Skip to main content
Log in

Analysis of Micropolar Fluids: Existence of Potential Microflow Solutions, Nearby Global Well-Posedness, and Asymptotic Stability

  • Published:
Acta Applicandae Mathematicae Aims and scope Submit manuscript

Abstract

In this paper we concern ourselves with an incompressible, viscous, isotropic, and periodic micropolar fluid. We find that in the absence of forcing and microtorquing there exists an infinite family of well-behaved solutions, which we call potential microflows, in which the fluid velocity vanishes identically, but the angular velocity of the microstructure is conservative and obeys a linear parabolic system. We then prove that nearby each potential microflow, the nonlinear equations of motion are well-posed globally-in-time, and solutions are stable. Finally, we prove that in the absence of force and microtorque, solutions decay exponentially, and in the presence of force and microtorque obeying certain conditions, solutions have quantifiable decay rates.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

References

  1. Allen, S.J., Kline, K.A.: Lubrication theory for micropolar fluids. J. Appl. Mech. 38(3), 646 (1971)

    Article  Google Scholar 

  2. Abraham, R., Marsden, J.E., Ratiu, T.: Manifolds, Tensor Analysis, and Applications, 2nd edn. Applied Mathematical Sciences, vol. 75. Springer, New York (1988)

    Book  Google Scholar 

  3. Bég, O., Bhargava, R., Rawat, S., Halim, K., Takhar, H.S.: Computational modeling of biomagnetic micropolar blood flow and heat transfer in a two-dimensional non-Darcian porous medium. Meccanica 43(4), 391–410 (2008)

    Article  MathSciNet  Google Scholar 

  4. Bayada, G., Łukaszewicz, G.: On micropolar fluids in the theory of lubrication. Rigorous derivation of an analogue of the Reynolds equation. Int. J. Eng. Sci. 34(13), 1477–1490 (1996)

    Article  MathSciNet  Google Scholar 

  5. Chen, Q., Miao, C.: Global well-posedness for the micropolar fluid system in critical Besov spaces. J. Differ. Equ. 252(3), 2698–2724 (2012)

    Article  MathSciNet  Google Scholar 

  6. Eringen, A.: Theory of micropolar fluids. J. Math. Mech. 16, 1–18 (1966)

    MathSciNet  Google Scholar 

  7. Eringen, A.: Microcontinuum Field Theories. I. Foundations and Solids. Springer, New York (1999)

    Book  Google Scholar 

  8. Eringen, A.: Microcontinuum Field Theories: II. Fluent Media, vol. 2. Springer, New York (2001)

    MATH  Google Scholar 

  9. Folland, G.: Introduction to Partial Differential Equations, 2nd edn. Princeton University Press, Princeton (1995)

    MATH  Google Scholar 

  10. Ferreira, L., Villamizar-Roa, E.J.: Micropolar fluid system in a space of distributions and large time behavior. J. Math. Anal. Appl. 332(2), 1425–1445 (2007)

    Article  MathSciNet  Google Scholar 

  11. Gay-Balmaz, F., Ratiu, T., Tronci, C.: Equivalent theories of liquid crystal dynamics. Arch. Ration. Mech. Anal. 210(3), 773–811 (2013)

    Article  MathSciNet  Google Scholar 

  12. Galdi, G., Rionero, S.: A note on the existence and uniqueness of solutions of the micropolar fluid equations. Int. J. Eng. Sci. 15(2), 105–108 (1977)

    Article  MathSciNet  Google Scholar 

  13. Lions, J.-L., Magenes, E.: Non-homogeneous Boundary Value Problems and Applications. Vol. I. Die Grundlehren der mathematischen Wissenschaften, vol. 181. Springer, New York (1972). Translated from the French by P. Kenneth

    Book  Google Scholar 

  14. Lhuillier, D., Rey, A.: Nematic liquid crystals and ordered micropolar fluids. J. Non-Newton. Fluid Mech. 120(1–3), 169–174 (2004)

