_0/2+3,\infty }} + {\Vert A_2 \Vert }_{W^{N_0/2+4,\infty }} \right) \lesssim {\varepsilon }_1^3 {(1+t)}^{-1-{\gamma }} . \end{aligned}$$

We have therefore obtained (9.34). The \(H^{N_0/2+4}\) estimate (9.35) can be obtained similarly, estimating \(K_1\) and \(K_2\) in \(H^{N_0/2+4}\), by using the bounds on \(A_1\) and \(A_2\) given by (9.20), (9.22), (9.36), (9.37) and (9.38), and boundedness properties of the Cauchy integral on Sobolev spaces, see (11.3).

Step 4: Approximation by \(T_0\). Let us denote

$$\begin{aligned} \phi _0&:= \partial _x (I-H_0) \phi \end{aligned}$$
(9.39)
$$\begin{aligned} h_0&:= \partial _x (I-H_0) h. \end{aligned}$$
(9.40)

To eventually obtain (9.18)–(9.19) it suffices to combine (9.34)–(9.35) with

$$\begin{aligned} {\left\| \partial _{\alpha }Q_0 \left( \phi _0 \circ \text{ Re }{\zeta }, h_0 \circ \text{ Re }{\zeta }\, \text{ Re }{\zeta }_{\alpha }\right) - T_0 (h,\phi ) \circ \text{ Re }{\zeta }\right\| }_{W^{N_0/2+3,\infty }}&\lesssim {\varepsilon }_1^2 {(1+t)}^{-1-{\gamma }} \end{aligned}$$
(9.41)
$$\begin{aligned} {\left\| \partial _{\alpha }Q_0 \left( \phi _0 \circ \text{ Re }{\zeta }, h_0 \circ \text{ Re }{\zeta }\, \text{ Re }{\zeta }_{\alpha }\right) - T_0 (h,\phi ) \circ \text{ Re }{\zeta }\right\| }_{H^{N_0/2+4}}&\lesssim {\varepsilon }_1^2 {(1+t)}^{-1/2-{\gamma }} . \end{aligned}$$
(9.42)

With the notation (9.39)–(9.40) we can write \(T_0(h,\phi )\) in (9.8) as

$$\begin{aligned} T_0(h,\phi ) \circ \text{ Re }{\zeta }({\alpha })&= \partial _x [ \phi _0, H_0 ] h_0 \circ \text{ Re }{\zeta }({\alpha })\nonumber \\&= \frac{1}{i\pi } \partial _x \int \frac{ \phi _0 (x) - \phi _0 (y) }{ x - y} h_0(y) \, dy \circ \text{ Re }{\zeta }({\alpha }) , \end{aligned}$$
(9.43)

whereas writing explicitly \(Q_0\) and changing variables we can write

$$\begin{aligned}&\partial _{\alpha }Q_0 \left( \phi _0 \circ \text{ Re }{\zeta }, h_0 \circ \text{ Re }{\zeta }\, \text{ Re }{\zeta }_{\alpha }\right) \nonumber \\&\quad = \frac{1}{i\pi } \partial _{\alpha }\int \frac{ \phi _0 \circ \text{ Re }{\zeta }({\alpha }) - \phi _0 \circ \text{ Re }{\zeta }({\beta }) }{{\zeta }({\alpha }) - {\zeta }({\beta })}h_0 \circ \text{ Re }{\zeta }({\beta }) \, \text{ Re }{\zeta }_{\beta }({\beta }) \, d{\beta }\nonumber \\&\quad = \frac{1}{i\pi } \partial _{\alpha }\left( \int \frac{ \phi _0 (x) - \phi _0 (y) }{ x+ih(x) - (y+ih(y))} h_0(y) \, dy \circ \text{ Re }{\zeta }({\alpha }) \right) . \end{aligned}$$
(9.44)

Then the difference we are interested in is given by

$$\begin{aligned}&\partial _{\alpha }Q_0 \left( \phi _0 \circ \text{ Re }{\zeta }, h_0 \circ \text{ Re }{\zeta }\, \text{ Re }{\zeta }_{\alpha }\right) - T_0(h,\phi ) \circ \text{ Re }{\zeta }({\alpha })\\&\quad = T_1 \circ \text{ Re }{\zeta }\, (\text{ Re }{\zeta }_{\alpha }-1) + T_2 \circ \text{ Re }{\zeta }\, \text{ Re }{\zeta }_{\alpha }, \end{aligned}$$

where

$$\begin{aligned} T_1&:= \frac{1}{i\pi } \partial _x \int \frac{ \phi _0 (x) - \phi _0 (y) }{x - y} h_0(y) \, dy \end{aligned}$$
(9.45)
$$\begin{aligned} T_2&:= \frac{1}{i\pi } \partial _x \int H \left( \frac{ h(x)-h(y) }{x-y} \right) \frac{ (h(x)-h(y))(\phi _0 (x) - \phi _0 (y)) }{ {(x - y)}^2 } h_0(y) \, dy , \end{aligned}$$
(9.46)

for some smooth function \(H\). We have expanded the denominator in (9.44) to obtain the above identity. Then, since \(\text{ Re }{\zeta }_{\alpha }-1\) decays like \({(1+t)}^{-1/2}\) in \(L^\infty \), in order to get (9.41) it suffices to prove

$$\begin{aligned} {\Vert T_1 \Vert }_{W^{N_0/2+3,\infty }}&\lesssim {\varepsilon }_1^2 {(1+t)}^{-1/2-{\gamma }} \end{aligned}$$
(9.47)
$$\begin{aligned} {\Vert T_2 \Vert }_{W^{N_0/2+3,\infty }}&\lesssim {\varepsilon }_1^2 {(1+t)}^{-1-{\gamma }} . \end{aligned}$$
(9.48)

Applying \(\partial _x\) to the integrand in (9.45), and using (11.40), we see that

$$\begin{aligned} {\Vert T_1 \Vert }_{W^{N_0/2+3,\infty }}&\lesssim {\left\| \int \frac{ \phi _0 (x) - \phi _0 (y) }{{(x - y)}^2} h_0(y) \, dy \right\| }_{W^{N_0/2+3,\infty }} + {\left\| \partial _x \phi _0 \, H_0 h_0 \right\| }_{W^{N_0/2+3,\infty }} \\&\lesssim {\left\| \phi _0 \right\| }_{W^{N_0/2+5,\infty }} \left( {\left\| h_0 \right\| }_{W^{N_0/2+4,\infty }} + {\left\| H_0 h_0 \right\| }_{W^{N_0/2+4,\infty }} \right) \\&\lesssim {\left\| \Lambda \phi \right\| }_{W^{N_0/2+6,\infty }} {\left\| h \right\| }_{W^{N_0/2+5,\infty }} \lesssim {\varepsilon }_1^2 {(1+t)}^{-1} . \end{aligned}$$

Similarly, again using (11.40), it is not hard to see that

$$\begin{aligned} {\Vert T_2 \Vert }_{W^{N_0/2+3,\infty }}&\lesssim {\left\| h \right\| }_{W^{N_0/2+5,\infty }} {\left\| \phi _0 \right\| }_{W^{N_0/2+5,\infty }} {\left\| h_0 \right\| }_{W^{N_0/2+4,\infty }} \lesssim {\varepsilon }_1^3 {(1+t)}^{-3/2} . \end{aligned}$$

This gives us (9.48) and concludes the proof of (9.41).

