Log in

Free Boundary Problem for a Gas Bubble in a Liquid, and Exponential Stability of the Manifold of Spherically Symmetric Equilibria

  • Published:
Archive for Rational Mechanics and Analysis Aims and scope Submit manuscript

Abstract

We consider the dynamics of a gas bubble immersed in an incompressible fluid of fixed temperature, and focus on the relaxation of an expanding and contracting spherically symmetric bubble due to thermal effects. We study two models, both systems of PDEs with an evolving free boundary: the full mathematical model and an approximate model, arising, for example, in the study of sonoluminescence. For fixed physical parameters (surface tension of the gas–liquid interface, liquid viscosity, thermal conductivity of the gas, etc.), both models share a family of spherically symmetric equilibria, smoothly parametrized by the mass of the gas bubble. Our main result concerns the approximate model. We prove the nonlinear asymptotic stability of the manifold of equilibria with respect to small spherically symmetric perturbations. The rate of convergence is exponential in time. To prove this result we first prove a weak form of nonlinear asymptotic stability –with no explicit rate of time-decay– using the energy dissipation law, and then, via a center manifold analysis, bootstrap the weak time-decay to exponential time-decay. We also study the uniqueness of the family of spherically symmetric equilibria within each model. The family of spherically symmetric equilibria captures all spherically symmetric equilibria of the approximate system. However within the full model, this family is embedded in a larger family of spherically symmetric solutions. For the approximate system, we prove that all equilibrium bubbles are spherically symmetric, by an application of Alexandrov’s theorem on closed surfaces of constant mean curvature.

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

Access this article

Subscribe and save

Springer+ Basic
EUR 32.99 /Month
  • Get 10 units per month
  • Download Article/Chapter or Ebook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime
Subscribe now

Buy Now

Price includes VAT (Germany)

Instant access to the full article PDF.

Similar content being viewed by others

References

  1. Alexandrov, A.D.: A characteristic property of spheres. Ann. Mat. Pura Appl. 4(58), 303–315, 1962

    MathSciNet  Google Scholar 

  2. Bang, J.: Stationary Navier-Stokes Equations in an Exterior Domain, and Some Integral Identities for Euler and Navier-Stokes Equations. ProQuest LLC, Ann Arbor, MI, 2021. Thesis (Ph.D.)–Rutgers The State University of New Jersey, School of Graduate Studies

  3. Barber, B.P., Putterman, S.J.: Observation of synchronous picosecond sonoluminescence. Nature 352(6333), 318–320, 1991

    ADS  Google Scholar 

  4. Barber, B.P., Putterman, S.J.: Light scattering measurements of the repetitive supersonic implosion of a sonoluminescing bubble. Phys. Rev. Lett. 69(26), 3839, 1992

    ADS  Google Scholar 

  5. Barber, B.P., Wu, C., Löfstedt, R., Roberts, P.H., Putterman, S.J.: Sensitivity of sonoluminescence to experimental parameters. Phys. Rev. Lett. 72(9), 1380, 1994

    ADS  Google Scholar 

  6. Biro, Z., Velazquez, J.J.L.: Analysis of a free boundary problem arising in bubble dynamics. SIAM J. Math. Anal. 32(1), 142–171, 2000

    MathSciNet  MATH  Google Scholar 

  7. Bramson, M.: Convergence of solutions of the Kolmogorov equation to travelling waves. Mem. Amer. Math. Soc. 44(285), iv+190, 1983

    MathSciNet  MATH  Google Scholar 

  8. Brenner, M.P., Lohse, D., Dupont, T.F.: Bubble shape oscillations and the onset of sonoluminescence. Phys. Rev. Lett. 75(5), 954, 1995

    ADS  Google Scholar 

  9. Calvisi, M., Iloreta, J., Szeri, A.: Dynamics of bubbles near a rigid surface subjected to a lithotripter shock wave. Part 2. Reflected shock intensifies non-spherical cavitation collapse. J. Fluid Mech. 616, 63–97, 2008

    ADS  MATH  Google Scholar 

  10. Carr, J.: Applications of centre manifold theory, volume 35 of Applied Mathematical Sciences. Springer-Verlag, New York-Berlin (1981)

  11. Coddington, E.A., Levinson, N.: Theory of Ordinary Differential Equations. McGraw-Hill Book Co., Inc., New York-Toronto-London (1955)

    MATH  Google Scholar 

  12. Costin, O., Tanveer, S., Weinstein, M.I.: The lifetime of shape oscillations of a bubble in an unbounded, inviscid, and compressible fluid with surface tension. SIAM J. Math. Anal. 45(5), 2924–2936, 2013

    MathSciNet  MATH  Google Scholar 

  13. Coussios, C.C., Roy, R.A.: Applications of acoustics and cavitation to noninvasive therapy and drug delivery. Annu. Rev. Fluid Mech. 40, 395–420, 2008

    ADS  MathSciNet  MATH  Google Scholar 

  14. Curtiss, G., Leppinen, D., Wang, Q., Blake, J.: Ultrasonic cavitation near a tissue layer. J. Fluid Mech. 730, 245–272, 2013

    ADS  MathSciNet  MATH  Google Scholar 

  15. Epstein, D., Keller, J.: Expansion and contraction of planar, cylinderical and spherical underwater gas bubbles. J. Acoust. Soc. Am. 52, 975–980, 1972

    ADS  Google Scholar 

  16. Escher, Joachim, Simonett, Gieri: A center manifold analysis for the Mullins–Sekerka model. J. Differ. Equ. 143(2), 267–292, 1998

    ADS  MathSciNet  MATH  Google Scholar 

  17. Feireisl, E.: Dynamics of viscous compressible fluids, volume 26 of Oxford Lecture Series in Mathematics and its Applications. Oxford University Press, Oxford (2004)

  18. Feng, Z.C., Leal, L.G.: Nonlinear bubble dynamics. In Annual review of fluid mechanics, Vol. 29, volume 29 of Annu. Rev. Fluid Mech., pages 201–243. Annual Reviews, Palo Alto, CA (1997)

  19. Ferrara, K., Pollard, R., Borden, M.: Ultrasound microbubble contrast agents: fundamentals and application to gene and drug delivery. Annu. Rev. Biomed. Eng. 9, 415–447, 2007

    Google Scholar 

  20. Fourest, T., Deletombe, E., Faucher, V., Arrigoni, M., Dupas, J., Laurens, J.-M.: Comparison of Keller-Miksis model and finite element bubble dynamics simulations in a confined medium. Application to the hydrodynamic ram. Eur. J. Mech. B Fluids 68, 66–75, 2018

    ADS  MathSciNet  MATH  Google Scholar 

  21. Funaki, T., Ohnawa, M., Suzuki, Y., Yokoyama, S.: Existence and uniqueness of solutions to stochastic Rayleigh–Plesset equations. J. Math. Anal. Appl. 425(1), 20–32, 2015

    MathSciNet  MATH  Google Scholar 

  22. Galdi, G.P. An introduction to the mathematical theory of the Navier-Stokes equations. Vol. II, volume 39 of Springer Tracts in Natural Philosophy. Springer-Verlag, New York, 1994. Nonlinear steady problems

  23. Gearhart, Larry: Spectral theory for contraction semigroups on Hilbert space. Trans. Am. Math. Soc. 236, 385–394, 1978

    MathSciNet  MATH  Google Scholar 

  24. Goldsztein, G.H.: Collapse and rebound of a gas bubble. Stud. Appl. Math. 112(2), 101–132, 2004

    MathSciNet  MATH  Google Scholar 

  25. Goodman, J.: Nonlinear asymptotic stability of viscous shock profiles for conservation laws. Arch. Rational Mech. Anal. 95(4), 325–344, 1986

    ADS  MathSciNet  MATH  Google Scholar 

  26. Iloreta, J., Fung, N., Szeri, A.: Dynamics of bubbles near a rigid surface subjected to a lithotripter shock wave. part 1. consequences of interference between incident and reflected waves. J. Fluid Mech. 616, 43–61, 2008

    ADS  MATH  Google Scholar 

  27. Keller, J.B., Kolodner, I.I.: Dam** of underwater explosion bubble oscillations. J. Appl. Phys. 27(10), 1152–1161, 1956

    ADS  Google Scholar 

  28. Keller, J.B., Miksis, M.: Bubble oscillations of large amplitude. J. Acoust. Soc. Am. 68(2), 628–633, 1980

    ADS  MATH  Google Scholar 

  29. Klaseboer, E., Fong, S.W., Turangan, C.K., Khoo, B.C., Szeri, A.J., Calvisi, M.L., Sankin, G.N., Zhong, P.: Interaction of lithotripter shockwaves with single inertial cavitation bubbles. J. Fluid Mech. 593, 33–56, 2007

    ADS  MATH  Google Scholar 

  30. Kornfeld, M., Suvorov, L.: On the destructive action of cavitation. J. Appl. Phys. 15(6), 495–506, 1944

    ADS  Google Scholar 

  31. Ladyženskaja, O.A., Solonnikov, V.A., Ural’ceva, N.N.: Linear and quasi-linear equations of parabolic type. Izdat. “Nauka”, Moscow (1967)

  32. Lai, C.-C., Weinstein, M.I.: Thermal relaxation toward equilibrium and periodically pulsating spherical bubbles in an incompressible liquid. ar**v:2305.03569

  33. Leighton, T.: From seas to surgeries, from babbling brooks to baby scans: the acoustics of gas bubbles in liquids. Internat. J. Modern Phys. B 18(25), 3267–3314, 2004

    ADS  Google Scholar 

  34. Leighton, T.: The Acoustic Bubble. Academic Press, London (2012)

    Google Scholar 

  35. Leighton, T., Fedele, F., Coleman, A., McCarthy, C., Ryves, S., Hurrell, A., De Stefano, A., White, P.: A passive acoustic device for real-time monitoring of the efficacy of shockwave lithotripsy treatment. Ultrasound Med. Biol. 34(10), 1651–1665, 2008

    Google Scholar 

  36. Leighton, T.G., Turangan, C.K., Jamaluddin, A.R., Ball, G.J., White, P.R.: Prediction of far-field acoustic emissions from cavitation clouds during shock wave lithotripsy for development of a clinical device. Proc. R. Soc. Lond. Ser. A Math. Phys. Eng.Sci. 469(2150), 20120538, 21, 2013

    MathSciNet  MATH  Google Scholar 

  37. Leslie, T.A., Kennedy, J.E.: High-intensity focused ultrasound principles, current uses, and potential for the future. Ultrasound Q. 22(4), 263–272, 2006

    Google Scholar 

  38. Liu, W.: Bubble dynamics in a compressible viscous liquid. PhD thesis, University of Birmingham (2018)

  39. Longuet-Higgins, M.S.: Monopole emission of sound by asymmetric bubble oscillations. I. Normal modes. J. Fluid Mech. 201, 525–541, 1989

    ADS  MathSciNet  MATH  Google Scholar 

  40. Majda, A.J., Bertozzi, A.L.: Vorticity and incompressible flow, volume 27 of Cambridge Texts in Applied Mathematics. Cambridge University Press, Cambridge (2002)

  41. McKean, H.P.: Stabilization of solutions of a caricature of the FitzHugh-Nagumo equation. Comm. Pure Appl. Math. 36(3), 291–324, 1983

    MathSciNet  MATH  Google Scholar 

  42. McKean, H.P.: Stabilization of solutions of a caricature of the FitzHugh-Nagumo equation. II. Commun. Pure Appl. Math. 37(3), 299–301, 1984

    MathSciNet  MATH  Google Scholar 

  43. Novotný, A., Straškraba, I.: Introduction to the mathematical theory of compressible flow, volume 27 of Oxford Lecture Series in Mathematics and its Applications. Oxford University Press, Oxford (2004)

  44. Ohl, C.-D., Arora, M., Ikink, R., De Jong, N., Versluis, M., Delius, M., Lohse, D.: Sonoporation from jetting cavitation bubbles. Biophys. J. 91(11), 4285–4295, 2006

    ADS  Google Scholar 

  45. Ohnawa, M., Suzuki, Y.: Mathematical and numerical analysis of the Rayleigh-Plesset and the Keller equations. In Mathematical fluid dynamics, present and future, volume 183 of Springer Proc. Math. Stat., pages 159–180. Springer, Tokyo (2016)

  46. Plesset, M.: The dynamics of cavitation bubbles. J. Appl. Mech. 16(3), 277–282, 1949

    ADS  Google Scholar 

  47. Plesset, M.S., Chapman, R.B.: Collapse of an initially spherical vapour cavity in the neighbourhood of a solid boundary. J. Fluid Mech. 47(2), 283–290, 1971

    ADS  Google Scholar 

  48. Plesset, M.S., Prosperetti, A.: Bubble dynamics and cavitation. Annu. Rev. Fluid Mech. 9(1), 145–185, 1977

    ADS  MATH  Google Scholar 

  49. Prosperetti, A.: The thermal behaviour of oscillating gas bubbles. J. Fluid Mech. 222, 587–616, 1991

    ADS  MathSciNet  MATH  Google Scholar 

  50. Prosperetti, A.: Bubbles. Phys. Fluids 16(6), 1852–1865, 2004

    ADS  MathSciNet  MATH  Google Scholar 

  51. Prüss, Jan: On the spectrum of \(C_{0}\)-semigroups. Trans. Am. Math. Soc. 284(2), 847–857, 1984

    MATH  Google Scholar 

  52. Putterman, S.J., Roberts, P.H.: Comment on “bubble shape oscillations and the onset of sonoluminescence’’. Phys. Rev. Lett. 80, 3666–3667, 1998

    ADS  Google Scholar 

  53. Rayleigh, L.: Viii. On the pressure developed in a liquid during the collapse of a spherical cavity. Lond. Edinb. Dublin Philos. Mag. J. Sci. 34(200), 94–98, 1917

  54. Ripepe, M., Gordeev, E.: Gas bubble dynamics model for shallow volcanic tremor at stromboli. J. Geophys. Res. Solid Earth 104(B5), 10639–10654, 1999

    Google Scholar 

  55. Roberts, W.W., Hall, T.L., Ives, K., Wolf, J.S., Fowlkes, J.B., Cain, C.A.: Pulsed cavitational ultrasound: a noninvasive technology for controlled tissue ablation (histotripsy) in the rabbit kidney. J. Urol. 175(2), 734–738, 2006

    Google Scholar 

  56. Shapiro, A.M., Weinstein, M.I.: Radiative decay of bubble oscillations in a compressible fluid. SIAM J. Math. Anal. 43(2), 828–876, 2011

