Skip to main content
Log in

Expanding large global solutions of the equations of compressible fluid mechanics

  • Published:
Inventiones mathematicae Aims and scope

Abstract

Without any symmetry assumptions on the initial data we construct global-in-time unique solutions to the vacuum free boundary three-dimensional isentropic compressible Euler equations when the adiabatic exponent \(\gamma \) lies in the interval \((1,\frac{5}{3}]\). Our initial data lie sufficiently close to the expanding compactly supported affine motions recently constructed by Sideris and they satisfy the physical vacuum boundary condition.

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

Notes

  1. Affine motions in the absence of free boundaries have been used before to understand qualitative behavior of solutions in fluid mechanics, for instance see Majda [32].

  2. In the absence of free boundaries, gas expansion also plays an important role in the global results in [12, 39]. The full nonlinear analysis of stabilizing effects of the fluid expansion in the context of general relativistic cosmological models was initiated by Rodnianski and Speck [36] and extended further in [11, 14, 27, 35, 41].

  3. By definition, Type I self-similar solutions approach an equilibrium exponentially fast in the logarithmic time variable (1.22). We refer to the review article [10] for a detailed discussion of this terminology.

  4. One could introduce a more geometric language so that \(\text {curl}_{\Lambda }\) and \(\Lambda \nabla \) are “natural” operators with respect to a given metric structure, but for the sake of conciseness we choose not to.

References

  1. Chen, G.-Q.: Remarks on R. J. DiPerna’s paper: convergence of the viscosity method for isentropic gas dynamics [Comm. Math. Phys. 91 (1983), no. 1, 1–30; MR0719807 (85i:35118)]. Proc. Am. Math. Soc. 125(10), 2981–2986 (1997)

    Article  MathSciNet  Google Scholar 

  2. Chiodaroli, E., De Lellis, C., Kreml, O.: Global ill-posedness of the isentropic system of gas dynamics. Commun. Pure Appl. Math. 68(7), 1157–1190 (2015)

    Article  MathSciNet  Google Scholar 

  3. Christodoulou, D.: The Formation of Shocks in 3-Dimensional Fluids. EMS Monographs in Mathematics. EMS Publishing House, Zürich (2007)

    Book  Google Scholar 

  4. Christodoulou, D., Miao, S.: Compressible Flow and Euler’s Equations, Surveys in Modern Mathematics, vol. 9. International Press, Vienna (2014)

    MATH  Google Scholar 

  5. Coutand, D., Shkoller, S.: Well-posedness in smooth function spaces for the moving-boundary 1-D compressible Euler equations in physical vacuum. Commun. Pure Appl. Math. 64(3), 328–366 (2011)

    Article  MathSciNet  Google Scholar 

  6. Coutand, D., Shkoller, S.: Well-posedness in smooth function spaces for the moving boundary three-dimensional compressible Euler equations in physical vacuum. Arch. Ration. Mech. Anal. 206(2), 515–616 (2012)

    Article  MathSciNet  Google Scholar 

  7. Dacorogna, B., Moser, J.: On a partial differential equation involving the Jacobian determinant. Ann. Inst. H. Poincaré Anal. Non Linéaire 7(1), 1–26 (1990)

    Article  MathSciNet  Google Scholar 

  8. Dafermos, C.: Hyperbolic Conservation Laws in Continuum Physics. Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], vol. 325, 2nd edn. Springer, Berlin (2005)

    Google Scholar 

  9. DiPerna, R.J.: Convergence of the viscosity method for isentropic gas dynamics. Commun. Math. Phys. 91, 1–30 (1983)

    Article  MathSciNet  Google Scholar 

  10. Eggers, J., Fontelos, A.M.: The role of self-similarity in singularities of partial differential equations. Nonlinearity 22, R1–R44 (2009)

    Article  MathSciNet  Google Scholar 

  11. Friedrich, H.: Sharp asymptotics for Einstein-\(\lambda \)-dust flows. Commun. Math. Phys. 350, 803–844 (2017)

    Article  MathSciNet  Google Scholar 

  12. Grassin, M.: Global smooth solutions to Euler equations for a perfect gas. Indiana Univ. Math. J. 47, 1397–1432 (1998)