The remaining estimate (9.42) can be obtained similarly, using the \(L^2\) bounds of Theorem 11.3 instead of the \(L^\infty \) bound (11.40). In particular it suffices to show

$$\begin{aligned} {\Vert T_1 \Vert }_{H^{N_0/2+4}} \lesssim {\varepsilon }_1^2 {(1+t)}^{-{\gamma }} \qquad \text{ and } \qquad {\Vert T_2 \Vert }_{W^{N_0/2+3,\infty }} \lesssim {\varepsilon }_1^2 {(1+t)}^{-1/2-{\gamma }} . \end{aligned}$$

We only detail the bound for \(T_1\) as the one for \(T_2\) can proved similarly (notice that \(T_2\) is a cubic term and therefore its bounds are \({(1+t)}^{-1/2}\) better than those of \(T_1\)). Applying \(\partial _x\) to the integrand in (9.45), commuting derivatives via (11.39b), and using the estimates in Theorem 11.3, one sees that

$$\begin{aligned} {\Vert T_1 \Vert }_{H^{N_0/2+4}}&\lesssim {\left\| \int \frac{ \phi _0 (x) - \phi _0 (y) }{{(x - y)}^2} h_0(y) \, dy \right\| }_{H^{N_0/2+4}} + {\left\| \partial _x \phi _0 \, H_0 h_0 \right\| }_{H^{N_0/2+4}} \\&\lesssim {\left\| \phi _0 \right\| }_{H^{N_0/2+5}} \left( {\left\| h_0 \right\| }_{W^{N_0/2+4,\infty }} + {\left\| H_0 h_0 \right\| }_{W^{N_0/2+4,\infty }} \right) \lesssim {\varepsilon }_1^2 {(1+t)}^{-1/2} \end{aligned}$$

which is more than sufficient. \(\square \)

9.2 Proof of Lemma 9.2

The proof proceeds in several steps.

9.2.1 Step 1: The operator in Fourier space

Introduce the notations

$$\begin{aligned} \widetilde{\phi }&:= (I-H_0) \phi , \quad \widetilde{h} := (I-H_0) h \end{aligned}$$

so that from (9.8) we can write

$$\begin{aligned} \partial _x T_0(h,\phi ) = \partial _x [ \widetilde{\phi }_x ,H_0] \widetilde{h}_x . \end{aligned}$$

By taking Fourier transform, and using the notation (5.1), we see that

$$\begin{aligned} \partial _x T_0(h,\phi ) = M \big ( \widetilde{h}, \widetilde{\phi } \big ) \end{aligned}$$
(9.49)

where the symbol of the operator \(M\) is given by

$$\begin{aligned} m(\xi ,\eta ) = \xi \eta |\xi -\eta | -|\xi | \eta (\xi -\eta ) . \end{aligned}$$
(9.50)

We then want to show

$$\begin{aligned} {\big \Vert M \big ( \widetilde{h}, \widetilde{\phi } \big ) \big \Vert }_{W^{N_0/2+3,\infty }}&\lesssim {\varepsilon }_1^2 {(1+t)}^{-1-{\gamma }} \end{aligned}$$
(9.51)
$$\begin{aligned} {\big \Vert M \big ( \widetilde{h}, \widetilde{\phi } \big ) \big \Vert }_{H^{N_0/2+4}}&\lesssim {\varepsilon }_1^2 {(1+t)}^{-1/2-{\gamma }} \end{aligned}$$
(9.52)

for some \({\gamma }>0\), under the a priori assumptions (9.2) and (9.3).

9.2.2 Step 2: Approximation

Let \(H\) and \(\Psi \) be the functions defined in (4.17) in Proposition 4.1. Define

$$\begin{aligned} \widetilde{H} := (I-H_0) H , \quad \widetilde{\Psi } := (I-H_0) \Psi . \end{aligned}$$

We then claim that the following hold true:

$$\begin{aligned} {\big \Vert M \big ( \widetilde{h}, \widetilde{\phi } \big ) - M \big ( \widetilde{H}, \widetilde{\Psi } \big ) \big \Vert }_{W^{N_0/2+3,\infty }}&\lesssim {\varepsilon }_1^2 {(1+t)}^{-1-{\gamma }} , \end{aligned}$$
(9.53)
$$\begin{aligned} {\big \Vert M \big ( \widetilde{h}, \widetilde{\phi } \big ) - M \big ( \widetilde{H}, \widetilde{\Psi } \big ) \big \Vert }_{H^{N_0/2+4}}&\lesssim {\varepsilon }_1^2 {(1+t)}^{-1/2-{\gamma }} . \end{aligned}$$
(9.54)

Using (4.17) we see that

$$\begin{aligned} M \big ( \widetilde{H}, \widetilde{\Psi } \big ) - M \big ( \widetilde{h}, \widetilde{\phi } \big )&= M \big ( (I-H_0)A, \widetilde{\phi } \big ) + M \big ( \widetilde{H}, (I-H_0) B \big ) . \end{aligned}$$
(9.55)

From the definition of \(m\) in (9.50), the definitions (5.14) and (5.19), we see that

$$\begin{aligned} {\Vert m^{k,k_1,k_2}\Vert }_{\mathcal {S}^\infty } \lesssim 2^k 2^{k_1} 2^{k_2} . \end{aligned}$$
(9.56)

Combining this with the \(L^2\) and \(L^\infty \) bounds on \(A\) and \(B\) in (5.26) and (5.27), the a priori bounds (5.23), and using (5.17) in Lemma 5.2, one can easily verify that

$$\begin{aligned} {\big \Vert M \big ( (I-H_0)A, \widetilde{\phi } \big ) \big \Vert }_{W^{N_0/2+3,\infty }}&\lesssim {\varepsilon }_1^3 {(1+t)}^{-5/4} \\ {\big \Vert M \big ( \widetilde{H}, (I-H_0) B \big ) \big \Vert }_{W^{N_0/2+4,\infty }}&\lesssim {\varepsilon }_1^3 {(1+t)}^{-5/4} . \end{aligned}$$

These and (9.55) give us (9.53). Similarly one can show (9.54) again by using (9.56), Lemma 5.2, (5.23) and the bounds on the Sobolev norms of \(A\) and \(B\) provided by (5.26).