    MathSciNet  MATH  Google Scholar 

  57. Smith, W.R., Wang, Q.: Radiative decay of the nonlinear oscillations of an adiabatic spherical bubble at small Mach number. J. Fluid Mech. 837, 1–18, 2018

    ADS  MathSciNet  MATH  Google Scholar 

  58. Song, W., Hong, M., Lukyanchuk, B., Chong, T.: Laser-induced cavitation bubbles for cleaning of solid surfaces. J. Appl. Phys. 95(6), 2952–2956, 2004

    ADS  Google Scholar 

  59. Suslick, K.S.: Sonochemistry. Science 247(4949), 1439–1445, 1990

    ADS  Google Scholar 

  60. Tomita, Y., Shima, A.: On the behavior of a spherical bubble and the impulse pressure in a viscous compressible liquid. Bull. JSME 20(149), 1453–1460, 1977

    Google Scholar 

  61. Van Gorder, R.A.: Dynamics of the Rayleigh–Plesset equation modelling a gas-filled bubble immersed in an incompressible fluid. J. Fluid Mech. 807, 478–508, 2016

    ADS  MathSciNet  MATH  Google Scholar 

  62. Vokurka, K.: Comparison of Rayleigh’s, Herring’s, and Gilmore’s models of gas bubbles. Acta Acust. United Acust. 59(3), 214–219, 1986

    MATH  Google Scholar 

  63. Winkler, M.: Stabilization in a two-dimensional chemotaxis-Navier–Stokes system. Arch. Ration. Mech. Anal. 211(2), 455–487, 2014

    MathSciNet  MATH  Google Scholar 

  64. Xu, J., Attinger, D.: Acoustic excitation of superharmonic capillary waves on a meniscus in a planar microgeometry. Phys. Fluids 19(10), 108107, 2007

    ADS  MATH  Google Scholar 

  65. Zhang, Y.-N., Li, S.-C.: Effects of liquid compressibility on radial oscillations of gas bubbles in liquids. J. Hydrodyn. 24(5), 760–766, 2012

    ADS  Google Scholar 

  66. Zhou, G., Prosperetti, A.: Modelling the thermal behaviour of gas bubbles. J. Fluid Mech. 901, R3, 15, 2020

    MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

The authors thank Juan J. L. Velázquez for detailed discussions concerning the article [6], which motivated the present work. We also thank Panagiota Daskalopoulos, Qiang Du and Christophe Josserand for very stimulating discussions. CL acknowledges support by Simons Foundation as well as support from Department of Mathematics at Columbia University. MIW was supported in part by National Science Foundation Grant DMS-1908657 and Simons Foundation Math + X Investigator Award #376319 (Michael I. Weinstein).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Michael I. Weinstein.

Additional information

Communicated by P. Constantin.

Publisher's Note

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

Appendices

Appendix A: Spherically Symmetric Equilibria of the Full Liquid/Gas Model

In this appendix, we prove Proposition 4.1. That is, we show that (4.3) is the unique regular spherically symmetric equilibrium solution to the system (2.1)–(2.4) under the radiation condition (4.2) for \(T_l\).

Proof of Proposition 4.1

We consider the full liquid/gas model (2.1)–(2.4) and prove Proposition 4.1. Steady-state solutions of (2.1)–(2.4) solve

figure j
figure k
figure l

For the spherically symmetric case, (A.1b) we have \({\textbf{v}}_{l,*}(x) = v_{l,*}(r) \frac{x}{r}\). Therefore, \(0={\textrm{div}}{\textbf{v}}_{l,*}= \partial _r v_{l,*}(r) + (2/r)v_{l,*}(r)\) and hence

$$\begin{aligned} \begin{aligned} \frac{1}{r^2}\partial _r(r^2 v_{l,*}(r)) = 0,\quad r\ge R_*. \end{aligned}\end{aligned}$$
(A.4)

Therefore,

$$\begin{aligned} \begin{aligned} v_{l,*}(r) = \frac{a}{r^2},\quad r\ge R_*, \end{aligned}\end{aligned}$$
(A.5)

for constant a. But the boundary condition (A.3a) implies \(v_{l,*}(R_*) = 0\). So \(a=0\), and thus \(v_{l,*}\equiv 0\). Therefore, (A.1a) becomes \(\nabla p_{l,*} = 0\). So the pressure \(p_{l,*}\) is a constant and equal to its value at infinity, \(p_{\infty ,*}\).

For the gas velocity \(v_{g,*}\), (A.2a) becomes

$$\begin{aligned} \frac{1}{r^2} \partial _r(\rho _{g,*} v_{g,*})= 0,\quad 0\le r\le R_*, \end{aligned}$$

which implies \(\rho _{g,*} v_{g,*}\) is a constant. Again, the boundary condition (A.3a) implies \(v_{g,*}(R_*) = 0\). So \(\rho _{g,*} v_{g,*}\equiv 0\). But \(\rho _{g,*}\ne 0\) since otherwise \(s_*\) is singular in (A.2e). Therefore, \({\textbf{v}}_{g,*}\equiv {\textbf{0}}\) and thus (A.2b) becomes \(\nabla p_{g,*} = 0\). So \(p_{g,*}\) is a constant. Moreover, by \(v_{l,*}=v_{g,*}\equiv 0\), (A.3b) yields \(-p_{l,*} + p_{g,*} = \frac{2\sigma }{R_*}\). So \(p_{g,*} = p_{\infty ,*} + \frac{2\sigma }{R_*}\).

For the equations of the temperatures, due to \(v_{l,*}=0\) (A.1c) becomes \(\Delta T_{l,*} = 0\) in \({\mathbb {R}}^3{\setminus } B_{R_*}\), or, in spherical coordinates,

$$\begin{aligned} \frac{1}{r^2} \partial _r(r^2 \partial _r T_{l,*}) = 0,\quad r\ge R_*, \end{aligned}$$

which implies

$$\begin{aligned} \partial _r T_{l,*}=\frac{a_1}{r^2},\quad r\ge R_* \end{aligned}$$

for some constant \(a_1\). Integrating over \((r,\infty )\) we get

$$\begin{aligned} T_{l,*}(r) = T_\infty - \frac{a_1}{r},\quad r\ge R_*. \end{aligned}$$

By the radiation condition (4.2), \(T_{l,*}(r) = T_\infty + o(r^{-1})\) as \(r\rightarrow \infty \). This gives \(a_1=0\) and thus \(T_{l,*}\equiv T_\infty \). On the other hand, (A.2c) becomes \(\Delta T_{g,*} = 0\) in \(B_{R_*}\) since \(v_{g,*}=0\). Since \(T_{g,*}\) is regular, \(T_{g,*}\equiv T_{g,*}(R_*)=T_\infty \) by the maximum principle.

For the gas density \(\rho _{g,*}\), by (A.2d) \(\rho _{g,*} = \frac{p_{g,*}}{{\mathscr {R}}_gT_\infty } = \frac{1}{{\mathscr {R}}_gT_\infty } \left( p_{\infty ,*}+\frac{2\sigma }{R_*} \right) \). Due to the conservation of mass (7.1),

$$\begin{aligned} M:= \int _{B_{R_0}} \rho _0 = \lim _{t\rightarrow \infty } \int _{B_{R(t)}} \rho _g(\cdot ,t) = \frac{4\pi }{3} \rho _{g,*} R_*^3, \end{aligned}$$

where \((\rho _0(x),R_0)\), \(\rho _0(x)\ge 0, R_0>0\), is the initial data. Therefore, the steady state \((\rho _{g,*},R_*)\) can be obtained by solving

figure m

In particular, the stationary radius \(R_*\) satisfies the cubic equation

$$\begin{aligned} \begin{aligned} p_{\infty ,*} R_*^3 + 2\sigma R_*^2 - \frac{3{\mathscr {R}}_gT_\infty M}{4\pi } = 0. \end{aligned}\end{aligned}$$
(A.7)

It is readily seen that for any \(M>0\), the cubic equation has a unique positive root \(R_*[M]\). This proves Proposition 4.1. \(\square \)

Appendix B: Derivation of the Reduced System for \(\rho (r,t)\) and R(t): Proof of Proposition 5.1

Considering the uniformity of the pressure \(p_g\) in (3.2b), we can eliminate \(T_g\) by plugging (3.2d) into (3.2c) and deduce

$$\begin{aligned} \begin{aligned} \partial _t s + {\textbf{v}}_g\cdot \nabla s = \kappa _g \Delta \left( \frac{1}{\rho _g} \right) . \end{aligned}\end{aligned}$$
(B.1)

Plugging (3.2e) into the left hand side of (B.1) and using (3.2b), we have

$$\begin{aligned} \begin{aligned} c_v \left\{ \frac{\partial _t p_g}{p_g} - \frac{\gamma }{\rho _g} \left[ \partial _t\rho _g + {\textbf{v}}_g\cdot \nabla \rho _g \right] \right\} = \kappa _g \Delta \left( \frac{1}{\rho _g} \right) . \end{aligned}\end{aligned}$$
(B.2)

Using (3.2a) in (B.2), we obtain

$$\begin{aligned} \begin{aligned} c_v \left\{ \frac{\partial _t p_g}{p_g} + \gamma {\textrm{div}}{\textbf{v}}_g \right\} = \kappa _g \Delta \left( \frac{1}{\rho _g} \right) , \end{aligned}\end{aligned}$$
(B.3)

Therefore, the system (3.2) is reduced to

figure n

Expanding the term \({\textrm{div}}(\rho _g{\textbf{v}}_g)\) in (B.4a) and substituting \({\textrm{div}}{\textbf{v}}_g\) using (B.4b), and using the elementary identity

$$\begin{aligned} \begin{aligned} \rho _g \Delta \left( \frac{1}{\rho _g} \right) = -\Delta \log \rho _g + \frac{|\nabla \rho _g|^2}{\rho _g^2}, \end{aligned}\end{aligned}$$
(B.5)

we get

$$\begin{aligned} \begin{aligned} \partial _t\rho _g = \frac{\kappa }{\gamma c_v} \Delta \log \rho _g - \frac{\kappa }{\gamma c_v} \frac{|\nabla \rho _g|^2}{\rho _g^2} - {\textbf{v}}_g\cdot \nabla \rho _g + \frac{\partial _tp_g}{\gamma p_g} \rho _g. \end{aligned}\end{aligned}$$
(B.6)

Assuming the bubble is a sphere \(B_{R(t)}\) and solutions are spherically symmetric, and recalling we denoted the radial components of the gas and liquid velocity by \(v_g(r,t)\) and \(v_l(r,t)\), respectively, the systems (3.1) and (B.4) become

figure o

and

figure p

and the boundary condition (3.3) becomes

figure q

The liquid velocity and pressure \((v_l,p_l)\) can be directly solved in terms of R(t), \(\dot{R}(t)\), the liquid pressure \(p_l(R(t),t)\) on the bubble wall, and the far-field liquid pressure \(p_\infty (t):= p_l(r=\infty ,t)\). In fact, the incompressibility condition (B.7b) and the kinematic boundary condition (B.9a) imply

$$\begin{aligned} \begin{aligned} v_l(r,t) = \frac{(R(t))^2\dot{R}(t)}{r^2},\quad r\ge R(t),\ t>0. \end{aligned}\end{aligned}$$
(B.10)

Plugging Equation (B.10) into Equation (B.7a), we have

$$\begin{aligned} \begin{aligned} \frac{2R\dot{R}^2 {+} R^2 {\ddot{R}}}{r^2} {=} \nu _l\left( \frac{2R^2\dot{R}}{r^4} {-} 2\frac{R^2\dot{R}}{r^4} \right) {+} 2 \frac{R^4\dot{R}^2}{r^5} {-} \frac{1}{\rho _l} \partial _rp_l,\quad r\ge R(t),\ t>0. \end{aligned}\nonumber \\\end{aligned}$$
(B.11)

Note that the diffusion term in (B.11) vanishes. So the reduction using spherical symmetry assumption also works for Euler equation, i.e., we can take \(\nu _l=0\) in (3.1a). Integrating Equation (B.11) over \(r>R(t)\), we deduce

$$\begin{aligned} \begin{aligned} p_l(r,t) = p_\infty (t) + \rho _l \left( \frac{2R(t) (\dot{R}(t))^2 + (R(t))^2{\ddot{R}}(t)}{r} - \frac{(R(t))^4(\dot{R}(t))^2}{2r^4} \right) ,\quad r\ge R(t),\ t>0. \end{aligned}\nonumber \\\end{aligned}$$
(B.12)

In particular, on the boundary the liquid pressure is

$$\begin{aligned} p_l(R(t),t) = p_\infty (t) + \rho _l\left( \frac{3}{2}\dot{R}^2 + R{\ddot{R}} \right) ,\quad t>0. \end{aligned}$$

Moreover, (B.10) implies \(\partial _rv_l(r,t) = -2(R(t))^2\dot{R}(t)/r^3\) so that

$$\begin{aligned} \partial _rv_l(R(t),t) = -2\, \frac{\dot{R}(t)}{R(t)}. \end{aligned}$$

This implies

$$\begin{aligned} \begin{aligned} R{\ddot{R}} + \frac{3}{2} \dot{R}^2 = \frac{1}{\rho _l} \left( p_g(t) - p_\infty (t) - \frac{2\sigma }{R} -4\mu _l \frac{\dot{R}}{R} \right) ,\quad t>0, \end{aligned}\end{aligned}$$
(B.13)

where the Young–Laplace boundary condition (B.9b) has been used.