    Article  MathSciNet  Google Scholar 

  13. Hadžić, M., Jang, J.: Nonlinear stability of expanding star solutions in the radially-symmetric mass-critical Euler-Poisson system. Commun. Pure Appl. Math. 71(5), 827–891 (2018)

    Article  Google Scholar 

  14. Hadžić, M., Speck, J.: The global future stability of the FLRW solutions to the Dust-Einstein system with a positive cosmological constant. J. Hyp. Differ. Equ. 12(1), 87–188 (2015)

    Article  MathSciNet  Google Scholar 

  15. Huang, H., Marcati, P., Pan, R.: Convergence to the Barenblatt solution for the compressible Euler equations with damping and vacuum. Arch. Ration. Mech. Anal. 176, 1–24 (2005)

    Article  MathSciNet  Google Scholar 

  16. Jang, J., Masmoudi, N.: Well-posedness for compressible Euler equations with physical vacuum singularity. Commun. Pure Appl. Math. 62, 1327–1385 (2009)

    Article  MathSciNet  Google Scholar 

  17. Jang, J., Masmoudi, N.: Vacuum in gas and fluid dynamics. In: Proceedings of the IMA Summer School on Nonlinear Conservation Laws and Applications, pp. 315–329. Springer (2011)

  18. Jang, J., Masmoudi, N.: Well and ill-posedness for compressible Euler equations with vacuum. J. Math. Phys. 53(11), 115625 (2012)

    Article  MathSciNet  Google Scholar 

  19. Jang, J., Masmoudi, N.: Well-posedness of compressible Euler equations in a physical vacuum. Commun. Pure Appl. Math. 68(1), 61–111 (2015)

    Article  MathSciNet  Google Scholar 

  20. Kufner, A., Malgranda, L., Persson, L.-E.: The Hardy Inequality. Vydavatelský Servis, Plzen (2007)

    Google Scholar 

  21. Lindblad, H.: Well posedness for the motion of a compressible liquid with free surface boundary. Commun. Math. Phys. 260, 319–392 (2005)

    Article  MathSciNet  Google Scholar 

  22. Lions, P.L., Perthame, B., Souganidis, P.E.: Existence and stability of entropy solutions for the hyperbolic systems of isentropic gas dynamics in Eulerian and Lagrangian coordinates. Commun. Pure Appl. Math. 49(6), 599–638 (1996)

    Article  MathSciNet  Google Scholar 

  23. Liu, T.-P.: Compressible flow with damping and vacuum. Jpn. J. Appl. Math. 13, 25–32 (1996)

    Article  MathSciNet  Google Scholar 

  24. Liu, T.-P., Smoller, J.: On the vacuum state for isentropic gas dynamics equations. Adv. Math. 1, 345–359 (1980)

    Article  MathSciNet  Google Scholar 

  25. Liu, T.-P., Yang, T.: Compressible Euler equations with vacuum. J. Differ. Equ. 140, 223–237 (1997)

    Article  MathSciNet  Google Scholar 

  26. Liu, T.-P., Yang, T.: Compressible flow with vacuum and physical singularity. Methods Appl. Anal. 7, 495–509 (2000)

    MathSciNet  MATH  Google Scholar 

  27. Lübbe, C., Valiente-Kroon, J.A.: A conformal approach for the analysis of the nonlinear stability of pure radiation cosmologies. Ann. Phys. 328, 1–25 (2013)

    Article  Google Scholar 

  28. Luk, J., Speck, J.: Shock formation in solutions to the \(2D\) compressible Euler equations in the presence of non-zero vorticity. Invent. Math. (to appear). arXiv:1610.00737

  29. Luo, T., Xin, Z., Zeng, H.: Well-posedness for the motion of physical vacuum of the three-dimensional compressible Euler equations with or without self-gravitation. Arch. Ration. Mech. Anal. 213(3), 763–831 (2014)