9.2.3 Step 3: Reduction to bilinear estimates

We are left with proving

$$\begin{aligned} {\big \Vert M \big ( \widetilde{H}, \widetilde{\Psi } \big ) \big \Vert }_{W^{N_0/2+3,\infty }}&\lesssim {\varepsilon }_1^2 {(1+t)}^{-1-{\gamma }} , \end{aligned}$$
(9.57)
$$\begin{aligned} {\big \Vert M \big ( \widetilde{H}, \widetilde{\Psi } \big ) \big \Vert }_{H^{N_0/2+4}}&\lesssim {\varepsilon }_1^2 {(1+t)}^{-1/2-{\gamma }} . \end{aligned}$$
(9.58)

Recalling the definition of \(V = H+i\Lambda \Psi \) from (5.6), we have \(\widetilde{V}_\pm = \widetilde{H} \pm i\Lambda \widetilde{\Psi }\), so that \(\widetilde{H} = (\widetilde{V}_+ + \widetilde{V}_-)/2\) and \(\widetilde{\Psi } = (i\Lambda )^{-1} (\widetilde{V}_+ - \widetilde{V}_-)/2\). We then have

$$\begin{aligned} M \big ( \widetilde{H},\widetilde{\Psi } \big ) = \sum _{{\varepsilon }_1,{\varepsilon }_2 \in \{+,-\}} c_{{\varepsilon }_1,{\varepsilon }_2} Q \big ( \widetilde{V}_{{\varepsilon }_1}, \widetilde{V}_{{\varepsilon }_2} \big ) \end{aligned}$$
(9.59)

where, again according to the notation (5.1),

$$\begin{aligned} q (\xi ,\eta ) = {|\eta |}^{-1/2} \left[ \xi \eta |\xi -\eta | -|\xi | \eta (\xi -\eta ) \right] , \end{aligned}$$
(9.60)

and \(c_{{\varepsilon }_1,{\varepsilon }_2}\) are some constants. With the notation (5.14) and (5.19) we have

$$\begin{aligned} {\Vert q^{k,k_1,k_2}\Vert }_{\mathcal {S}^\infty } \lesssim 2^k 2^{k_1} 2^{k_2/2} . \end{aligned}$$
(9.61)

From the definition of \(\widetilde{V}\) and the bounds provided by (4.25), (4.26) and (4.30) on \(V = H + i \Lambda \Psi \), we see that the desired bounds (9.57)–(9.58) reduce to showing the following bilinear estimates:

$$\begin{aligned} {\left\| Q (v_\pm (t), v_\pm (t)) \right\| }_{W^{N_0/2+3,\infty }}&\lesssim {\varepsilon }_1^2 {(1+t)}^{-1-{\gamma }} , \end{aligned}$$
(9.62)
$$\begin{aligned} {\left\| Q (v_\pm (t), v_\pm (t)) \right\| }_{H^{N_0/2+4}}&\lesssim {\varepsilon }_1^2 {(1+t)}^{-1/2-{\gamma }} , \end{aligned}$$
(9.63)

where \(q\) is as in (9.60), and \(v_\pm = e^{\mp it\Lambda }f_\pm \) satisfies, for all \(k \in {\mathbb {Z}}\),

$$\begin{aligned} {\Vert P_k v_\pm (t) \Vert }_{W^{N_1,\infty }}&\lesssim {\varepsilon }_1 {(1+t)}^{-1/2} , \end{aligned}$$
(9.64)
$$\begin{aligned} {\Vert P_k f_\pm (t) \Vert }_{H^{N_0-10}}&\lesssim {\varepsilon }_1 {(1+t)}^{5p_0} , \end{aligned}$$
(9.65)
$$\begin{aligned} {\Vert x \partial _x P_k f_\pm (t) \Vert }_{H^{\frac{N_0}{2}-20}}&\lesssim {\varepsilon }_1 {(1+t)}^{5p_0} . \end{aligned}$$
(9.66)

9.2.4 Step 4: Proof of the bilinear estimates (9.62)–(9.63)

Step 1: Frequency decomposition. It suffices to show that for all \(t \in [2^m-2, 2^{m+1}]\) and \(m \in \{1,2,\dots \}\), there exists a constant \({\gamma }>0\) such that

$$\begin{aligned} \sum _{k,k_1,k_2} {\left\| P_k Q (P_{k_1} v_\pm (t), P_{k_2} v_\pm (t)) \right\| }_{W^{N_0/2+3,\infty }}&\lesssim {\varepsilon }_1^2 2^{-(1+{\gamma })m} , \end{aligned}$$
(9.67)
$$\begin{aligned} \sum _{k,k_1,k_2} {\left\| P_k Q (P_{k_1} v_\pm (t), P_{k_2} v_\pm (t)) \right\| }_{H^{N_0/2+4}}&\lesssim {\varepsilon }_1^2 2^{-(1/2+{\gamma })m} . \end{aligned}$$
(9.68)

By symmetry and conjugation it is clear that we can reduce matters to proving the estimates (9.67)–(9.68) for the two bilinear operators

$$\begin{aligned} T_+(f,f)(t)&:= {\mathcal {F}}^{-1} \int _{{\mathbb {R}}} e^{it\Phi _+} q(\xi ,\eta ) \widehat{f}_+(t,\xi -\eta ) \widehat{f}_+(t,\eta ) \, d\eta , \end{aligned}$$
(9.69)
$$\begin{aligned} T_-(f,f)(t)&:= {\mathcal {F}}^{-1} \int _{{\mathbb {R}}} e^{it\Phi _-} q(\xi ,\eta ) \widehat{f}_+(t,\xi -\eta ) \widehat{f}_-(t,\eta ) \, d\eta , \end{aligned}$$
(9.70)

where \(f = (f_+, f_-)\), and

$$\begin{aligned} \Phi _\pm (\xi ,\eta ) \!=\! {|\eta |}^{1/2} \pm {|\xi - \eta |}^{1/2} , \quad q (\xi ,\eta ) \!=\! {|\eta |}^{-1/2} \left[ \xi \eta |\xi \!-\!\eta | \!-\!|\xi | \eta (\xi -\eta ) \right] . \end{aligned}$$
(9.71)

Notice that under the a priori assumptions these bilinear terms have decay rates which barely fail to give (9.67) and (9.68). The key to obtaining the extra necessary decay is the vanishing of the symbol \(q(\xi ,\eta )\) on the space-resonant sets, i.e. for those \((\xi ,\eta )\) such that \(\partial _\eta \Phi _\pm (\xi ,\eta ) = 0\). One can then use integration by parts in frequency, and the weighted bound (9.66), to derive the desired estimate.