    Article  Google Scholar 

  15. Łukaszewicz, G.: On the existence, uniqueness and asymptotic properties for solutions of flows of asymmetric fluids. Rend. Accad. Naz. Sci. Detta Accad. XL, Parte I, Mem. Mat. 13(1), 105–120 (1989)

    MathSciNet  MATH  Google Scholar 

  16. Łukaszewicz, G.: On non-stationary flows of incompressible asymmetric fluids. Math. Methods Appl. Sci. 13(3), 219–232 (1990)

    Article  MathSciNet  Google Scholar 

  17. Łukaszewicz, G.: Micropolar Fluids. Modeling and Simulation in Science, Engineering and Technology. Birkhäuser Boston, Inc., Boston (1999)

    Google Scholar 

  18. Łukaszewicz, G.: Long time behavior of 2d micropolar fluid flows. Math. Comput. Model. 34(5–6), 487–509 (2001)

    Article  MathSciNet  Google Scholar 

  19. Mekheimer, Kh.S., El Kot, M.A.: The micropolar fluid model for blood flow through a tapered artery with a stenosis. Acta Mech. Sin. 24(6), 637–644 (2008)

    Article  MathSciNet  Google Scholar 

  20. Nicodemus, E., Sharma, S.: Performance characteristics of micropolar lubricated membrane-compensated worn hybrid journal bearings. Tribol. Trans. 55(1), 59–70 (2012)

    Article  Google Scholar 

  21. Nochetto, R., Salgado, A., Tomas, I.: The equations of ferrohydrodynamics: modeling and numerical methods. Math. Models Methods Appl. Sci. 26(13), 2393–2449 (2016)

    Article  MathSciNet  Google Scholar 

  22. Ramkissoon, H.: Flow of a micropolar fluid past a Newtonian fluid sphere. Z. Angew. Math. Mech. 65(12), 635–637 (1985)

    Article  MathSciNet  Google Scholar 

  23. Rojas-Medar, M.A., Ortega-Torres, E.E.: The equations of a viscous asymmetric fluid: an interactive approach. Z. Angew. Math. Mech. 85(7), 471–489 (2005)

    Article  MathSciNet  Google Scholar 

  24. Sinha, P., Singh, C., Prasad, K.R.: Lubrication of human joints – a microcontinuum approach. Wear 80(2), 159–181 (1982)

    Article  Google Scholar 

  25. Villamizar-Roa, E.J., Rodríguez-Bellido, M.A.: Global existence and exponential stability for the micropolar fluid system. Z. Angew. Math. Phys. 59(5), 790–809 (2008)

    Article  MathSciNet  Google Scholar 

  26. Yamaguchi, N.: Existence of global strong solution to the micropolar fluid system in a bounded domain. Math. Methods Appl. Sci. 28(13), 1507–1526 (2005)

    Article  MathSciNet  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Ian Tice.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

I. Tice was supported by an NSF CAREER Grant (DMS #1653161). N. Stevenson was supported by the summer research support provided by this grant.

Appendices

Appendix A: Reduction to Velocity Fields with Vanishing Average

The natural setting for the initial data, force, and microtorque in (1.1) is the space

$$\begin{aligned} &H^{1+s}_{\perp }\left (\mathbb{T}^{3};\mathbb{R}^{3}\right )\times H^{1+s} \left (\mathbb{T}^{3};\mathbb{R}^{3}\right )\times L^{2}\left ( \mathbb{R}^{+};\mathring{H}^{s}\left (\mathbb{T}^{3};\mathbb{R}^{3} \right )\right )\times L^{2}\left (\mathbb{R}^{+};H^{s}\left ( \mathbb{T}^{3};\mathbb{R}^{3}\right )\right ) \\ &\quad \text{ for } s\in \mathbb{R}^{+}\cup \left \{ {0}\right \} . \end{aligned}$$
(A.1)

However, as we will see below, the corresponding solutions are globally integrable in time if and only if \(u_{0}\) has vanishing average. It is thus convenient to introduce a change of unknowns that allows us to reduce to studying this case. This is possible due to the invariance of the micropolar equations (1.1) under Galilean transformations.