    Article  MathSciNet  Google Scholar 

  30. Luo, T., Zeng, H.: Global existence of smooth solutions and convergence to barenblatt solutions for the physical vacuum free boundary problem of compressible euler equations with damping. Commun. Pure Appl. Math. 69(7), 1354–1396 (2016)

    Article  MathSciNet  Google Scholar 

  31. Majda, A.: Compressible Fluid Flow and Systems of Conservation Laws in Several Space Variables, Volume 53 of Applied Mathematical Sciences. Springer, New York (1984)

    Book  Google Scholar 

  32. Majda, A.: Vorticity and the mathematical theory of incompressible fluid flow. Commun. Pure Appl. Math. 39(S, suppl.), S187–S220 (1986)

    Article  MathSciNet  Google Scholar 

  33. Makino, T., Ukai, S., Kawashima, S.: Sur la solution à support compact de l’équations d’Euler compressible. Jpn. J. Appl. Math. 3, 249–257 (1986)

    Article  Google Scholar 

  34. Merle, F., Raphaël, P., Szeftel, J.: Stable self similar blow up dynamics for \(L^2\)-supercritical NLS equations. Geom. Funct. Anal. 20(4), 1028–1071 (2010)

    Article  MathSciNet  Google Scholar 

  35. Oliynyk, T.: Future stability of the FLRW fluid solutions in the presence of a positive cosmological constant. Commun. Math. Phys. 346, 293–312 (2016)

    Article  MathSciNet  Google Scholar 

  36. Rodnianski, I., Speck, J.: The nonlinear future stability of the FLRW family of solutions to the irrotational Euler–Einstein system with a positive cosmological constant. J. Eur. Math. Soc. 15(6), 2369–2462 (2013)

    Article  MathSciNet  Google Scholar 

  37. Rozanova, O.: Solutions with Linear Profile of Velocity to the Euler Equations in Several Dimensions. Hyperbolic Problems: Theory, Numerics, Applications, pp. 861–870. Springer, Berlin (2003)

    MATH  Google Scholar 

  38. Shkoller, S., Sideris, T.C.: Global existence of near-affine solutions to the compressible Euler equations. Preprint arXiv:1710.08368

  39. Serre, D.: Solutions classiques globales des équations d’Euler pour un fluide parfait compressible. Ann. l’Inst. Fourier 47, 139–153 (1997)

    Article  Google Scholar 

  40. Serre, D.: Expansion of a compressible gas in vacuum. Bull. Inst. Math. Acad. Sin. Taiwan 10, 695–716 (2015)

    MathSciNet  MATH  Google Scholar 

  41. Speck, J.: The nonlinear future stability of the FLRW family of solutions to the Euler–Einstein system with a positive cosmological constant. Sel. Math. 18(3), 633–715 (2012)

    Article  MathSciNet  Google Scholar 

  42. Sideris, T.C.: Formation of singularities in three-dimensional compressible fluids. Commun. Math. Phys. 101(4), 475–485 (1985)

    Article  MathSciNet  Google Scholar 

  43. Sideris, T.C.: Spreading of the free boundary of an ideal fluid in a vacuum. J. Differ. Equ. 257(1), 1–14 (2014)

    Article  MathSciNet  Google Scholar 

  44. Sideris, T.C.: Global existence and asymptotic behavior of affine motion of 3D ideal fluids surrounded by vacuum. Arch. Ration. Mech. Anal. 225(1), 141–176 (2017)

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

The authors express their gratitude to P. Raphaël for fruitful discussions and for pointing out connections to the treatment of self-similar singular behavior for nonlinear Schrödinger equations. They also thank C. Dafermos for his feedback and pointing out important references. JJ is supported in part by NSF Grants DMS-1608492 and DMS-1608494 and a von Neumann fellowship of the Institute for Advanced Study through the NSF grant DMS-1128155. MH acknowledges the support of the EPSRC Grant EP/N016777/1.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Mahir Hadžić.