First let us observe that using (9.61), Lemma 5.2, and (9.64)–(9.66), one can bound as desired the sums in (9.67) and (9.68), for all those frequencies \((k,k_1,k_2)\) such that \(\min (k,k_1,k_2) \le -m/N_0\) and \(\max (k,k_1,k_2) \ge 3m/N_0\). The remaining sums have only \(C m^3\) terms. Therefore it suffices to show the estimates for each \((k,k_1,k_2)\) fixed satisfying

$$\begin{aligned} k,k_1,k_2 \in [-m/N_0,3m/N_0] \cap {\mathbb {Z}}, \quad \max (k_1, k_2) \ge k - 10 . \end{aligned}$$
(9.72)

Step 2: Spatial decomposition. Let us define \(\rho :\mathbb {R}\rightarrow [0,1]\) to be an even compactly supported function which equals \(1\) on \([0,1]\) and vanishes on \([2,\infty )\). Let \(R := 2^{3m/4}\). We decompose the profiles \(f = (f_+,f_-)\) into two pieces: \(f = f_{\ge R} + f_{\le R} = (f_{+,\ge R}, f_{-,\ge R}) + (f_{+,\le R}, f_{-,\le R})\) where

$$\begin{aligned} f_{\le R}(x) = f(x) \rho \left( \frac{x}{R} \right) , \quad f_{\ge R}(x) = f(x) - f_{\le R}(x) . \end{aligned}$$

We then want to show that for all \(t \in [2^m-2, 2^{m+1}]\), \(m \in \{1,2,\dots \}\), and \(k,k_1,k_2 \in [-m/N_0,3m/N_0] \cap {\mathbb {Z}}\)

$$\begin{aligned}&{\left\| P_k T_\pm (P_{k_1} f_{\ge R}(t), P_{k_2} f(t)) \right\| }_{L^\infty } + {\left\| P_k T_\pm (P_{k_1} f_{\le R}(t), P_{k_2} f_{\ge R}(t)) \right\| }_{L^\infty } \nonumber \\&\quad \lesssim {\varepsilon }_1^2 2^{-(1+{\gamma })m} 2^{-(N_0/2+3)k_+} , \end{aligned}$$
(9.73)
$$\begin{aligned}&{\left\| P_k T_\pm (P_{k_1} f_{\le R}(t), P_{k_2} f_{\le R}(t)) \right\| }_{L^\infty } \lesssim {\varepsilon }_1^2 2^{-(1+{\gamma })m} 2^{-(N_0/2+3)k_+} . \end{aligned}$$
(9.74)

We also need to prove the \(H^{N_0/2+4}\) versions of the above estimates corresponding to (9.68), but since those can be obtained analogously we will skip them.

Step 3: Proof of (9.73). First notice that both terms in (9.73) have (at least) one profile supported at a distance \(R\) from the origin. Since this is the only important aspect that we will use to gain the necessary decay, we only show how to estimate one of the terms, the other being analogous. We then want to prove

$$\begin{aligned} {\left\| P_k T_\pm (P_{k_1} f(t), P_{k_2} f_{\ge R}(t)) \right\| }_{L^\infty } \lesssim {\varepsilon }_1^2 2^{-9m/8} 2^{-(N_0/2+3)k_+} \end{aligned}$$

for any \(k,k_1,k_2 \in [-m/N_0, 3m/N_0] \cap \mathbb {Z}\), and any \(t \in [2^m-2, 2^{m+1}]\). Using (9.61), Sobolev’s embedding, and the bounds (9.64) and (9.66), we see that

$$\begin{aligned} {\left\| P_k T_\pm (P_{k_1} f(t), P_{k_2} f_{\ge R}(t)) \right\| }_{L^\infty }&\lesssim 2^k 2^{k_1} 2^{k_2/2} {\left\| P_{k_1} v_\pm (t) \right\| }_{L^\infty } {\left\| e^{\pm it\Lambda } P_{k_2} f_{\ge R}(t) \right\| }_{L^\infty } \\&\lesssim 2^k 2^{k_1} {\varepsilon }_1 2^{-m/2} 2^{-N_1 \max (k_1,0)} {\left\| \partial _x f_{\ge R}(t) \right\| }_{L^2} \\&\lesssim {\varepsilon }_1^2 2^k 2^{k_1} 2^{-m/2} R^{-1} 2^{-(N_0/2-20) \max (k_1,k_2,0)} 2^{mp_0} \\&\lesssim {\varepsilon }_1^2 2^{-9m/8} 2^{-(N_0/2+3)k_+} , \end{aligned}$$

in view of the frequency constraints (9.72), \(p_0 \le 1/1000\) and \(R = 2^{3m/4}\).

Step 4: Proof of (9.74)

It suffices to show

$$\begin{aligned} {\left\| P_k T_\pm (P_{k_1} f_{\le R}(t), P_{k_2} f_{\le R}(t)) \right\| }_{L^\infty } \lesssim {\varepsilon }_1^2 2^{-9m/8} 2^{-(N_0/2+3)k_+} , \end{aligned}$$

for any \(k,k_1,k_2\) as in (9.72) and \(t \in [2^m-2, 2^{m+1}]\). We distinguish the two cases of \(T_+\) and \(T_-\). In the first case we introduce and extra cutoff in \(\xi - 2\eta \) by writing

$$\begin{aligned}&P_k T_+ ( P_{k_1} f_{\le R}(t), P_{k_2} f_{\le R}(t) )\nonumber \\&\quad = \sum _{l \in {\mathbb {Z}}} P_k T^l_+ ( P_{k_1} f_{\le R}(t), P_{k_2} f_{\le R}(t) ) \nonumber \\&\quad := \sum _{l \in {\mathbb {Z}}} {\mathcal {F}}^{-1} \int _{{\mathbb {R}}} e^{it\Phi _+(\xi ,\eta )} \varphi _{l}(\xi -2\eta )\nonumber \\&\quad \quad \,\times \varphi _{k}(\xi ) \varphi _{k_1}(\xi -\eta ) \varphi _{k_2}(\eta ) q(\xi ,\eta ) \widehat{f_{+,\le R}}(t,\xi -\eta ) \widehat{f_{+,\le R}}(t,\eta ) \, d\eta \end{aligned}$$
(9.75)

where \(\Phi _+\) is as in (9.71). First notice that the contribution in the summation over \(l\) in (9.75) is zero if \(l \ge 3m/N_0 + 100\). (9.75) also vanishes if \(l \le -m/N_0 - 100\), because in this case \(|\xi - 2\eta | \le |\eta |/10\), which implies that \(\xi - \eta \) and \(\xi \) have the same sign, and therefore \(q(\xi ,\eta ) = 0\), see (9.71). The summation over \(l\) can then be disregarded and it is enough to show

$$\begin{aligned} {\left\| P_k T_+^l (P_{k_1} f_{\le R}(t), P_{k_2} f_{\le R}(t)) \right\| }_{L^\infty } \lesssim {\varepsilon }_1^2 2^{-5m/4} 2^{-(N_0/2+3)k_+} \end{aligned}$$

for any \(l,k,k_1,k_2 \in [-m/N_0 - 100, 3m/N_0 + 100] \cap \mathbb {Z}\) and \(t \in [2^m-2, 2^{m+1}]\). Observe that for any integer \(j\) we have

$$\begin{aligned} {\big | \partial _\eta ^j \partial _\eta \Phi _+(\xi ,\eta ) \big |} \gtrsim 2^{- j m/10} \end{aligned}$$

and

$$\begin{aligned} {\big \Vert \partial _\eta ^j \varphi _{k_i}(\cdot ) \widehat{f_{\le R}}(t,\cdot ) \big \Vert }_{L^1} \lesssim R^j 2^{k_i/2} 2^{mp_0} , \quad \left| \partial _\eta ^j ( q(\xi ,\eta ) \varphi _{l}(\xi -2\eta ) )\right| \lesssim 2^{(j + 5/2)m/N_0} . \end{aligned}$$