Lemma 1.1

Let \(s\in \mathbb{R}^{+}\cup \left \{ {0}\right \} \). Suppose that \(u_{0},\omega _{0}\in H^{1+s}\left (\mathbb{T}^{3};\mathbb{R}^{3} \right )\) are initial data with \(\operatorname{div}u_{0}=0\), \(f,g\in L^{2}\left (\mathbb{R}^{+};H^{s}\left (\mathbb{T}^{3}; \mathbb{R}^{3}\right )\right )\) are forcing and microtorquing with \(\int _{\mathbb{T}^{3}}\!f\!\left (t,x\right )\mathrm{d}x\!=0\) for almost every \(t\in \mathbb{R}^{+}\). Suppose that \(\left (u,\omega ,p\right )\) are a corresponding strong solution triple to system (1.1). Then the following hold.

  1. 1.

    For almost every \(t\in \mathbb{R}^{+}\) it holds that \(\int _{\mathbb{T}^{3}}u\left (t,x\right )\;\mathrm{d}x=\int _{ \mathbb{T}^{3}}u_{0}\left (x\right )\;\mathrm{d}x\).

  2. 2.

    Given any \(b\in \mathbb{R}^{3}\), the triple \(\left (v,\chi ,q\right ): \mathbb{R}^{+} \times \mathbb{T}^{3} \to \mathbb{R}^{3} \times \mathbb{R}^{3} \times \mathbb{R}\) defined by

    $$ \textstyle\begin{cases} v\left (t,x\right )=-b+u\left (t,x+tb\right ) \\ \chi \left (t,x\right )=\omega \left (t,x+tb\right ) \\ q\left (t,x\right )=p\left (t,x+tb\right ) \\ \varphi \left (t,x\right )=f\left (t,x+tb\right ) \\ \psi \left (t,x\right )=g\left (t,x+tb\right ) \end{cases} $$
    (A.2)

    is a strong solution to system (1.1) with initial data \(\left (u_{0}-b,\omega _{0}\right )\) and forcing/microtorquing \(\left (\varphi ,\psi \right )\). Moreover, if we posit the space-time regularity of \(\left (u,\omega ,p\right )\) to be encoded with the inclusion:

    $$\begin{aligned} &\left (u-\int _{\mathbb{T}^{3}}u_{0},\omega ,p\right )\in \left (L^{2} \left (\mathbb{R}^{+};\mathring{H}^{2+s}_{\perp }\left (\mathbb{T}^{3}; \mathbb{R}^{3}\right )\right )\cap H^{1}\left (\mathbb{R}^{+}; \mathring{H}^{s}_{\perp }\left (\mathbb{T}^{3};\mathbb{R}^{3}\right ) \right )\right ) \\ &\quad {}\times \left (L^{2}\left (\mathbb{R}^{+};H^{2+s}\left (\mathbb{T}^{3}; \mathbb{R}^{3}\right )\right )\cap H^{1}\left (\mathbb{R}^{+};H^{s} \left (\mathbb{T}^{3};\mathbb{R}^{3}\right )\right )\right )\!\times\! \left (L^{2}\left (\mathbb{R}^{+};\mathring{H}^{1+s}\left ( \mathbb{T}^{3};\mathbb{R}^{3}\right )\!\right )\!\right ), \end{aligned}$$
    (A.3)

    then the same inclusion is true for \(\left (v-\int _{\mathbb{T}^{3}}\left (u_{0}-b\right ),\chi ,q \right )\). Also, \(\left (\varphi ,\psi \right )\) belong to the same space-time regularity class as \(\left (f,g\right )\).