For the gas dynamics in the bubble, by integrating (B.8b) in r, the radial component of the gas velocity \(v_g\) can be expressed in terms of \(\rho _g(r,t)\), \(\partial _r\rho _g(r,t)\), \(p_g(t)\), and \(\partial _tp_g(t)\). To be more precise,

$$\begin{aligned} \begin{aligned} v_g(r,t) = \frac{\kappa }{\gamma c_v} \partial _r\left( \frac{1}{\rho _g(r,t)} \right) - \frac{\partial _tp_g(t)}{p_g(t)}\frac{r}{3\gamma },\quad 0\le r\le R(t),\ t>0. \end{aligned}\nonumber \\\end{aligned}$$
(B.14)

Using (B.14) we can eliminate \({\textbf{v}}_g\) in (B.6) and obtain

$$\begin{aligned} \begin{aligned} \partial _t\rho _g = \frac{\kappa }{\gamma c_v} \Delta _r\log \rho _g + \frac{\partial _t p_g}{3\gamma p_g} r \partial _r\rho _g + \frac{\partial _t p_g}{\gamma p_g} \rho _g,\quad 0\le r\le R(t),\ t>0, \end{aligned}\nonumber \\\end{aligned}$$
(B.15)

where \(\Delta _r f = \frac{1}{r^2}\partial _r(r^2\partial _rf)\) for spherically symmetric functions f. From the boundary condition (B.9c) for the gas temperature and the equation of state (3.2d),

$$\begin{aligned} \begin{aligned} p_g(t) = {\mathscr {R}}_g\rho _g(R(t),t) T_\infty . \end{aligned}\end{aligned}$$
(B.16)

Taking time derivative of (B.16) we obtain

$$\begin{aligned} \begin{aligned} \frac{\partial _t p_g}{p_g} = \frac{\partial _t\rho _g(R(t),t)}{\rho _g(R(t),t)} + \frac{\partial _r\rho _g(R(t),t)}{\rho _g(R(t),t)} \dot{R}(t). \end{aligned}\end{aligned}$$
(B.17)

Evaluating (B.14) at \(r=R(t)\) and using the kinematic boundary condition (B.9a) it follows that

$$\begin{aligned} \begin{aligned} \dot{R}(t) = -\frac{\kappa }{\gamma c_v} \frac{\partial _r\rho _g(R(t),t)}{\left( \rho _g(R(t),t) \right) ^2} - \frac{R(t)}{3\gamma } \frac{\partial _t p_g}{p_g}. \end{aligned}\end{aligned}$$
(B.18)

For the boundary data for the gas density, we use (B.16) and (B.13) to deduce

$$\begin{aligned} \begin{aligned} \rho _g(R(t),t) = \frac{1}{{\mathscr {R}}_gT_\infty } \left[ p_\infty + \frac{2\sigma }{R} + \rho _l\left( R{\ddot{R}} + \frac{3}{2} (\dot{R})^2 \right) \right] . \end{aligned}\end{aligned}$$
(B.19)

Collecting the results (B.15), (B.16), (B.18), (B.19), we conclude that, under the spherical symmetry assumption, the system (3.1)–(3.3) is reduced to a system of \((\rho (r,t),R(t))\):

$$\begin{aligned} \partial _t\rho= & {} \frac{\kappa }{\gamma c_v} \Delta _r\log \rho + \frac{\partial _t p}{3\gamma p} r \partial _r\rho + \frac{\partial _t p}{\gamma p} \rho ,\quad 0\le r\le R(t),\ t>0, \end{aligned}$$
(B.20)
$$\begin{aligned} p(t)= & {} {\mathscr {R}}_gT_\infty \rho (R(t),t),\quad t>0, \end{aligned}$$
(B.21)
$$\begin{aligned} \dot{R}(t)= & {} -\frac{\kappa }{\gamma c_v} \frac{\partial _r\rho (R(t),t)}{\left( \rho (R(t),t) \right) ^2} - \frac{R(t)}{3\gamma } \frac{\partial _t p}{p},\quad t>0, \end{aligned}$$
(B.22)
$$\begin{aligned} \rho (R(t),t)= & {} \frac{1}{{\mathscr {R}}_gT_\infty } \left[ p_\infty + \frac{2\sigma }{R} + 4\mu _l \frac{\dot{R}}{R} + \rho _l\left( R{\ddot{R}} + \frac{3}{2} (\dot{R})^2 \right) \right] , \quad t>0,\nonumber \\ \end{aligned}$$
(B.23)

where \(\rho \equiv \rho _g\), \(p\equiv p_g\), \(\kappa =\kappa _g\). This is the reduced system (5.1a)–(5.1c).

Appendix C: A Perspective on Coercive Energy Estimate of Biro–Velázquez, and an Extension

In this appendix, we prove Theorem 7.5, which extends the coercivity estimate of Biro-Velázquez to the case where \(p_\infty -p_{\infty ,*}\) is small in norm.

Proof of Theorem 7.5

Let us recall the total energy

$$\begin{aligned} {\mathcal {E}}_{\textrm{total}} = FE + KE_l + U_{g-l} + PV_{p_\infty }, \end{aligned}$$

where

figure r

The energy is a functional of state variables, which are defined on a deforming regime, \(B_{R}\). We fix the region to be \(B_1\) by setting \(x=Ry\), where \(y\in B_1\). Defining \({\overline{\rho }}(y)=\rho (Rr)\) and using the constitutive relation \(p = {\mathscr {R}}_gT_\infty \rho (R) = {\mathscr {R}}_gT_\infty {\overline{\rho }}(1)\) we have that

$$\begin{aligned} FE= & {} \frac{4\pi c_v T_\infty }{3} {\overline{\rho }}(1) R^3 - c_v T_\infty M_0 \log ({\mathscr {R}}_gT_\infty ) - c_v T_\infty M_0 \log {\overline{\rho }}(1)\\{} & {} +\, c_v\gamma T_\infty R^3 \int _{B_1} {\overline{\rho }} \log {\overline{\rho }}. \end{aligned}$$

Thus, \({\mathcal {E}}_{\textrm{total}} \) is a functional of \(({\overline{\rho }}, R,\dot{R})\):

$$\begin{aligned} \begin{aligned} {\mathcal {E}}_{\textrm{total}}&= {\mathcal {E}}_{\textrm{total}} [{\overline{\rho }}, R,\dot{R}]\\&= \frac{4\pi c_v T_\infty }{3} {\overline{\rho }}(1)R^3 - c_vT_\infty M_0\log ({\mathscr {R}}_gT_\infty ) - c_vT_\infty M_0 \log {\overline{\rho }}(1)\\&\quad + c_v\gamma T_\infty R^3 \int _{B_1} {\overline{\rho }} \log {\overline{\rho }}\\&\quad + 2\pi \rho _lR^3\dot{R}^2 + 4\pi \sigma R^2 + \frac{4\pi }{3}R^3p_\infty . \end{aligned} \end{aligned}$$

We set \({\overline{\rho }} = \rho _* + \varrho \), \(R = R_* + {\mathcal {R}}\) and expand the total energy \({\mathcal {E}}_{\textrm{total}}[{\overline{\rho }}, R,\dot{R}]\) at \((\rho _*,R_*,\dot{R}_*=0)\) along the mass preserving hypersurface \(M_0 = \text {Mass}[\rho ,R]\):

$$\begin{aligned} \begin{aligned}&{\mathcal {E}}_{\textrm{total}} [{\overline{\rho }},R,\dot{R}] = {\mathcal {E}}_{*} + d{\mathcal {E}}_{*}[ \varrho , {\mathcal {R}}, \dot{{\mathcal {R}}}]+ \frac{1}{2} d^2{\mathcal {E}}_{*}[ \varrho , {\mathcal {R}}, \dot{{\mathcal {R}}}] + O(|( \varrho , {\mathcal {R}}, \dot{{\mathcal {R}}})|^3),\quad \textrm{where}\\&{\mathcal {E}}_{*}={\mathcal {E}}_{\textrm{total}} [\rho _*,R_*,\dot{R}_*=0]\\&d{\mathcal {E}}_{*}[ \varrho , {\mathcal {R}}, \dot{{\mathcal {R}}}]= \frac{\textrm{d}}{\textrm{d}\varepsilon }\Big |_{\varepsilon =0} {\mathcal {E}}_{\textrm{total}} (\rho _*,R_*+\varepsilon {\mathcal {R}},\varepsilon \dot{{\mathcal {R}}})\\&d^2{\mathcal {E}}_{*}[ \varrho , {\mathcal {R}}, \dot{{\mathcal {R}}}]= \frac{\textrm{d}^2}{\textrm{d}\varepsilon ^2}\Big |_{\varepsilon =0} {\mathcal {E}}_{\textrm{total}} (\rho _*,R_*+\varepsilon {\mathcal {R}},\varepsilon \dot{{\mathcal {R}}})\ \end{aligned} \end{aligned}$$
(C.2)

To expand along the mass preserving hypersurface, we first use \(M_0 = \int _{B_R} \rho \, \textrm{d}x\) to rewrite

$$\begin{aligned} \begin{aligned} R^3 \int _{B_1} {\overline{\rho }} \log {\overline{\rho }}&= \int _{B_R} \rho \log \rho = \int _{B_R} \rho \log \rho _* + \int _{B_R} \rho \log \frac{\rho }{\rho _*}\\&= \log \rho _*\int _{B_R} \rho + \int _{B_R} \rho + \int _{B_R} \rho \left( \log \frac{\rho }{\rho _*} - 1 \right) \\&= M_0\log \rho _* + M_0 + R^3 \int _{B_1} {\overline{\rho }} \left( \log \frac{{\overline{\rho }}}{\rho _*} - 1 \right) , \end{aligned} \end{aligned}$$

giving the following expression for the total energy:

$$\begin{aligned} \begin{aligned} {\mathcal {E}}_{\textrm{total}} [{\overline{\rho }}, R,\dot{R}]&= \frac{4\pi c_v T_\infty }{3} {\overline{\rho }}(1,t)R^3 - c_vT_\infty M_0\log ({\mathscr {R}}_gT_\infty ) - c_vT_\infty M_0 \log {\overline{\rho }}(1,t) \\&\quad + c_v\gamma T_\infty M_0\log \rho _* + c_v\gamma T_\infty M_0 + c_v\gamma T_\infty R^3 \int _{B_1} {\overline{\rho }} \left( \log \frac{{\overline{\rho }}}{\rho _*} - 1 \right) \\&\quad + 2\pi \rho _lR^3\dot{R}^2 + 4\pi \sigma R^2 + \frac{4\pi }{3}R^3 p_\infty . \end{aligned} \end{aligned}$$

To expand the logarithmic terms we note that for \(z_*\ne 0\) and \(|z-z_*|<\frac{1}{2} |z_*|\):

$$\begin{aligned} \Big | \Big ( z\left( \log \left( \frac{z}{z_*}\right) - 1 \Big ) - \Big (-z_* + \frac{1}{2z_*} (z-z_*)^2 \right) \Big |&\le \frac{2}{|z_*|}|z-z_*|^3 \end{aligned}$$
(C.3)
$$\begin{aligned} \Big | \log z - \left( \log z_* + \frac{1}{z_*}(z-z_*)- \frac{1}{2 z_*^2} (z-z_*)^2 \right) \Big |&\le \frac{2}{3|z_*|^3}|z-z_*|^3 \end{aligned}$$
(C.4)

Applying (C.3) and (C.4) we have

$$\begin{aligned} \begin{aligned} {\mathcal {E}}_{\textrm{total}} [{\overline{\rho }}, R,\dot{R}]&= \frac{4\pi c_v T_\infty }{3} {\overline{\rho }}(1,t)R^3 - c_vT_\infty M_0\log ({\mathscr {R}}_gT_\infty )\\&\quad - c_vT_\infty M_0\left( \log \rho _* + \frac{1}{\rho _*} \varrho (1)- \frac{1}{2 \rho _*^2} \varrho (1)^2\right) + O\left( \varrho (1)^3\right) \\&\quad + c_v\gamma T_\infty M_0\log \rho _* + c_v\gamma T_\infty M_0\\&\quad + c_v\gamma T_\infty R^3\left( -\frac{4\pi }{3}\rho _* +\frac{1}{2\rho _*}\int _{B_1} \varrho ^2 \right) + O\left( R^3\int _{B_1}| \varrho |^3 \right) \\&\quad + 2\pi \rho _lR^3\dot{R}^2 + 4\pi \sigma R^2 + \frac{4\pi }{3}R^3 p_\infty . \end{aligned} \end{aligned}$$

Rearranging and simplifying gives

$$\begin{aligned} {\mathcal {E}}_{\textrm{total}} [{\overline{\rho }}, R,\dot{R}]&= - c_vT_\infty M_0\log ({\mathscr {R}}_gT_\infty ) + c_v(\gamma -1) T_\infty M_0\log \rho _* + c_v\gamma T_\infty M_0 \nonumber \\&\quad +\, \frac{4\pi c_v T_\infty }{3} {\overline{\rho }}(1)R^3 + c_vT_\infty M_0\left( -\frac{1}{\rho _*} \varrho (1) + \frac{1}{2 \rho _*^2} \varrho (1)^2\right) + O\left( \varrho (1)^3\right) \nonumber \\&\quad + c_v\gamma T_\infty R^3\left( -\frac{4\pi }{3}\rho _* +\frac{1}{2\rho _*}\int _{B_1} \varrho ^2 \right) + O\left( R^3\int _{B_1} | \varrho |^3 \right) \nonumber \\&\quad + 2\pi \rho _lR^3\dot{R}^2 + 4\pi \sigma R^2 + \frac{4\pi }{3}R^3 p_\infty . \end{aligned}$$
(C.5)

Verification that \(d{\mathcal {E}}_*[\varrho ,{\mathcal {R}},\dot{{\mathcal {R}}}]=0\) when \(p_\infty = p_{\infty ,*}\). Starting with (C.5) we calculate:

$$\begin{aligned} d{\mathcal {E}}_*[ \varrho ,{\mathcal {R}},\dot{{\mathcal {R}}}]&= \frac{4\pi c_v T_\infty }{3}\left( 3\rho _*R_*^2{\mathcal {R}} +R_*^3 \varrho (1)\right) -\frac{c_vT_\infty }{\rho _*}\left( \frac{4\pi }{3}\rho _*R_*^3\right) \varrho (1) \nonumber \\&\quad -4\pi c_v\gamma T_\infty R_*^2\rho _*{\mathcal {R}}+8\pi \sigma R_*{\mathcal {R}}+4\pi R_*^2 p_\infty {\mathcal {R}}\nonumber \\&= 4\pi R_*^2\left( c_vT_\infty \rho _*(1-\gamma )+ \frac{2\sigma }{R_*}+p_\infty \right) {\mathcal {R}}\nonumber \\&= 4\pi R_*^2\left( -{\mathscr {R}}_gT_\infty \rho _*+ \frac{2\sigma }{R_*}+p_\infty \right) {\mathcal {R}}\qquad \left[ \gamma -1=\frac{{\mathscr {R}}_g}{c_v}\ \text {by (2.3)}\right] \nonumber \\&= 4\pi R_*^2 {\mathcal {P}}_\infty {\mathcal {R}} \qquad \left[ \text {by}\, (4.14b) \right] , \end{aligned}$$
(C.6)

where \({\mathcal {P}}_\infty = p_\infty - p_{\infty ,*}\). It is readily to see that \(d{\mathcal {E}}_*[ \varrho ,{\mathcal {R}},\dot{{\mathcal {R}}}]=0\) when \(p_\infty = p_{\infty ,*}\).