Integrating by parts \(L\) times in \(\eta \) in the integral in (9.75) we see that

$$\begin{aligned}&{\big \Vert P_k T_+^l (P_{k_1} f_{\le R}(t), P_{k_2} f_{\le R}(t)) \big \Vert }_{L^\infty }\\&\quad \lesssim {\left( \frac{1}{2^m 2^{-m/10}} \right) }^{L} R^{L} \, 2^{k_1/2} 2^{k_2/2} 2^{2mp_0} 2^{- (N_0 - 10)\max (k_1,k_2,0)} {\varepsilon }_1^2 \\&\quad \lesssim {\varepsilon }_1^2 2^{-3m/2} 2^{- (N_0/2 + 3)k_+} , \end{aligned}$$

where the second inequality holds provided \(L\) is large enough.

In the case of the operator \(T_-\) it is enough to observe that \(\partial _\eta \Phi _-\) vanishes (linearly) only when \(\xi = 0\), and, also in this case, we have

$$\begin{aligned} {|\partial _\eta ^j \partial _\eta \Phi _-|} \gtrsim 2^{- j m/10} , \end{aligned}$$

under the frequency constraints (9.72). One can then use the same integration by parts argument as above (without the need to resort to a further splitting) and obtain

$$\begin{aligned} {\left\| P_k T_- (P_{k_1} f_{\le R}(t), P_{k_2} f_{\le R}(t)) \right\| }_{L^\infty } \lesssim {\varepsilon }_1^2 2^{-3m/2} 2^{- (N_0/2+3)k_+} . \end{aligned}$$

This concludes the proof of (9.74) and therefore of Lemma 9.2. \(\square \)

10 Proof of Proposition 2.4: transition to Eulerian coordinates

Here we want to transfer the a priori bounds from the modified Lagrangian coordinates to Eulerian coordinates. Recall that

$$\begin{aligned} \widetilde{L}(t,{\alpha }) = ( {\zeta }_{\alpha }(t,{\alpha }) - 1, u(t,{\alpha }), w(t,{\alpha }), \text{ Im }{\zeta }(t,{\alpha }) ) . \end{aligned}$$

Also, recall the a priori assumptions (2.9) and (2.8), that is

$$\begin{aligned} \sup _{t\in [0,T]} \left( (1+t)^{-p_0} {\Vert \widetilde{L}(t) \Vert }_{X_{N_0}} + {\Vert \widetilde{L}(t) \Vert }_{H^{N_1+5}} + \sqrt{1+t} {\Vert \widetilde{L}(t) \Vert }_{W^{N_1,\infty } } \right) \le {\varepsilon }_1 , \end{aligned}$$
(10.1)

and

$$\begin{aligned}&\sup _{[0,T]} \left( (1+t)^{-p_0} {\Vert (h(t), \phi _x(t)) \Vert }_{X_{N_0}} + {\Vert h(t) + i\Lambda \phi (t) \Vert }_{H^{N_1 + 10}}\right. \nonumber \\&\quad \left. + \sqrt{1+t} {\Vert (h(t), \Lambda \phi (t)) \Vert }_{W^{N_1+4,\infty }} \right) \le {\varepsilon }_1 , \end{aligned}$$
(10.2)

where \({\Vert f \Vert }_{X_{N_0}} := {\Vert f \Vert }_{H^{N_0}} + {\Vert Sf \Vert }_{H^{N_0/2}}\). To prove Proposition 2.4 we need to show, under the above a priori assumptions, that

$$\begin{aligned} {\Vert \widetilde{L}(t) \Vert }_{W^{N_1,\infty }}&\lesssim {\Vert (h(t) , \partial _x \phi (t)) \Vert }_{Z^{\prime }} \end{aligned}$$
(10.3)
$$\begin{aligned} {\Vert \widetilde{L}(t) \Vert }_{H^{N_1+5}}&\lesssim {\Vert (h(t) , \partial _x \phi (t)) \Vert }_{H^{N_1+7}} + {\varepsilon }_1^2 \end{aligned}$$
(10.4)
$$\begin{aligned} {\Vert (h(t) , \partial _x \phi (t)) \Vert }_{X_{N_0}}&\lesssim {\Vert \widetilde{L}(t) \Vert }_{X_{N_0}} . \end{aligned}$$
(10.5)

10.1 Proof of (10.3)

From Proposition 8.3 we have

$$\begin{aligned} {\Vert L(t) \Vert }_{W^{N_1,\infty }} \lesssim {\Vert (h(t),\partial _x \phi (t)) \Vert }_{Z^{\prime }} \end{aligned}$$

which is stronger than (10.3).

10.2 Proof of (10.4)

Using \(h (\text{ Re }{\zeta }) = \text{ Im }{\zeta }\) it is clear that

$$\begin{aligned} {\Vert \text{ Im }{\zeta }\Vert }_{H^{N_1+5}}&\lesssim {\Vert h \Vert }_{H^{N_1+5}} . \end{aligned}$$

Also, from (8.126)–(8.127) in Lemma 8.9, and the a priori assumptions (8.1), we see that

$$\begin{aligned}&{\Vert ({\zeta }_{\alpha }(t)-1,u(t),w(t)) \Vert }_{H^{N_1+5}}\\&\quad \lesssim {\Vert (\partial _{\alpha }\chi (t), \partial _{\alpha }{\lambda }(t)) \Vert }_{H^{N_1+5}} + {\Vert \widetilde{L} (t) \Vert }_{H^{N_1+6}} {\Vert \widetilde{L} (t) \Vert }_{W^{\frac{N_1+5}{2},\infty }} \\&\quad \lesssim {\Vert (\partial _{\alpha }\chi (t), \partial _{\alpha }{\lambda }(t)) \Vert }_{H^{N_1+5}} + {\varepsilon }_1 {(1+t)}^{p_0} {\varepsilon }_1 {(1+t)}^{-1/2} \\&\quad \lesssim {\Vert (\partial _{\alpha }\chi (t), \partial _{\alpha }{\lambda }(t)) \Vert }_{H^{N_1+5}} + {\varepsilon }_1^2. \end{aligned}$$