As a consequence, to understand the solvability of (1.1) with respect to data and forcing / microtorquing quadruples belonging to the space

$$ H^{1+s}_{\perp }\left (\mathbb{T}^{3};\mathbb{R}^{3}\right )\times H^{1+s} \left (\mathbb{T}^{3};\mathbb{R}^{3}\right )\times L^{2}\left ( \mathbb{R}^{+};\mathring{H}^{s}\left (\mathbb{T}^{3};\mathbb{R}^{3} \right )\right )\times L^{2}\left (\mathbb{R}^{+};H^{s}\left ( \mathbb{T}^{3};\mathbb{R}^{3}\right )\right ), $$
(A.4)

it is sufficient to understand the system’s solvability for data/forcing/microtorquing belonging to the smaller space

$$ \mathring{H}^{1+s}_{\perp }\left (\mathbb{T}^{3};\mathbb{R}^{3} \right )\times H^{1+s}\left (\mathbb{T}^{3};\mathbb{R}^{3}\right ) \times L^{2}\left (\mathbb{R}^{+};\mathring{H}^{s}\left ( \mathbb{T}^{3};\mathbb{R}^{3}\right )\right )\times L^{2}\left ( \mathbb{R}^{+};H^{s}\left (\mathbb{T}^{3};\mathbb{R}^{3}\right ) \right ). $$
(A.5)

Proof

Suppose that we are given a strong solution triple \(\left (u,\omega ,p\right )\) corresponding to the data / forcing / microtorquing quadruple \(\left (u_{0},\omega _{0},f,g\right )\) as in the hypotheses. Averaging the second equation in (1.1), and then integration by parts yields the identity for almost everywhere on \(\mathbb{R}^{+}\):

$$\begin{aligned} 0 =&\int _{\mathbb{T}^{3}}\varrho \left (\partial _{t}u+u\cdot \nabla u \right )-\left (\varepsilon +\frac{\kappa }{2}\right )\Delta u- \kappa \operatorname{curl}\omega +\nabla p \\ =&\varrho \left (\int _{\mathbb{T}^{3}}u \right )'+\int _{\mathbb{T}^{3}}u\operatorname{div}u=\varrho \left (\int _{ \mathbb{T}^{3}}u\right )'. \end{aligned}$$
(A.6)

Thus, the first item now follows from the fundamental theorem of calculus.

Now, if \(\mu \in \mathbb{N}^{2}\) is a multi-index with \(\left |\mu \right |\le 2\), then for almost every \(\left (t,x\right )\in \mathbb{R}^{+}\times \mathbb{T}^{3}\) we have that \(\partial ^{\mu }v\left (t,x\right )=\partial ^{\mu }u\left (t,x+tb \right )\) and \(\partial ^{\mu }\chi \left (t,x\right )=\partial ^{\mu }\omega \left (t,x+tb \right )\), and \(\nabla q\left (t,x\right )=\nabla p\left (t,x+tb\right )\). We next compute the discrepancy in the time derivatives: \(\partial _{t}v\left (t,x\right )=\partial _{t}u\left (t,x+tb\right )+b \cdot \nabla u\left (t,x+tb\right )\), and \(\partial _{t}\chi \left (t,x\right )=\partial _{t}\omega \left (t,x \right )+b\cdot \nabla \omega \left (t,x+tb\right )\). Finally, using the previous remarks we verify the following equality between the two material derivative terms:

$$ \textstyle\begin{cases} \left (\partial _{t}+v\left (t,x\right )\cdot \nabla \right )v\left (t,x \right )=\partial _{t}u\left (t,x+tb\right )+u\left (t,x+tb\right ) \cdot \nabla u\left (t,x+tb\right ) \\ \left (\partial _{t}+v\left (t,x\right )\cdot \nabla \right )v\left (t,x \right )=\partial _{t}\omega \left (t,x+tb\right )+u\left (t,x+tb \right )\cdot \nabla \omega \left (t,x+tb\right ). \end{cases} $$
(A.7)

Thus, evaluation of the system (1.1) at points \(\left (t,x+tb\right )\) for \(\left (t,x\right )\in \mathbb{R}^{+}\times \mathbb{T}^{3}\) yields that the triple \(\left (v,\chi ,q\right )\) is a strong solution to (1.1) with data \(\left (u_{0}-b,\omega _{0}\right )\) and forcing/microtorquing \(\left (\varphi ,\psi \right )\). The space-time regularity assertions in the latter part of the second item are now a trivial verification given that the first item shows that the velocity’s average is constant in time. The consequence follows by: changing coordinates and unknowns, taking \(b\) equal to the spatial average of \(u_{0}\); solving the new system using the hypothesized solvability of (1.1) in the average-zero spaces; and finally transforming back with \(b\) equal to the negative of the average of \(u_{0}\) and citing the conclusion of the second item. □

Appendix B: Tools from Analysis

In this section we record a number of analysis results used throughout the paper.