Computation of \(\frac{1}{2}d^2{\mathcal {E}}[ \varrho , {\mathcal {R}}, \dot{{\mathcal {R}}}]\).

From (C.5) we compute the quadratic terms:

$$\begin{aligned} \frac{1}{2}d^2{\mathcal {E}}[ \varrho , {\mathcal {R}}, \dot{{\mathcal {R}}}]&= 4\pi c_v T_\infty \left( \rho _*R_* {\mathcal {R}}^2 + R_*^2 \varrho (1) {\mathcal {R}}\right) + \frac{c_v T_\infty M_0}{2\rho _*^2} \varrho (1)^2 \nonumber \\&\quad + \frac{c_v\gamma T_\infty R_*^3}{2\rho _*} \int _{B_1} \varrho ^2 - 4\pi c_v\gamma \rho _* T_\infty R_* {\mathcal {R}}^2\nonumber \\&\quad + 2\pi \rho _l R_*^3 \dot{{\mathcal {R}}}^2 + 4\pi \sigma {\mathcal {R}}^2 + 4\pi R_* p_\infty {\mathcal {R}}^2. \end{aligned}$$
(C.7)

Next, using that the perturbed bubble is assumed to have mass equal to \(M_0=\textrm{Mass}(\rho _*,R_*)\), we express the cross-term just above in terms of a quadratic expression in \(\varrho \) as follows:

$$\begin{aligned} M_0=R^3 \int _{B_1}{\overline{\rho }}&= (R_*+{\mathcal {R}})^3 \int _{B_1}\left( \rho _*+ \varrho \right) \\&= M_0 + 4\pi R_*^2 \rho _* {\mathcal {R}} + R_*^3 \int _{B_1} \varrho + O\Big ({\mathcal {R}}^2 + \left( \int _{B_1} \varrho \right) ^2 \Big ) \end{aligned}$$

and therefore

$$\begin{aligned} {\mathcal {R}} = -\frac{R_*}{4\pi \rho _*}\int _{B_1} \varrho + O\Big ({\mathcal {R}}^2 + \left( \int _{B_1} \varrho \right) ^2 \Big ). \end{aligned}$$
(C.8)

Substitution of (C.8) into (C.7) we obtain a leading expression entirely in terms of the perturbed density \( \varrho \). We list the various terms that we rewrite exclusively in terms of \( \varrho \):

$$\begin{aligned} 4\pi c_v T_\infty \rho _*R_* {\mathcal {R}}^2&= \frac{c_v T_\infty \rho _* R_*^3}{4\pi \rho _*^2} \left( \int _{B_1} \varrho \right) ^2 + O\left( |{\mathcal {R}}|^3 + \left( \int _{B_1} | \varrho |\right) ^3 \right) \\ 4\pi c_v T_\infty R_*^2 \varrho (1) {\mathcal {R}}&= -\frac{c_v T_\infty R_*^3}{\rho _*}\ \varrho (1) \int _{B_1} \varrho + O\left( |{\mathcal {R}}|^3 + | \varrho (1)|^3 + \left( \int _{B_1} | \varrho |\right) ^3 \right) \\ - 4\pi c_v\gamma \rho _* T_\infty R_* {\mathcal {R}}^2&= -\frac{c_v\gamma T_\infty R_*^3}{4\pi \rho _*} \left( \int _{B_1} \varrho \right) ^2 + O\left( |{\mathcal {R}}|^3 + \left( \int _{B_1} | \varrho |\right) ^3 \right) \\ 4\pi \left( \sigma + R_* p_\infty \right) {\mathcal {R}}^2&= \frac{1}{4\pi \rho _*^2} \left( \frac{\sigma }{R_*} + p_\infty \right) R_*^3\left( \int _{B_1} \varrho \right) ^2 + O\left( |{\mathcal {R}}|^3 + \left( \int _{B_1} | \varrho |\right) ^3 \right) . \end{aligned}$$

Inserting these expressions into (C.7), we obtain

$$\begin{aligned} \frac{1}{2}d^2{\mathcal {E}}[ \varrho , {\mathcal {R}}, \dot{{\mathcal {R}}}]&= \frac{c_v T_\infty M_0}{2\rho _*^2} \varrho (1)^2 + \frac{c_v\gamma T_\infty R_*^3}{2\rho _*} \int _{B_1} \varrho ^2 + 2\pi \rho _l R_*^3 \dot{{\mathcal {R}}}^2 \nonumber \\&\quad + \frac{R_*^3}{4\pi \rho _*^2}\left( c_v(1-\gamma )T_\infty \rho _* + \frac{\sigma }{R_*}+p_\infty \right) \left( \int _{B_1} \varrho \right) ^2 - \frac{c_v T_\infty R_*^3}{\rho _*} \varrho (1) \int _{B_1} \varrho \nonumber \\&\quad + O\left( |{\mathcal {R}}|^3 + | \varrho (1)|^3 +\left( \int _{B_1} | \varrho |\right) ^3 \right) \end{aligned}$$
(C.9)

The coefficient of the fourth term on the right of (C.9) can be simplified using the relation \(1-\gamma = -{\mathscr {R}}_g/c_v\) and the relation \({\mathscr {R}}_gT_\infty \rho _* = p_{\infty ,*} + 2\sigma /R_*\) between the equilibrium density and bubble radius:

$$\begin{aligned} c_v(1-\gamma )T_\infty \rho _* + \frac{\sigma }{R_*}+p_\infty = -{\mathscr {R}}_gT_\infty \rho _* + \frac{\sigma }{R_*}+p_\infty =-\frac{\sigma }{R_*} + {\mathcal {P}}_\infty , \end{aligned}$$

where \({\mathcal {P}}_\infty = p_\infty - p_{\infty ,*}\). Thus,

$$\begin{aligned} \frac{1}{2}d^2{\mathcal {E}}[ \varrho , {\mathcal {R}}, \dot{{\mathcal {R}}}]&= \frac{c_v T_\infty M_0}{2\rho _*^2} \varrho (1)^2 + \frac{c_v\gamma T_\infty R_*^3}{2\rho _*} \int _{B_1} \varrho ^2 + 2\pi \rho _l R_*^3 \dot{{\mathcal {R}}}^2 \nonumber \\&\quad {-} \frac{\sigma R_*^2}{4\pi \rho _*^2}\left( \int _{B_1} \varrho \right) ^2 {+} \frac{R_*^3}{4\pi \rho _*^2} {\mathcal {P}}_\infty \left( \int _{B_1} \varrho \right) ^2 {-} \frac{c_v T_\infty R_*^3}{\rho _*} \varrho (1) \int _{B_1} \varrho \nonumber \\&\quad + O\left( |{\mathcal {R}}|^3 + | \varrho (1)|^3 +\left( \int _{B_1} | \varrho |\right) ^3 \right) \end{aligned}$$
(C.10)

Using that \(M_0=\rho _* R_*^3 |B_1|\), we may rewrite (C.10) as

$$\begin{aligned}&\frac{1}{2}d^2{\mathcal {E}}[ \varrho , {\mathcal {R}}, \dot{{\mathcal {R}}}] \nonumber \\&\quad = \frac{c_v T_\infty R_*^3 |B_1|}{2} \left( \frac{ \varrho (1)^2}{\rho _*} - 2\frac{ \varrho (1)}{\rho _*} \frac{1}{|B_1|}\int _{B_1} \varrho \right) + \frac{c_v\gamma T_\infty R_*^3 }{2\rho _*} \int _{B_1} \varrho ^2 + 2\pi \rho _l R_*^3 \dot{{\mathcal {R}}}^2 \nonumber \\&\qquad - \frac{\sigma R_*^2}{4\pi \rho _*^2}\left( \int _{B_1} \varrho \right) ^2 + \frac{R_*^3}{4\pi \rho _*^2} {\mathcal {P}}_\infty \left( \int _{B_1} \varrho \right) ^2 + O\left( |{\mathcal {R}}|^3 + | \varrho (1)|^3 +\left( \int _{B_1} | \varrho |\right) ^3 \right) \end{aligned}$$
(C.11)

or

$$\begin{aligned} \frac{1}{2}d^2{\mathcal {E}}[ \varrho , {\mathcal {R}}, \dot{{\mathcal {R}}}]&= \frac{c_v T_\infty R_*^3 |B_1|}{2\rho _*} \left( \varrho (1)- \frac{1}{|B_1|}\int _{B_1} \varrho \right) ^2 + \frac{c_v\gamma T_\infty R_*^3}{2\rho _*} \int _{B_1} \varrho ^2 + 2\pi \rho _l R_*^3 \dot{{\mathcal {R}}}^2 \nonumber \\&\quad - \left[ \frac{\sigma R_*^2}{4\pi \rho _*^2} + \frac{c_vT_\infty R_*^3}{2 \rho _*|B_1|} \right] \left( \int _{B_1} \varrho \right) ^2 + \frac{R_*^3}{4\pi \rho _*^2} {\mathcal {P}}_\infty \left( \int _{B_1} \varrho \right) ^2 \nonumber \\&\quad + O\left( |{\mathcal {R}}|^3 + | \varrho (1)|^3 +\left( \int _{B_1} | \varrho |\right) ^3 \right) \end{aligned}$$
(C.12)

By the Cauchy-Schwarz inequality \(\left( \int _{B_1} \varrho \right) ^2\le |B_1| \int _{B_1} \varrho ^2 \) and therefore

$$\begin{aligned} \frac{1}{2}d^2{\mathcal {E}}[ \varrho , {\mathcal {R}}, \dot{{\mathcal {R}}}]&\ge \frac{c_v T_\infty R_*^3 |B_1|}{2\rho _*} \left( \varrho (1)- \frac{1}{|B_1|}\int _{B_1} \varrho \right) ^2 + 2\pi \rho _l R_*^3 \dot{{\mathcal {R}}}^2 \nonumber \\&\quad + \left( \frac{c_v\gamma T_\infty R_*^3}{2\rho _*}- \left[ \frac{\sigma R_*^2|B_1|}{4\pi \rho _*^2} + \frac{c_vT_\infty R_*^3}{2 \rho _*} \right] \right) \int _{B_1} \varrho ^2 + \frac{R_*^3}{4\pi \rho _*^2}{\mathcal {P}}_\infty \left( \int _{B_1} \varrho \right) ^2 \nonumber \\&\quad + O\left( |{\mathcal {R}}|^3 + | \varrho (1)|^3 +\left( \int _{B_1} | \varrho |\right) ^3 \right) \end{aligned}$$
(C.13)

Finally, we find for the constant in (C.13) that

$$\begin{aligned} \frac{c_v\gamma T_\infty R_*^3}{ 2\rho _*}- \left[ \frac{\sigma R_*^2|B_1|}{4\pi \rho _*^2} + \frac{c_vT_\infty R_*^3}{2 \rho _*} \right] = \frac{R_*^3}{ \rho _*^2} \left( \frac{p_{\infty ,*}}{2} + \frac{2\sigma }{3R_*}\right) . \end{aligned}$$
(C.14)

This follows, yet again, from the relations \(\gamma - 1= {\mathscr {R}}_g/c_v\) and \({\mathscr {R}}_gT_\infty \rho _* = p_{\infty ,*} + 2\sigma /R_*\). For the term involving \({\mathcal {P}}_\infty \), using the Cauchy-Schwarz inequality \(\left( \int _{B_1} \varrho \right) ^2\le |B_1| \int _{B_1} \varrho ^2 \),

$$\begin{aligned} {\mathcal {P}}_\infty \left( \int _{B_1} \varrho \right) ^2 \ge -|{\mathcal {P}}_\infty | |B_1| \int _{B_1} \varrho ^2. \end{aligned}$$

Summarizing

$$\begin{aligned} {\mathcal {E}}_{\textrm{total}} - {\mathcal {E}}_*&\ge \frac{c_v T_\infty R_*^3 |B_1|}{2\rho _*} \left( \varrho (1)- \frac{1}{|B_1|}\int _{B_1} \varrho \right) ^2\nonumber \\&\quad + 2\pi \rho _l R_*^3 \dot{{\mathcal {R}}}^2 + \frac{R_*^3}{\rho _*^2} \left( \frac{p_{\infty ,*}}{2} + \frac{2\sigma }{3R_*}\right) \int _{B_1} \varrho ^2 \nonumber \\&\quad - 4\pi R_*^2 |{\mathcal {P}}_\infty | |{\mathcal {R}}| - \frac{R_*^3}{4\pi \rho _*^2} |{\mathcal {P}}_\infty | |B_1| \int _{B_1} \varrho ^2\nonumber \\&\quad + O\left( |{\mathcal {R}}|^3 + | \varrho (1)|^3 +\left( \int _{B_1} | \varrho |\right) ^3 \right) , \end{aligned}$$
(C.15)

where all explicit terms are non-negative except for the terms involving \({\mathcal {P}}_\infty \) which can be made small since \(|{\mathcal {P}}_\infty | = |p_\infty - p_{\infty ,*}|\le \delta _0\).

We now conclude the proof by bounding the error term in (C.15) from above by a sufficiently small constant times \(\int _{B_1} \varrho ^2\).

For the fourth term on the right hand side of (C.15), using (C.8), in terms of the perturbed density \( \varrho \),

$$\begin{aligned} - 4\pi R_*^2 |{\mathcal {P}}_\infty | |{\mathcal {R}}| \ge - \frac{R_*^3}{\rho _*} |{\mathcal {P}}_\infty | \int _{B_1} | \varrho | - C_0|{\mathcal {P}}_\infty | \left( \int _{B_1} \varrho \right) ^2 \end{aligned}$$

for some constant \(C_0>0\). Since \(| \varrho |\le \delta _0\le 1\) and \(|{\mathcal {P}}_\infty | = |p_\infty - p_{\infty ,*}|\le \delta _0\), by the Cauchy-Schwarz inequality \(\left( \int _{B_1} \varrho \right) ^2\le |B_1| \int _{B_1} \varrho ^2 \)

$$\begin{aligned} - 4\pi R_*^2 |{\mathcal {P}}_\infty | |{\mathcal {R}}| \ge - \frac{R_*^3}{\rho _*} |{\mathcal {P}}_\infty | \int _{B_1} \varrho ^2 - C_0|{\mathcal {P}}_\infty | |B_1| \int _{B_1} \varrho ^2 \ge - C_1 \delta _0 \int _{B_1} \varrho ^2 \end{aligned}$$

for some constant \(C_1>0\).