Appendices

A Asymptotic-in-\(\tau \) behavior of affine solutions

In this section we collect some of the basic properties of affine motions that are used at many places in our estimates. Their proofs are rather straightforward and follow directly from the description of the asymptotic behavior of the solutions of (2.18) from [44].

We remind the reader that for any \(M\in {\mathbb {M}}^{3\times 3}\) we denote by \(\Vert M\Vert \) the Hilbert–Schmidt norm of the matrix M.

Lemma A.1

(Asymptotic behavior of A, \(\Gamma ^*= O^{-1} O_\tau \), and \(\Lambda = O^{-1} O^{-\top }\)) For any \(\gamma \in (1,\frac{5}{3}]\) and any pair of initial conditions \((A(0),\dot{A}(0))\in {\mathrm{GL}}^+(3)\times {\mathbb {M}}^{3\times 3}\) there exist matrices \(A_0,A_1,M(t)\) such that the unique solution A(t) to the Cauchy problem

$$\begin{aligned} A_{tt}&= \delta \det {A}^{1-\gamma } A^{-\top } \end{aligned}$$
(A.1)
$$\begin{aligned} A(0)&= A(0), \quad A_t(0) = \dot{A}(0) \end{aligned}$$
(A.2)

can be written in the form

$$\begin{aligned} A(t) = A_0 + t A_1 + M(t), \quad t\ge 0, \end{aligned}$$
(A.3)

where \(A_0,A_1\) are both time-independent and M(t) satisfies the bounds

$$\begin{aligned} \Vert M(t)\Vert = o_{t\rightarrow \infty }(1+t), \quad \Vert \partial _t M(t)\Vert \lesssim (1+t)^{3-3\gamma }. \end{aligned}$$
(A.4)

Moreover

$$\begin{aligned} e^{\mu _1\tau } \lesssim \mu (\tau ) \lesssim e^{\mu _1\tau }, \quad \tau \ge 0, \end{aligned}$$
(A.5)
$$\begin{aligned} \lim _{\tau \rightarrow \infty } \Vert \Gamma ^* - e^{-\mu _1\tau } \mu _1 A_0 A_1^{-1} \Vert = 0, \end{aligned}$$
(A.6)

where

$$\begin{aligned} \Gamma ^* = O^{-1} O_\tau , \quad \mu _1 = (\det A_1)^{\frac{1}{3}}>0. \end{aligned}$$

Furthermore there exists a constant \(C>0\) such that

$$\begin{aligned}&\Vert \Lambda _\tau \Vert \le C e^{-\mu _1\tau }, \quad \Vert \Lambda _{\tau \tau }\Vert \le C e^{-2\mu _0\tau } , \quad \Vert \Lambda \Vert + \Vert \Lambda ^{-1}\Vert \le C, \end{aligned}$$
(A.7)
$$\begin{aligned}&\sum _{i=1}^3\left( d_i+\frac{1}{d_i}\right) \le C \end{aligned}$$
(A.8)
$$\begin{aligned}&\sum _{i=1}^3|\partial _\tau d_i| + \Vert \partial _\tau P\Vert \le C e^{-\mu _1\tau } \end{aligned}$$
(A.9)
$$\begin{aligned}&\frac{1}{C}|\mathbf{w}|^2 \le \langle \Lambda ^{-1}{} \mathbf{w}, \mathbf{w}\rangle \le C |\mathbf{w}|^2 , \ \mathbf{w}\in {\mathbb {R}}^3, \end{aligned}$$
(A.10)

where \(d_i\), \(i=1,2,3\), are the eigenvalues of the matrix \(\Lambda \) and \(P\in \text {SO}(3)\) satisfies \(\Lambda = P^\top Q P,\) \(Q = {\mathrm{diag}}(d_i)\).

Proof

Asymptotic behavior (A.3)–(A.4) and bound (A.5) are a consequence of Theorem 3 and Lemma 6 from [44].