Thus, to obtain (10.4) it is enough to show

$$\begin{aligned} {\Vert (\partial _{\alpha }\chi (t), \partial _{\alpha }{\lambda }(t)) \Vert }_{H^{N_1+5}}&\lesssim {\Vert (h(t) , \partial _x \phi (t)) \Vert }_{H^{N_1+7}} + {\varepsilon }_1^2 . \end{aligned}$$
(10.6)

From the definition of \(\chi \) in (7.4), and \(\text{ Im }{\zeta }= h \circ \text{ Re }{\zeta }\), we see that

$$\begin{aligned} \chi = (I- {\mathcal {H} } ) ({\zeta }-\overline{{\zeta }}) = 2i (I- {\mathcal {H} } ) \text{ Im }{\zeta }= 2i (I- {\mathcal {H} } ) (h \circ \text{ Re }{\zeta }) . \end{aligned}$$

Using the bounds on the Hilbert transform (11.1) with the a priori assumptions (10.1)–(10.2), we get

$$\begin{aligned} {\Vert \partial _{\alpha }\chi (t) \Vert }_{H^{N_1+5}}&\lesssim {\Vert h \circ \text{ Re }{\zeta }(t) \Vert }_{H^{N_1+6}} + {\varepsilon }_1^2 \lesssim {\Vert h(t) \Vert }_{H^{N_1+6}} + {\varepsilon }_1^2 \, \end{aligned}$$

which gives the bound (10.6) for the component \(\chi _{\alpha }\).

From (8.142) and (8.138) in the proof of Lemma 8.10, we see that

$$\begin{aligned}&{\Vert \partial _{\alpha }{\lambda }\Vert }_{H^{N_1+5}} \lesssim {\Vert \partial _x [\phi (1+h^{\prime })] \Vert }_{H^{N_1+5}}\nonumber \\&\quad + {\Vert R_\phi \Vert }_{H^{N_1+6}} \lesssim {\Vert \partial _x \phi \Vert }_{H^{N_1+5}} + {\Vert \phi h^{\prime }\Vert }_{H^{N_1+6}} + {\Vert R_\phi \Vert }_{H^{N_1+6}} , \end{aligned}$$

where we recall

$$\begin{aligned} R_\phi (t,x) := \frac{1}{i\pi } \int H\left( \frac{h(x) - h(y)}{x-y} \right) \frac{h(x) - h(y)}{ {(x-y)}^2 } \phi (t,y) \, dy. \end{aligned}$$

Applying \(N_1 + 6\) derivatives to the above expression, commuting them via (11.39b), and using Theorem 11.3, one can obtain

$$\begin{aligned} {\Vert R_\phi \Vert }_{H^{N_1+6}} \lesssim {\Vert h \Vert }_{H^{N_1+7}} {\Vert \phi \Vert }_{W^{N_1+4,\infty }} + {\Vert h \Vert }_{W^{N_1+4,\infty }} {\Vert \phi _x \Vert }_{H^{N_1+5}} \lesssim {\varepsilon }_1^2 . \end{aligned}$$

Here we have used the a priori assumptions (10.2) and the bound (4.35) in Corollary 4.5. A similar estimate can be easily obtained for \({\Vert \phi h^{\prime }\Vert }_{H^{N_1+6}}\). It follows that

$$\begin{aligned} {\Vert \partial _{\alpha }{\lambda }\Vert }_{H^{N_1+6}} \lesssim {\Vert (h, \phi _x) \Vert }_{H^{N_1+7}} + {\varepsilon }_1^2 , \end{aligned}$$

which gives (10.6), completing the proof of (10.4).

10.3 Proof of (10.5)

To show (10.5) we will exploit the identities

$$\begin{aligned} h(t, \text{ Re }{\zeta }(t,{\alpha })) = \text{ Im }{\zeta }(t,{\alpha }) \qquad \text{ and } \qquad (I- {\mathcal {H} } _{\zeta }) \phi (t, \text{ Re }{\zeta }(t,{\alpha })) = {\lambda }(t,{\alpha }) \end{aligned}$$

to prove the following:

Lemma 10.1

Let \(S\) be the scaling vector field. Then

$$\begin{aligned}&{\Vert h(t) \Vert }_{X_{N_0}} \lesssim {\Vert \text{ Im }{\zeta }(t) \Vert }_{X_{N_0}} + {\Vert {\zeta }_{\alpha }(t) - 1 \Vert }_{X_{N_0}} \end{aligned}$$
(10.7)

and

$$\begin{aligned}&{\Vert \partial _x \phi (t) \Vert }_{X_{N_0}} \lesssim {\Vert {\lambda }_{\alpha }(t) \Vert }_{X_{N_0}} + {\Vert {\zeta }_{\alpha }(t) - 1 \Vert }_{X_{N_0}} . \end{aligned}$$
(10.8)

The estimates in the above Lemma, together with (8.33), imply (10.5). To prove the estimates involving \(S\) we will need the two auxiliary Lemmas below:

Lemma 10.2

Let \(S\) be the scaling vector field. Then for any two functions \(f: {\mathbb {R}}_t \times {\mathbb {R}}_x \rightarrow {\mathbb {R}}\) and \(g: {\mathbb {R}}_t \times {\mathbb {R}}_\alpha \rightarrow {\mathbb {R}}_x \), the following formula holds:

$$\begin{aligned} S f \circ g = S (f \circ g) - (f^{\prime }\circ g) (S g - g) . \end{aligned}$$
(10.9)

Here, for a function \(c:{\mathbb {R}}_t \times {\mathbb {R}}_x \rightarrow {\mathbb {R}}\), the operation \(c \circ g\) is to be understood as composition in the space variable:

$$\begin{aligned} (c \circ g) (t,\alpha ) = c(t,g(t,\alpha )) , \end{aligned}$$

and \(f^{\prime }\) denotes the derivative with respect to the space variable.

The proof of the above statement is by direct computation.

