1 Introduction

We investigate the fluid flow past a rigid body \({\mathcal B}\) that moves through an infinite three-dimensional liquid reservoir with prescribed velocity

$$\begin{aligned} V(t,x)=\xi (t)+\eta \wedge (x-x_\mathrm {C}(t)) \end{aligned}$$

with respect to its center of mass \(x_\mathrm {C}\). Here \(t\in \mathbb {R}\) and \(x\in \mathbb {R}^3\) denote time and spatial variable, respectively, \(\xi :=\tfrac{{\mathrm d}}{{\mathrm d}t}x_\mathrm {C}\) is the translation velocity and \(\eta \) the angular velocity of \({\mathcal B}\) with respect to its center of mass. We only consider the case where the angular velocity \(\eta \) is constant, but the translation velocity \(\xi \) may depend on time. In a frame attached to the body, with origin at its center of mass \(x_\mathrm {C}\), the motion of an incompressible Navier–Stokes fluid around \({\mathcal B}\) that adheres to \({\mathcal B}\) at the boundary is described by the equations

$$\begin{aligned} \left\{ \begin{array}{r@{\quad }l} \rho \big (\partial _t u+ \eta \wedge u-\eta \wedge x \cdot \nabla u-\xi \cdot \nabla u+ u\cdot \nabla u\big ) = f +\mu \Delta u- \nabla \mathfrak {p}&{}\quad \text {in }\mathbb {R}\times \Omega , \\ {{\,\mathrm{div}\,}}u=0 &{}\quad \text {in }\mathbb {R}\times \Omega , \\ u=\xi +\eta \wedge x &{} \quad \text {on }\mathbb {R}\times \partial \Omega , \\ \lim _{{|x |}\rightarrow \infty } u(t,x) = 0 &{}\quad \text {for }t\in \mathbb {R}; \end{array}\right. \end{aligned}$$
(1.1)

see [12, Section 1]. Here \(\Omega :=\mathbb {R}^3\setminus {\overline{{\mathcal B}}}\) is the exterior domain surrounding \({\mathcal B}\), and \(\mathbb {R}\) represents the time axis. The functions \(u:\mathbb {R}\times \Omega \rightarrow \mathbb {R}^3\) and \(\mathfrak {p}:\mathbb {R}\times \Omega \rightarrow \mathbb {R}\) describe velocity and pressure fields of the fluid. The constants \(\rho >0\) and \(\mu >0\) denote density and viscosity, respectively. For the sake of generality, we additionally consider an external body force \(f:\mathbb {R}\times \Omega \rightarrow \mathbb {R}^3\).

In this paper, we investigate a configuration where the rigid body \({\mathcal B}\) translates periodically with some prescribed time period \({\mathcal T}>0\). More precisely, we assume the data

$$\begin{aligned} \xi (t+{\mathcal T})=\xi (t),\qquad f(t+{\mathcal T},x)=f(t,x) \end{aligned}$$

to be \({\mathcal T}\)-time-periodic As the main theorem we show existence of a solution \((u,\mathfrak {p})\) to (1.1) that shares this time periodicity.

We consider a prescribed motion of \({\mathcal B}\) where the axes of translation and rotation do not vary over time and are parallel. Without loss of generality, both are directed along the \(x_1\)-axis such that

$$\begin{aligned} \xi (t)=\alpha (t){{\,\mathrm{e}\,}}_1, \qquad \eta =\omega {{\,\mathrm{e}\,}}_1\end{aligned}$$

for some \({\mathcal T}\)-periodic function \(\alpha :\mathbb {R}\rightarrow \mathbb {R}\) and a constant \(\omega \in \mathbb {R}\). Note that, at least in the case where \(\xi \) is time-independent, this assumption can be made without loss of generality as long as \(\xi \cdot \eta \ne 0\) due to the Mozzi–Chasles theorem.

We assume that the mean translational velocity of the body over one time period is non-zero:

$$\begin{aligned} \lambda :=\frac{1}{{\mathcal T}}\int _0^{\mathcal T}\alpha (t)\,{\mathrm d}t\ne 0. \end{aligned}$$
(1.2)

The case of vanishing mean translational velocity shall not be treated here. Not only does the fluid flow exhibit different physical properties when (1.2) is not satisfied, due to the absence of a wake region in this case, but also the mathematical properties of the linearization of (1.1) differ significantly. If (1.2) is satisfied, the linearization of (1.1) is a time-periodic generalized Oseen system, for which we shall establish suitable \(\mathrm {L}^{q}\) estimates in order to show existence of a solution to (1.1). If (1.2) is not satisfied, the linearization of (1.1) is a time-periodic generalized Stokes system, for which similar estimates cannot be derived. In this case, problem (1.1) thus has to be approached in a different way, which has recently been done by Galdi [15].

Since the case \(\eta =0\) was treated in [18], we only consider the case \(\eta \ne 0\) in the following. Observe that then \(\eta \wedge x\cdot \nabla \) is a differential operator with unbounded coefficient. Therefore, the linearization of (1.1) cannot be treated as a lower-order perturbation of the time-periodic Oseen problem, even if \(\eta \) is “small”. In particular, as we will see below, the corresponding resolvent problem also requires an analysis in a different functional setting. This observation reflects the properties of the corresponding stationary problem (see [13, Chapter VIII]), which can be regarded as a special case of the time-periodic problem. In order to find a framework in which the time-periodic generalized Oseen problem is well posed, we employ the idea from [16, 17], where the steady-state problem corresponding to (1.1) was considered, and the rotation term \(\eta \wedge u-\eta \wedge x \cdot \nabla u\) was handled by a change of coordinates into a non-rotating frame. This procedure only yields suitable estimates for time-periodic solutions when the change of coordinates maintains the time periodicity of the involved functions. This is the case if the angular velocity \(\omega \) is an integer multiple of the angular frequency \(2\pi /{\mathcal T}\) of the time-periodic data. For simplicity, we assume

$$\begin{aligned} \omega =2\pi /{\mathcal T}. \end{aligned}$$
(1.3)

This condition means that during one period the rigid body completes one full revolution. In other words, the rotation and the time-periodic data, which may be regarded as two different sources of time-periodic forcing, have to be compatible.

The equations governing the fluid flow around a rigid body that performs a prescribed rigid motion have been studied by many researchers during the last decades. The first successful attempts of a rigorous mathematical treatment date back to the fundamental works of Oseen [44], Leray [36, 37] and Ladyžhenskaya [34, 35]. The study of time-periodic Navier–Stokes flows was proposed in a short note by Serrin [47], which induced Prodi [45], Yudovich [56] and Prouse [46] to initiate the examination in bounded domains. Through the years, this investigation has been continued and extended to other types of domains and fluid-flow configurations by several authors; see for example [5, 10, 11, 14, 18, 21,22,23, 28,29,30,31, 33, 38,39,43, 49, 51,52,55]. We refer to [19] for a more detailed overview. Concerning in particular time-periodic Navier–Stokes flows around rigid bodies, more specifically the three-dimensional exterior-domain configuration, we emphasize the fundamental work of Yamazaki [55], who introduced a setting of \(\mathrm {L}^{3,\infty }(\Omega )\) spaces to obtain time-periodic solutions in the case \(\xi =\eta =0\). The main estimates in [55] are based on well-known \(\mathrm {L}^{p}\)-\(\mathrm {L}^{q}\) estimates of the Stokes semigroup. If one replaces these estimates with the \(\mathrm {L}^{p}\)-\(\mathrm {L}^{q}\) estimates obtained by Shibata [48] for an Oseen semigroup with rotational effects, the approach in [55] also seems to yield existence of time-periodic solutions to (1.1) in an \(\mathrm {L}^{3,\infty }(\Omega )\) framework in the case of constant non-zero parameters \(\xi \ne 0, \eta \ne 0\). This analysis was recently carried out by Geissert, Hieber and Nguyen [23], who introduced a general semigroup-based approach to show existence of mild solutions to time-periodic problems. Using a recent result by Hishida [27], who established \(\mathrm {L}^{p}\)-\(\mathrm {L}^{q}\) estimates for an evolution operator corresponding to a linearization of the Navier–Stokes equations in the case of time-dependent \(\xi (t)\) and \(\eta (t)\), the approach of Yamazaki [55] even leads to time-periodic solutions in an \(\mathrm {L}^{3,\infty }(\Omega )\) framework for general time-periodic \(\xi (t)\) and \(\eta (t)\). In this general case, Galdi and Silvestre [21] already established the existence of time-periodic solutions in an \(\mathrm {L}^{2}(\Omega )\) setting via a Galerkin approach.

As the main novelty of the present paper, we establish existence of strong solutions to (1.1) in an \(\mathrm {L}^{q}(\Omega )\) setting for a certain range of exponents \(q\in (1,\infty )\). In this setting, better information on the spatial decay of the solutions can be derived compared to the \(\mathrm {L}^{3,\infty }(\Omega )\) and \(\mathrm {L}^{2}(\Omega )\) frameworks described above. Our approach is based on an analysis of the linearization of (1.1) and the associated resolvent problem

$$\begin{aligned} \left\{ \begin{array}{r@{\quad }l} isv+\omega ( {{\,\mathrm{e}\,}}_1\wedge v - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla v)- \Delta v- \lambda \partial _1 v+ \nabla p= F&{} \quad \text {in }\Omega , \\ {{\,\mathrm{div}\,}}v=0 &{}\quad \text {in }\Omega , \\ v=0 &{}\quad \text {on }\partial \Omega \end{array}\right. \end{aligned}$$
(1.4)

for suitable \(s\in \mathbb {R}\) and \(F\in \mathrm {L}^{q}(\Omega )^3\), \(1<q<\infty \). At first glance, it seems reasonable to regard (1.4) as a resolvent problem \((is-A)v=F\) for a closed operator A on the space of solenoidal vector fields in \(\mathrm {L}^{q}(\Omega )^3\). However, the spectral analysis in this setting, which was carried out by Farwig and Neustupa [7, 8], reveals that is, \(s\in \mathbb {R}\), belongs to the spectrum of A when \(s\in \omega \mathbb {Z}\), whereas well-posedness of the time-periodic problem requires invertibility of (1.4) for \(s\in \omega \mathbb {Z}\). Therefore, we propose to investigate the problem in homogeneous Sobolev spaces instead. Although it is merely possible to derive the non-classical resolvent estimate (2.4) in this setting (see Theorem 2.1 below), we are nevertheless able to conclude a suitable solution theory for the linearization of (1.1). To this end, we shall employ a framework of functions with absolutely convergent Fourier series. Finally, a fixed-point argument yields the existence of a solution to the nonlinear problem (1.1) when the data \(f\), \(\xi \) and \(\eta \) are “sufficiently small”.

2 Main Results

In virtue of (1.2) we may assume \(\lambda >0\) without loss of generality, and by (1.3) we have \(\omega =2\pi /{\mathcal T}>0\). To reformulate (1.1) in a non-dimensional way, we let the diameter \(d>0\) of \({\mathcal B}\) serve as a characteristic length scale. We introduce the Reynolds number \(\lambda ':=\lambda \rho d/\mu \), the Taylor number \(\omega ':=\omega \rho d^2/\mu \), and the non-dimensional time and spatial variables \(t'=\omega t\) and \(x'=x/d\). In particular, \(\Omega \) is transformed to \( \Omega ' :=\{x/d\ \vert \ x \in \Omega \}. \) We define \(\alpha '(t'):=\alpha (t)\rho d/\mu \) and the non-dimensional functions

$$\begin{aligned} u'(t',x'):=\frac{\rho d}{\mu }u(t,x), \quad \mathfrak {p}'(t',x'):=\frac{\rho d^2}{\mu ^2}\mathfrak {p}(t,x), \quad f'(t',x'):=\frac{\rho d^3}{\mu ^2}f(t,x), \end{aligned}$$

which are time-periodic with period \({\mathcal T}'=2\pi \) and can thus be identified with functions on the torus group \({\mathbb T}=\mathbb {R}/2\pi \mathbb {Z}\) with respect to time. Expressing (1.1) in these new quantities and omitting the primes, we obtain the non-dimensional formulation

$$\begin{aligned} \left\{ \begin{array}{r@{\quad }l} \omega (\partial _tu+ {{\,\mathrm{e}\,}}_1\wedge u - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla u ) - \alpha \partial _1 u+ u\cdot \nabla u= f + \Delta u- \nabla \mathfrak {p}&{}\quad \text {in }{\mathbb T}\times \Omega , \\ {{\,\mathrm{div}\,}}u=0 &{}\quad \text {in }{\mathbb T}\times \Omega , \\ u=\alpha {{\,\mathrm{e}\,}}_1+ \omega {{\,\mathrm{e}\,}}_1\wedge x &{}\quad \text {on }{\mathbb T}\times \partial \Omega , \\ \lim _{{|x |}\rightarrow \infty } u(t,x) = 0 &{}\quad \text {for }t\in {\mathbb T}. \end{array}\right. \end{aligned}$$
(2.1)

Our analysis of (2.1) is based on the study of the linear time-periodic problem

$$\begin{aligned} \left\{ \begin{array}{rl} \omega (\partial _t u + {{\,\mathrm{e}\,}}_1\wedge u - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla u )- \Delta u- \lambda \partial _1 u+ \nabla \mathfrak {p}= f &{}\quad \text {in }{\mathbb T}\times \Omega , \\ {{\,\mathrm{div}\,}}u=0 &{}\quad \text {in }{\mathbb T}\times \Omega , \\ u=0 &{}\quad \text {on }{\mathbb T}\times \partial \Omega , \end{array}\right. \end{aligned}$$
(2.2)

and of the corresponding resolvent problem

$$\begin{aligned} \left\{ \begin{array}{r@{\quad }l} \omega (ik v + {{\,\mathrm{e}\,}}_1\wedge v - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla v )- \Delta v- \lambda \partial _1 v+ \nabla p= F&{}\quad \text {in }\Omega , \\ {{\,\mathrm{div}\,}}v=0 &{}\quad \text {in }\Omega , \\ v=0 &{}\quad \text {on }\partial \Omega \end{array}\right. \end{aligned}$$
(2.3)

for \(k\in \mathbb {Z}\). For the latter we shall derive the following well-posedness result.

Theorem 2.1

Let \(\Omega \subset \mathbb {R}^3\) be an exterior domain of class \(\mathrm {C}^{3}\). Let \(q\in (1,2)\), \(k\in \mathbb {Z}\) and \(\lambda ,\,\omega , \,\theta ,\,B>0\) with \(\lambda ^2\le \theta \omega \le B\). For every \(F\in \mathrm {L}^{q}(\Omega )^3\) there exists a solution \((v,p)\in \mathrm {W}^{2,q}_{\mathrm {loc}}({\overline{\Omega }})^3\times \mathrm {W}^{1,q}_{\mathrm {loc}}({\overline{\Omega }})\) to (2.3) subject to the estimate

$$\begin{aligned} \begin{aligned}&\omega \Vert ik v + {{\,\mathrm{e}\,}}_1\wedge v - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla v \Vert _{q} +\Vert \nabla ^2v\Vert _{q} +\lambda \Vert \partial _1v\Vert _{q} \\&\qquad +\lambda ^{1/2}\Vert v\Vert _{s_1} +\lambda ^{1/4}\Vert \nabla v\Vert _{s_2} +\Vert \nabla p\Vert _{q} \le {C}_{1} \Vert F\Vert _{q} \end{aligned} \end{aligned}$$
(2.4)

for a constant \({C}_{1} ={C}_{1}(\Omega ,q,\lambda ,\omega )>0\) and \(s_1=2q/(2-q)\), \(s_2=4q/(4-q)\). Additionally, if \((w,\mathfrak {q})\) is another solution to (2.3) in the function class defined by the norms on the left-hand side of (2.4), then \(v=w\), and \(p-\mathfrak {q}\) is a constant. Moreover, if \(q\in (1,\frac{3}{2})\), then the constant \({C}_{1}\) can be chosen independently of \(\lambda \) and \(\omega \) such that \({C}_{1} ={C}_{1}(\Omega ,q,\theta ,B)\).

Note that for \(k=0\) we recover the well-known \(\mathrm {L}^{q}\) theory for the corresponding stationary problem; see [13, Theorem VIII.8.1].