2.1 B.1 Real-Valued Distributions and Fourier Coefficients

The following lemma characterizes when a distribution \(T \in \mathcal{D}^{\ast }\left (\mathbb{T}^{d};\mathbb{C}^{\ell }\right )\) is actually \(\mathbb{R}^{\ell }\)-valued. To state the result we recall a few definitions. First, for a sequence \(z: \mathbb{Z}^{d} \to \mathbb{C}^{\ell }\) we define its reflection \(Rz : \mathbb{Z}^{d} \to \mathbb{C}^{\ell }\) via \(Rz(k) = z(-k)\) for \(k\in \mathbb{Z}^{d}\). Second, for a distribution \(T \in \mathcal{D}^{\ast }\left (\mathbb{T}^{d};\mathbb{C}^{\ell }\right )\) we define its complex conjugate as the distribution \(\overline{T}\in \mathcal{D}^{\ast }\left (\mathbb{T}^{d}; \mathbb{C}^{\ell }\right )\) given by \(\left \langle {\overline{T}, \psi }\right \rangle = \overline{\left \langle {T,\overline{\psi }}\right \rangle }\) for each \(\psi \in \mathcal{D}(\mathbb{T}^{d};\mathbb{C})\).

Lemma B.1

Let \(\ell ,d \in \mathbb{N}^{+}\) The following hold.

  1. 1.

    If \(f \in L^{2}(\mathbb{T}^{d};\mathbb{C}^{\ell })\), then \(f\) is \(\mathbb{R}^{\ell }\)-valued, i.e. \(f = \overline{f}\), if and only if \(\hat{f} \in \ell ^{2}(\mathbb{Z}^{d};\mathbb{C}^{\ell })\) satisfies \(\overline{\hat{f}} = R \hat{f}\).

  2. 2.

    If \(T \in \mathcal{D}^{\ast }\left (\mathbb{T}^{d};\mathbb{C}^{\ell }\right )\), then \(T\) is \(\mathbb{R}^{\ell }\)-valued, i.e. \(T = \overline{T}\), if and only if \(\overline{\hat{T}} = R \hat{T}\) holds on the lattice \(\mathbb{Z}^{d}\).

Proof

If \(f = \overline{f}\), then we have that

$$ \overline{\hat{f}(k)} = \int _{\mathbb{T}^{d}} \overline{f(x)} \mathbf{e}_{k}\left (x\right )\;\mathrm{d}x = \int _{\mathbb{T}^{d}} f(x) \mathbf{e}_{k}\left (x\right )\mathrm{d}x = \hat{f}(-k), $$
(B.1)

and if \(\overline{\hat{f}} = R \hat{f}\), then for almost every \(x\in \mathbb{T}^{d}\)

$$ \overline{f(x)} = \sum _{k \in \mathbb{Z}^{d}} \overline{\hat{f}(k)} \mathbf{e}_{-k}\left (x\right ) = \sum _{k \in \mathbb{Z}^{d}} \hat{f}(-k)\mathbf{e}_{-k}\left (x\right ) = \sum _{k \in \mathbb{Z}^{d}} \hat{f}(k)\mathbf{e}_{k} = f(x). $$
(B.2)

This proves the first item.

We now turn to the proof of the second item. If \(\psi \in \mathcal{D}\left (\mathbb{T}^{d};\mathbb{C}\right )\), then the series \(\sum _{k\in \mathbb{Z}^{d}}\hat{\psi }\left (k\right )\mathbf{e}_{k}\) converges absolutely in the topology of \(H^{s}\left (\mathbb{T}^{3};\mathbb{C}^{\ell }\right )\) for all \(s\in \mathbb{R}\). In particular, for each \(m\in \mathbb{N}\) we have the convergence:

$$ \lim _{K\to \infty }\left [{\psi -\sum _{ \substack{k\in \mathbb{Z}^{d}\\\left |k\right |\le K}}\hat{\psi } \left (k\right )\mathbf{e}_{k}}\right ]_{m}=0, $$
(B.3)

where \(\left [{\cdot }\right ]_{m}\) is the seminorm from Definition 1.1. Thus, given \(T\in \mathcal{D}^{\ast }\left (\mathbb{T}^{d};\mathbb{C}^{\ell }\right )\) such that \(\overline{\hat{T}}=R\hat{T}\), we can compute its action on \(\psi \) via the series:

$$\begin{aligned} &\left \langle {T,\psi }\right \rangle \\ &\quad =\lim _{K\to \infty }\sum _{ \substack{k\in \mathbb{Z}^{d}\\\left |k\right |\le K}}\hat{\psi } \left (k\right )\left \langle {T,\mathbf{e}_{k}}\right \rangle = \lim _{K\to \infty }\sum _{ \substack{k\in \mathbb{Z}^{d}\\\left |k\right |\le K}}\hat{\psi } \left (k\right )\hat{T}\left (-k\right ) \\ &\quad =\lim _{K\to \infty }\sum _{ \substack{k\in \mathbb{Z}^{d}\\\left |k\right |\le K}}\hat{\psi } \left (k\right ) R\hat{T}\left (k\right )=\lim _{K\to \infty }\sum _{ \substack{k\in \mathbb{Z}^{d}\\\left |k\right |\le K}}\hat{\psi } \left (k\right ) \overline{\hat{T}\left (k\right )} \\ &\quad =\lim _{K\to \infty } \overline{\sum _{\substack{k\in \mathbb{Z}^{d}\\\left |k\right |\le K}}\hat{\overline{\psi }}\left (-k\right )\hat{T}\left (k\right )}= \lim _{K\to \infty } \overline{\sum _{\substack{k\in \mathbb{Z}^{d}\\\left |k\right |\le K}}\hat{\overline{\psi }}\left (-k\right )\left \langle {T,\mathbf{e}_{-k}}\right \rangle }= \overline{\left \langle {T,\lim _{K\to \infty }\sum _{\substack{k\in \mathbb{Z}^{d}\\\left |k\right |\le K}}\hat{\overline{\psi }}\left (k\right )\mathbf{e}_{k}}\right \rangle } \\ &\quad = \overline{\left \langle {T,\overline{\psi }}\right \rangle } =\left \langle {\overline{T},\psi }\right \rangle . \end{aligned}$$
(B.4)

This gives the sufficient condition for the second item. To see that it is also necessary, suppose now that we have \(T\in \mathcal{D}^{\ast }\left (\mathbb{T}^{d};\mathbb{C}^{\ell }\right )\) satisfying \(T=\overline{T}\). We compute for \(k\in \mathbb{Z}^{d}\):

$$ \overline{\hat{T}\left (k\right )}= \overline{\left \langle {T,\mathbf{e}_{-k}}\right \rangle }= \overline{\left \langle {T,\overline{\mathbf{e}_{k}}}\right \rangle }= \left \langle {\overline{T},\mathbf{e}_{k}}\right \rangle =\left \langle {T,\mathbf{e}_{k}}\right \rangle =\hat{T}\left (-k\right )=R \hat{T}\left (k\right ). $$
(B.5)

 □

2.2 B.2 Fractional Sobolev Spaces

Here we record some results in fractional Sobolev spaces, as defined in Definition 1.3.

Proposition B.2

Let \(s \in \mathbb{R}\). Then the map \(\Pi : H^{s}_{\|} \left (\mathbb{T}^{3};\mathbb{R}^{3}\right ) \to \mathring{H}^{1+s}\left (\mathbb{T}^{3};\mathbb{R}\right )\) defined by

$$ \Pi f = \sum _{k\in \mathbb{Z}^{3} \backslash \{0\}} \left ( \frac{k \cdot \hat{f}(k)}{2\pi i \left |k\right |^{2}} \right ) \mathbf{e}_{k} $$
(B.6)

is well-defined, bounded, and linear, and \(\nabla \Pi f = f\) for all \(f \in H^{s}_{\|} \left (\mathbb{T}^{3};\mathbb{R}^{3}\right )\). Moreover, \(\Pi \) extends to a bounded linear map from \(L^{2}\left (\mathbb{R}^{+};H^{s}_{\|} \left (\mathbb{T}^{3}; \mathbb{R}^{3}\right )\right )\) to \(L^{2}\left (\mathbb{R}^{+};\mathring{H}^{1+s}\left (\mathbb{T}^{3}; \mathbb{R}\right )\right )\) with the same properties.