Lemma 10.3

Let \({\zeta }\) be the Lagrangian map in the modified Lagrangian coordinates, then

$$\begin{aligned} (I- {\mathcal {H} } )(\text{ Re }{\zeta }- {\alpha }) = i (I- {\mathcal {H} } ) \text{ Im }{\zeta }. \end{aligned}$$
(10.10)

Under the a priori assumptions (10.1)–(10.2), it follows that

$$\begin{aligned} {\Vert \text{ Re }{\zeta }- {\alpha }\Vert }_{X_{N_0}}&\lesssim {\Vert \text{ Im }{\zeta }\Vert }_{X_{N_0}} + {\Vert {\zeta }_{\alpha }-1 \Vert }_{X_{N_0}} . \end{aligned}$$
(10.11)

In particular

$$\begin{aligned} {\Vert S \text{ Re }{\zeta }- \text{ Re }{\zeta }\Vert }_{H^{\frac{N_0}{2}}}&\lesssim {\Vert \text{ Im }{\zeta }\Vert }_{X_{N_0}} + {\Vert {\zeta }_{\alpha }-1 \Vert }_{X_{N_0}} \lesssim {\Vert \widetilde{L} \Vert }_{X_{N_0}} . \end{aligned}$$
(10.12)

Proof

The identity (10.10) follows from the fact that \((I- {\mathcal {H} } _{\zeta })(\overline{{\zeta }}-{\alpha }) = 0\), which comes from the identity \((I- {\mathcal {H} } _z)(k-\overline{z}) = 0\) upon composition with \(k^{-1}\).

The estimate (10.11) follows from an application of Lemma 11.2 with \(f = \text{ Re }{\zeta }-{\alpha }\) and \(g = i (I- {\mathcal {H} } ) \text{ Im }{\zeta }\), Sobolev’s embedding, and from the boundedness properties of \( {\mathcal {H} } \) in Lemma 11.1:

$$\begin{aligned} {\Vert \text{ Re }{\zeta }- {\alpha }\Vert }_{X_{N_0}}&\lesssim {\Vert (I- {\mathcal {H} } ) \text{ Im }{\zeta }\Vert }_{X_{N_0}} + {\Vert {\zeta }_{\alpha }- 1 \Vert }_{X_{N_0}} {\Vert (I- {\mathcal {H} } ) \text{ Im }{\zeta }\Vert }_{H^{N_1}} \\&\lesssim {\Vert \text{ Im }{\zeta }\Vert }_{X_{N_0}} \!+\! {\Vert {\zeta }_{\alpha }- 1 \Vert }_{X_{N_0}} ( {\Vert \text{ Im }{\zeta }\Vert }_{H^{N_1}} \!+\! {\Vert (I\!-\! {\mathcal {H} } ) \text{ Im }{\zeta }\Vert }_{H^{N_1}} )\\&\lesssim {\Vert \text{ Im }{\zeta }\Vert }_{X_{N_0}} + {\Vert {\zeta }_{\alpha }- 1 \Vert }_{X_{N_0}} . \end{aligned}$$

(10.12) immediately follows since \(S\text{ Re }{\zeta }- \text{ Re }{\zeta }= S(\text{ Re }{\zeta }- {\alpha }) + ({\alpha }-\text{ Re }{\zeta })\). \(\square \)

Proof of Lemma 10.1

Recall that \(h(t, \text{ Re }{\zeta }(t,{\alpha })) = \text{ Im }{\zeta }(t,{\alpha })\). Since \({\Vert {\zeta }_{\alpha }- 1 \Vert }_{H^{N_1+5}} \le {\varepsilon }_1\), in particular we see that for \({\varepsilon }_1 \le 1/2\), the map \(\text{ Re }{\zeta }\) is a diffeomorphism with

$$\begin{aligned} \big | \partial _{\alpha }\text{ Re }{\zeta }\big | \ge 1/2 , \quad \big | \partial _{\alpha }^k \text{ Re }{\zeta }\big | \le 3/2 \end{aligned}$$

for \(1 \le k \le N_1\). It immediately follows that

$$\begin{aligned} {\Vert h \Vert }_{H^{N_0}} = {\Vert \text{ Im }{\zeta }\circ \text{ Re }{\zeta }^{-1} \Vert }_{H^{N_0}}&\lesssim {\Vert \text{ Im }{\zeta }(t) \Vert }_{H^{N_0}} + {\Vert \text{ Re }{\zeta }_{{\alpha }{\alpha }}(t) \Vert }_{H^{N_0-1}}\\&\lesssim {\Vert \text{ Im }{\zeta }\Vert }_{H^{N_0}} + {\Vert {\zeta }_{\alpha }(t) - 1\Vert }_{H^{N_0}} . \end{aligned}$$

This takes care of the Sobolev component of the norm to bound in (10.7). To estimate the weighted component we apply Lemma 10.2 to get

$$\begin{aligned} S h \circ \text{ Re }{\zeta }= S \text{ Im }{\zeta }- (h^{\prime }\circ \text{ Re }{\zeta }) (S \text{ Re }{\zeta }- \text{ Re }{\zeta }) . \end{aligned}$$
(10.13)

Using (10.12) we have

$$\begin{aligned} {\Vert S h \Vert }_{H^\frac{N_0}{2}} \lesssim {\Vert S \text{ Im }{\zeta }\Vert }_{H^\frac{N_0}{2}} \!+\! {\Vert h^{\prime }\Vert }_{H^\frac{N_0}{2}} {\Vert S \text{ Re }{\zeta }\!-\! \text{ Re }{\zeta }\Vert }_{H^\frac{N_0}{2}} \lesssim {\Vert \text{ Im }{\zeta }\Vert }_{X_{N_0}} \!+\! {\Vert {\zeta }_{\alpha }-1 \Vert }_{X_{N_0}} . \end{aligned}$$

This gives us (10.7).

Recall the relation between the trace of the velocity potential in Eulerian variables and in modified Lagrangian variables \({\lambda }\):

$$\begin{aligned} (I- {\mathcal {H} } _{\zeta }) \phi (t, \text{ Re }{\zeta }(t,{\alpha }))&= {\lambda }(t,{\alpha }) . \end{aligned}$$
(10.14)

Since

$$\begin{aligned} {[\partial _{\alpha }, {\mathcal {H} } ]} f = [{\zeta }_{\alpha }, {\mathcal {H} } ] \frac{f_{\alpha }}{{\zeta }_{\alpha }} = \frac{1}{i\pi } Q_0 ({\zeta }_{\alpha }-1, f_{\alpha }) , \end{aligned}$$

where \(Q_0\) is defined in (11.18), it follows that

$$\begin{aligned} (I- {\mathcal {H} } ) \partial _{\alpha }(\phi \circ \text{ Re }{\zeta })&= \partial _{\alpha }{\lambda }+ Q_0 \left( {\zeta }_{\alpha }-1, \partial _{\alpha }(\phi \circ \text{ Re }{\zeta }) \right) . \end{aligned}$$
(10.15)