In order to transfer estimate (2.4) to the time-periodic setting without losing information on the dependencies of the constant \({C}_{1}\), we work within spaces \(\mathrm {A}({\mathbb T};X)\) of absolutely convergent X-valued Fourier series for suitable Banach spaces X; see (3.1) below. We establish the following solution theory for the time-periodic problem (2.2).

Theorem 2.2

Let \(\Omega \subset \mathbb {R}^3\) be an exterior domain of class \(\mathrm {C}^{3}\). Let \(q\in (1,2)\) and \(\lambda ,\,\omega ,\,\theta ,\,B>0\) with \(\lambda ^2\le \theta \omega \le B\). For every \(f\in \mathrm {A}({\mathbb T};\mathrm {L}^{q}(\Omega ))^3\) there exists a solution \((u,\mathfrak {p})\) to (2.2) subject to the estimate

$$\begin{aligned}&\omega \Vert \partial _t u + {{\,\mathrm{e}\,}}_1\wedge u - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla u \Vert _{\mathrm {A}({\mathbb T};\mathrm {L}^{q}(\Omega ))} +\Vert \nabla ^2u\Vert _{\mathrm {A}({\mathbb T};\mathrm {L}^{q}(\Omega ))} +\lambda \Vert \partial _1u\Vert _{\mathrm {A}({\mathbb T};\mathrm {L}^{q}(\Omega ))} \nonumber \\&\quad +\lambda ^{1/2}\Vert u\Vert _{\mathrm {A}({\mathbb T};\mathrm {L}^{s_1}(\Omega ))} +\lambda ^{1/4}\Vert \nabla u\Vert _{\mathrm {A}({\mathbb T};\mathrm {L}^{s_2}(\Omega ))} +\Vert \nabla \mathfrak {p}\Vert _{\mathrm {A}({\mathbb T};\mathrm {L}^{q}(\Omega ))}\nonumber \\&\quad \le {C}_{1}\Vert f\Vert _{\mathrm {A}({\mathbb T};\mathrm {L}^{q}(\Omega ))} \end{aligned}$$
(2.5)

for the constant \({C}_{1}\) from Theorem 2.1, and \(s_1=2q/(2-q)\), \(s_2=4q/(4-q)\). Additionally, if \((w,\mathfrak {q})\) is another solution to (2.2) in the function class defined by the norms on the left-hand side of (2.5), then \(u=w\) and \(\mathfrak {p}=\mathfrak {q}+\mathfrak {q}_0\) for some (spatially constant) function \(\mathfrak {q}_0:{\mathbb T}\rightarrow \mathbb {R}\).

In Sect. 6, we finally prove the following existence result on solutions to the nonlinear system (2.1).

Theorem 2.3

Let \(\Omega \subset \mathbb {R}^3\) be an exterior domain of class \(\mathrm {C}^{3}\), and let \(q\in \big [\frac{12}{11},\frac{4}{3}\big ]\), \(\rho \in \big (\frac{3q-3}{q},1\big )\) and \(\theta >0\). Then there are constants \(\kappa >0\) and \(\lambda _0>0\) such that for all

$$\begin{aligned} \lambda \in (0,\lambda _0),\qquad \omega \in \big (\frac{\lambda ^2}{\theta },\infty \big ) \end{aligned}$$
(2.6)

there exists \(\varepsilon >0\) such that for all \(f\in \mathrm {A}({\mathbb T};\mathrm {L}^{q}(\Omega ))^3\) and \(\alpha \in \mathrm {A}({\mathbb T};\mathbb {R})\) with \(\frac{{\mathrm d}}{{\mathrm d}t}\alpha \in \mathrm {A}({\mathbb T};\mathbb {R})\) and

$$\begin{aligned} \lambda =\frac{1}{2\pi }\int _0^{2\pi }\alpha (t)\,{\mathrm d}t, \quad \omega \Vert \tfrac{{\mathrm d}}{{\mathrm d}t}\alpha \Vert _{\mathrm {A}({\mathbb T};\mathbb {R})} \le \kappa \lambda ^\rho , \quad \Vert \alpha -\lambda \Vert _{\mathrm {A}({\mathbb T};\mathbb {R})} +\Vert f\Vert _{\mathrm {A}({\mathbb T};\mathrm {L}^{q}(\Omega ))} \le \varepsilon \end{aligned}$$

there is a solution \((u,\mathfrak {p})\) to (2.1) with

$$\begin{aligned}&u\in \mathrm {A}({\mathbb T};\mathrm {L}^{2q/(2-q)}(\Omega ))^3, \quad \nabla u\in \mathrm {A}({\mathbb T};\mathrm {L}^{4q/(4-q)}(\Omega ))^{3\times 3}, \\&\quad \nabla ^2u\in \mathrm {A}({\mathbb T};\mathrm {L}^{q}(\Omega ))^{3\times 3\times 3},\\&\quad \partial _t u + {{\,\mathrm{e}\,}}_1\wedge u - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla u ,\ \partial _1u,\ \nabla \mathfrak {p}\in \mathrm {A}({\mathbb T};\mathrm {L}^{q}(\Omega ))^3. \end{aligned}$$

Remark 2.4

The lower bound \(\frac{\lambda ^2}{\theta }\le \omega \) on the angular velocity in (2.6) may seem strange in light of the underlying physics of the problem since from a physical point of view, the limit \(\omega \rightarrow 0\) towards the case of a non-rotating body seems uncritical. The lower bound on \(\omega \) in (2.6) is an artifact of the change of coordinates into the rotating frame of reference employed in the mathematical analysis of the problem, which leads to a priori estimates with constants exhibiting a singular behavior as \(\omega \rightarrow 0\). As a consequence, a lower bound on \(\omega \) is required in Theorem 2.3 to obtain existence of a solution via a fixed-point iteration. A similar observation was made in the investigation of a steady flow past a rotating and translating obstacle carried out in [6]. Therefore, it is not surprising to see the same effect appearing in the more general time-periodic case investigated here.

3 Preliminaries

We use capital letters to denote global constants, while constants in small letters are local to the respective proof. When we want to emphasize that a constant C depends on the quantities \(\alpha ,\beta ,\gamma ,\dots \), we write \(C(\alpha ,\beta ,\gamma ,\dots )\).

We denote points in \({\mathbb T}\times \mathbb {R}^3\) by (tx), where t and \(x=(x_1,x_2,x_3)\) are referred to as time and spatial variable. The symbol \(\Omega \) always denotes an exterior domain, that is, \(\Omega \subset \mathbb {R}^3\) is connected and the complement of a non-empty compact set. We always assume that the origin is not contained in \(\Omega \).

Inner and outer product of two vectors \(a,b\in \mathbb {R}^3\) are denoted by \(a\cdot b\) and \(a\wedge b\), respectively. For any radius \(R>0\) we set \(\mathrm {B}_R:=\bigl \{x\in \mathbb {R}^3\ \big \vert \ {|x |}<R\bigr \}\), \(\mathrm {B}^R:=\bigl \{x\in \mathbb {R}^3\ \big \vert \ {|x |}>R\bigr \}\), and for a domain \(D\subset \mathbb {R}^3\) we define \(D_R:=D\cap \mathrm {B}_R\) and \(D^R:=D\cap \mathrm {B}^R\).

For \(q\in [1,\infty ]\) and \(k\in \mathbb {N}_0\), the symbols \(\mathrm {L}^{q}(D)\) and \(\mathrm {W}^{k,q}(D)\) denote usual Lebesgue and Sobolev spaces with associated norms \(\Vert \cdot \Vert _{q}=\Vert \cdot \Vert _{q;D}\) and \(\Vert \cdot \Vert _{k,q}=\Vert \cdot \Vert _{k,q;D}\), respectively. Furthermore, \(\mathrm {W}^{1,q}_0(D)\) denotes the subset of functions in \(\mathrm {W}^{1,q}(D)\) with vanishing boundary trace, and \(\mathrm {W}^{-1,q}(D)\) (with norm \(\Vert \cdot \Vert _{-1,q;D}\)) is the dual space of \(\mathrm {W}^{1,q'}_0(D)\) where \(1/q+1/q'=1\) with the usual convention \(1/\infty :=0\). Moreover, \(\mathrm {L}^{2}_{\sigma }(D)\) denotes the set of solenoidal vector fields in \(\mathrm {L}^{2}(D)^3\), that is,

$$\begin{aligned} \mathrm {L}^{2}_{\sigma }(D):=\overline{\bigl \{\varphi \in \mathrm {C}^{\infty }_0(D)^3\ \big \vert \ {{\,\mathrm{div}\,}}\varphi =0\bigr \}}^{\Vert \cdot \Vert _{2}}, \end{aligned}$$

and \({\mathcal P}_\mathrm {H}\) is the corresponding Helmholtz projection that maps \(\mathrm {L}^{2}(D)^3\) onto \(\mathrm {L}^{2}_{\sigma }(D)\).

We always identify \(2\pi \)-periodic functions with functions on the torus group \({\mathbb T}:=\mathbb {R}/2\pi \mathbb {Z}\), which is usually represented by the set \([0,2\pi )\). We consider \({\mathbb T}\) and \(G:={\mathbb T}\times \mathbb {R}^3\) as locally compact abelian groups. The (normalized) Haar measure on \({\mathbb T}\) is given by

$$\begin{aligned} \forall f\in \mathrm {C}^{}({\mathbb T}):\quad \int _{\mathbb T}f\,{\mathrm d}t:=\frac{1}{2\pi }\int _0^{2\pi } f(t)\,{\mathrm d}t, \end{aligned}$$

and \(G\) is equipped with the corresponding product measure. Recall that the dual group of \({\mathbb T}\) can be identified with \(\widehat{{\mathbb T}}=\mathbb {Z}\) and that of \(G\) with \(\widehat{G}:=\mathbb {Z}\times \mathbb {R}^3\).

For \(H={\mathbb T}\) or \(H=G\), the space \(\mathscr {S}(H)\) is the Schwartz–Bruhat space of generalized Schwartz functions on H, and \(\mathscr {S^\prime }(H)\) denotes the corresponding dual space of tempered distributions; see [1, 4] for precise definitions. The Fourier transform on \({\mathbb T}\) and \(G\) and the respective inverses are given by

$$\begin{aligned}&\mathscr {F}_{\mathbb T}:\mathscr {S}({\mathbb T})\rightarrow \mathscr {S}(\mathbb {Z}),&\mathscr {F}_{\mathbb T}[u](k):=\int _{\mathbb T}u(t)\,{{\,\mathrm{e}\,}}^{-ik t}\,{\mathrm d}t,\\&\mathscr {F}^{-1}_{\mathbb T}:\mathscr {S}(\mathbb {Z})\rightarrow \mathscr {S}({\mathbb T}),&\mathscr {F}^{-1}_{\mathbb T}[w](t):=\sum _{k\in \mathbb {Z}} w(k)\,{{\,\mathrm{e}\,}}^{ik t}, \end{aligned}$$

and

$$\begin{aligned}&\mathscr {F}_G:\mathscr {S}(G)\rightarrow \mathscr {S}(\widehat{G}),&\mathscr {F}_G[u](k,\xi ):=\int _{\mathbb T}\int _{\mathbb {R}^n} u(t,x)\,{{\,\mathrm {e}\,}}^{-ix\cdot \xi -ik t}\,{\mathrm d}x{\mathrm d}t,\\&\mathscr {F}^{-1}_G:\mathscr {S}(\widehat{G})\rightarrow \mathscr {S}(G),&\mathscr {F}^{-1}_G[w](t,x):=\sum _{k\in \mathbb {Z}}\,\int _{\mathbb {R}^n} w(k,\xi )\,{{\,\mathrm {e}\,}}^{ix\cdot \xi +ik t}\,{\mathrm d}\xi , \end{aligned}$$

provided the Lebesgue measure \({\mathrm d}\xi \) is correctly normalized. By duality, \(\mathscr {F}_{\mathbb T}\) and \(\mathscr {F}_G\) extend to homeomorphisms \(\mathscr {F}_{\mathbb T}:\mathscr {S^\prime }({\mathbb T})\rightarrow \mathscr {S^\prime }(\mathbb {Z})\) and \(\mathscr {F}_G:\mathscr {S^\prime }(G)\rightarrow \mathscr {S^\prime }(\widehat{G})\), respectively.

Furthermore, we introduce the Sobolev space

$$\begin{aligned}&\mathrm {W}^{1,2,q}({\mathbb T}\times D) :=\overline{\mathrm {C}^{\infty }_0({\mathbb T}\times \overline{D})}^{\Vert \cdot \Vert _{1,2,q}},\quad \Vert f\Vert _{1,2,q}:=\bigg ( \Vert \partial _t f\Vert _{q}^q + \sum _{k=0}^2 \Vert \nabla ^k f\Vert _{q}^q \bigg )^{\frac{1}{q}}, \end{aligned}$$

where \(\mathrm {C}^{\infty }_0({\mathbb T}\times \overline{D})\) denotes the space of smooth functions of compact support on \({\mathbb T}\times \overline{D}\) .

Let X be a Banach space. We introduce the projections \({\mathcal P}\) and \({\mathcal P}_\bot \) by

$$\begin{aligned} {\mathcal P}u:=\int _{{\mathbb T}}u(t)\,{\mathrm d}t, \qquad {\mathcal P}_\bot :={{\,\mathrm{Id}\,}}-{\mathcal P}\end{aligned}$$

for \(u\in \mathrm {L}^{1}({\mathbb T};X)\). Note that \({\mathcal P}u\in X\) is time-independent, and we have the decomposition \(u={\mathcal P}u+{\mathcal P}_\bot u\) into the steady-state part \({\mathcal P}u\) and the purely periodic part \({\mathcal P}_\bot u\) of \(u\).

Our analysis of the time-periodic problems (2.1) and (2.2) will be carried out within spaces of functions with absolutely convergent Fourier series defined by

$$\begin{aligned} \begin{aligned} \mathrm {A}({\mathbb T};X)&:=\biggl \{f:{\mathbb T}\rightarrow X\ \bigg \vert \ f(t)=\sum _{k\in \mathbb {Z}}f_k {{\,\mathrm{e}\,}}^{ikt}, \ f_k\in X, \ \sum _{k\in \mathbb {Z}}\Vert f_k\Vert _{X}<\infty \biggr \}, \\ \Vert f\Vert _{\mathrm {A}({\mathbb T};X)}&:=\sum _{k\in \mathbb {Z}}\Vert f_k\Vert _{X}. \end{aligned} \end{aligned}$$
(3.1)

Observe that \(\mathrm {A}({\mathbb T};X)\) is the Banach space that coincides with \(\mathscr {F}^{-1}_{\mathbb T}\big [\ell ^{1}(\mathbb {Z};X)\big ]\), which embeds into the X-valued continuous functions on \({\mathbb T}\). It is well known that the scalar-valued space \(\mathrm {A}({\mathbb T};\mathbb {R})\) is an algebra with respect to pointwise multiplication, the so-called Wiener algebra. One can exploit this property to derive estimates in the X-valued case. For example, one readily shows the following correspondences of Hölder’s inequality and interpolation inequalities.