Proof

First we compute

$$ \left \lVert \Pi f\right \rVert _{\mathring{H}^{1+s}}^{2} = \sum _{k \in \mathbb{Z}^{3} \backslash \{0\}} \left |k\right |^{2+ 2s} \left | \frac{k \cdot \hat{f}(k)}{2\pi i \left |k\right |^{2}} \right |^{2} \le \frac{1}{4\pi ^{2}} \sum _{k \in \mathbb{Z}^{3} \backslash \{0\}} \left |k\right |^{2s} \left | \hat{f}(k) \right |^{2} = \frac{1}{4\pi ^{2}} \left \lVert f\right \rVert _{\mathring{H}^{s}}^{2}, $$
(B.7)

and then we use Lemma B.1 to see that

$$ \overline{\widehat{\Pi f}(k)} = \frac{-k \cdot \overline{\hat{f}(k) } }{2\pi i \left |k\right |^{2} } = \frac{-k \cdot \hat{f}(-k) }{2\pi i \left |-k\right |^{2} } = \widehat{\Pi f}(-k), $$
(B.8)

which then implies that \(\Pi f\) is a real-valued distribution. From this we deduce that \(\Pi \) is a well-defined bounded linear map. Then for \(f \in H^{s}_{\|} \left (\mathbb{T}^{3};\mathbb{R}^{3}\right )\) we compute

$$ \widehat{\nabla \Pi f}(k) = 2\pi i k \frac{k \cdot \hat{f}(k)}{2\pi i \left |k\right |^{2}} = \frac{k \otimes k}{\left |k\right |^{2}} \hat{f}(k), $$
(B.9)

to deduce that \(\nabla \Pi f = (I- \mathbb{P}) f =f\). The extension result then follows trivially. □

Next we record a useful product estimate in fractional Sobolev spaces.

Proposition B.3

Let \(d\in \mathbb{N}^{+}\), and suppose that \(s,t\in \mathbb{R}^{+}\cup \left \{ {0}\right \} \) satisfy \(s>\frac{d}{2}\) and \(s\ge t\). Then there exists a constant \(C >0\), depending on \(s\) and \(t\), such that if \(f\in H^{s}\left (\mathbb{T}^{d}\right )\) and \(h\in H^{t}\left (\mathbb{T}^{d}\right )\), then \(fh\in H^{t}\left (\mathbb{T}^{d}\right )\) and

$$ \left \lVert fh\right \rVert _{H^{t}}\le C\left \lVert f\right \rVert _{H^{s}}\left \lVert h\right \rVert _{H^{t}}. $$
(B.10)

Proof

Since \(s > d/2\) we have that \(H^{s}(\mathbb{T}^{d})\) is an algebra and that \(H^{s}(\mathbb{T}^{d}) \hookrightarrow C^{0}_{b}(\mathbb{T}^{d})\). Consequently, for a fixed \(f \in H^{s}(\mathbb{T}^{d})\), if we define the linear operator \(T_{f}\) via \(T_{f} g = fg\), then \(T_{f}\) is bounded linear map from \(L^{2}\left (\mathbb{T}^{d}\right )\) to \(L^{2}\left (\mathbb{T}^{d}\right )\) and from \(H^{s}\left (\mathbb{T}^{d}\right )\) to \(H^{s}\left (\mathbb{T}^{d}\right )\) satisfying

$$ \left \lVert T_{f}\right \rVert _{\mathcal{L}(L^{2})} \le \left \lVert f\right \rVert _{C^{0}_{b}} \lesssim \left \lVert f\right \rVert _{H^{s}} \text{ and } \left \lVert T_{f}\right \rVert _{ \mathcal{L}(H^{s})} \lesssim \left \lVert f\right \rVert _{H^{s}}. $$
(B.11)