Now we estimate the cubic term in the third line on the right hand side of (C.15). Since \(M_0 = \text {Mass}[\rho ,R]\),

$$\begin{aligned} \int _{B_R} (\rho - \rho _*)\, \textrm{d}x = \int _{B_R} \varrho \, \textrm{d}x = M_0 - \frac{4\pi R^3}{3} \rho _* = \frac{4\pi R_*^3}{3} \rho _* - \frac{4\pi R^3}{3} \rho _*, \end{aligned}$$

or

$$\begin{aligned} \begin{aligned} \frac{4\pi \rho _*}{3} (R^3 - R_*^3) = -\int _{B_R} (\rho -\rho _*)\, \textrm{d}x, \end{aligned}\end{aligned}$$
(C.16)

which implies

$$\begin{aligned} \begin{aligned} |{\mathcal {R}}|&= |R - R_*| \le \frac{3}{4\pi \rho _*(R^2+RR_*+R_*^2)} |B_R|^{\frac{1}{2}} \left( \int _{B_R} |\rho -\rho _*|^2\, \textrm{d}x \right) ^{\frac{1}{2}} \\&\le C_2 \left( \int _{B_R} |\rho -\rho _*|^2\, \textrm{d}x \right) ^{\frac{1}{2}}, \end{aligned}\end{aligned}$$
(C.17)

where \(C_2>0\) depends only on \(\nu , M_0, T_\infty \). We now control \(| \varrho (1,t)|^3\) by the first and the third terms on the right hand side of (C.15). Indeed,

$$\begin{aligned} \begin{aligned} | \varrho (1)|&= \left| \left( \varrho (1) - \frac{1}{|B_1|} \int _{B_1} \varrho \right) + \frac{1}{|B_1|} \int _{B_1} \varrho \, \right| \\&\le \left| \varrho (1) - \frac{1}{|B_1|} \int _{B_1} \varrho \, \right| + \frac{1}{|B_1|^{\frac{1}{2}}} \left( \int _{B_1} \varrho ^2 \right) ^{\frac{1}{2}}\\&\le C_3\left\{ \left| \varrho (1) - \frac{1}{|B_1|} \int _{B_1} \varrho \, \right| + \left( \int _{B_1} \varrho ^2 \right) ^{\frac{1}{2}} \right\} \end{aligned} \end{aligned}$$

for some \(C_3>0\) depending only on \(\nu , M_0, T_\infty \). Since \(| \varrho (1)| = |\rho (R) - \rho _*| \le \delta _0\),

$$\begin{aligned} \begin{aligned} | \varrho (1)|^3 = | \varrho (1)| |\varrho (1)|^2&\le \delta _0 C_3 \left\{ \left| \varrho (1) - \frac{1}{|B_1|} \int _{B_1} \varrho \, \right| + \left( \int _{B_1} \varrho ^2 \right) ^{\frac{1}{2}} \right\} ^2\\&\le 2 \delta _0 C_3 \left| \varrho (1) - \frac{1}{|B_1|} \int _{B_1} \varrho \, \right| ^2 + 2\delta _0 C_3 \int _{B_1} \varrho ^2. \end{aligned}\end{aligned}$$
(C.18)

Using (C.17) and (C.18), one has

$$\begin{aligned} \begin{aligned}&O\left( |{\mathcal {R}}|^3 + | \varrho (1)|^3 +\left( \int _{B_1} | \varrho |\right) ^3 \right) \\&\quad \ge - C \left( \int _{B_R} (\rho -\rho _*)^2 \right) ^{\frac{3}{2}} - C\delta _0 \left| \varrho (1) - \frac{1}{|B_1|} \int _{B_1} \varrho \, \right| ^2 - C\delta _0 \int _{B_1} \varrho ^2 - C \int _{B_1}| \varrho |^3 \end{aligned} \end{aligned}$$

for some \(C>0\) depending only on \(\nu , M_0, T_\infty \).

Consequently, using \(|{\mathcal {P}}_\infty | = |p_\infty - p_{\infty ,*}|\le \delta _0\), (C.15) can be further computed as

$$\begin{aligned} \begin{aligned} {\mathcal {E}}_{\textrm{total}} - {\mathcal {E}}_*&\ge - C_1 \delta _0 \int _{B_1} \varrho ^2 + \left( \frac{c_v T_\infty R_*^3 |B_1|}{2\rho _*} - C\delta _0 \right) \left( \varrho (1) - \frac{1}{|B_1|} \int _{B_1} \varrho \right) ^2 \\&\quad + \frac{R_*^3}{\rho _*}\left( \frac{p_{\infty ,*}}{2} + \frac{2\sigma }{3R_*} \right) \int _{B_1} \varrho ^2\\&\quad - C\delta _0 |B_R|^{\frac{1}{2}} \int _{B_R} (\rho -\rho _*)^2 \,\textrm{d}x - 2C\delta _0 \int _{B_1} \varrho ^2 - \delta _0 \frac{R_*^3}{4\pi \rho _*} |B_1| \int _{B_1} \varrho ^2\\&\ge \Theta \left( \int _{B_R} (\rho -\rho _*)^2 \right) \end{aligned} \end{aligned}$$

for some constant \(\Theta >0\), provided \(\delta _0>0\) is sufficiently small. Note that we’ve used \(\int _{B_1} \varrho ^2\, \textrm{d}y = R^{-3}\int _{B_R}(\rho - \rho _*)^2\, \textrm{d}x\) in which \(R^{-3}\ge \nu ^3\) above. This completes the proof of Theorem 7.5.

Appendix D: An Interpolation Lemma

Lemma D.1

Let \(\Omega \) be a bounded Lipschitz domain in \({\mathbb {R}}^n\), \(k< m\), and \(0<\gamma \le 1\). For \(u\in C^\infty (\Omega )\),

$$\begin{aligned} \left\| \nabla ^k u \right\| _{L^\infty (\Omega )} \le C_1 \left\| u \right\| _{L^p(\Omega )}^\lambda \left\| u \right\| _{C^{m,\gamma }(\Omega )}^{1-\lambda } + C_2 \left\| u \right\| _{L^s(\Omega )} \end{aligned}$$

for arbitrary \(s\ge 1\), where \(-\frac{k}{n} = \frac{\lambda }{p} - (1-\lambda ) \frac{m}{n}\), and the constants \(C_1\), \(C_2\) depend on the domain \(\Omega \) and on s in addition to the other parameters.

Proof

By Gagliardo–Nirenberg interpolation inequality,

$$\begin{aligned} \left\| \nabla ^k u \right\| _{L^\infty (\Omega )} \le C_1 \left\| u \right\| _{L^p(\Omega )}^{\lambda } \left\| \nabla ^m u \right\| _{L^\infty (\Omega )}^{1-\lambda } + C_2 \left\| u \right\| _{L^s(\Omega )} \end{aligned}$$

for arbitrary \(s\ge 1\), where

$$\begin{aligned} 0 = \frac{k}{n} - \frac{m}{n}(1-\lambda ) + \frac{\lambda }{p}, \end{aligned}$$

and the constants \(C_1\), \(C_2\) depend on the domain \(\Omega \) and on s in addition to the other parameters. The lemma then follows since \(\left\| \nabla ^m u \right\| _{L^\infty (\Omega )} \le \left\| u \right\| _{C^{m,\gamma }(\Omega )}\). \(\quad \square \)

Appendix E: Estimate of the Exponential Decay Rate \(\beta \) in the Linearized System

In this appendix, we prove parts (2) and (3) of Theorem 9.3. In particular, we investigate the location of the roots of the meromorphic function \(Q(\tau )\) defined in (9.20) which corresponds to the spectrum of the linear operator \({\mathcal {L}}\).

Lemma E.1

There exists a negative upper bound for the real parts of the roots of the meromorphic function \(Q(\tau )\) in (9.20). More precisely, there exists \(\beta >0\) such that \(\xi <-\beta \) for all roots \(\tau =\xi +i\eta \) of \(Q(\tau )\). The constant \(\beta \) can be chosen as

$$\begin{aligned} \begin{aligned} \beta&= \min \Bigg \{ \left( 1 - \sqrt{\dfrac{1-\frac{1}{\gamma }}{\frac{3p_{\infty ,*}R_*+6\sigma }{2p_{\infty ,*}R_*+6\sigma } - \frac{1}{\gamma }}} \right) \pi ^2\overline{\kappa },\, \sqrt{ \frac{{\mathscr {R}}_gT_\infty \rho _*}{\rho _lR_*^2} },\\&\quad \frac{2\mu _l}{\rho _lR_*^2} + {{\mathbbm {1}}}_{\Delta \le 0}\, \frac{{\mathscr {R}}_gT_\infty \rho _*}{\pi ^4\overline{\kappa }\rho _lR_*^2} \left( 1-\frac{1}{\gamma } \right) \left[ \frac{\pi ^4}{90} + O\left( \left( \frac{1}{\pi ^2\overline{\kappa }}\sqrt{ \frac{{\mathscr {R}}_gT_\infty \rho _*}{\rho _lR_*^2} } \right) ^{3/2} \right) \right] \\&\quad - {{\mathbbm {1}}}_{\Delta >0}\, \frac{\sqrt{\Delta }}{2\rho _lR_*} \Bigg \}, \end{aligned}\nonumber \\\end{aligned}$$
(E.1)

in which

$$\begin{aligned} \Delta := \left( \frac{4\mu _l}{R_*} \right) ^2 - 8\rho _l{\mathscr {R}}_gT_\infty \rho _*. \end{aligned}$$

Proof of Lemma E.1

Let \(\tau = \xi + i\eta \) be a root of \(Q(\tau )\), i.e., \(Q(\tau )=0\). Plugging \(\tau = \xi + i\eta \), \(\xi \in {\mathbb {R}},\eta \in {\mathbb {R}}\), into (9.20), we have

$$\begin{aligned} Q(\xi + i\eta ) = \frac{1}{{\mathscr {R}}_gT_\infty } \left( \** _1 + iH_1 \right) \left( \** _2 + iH_2 \right) + 4\pi \,\frac{\rho _*}{R_*}, \end{aligned}$$

where

$$\begin{aligned} \begin{aligned} \** _1&= \frac{4\pi }{3\gamma } + \frac{8(\gamma -1)}{\pi \gamma }\sum _{j=1}^\infty \frac{\pi ^2\overline{\kappa }\left( \pi ^2\overline{\kappa }j^2 + \xi \right) }{\left( \pi ^2\overline{\kappa }j^2 + \xi \right) ^2 + \eta ^2},\\ H_1&= - \frac{8(\gamma -1)}{\pi \gamma } \sum _{j=1}^\infty \frac{\pi ^2\overline{\kappa }\eta }{\left( \pi ^2\overline{\kappa }j^2 + \xi \right) ^2 + \eta ^2},\\ \** _2&= \rho _lR_*\left( \xi ^2 - \eta ^2 \right) + \frac{4\mu _l}{R_*}\, \xi - \frac{2\sigma }{R_*^2},\\ H_2&= \rho _lR_*(2\xi \eta ) + \frac{4\mu _l}{R_*}\, \eta . \end{aligned}\end{aligned}$$
(E.2)

Setting real and imaginary parts of Q equal to zero, we obtain

$$\begin{aligned} \begin{aligned} \text {real part: }&\frac{1}{{\mathscr {R}}_gT_\infty } \left( \** _1\** _2 - H_1H_2 \right) + 4\pi \, \frac{\rho _*}{R_*} = 0,\\ \text {imaginary part: }&\frac{1}{{\mathscr {R}}_gT_\infty } \left( \** _1H_2 + H_1\** _2 \right) = 0. \end{aligned}\end{aligned}$$
(E.3)

The real part in (E.3) reads

$$\begin{aligned} \begin{aligned} 0&= \frac{1}{{\mathscr {R}}_gT_\infty } \bigg [ \left( \frac{4\pi }{3\gamma } + \frac{8(\gamma -1)}{\pi \gamma }\sum _{j=1}^\infty \frac{\pi ^2\overline{\kappa }\left( \pi ^2\overline{\kappa }j^2 + \xi \right) }{\left( \pi ^2\overline{\kappa }j^2 + \xi \right) ^2 + \eta ^2} \right) \left( \rho _lR_*\left( \xi ^2-\eta ^2 \right) + \frac{4\mu _l}{R_*}\, \xi - \frac{2\sigma }{R_*^2} \right) \\&\quad + \frac{8(\gamma -1)}{\pi \gamma } \sum _{j=1}^\infty \frac{\pi ^2\overline{\kappa }\eta ^2}{\left( \pi ^2\overline{\kappa }j^2 + \xi \right) ^2 + \eta ^2 } \left( \rho _lR_*(2\xi ) + \frac{4\mu _l}{R_*} \right) \bigg ] + 4\pi \, \frac{\rho _*}{R_*}\\&= \frac{1}{{\mathscr {R}}_gT_\infty } \bigg [ \left( \frac{4\pi }{3\gamma } + \frac{8(\gamma -1)}{\pi \gamma }\sum _{j=1}^\infty \frac{\pi ^4\overline{\kappa }^2j^2}{\left( \pi ^2\overline{\kappa }j^2 + \xi \right) ^2 + \eta ^2} \right) \left( \rho _lR_*\left( \xi ^2-\eta ^2 \right) + \frac{4\mu _l}{R_*}\, \xi - \frac{2\sigma }{R_*^2} \right) \\&\quad + \frac{8(\gamma -1)}{\pi \gamma } \sum _{j=1}^\infty \frac{\pi ^2\overline{\kappa }}{\left( \pi ^2\overline{\kappa }j^2 + \xi \right) ^2 + \eta ^2 } \left( \rho _lR_*\xi (\xi ^2+\eta ^2) + \frac{4\mu _l}{R_*}(\xi ^2+\eta ^2) - \frac{2\sigma }{R_*^2}\, \xi \right) \bigg ] + 4\pi \, \frac{\rho _*}{R_*}. \end{aligned}\end{aligned}$$
(E.4)