Proposition 3.1

Let \(D\subset \mathbb {R}^n\), \(n\in \mathbb {N}\), be an open set and \(p,q,r\in [1,\infty ]\) such that \(1/p+1/q=1/r\). Moreover, let \(f\in \mathrm {A}({\mathbb T};\mathrm {L}^{p}(D))\) and \(g\in \mathrm {A}({\mathbb T};\mathrm {L}^{q}(D))\). Then \(fg\in \mathrm {A}({\mathbb T};\mathrm {L}^{r}(D))\) and

$$\begin{aligned} \Vert fg\Vert _{\mathrm {A}({\mathbb T};\mathrm {L}^{r}(D))} \le \Vert f\Vert _{\mathrm {A}({\mathbb T};\mathrm {L}^{p}(D))} \Vert g\Vert _{\mathrm {A}({\mathbb T};\mathrm {L}^{q}(D))}. \end{aligned}$$
(3.2)

Proof

By assumption we have \(f=\mathscr {F}^{-1}_{\mathbb T}[(f_k)]\) and \(g=\mathscr {F}^{-1}_{\mathbb T}[(g_k)]\) for elements \((f_k)\in \ell ^{1}(\mathbb {Z};\mathrm {L}^{p}(D))\) and \((g_k)\in \ell ^{1}(\mathbb {Z};\mathrm {L}^{q}(D))\). Then \(fg=\mathscr {F}^{-1}_{\mathbb T}\big [(f_k)*_\mathbb {Z}(g_k)\big ]\) and

$$\begin{aligned} \Vert fg\Vert _{\mathrm {A}({\mathbb T};\mathrm {L}^{r}(D))}&=\sum _{k\in \mathbb {Z}} \Bigl \Vert \sum _{\ell \in \mathbb {Z}} f_\ell g_{k-\ell }\Bigr \Vert _{\mathrm {L}^{r}(D)} \le \sum _{k\in \mathbb {Z}}\sum _{\ell \in \mathbb {Z}}\Vert f_\ell g_{k-\ell }\Vert _{\mathrm {L}^{r}(D)}\\&\le \sum _{k\in \mathbb {Z}}\sum _{\ell \in \mathbb {Z}} \Vert f_\ell \Vert _{\mathrm {L}^{p}(D)}\Vert g_{k-\ell }\Vert _{\mathrm {L}^{q}(D)} =\Vert f\Vert _{\mathrm {A}({\mathbb T};\mathrm {L}^{p}(D))} \Vert g\Vert _{\mathrm {A}({\mathbb T};\mathrm {L}^{q}(D))}, \end{aligned}$$

where the last estimate is due to Hölder’s inequality. \(\square \)

Proposition 3.2

Let \(D\subset \mathbb {R}^n\), \(n\in \mathbb {N}\), be an open set and \(p,q,r\in [1,\infty ]\) such that \((1-\theta )/p+\theta /q=1/r\) for some \(\theta \in [0,1]\), and let \(f\in \mathrm {A}({\mathbb T};\mathrm {L}^{p}(D))\cap \mathrm {A}({\mathbb T};\mathrm {L}^{q}(D))\). Then \(f\in \mathrm {A}({\mathbb T};\mathrm {L}^{r}(D))\) and

$$\begin{aligned} \Vert f\Vert _{\mathrm {A}({\mathbb T};\mathrm {L}^{r}(D))} \le \Vert f\Vert _{\mathrm {A}({\mathbb T};\mathrm {L}^{p}(D))}^{1-\theta } \Vert f\Vert _{\mathrm {A}({\mathbb T};\mathrm {L}^{q}(D))}^\theta . \end{aligned}$$
(3.3)

Proof

We have \(f=\mathscr {F}^{-1}_{\mathbb T}[(f_k)]\) for an element \((f_k)\in \ell ^{1}(\mathbb {Z};\mathrm {L}^{p}(D)\cap \mathrm {L}^{q}(D))\). The classical interpolation inequality for Lebesgue spaces yields

$$\begin{aligned} \begin{aligned} \Vert f\Vert _{\mathrm {A}({\mathbb T};\mathrm {L}^{r}(D))} =\sum _{k\in \mathbb {Z}} \Vert f_k\Vert _{\mathrm {L}^{r}(D)}&\le \sum _{k\in \mathbb {Z}}\Vert f_k\Vert _{\mathrm {L}^{p}(D)}^{1-\theta } \Vert f_k\Vert _{\mathrm {L}^{q}(D)}^\theta \\&\le \Vert f\Vert _{\mathrm {A}({\mathbb T};\mathrm {L}^{p}(D))}^{1-\theta } \Vert f\Vert _{\mathrm {A}({\mathbb T};\mathrm {L}^{q}(D))}^\theta , \end{aligned} \end{aligned}$$

where the last estimate follows from Hölder’s inequality on \(\mathbb {Z}\). \(\square \)

4 Embedding Theorem

This section deals with embedding properties of Sobolev spaces of time-periodic functions. The embedding theorem below is a refinement of [18, Theorem 4.1] adapted to the time-scaling employed in (2.1). Clearly, embeddings of the steady-state part \({\mathcal P}u\) are independent of the actual period. Therefore, we only consider the case of purely periodic functions. For the sake of generality, we establish the following theorem in arbitrary dimension \(n\ge 2\).

Theorem 4.1

Let \(n\ge 2\), \(\omega >0\) and \(q\in (1,\infty )\). For \(\alpha \in [0,2]\) with \(\alpha q<2\) and \((2-\alpha ) q <n\) let

$$\begin{aligned} r_0:=\frac{2q}{2-\alpha q}, \qquad p_0:=\frac{nq}{n-(2-\alpha )q}, \end{aligned}$$

and for \(\beta \in [0,1]\) with \(\beta q<2\) and \((1-\beta ) q<n\) let

$$\begin{aligned} r_1:=\frac{2q}{2-\beta q}, \qquad p_1:=\frac{nq}{n-(1-\beta )q}. \end{aligned}$$

Then the inequality

$$\begin{aligned} \omega ^{\alpha /2}\Vert u\Vert _{\mathrm {L}^{r_0}({\mathbb T};\mathrm {L}^{p_0}({\mathbb {R}^n}))} +\omega ^{\beta /2}\Vert \nabla u\Vert _{\mathrm {L}^{r_1}({\mathbb T};\mathrm {L}^{p_1}({\mathbb {R}^n}))}&\le {C}_{2} \big ( \omega \Vert \partial _tu\Vert _q +\Vert \nabla ^2u\Vert _q \big ) \end{aligned}$$
(4.1)

holds for all \(u\in {\mathcal P}_\bot \mathrm {W}^{1,2,q}({\mathbb T}\times {\mathbb {R}^n})\) and a constant \({C}_{2}={C}_{2}(n,q,\alpha ,\beta )>0\).

Proof

Since the proof is analogue to [18, Proof of Theorem 4.1], we merely give a brief sketch here. Without restriction we may assume \(u\in \mathscr {S}(G)\). Due to the assumption \(u={\mathcal P}_\bot u\), we have \(\mathscr {F}_G[u]=(1-\delta _\mathbb {Z})\mathscr {F}_G[u]\), where \(\delta _\mathbb {Z}\) is the delta distribution on \(\mathbb {Z}\). Utilizing the Fourier transform, we thus derive the identity

$$\begin{aligned} \begin{aligned} u&=\mathscr {F}^{-1}_G\bigg [ \frac{1-\delta _\mathbb {Z}(k)}{{|\xi |}^2+i\omega k} \mathscr {F}_G\big [\omega \partial _tu-\Delta u\big ] \bigg ] \\&=\omega ^{-\alpha /2} \mathscr {F}^{-1}_{{\mathbb {R}^n}}\big [ {|\xi |}^{\alpha -2} \big ] *_{{\mathbb {R}^n}} \mathscr {F}^{-1}_{{\mathbb T}}\big [ (1-\delta _\mathbb {Z}){|k |}^{-\alpha /2} \big ] *_{{\mathbb T}} F, \end{aligned} \end{aligned}$$
(4.2)

where

$$\begin{aligned} F:=\mathscr {F}^{-1}_G\bigg [M_\omega (k,\xi )\mathscr {F}_G\big [\omega \partial _tu-\Delta u\big ]\bigg ], \quad M_\omega (k,\xi ) :=\frac{{|\omega k |}^{\alpha /2}{|\xi |}^{2-\alpha }(1-\delta _\mathbb {Z}(k))}{{|\xi |}^2+i\omega k}. \end{aligned}$$

Employing the so-called transference principle for Fourier multipliers (see [3, 4]) together with the Marcinkiewicz multiplier theorem, one readily verifies that \(M_\omega \) is an \(\mathrm {L}^{q}(G)\) multiplier for any \(q\in (1,\infty )\) such that

$$\begin{aligned} \Vert F\Vert _{q}\le {c}_{0}\Vert \omega \partial _tu-\Delta u\Vert _{q} \le {c}_{0}\big (\omega \Vert \partial _tu\Vert _q+\Vert \nabla ^2u\Vert _{q}\big ) \end{aligned}$$

with \({c}_{0}\) independent of \(\omega \). Moreover, when we choose \([-\pi ,\pi )\) as a realization of \({\mathbb T}\), we obtain

$$\begin{aligned} \gamma _\alpha (t):=\mathscr {F}^{-1}_{{\mathbb T}}\big [ (1-\delta _\mathbb {Z}){|k |}^{-\alpha /2} \big ](t) ={c}_{1}t^{-1+\alpha /2}+h(t), \end{aligned}$$

for some \(h\in \mathrm {C}^{\infty }({\mathbb T})\); see for example [24, Example 3.1.19]. In particular, this yields \(\gamma _\alpha \in \mathrm {L}^{\frac{1}{1-\alpha /2},\infty }({\mathbb T})\), so that Young’s inequality implies that the mapping \(\varphi \mapsto \gamma _\alpha *\varphi \) extends to a bounded operator \(\mathrm {L}^{q}({\mathbb T})\rightarrow \mathrm {L}^{r_0}({\mathbb T})\). Moreover, it is well known that the mapping \(\varphi \mapsto \mathscr {F}^{-1}_{{\mathbb {R}^n}}\big [{|\xi |}^{\alpha -2}\big ]*\varphi \) extends to a bounded operator \(\mathrm {L}^{q}({\mathbb {R}^n})\rightarrow \mathrm {L}^{p_0}({\mathbb {R}^n})\); see [25, Theorem 6.1.13]. Recalling (4.2), we thus have

$$\begin{aligned}&\omega ^{\alpha /2}\Vert u\Vert _{\mathrm {L}^{r_0}({\mathbb T};\mathrm {L}^{p_0}({\mathbb {R}^n}))} =\bigg (\int _{{\mathbb T}}\Bigl \Vert \mathscr {F}^{-1}_{{\mathbb {R}^n}}\big [ {|\xi |}^{\alpha -2} \big ] *_{{\mathbb {R}^n}} \gamma _\alpha *_{{\mathbb T}} F(t,\cdot )\Bigr \Vert _{p_0}^{r_0}\,{\mathrm d}t\bigg )^{\frac{1}{r_0}} \\&\qquad \le {c}_{2}\bigg (\int _{{\mathbb T}}\Vert \gamma _\alpha *_{{\mathbb T}} F(t,\cdot )\Vert _{q}^{r_0}\,{\mathrm d}t\bigg )^{\frac{1}{r_0}} \le {c}_{3}\bigg (\int _{{\mathbb {R}^n}}\Vert \gamma _\alpha *_{{\mathbb T}} F(\cdot ,x)\Vert _{r_0}^{q}\,{\mathrm d}x\bigg )^{\frac{1}{q}} \\&\qquad \le {c}_{4}\Vert F\Vert _{q} \le {c}_{5}\big (\omega \Vert \partial _tu\Vert _{q}+\Vert \nabla ^2u\Vert _{q}\big ), \end{aligned}$$

where Minkowski’s integral inequality is used in the second estimate. This is the asserted inequality for \(u\). The estimate of \(\nabla u\) follows in the same way. \(\square \)

Remark 4.2

Note that the term on the right-hand side of (4.1) defines a norm equivalent to \(\Vert \cdot \Vert _{1,2,q}\) on \({\mathcal P}_\bot \mathrm {W}^{1,2,q}({\mathbb T}\times \Omega )\) due to Poincaré’s inequality on \({\mathbb T}\).

Remark 4.3

Theorem 4.1 can be generalized to the setting of an exterior domain \(\Omega \subset {\mathbb {R}^n}\) by means of Sobolev extensions. However, to maintain estimate (4.1), one has to construct a specific extension operator that respects the homogeneous second-order Sobolev norm. To this end, one can make use of results from [2].

5 Linear Theory

This section is dedicated to the investigation of the resolvent problem (2.3) and the linear time-periodic problem (2.2). After having shown Theorem 2.1, we establish Theorem 2.2 as an immediate consequence hereof.

5.1 The Whole Space

To study the problems (2.2) and (2.3) in an exterior domain, we first consider the case \(\Omega =\mathbb {R}^3\). In this whole-space setting one can namely change coordinates back to the non-rotating inertial frame and thereby reduce the study of (2.2) to an investigation of the time-periodic Oseen problem without rotation terms, which was analyzed in [18, 32]. In this section, we set

$$\begin{aligned} s_1:=\frac{2q}{2-q}, \qquad s_2:=\frac{4q}{4-q}, \qquad s_3:=\frac{8q}{8-q}. \end{aligned}$$

for appropriately fixed q.

Theorem 5.1

Let \(q\in (1,2)\) and \(\lambda ,\,\omega ,\,\theta >0\) with \(\lambda ^2\le \theta \omega \). For every \(f\in \mathrm {L}^{q}({\mathbb T}\times \mathbb {R}^3)^3\) there exists a solution \((u,\mathfrak {p})\in \mathscr {S^\prime }({\mathbb T}\times \mathbb {R}^3)^{3+1}\) to

$$\begin{aligned} \left\{ \begin{aligned} \omega \partial _t u- \Delta u- \lambda \partial _1 u+ \nabla \mathfrak {p}&= f&\text {in }{\mathbb T}\times \mathbb {R}^3, \\ {{\,\mathrm{div}\,}}u&=0&\text {in }{\mathbb T}\times \mathbb {R}^3, \end{aligned}\right. \end{aligned}$$
(5.1)

with \(\partial _tu,\nabla ^2u,\,\nabla \mathfrak {p}\in \mathrm {L}^{q}({\mathbb T}\times \mathbb {R}^3)\). Moreover, there exist constants \({C}_{3}={C}_{3}(q)>0\) and \({C}_{4}={C}_{4}(q,\theta )>0\) such that

$$\begin{aligned}&\Vert \nabla ^2{\mathcal P}u\Vert _{q} +\lambda \Vert \partial _1{\mathcal P}u\Vert _{q} +\lambda ^{1/2}\Vert {\mathcal P}u\Vert _{s_1} +\lambda ^{1/4}\Vert \nabla {\mathcal P}u\Vert _{s_2} +\Vert \nabla {\mathcal P}\mathfrak {p}\Vert _{q} \le {C}_{3}\Vert {\mathcal P}f\Vert _{q}, \end{aligned}$$
(5.2)
$$\begin{aligned}&\omega \Vert \partial _t{\mathcal P}_\bot u\Vert _{q} +\Vert \nabla ^2{\mathcal P}_\bot u\Vert _{q} +\lambda \Vert \partial _1{\mathcal P}_\bot u\Vert _{q} +\Vert \nabla {\mathcal P}_\bot \mathfrak {p}\Vert _{q} \le {C}_{4}\Vert {\mathcal P}_\bot f\Vert _{q}. \end{aligned}$$
(5.3)

Additionally, if \((w,\mathfrak {q})\in \mathscr {S^\prime }({\mathbb T}\times \mathbb {R}^3)^{3+1}\) is another solution to (5.1), then \({\mathcal P}_\bot u={\mathcal P}_\bot w\), and \({\mathcal P}u-{\mathcal P}w\) is a polynomial in each component, and \(\mathfrak {p}-\mathfrak {q}=\mathfrak {p}_0\), where \(\mathfrak {p}_0(t,\cdot )\) is a polynomial for each \(t\in {\mathbb T}\).