Proof of (A.6) Since \( O = \frac{A}{\mu }\) we have

$$\begin{aligned} O_t = \frac{A_t}{\mu } - \frac{A\mu _t}{\mu ^2} = \frac{A_t}{\mu } - \frac{A\text {Tr}(A^{-1}A_t)}{3\mu }, \end{aligned}$$
(A.11)

where we used the formula

$$\begin{aligned} \mu _t {=}\partial _t\left( (\det A)^{\frac{1}{3}}\right) {=} \frac{1}{3} \mu ^{-2} \partial _t\det A = \frac{1}{3} \mu ^{-2} \mu ^3 \text {Tr}(A^{-1}A_t) = \frac{1}{3} \mu \text {Tr}(A^{-1}A_t). \end{aligned}$$

Therefore, from (A.3) and (A.4) it is easy to obtain the following asymptotics:

$$\begin{aligned} \text {Tr}(A^{-1}A_t)&\sim \text {Tr}((A_0+tA_1 + M(t))^{-1}(A_1+\partial _t M(t))) \nonumber \\&=t^{-1}\text {Tr}\left( \left( \frac{A_0}{t}+A_1\right) ^{-1}\left( A_1+O\left( t^{3-3\gamma }\right) \right) \right. \nonumber \\&\sim t^{-1}\text {Tr}\left( A_1^{-1}\left( A_1+O\left( t^{3-3\gamma }\right) \right) = \frac{3}{t} + O(t^{2-3\gamma }). \right. \end{aligned}$$
(A.12)

This implies that

$$\begin{aligned} O_t\sim \frac{A_1}{\mu } - \frac{(A_0+tA_1)\frac{3}{t}}{3\mu } = \frac{A_0}{t\mu }, \end{aligned}$$

where we made use of (A.4) again. Recalling that \( \frac{d\tau }{dt} = \frac{1}{\mu } \) we obtain the following asymptotic behavior \( O_\tau \sim _{\tau \rightarrow \infty } \frac{A_0}{t(\tau )}. \) Using (A.3) again it follows that

$$\begin{aligned} \det A\sim \det (A_0+t A_1) = t^3 \det \left( \frac{A_0}{t} + A_1\right) \sim t^3\det A_1, \quad \mu \sim t (\det A_1)^{\frac{1}{3}}. \end{aligned}$$
(A.13)

Since

$$\begin{aligned} t = t(\tau ) \sim _{\tau \rightarrow \infty } e^{(\det A_1)^{\frac{1}{3}}\tau }, \end{aligned}$$

and \( O \sim \frac{A_0+tA_1}{\mu }\) we conclude that

$$\begin{aligned} O^{-1}\sim \frac{\mu }{t} \left( \frac{A_0}{t}+A_1\right) ^{-1}\sim \frac{\mu }{t} A_1^{-1} \sim (\det A_1)^{\frac{1}{3}}A_1^{-1}, \end{aligned}$$

where we used (A.13). Therefore

$$\begin{aligned} \Gamma ^* = O^{-1} O_\tau \sim (\det A_1)^{\frac{1}{3}}A_1^{-1}\frac{A_0}{t(\tau )} = e^{-\mu _1\tau } \mu _1 A_0 A_1^{-1}, \end{aligned}$$

where \( \mu _1 = (\det A_1)^{\frac{1}{3}}, \) and this completes the proof of (A.6). Proof of (A.7) is similar.

Proof of (A.7)–(A.9) From the definition of \(\Lambda \) we have

$$\begin{aligned} \Lambda _\tau&= - O^{-1} O_\tau O^{-1} O^{-\top } - O^{-1} O^{-\top } O^\top _\tau O^{-\top } \nonumber \\&= -\Gamma ^*\Lambda - \Lambda (\Gamma ^*)^\top = -2 \Lambda (\Gamma ^*)^\top , \end{aligned}$$
(A.14)