When \(\eta \ne 0\), the imaginary part in (E.3) reads

$$\begin{aligned} 0&= \left( \frac{4\pi }{3\gamma } + \frac{8(\gamma -1)}{\pi \gamma } \sum _{j=1}^\infty \frac{\pi ^2\overline{\kappa }\left( \pi ^2\overline{\kappa }j^2 + \xi \right) }{\left( \pi ^2\overline{\kappa }j^2 + \xi \right) ^2 + \eta ^2} \right) \left( \rho _lR_*(2\xi ) + \frac{4\mu _l}{R_*} \right) \nonumber \\&\quad - \frac{8(\gamma -1)}{\pi \gamma } \sum _{j=1}^\infty \frac{\pi ^2\overline{\kappa }}{\left( \pi ^2\overline{\kappa }j^2 + \xi \right) ^2 + \eta ^2 } \left( \rho _lR_*(\xi ^2 - \eta ^2) + \frac{4\mu _l}{R_*}\, \xi - \frac{2\sigma }{R_*^2} \right) \nonumber \\&= \left( \frac{4\pi }{3\gamma } + \frac{8(\gamma -1)}{\pi \gamma } \sum _{j=1}^\infty \frac{\pi ^4\overline{\kappa }^2 j^2}{\left( \pi ^2\overline{\kappa }j^2 + \xi \right) ^2 + \eta ^2} \right) \left( \rho _lR_*(2\xi ) + \frac{4\mu _l}{R_*} \right) \nonumber \\&\quad + \frac{8(\gamma -1)}{\pi \gamma } \sum _{j=1}^\infty \frac{\pi ^2\overline{\kappa }}{\left( \pi ^2\overline{\kappa }j^2 + \xi \right) ^2 + \eta ^2 } \left( \rho _lR_*(\xi ^2 + \eta ^2) + \frac{2\sigma }{R_*^2} \right) . \end{aligned}$$
(E.5)

For \(\eta \ne 0\), the equation (E.5) implies

$$\begin{aligned} \begin{aligned} \rho _lR_*(2\xi ) + \frac{4\mu _l}{R_*} < 0, \end{aligned}\end{aligned}$$
(E.6)

which gives

$$\begin{aligned} \begin{aligned} \xi < - \frac{2\mu _l}{\rho _lR_*^2} \le 0,\qquad \eta \ne 0. \end{aligned}\end{aligned}$$
(E.7)

The equation (E.5) also implies

$$\begin{aligned} \begin{aligned}&\frac{8(\gamma -1)}{\pi \gamma } \sum _{j=1}^\infty \frac{\pi ^2\overline{\kappa }}{\left( \pi ^2\overline{\kappa }j^2 + \xi \right) ^2 + \eta ^2 }\\&\quad = - \left( \frac{4\pi }{3\gamma } + \frac{8(\gamma -1)}{\pi \gamma } \sum _{j=1}^\infty \frac{\pi ^4\overline{\kappa }^2 j^2}{\left( \pi ^2\overline{\kappa }j^2 + \xi \right) ^2 + \eta ^2} \right) \dfrac{\rho _lR_*(2\xi ) + \frac{4\mu _l}{R_*}}{\rho _lR_*(\xi ^2+\eta ^2) + \frac{2\sigma }{R_*^2}}. \end{aligned}\end{aligned}$$
(E.8)

Plugging (E.8) into the real part (E.4), we derive

$$\begin{aligned} \begin{aligned} 0&= \frac{1}{{\mathscr {R}}_gT_\infty } \left( \frac{4\pi }{3\gamma } + \frac{8(\gamma -1)}{\pi \gamma }\sum _{j=1}^\infty \frac{\pi ^4\overline{\kappa }^2j^2}{\left( \pi ^2\overline{\kappa }j^2 + \xi \right) ^2 + \eta ^2} \right) \Bigg [ \rho _lR_*\left( \xi ^2-\eta ^2 \right) + \frac{4\mu _l}{R_*}\, \xi - \frac{2\sigma }{R_*^2} \\&\quad - \dfrac{\left( \rho _lR_*(2\xi ) + \frac{4\mu _l}{R_*} \right) \left( \rho _lR_*\xi (\xi ^2+\eta ^2) + \frac{4\mu _l}{R_*}(\xi ^2+\eta ^2) - \frac{2\sigma }{R_*^2}\, \xi \right) }{\rho _lR_*(\xi ^2+\eta ^2) + \frac{2\sigma }{R_*^2}} \Bigg ] + 4\pi \, \frac{\rho _*}{R_*}\\&=: \frac{1}{{\mathscr {R}}_gT_\infty } \left( \frac{4\pi }{3\gamma } + \frac{8(\gamma -1)}{\pi \gamma }\sum _{j=1}^\infty \frac{\pi ^4\overline{\kappa }^2j^2}{\left( \pi ^2\overline{\kappa }j^2 + \xi \right) ^2 + \eta ^2} \right) \Lambda + 4\pi \, \frac{\rho _*}{R_*}, \end{aligned}\end{aligned}$$
(E.9)

where \(\Lambda \) is the square bracket on the right hand side of the first equation. A straightforward calculation shows that

$$\begin{aligned} \begin{aligned}&\left( \rho _lR_*(\xi ^2+\eta ^2) + \frac{2\sigma }{R_*^2} \right) \Lambda \\&\quad = -\rho _l^2R_*^2 (\xi ^2 + \eta ^2)^2 - \frac{4\mu _l}{R_*} \left( \rho _lR_*(2\xi ) + \frac{4\mu _l}{R_*} \right) (\xi ^2 + \eta ^2)\\&\qquad + \frac{2\sigma }{R_*^2} \left( 2\rho _lR_*\xi ^2 + 2\,\frac{4\mu _l}{R_*}\, \xi - 2\rho _lR_*\eta ^2 \right) - \left( \frac{2\sigma }{R_*^2} \right) ^2\\&\quad > -\rho _l^2R_*^2 (\xi ^2 + \eta ^2)^2 - \frac{2\sigma }{R_*^2}\, 2\rho _lR_*(\xi ^2+\eta ^2) - \left( \frac{2\sigma }{R_*^2} \right) ^2\\&\quad = -\left( \rho _lR_*(\xi ^2+\eta ^2) + \frac{2\sigma }{R_*^2} \right) ^2, \end{aligned} \end{aligned}$$

where we’ve used (E.6) and so \(2\rho _lR_*\xi ^2 + 2\,\frac{4\mu _l}{R_*}\, \xi = \xi \left( \rho _lR_*(2\xi ) + 2\,\frac{4\mu _l}{R_*} \right) > -2\rho _lR_*\xi ^2\) in the last inequality. This implies

$$\begin{aligned} \begin{aligned} \Lambda > -\left( \rho _lR_*(\xi ^2+\eta ^2) + \frac{2\sigma }{R_*^2} \right) . \end{aligned}\end{aligned}$$
(E.10)

Using (E.10) in (E.9), we get

$$\begin{aligned} \begin{aligned} 0&> - \frac{1}{{\mathscr {R}}_gT_\infty } \left( \frac{4\pi }{3\gamma } + \frac{8(\gamma -1)}{\pi \gamma }\sum _{j=1}^\infty \frac{\pi ^4\overline{\kappa }^2j^2}{\left( \pi ^2\overline{\kappa }j^2 + \xi \right) ^2 + \eta ^2} \right) \left( \rho _lR_*(\xi ^2+\eta ^2) + \frac{2\sigma }{R_*^2} \right) \\&\quad + 4\pi \, \frac{\rho _*}{R_*}. \end{aligned}\end{aligned}$$
(E.11)

Suppose

$$\begin{aligned} \begin{aligned} \xi \ge -\theta \pi ^2\overline{\kappa }, \end{aligned}\end{aligned}$$
(E.12)

where \(\theta \in (0,1)\), to be chosen. Then \(\xi \ge -\theta \pi ^2\overline{\kappa }j^2\) for all \(j=1,2,\ldots \). We further assume that

$$\begin{aligned} \begin{aligned} \xi \ge - \sqrt{\frac{2p_{\infty ,*}}{\rho _lR_*^2} + \frac{4\sigma }{\rho _lR_*^3} - \eta ^2}, \end{aligned}\end{aligned}$$
(E.13)

provided \(\eta ^2 \le \frac{2p_{\infty ,*}}{\rho _lR_*^2} + \frac{4\sigma }{\rho _lR_*^3}\), so that \(\rho _lR_*(\xi ^2+\eta ^2) + \frac{2\sigma }{R_*^2} \le \frac{2p_{\infty ,*}}{R_*} + \frac{6\sigma }{R_*^2}\). Then (E.11) gives

$$\begin{aligned} \begin{aligned} 0&> - \frac{1}{{\mathscr {R}}_gT_\infty } \left( \frac{4\pi }{3\gamma } + \frac{8(\gamma -1)}{\pi \gamma }\, \frac{1}{(1-\theta )^2} \sum _{j=1}^\infty \frac{1}{j^2} \right) \left( \frac{2p_{\infty ,*}}{R_*} + \frac{6\sigma }{R_*^2} \right) + 4\pi \, \frac{\rho _*}{R_*}. \end{aligned} \end{aligned}$$

Using \(\sum _{j=1}^\infty j^{-2} = \pi ^2/6\), one has

$$\begin{aligned} 3{\mathscr {R}}_gT_\infty \, \frac{\rho _*}{R_*} < \left[ \frac{1}{\gamma }+ \left( 1-\frac{1}{\gamma } \right) \frac{1}{(1-\theta )^2} \right] \left( \frac{2p_{\infty ,*}}{R_*} + \frac{6\sigma }{R_*^2} \right) , \end{aligned}$$

or, equivalently, using \({\mathscr {R}}_gT_\infty \rho _* = p_{\infty ,*} + 2\sigma /R_*\),

$$\begin{aligned} \theta > 1 - \sqrt{\dfrac{1-\frac{1}{\gamma }}{\frac{3p_{\infty ,*}R_*+6\sigma }{2p_{\infty ,*}R_*+6\sigma } - \frac{1}{\gamma }}}. \end{aligned}$$

We simply choose

$$\begin{aligned} \theta = 1 - \sqrt{\dfrac{1-\frac{1}{\gamma }}{\frac{3p_{\infty ,*}R_*+6\sigma }{2p_{\infty ,*}R_*+6\sigma } - \frac{1}{\gamma }}} \in (0,1) \end{aligned}$$

to reach a contradiction to (E.12) and (E.13). Therefore, we have for \(\eta \in {\mathbb {R}}\) with \(\eta \ne 0\) and \(\eta ^2\le \frac{2p_{\infty ,*}}{\rho _lR_*^2} + \frac{4\sigma }{\rho _lR_*^3}\) that

$$\begin{aligned} \xi < - \min \left\{ \left( 1 - \sqrt{\dfrac{1-\frac{1}{\gamma }}{\frac{3p_{\infty ,*}R_*+6\sigma }{2p_{\infty ,*}R_*+6\sigma } - \frac{1}{\gamma }}} \right) \pi ^2\overline{\kappa },\, \sqrt{\frac{2p_{\infty ,*}}{\rho _lR_*^2} + \frac{4\sigma }{\rho _lR_*^3} - \eta ^2} \right\} . \end{aligned}$$

Combining (E.7), for \(\eta \in {\mathbb {R}}\) with \(\eta \ne 0\) and \(\eta ^2\le \frac{p_{\infty ,*}}{\rho _lR_*^2} + \frac{2\sigma }{\rho _lR_*^3} \) we have

$$\begin{aligned} \begin{aligned} \xi < - \max \left\{ \frac{2\mu _l}{\rho _lR_*^2},\, \min \left\{ \left( 1 - \sqrt{\dfrac{1-\frac{1}{\gamma }}{\frac{3p_{\infty ,*}R_*+6\sigma }{2p_{\infty ,*}R_*+6\sigma } - \frac{1}{\gamma }}} \right) \pi ^2\overline{\kappa },\, \sqrt{ \frac{p_{\infty ,*}}{\rho _lR_*^2} + \frac{2\sigma }{\rho _lR_*^3} } \right\} \right\} . \end{aligned}\nonumber \\\end{aligned}$$
(E.14)

Now, we consider the case \(\eta ^2 > \frac{p_{\infty ,*}}{\rho _lR_*^2} + \frac{2\sigma }{\rho _lR_*^3} \). Since \(\eta \ne 0\), the imaginary part in (E.3) gives the identity \(\** _1 = - \frac{H_1}{H_2}\, \** _2\). Using this identity in the real part in (E.3), we derive

$$\begin{aligned} 4\pi \, \frac{\rho _*}{R_*}\, {\mathscr {R}}_gT_\infty H_2 = H_1\left( \** _2^2 + H_2^2 \right) , \end{aligned}$$

which implies

$$\begin{aligned} \begin{aligned}&4\pi \, \frac{\rho _*}{R_*}\, {\mathscr {R}}_gT_\infty \left( \rho _lR_*(2\xi ) + \frac{4\mu _l}{R_*} \right) \\&\quad = -\frac{8(\gamma -1)}{\pi \gamma } \sum _{j=1}^\infty \frac{\pi ^2\overline{\kappa }}{\left( \pi ^2\overline{\kappa }j^2 + \xi \right) ^2 + \eta ^2}\, \left( \** _2^2 + H_2^2 \right) . \end{aligned}\end{aligned}$$
(E.15)

To find a positive lower bound for \(\** _2^2 + H_2^2\), note that

$$\begin{aligned} \** _2^2 + H_2^2 = \left| \rho _lR_*\tau ^2 + \frac{4\mu _l}{R_*}\, \tau - \frac{2\sigma }{R_*^2} \right| ^2 = \rho _l^2R_*^2 \left| \tau - \tau _+ \right| ^2 \left| \tau - \tau _- \right| ^2, \end{aligned}$$

where

$$\begin{aligned} \tau _{\pm } = \dfrac{-\dfrac{4\mu _l}{R_*} \pm \sqrt{\left( \dfrac{4\mu _l}{R_*} \right) ^2 + 4\rho _lR_*\, \dfrac{2\sigma }{R_*^2}}}{2\rho _lR_*} \end{aligned}$$

are on the real axis. By the triangular inequality \(|\tau - \tau _{\pm }| > |\eta |\), and so

$$\begin{aligned} \** _2^2 + H_2^2 > \rho _l^2 R_*^2 \eta ^4. \end{aligned}$$

Therefore, (E.15) yields

$$\begin{aligned} 4\pi \, \frac{\rho _*}{R_*}\, {\mathscr {R}}_gT_\infty \left( \rho _lR_*(2\xi ) + \frac{4\mu _l}{R_*} \right) < -\frac{8(\gamma -1)}{\pi \gamma } \sum _{j=1}^\infty \frac{\pi ^2\overline{\kappa }}{\pi ^4\overline{\kappa }^2 j^4 + \eta ^2}\, \rho _l^2R_*^2\eta ^4. \end{aligned}$$