Proof

We decompose (5.1) into two problems by splitting and . For the steady-state part \((u_{\mathrm {s}},\mathfrak {p}_{\mathrm {s}})\) we obtain the system

$$\begin{aligned} \left\{ \begin{aligned} - \Delta u_{\mathrm {s}}- \lambda \partial _1 u_{\mathrm {s}}+ \nabla \mathfrak {p}_{\mathrm {s}}&= {\mathcal P}f&\text {in }\mathbb {R}^3, \\ {{\,\mathrm{div}\,}}u_{\mathrm {s}}&=0&\text {in }\mathbb {R}^3, \end{aligned}\right. \end{aligned}$$

which is the classical steady-state Oseen problem. The existence of a time-independent solution \((u_{\mathrm {s}},\mathfrak {p}_{\mathrm {s}})\) satisfying estimate (5.2) is well known; see for example [13, Theorem VII.4.1]. The remaining purely periodic part \((u_{\mathrm {p}},\mathfrak {p}_{\mathrm {p}})\) must solve (5.1), but with purely periodic right-hand side \({\mathcal P}_\bot f\). We define

$$\begin{aligned} \begin{aligned} U(t,x)&:=\,u_{\mathrm {p}}(t, \omega ^{-1/2} x), \\ \mathfrak {P}(t,x)&:=\omega ^{-1/2}\,\mathfrak {p}_{\mathrm {p}}(t, \omega ^{-1/2}x), \\ F(t,x)&:=\omega ^{-1}\,{\mathcal P}_\bot f(t, \omega ^{-1/2}x), \end{aligned} \end{aligned}$$

which leads to the system

$$\begin{aligned} \left\{ \begin{aligned} \partial _t U- \Delta U- \widetilde{\lambda } \partial _1 U+ \nabla \mathfrak {P}&= F&\text {in }{\mathbb T}\times \mathbb {R}^3, \\ {{\,\mathrm{div}\,}}U&=0&\text {in }{\mathbb T}\times \mathbb {R}^3, \end{aligned}\right. \end{aligned}$$

where \(\widetilde{\lambda }=\lambda \omega ^{-1/2}\). From [32, Theorem 2.1] we conclude the existence of a unique solution \((U,\mathfrak {P})\) that satisfies the estimate

$$\begin{aligned} \Vert U\Vert _{1,2,q}+\Vert \nabla \mathfrak {P}\Vert _{q} \le {c}_{0}\Vert F\Vert _{q}, \end{aligned}$$

where \({c}_{0}\) is a polynomial in \(\widetilde{\lambda }\) and can thus be bounded uniformly in \(\widetilde{\lambda }\in (0,\sqrt{\theta }]\). Estimate (5.3) with the asserted dependency of the constant \({C}_{4}\) follows after reversing the applied scaling.

The uniqueness statement is readily shown by means of the Fourier transform on \(G={\mathbb T}\times \mathbb {R}^3\). We consider (5.1) with \(f=0\) and apply the divergence operator to (5.1)\(_{1}\). This yields \(\Delta \mathfrak {p}=0\) and thus \({|\xi |}^2\mathscr {F}_{\mathbb {R}^3}[\mathfrak {p}(t,\cdot )]=0\) for all \(t\in {\mathbb T}\). Therefore, we obtain \({{\,\mathrm{supp}\,}}\mathscr {F}_{\mathbb {R}^3}[\mathfrak {p}(t,\cdot )] \subset \{0\}\), so that \(\mathfrak {p}(t,\cdot )\) is a polynomial for all \(t\in {\mathbb T}\). Next we apply the Fourier transform to (5.1)\(_{1}\) to deduce \((i\omega k+{|\xi |}^2-i\xi _1)\mathscr {F}_G[u]+i\xi \mathscr {F}_G[\mathfrak {p}]=0\). Multiplying with the symbol of the Helmholtz projection \(\mathrm {I}-\xi \otimes \xi /{|\xi |}^2\) and utilizing \({{\,\mathrm{div}\,}}u=0\), we obtain \((i\omega k+{|\xi |}^2-i\xi _1)\mathscr {F}_G[u]=0\), which yields \({{\,\mathrm{supp}\,}}\mathscr {F}_G[u]\subset \{(0,0)\}\). Since \({{\mathcal P}_\bot u}=\mathscr {F}^{-1}_G\big [(1-\delta _\mathbb {Z})\mathscr {F}_G[u]\big ]\), it follows that \({\mathcal P}_\bot u=0\), and that each component of \({\mathcal P}u\) is a polynomial. This completes the proof. \(\square \)

Remark 5.2

In the setting of Theorem 5.1 we can write the estimate for the steady-state part \((u_{\mathrm {s}},\mathfrak {p}_{\mathrm {s}})=({\mathcal P}u,{\mathcal P}\mathfrak {p})\) and the purely periodic part \((u_{\mathrm {p}},\mathfrak {p}_{\mathrm {p}})=({\mathcal P}_\bot u,{\mathcal P}_\bot \mathfrak {p})\) in a more condensed way: From the embeddings established in Theorem 4.1 we deduce

$$\begin{aligned}&\omega ^{1/4}\Vert u_{\mathrm {p}}\Vert _{\mathrm {L}^{s_2}({\mathbb T};\mathrm {L}^{s_1}(\mathbb {R}^3))} +\omega ^{1/8}\Vert \nabla u_{\mathrm {p}}\Vert _{\mathrm {L}^{s_3}({\mathbb T};\mathrm {L}^{s_2}(\mathbb {R}^3))}\\&\quad \le {C}_{5}\big (\omega \Vert \partial _tu_{\mathrm {p}}\Vert _{\mathrm {L}^{q}({\mathbb T}\times \mathbb {R}^3)} +\Vert u_{\mathrm {p}}\Vert _{\mathrm {L}^{q}({\mathbb T}\times \mathbb {R}^3)}\big ). \end{aligned}$$

Recalling Remark 4.2, we see that (5.2) and (5.3) can be formulated as

$$\begin{aligned} \omega \Vert \partial _tu\Vert _{q} +\Vert \nabla ^2u\Vert _{q} +\lambda \Vert \partial _1u\Vert _{q} +\lambda ^{1/2}\Vert u\Vert _{\mathrm {L}^{s_2}({\mathbb T};\mathrm {L}^{s_1}(\mathbb {R}^3))}\nonumber \\ +\lambda ^{1/4}\Vert \nabla u\Vert _{\mathrm {L}^{s_3}({\mathbb T};\mathrm {L}^{s_2}(\mathbb {R}^3))} +\Vert \nabla \mathfrak {p}\Vert _{q} \le {C}_{6} \Vert f\Vert _{q} \end{aligned}$$
(5.4)

for a constant \(C_{6}=C_{6}(q,\theta )\) as long as \(\lambda ^2\le \theta \omega \).

With Theorem 5.1 we now solve the linear problem (2.2) for \(\Omega =\mathbb {R}^3\) and \(f\in \mathrm {L}^{q}({\mathbb T}\times \mathbb {R}^3)^3\).

Theorem 5.3

Let \(q\in (1,2)\) and \(\lambda ,\,\omega , \,\theta >0\) with \(\lambda ^2\le \theta \omega \). For every \(f\in \mathrm {L}^{q}({\mathbb T}\times \mathbb {R}^3)^3\) there exists a solution \((u,\mathfrak {p})\in \mathscr {S^\prime }({\mathbb T}\times \mathbb {R}^3)^{3+1}\) to

$$\begin{aligned} \left\{ \begin{aligned} \omega (\partial _t u+{{\,\mathrm{e}\,}}_1\wedge u-{{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla u) -\Delta u- \lambda \partial _1 u+ \nabla \mathfrak {p}&= f&\text {in }{\mathbb T}\times \mathbb {R}^3, \\ {{\,\mathrm{div}\,}}u&=0&\text {in }{\mathbb T}\times \mathbb {R}^3, \end{aligned}\right. \end{aligned}$$
(5.5)

with \(\nabla ^2u,\,\partial _1u,\,\nabla \mathfrak {p}\in \mathrm {L}^{q}({\mathbb T}\times \mathbb {R}^3)\). Moreover, there exists a constant \({C}_{7} ={C}_{7}(q,\theta )>0\) such that

$$\begin{aligned} \begin{aligned}&\omega \Vert \partial _t u + {{\,\mathrm{e}\,}}_1\wedge u - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla u \Vert _{\mathrm {L}^{q}({\mathbb T}\times \mathbb {R}^3)} +\Vert \nabla ^2u\Vert _{\mathrm {L}^{q}({\mathbb T}\times \mathbb {R}^3)} +\lambda \Vert \partial _1u\Vert _{\mathrm {L}^{q}({\mathbb T}\times \mathbb {R}^3)} \\&\quad +\lambda ^{1/2}\Vert u\Vert _{\mathrm {L}^{s_2}({\mathbb T};\mathrm {L}^{s_1}(\mathbb {R}^3))} +\lambda ^{1/4}\Vert \nabla u\Vert _{\mathrm {L}^{s_3}({\mathbb T};\mathrm {L}^{s_2}(\mathbb {R}^3))} +\Vert \nabla \mathfrak {p}\Vert _{\mathrm {L}^{q}({\mathbb T}\times \mathbb {R}^3)}\\&\qquad \le {C}_{7} \Vert f\Vert _{\mathrm {L}^{q}({\mathbb T}\times \mathbb {R}^3)}. \end{aligned} \end{aligned}$$
(5.6)

Additionally, if \((w,\mathfrak {q})\in \mathscr {S^\prime }({\mathbb T}\times \mathbb {R}^3)^{3+1}\) is another solution to (5.5) with \(w\in \mathrm {L}^{r}({\mathbb T}\times \mathbb {R}^3)\) for some \(r\in [1,\infty )\), then \(u=w\), and \(\mathfrak {p}-\mathfrak {q}=\mathfrak {q}_0\) for some spatially constant function \(\mathfrak {q}_0:{\mathbb T}\rightarrow \mathbb {R}\).

Proof

Let

$$\begin{aligned} Q(t):=\begin{pmatrix} 1 &{}\quad 0 &{}\quad 0 \\ 0 &{}\quad \cos (t) &{}\quad -\sin (t) \\ 0 &{}\quad \sin (t) &{}\quad \cos (t) \end{pmatrix} \end{aligned}$$

be the matrix corresponding to the rotation with angular velocity \({{\,\mathrm{e}\,}}_1\). Define

$$\begin{aligned} \begin{aligned} U(t,y)&:=Q(t)u(t,Q(t)^\top y), \\ \mathfrak {P}(t,y)&:=\mathfrak {p}(t,Q(t)^\top y), \\ F(t,y)&:=Q(t) f(t,Q(t)^\top y) \end{aligned} \end{aligned}$$

with the new spatial variable \(y=Q(t)x\). Due to

$$\begin{aligned} \partial _tU(t,y)=Q(t) (\partial _t u(t,x) + {{\,\mathrm{e}\,}}_1\wedge u(t,x) - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla u(t,x) ), \end{aligned}$$

the functions \(u\), \(\mathfrak {p}\) and f satisfy (5.5) if and only if

$$\begin{aligned} \left\{ \begin{aligned} \omega \partial _t U- \Delta U- \lambda \partial _1 U+ \nabla \mathfrak {P}&= F&\text {in }{\mathbb T}\times \mathbb {R}^3, \\ {{\,\mathrm{div}\,}}U&=0&\text {in }{\mathbb T}\times \mathbb {R}^3. \end{aligned}\right. \end{aligned}$$

The assertions in Theorem 5.3 are now a direct consequence of Theorem 5.1 and estimate (5.4). \(\square \)

Remark 5.4

As for the corresponding steady-state problem (see for example [13, Theorem VIII.8.1]), one can extend Theorem 5.3 to the case of an exterior domain \(\Omega \) for \(f\in \mathrm {L}^{q}({\mathbb T}\times \Omega )\), but it is not clear to the authors whether or not the constant in the resulting a priori estimate can then be chosen independently of \(\lambda \) and \(\omega \). Observe that such an independence is obtained in the functional setting of Theorem 2.2 where \(f\in \mathrm {A}({\mathbb T};\mathrm {L}^{q}(\Omega ))\). Since we solve the nonlinear problem (2.1) via a fixed-point iteration which requires \(\lambda \) and \(\omega \) to be chosen sufficiently small, it crucial to obtain an estimate with the constant independent of \(\lambda \) and \(\omega \).

From Theorem 5.3 we can extract a similar result for the resolvent problem (2.3) in the whole space.

Theorem 5.5

Let \(q\in (1,2)\), \(k\in \mathbb {Z}\) and \(\lambda ,\,\omega , \,\theta >0\) with \(\lambda ^2\le \theta \omega \). For every \(F\in \mathrm {L}^{q}(\mathbb {R}^3)^3\) there exists a solution \((v,p)\in \mathscr {S^\prime }(\mathbb {R}^3)^{3+1}\) to

$$\begin{aligned} \left\{ \begin{aligned} \omega (ik v + {{\,\mathrm{e}\,}}_1\wedge v - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla v ) -\Delta v- \lambda \partial _1 v+ \nabla p&= F&\text {in }\mathbb {R}^3, \\ {{\,\mathrm{div}\,}}v&=0&\text {in }\mathbb {R}^3, \end{aligned}\right. \end{aligned}$$
(5.7)

and a constant \({C}_{8} ={C}_{8}(q,\theta )>0\) with

$$\begin{aligned} \begin{aligned} \omega&\Vert ik v + {{\,\mathrm{e}\,}}_1\wedge v - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla v \Vert _{q} +\Vert \nabla ^2v\Vert _{q} +\lambda \Vert \partial _1v\Vert _{q} \\&\qquad \qquad \qquad \qquad +\lambda ^{1/2}\Vert v\Vert _{s_1} +\lambda ^{1/4}\Vert \nabla v\Vert _{s_2} +\Vert \nabla p\Vert _{q} \le {C}_{8} \Vert F\Vert _{q}. \end{aligned} \end{aligned}$$
(5.8)

Additionally, if \((w,\mathfrak {q})\in \mathscr {S}(\mathbb {R}^3)^{3+1}\) is another solution to (5.1) with \(w\in \mathrm {L}^{r}(\Omega )\) for some \(r\in [1,\infty )\), then \(v=w\), and \(p-\mathfrak {q}\) is constant.

Proof

First consider a solution \((v,p)\) in the described function class. Then the fields

$$\begin{aligned} u(t,x):={{\,\mathrm{e}\,}}^{ikt}v(x), \qquad \mathfrak {p}(t,x):={{\,\mathrm{e}\,}}^{ikt}p(x), \qquad f(t,x):={{\,\mathrm{e}\,}}^{ikt}F(x), \end{aligned}$$

satisfy (5.5). Therefore, uniqueness of \((v,\nabla p)\) follows from the uniqueness statement in Theorem 5.3. To show existence, let \(F\in \mathrm {L}^{q}(\mathbb {R}^3)\) and define \(f\in \mathrm {L}^{q}({\mathbb T}\times \mathbb {R}^3)\) as above. Theorem 5.3 yields the existence of a pair \((u,\mathfrak {p})\) that solves (5.5). Then the k-th Fourier coefficients \(v(x):=\mathscr {F}_{\mathbb T}[u(\cdot ,x)](k)\) and \(p(x):=\mathscr {F}_{\mathbb T}[\mathfrak {p}(\cdot ,x)](k)\) satisfy (5.7), and estimate (5.8) is a direct consequence of (5.6). \(\square \)

5.2 Uniqueness

Next we show a uniqueness result for the resolvent problem (2.3).

Lemma 5.6

Let \(\lambda \ge 0\), \(\omega >0\), \(k\in \mathbb {Z}\), and let \((v,p)\) be a distributional solution to (2.3) with \(F=0\) and \(\nabla ^2v,\,\partial _1v,\,\nabla p\in \mathrm {L}^{q}(\Omega )\) for some \(q\in (1,\infty )\) and \(v\in \mathrm {L}^{s}(\Omega )\) for some \(s\in (1,\infty )\). Then \(v=0\) and \(p\) is constant.