where we used the symmetry of \(\Lambda \) in the last equality. Since \(\Vert \Lambda \Vert \lesssim 1\) it follows by part (i) that \(\Vert \Lambda _\tau \Vert \lesssim e^{-\mu _1\tau }\). To bound \(\Lambda _{\tau \tau }\) we note that by (A.14) \( \Lambda _{\tau \tau } = - 2 \Lambda _\tau (\Gamma ^*)^\top - 2\Lambda (\Gamma _\tau ^*)^\top . \) Since both \(\Vert \Lambda _\tau \Vert \) and \(\Vert \Gamma ^*\Vert \) decay exponentially, it remains to prove the decay of \(\Vert \Gamma _\tau ^*\Vert \). From \(\Gamma ^* = O^{-1}O_\tau \) it follows that \(\Gamma _\tau ^*=-(\Gamma ^*)^2 + O^{-1}O_{\tau \tau }\), and therefore it remains to prove the decay of \(\Vert O_{\tau \tau }\Vert \). A simple calculation shows that

$$\begin{aligned} O_{\tau \tau } = \mu \left( A_{tt} - \frac{A_t \text {Tr}(A^{-1}A_t)+A \partial _t\text {Tr}(A^{-1}A_t)}{3} \right) . \end{aligned}$$

Using the asymptotic behavior (A.12), (2.18), we can refine the asymptotics (A.12) to show that \(A_t \text {Tr}(A^{-1}A_t)+A \partial _t\text {Tr}(A^{-1}A_t) = O(t^{2-3\gamma })\) and therefore from the above equation it follows that

$$\begin{aligned} \Vert O_{\tau \tau }\Vert \lesssim (1+t)^{3-3\gamma } \lesssim e^{-2\mu _0\tau }. \end{aligned}$$

This yields the second bound in (A.7). Bounds (A.8) (A.9) follow by similar arguments using (A.3), while (A.10) is a direct consequence of (A.8). \(\square \)

B Commutators

In order to evaluate various commutator terms that arise from commuting differential operators with the usual Cartesian derivatives or apply the Leibniz rule we shall rely on the fact that the high-order Sobolev norms expressed in polar coordinates are equivalent to the usual high-order Sobolev norms on the support of function \(\psi \).

Lemma B.1

Let be a collection of the standard Cartesian, normal, and tangential vector-fields. For any two vector fields \(X_k,X_\ell \in {\mathcal {X}}\), \(k,\ell = 1,\ldots ,7\) there commutator satisfies the following relationship

where the functions \(c^m_{k\ell }\), \(c_{k\ell }\), are \(C^\infty \) on the exterior of any ball around the origin \(r=0\).

Proof

The proof is a simple consequence of the following direct calculations. For any \(i,j\in \{1,2,3\}\) we have

(B.1)

\(\square \)

Lemma B.1 is a technical tool allowing us to bound the lower order commutators in our energy estimates.

Using the product rule and the relationship \(\mathscr {A}=[D\eta ]^{-1}\) it is easy to see that the following formulas hold

$$\begin{aligned} \partial _r\mathscr {A}^k_i&= -\mathscr {A}^k_s\partial _r\partial _m\eta ^s\mathscr {A}^m_i = -\mathscr {A}^k_s\partial _r\left( \partial _m\uptheta ^s+\delta ^s_m\right) \mathscr {A}^m_i \nonumber \\&= -\mathscr {A}^k_s(\partial _r\uptheta ^s),_m\mathscr {A}^m_i + \mathscr {A}^k_s [\partial _m,\partial _r]\uptheta ^s\mathscr {A}^m_i. \end{aligned}$$
(B.2)

Similarly, for any \(j=1,2,3\) we have

(B.3)

Therefore, if \(\beta =(0,0,0)\) we have the formula

(B.4)

If \(|\beta | > 0\), then with \(e_1=(1,0,0)\), \(e_2 = (0,1,0)\), and \(e_3=(0,0,1)\) we obtain

(B.5)

where \(c_{a'\beta '},c_{\beta '}\) are positive universal constants and the Einstein summation convention does not apply to the index j. Finally, the high-order commutators appearing on the right-hand side in the identities (B.4)–(B.5) can be expressed as a linear combination of the elements of \({\mathcal {X}}\) with smooth coefficients away from zero. In other words, for any \(j=1,2,3\) the following commutator identity holds:

(B.6)

for some universal coefficients \(C^j_{a',\beta ',\ell }\) which are smooth on the exterior of any ball around the origin \(r=0\). Formula (B.6) is a direct consequence of Lemma B.1.