Denoting \(f = \phi \circ \text{ Re }{\zeta }\) and \(g = \partial _{\alpha }{\lambda }+ \frac{1}{i\pi } Q_0 \left( {\zeta }_{\alpha }-1, \partial _{\alpha }(\phi \circ \text{ Re }{\zeta }) \right) \) we have \((I- {\mathcal {H} } ) \partial _{\alpha }f = g\), so that (11.7) gives

$$\begin{aligned} {\Vert \partial _{\alpha }f \Vert }_{H^{N_1}}&\lesssim {\Vert g \Vert }_{H^{N_1}} \lesssim {\Vert \partial _{\alpha }{\lambda }\Vert }_{H^{N_1}} + {\Vert Q_0 ({\zeta }_{\alpha }-1, \partial _{\alpha }f) \Vert }_{H^{N_1}} , \end{aligned}$$
(10.16)

while (11.9) gives

$$\begin{aligned} {\Vert \partial _{\alpha }f \Vert }_{X_{N_0}}&\lesssim {\Vert g \Vert }_{X_{N_0}} + {\Vert {\zeta }_{\alpha }-1 \Vert }_{X_{N_0}} {\Vert g \Vert }_{H^{N_1}} \lesssim {\Vert \partial _{\alpha }{\lambda }\Vert }_{X_{N_0}} + {\Vert Q_0 ({\zeta }_{\alpha }-1, \partial _{\alpha }f) \Vert }_{X_{N_0}} \nonumber \\&\quad + {\Vert {\zeta }_{\alpha }-1 \Vert }_{X_{N_0}} \left( {\Vert \partial _{\alpha }{\lambda }\Vert }_{H^{N_1}} + {\Vert Q_0 ({\zeta }_{\alpha }-1, \partial _{\alpha }f) \Vert }_{H^{N_1}} \right) . \end{aligned}$$
(10.17)

From (8.32)–(8.33) we have

$$\begin{aligned} {\Vert \partial _{\alpha }{\lambda }\Vert }_{H^{N_1}} \lesssim {\Vert \widetilde{L} \Vert }_{H^{N_1}} \lesssim {\varepsilon }_1 , \quad {\Vert \partial _{\alpha }{\lambda }\Vert }_{X_{N_0}} \lesssim {\Vert \widetilde{L} \Vert }_{X_{N_0}} . \end{aligned}$$

From estimate (11.31) we get

$$\begin{aligned} {\Vert Q_0 ({\zeta }_{\alpha }-1, \partial _{\alpha }f) \Vert }_{X_{N_0}}&\lesssim {\Vert {\zeta }_{\alpha }- 1 \Vert }_{W^{\frac{N_0}{2}+1,\infty }} {\Vert \partial _{\alpha }f \Vert }_{X_{N_0}}\\&\quad + {\Vert (\partial _{\alpha }f, {\mathcal {H} } \partial _{\alpha }f) \Vert }_{W^{\frac{N_0}{2}+1,\infty }} {\Vert {\zeta }_{\alpha }- 1 \Vert }_{X_{N_0}} \\&\quad + {\Vert {\zeta }_{\alpha }- 1 \Vert }_{W^{\frac{N_0}{2}+1,\infty }} {\Vert \partial _{\alpha }f \Vert }_{W^{\frac{N_0}{2}+1,\infty }} {\Vert {\zeta }_{\alpha }-1 \Vert }_{X_{N_0}} \\&\lesssim {\varepsilon }_1 {\Vert \partial _{\alpha }f \Vert }_{X_{N_0}} + {\Vert (\partial _{\alpha }f, {\mathcal {H} } \partial _{\alpha }f) \Vert }_{W^{\frac{N_0}{2}+1,\infty }} {\Vert \widetilde{L} \Vert }_{X_{N_0}} . \end{aligned}$$

Next we claim that

$$\begin{aligned} {\Vert (\partial _{\alpha }f, {\mathcal {H} } \partial _{\alpha }f) \Vert }_{W^{\frac{N_0}{2}+1,\infty }} \lesssim {\varepsilon }_1 . \end{aligned}$$
(10.18)

Assuming this estimate for now, and using the fact that \({\Vert Q_0 ({\zeta }_{\alpha }-1, \partial _{\alpha }f) \Vert }_{H^{N_1}}\) can be bounded uniformly in time in a straightforward fashion, we can use the last four bounds in (10.17) to deduce that

$$\begin{aligned}&{\Vert \partial _{\alpha }f \Vert }_{X_{N_0}} \lesssim {\Vert \widetilde{L}\Vert }_{X_{N_0}} . \end{aligned}$$
(10.19)

Them, since \(\phi _x = (\partial _{\alpha }f/\text{ Re }{\zeta }_{\alpha }) \circ \text{ Re }{\zeta }^{-1}\), we have

$$\begin{aligned}&{\Vert \phi _x \Vert }_{H^{N_0}} \lesssim {\Vert \partial _{\alpha }f \Vert }_{H^{N_0}} + {\Vert \text{ Re }{\zeta }_{{\alpha }{\alpha }} \Vert }_{H^{N_0-1}} \lesssim {\Vert \widetilde{L}\Vert }_{X_{N_0}} . \end{aligned}$$
(10.20)

This takes care of the Sobolev component of the \(X_{N_0}\)-norm. To bound the weighted component we use \(\phi _x \circ \text{ Re }{\zeta }= \partial _{\alpha }f / \text{ Re }{\zeta }_{\alpha }\) in combination with (10.9), estimate (10.12), and (10.19), to get

$$\begin{aligned} {\Vert S \phi _x \Vert }_{H^\frac{N_0}{2}}&\lesssim {\Vert (S \phi _x) \circ \text{ Re }{\zeta }\Vert }_{H^\frac{N_0}{2}}\\&\lesssim \! {\left\| S (f_{\alpha }/\text{ Re }{\zeta }_{\alpha }) \right\| }_{H^\frac{N_0}{2}} \!+\! {\Vert \phi _x \Vert }_{H^{\frac{N_0}{2}+2}} {\Vert S \text{ Re }{\zeta }\!-\! \text{ Re }{\zeta }\Vert }_{H^\frac{N_0}{2}} \!\lesssim \! {\Vert \widetilde{L} \Vert }_{X_{N_0}}. \end{aligned}$$

This shows (10.5) provided we verify (10.18). Observe that

$$\begin{aligned} {\Vert \partial _{\alpha }f \Vert }_{W^{\frac{N_0}{2}+1,\infty }} \lesssim {\Vert \partial _x \phi \Vert }_{W^{\frac{N_0}{2}+1,\infty }} \lesssim {\varepsilon }_1 , \end{aligned}$$

directly from the a priori assumptions. To bound \( {\mathcal {H} } \partial _{\alpha }f\) instead, we use Lemma 11.8, the a priori assumptions and (4.35) to obtain

$$\begin{aligned} {\Vert {\mathcal {H} } \partial _{\alpha }f \Vert }_{W^{\frac{N_0}{2}+1,\infty }} \lesssim {\Vert f \Vert }_{W^{\frac{N_0}{2}+3,\infty }} \lesssim {\Vert \phi \Vert }_{W^{\frac{N_0}{2}+3,\infty }} \lesssim {\varepsilon }_1 . \end{aligned}$$

This shows (10.18) and concludes the proof of (10.5), hence of Proposition 2.4. \(\square \)