Proof

We only consider the case \(\lambda >0\) here. The proof for \(\lambda =0\) can be shown in exactly the same way. Fix a radius \(R>0\) such that \(\partial \mathrm {B}_R\subset \Omega \), and define a “cut-off” function \(\chi _0\in \mathrm {C}^{\infty }_0(\mathbb {R}^3)\) with \(\chi _0(x)=1\) for \({|x |}\le 2R\) and \(\chi _0(x)=0\) for \({|x |}\ge 4R\). Set

$$\begin{aligned} w:=\chi _0v-\mathfrak {B}(v\cdot \nabla \chi _0), \qquad \mathfrak {q}:=\chi _0p\end{aligned}$$
(5.9)

where \(\mathfrak {B}\) denotes the Bogovskiĭ operator; see for example [13, Section III.3]. Then

$$\begin{aligned} \left\{ \begin{array}{r@{\quad }l} - \Delta w+ \nabla \mathfrak {q}= h&{}\quad \text {in }\Omega _{4R}, \\ {{\,\mathrm{div}\,}}w=0 &{}\quad \text {in }\Omega _{4R}, \\ w=0 &{}\quad \text {on }\partial \Omega _{4R}, \end{array}\right. \end{aligned}$$

with

$$\begin{aligned} \begin{aligned}&h:=\big (-\omega (ik v + {{\,\mathrm{e}\,}}_1\wedge v - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla v )-\lambda \partial _1v\big )\chi _0 \\&\qquad -2\nabla \chi _0\cdot \nabla v-\Delta \chi _0v+\nabla \chi _0p+\Delta \mathfrak {B}(\nabla \chi _0\cdot v). \end{aligned} \end{aligned}$$

From the assumptions, we obtain \(v\in \mathrm {W}^{2,q}(\Omega _{4R})\) and \(p\in \mathrm {W}^{1,q}(\Omega _{4R})\). Standard Sobolev embeddings imply \(v, \nabla v, p\in \mathrm {L}^{\frac{3}{2} q}(\Omega _{4R})\). Therefore, we also have \(h\in \mathrm {L}^{r}(\Omega _{4R})\) for all \(1< r\le \frac{3}{2}q\). From well-known regularity results for the Stokes problem in bounded domains (see [13, Theorem IV.6.1]) we obtain \(w\in \mathrm {W}^{2,r}(\Omega _{4R})\) and \(\nabla \mathfrak {q}\in \mathrm {L}^{r}(\Omega _{4R})\). Since \(v=w\) and \(p=\mathfrak {q}\) on \(\Omega _{2R}\), this yields

$$\begin{aligned} (v,p)\in \mathrm {W}^{2,r}(\Omega _{2R})\times \mathrm {W}^{1,r}(\Omega _{2R}) \end{aligned}$$
(5.10)

for all \(1< r\le \frac{3}{2}q\).

Next consider another “cut-off” function \(\chi _1\in \mathrm {C}^{\infty }(\mathbb {R}^3)\) with \(\chi _1(x)=1\) for \({|x |}\ge 2R\) and \(\chi _1(x)=0\) for \({|x |}\le R\). As above, we define

$$\begin{aligned} u:=\chi _1v-\mathfrak {B}(v\cdot \nabla \chi _1), \qquad \mathfrak {p}:=\chi _1p, \end{aligned}$$
(5.11)

which satisfy the system

$$\begin{aligned} \left\{ \begin{array}{r@{\quad }l} \omega (ik u + {{\,\mathrm{e}\,}}_1\wedge u - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla u ) -\Delta u-\lambda \partial _1u+ \nabla \mathfrak {p}= f&{} \quad \text {in }\mathbb {R}^3, \\ {{\,\mathrm{div}\,}}u=0 &{}\quad \text {in }\mathbb {R}^3, \end{array}\right. \end{aligned}$$
(5.12)

with

$$\begin{aligned} f&:=\omega ({{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla \chi _1)v-2\nabla \chi _1\cdot \nabla v-\Delta \chi _1v+\lambda \partial _1\chi _1v+\nabla \chi _1p-\Delta \mathfrak {B}(v\cdot \nabla \chi _1) \\&+\lambda \partial _1\mathfrak {B}(v\cdot \nabla \chi _1) +\omega (ik \mathfrak {B}(v\cdot \nabla \chi _1) + {{\,\mathrm{e}\,}}_1\wedge \mathfrak {B}(v\cdot \nabla \chi _1) - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla \mathfrak {B}(v\cdot \nabla \chi _1) ). \end{aligned}$$

As above, we see \(f\in \mathrm {L}^{r}(\mathbb {R}^3)\) for all \(1< r\le \frac{3}{2}q\). Since we also have \(u\in \mathrm {L}^{s}(\mathbb {R}^3)\), Theorem 5.5 implies

$$\begin{aligned} ik u + {{\,\mathrm{e}\,}}_1\wedge u - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla u ,\,\nabla ^2u,\,\partial _1u,\,\nabla \mathfrak {p}\in \mathrm {L}^{r}(\mathbb {R}^3) \end{aligned}$$

if additionally \(r<2\). Due to \(v=u\) and \(p=\mathfrak {p}\) on \(\mathrm {B}^{2R}\), we have

$$\begin{aligned} ik v + {{\,\mathrm{e}\,}}_1\wedge v - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla v ,\,\nabla ^2v,\,\partial _1v,\,\nabla p\in \mathrm {L}^{r}(\mathrm {B}^{2R}) \end{aligned}$$
(5.13)

for \(1< r\le \frac{3}{2}q\) with \(r<2\).

We combine (5.10) and (5.13) to deduce

$$\begin{aligned} ik v + {{\,\mathrm{e}\,}}_1\wedge v - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla v ,\,\nabla ^2v,\,\partial _1v,\,\nabla p\in \mathrm {L}^{r}(\Omega ) \end{aligned}$$
(5.14)

for \(1<r\le \frac{3}{2}q\) with \(r<2\). After repeating the above argument a sufficient number of times, we obtain (5.14) for all \(r\in (1,2)\). Since \(v\in \mathrm {L}^{s}(\Omega )\), the Sobolev inequality further yields

$$\begin{aligned} \forall r\in \big (\frac{3}{2},6\big ): \ \nabla v\in \mathrm {L}^{r}(\Omega ), \qquad \forall r\in (3,\infty ) : \ v\in \mathrm {L}^{r}(\Omega ). \end{aligned}$$

In particular, we can employ the divergence theorem to compute

$$\begin{aligned} \int _{\Omega _R} {{\,\mathrm{div}\,}}\big [({{\,\mathrm{e}\,}}_1\wedge x){|v |}^2\big ] \,{\mathrm d}x=\int _{\partial \Omega _R} ({{\,\mathrm{e}\,}}_1\wedge x)\cdot \mathrm {n}{|v |}^2 \,{\mathrm d}S=\int _{\partial \mathrm {B}_R} ({{\,\mathrm{e}\,}}_1\wedge x)\cdot x R^{-1}{|v |}^2 \,{\mathrm d}S=0 \end{aligned}$$

for any \(R>0\) with \(\partial \mathrm {B}_R\subset \Omega \). Passing to the limit \(R\rightarrow \infty \), we obtain

$$\begin{aligned} \int _{\Omega } {{\,\mathrm{div}\,}}\big [({{\,\mathrm{e}\,}}_1\wedge x){|v |}^2\big ] \,{\mathrm d}x=0. \end{aligned}$$
(5.15)

By the above integrability properties, we can further multiply (2.3)\(_{1}\) by \(v\) and integrate over \(\Omega \). By utilizing (5.15) and integration by parts, we conclude

$$\begin{aligned} 0&= \int _{\Omega } \big ( \omega (ik v + {{\,\mathrm{e}\,}}_1\wedge v - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla v ) -\Delta v+\lambda \partial _1v+\nabla p\big ) \cdot v\,{\mathrm d}x\\&= \int _{\Omega } \omega ik \,{|v |}^2 +\frac{1}{2}\omega {{\,\mathrm{div}\,}}\big [({{\,\mathrm{e}\,}}_1\wedge x){|v |}^2\big ] -\Delta v\cdot v+\frac{1}{2}\lambda \partial _1 {|v |}^2 +\nabla p\cdot v\,{\mathrm d}x\\&= \omega ik\int _{\Omega }{|v |}^2\,{\mathrm d}x+\int _{\Omega } {|\nabla v |}^2 \,{\mathrm d}x. \end{aligned}$$

This implies \(\nabla v=0\). The imposed boundary conditions thus yield \(v=0\). Finally, (2.3)\(_{1}\) leads to \(\nabla p=0\), and the proof is complete. \(\square \)

5.3 A Priori Estimate

Next we establish an a priori estimate for the solution to the resolvent problem (2.3).

Lemma 5.7

Let \(q\in (1,2)\), \(k\in \mathbb {Z}\) and \(\lambda ,\,\omega , \,\theta >0\) with \(\lambda ^2\le \theta \omega \). Moreover, let \(F\in \mathrm {L}^{q}(\Omega )\) and \(R>0\) such that \(\partial \mathrm {B}_R\subset \Omega \). Let \((v,p)\in \mathrm {L}^{1}_{\mathrm {loc}}(\Omega )\) with

$$\begin{aligned} \begin{aligned} ik v + {{\,\mathrm{e}\,}}_1\wedge v - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla v ,\,&\nabla ^2v,\, \partial _1v,\, \nabla p\in \mathrm {L}^{q}(\Omega ),\\ v\in \mathrm {L}^{s_1}(\Omega ),\qquad&\nabla v\in \mathrm {L}^{s_2}(\Omega ) \end{aligned} \end{aligned}$$
(5.16)

be a solution to (2.3). Then there exists a constant \({C}_{9} ={C}_{9}(\Omega ,q,\theta ,R)>0\) such that

$$\begin{aligned} \begin{aligned}&\omega \Vert ik v + {{\,\mathrm{e}\,}}_1\wedge v - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla v \Vert _{q} +\Vert \nabla ^2v\Vert _{q} \\&\quad +\lambda \Vert \partial _1v\Vert _{q} +\lambda ^{1/2}\Vert v\Vert _{s_1} +\lambda ^{1/4}\Vert \nabla v\Vert _{s_2} +\Vert \nabla p\Vert _{q} \\&\qquad \le {C}_{9} \big (\Vert F\Vert _{q}+(1+\lambda +\omega )\Vert v\Vert _{1,q;\Omega _{4R}} +\omega {|k |}\,\Vert v\Vert _{-1,q;\Omega _{4R}} +\Vert p\Vert _{q;\Omega _{4R}}\big ). \end{aligned} \end{aligned}$$
(5.17)

Proof

Let \(\chi _0\), \(\chi _1\) be the “cut-off” functions from the proof of Lemma 5.6. Define \(w\in \mathrm {W}^{2,q}(\Omega )\) and \(\mathfrak {q}\in \mathrm {W}^{1,q}(\Omega )\) as in (5.9). Then

$$\begin{aligned} \left\{ \begin{array}{r@{\quad }l} i k\omega \,w- \Delta w+ \nabla \mathfrak {q}= h&{}\quad \text {in }\Omega _{4R}, \\ {{\,\mathrm{div}\,}}w=0 &{}\quad \text {in }\Omega _{4R}, \\ w=0 &{}\quad \text {on }\partial \Omega _{4R}, \end{array}\right. \end{aligned}$$

with

$$\begin{aligned} \begin{aligned} h&:=\big (F-\omega ( {{\,\mathrm{e}\,}}_1\wedge v - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla v)-\lambda \partial _1v\big )\chi _0 \\&\qquad -2\nabla \chi _0\cdot \nabla v-\Delta \chi _0v+\nabla \chi _0p-(ik\omega -\Delta )\mathfrak {B}(v\cdot \nabla \chi _0). \end{aligned} \end{aligned}$$

Well-known theory for the Stokes resolvent problem (see for example [9]) yields

$$\begin{aligned} \begin{aligned}&\Vert v\Vert _{2,q;\Omega _{2R}} +\Vert \nabla p\Vert _{q;\Omega _{2R}} \le \Vert w\Vert _{2,q;\Omega _{4R}} +\Vert \nabla \mathfrak {q}\Vert _{q;\Omega _{4R}} \le {c}_{0}\Vert h\Vert _{q;\Omega _{4R}}\\&\quad \le {c}_{1}\big ( \Vert F\Vert _q +(1+\lambda +\omega ) \Vert v\Vert _{1,q;\Omega _{4R}} +\Vert p\Vert _{q;\Omega _{4R}} +\omega {|k |}\,{|v\cdot \nabla \chi _0 |}_{-1,q;\Omega _{4R}}^*\big ). \end{aligned} \end{aligned}$$
(5.18)

In the last estimate we used mapping properties of the Bogovskiĭ operator (see [13, Section III.3]), namely

$$\begin{aligned} \Vert \nabla \mathfrak {B}h\Vert _{m,q;\Omega _{4R}}\le {c}_{2}\Vert h\Vert _{m,q;\Omega _{4R}}, \qquad \Vert \mathfrak {B}h\Vert _{q;\Omega _{4R}}\le {c}_{3}{|h |}_{-1,q;\Omega _{4R}}^*\end{aligned}$$

for \(m\in \mathbb {N}_0\), where

$$\begin{aligned} {|h |}_{-1,q;D}^*:=\sup \biggl \{{\bigl |\int _{D}h\psi \,{\mathrm d}x \big |}\ \bigg \vert \ \psi \in \mathrm {C}^{\infty }_0(\mathbb {R}^3),\ \Vert \nabla \psi \Vert _{q/(q-1);D}=1\biggr \}. \end{aligned}$$

To estimate the last term in (5.18), we introduce the notation

$$\begin{aligned} {\overline{\psi }}:=\psi -\frac{1}{{\bigl |\Omega _{4R} \big |}}\int _{\Omega _{4R}}\psi \,{\mathrm d}x\end{aligned}$$

for \(\psi \in \mathrm {C}^{\infty }_0(\mathbb {R}^3)\), and we employ that \({{\,\mathrm{div}\,}}v=0\) in \(\Omega \) and \(v=0\) on \(\partial \Omega \) to deduce the identity

$$\begin{aligned} \int _{\Omega _{4R}}v\cdot \nabla \chi _0 \psi \,{\mathrm d}x&=\int _{\Omega _{4R}}{{\,\mathrm{div}\,}}(v\chi _0) \psi \,{\mathrm d}x=-\int _{\Omega _{4R}}\chi _0 v\cdot \nabla {\overline{\psi }} \,{\mathrm d}x\\&=\int _{\Omega _{4R}}{{\,\mathrm{div}\,}}(v\chi _0){\overline{\psi }} \,{\mathrm d}x=\int _{\Omega _{4R}}v\cdot \nabla \chi _0{\overline{\psi }} \,{\mathrm d}x. \end{aligned}$$

Since Poincaré’s inequality yields

$$\begin{aligned} \Vert {\overline{\psi }}{\nabla \chi _0}\Vert _{1,q';\Omega _{4R}} \le {c}_{4}\Vert {\overline{\psi }}\Vert _{1,q';\Omega _{4R}} \le {c}_{5}\Vert \nabla {\psi }\Vert _{q';\Omega _{4R}}, \end{aligned}$$

we have

$$\begin{aligned}&{|v\cdot \nabla \chi _0 |}_{-1,q;\Omega _{4R}}^*\\&\quad \le \sup \bigl \{\Vert v\Vert _{-1,q;\Omega _{4R}}\Vert {\overline{\psi }}{\nabla \chi _0}\Vert _{1,q';\Omega _{4R}}\ \big \vert \ \psi \in \mathrm {C}^{\infty }_0(\mathbb {R}^3),\ \Vert \nabla \psi \Vert _{q';\Omega _{4R}}=1\bigr \}\\&\quad \le {c}_{6} \Vert v\Vert _{-1,q;\Omega _{4R}}. \end{aligned}$$

Applying this estimate to the last term in (5.18), we obtain

$$\begin{aligned} \begin{aligned}&\Vert v\Vert _{2,q;\Omega _{2R}} +\Vert \nabla p\Vert _{q;\Omega _{2R}} \\&\qquad \le {c}_{7}\big ( \Vert F\Vert _q +(1+\lambda +\omega ) \Vert v\Vert _{1,q;\Omega _{4R}} +\Vert p\Vert _{q;\Omega _{4R}} +\omega {|k |}\,\Vert v\Vert _{-1,q;\Omega _{4R}} \big ). \end{aligned} \end{aligned}$$
(5.19)

Next define (\(u,\mathfrak {p})\) as in (5.11), which satisfies the system

$$\begin{aligned} \left\{ \begin{array}{r@{\quad }l} \omega (ik u + {{\,\mathrm{e}\,}}_1\wedge u - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla u ) -\Delta u-\lambda \partial _1u+ \nabla \mathfrak {p}= f&{} \quad \text {in }\mathbb {R}^3, \\ {{\,\mathrm{div}\,}}u=0 &{}\quad \text {in }\mathbb {R}^3, \end{array}\right. \end{aligned}$$

with

$$\begin{aligned} f&:=\chi _1F-\omega ({{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla \chi _1)v-2\nabla \chi _1\cdot \nabla u-\Delta \chi _1v\\&\qquad +\lambda \partial _1\chi _1v+\nabla \chi _1p-\Delta \mathfrak {B}(v\cdot \nabla \chi _1) +\lambda \partial _1\mathfrak {B}(v\cdot \nabla \chi _1) \\&\qquad +\omega (ik \mathfrak {B}(v\cdot \nabla \chi _1) + {{\,\mathrm{e}\,}}_1\wedge \mathfrak {B}(v\cdot \nabla \chi _1) - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla \mathfrak {B}(v\cdot \nabla \chi _1) ). \end{aligned}$$

Theorem 5.5 implies

$$\begin{aligned} \begin{aligned} \omega&\Vert ik v + {{\,\mathrm{e}\,}}_1\wedge v - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla v \Vert _{q;\Omega ^{2R}} +\Vert \nabla ^2v\Vert _{q;\Omega ^{2R}} +\lambda \Vert \partial _1v\Vert _{q;\Omega ^{2R}}\\&\qquad +\lambda ^{1/4}\Vert \nabla v\Vert _{s_2;\Omega ^{2R}} +\lambda ^{1/2}\Vert v\Vert _{s_1;\Omega ^{2R}} +\Vert \nabla p\Vert _{q;\Omega ^{2R}} \\&\le \omega \Vert ik u + {{\,\mathrm{e}\,}}_1\wedge u - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla u \Vert _{q} +\Vert \nabla ^2u\Vert _{q} +\lambda \Vert \partial _1u\Vert _{q} \\&\qquad +\lambda ^{1/4}\Vert \nabla u\Vert _{s_2} +\lambda ^{1/2}\Vert u\Vert _{s_1} +\Vert \nabla \mathfrak {p}\Vert _{q} \\&\le {c}_{8}\big ( \Vert F\Vert _q +(1+\lambda +\omega )\Vert v\Vert _{1,q;\Omega _{2R}} +\Vert p\Vert _{q;\Omega _{2R}} +\omega {|k |}\Vert v\Vert _{-1,q;\Omega _{2R}} \big ), \end{aligned} \end{aligned}$$

where we estimated the terms containing the Bogovskiĭ operator as above. Combining this estimate with (5.19), we conclude (5.17). \(\square \)

In the next step we improve estimate (5.17) by showing that the lower-order terms on the right-hand side can be omitted. This leads to the desired estimate (2.4) with the asserted dependencies of the constant \({C}_{1}\).