C Weighted spaces, Hardy inequalities, and Sobolev embeddings

In this section, we recall Hardy inequalities and embedding results of weighted function spaces. First of all, we state the Hardy inequality near \(r=1\).

Lemma C.1

(Hardy inequality [20]) Let k be a real number and g a function satisfying \(\int _0^1 (1-r)^{k +2}(g^2 + g'^2) dr < \infty \).

If \(k > -1\), then we have \( \int _0^1 (1-r)^{k} g^2 dr \le C \int _0^1 (1-r)^{k+2} (g^2 + |g'|^2) dr \).

If \(k < -1\), then g has a trace at \(r=1\) and \( \int _0^1 (1-r)^{k} (g - g(1))^2 dr \le C \int _0^1 (1-r)^{k+2} |g'|^2 dr\).

Since w depends only on r and w behaves like a distance function \(1-r\), using Lemma C.1, we in particular get

$$\begin{aligned} \int _{B_1(\mathbf{0}){\setminus } B_{\frac{1}{4}}(\mathbf{0})} \psi w^{k} | u |^2 dy \, \lesssim \int _{B_1(\mathbf{0}){\setminus } B_{\frac{1}{4}}(\mathbf{0})} \psi w^{k+2} \left( |\partial _r u |^2 + |u|^2 \right) dy \end{aligned}$$
(C.1)

for any nonnegative real number \(k> -1\) and for any \(u\in C^\infty (B_1(\mathbf{0}){\setminus } B_{\frac{1}{4}}(\mathbf{0}))\). We can apply (C.1) to and thereafter apply (C.1) repeatedly to the right-hand side. As a consequence for any \(-1< k < \alpha +a\),

(C.2)

Upon choosing \(m = \lceil a+\alpha -k \rceil \) where \(\lceil \ \rceil \) is the ceiling function, we obtain

(C.3)

for any \(u\in C^\infty (B_1(\mathbf{0}){\setminus } B_{\frac{1}{4}}(\mathbf{0}))\).

As a consequence of (C.3), we obtain the weighted Sobolev–Hardy inequality:

Proposition C.2

For any \(u\in C^\infty (B_1(\mathbf{0}))\), we have

(C.4)
(C.5)

We omit the technical details of the proof, as it follows from standard estimates relying on the \(H^2(\Omega )\hookrightarrow L^\infty (\Omega )\) continuous embedding and the Hardy inequality (C.3).

D Starting from Lagrangian formulation

Consider the Lagrangian formulation of the E\(_\gamma \)-system [6, 19]:

$$\begin{aligned} \zeta _{tt} + \tfrac{\gamma }{\gamma -1}\mathscr {A}_\zeta ^T \nabla (w {\mathscr {J}}_\zeta ^{1-\gamma })=0 \end{aligned}$$
(D.1)

where \(\mathscr {A}_\zeta \) and \({\mathscr {J}}_\zeta \) are induced by the flow map \(\partial _t\zeta (t,y)=\mathbf{{u}}(t,\zeta (t,y))\) (\(\mathbf{u}\) is the original fluid velocity appearing in (1.1)) and \(w=(\rho _0 {\mathscr {J}}_\zeta (0))^{\gamma -1}\ge 0\) is an enthalpy function that depends only on the initial data. To find affine motions (with center of mass at 0) one makes the ansatz \( \zeta (t,y)= A(t) y\) [44]. With \(\mathscr {A}_{\zeta }^\top =[D\zeta ]^{-\top }= A(t)^{-\top }\) and \({\mathscr {J}}_{\zeta }=\det A\), by plugging this ansatz into (D.1), we obtain

$$\begin{aligned} A_{tt} y + \tfrac{\gamma }{\gamma -1} (\det A)^{1-\gamma } A^{-\top } \nabla w=0 \end{aligned}$$
(D.2)

Since w is independent of t, (D.2) will be satisfied if we demand

$$\begin{aligned} A_{tt}= \delta (\det A)^{1-\gamma } A^{-\top } , \quad \delta y = - \tfrac{\gamma }{\gamma -1} \nabla w, \quad \delta >0, \end{aligned}$$
(D.3)

which precisely yields the affine motions (1.5)–(1.6) (discovered by Sideris when \(\mathbf{a}\equiv 0\)) which form the set \({\mathscr {S}}\) with the center of mass fixed at the origin.