Since \(\eta ^2 > \frac{p_{\infty ,*}}{\rho _lR_*^2} + \frac{2\sigma }{\rho _lR_*^3}\) and \(\frac{a^2}{\pi ^4\overline{\kappa }^2 j^4 + a}\) is increasing in a for \(a = \eta ^2 \ge 0\), we further derive

$$\begin{aligned} \begin{aligned}&4\pi \, \frac{\rho _*}{R_*}\, {\mathscr {R}}_gT_\infty \left( \rho _lR_*(2\xi ) + \frac{4\mu _l}{R_*} \right) \\&\quad < -\frac{8(\gamma -1)}{\pi \gamma } \sum _{j=1}^\infty \frac{\pi ^2\overline{\kappa }}{\pi ^4\overline{\kappa }^2 j^4 + \left( \frac{p_{\infty ,*}}{\rho _lR_*^2} + \frac{2\sigma }{\rho _lR_*^3} \right) }\, \rho _l^2R_*^2 \left( \frac{p_{\infty ,*}}{\rho _lR_*^2} + \frac{2\sigma }{\rho _lR_*^3} \right) ^2\\&\quad = -\frac{8(\gamma -1)}{\pi \gamma }\, \frac{1}{\pi ^2\overline{\kappa }} \sum _{j=1}^\infty \frac{1}{j^4 + B^2}\, \rho _l^2R_*^2 \left( \frac{p_{\infty ,*}}{\rho _lR_*^2} + \frac{2\sigma }{\rho _lR_*^3} \right) ^2, \end{aligned}\end{aligned}$$
(E.16)

where

$$\begin{aligned} B:= \frac{1}{\pi ^2\overline{\kappa }}\sqrt{ \frac{p_{\infty ,*}}{\rho _lR_*^2} + \frac{2\sigma }{\rho _lR_*^3}}. \end{aligned}$$

Using

$$\begin{aligned} \begin{aligned} \sum _{j=1}^\infty \frac{1}{j^4 + B^2}&= \frac{e^{\frac{i\pi }{4}}\pi \cot \left( e^{\frac{i\pi }{4}}\pi \sqrt{B} \right) + e^{\frac{3i\pi }{4}}\pi \cot \left( e^{\frac{3i\pi }{4}}\pi \sqrt{B} \right) }{4B^{3/2}} - \frac{1}{2B^2}\\&= \dfrac{\frac{2\pi }{\sqrt{2}} \left( \frac{\cot \left( \pi \sqrt{B/2} \right) {{\,\textrm{csch}\,}}^2\left( \pi \sqrt{B/2} \right) + \csc ^2\left( \pi \sqrt{B/2} \right) \coth \left( \pi \sqrt{B/2} \right) }{\cot ^2\left( \pi \sqrt{B/2} \right) + \coth ^2\left( \pi \sqrt{B/2} \right) } \right) }{4B^{3/2}} - \frac{1}{2B^2}\\&= \dfrac{\frac{2\pi }{\sqrt{2}} \left( \frac{1}{\pi \sqrt{B/2}} + \frac{4}{45} \left( \pi \sqrt{B/2} \right) ^3 + O\left( \left( \pi \sqrt{B/2} \right) ^6 \right) \right) }{4B^{\frac{3}{2}}} - \frac{1}{2B^2}\\&= \frac{\pi ^4}{90} + O\left( B^{3/2} \right) \ \text { for }B\ll 1, \end{aligned} \end{aligned}$$

we have from (E.16) that

$$\begin{aligned} \begin{aligned}&4\pi \, \frac{\rho _*}{R_*}\, {\mathscr {R}}_gT_\infty \left( \rho _lR_*(2\xi ) + \frac{4\mu _l}{R_*} \right) \\&\quad < -\frac{8(\gamma -1)}{\pi \gamma }\, \frac{1}{\pi ^2\overline{\kappa }} \left( \frac{\pi ^4}{90} + O\left( \left( \frac{1}{\pi ^2\overline{\kappa }}\sqrt{ \frac{p_{\infty ,*}}{\rho _lR_*^2} + \frac{2\sigma }{\rho _lR_*^3} } \right) ^{3/2} \right) \right) \rho _l^2R_*^2 \left( \frac{p_{\infty ,*}}{\rho _lR_*^2} + \frac{2\sigma }{\rho _lR_*^3} \right) ^2. \end{aligned} \end{aligned}$$

Consequently, we have for \(\eta ^2 > \frac{p_{\infty ,*}}{\rho _lR_*^2} + \frac{2\sigma }{\rho _lR_*^3} \) that

$$\begin{aligned} \begin{aligned} \xi < -\frac{2\mu _l}{\rho _lR_*^2} - \frac{{\mathscr {R}}_gT_\infty \rho _*}{\pi ^4\overline{\kappa }\rho _lR_*^2} \left( 1-\frac{1}{\gamma } \right) \left( \frac{\pi ^4}{90} + O\left( \left( \frac{1}{\pi ^2\overline{\kappa }}\sqrt{\frac{p_{\infty ,*}}{\rho _lR_*^2} + \frac{2\sigma }{\rho _lR_*^3}} \right) ^{3/2} \right) \right) , \end{aligned}\nonumber \\\end{aligned}$$
(E.17)

where \({\mathscr {R}}_gT_\infty \rho _* = p_{\infty ,*} + 2\sigma /R_*\) is used.

It remains to consider the case \(\eta =0\). We first show that \(\xi <0\). Suppose for the sake of contradiction that \(\xi \ge 0\) then

$$\begin{aligned} \begin{aligned} Q(\xi )&= \frac{1}{{\mathscr {R}}_gT_\infty } \left( \frac{4\pi }{3\gamma } + \frac{8(\gamma -1)}{\pi \gamma } \sum _{j=1}^\infty \frac{\pi ^2\overline{\kappa }}{\pi ^2\overline{\kappa }j^2 + \xi } \right) \left( \rho _lR_*\xi ^2 + \frac{4\mu _l}{R_*}\, \xi \right) \\&\quad - \frac{2\sigma }{{\mathscr {R}}_gT_\infty R_*^2} \left( \frac{4\pi }{3\gamma } + \frac{8(\gamma -1)}{\pi \gamma } \sum _{j=1}^\infty \frac{\pi ^2\overline{\kappa }}{\pi ^2\overline{\kappa }j^2 + \xi } \right) + 4\pi \,\frac{\rho _*}{R_*}, \end{aligned} \end{aligned}$$

where the second line is greater than

$$\begin{aligned} - \frac{2\sigma }{{\mathscr {R}}_gT_\infty R_*^2} \left( \frac{4\pi }{3\gamma } + \frac{8(\gamma -1)}{\pi \gamma } \sum _{j=1}^\infty \frac{\pi ^2\overline{\kappa }}{\pi ^2\overline{\kappa }j^2} \right) + 4\pi \,\frac{\rho _*}{R_*} = \frac{8\pi \rho _*}{3R_*} + \frac{4\pi p_{\infty ,*}}{3{\mathscr {R}}_gT_\infty R_*} > 0. \end{aligned}$$

Here the identities \(\sum _{j=1}^\infty j^{-2} = \frac{\pi ^2}{6}\) and \({\mathscr {R}}_gT_\infty \rho _* = p_{\infty ,*} + \frac{2\sigma }{R_*}\) are used. This yields \(Q(\xi )>0\), which contradicts to the fact that \(Q(\tau )=0\). Thus, we have \(\xi <0\) for \(\eta =0\).

Now we search for a negative upper bound for \(\xi \) when \(\eta =0\). Suppose

$$\begin{aligned} \begin{aligned} \xi \ge -\theta _0\pi ^2\overline{\kappa }, \end{aligned}\end{aligned}$$
(E.18)

where \(0<\theta _0<1\) to be chosen. Suppose further that

$$\begin{aligned} \begin{aligned} \xi> \frac{-\frac{4\mu _l}{R_*} + \sqrt{\Delta }}{2\rho _lR_*}\quad \text { if }\Delta := \left( \frac{4\mu _l}{R_*} \right) ^2 - 4\rho _lR_* \left( \frac{2p_{\infty ,*}}{R_*} + \frac{4\sigma }{R_*^2} \right) >0 \end{aligned}\end{aligned}$$
(E.19)

such that

$$\begin{aligned} \begin{aligned} \rho _lR_*\xi ^2 + \frac{4\mu _l}{R_*}\, \xi - \frac{2\sigma }{R_*^2} \ge -\frac{2p_{\infty ,*}}{R_*} - \frac{6\sigma }{R_*^2}. \end{aligned}\end{aligned}$$
(E.20)

Note that the inequality (E.20) always holds when \(\Delta \le 0\). Then

$$\begin{aligned} \begin{aligned} 0 = Q(\xi )&= \frac{1}{{\mathscr {R}}_gT_\infty } \left( \frac{4\pi }{3\gamma } + \frac{8(\gamma -1)}{\pi \gamma } \sum _{j=1}^\infty \frac{\pi ^2\overline{\kappa }}{\pi ^2\overline{\kappa }j^2 + \xi } \right) \left( \rho _lR_*\xi ^2 + \frac{4\mu _l}{R_*}\, \xi - \frac{2\sigma }{R_*^2} \right) \\&\quad + 4\pi \,\frac{\rho _*}{R_*}\\&> - \frac{1}{{\mathscr {R}}_gT_\infty } \left( \frac{4\pi }{3\gamma } + \frac{8(\gamma -1)}{\pi \gamma (1-\theta _0)} \sum _{j=1}^\infty \frac{\pi ^2\overline{\kappa }}{\pi ^2\overline{\kappa }j^2} \right) \left( \frac{2p_{\infty ,*}}{R_*} + \frac{6\sigma }{R_*^2} \right) + 4\pi \,\frac{\rho _*}{R_*}\\&= - \frac{1}{{\mathscr {R}}_gT_\infty } \left( \frac{4\pi }{3\gamma } + \frac{4\pi }{3} \left( 1-\frac{1}{\gamma } \right) \frac{1}{1-\theta _0} \right) \left( \frac{2p_{\infty ,*}}{R_*} + \frac{6\sigma }{R_*^2} \right) + 4\pi \,\frac{\rho _*}{R_*} \end{aligned} \end{aligned}$$

So

$$\begin{aligned} 4\pi {\mathscr {R}}_gT_\infty \, \frac{\rho _*}{R_*} < \left( \frac{4\pi }{3\gamma } + \frac{4\pi }{3} \left( 1-\frac{1}{\gamma } \right) \frac{1}{1-\theta _0} \right) \left( \frac{2p_{\infty ,*}}{R_*} + \frac{6\sigma }{R_*^2} \right) , \end{aligned}$$

or, equivalently,

$$\begin{aligned} \theta _0 > 1 - \dfrac{1-\dfrac{1}{\gamma }}{\dfrac{3{\mathscr {R}}_gT_\infty \, \frac{\rho _*}{R_*}}{\frac{2p_{\infty ,*}}{R_*} + \frac{6\sigma }{R_*^2}} - \dfrac{1}{\gamma } } = 1 - \dfrac{1-\dfrac{1}{\gamma }}{ \frac{3p_{\infty ,*}R_*+6\sigma }{2p_{\infty ,*}R_*+6\sigma } - \dfrac{1}{\gamma } }, \end{aligned}$$

where \({\mathscr {R}}_gT_\infty \rho _* = p_{\infty ,*} + \frac{2\sigma }{R_*}\) has been used in the last equation. We then choose

$$\begin{aligned} \theta _0 = 1 - \dfrac{1-\dfrac{1}{\gamma }}{ \dfrac{3p_{\infty ,*}R_*+6\sigma }{2p_{\infty ,*}R_*+6\sigma } - \dfrac{1}{\gamma } } \in (0,1) \end{aligned}$$

to reach a contradiction to (E.18) and (E.19). Hence we derive for \(\eta =0\)

$$\begin{aligned} \begin{aligned} \xi < {\left\{ \begin{array}{ll} - \min \left\{ \left( 1 - \dfrac{1-\dfrac{1}{\gamma }}{ \dfrac{3p_{\infty ,*}R_*+6\sigma }{2p_{\infty ,*}R_*+6\sigma } - \dfrac{1}{\gamma } } \right) \pi ^2\overline{\kappa },\, \dfrac{ \dfrac{4\mu _l}{R_*} - \sqrt{\Delta }}{2\rho _lR_*} \right\} &{}\ \text { if }\Delta >0,\\ - \left( 1 - \dfrac{1-\dfrac{1}{\gamma }}{ \dfrac{3p_{\infty ,*}R_*+6\sigma }{2p_{\infty ,*}R_*+6\sigma } - \dfrac{1}{\gamma } } \right) \pi ^2\overline{\kappa }&{}\ \text { if }\Delta \le 0. \end{array}\right. } \end{aligned}\end{aligned}$$
(E.21)

Combining the upper bounds (E.14), (E.17), and (E.21) for different cases of \(\eta \) and using the identity \({\mathscr {R}}_gT_\infty \rho _* = p_{\infty ,*} + \frac{2\sigma }{R_*}\), the lemma follows. \(\quad \square \)

Appendix F: Rate of Convergence of Slow Solutions Approaching to Center Manifold for a Class of Fully Nonlinear Autonomous Systems

As mentioned in the paragraph below Theorem 6.7, there are several obstacles preventing us from direct applying center manifold theorem to prove the exponential decay in nonlinear bubble oscillations. One of which is the quasilinear character of the problem (9.5). For this purpose, we develop in this appendix a geometric theory for a class of fully nonlinear autonomous systems which covers the quasilinear system (9.5).

We study a larger class of fully nonlinear autonomous systems of the form \(\dot{\textbf{w}} = {\mathcal {L}}{\textbf{w}} + {\mathcal {N}}({\textbf{w}}, \dot{\textbf{w}})\) that includes the quasilinear autonomous system (9.5) of our concern. Assuming that the solution of such equation converges toward a given center manifold and that the time derivative of the solution is sufficiently small for all time, we prove that the convergence rate is exponential. The proof is an adaptation of [10, Sections 2.4 and 6.3], where the existence and stability of center manifold for semilinear equations are established.

Setup of the fully nonlinear autonomous system, assumptions on the solution, and the center manifold. Let Z be a Banach space with norm \(\left\| \,\cdot \, \right\| \). We consider the evolution equation

$$\begin{aligned} \begin{aligned} \dot{\textbf{w}} = {\mathcal {L}}{\textbf{w}} + {\mathcal {N}}({\textbf{w}}, \dot{\textbf{w}}),\quad {\textbf{w}}(0) \in Z, \end{aligned}\end{aligned}$$
(F.1)

where \({\mathcal {N}}({\textbf{w}},\textbf{p}):Z\times Z\rightarrow Z\) has a uniformly continuous second derivative with \({\mathcal {N}}({\textbf{0}},{\textbf{p}}) = {\textbf{0}}\) and \(\partial _{({\textbf{w}}, {\textbf{p}})}{\mathcal {N}}({\textbf{0}}, {\textbf{0}}) = {\textbf{O}}\).