Lemma 5.8

Let \(q\in (1,2)\), \(k\in \mathbb {Z}\) and \(\lambda ,\,\omega >0\), and let \(F\in \mathrm {L}^{q}(\Omega )\). Let \((v,p)\in \mathrm {L}^{1}_{\mathrm {loc}}(\Omega )\) be a solution to (2.3) in the class (5.16). Then estimate (2.4) holds for a constant \({C}_{1} ={C}_{1}(\Omega ,q,\lambda ,\omega )>0\). If \(q\in (1,\frac{3}{2})\) and \(\lambda ^2\le \theta \omega \le B\), then this constant can be chosen independently of \(\lambda \) and \(\omega \) such that \({C}_{1} ={C}_{1}(\Omega ,q,\theta ,B)\).

Proof

We employ a contradiction argument. At first, consider the case \(q\in (1,\frac{3}{2})\) and assume that (2.4) is not valid for a constant \({C}_{1}={C}_{1}(\Omega ,q,\theta ,B)\). Then there exist sequences of numbers \((\lambda _j)\subset (0,\sqrt{B}]\), \((\omega _j)\subset (0,B/\theta ]\) with \(\lambda _j^2\le \theta \omega _j\), and \((k_j)\subset \mathbb {Z}\), and of functions \((v_j)\), \((p_j)\), \((F_j)\) that satisfy

$$\begin{aligned} \begin{aligned}&\omega _j\Vert ik_j v_j + {{\,\mathrm{e}\,}}_1\wedge v_j - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla v_j \Vert _{q} +\Vert \nabla ^2v_j\Vert _{q}\\&\qquad +\lambda _j\Vert \partial _1v_j\Vert _{q} +\lambda _j^{1/2}\Vert v_j\Vert _{s_1} +\lambda _j^{1/4}\Vert \nabla v_j\Vert _{s_2} +\Vert \nabla p_j\Vert _{q} =1, \end{aligned} \end{aligned}$$
(5.20)

\(\Vert F_j\Vert _{q}\rightarrow 0\) as \(j\rightarrow \infty \), and

$$\begin{aligned} \left\{ \begin{array}{r@{\quad }l} \omega _j(ik_j v_j + {{\,\mathrm{e}\,}}_1\wedge v_j - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla v_j ) - \Delta v_j - \lambda _j \partial _1 v_j + \nabla p_j = F_j &{} \quad \text {in }\Omega , \\ {{\,\mathrm{div}\,}}v_j=0 &{}\quad \text {in }\Omega ,\\ v_j=0 &{}\quad \text {on }\partial \Omega , \end{array}\right. \end{aligned}$$
(5.21)

for all \(j\in \mathbb {N}\). Furthermore, without loss of generality, we assume \(\int _{\Omega _R}p_j\,{\mathrm d}x=0\) for \(R>0\) as in Lemma 5.7. Then, \((\lambda _j)\), \((\omega _j)\) and \((k_j)\) contain (improper) convergent subsequences with limits \(\lambda \in [0,\sqrt{B}]\), \(\omega \in [0,B/\theta ]\) and \(k\in \mathbb {Z}\cup \{\pm \infty \}\), respectively, and we have \(\lambda ^2\le \theta \omega \). For simplicity, we identify selected subsequences with the actual sequences. Moreover, (5.20) implies that \(U_j:=(i\omega _j k_jv_j,v_j,p_j)\) is bounded in \(\mathrm {L}^{q}(\Omega _\rho )\times \mathrm {W}^{2,q}(\Omega _\rho )\times \mathrm {W}^{1,q}(\Omega _\rho )\) for any \(\rho >R\). Hence, by a Cantor diagonalization argument, there exists a subsequence that converges weakly in \(\mathrm {L}^{q}(\Omega _\rho )\times \mathrm {W}^{2,q}(\Omega _\rho )\times \mathrm {W}^{1,q}(\Omega _\rho )\) to some \(U:=(w,v,p)\) for each \(\rho >R\). Consequently, passing to the limit \(j\rightarrow \infty \) in (5.21), we obtain

$$\begin{aligned} \left\{ \begin{aligned} w+\omega ( {{\,\mathrm{e}\,}}_1\wedge v - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla v) - \Delta v- \lambda \partial _1 v+ \nabla \mathfrak {p}&= 0&\text {in }\Omega , \\ {{\,\mathrm{div}\,}}v&=0&\text {in }\Omega ,\\ v&=0&\text {on }\partial \Omega . \end{aligned}\right. \end{aligned}$$
(5.22)

Moreover, by the compact embeddings

$$\begin{aligned} \mathrm {W}^{2,q}(\Omega _{4R})\hookrightarrow \mathrm {W}^{1,q}(\Omega _{4R})\hookrightarrow \mathrm {L}^{q}(\Omega _{4R})\hookrightarrow \mathrm {W}^{-1,q}(\Omega _{4R}), \end{aligned}$$

we deduce that \(U\) is the strong limit of \((U_j)\) in the topology of \(\mathrm {W}^{-1,q}(\Omega _{4R})\times \mathrm {W}^{1,q}(\Omega _{4R})\times \mathrm {L}^{q}(\Omega _{4R})\). By Lemma 5.7,

$$\begin{aligned} \begin{aligned}&\omega _j\Vert ik_j v_j + {{\,\mathrm{e}\,}}_1\wedge v_j - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla v_j \Vert _{q} +\Vert \nabla ^2v_j\Vert _{q}\\&\quad +\lambda _j\Vert \partial _1v_j\Vert _{q} +\lambda _j^{1/2}\Vert v_j\Vert _{s_1} +\lambda _j^{1/4}\Vert \nabla v_j\Vert _{s_2} +\Vert \nabla p_j\Vert _{q}\\&\qquad \le {C}_{9} \big (\Vert F_j\Vert _{q}+(1+\lambda _j+\omega _j)\Vert v_j\Vert _{1,q;\Omega _{4R}} +\omega {|k_j |}\,\Vert v_j\Vert _{-1,q;\Omega _{4R}} +\Vert p_j\Vert _{q;\Omega _{4R}}\big ). \end{aligned} \end{aligned}$$

Passing to the limit \(j\rightarrow \infty \) in this estimate, we conclude in virtue of (5.20) that

$$\begin{aligned} 1 \le {C}_{9} \big ((1+\lambda +\omega )\Vert v\Vert _{1,q;\Omega _{4R}} +\Vert w\Vert _{-1,q;\Omega _{4R}} +\Vert p\Vert _{q;\Omega _{4R}}\big ). \end{aligned}$$
(5.23)

Moreover,

$$\begin{aligned} \Vert w+ \omega ( {{\,\mathrm{e}\,}}_1\wedge v - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla v)\Vert _{q} +\Vert \nabla ^2v\Vert _{q} +\lambda \Vert \partial _1v\Vert _{q}\nonumber \\ +\lambda ^{1/2}\Vert v\Vert _{s_1} +\lambda ^{1/4}\Vert \nabla v\Vert _{s_2} +\Vert \nabla p\Vert _{q}< \infty . \end{aligned}$$
(5.24)

Now we distinguish between several cases:

  1. i.

    If \(\omega _j k_j\rightarrow s\in \mathbb {R}\) and \(\omega =0\), then \(\lambda =0\) and \(w=isv\), so that (5.22) reduces to a Stokes resolvent problem. If \(s\ne 0\), we also have \(v\in \mathrm {L}^{q}(\Omega )\) and we conclude \(v=\nabla p=0\) from a well-known uniqueness result; see for example [9]. If \(s=0\), we utilize that \(q<\frac{3}{2}\) and \(v_j\in \mathrm {L}^{s_1}(\Omega )\), \(\nabla v_j\in \mathrm {L}^{s_2}(\Omega )\), so that Sobolev’s inequality implies

    $$\begin{aligned} \Vert v_j\Vert _{3q/(3-2q)}\le {c}_{0} \Vert \nabla v_j\Vert _{3q/(3-q)}\le {c}_{1}\Vert \nabla ^2v_j\Vert _{q}, \end{aligned}$$

    and thus \(v\in \mathrm {L}^{3q/(3-2q)}(\Omega )\). Now \(v=\nabla p=0\) follows from classical uniqueness properties of the steady-state Stokes problem, see for example [13, Theorem V.4.6].

  2. ii.

    If \(\omega _j k_j\rightarrow s\in \mathbb {R}\) and \(\omega \ne 0\) but \(\lambda =0\), then \(k_j\rightarrow k\in \mathbb {Z}\) and \(w=i\omega kv\), so that (5.22) reduces to (2.3) with \(\lambda =0\). As above, we deduce \(v\in \mathrm {L}^{3q/(3-2q)}(\Omega )\). From Lemma 5.6 we conclude \(v=\nabla p=0\).

  3. iii.

    If \(\omega _j k_j\rightarrow s\in \mathbb {R}\) and \(\omega \ne 0\) and \(\lambda \ne 0\), then \(k_j\rightarrow k\in \mathbb {Z}\) and \(w=i\omega kv\), so that \((v,p)\) satisfies (2.3). Since \(\lambda \ne 0\), it follows from (5.24) that \(v\in \mathrm {L}^{s_1}(\Omega )\). Lemma 5.6 thus implies \(v=\nabla p=0\).

  4. iv.

    If \(\omega _j {|k_j |}\rightarrow \infty \), we recall (5.20) and estimate

    $$\begin{aligned} \omega _j{|k_j |}\Vert v_j\Vert _{q;\Omega _\rho } \le \omega _j\Vert ik_j v_j + {{\,\mathrm{e}\,}}_1\wedge v_j - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla v_j \Vert _{q;\Omega _\rho } +{c}_{2}(\rho )\Vert v_j\Vert _{1,q;\Omega _\rho } \le {c}_{3}(\rho ) \end{aligned}$$

    for any \(\rho >R\). Passing to the limit \(j\rightarrow \infty \), we thus obtain \(v=0\) on \(\Omega _\rho \) for each \(\rho >R\), whence \(v= 0\) on \(\Omega \). Hence, (5.22)\(_{1}\) reduces to \(w+\nabla p=0\). Clearly, we also have \({{\,\mathrm{div}\,}}w=0\) and \(w\big | _{\partial \Omega }=0\), so that \(w+\nabla p=0\) corresponds to the Helmholtz decomposition of 0 in \(\mathrm {L}^{q}(\Omega )\). Since this decomposition is unique, we conclude \(w=\nabla p=0\).

Consequently, all four cases lead to \(w=v=\nabla p=0\), which contradicts (5.23). This completes the proof in the case \(1<q<\frac{3}{2}\).

In the more general case \(q\in (1,2)\), where we do not assert the constant \({C}_{1}\) to be independent of \(\lambda \) and \(\omega \), these parameters remain fixed in the contradiction argument above. Consequently, only the last two cases above have to be considered. The conclusion in both of these cases is valid for all \(q\in (1,2)\), and we thus conclude the lemma. \(\square \)

5.4 Existence

To complete the proof of Theorem 2.1, it remains to show existence of a solution. For this purpose, recall the following property of the Stokes operator.

Lemma 5.9

Let \(D\subset \mathbb {R}^3\) be a bounded domain with \(\mathrm {C}^{3}\)-boundary. Every \(u\in \mathrm {L}^{2}_{\sigma }(D)\cap \mathrm {W}^{1,2}_0(D)\cap \mathrm {W}^{2,2}(D)\) satisfies

$$\begin{aligned} \Vert \nabla ^2u\Vert _2 \le {C}_{10} \big (\Vert {\mathcal P}_\mathrm {H}\Delta u\Vert _2+\Vert \nabla u\Vert _2\big ) \end{aligned}$$

for a constant \({C}_{10}={C}_{10}(D)>0\) that does not depend on the “size” of D but solely on its “regularity”. In particular, if \(D=\Omega _R\) for an exterior domain \(\Omega \) with \(\partial \Omega \subset \mathrm {B}_R\), the constant \({C}_{10}\) is independent of R and solely depends on \(\Omega \).

Proof

See [26, Lemma 1]. \(\square \)

We further need the following identity from [20].

Lemma 5.10

Let \(u\in \mathrm {L}^{2}_{\sigma }(\Omega _R)\cap \mathrm {W}^{1,2}_0(\Omega _R)\cap \mathrm {W}^{2,2}(\Omega _R)\) with complex conjugate \(u^*\). Then \( {{\,\mathrm{e}\,}}_1\wedge u - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla u \in \mathrm {L}^{2}_{\sigma }(\Omega _R)\) and

$$\begin{aligned}&\int _{\Omega _R}( {{\,\mathrm{e}\,}}_1\wedge u - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla u)\cdot {\mathcal P}_\mathrm {H}\Delta u^*\,{\mathrm d}x\\&=\int _{\partial \Omega } \frac{1}{2}{|\nabla u |}^2 ({{\,\mathrm{e}\,}}_1\wedge x)\cdot \mathrm {n}-\mathrm {n}\cdot \nabla u^*\cdot ({{\,\mathrm{e}\,}}_1\wedge x \cdot \nabla u)\,{\mathrm d}S- \int _{\Omega _R} \nabla ({{\,\mathrm{e}\,}}_1\wedge u):\nabla u^*\,{\mathrm d}x. \end{aligned}$$

Proof

See [20, Lemma 3]. \(\square \)

Existence of a solution to the resolvent problem (2.3) can be shown via a Galerkin approach combined with an “invading domains” technique.

Lemma 5.11

Let \(\Omega \subset \mathbb {R}^3\) be an exterior domain of class \(\mathrm {C}^{3}\). Let \(\lambda ,\,\omega >0\), \(k\in \mathbb {Z}\), and let \(F\in \mathrm {C}^{\infty }_0(\Omega )\). Then there exists a solution \((v,p)\) to (2.3) with

$$\begin{aligned} \begin{aligned}&ik v + {{\,\mathrm{e}\,}}_1\wedge v - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla v ,\, \nabla ^2v,\, \partial _1v,\, \nabla p\in \mathrm {L}^{q}(\Omega ),\\&v\in \mathrm {L}^{2q/(2-q)}(\Omega ),\qquad \nabla v\in \mathrm {L}^{4q/(4-q)}(\Omega ) \end{aligned} \end{aligned}$$

for all \(q\in (1,2)\).