In order to study the stability of elements of \({\mathscr {S}}\) we want to realize them as time-independent background solutions. Given such an affine motion A we modify the flow map \(\zeta \) and define \(\eta := A^{-1} \zeta \). Since \(\mathscr {A}_\zeta ^\top = A^{-\top } \mathscr {A}_\eta ^\top \) and \({\mathscr {J}}_\zeta = (\det A) {\mathscr {J}}_\eta \), we obtain

$$\begin{aligned} \eta _{tt} + 2 A^{-1} A_t \eta _t + A^{-1} A_{tt}\eta + \tfrac{\gamma }{\gamma -1} (\det A)^{1-\gamma } A^{-1} A^{-\top } \mathscr {A}_\eta ^\top \nabla (w {\mathscr {J}}_\eta ^{1-\gamma }) =0 \end{aligned}$$

By using (D.3) we can rewrite the previous equation in the form

$$\begin{aligned} (\det A)^{\gamma -\frac{1}{3}} \eta _{tt} + 2 (\det A)^{\gamma -\frac{1}{3}} A^{-1} A_t \eta _t +\delta \Lambda \eta + \tfrac{\gamma }{\gamma -1} \Lambda \mathscr {A}_\eta ^\top \nabla (w {\mathscr {J}}_\eta ^{1-\gamma }) =0 \end{aligned}$$
(D.4)

where \(\Lambda =\det A^{\frac{2}{3}} A^{-1}A^{-T}\).

We rescale the time variable t so that \(1+t\sim e^{\mu _1\tau }\) by setting \(\frac{d\tau }{d t} = (\det A)^{-\frac{1}{3}}\). Then (D.4) can be written as

$$\begin{aligned}&(\det A)^{\gamma -1} \eta _{\tau \tau } -\tfrac{1}{3} (\det A)^{\gamma -2} (\det A)_\tau \eta _\tau + 2 (\det A)^{\gamma -1} A^{-1} A_\tau \eta _\tau \nonumber \\&\quad +\delta \Lambda \eta + \tfrac{\gamma }{\gamma -1} \Lambda \mathscr {A}_\eta ^\top \nabla (w {\mathscr {J}}_\eta ^{1-\gamma }) =0 \end{aligned}$$
(D.5)

We now recall \(A^{-1} A_\tau = \mu ^{-1} \mu _\tau I + O^{-1}O_\tau \) where \( A= \mu O\) and \(\mu = (\det A)^{\frac{1}{3}}\) (see Sect. 2.1). The equation for \(\eta \) reads

$$\begin{aligned} \mu ^{3\gamma -3}\eta _{\tau \tau } + \mu ^{3\gamma -4}\mu _\tau \eta _\tau + 2 \mu ^{3\gamma -3}\Gamma ^*\eta _\tau +\delta \Lambda \eta + \tfrac{\gamma }{\gamma -1} \Lambda \mathscr {A}_\eta ^\top \nabla (w {\mathscr {J}}_\eta ^{1-\gamma }) =0 \end{aligned}$$
(D.6)

where \(\Gamma ^*=O^{-1}O_\tau \). It is clear that \(\eta (y)\equiv y\) corresponds to Sideris’ affine motions, and equation (D.6) is nothing but (2.35). By considering \(\uptheta =\eta -y\), we obtain the \(\uptheta \)-Eq. (2.38).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Hadžić, M., Jang, J. Expanding large global solutions of the equations of compressible fluid mechanics. Invent. math. 214, 1205–1266 (2018). https://doi.org/10.1007/s00222-018-0821-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00222-018-0821-1

Navigation