If \(s=t\), then we’re done. Otherwise \(0 < t < s\), and so the theory of complex interpolation (see, for instance, Theorem 6.23 in [9]) implies that \(T_{f}\) is a bounded linear map from \(H^{t}(\mathbb{T}^{d})\) to \(H^{t}(\mathbb{T}^{d})\) and that there exists \(\theta \in (0,1)\), depending on \(t\), \(s\), such that

$$ \left \lVert T_{f}\right \rVert _{\mathcal{L}(H^{t})} \le C \left \lVert T_{f}\right \rVert _{\mathcal{L}(L^{2})}^{\theta } \left \lVert T_{f}\right \rVert _{\mathcal{L}(H^{s})}^{1-\theta } \le C \left \lVert f\right \rVert _{H^{s}} $$
(B.12)

for a constant \(C>0\) depending only on \(s\), \(t\). Then for \(h \in H^{t}(\mathbb{T}^{d})\) we have that \(fh = T_{f} h \in H^{t}(\mathbb{T}^{d})\) with the stated estimate. □

2.3 B.3 Space-Time Sobolev Spaces

Here we consider some useful embedding properties of the space-time Sobolev spaced defined in Definition 1.6.

Proposition B.4

Suppose that \(r\in \mathbb{R}\), \(n\in \mathbb{N}\), and \(I= (0,T)\) for \(T \in (0,\infty ]\). Then following hold.

  1. 1.

    For every \(\mathbb{N}\ni k\le n\) and for all \(f\in L^{2}\left (I;H^{2+r}\left (\mathbb{T}^{3};\mathbb{R}^{3} \right )\right )\cap H^{n+1}\left (I;H^{r-2n}\left (\mathbb{T}^{3}; \mathbb{R}^{3}\right )\right )\) we have that \(f\in H^{k}\left (I;H^{2+r-2k}\left (\mathbb{T}^{3};\mathbb{R}^{3} \right )\right )\). Moreover, there exists a constant \(C\) independent of \(f\) such that

    $$ \left \lVert f\right \rVert _{H^{k}H^{2+r-2k}}\le C\left (\left \lVert f\right \rVert _{L^{2}H^{2+r}}+\left \lVert f\right \rVert _{H^{n+1}H^{r-2n}} \right ). $$
    (B.13)
  2. 2.

    If \(\mathbb{N}\ni k\le n\), then for every \(f\in L^{2}\left (I;H^{2+r}\left (\mathbb{T}^{3};\mathbb{R}^{3} \right )\right )\cap H^{n+1}\left (I;H^{r-2n}\left (\mathbb{T}^{3}; \mathbb{R}^{3}\right )\right )\) we may redefine \(f\) on a null set to arrive at the inclusion \(f\in UC^{k}_{b}\left (I;H^{r-2k+1}\left (\mathbb{T}^{3};\mathbb{R}^{3} \right )\right )\). Moreover, there exists a constant \(C\) depending on only \(k\) such that

    $$ \left \lVert f\right \rVert _{C^{k}_{b}H^{r-2k+1}}\le C\left (\left \lVert f\right \rVert _{L^{2}H^{2+r}}+\left \lVert f\right \rVert _{H^{n+1}H^{r-2n}} \right ). $$
    (B.14)

    Finally, if \(T=\infty \), then we have \(\lim _{t\to \infty }\left \lVert f(t)\right \rVert _{H^{r-2k+1}}=0\).

Proof

This is proved in Theorem 2.3 and 3.1 of [13]. □

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Stevenson, N., Tice, I. Analysis of Micropolar Fluids: Existence of Potential Microflow Solutions, Nearby Global Well-Posedness, and Asymptotic Stability. Acta Appl Math 170, 903–945 (2020). https://doi.org/10.1007/s10440-020-00363-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10440-020-00363-5

Mathematics Subject Classification (2010)

Keywords

Navigation