Proof

Let \(R>0\) such that \(\partial \mathrm {B}_R\subset \Omega \), and take \(m\in \mathbb {N}\) with \(m>2R\). Since the Stokes operator in the bounded domain \(\Omega _m\) is a positive self-adjoint invertible operator (see [50, Chapter III, Theorem 2.1.1]), there exist sequences \((\psi _j)_{j\in \mathbb {N}}\) of (real-valued) eigenfunctions and \((\mu _j)_{j\in \mathbb {N}}\subset (0,\infty )\) of eigenvalues, that is,

$$\begin{aligned} -{\mathcal P}_\mathrm {H}\Delta \psi _j=\mu _j\psi _j, \qquad \psi _j\in \mathrm {L}^{2}_{\sigma }(\Omega _m)\cap \mathrm {W}^{1,2}_0(\Omega _m)\cap \mathrm {W}^{2,2}(\Omega _m), \end{aligned}$$

normalized such that

$$\begin{aligned} \int _{\Omega _m} \psi _j \cdot \psi _\ell \,{\mathrm d}x= \frac{1}{\mu _j}\delta _{j\ell }. \end{aligned}$$

We show the existence of a function \(u=u^m_n\in X^m_n:={{\,\mathrm{span}\,}}_{\mathbb {C}}\bigl \{\psi _j\ \big \vert \ j=1,\ldots ,n\bigr \}\) satisfying

$$\begin{aligned} \int _{\Omega _m} \big [ \omega (ik u + {{\,\mathrm{e}\,}}_1\wedge u - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla u ) -\Delta u-\lambda \partial _1u\big ] \cdot \psi _j \,{\mathrm d}x=\int _{\Omega _m} F\cdot \psi _j\,{\mathrm d}x\end{aligned}$$
(5.25)

for all \(j\in \{1,\ldots ,n\}\). Since

$$\begin{aligned} u=\sum _{\ell =1}^n\xi _\ell \psi _{\ell } \end{aligned}$$

for some \(\xi _1,\ldots ,\xi _n\in \mathbb {C}\), this is equivalent to solving the algebraic equation

$$\begin{aligned} (\mathrm {I}+ M) \xi = c \end{aligned}$$
(5.26)

with \(\xi =(\xi _1,\ldots ,\xi _n) \in \mathbb {C}^n\) and \(M=(M_{\ell j})\in \mathbb {C}^{n\times n}\), \(c=(c_j)\in \mathbb {C}^n\) with

$$\begin{aligned} M_{\ell j}&:=\int _{\Omega _m} \big ( \omega (ik \psi _\ell + {{\,\mathrm{e}\,}}_1\wedge \psi _\ell - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla \psi _\ell ) -\lambda \partial _1 \psi _\ell \big )\cdot \psi _j\,{\mathrm d}x, \\ c_j&:=\int _{\Omega _m} F\cdot \psi _j\,{\mathrm d}x. \end{aligned}$$

Note that (5.26) is a resolvent problem for the skew-Hermitian matrix M, which is uniquely solvable. Existence of a unique solution \(u=u^m_n\in X^m_n\) to (5.25) thus follows.

Next we need suitable estimates for \(u=u^m_n\). Multiplication of both sides of (5.25) by the complex conjugate coefficient \(\xi _j^*\) and summation over \(j=1,\dots ,n\) yields

$$\begin{aligned} \Vert \nabla u\Vert _2^2 + \int _{\Omega _m}\big ( \omega (ik u + {{\,\mathrm{e}\,}}_1\wedge u - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla u ) -\lambda \partial _1u\big ) \cdot u^*\,{\mathrm d}x= \int _{\Omega _m} F\cdot u^*\,{\mathrm d}x. \end{aligned}$$

Because the integral term on the left-hand side is purely imaginary, taking the real part of this equation leads to the estimate

$$\begin{aligned} \Vert \nabla u\Vert _2^2 \le \Vert F\Vert _{6/5}\Vert u\Vert _{6}. \end{aligned}$$

Recalling the Sobolev inequality \(\Vert u\Vert _{6}\le {c}_{0}\Vert \nabla u\Vert _2\), we obtain

$$\begin{aligned} \Vert u\Vert _{6} +\Vert \nabla u\Vert _2 \le {c}_{1} \Vert F\Vert _{6/5}, \end{aligned}$$
(5.27)

where \({c}_{1} \) is independent of m. If we multiply both sides of (5.25) by \(\mu _j\xi _j^*\) and sum over \(j=1,\dots ,n\), we obtain

$$\begin{aligned} \Vert {\mathcal P}_\mathrm {H}\Delta u\Vert _2^2 = \int _{\Omega _m} \big [F-\omega (ik u + {{\,\mathrm{e}\,}}_1\wedge u - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla u ) +\lambda \partial _1u\big ] \cdot {\mathcal P}_\mathrm {H}\Delta u^*\,{\mathrm d}x. \end{aligned}$$

Taking real part of both sides and observing that

$$\begin{aligned} {{\,\mathrm{Re}\,}}\int _{\Omega _m} iku\cdot {\mathcal P}_\mathrm {H}\Delta u^*\,{\mathrm d}x= - {{\,\mathrm{Re}\,}}\big (ik \Vert \nabla u\Vert _{2}^2\big ) =0, \end{aligned}$$

we conclude, using Hölder’s inequality, the estimate

$$\begin{aligned} \begin{aligned}&\Vert {\mathcal P}_\mathrm {H}\Delta u\Vert _2^2 \le \big ( \Vert F\Vert _2 +\lambda \Vert \partial _1u\Vert _2\big ) \Vert {\mathcal P}_\mathrm {H}\Delta u\Vert _2 \\&\qquad +{{\,\mathrm{Re}\,}}\int _{\Omega _m} \omega ( {{\,\mathrm{e}\,}}_1\wedge u - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla u) \cdot {\mathcal P}_\mathrm {H}\Delta u^*\,{\mathrm d}x. \end{aligned} \end{aligned}$$
(5.28)

Using Lemma 5.10, we estimate the remaining integral on the right-hand side to conclude

$$\begin{aligned} {{\,\mathrm{Re}\,}}\int _{\Omega _m}\omega ( {{\,\mathrm{e}\,}}_1\wedge u - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla u)\cdot {\mathcal P}_\mathrm {H}\Delta u^*\,{\mathrm d}x\le {c}_{2}\omega \big ( \Vert \nabla u\Vert _{2;\partial \Omega }^2 +\Vert \nabla u\Vert _{2;\Omega _m}^2 \big ) \end{aligned}$$

with \({c}_{2}\) independent of m. Employing the trace inequality [13, Theorem II.4.1] on the domain \(\Omega _{R}\), we further estimate

$$\begin{aligned} {{\,\mathrm{Re}\,}}\int _{\Omega _m}&\omega ( {{\,\mathrm{e}\,}}_1\wedge u - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla u)\cdot {\mathcal P}_\mathrm {H}\Delta u^*\,{\mathrm d}x\\&\le {c}_{3}\omega \big ( \Vert \nabla u\Vert _{2;\Omega _{R}} \Vert \nabla u\Vert _{1,2;\Omega _{R}} +\Vert \nabla u\Vert _{2;\Omega _m}^2 \big ) \\&\le {c}_{4}(\varepsilon ) (\omega +\omega ^2)\Vert \nabla u\Vert _{2;\Omega _m}^2 +\varepsilon \Vert \nabla ^2u\Vert _{2;\Omega _m}^2 \end{aligned}$$

for small \(\varepsilon >0\). From Lemma 5.9 we deduce

$$\begin{aligned} \begin{aligned}&{{\,\mathrm{Re}\,}}\int _{\Omega _m}\omega ( {{\,\mathrm{e}\,}}_1\wedge u - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla u)\cdot {\mathcal P}_\mathrm {H}\Delta u^*\,{\mathrm d}x\\&\qquad \le {c}_{5}(\varepsilon )(\omega +\omega ^2) \Vert \nabla u\Vert _{2;\Omega _m}^2 +\varepsilon {c}_{6}\Vert {\mathcal P}_\mathrm {H}\Delta u\Vert _{2;\Omega _m}^2 \end{aligned} \end{aligned}$$

with a constant \({c}_{6} >0\) independent of m. Combining this estimate with (5.28), choosing \(\varepsilon \) sufficiently small and employing estimate (5.27), we arrive at

$$\begin{aligned} \Vert {\mathcal P}_\mathrm {H}\Delta u\Vert _{2;\Omega _m} \le {c}_{7} \big (1+\lambda +\sqrt{\omega +\omega ^2}\big ) \big (\Vert F\Vert _2 + \Vert F\Vert _{6/5}\big ). \end{aligned}$$

Using Lemma 5.9 and estimate (5.27) once again and restoring the original notation, we end up with

$$\begin{aligned} \Vert \nabla ^2u^m_n\Vert _{2;\Omega _m} \le {c}_{8}\big ( \Vert {\mathcal P}_\mathrm {H}\Delta u^m_n\Vert _{2;\Omega _m} +\Vert \nabla u^m_n\Vert _{2;\Omega _m} \big ) \le {c}_{9} \big (\Vert F\Vert _2 + \Vert F\Vert _{6/5}\big ) \end{aligned}$$
(5.29)

with \({c}_{9}\) independent of m.

In particular, we see from (5.27), (5.29) and Poincaré’s inequality that \((u^m_n)\) is uniformly bounded in \(\mathrm {W}^{2,2}(\Omega _m)\) and thus contains a subsequence that converges weakly to some function \(v^m\in \mathrm {L}^{2}_{\sigma }(\Omega _m)\cap \mathrm {W}^{1,2}_0(\Omega _m)\cap \mathrm {W}^{2,2}(\Omega _m)\), which obeys the estimate

$$\begin{aligned} \Vert v^m\Vert _{6;\Omega _m} +\Vert \nabla v^m\Vert _{1,2;\Omega _m} \le {c}_{10}\big (\Vert F\Vert _{6/5}+\Vert F\Vert _{2}\big ) \end{aligned}$$
(5.30)

with \({c}_{10}\) independent of m. Moreover, \(v^m\) satisfies (5.25) for all \(j\in \mathbb {N}\), whence there exists \(p^m\in \mathrm {W}^{1,2}(\Omega _m)\) such that

$$\begin{aligned} \left\{ \begin{array}{r@{\quad }l} \omega (ik v^m + {{\,\mathrm{e}\,}}_1\wedge v^m - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla v^m ) - \Delta v^m - \lambda \partial _1 v^m + \nabla p^m =F&{}\quad \text {in }\Omega _m, \\ {{\,\mathrm{div}\,}}v^m=0 &{}\quad \text {in }\Omega _m,\\ v^m=0 &{}\quad \text {on }\partial \Omega _m; \end{array}\right. \end{aligned}$$
(5.31)

see [13, Corollary III.5.1]. Since \( {{\,\mathrm{e}\,}}_1\wedge v^m - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla v^m \in \mathrm {L}^{2}_{\sigma }(\Omega _m)\) by Lemma 5.10, we deduce from (5.31) and (5.30) the estimate

$$\begin{aligned} \omega&\Vert ik v^m + {{\,\mathrm{e}\,}}_1\wedge v^m - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla v^m \Vert _{2} =\omega \bigl \Vert {\mathcal P}_\mathrm {H}(ik v^m + {{\,\mathrm{e}\,}}_1\wedge v^m - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla v^m )\bigr \Vert _{2} \\&\le \Vert {\mathcal P}_\mathrm {H}F\Vert _{2} +\Vert {\mathcal P}_\mathrm {H}\Delta v^m\Vert _{2} + \lambda \Vert {\mathcal P}_\mathrm {H}\partial _1v^m\Vert _{2} \le {c}_{11}\big (\Vert F\Vert _{6/5}+\Vert F\Vert _{2}\big ). \end{aligned}$$

Combining the estimate above with (5.30), we conclude

$$\begin{aligned} \begin{aligned} \Vert v^m\Vert _{6;\Omega _m} +\Vert \nabla v^m\Vert _{1,2;\Omega _m}&+\omega \Vert ik v^m + {{\,\mathrm{e}\,}}_1\wedge v^m - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla v^m \Vert _{2;\Omega _m}\\&\qquad \qquad \qquad \qquad \quad \le {c}_{12}\big (\Vert F\Vert _{6/5}+\Vert F\Vert _{2}\big ) \end{aligned} \end{aligned}$$
(5.32)

with \({c}_{12}\) independent of m.

Now we introduce a sequence of rotationally symmetric “cut-off” functions \((\chi _m)\subset \mathrm {C}^{\infty }_0(\mathbb {R}^3)\) satisfying

$$\begin{aligned} \begin{aligned} \chi _m(x)&=1 \ \text {for }{|x |}\le \frac{m}{2},&\qquad {|\nabla \chi _m |}&\le \frac{{c}_{13}}{m}, \\ \chi _m(x)&=0 \ \text {for }{|x |}\ge \frac{3m}{4},&\qquad {|\nabla ^2\chi _m |}&\le \frac{{c}_{14}}{m^2}, \end{aligned} \end{aligned}$$

and we set \(w^m:=\chi _mv^m\). Then \(w^m\) is an element of \(\mathrm {W}^{2,2}(\Omega )\). Moreover, the rotational symmetry of \(\chi _m\) implies \({{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla \chi _m=0\). Therefore, from (5.32) and the properties of \(\chi _m\), we deduce the estimate

$$\begin{aligned} \Vert w^m\Vert _{6} +\Vert \nabla w^m\Vert _{1,2} +\omega \Vert ik w^m + {{\,\mathrm{e}\,}}_1\wedge w^m - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla w^m \Vert _{2}&\le {c}_{15} \big (\Vert F\Vert _{6/5}+\Vert F\Vert _{2}\big ) \end{aligned}$$

with \({c}_{15}\) independent of m. This implies the existence of a subsequence, still denoted by \((w^m)\), that converges in the sense of distributions to some function \(v\in \mathrm {W}^{2,2}_{\mathrm {loc}}(\Omega )\) that satisfies

$$\begin{aligned} \Vert v\Vert _{6} +\Vert \nabla v\Vert _{1,2} +\omega \Vert ik v + {{\,\mathrm{e}\,}}_1\wedge v - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla v \Vert _{2}&\le {c}_{12} \big (\Vert F\Vert _{6/5}+\Vert F\Vert _{2}\big ). \end{aligned}$$
(5.33)

Moreover, \(v\big | _{\partial \Omega }=0\). Let \(\varphi \in \mathrm {C}^{\infty }_0(\Omega )\). We choose \(m_0\in \mathbb {N}\) such that \({{\,\mathrm{supp}\,}}\varphi \) is contained in \(\Omega _{m_0/2}\). For \(m\ge m_0\) we have \(w^m=v^m\) on \(\Omega _{m_0/2}\) and thus

$$\begin{aligned} \int _{\Omega } w^m \cdot \nabla \varphi \,{\mathrm d}x=\int _{\Omega } v^m \cdot \nabla \varphi \,{\mathrm d}x= 0 \end{aligned}$$

by (5.31)\(_{2}\). Passing to the limit \(m\rightarrow \infty \), we conclude \({{\,\mathrm{div}\,}}v=0\). Now let \(\psi \in \mathrm {C}^{\infty }_{0,\sigma }(\Omega )\) and choose \(m_0\) such that \({{\,\mathrm{supp}\,}}\psi \subset \Omega _{m_0/2}\). With the same argument as above, for \(m\ge m_0\) we obtain from (5.31)\(_{1}\) that

$$\begin{aligned}&\int _{\Omega } \big ( \omega (ik w^m + {{\,\mathrm{e}\,}}_1\wedge w^m - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla w^m ) - \Delta w^m - \lambda \partial _1 w^m -F\big )\cdot \psi \,{\mathrm d}x\\&= \int _{\Omega } \big ( \omega (ik v^m + {{\,\mathrm{e}\,}}_1\wedge v^m - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla v^m ) - \Delta v^m - \lambda \partial _1 v^m +\nabla p^m -F\big )\cdot \psi \,{\mathrm d}x=0. \end{aligned}$$

Therefore, by passing to the limit \(m\rightarrow \infty \), we see

$$\begin{aligned} \int _{\Omega } \big ( \omega (ik v + {{\,\mathrm{e}\,}}_1\wedge v - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla v ) - \Delta v- \lambda \partial _1 v-F\big )\cdot \psi \,{\mathrm d}x=0 \end{aligned}$$

for all \(\psi \in \mathrm {C}^{\infty }_{0,\sigma }(\Omega )\). Consequently, by the Helmholtz decomposition, there exists a function \(p\) with \(\nabla p\in \mathrm {L}^{2}(\Omega )\) such that \((v,p)\) is a solution to (2.3).

It remains to show that \(v\) and \(p\) belong to the correct function spaces. By Hölder’s inequality, we directly find that

$$\begin{aligned} v\in \mathrm {W}^{2,q}(\Omega _{\rho }), \qquad p\in \mathrm {W}^{1,q}(\Omega _{\rho }) \end{aligned}$$
(5.34)

for any \(\rho >R\) and all \(q\in [1,2]\). Repeating the “cut-off” argument from (5.11), we obtain \((u,\mathfrak {p})\) which satisfy (5.12) for some function \(f\in \mathrm {L}^{2}(\mathbb {R}^3)\) with compact support. In particular, this implies \(f\in \mathrm {L}^{q}(\mathbb {R}^3)\) for all \(q\in (1,2)\). Theorem 5.5 yields existence of a solution to (5.12) satisfying (5.8). Since \(u\in \mathrm {L}^{6}(\mathbb {R}^3)\), Theorem 5.5 further ensures that \((u,\mathfrak {p})\) coincides with this solution. We thus have

$$\begin{aligned} \begin{aligned}&ik u + {{\,\mathrm{e}\,}}_1\wedge u - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla u ,\, \nabla ^2u,\, \partial _1u,\, \nabla \mathfrak {p}\in \mathrm {L}^{q}(\mathbb {R}^3),\\&u\in \mathrm {L}^{2q/(2-q)}(\mathbb {R}^3),\qquad \nabla u\in \mathrm {L}^{4q/(4-q)}(\mathbb {R}^3). \end{aligned} \end{aligned}$$

Since \(v=u\) and \(p=\mathfrak {p}\) on \(\mathrm {B}^{2R}\), the integrability properties above in combination with (5.34) show that \(v\) and \(p\) belong to the correct function spaces. \(\square \)

Combining Lemmas 5.6, 5.8 and 5.11, we can finally complete the proof of Theorem 2.1.

Proof of Theorem 2.1

The uniqueness statement is a direct consequence of Lemma 5.6. Estimate (2.4) has been proved in Lemma 5.8. It thus remains to show existence of a solution for \(F\in \mathrm {L}^{q}(\Omega )\). Consider a sequence \((F_j)\subset \mathrm {C}^{\infty }_0(\Omega )\) that converges to \(F\) in \(\mathrm {L}^{q}(\Omega )\). By Lemma 5.11, for each \(j\in \mathbb {N}\) there exists a solution \((v,p)=(v_j,p_j)\) to (2.3) with \(F=F_j\), which obeys estimate (2.4) by Lemma 5.8. Additionally, this implies that \((v_j,\nabla p_j)\) is a Cauchy sequence in the function space defined by the norm on the left-hand side of (2.4), and thus possesses a limit \((v,\nabla p)\), which satisfies (2.3) and (2.4). \(\square \)

5.5 The Time-Periodic Linear Problem

Proof of Theorem 2.2

An application of the Fourier transform \(\mathscr {F}_{\mathbb T}\) on \({\mathbb T}\) to (2.2) reduces the uniqueness statement to the corresponding uniqueness result for the resolvent problem established in Theorem 2.1. To show existence, consider \(f\in \mathrm {A}({\mathbb T};\mathrm {L}^{q}(\Omega ))\). Then

$$\begin{aligned} f(t,x)= \sum _{k\in \mathbb {Z}} f_k(x) {{\,\mathrm{e}\,}}^{ikt} \end{aligned}$$

with \(f_k\in \mathrm {L}^{q}(\Omega )\). Let \((u_k,\mathfrak {p}_k)=(v,p)\) be a solution to the resolvent problem (2.3) with \(F=f_k\) that exists due to Theorem 2.1. We define

$$\begin{aligned} u(t,x):=\sum _{k\in \mathbb {Z}} u_k(x) {{\,\mathrm{e}\,}}^{ikt}, \qquad \mathfrak {p}(t,x):=\sum _{k\in \mathbb {Z}} \mathfrak {p}_k(x) {{\,\mathrm{e}\,}}^{ikt}. \end{aligned}$$

By (2.4), \(u\) and \(\mathfrak {p}\) are well defined and satisfy (2.2). We directly conclude estimate (2.5) from estimate (2.4). \(\square \)

6 The Nonlinear Problem

We return to the nonlinear problem (2.1). At first, we reformulate it as a problem with homogeneous boundary conditions. To this end, fix \(R>0\) such that \(\partial \mathrm {B}_R\subset \Omega \). Let \(\varphi \in \mathrm {C}^{\infty }_0(\mathbb {R}^3)\) be a smooth function satisfying \(\varphi (x)=1\) if \({|x |}<R\), and \(\varphi (x)=0\) if \({|x |}>2R\), and define

$$\begin{aligned} U:{\mathbb T}\times \mathbb {R}^3\rightarrow \mathbb {R}^3, \qquad U(t,x)=\frac{1}{2}{{\,\mathrm{rot}\,}}\big [\big (\alpha (t){{\,\mathrm{e}\,}}_1\wedge x-\omega {{\,\mathrm{e}\,}}_1{|x |}^2\big )\varphi (x)\big ]. \end{aligned}$$

Then \(U(t,\cdot )\in \mathrm {C}^{\infty }_0(\mathbb {R}^3)^3\) for all \(t\in {\mathbb T}\), \(U\in \mathrm {C}^{1}({\mathbb T}\times \mathbb {R}^3)\), \({{\,\mathrm{div}\,}}U=0\), and a brief calculation shows \(U(t,x)=\alpha (t){{\,\mathrm{e}\,}}_1+\omega {{\,\mathrm{e}\,}}_1\wedge x\) for \((t,x)\in {\mathbb T}\times \partial \Omega \). Now define \(v:=u-U\) and \(p:=\mathfrak {p}\). Then \((u,\mathfrak {p})\) solves (2.1) if and only if \((v,p)\) solves

$$\begin{aligned} \left\{ \begin{array}{r@{\quad }l} \omega (\partial _t v + {{\,\mathrm{e}\,}}_1\wedge v - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla v ) -\Delta v- \lambda \partial _1 v+\nabla p= f+{\mathcal N}(v) &{}\quad \text {in }{\mathbb T}\times \Omega , \\ {{\,\mathrm{div}\,}}v=0 &{}\quad \text {in }{\mathbb T}\times \Omega , \\ v=0 &{}\quad \text {on }{\mathbb T}\times \partial \Omega , \\ \lim _{{|x |}\rightarrow \infty } v(t,x) = 0 &{}\quad \text {for }t\in {\mathbb T}, \end{array}\right. \end{aligned}$$
(6.1)

where

$$\begin{aligned} {\mathcal N}(v)&:=({\mathcal P}_\bot \alpha )\partial _1v-\omega (\partial _t U + {{\,\mathrm{e}\,}}_1\wedge U - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla U )\\&\quad +\Delta U+\alpha \partial _1U-v\cdot \nabla v-U\cdot \nabla v-v\cdot \nabla U-U\cdot \nabla U. \end{aligned}$$

Recall that \({\mathcal P}_\bot \alpha =\alpha -\lambda \). It thus remains to show existence of a solution to the nonlinear system (6.1).

Proof of Theorem 2.3

We define the function space

$$\begin{aligned} {\mathcal X}^q&:=\bigl \{v\in \mathrm {L}^{1}_{\mathrm {loc}}({\mathbb T}\times \Omega )\ \big \vert \ \Vert v\Vert _{{\mathcal X}^q}<\infty \bigr \}, \\ \Vert v\Vert _{{\mathcal X}^q}&:=\omega \Vert \partial _t v + {{\,\mathrm{e}\,}}_1\wedge v - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla v \Vert _{\mathrm {A}^{q}} +\Vert \nabla ^2v\Vert _{\mathrm {A}^q} \\&\qquad +\lambda \Vert \partial _1v\Vert _{\mathrm {A}^q} +\lambda ^{1/2}\Vert v\Vert _{\mathrm {A}^{s_1}} +\lambda ^{1/4}\Vert \nabla v\Vert _{\mathrm {A}^{s_2}}, \end{aligned}$$

where \(s_1=2q/(2-q)\), \(s_2=4q/(4-q)\) and

$$\begin{aligned} \Vert h\Vert _{\mathrm {A}^s}:=\Vert h\Vert _{\mathrm {A}({\mathbb T};\mathrm {L}^{s}(\Omega ))}. \end{aligned}$$

At first, we derive suitable estimates of \({\mathcal N}(v)\). For example, analogously to the proof of Proposition 3.1, we have

$$\begin{aligned} \Vert ({\mathcal P}_\bot \alpha )\partial _1v\Vert _{\mathrm {A}^q} \le \Vert {\mathcal P}_\bot \alpha \Vert _{\mathrm {A}({\mathbb T};\mathbb {R})}\Vert \partial _1v\Vert _{\mathrm {A}^q} \le \varepsilon \Vert \partial _1v\Vert _{\mathrm {A}^q} \le \varepsilon \lambda ^{-1} \Vert v\Vert _{{\mathcal X}^q}. \end{aligned}$$

Moreover, since \(\frac{2q}{2-q}\le 4\le \frac{3q}{3-2q}\), we can employ estimates (3.2) and (3.3) to obtain

$$\begin{aligned} \Vert v\cdot \nabla v\Vert _{\mathrm {A}^q} \le \Vert v\Vert _{\mathrm {A}^{4}}\Vert \nabla v\Vert _{\mathrm {A}^{4q/(4-q)}} \le {c}_{0}\Vert v\Vert _{\mathrm {A}^{2q/(2-q)}}^{1-\theta } \Vert v\Vert _{\mathrm {A}^{3q/(3-2q)}}^{\theta } \Vert \nabla v\Vert _{\mathrm {A}^{4q/(4-q)}} \end{aligned}$$

with \(\theta =\frac{12-9q}{2q}\). By the Sobolev inequality we thus deduce

$$\begin{aligned} \Vert v\cdot \nabla v\Vert _{\mathrm {A}^q} \le {c}_{1}\lambda ^{-1/4-(1-\theta )/2}\Vert v\Vert _{{\mathcal X}^q}^{2-\theta } \Vert \nabla ^2v\Vert _{\mathrm {A}^{q}}^\theta \le {c}_{2}\lambda ^{-(3q-3)/q}\Vert v\Vert _{{\mathcal X}^q}^2. \end{aligned}$$

The remaining terms in \({\mathcal N}(v)\) can be estimated in a similar fashion, which results in

$$\begin{aligned} \begin{aligned} \Vert {\mathcal N}(v)\Vert _{\mathrm {A}^q}&\le {c}_{3}\big ( \varepsilon \lambda ^{-1}\Vert v\Vert _{{\mathcal X}^q} +\lambda ^{-(3q-3)/q}\Vert v\Vert _{{\mathcal X}^q}^2 +\omega \Vert \tfrac{{\mathrm d}}{{\mathrm d}t}\alpha \Vert _{\mathrm {A}({\mathbb T};\mathbb {R})}\\&\qquad \qquad \qquad \qquad \qquad +(\lambda +\omega +\varepsilon ) (1+\lambda +\omega +\varepsilon +\Vert v\Vert _{{\mathcal X}^q})\big ). \end{aligned} \end{aligned}$$
(6.2)

Now consider the problem

$$\begin{aligned} \left\{ \begin{array}{r@{\quad }l} \omega (\partial _t w + {{\,\mathrm{e}\,}}_1\wedge w - {{\,\mathrm{e}\,}}_1\wedge x\cdot \nabla w ) -\Delta w- \lambda \partial _1 w+\nabla \mathfrak {q}= f+{\mathcal N}(v) &{}\quad \text {in }{\mathbb T}\times \Omega , \\ {{\,\mathrm{div}\,}}w=0 &{}\quad \text {in }{\mathbb T}\times \Omega , \\ w=0 &{}\quad \text {on }{\mathbb T}\times \partial \Omega , \end{array}\right. \end{aligned}$$
(6.3)

for given \(v\in {\mathcal X}^q\). Due to estimate (6.2) and Theorem 2.2 there exists a unique velocity field \(w\in {\mathcal X}^q\) and a pressure field \(\mathfrak {q}\) with \(\nabla \mathfrak {q}\in \mathrm {A}^q\) that satisfy (6.3) and the estimate

$$\begin{aligned} \Vert w\Vert _{{\mathcal X}^q}&\le {C}_{1} \big (\Vert f\Vert _{\mathrm {A}^q}+\Vert {\mathcal N}(v)\Vert _{\mathrm {A}^q}\big )\\&\le {c}_{4}\big ( \varepsilon +\varepsilon \lambda ^{-1}\Vert v\Vert _{{\mathcal X}^q} +\lambda ^{-(3q-3)/q}\Vert v\Vert _{{\mathcal X}^q}^2 +\omega \Vert \tfrac{{\mathrm d}}{{\mathrm d}t}\alpha \Vert _{\mathrm {A}({\mathbb T};\mathbb {R})}\\&\qquad \qquad \qquad \qquad \qquad +(\lambda +\omega +\varepsilon ) (1+\lambda +\omega +\varepsilon +\Vert v\Vert _{{\mathcal X}^q})\big ). \end{aligned}$$

We thereby obtain a solution map \({\mathcal S}:{\mathcal X}^q\rightarrow {\mathcal X}^q\), \(v\mapsto w\) which is a self-mapping on the ball

$$\begin{aligned} {\mathcal X}^q_\delta :=\bigl \{v\in {\mathcal X}^q\ \big \vert \ \Vert v\Vert _{{\mathcal X}^q}\le \delta \bigr \} \end{aligned}$$

provided

$$\begin{aligned} {c}_{4}\big ( \varepsilon +\varepsilon \lambda ^{-1}\delta +\lambda ^{-(3q-3)/q}\delta ^2 +\omega \Vert \tfrac{{\mathrm d}}{{\mathrm d}t}\alpha \Vert _{\mathrm {A}({\mathbb T};\mathbb {R})} +(\lambda +\omega +\varepsilon )(1+\lambda +\omega +\varepsilon +\delta )\big ) \le \delta . \end{aligned}$$

Recall that \(\rho \in \big (\frac{3q-3}{q},1\big )\). Choosing \(\delta :=\lambda ^\rho \), one readily verifies that there is a constant \(\kappa >0\) depending on \({c}_{4}\) such the condition above is satisfied with \(\omega \Vert \tfrac{{\mathrm d}}{{\mathrm d}t}\alpha \Vert _{\mathrm {A}({\mathbb T};\mathbb {R})} \le \kappa \lambda ^\rho \), \(\varepsilon =\lambda ^2\) and \(\lambda _0\) sufficiently small. In the same way, one derives the estimate

$$\begin{aligned} \Vert {\mathcal N}(v_1)-{\mathcal N}(v_2)\Vert _{\mathrm {A}^q} \le {c}_{5} \big (\varepsilon \lambda ^{-1}+\lambda +\omega +\varepsilon +\lambda ^{-(3q-3)/q}(\Vert v_1\Vert _{{\mathcal X}^q}+\Vert v_2\Vert _{{\mathcal X}^q})\big ) \Vert v_1-v_2\Vert _{{\mathcal X}^q}, \end{aligned}$$

which ensures that \({\mathcal S}\) is a contraction on \({\mathcal X}^q_\delta \) with a similar choice of parameters. Finally, the contraction mapping principle yields the existence of a fixed point \(v\in {\mathcal X}^q\) of \({\mathcal S}\), and hence of a solution \((v,p)\) to (6.1). Consequently, \((u,\mathfrak {p}):=(v+U,p)\) is a solution to (2.1). \(\square \)