Assume

(i) \(Z = X\,\oplus \, Y\) where X is finite dimensional and Y is closed.

(ii) X is \({\mathcal {L}}\)-invariant and that if \({\mathcal {A}}:= {\mathcal {L}}|_X\), then the real parts of the eigenvalues of \({\mathcal {A}}\) are all zeros.

(iii) Y is \(e^{{\mathcal {L}} t}\)-invariant. Let \(Q_1\) be a projection on X (not necessarily along Y) and \(Q_2:= I - Q_1\). For some positive constants bc,

$$\begin{aligned} \begin{aligned} \left\| e^{{\mathcal {L}} t} Q_2 \right\| \le ce^{-bt},\quad t\ge 0. \end{aligned}\end{aligned}$$
(F.2)

Let \({\textbf{w}}\) be a solution of (F.1). Decompose \({\textbf{w}}\) into \({\textbf{w}} = {\textbf{x}} + {\textbf{y}}\) where \(\textbf{x} = Q_1{\textbf{w}}\) and \({\textbf{y}} = Q_2{\textbf{w}}\). Let \({\mathcal {B}} = Q_2{\mathcal {L}}\). Then equation (9.12) can be written as

$$\begin{aligned} \begin{aligned} \dot{\textbf{x}}&= {\mathcal {A}}{\textbf{x}} + f({\textbf{x}}, {\textbf{y}}, \dot{\textbf{x}}, \dot{\textbf{y}}),\\ \dot{\textbf{y}}&= {\mathcal {B}}{\textbf{y}} + g({\textbf{x}}, {\textbf{y}}, \dot{\textbf{x}}, \dot{\textbf{y}}), \end{aligned}\end{aligned}$$
(F.3)

where

$$\begin{aligned} f({\textbf{x}}, {\textbf{y}}, \dot{\textbf{x}}, \dot{\textbf{y}}) = Q_1{\mathcal {N}}({\textbf{x}}+{\textbf{y}}, \dot{\textbf{x}}+\dot{\textbf{y}}),\quad g({\textbf{x}}, {\textbf{y}}, \dot{\textbf{x}}, \dot{\textbf{y}}) = Q_2{\mathcal {N}}({\textbf{x}}+{\textbf{y}}, \dot{\textbf{x}}+\dot{\textbf{y}}). \end{aligned}$$

A curve \({\textbf{y}} = h({\textbf{x}})\), defined for \(|{\textbf{x}}|\) small, is said to be an invariant manifold for (F.3) if the solution \(({\textbf{x}}(t),\textbf{y}(t))\) of (F.3) through \((\textbf{x}(0),h({\textbf{x}}(0)))\) satisfies \({\textbf{y}}(t) = h({\textbf{x}}(t))\). A center manifold for (F.3) is an invariant manifold that is tangent to X space at the origin.

By the assumption on the nonlinearity \({\mathcal {N}}(\textbf{w},\dot{\textbf{w}})\), there exists a continuous function \(k(\varepsilon )\) with \(k(0)=0\) such that

$$\begin{aligned} \begin{aligned} \left\| f({\textbf{x}}, {\textbf{y}}, \dot{\textbf{x}}, \dot{\textbf{y}}) \right\| + \left\| g({\textbf{x}}, {\textbf{y}}, \dot{\textbf{x}}, \dot{\textbf{y}}) \right\|&\le \varepsilon k(\varepsilon ),\\ \left\| f({\textbf{x}}, {\textbf{y}}, \dot{\textbf{x}}, \dot{\textbf{y}}) - f({\textbf{x}}', {\textbf{y}}', \dot{\textbf{x}}', \dot{\textbf{y}}') \right\|&\le k(\varepsilon ) \left[ \left\| \textbf{x} - {\textbf{x}}' \right\| + \left\| {\textbf{y}} - {\textbf{y}}' \right\| + \left\| \dot{\textbf{x}} - \dot{\textbf{x}}' \right\| + \left\| \dot{\textbf{y}} - \dot{\textbf{y}}' \right\| \right] ,\\ \left\| g({\textbf{x}}, {\textbf{y}}, \dot{\textbf{x}}, \dot{\textbf{y}}) - g({\textbf{x}}', {\textbf{y}}', \dot{\textbf{x}}', \dot{\textbf{y}}') \right\|&\le k(\varepsilon ) \left[ \left\| \textbf{x} - {\textbf{x}}' \right\| + \left\| {\textbf{y}} - {\textbf{y}}' \right\| + \left\| \dot{\textbf{x}} - \dot{\textbf{x}}' \right\| + \left\| \dot{\textbf{y}} - \dot{\textbf{y}}' \right\| \right] , \end{aligned}\nonumber \\\end{aligned}$$
(F.4)

for all \(\textbf{x}, {\textbf{x}}'\in X\), \({\textbf{y}}, {\textbf{y}}'\in Y\) and all \(\dot{\textbf{x}}, \dot{\textbf{x}}', \dot{\textbf{y}}, \dot{\textbf{y}}' \in Z\) with \(\left\| ({\textbf{x}}, {\textbf{y}}) \right\| , \left\| (\dot{\textbf{x}}, \dot{\textbf{y}}) \right\| , \left\| ({\textbf{x}}',\textbf{y}') \right\| , \left\| (\dot{\textbf{x}}', \dot{\textbf{y}}') \right\| < \varepsilon \).

Let \({\mathcal {M}}\) be a center manifold for (F.3) given by \({\textbf{y}} = h({\textbf{x}})\). If we substitute \({\textbf{y}}(t)=h({\textbf{x}}(t))\) into (F.3) we obtain

$$\begin{aligned} \begin{aligned} h'({\textbf{x}}) \left[ {\mathcal {A}}{\textbf{x}} + f({\textbf{x}}, h({\textbf{x}}), \dot{\textbf{x}}, h'({\textbf{x}})\dot{\textbf{x}}) \right] = \mathcal Bh({\textbf{x}}) + g({\textbf{x}}, h({\mathbf{x)}}, \dot{\textbf{x}}, h'({\textbf{x}})\dot{\textbf{x}}). \end{aligned}\end{aligned}$$
(F.5)

The equation on the center manifold is given by

$$\begin{aligned} \begin{aligned} \dot{\textbf{u}} = {\mathcal {A}}{\textbf{u}} + f({\textbf{u}}, h({\textbf{u}}), \dot{\textbf{u}}, h'({\textbf{u}})\dot{\textbf{u}}). \end{aligned}\end{aligned}$$
(F.6)

We assume that \({\textbf{w}}(t)\) converges to some element in \({\mathcal {M}}\), as \(t\rightarrow \infty \), and that \(\sup _{t\ge 0} \left\| \dot{\textbf{w}}(t) \right\| \) is sufficiently small.

Rate of convergence to the center manifold. The following lemma describes that the trajectory shadows the center manifold and corresponds to [10, Lemma 2.4.1].

Lemma F.1

Let \(({\textbf{x}}(t), {\textbf{y}}(t))\) be a solution of (F.3) with \(\left\| ({\textbf{x}}(0), \textbf{y}(0)) \right\| \) and \(\left\| (\dot{\textbf{x}}(t), \dot{\textbf{y}}(t)) \right\| \), for all \(t\ge 0\), sufficiently small. Then there exist positive \(C_1\) and \(\beta _1\) such that

$$\begin{aligned} \left\| {\textbf{y}}(t) - h({\textbf{x}}(t)) \right\| \le C_1 e^{-\beta _1 t} \left\| \textbf{y}(0) - h({\textbf{x}}(0)) \right\| \end{aligned}$$

for all \(t\ge 0\).

Proof

Let \(({\textbf{x}}(t),{\textbf{y}}(t))\) be a solution of (F.3) with \(({\textbf{x}}(0),{\textbf{y}}(0))\) sufficiently small. Let \({\textbf{z}}(t) = {\textbf{y}}(t) - h({\textbf{x}}(t))\), then

$$\begin{aligned} \begin{aligned} \dot{\textbf{z}} = {\mathcal {B}}{\textbf{z}} + {\mathcal {R}}({\textbf{x}},\textbf{z},\dot{\textbf{x}},\dot{\textbf{z}}) \end{aligned}\end{aligned}$$
(F.7)

where

$$\begin{aligned} \begin{aligned} {\mathcal {R}}({\textbf{x}},{\textbf{z}},\dot{\textbf{x}},\dot{\textbf{z}})&= h'(\textbf{x}) \left[ f({\textbf{x}}, h({\textbf{x}}), \dot{\textbf{x}}, h'({\textbf{x}})\dot{\textbf{x}}) - f({\textbf{x}},{\textbf{z}}+h({\textbf{x}}), \dot{\textbf{x}}, \dot{\textbf{z}} + h'(\textbf{x})\dot{\textbf{x}}) \right] \\&\quad + g({\textbf{x}},{\textbf{z}}+h({\textbf{x}}), \dot{\textbf{x}}, \dot{\textbf{z}} + h'({\textbf{x}})\dot{\textbf{x}}) - g({\textbf{x}}, h({\textbf{x}}), \dot{\textbf{x}}, h'({\textbf{x}})\dot{\textbf{x}}). \end{aligned}\end{aligned}$$
(F.8)

Using the hypotheses of f and g and the bounds on h,

$$\begin{aligned} \begin{aligned} \left\| {\mathcal {R}}({\textbf{x}},{\textbf{z}},\dot{\textbf{x}},\dot{\textbf{z}}) \right\|&\le \left\| h'({\textbf{x}}) \right\| \Big ( \left\| f({\textbf{x}}, h({\textbf{x}}), \dot{\textbf{x}}, h'({\textbf{x}})\dot{\textbf{x}}) - f({\textbf{x}},{\textbf{z}}+h({\textbf{x}}), \dot{\textbf{x}}, h'({\textbf{x}})\dot{\textbf{x}}) \right\| \\&\quad + \left\| f({\textbf{x}}, {\textbf{z}} + h({\textbf{x}}), \dot{\textbf{x}}, h'({\textbf{x}})\dot{\textbf{x}}) - f({\textbf{x}},{\textbf{z}}+h({\textbf{x}}), \dot{\textbf{x}}, \dot{\textbf{z}} + h'({\textbf{x}})\dot{\textbf{x}}) \right\| \Big )\\&\quad + \left\| g({\textbf{x}},{\textbf{z}}+h({\textbf{x}}), \dot{\textbf{x}}, \dot{\textbf{z}} + h'({\textbf{x}})\dot{\textbf{x}}) - g({\textbf{x}}, h({\textbf{x}}), \dot{\textbf{x}}, \dot{\textbf{z}} + h'({\textbf{x}})\dot{\textbf{x}}) \right\| \\&\quad + \left\| g({\textbf{x}}, h({\textbf{x}}), \dot{\textbf{x}}, \dot{\textbf{z}} + h'({\textbf{x}})\dot{\textbf{x}}) - g({\textbf{x}}, h({\textbf{x}}), \dot{\textbf{x}}, h'({\textbf{x}})\dot{\textbf{x}}) \right\| \\&\le \delta (\varepsilon ) \left[ \left\| \textbf{z} \right\| + \left\| \dot{\textbf{z}} \right\| \right] , \end{aligned}\end{aligned}$$
(F.9)

if \(\left\| \textbf{z} \right\| , \left\| \dot{\textbf{z}} \right\| < \varepsilon \), for some continuous function \(\delta (\varepsilon )\) with \(\delta (0)=0\). Using (F.2) we obtain, from (F.7),

$$\begin{aligned} \begin{aligned} \left\| {\textbf{z}}(t) \right\| \le ce^{-bt} \left\| {\textbf{z}}(0) \right\| + c\delta (\varepsilon ) \int _0^t e^{-b(t-s)} \left[ \left\| \mathbf{z(s)} \right\| + \left\| \dot{\textbf{z}}(s) \right\| \right] \textrm{d}s. \end{aligned}\end{aligned}$$
(F.10)

Using (F.9) in (F.7) one has

$$\begin{aligned} \left\| \dot{\textbf{z}} \right\| \le \left\| {\mathcal {B}} \right\| \left\| \textbf{z} \right\| + \left\| {\mathcal {R}}({\textbf{x}},{\textbf{z}},\dot{\textbf{x}},\dot{\textbf{z}}) \right\| \le \left\| {\mathcal {B}} \right\| \left\| \textbf{z} \right\| + \delta (\varepsilon )\left[ \left\| \textbf{z} \right\| + \left\| \dot{\textbf{z}} \right\| \right] , \end{aligned}$$

and so

$$\begin{aligned} \begin{aligned} \left\| \dot{\textbf{z}} \right\| \le C_0 \left\| \textbf{z} \right\| ,\qquad C_0 = \left( 1-\delta (\varepsilon ) \right) ^{-1} \left( \left\| {\mathcal {B}} \right\| + \delta (\varepsilon ) \right) . \end{aligned}\end{aligned}$$
(F.11)

Therefore, (F.10) yields

$$\begin{aligned} \begin{aligned} e^{bt} \left\| {\textbf{z}}(t) \right\|&\le c \left\| {\textbf{z}}(0) \right\| + c\delta (\varepsilon ) \left( 1 + C_0 \right) \int _0^t e^{bs} \left\| \mathbf{z(s)} \right\| \textrm{d}s \end{aligned}\end{aligned}$$
(F.12)

By Gronwall’s lemma,

$$\begin{aligned} e^{bt} \left\| {\textbf{z}}(t) \right\| \le c \left\| {\textbf{z}}(0) \right\| e^{c\delta (\varepsilon ) \left( 1 + C_0 \right) t }. \end{aligned}$$

The lemma follows. \(\square \)

Proposition F.2

Suppose that the zero solution of (F.6) is Lyapunov stable. Let \(({\textbf{x}}(t), {\textbf{y}}(t))\) be a solution of (F.3). There exists \(\varepsilon >0\) such that if \(\left\| ({\textbf{x}}(0), {\textbf{y}}(0)) \right\| < \varepsilon \) and if \(\left\| (\dot{\textbf{x}}(t), \dot{\textbf{y}}(t)) \right\| < \varepsilon \) for all \(t\ge 0\), then there exists a solution \({\textbf{u}}(t)\) of (F.6) such that as \(t\rightarrow \infty \),

$$\begin{aligned} \begin{aligned} {\textbf{x}}(t)&= {\textbf{u}}(t) + O(e^{-b_1 t}),\\ {\textbf{y}}(t)&= h({\textbf{u}}(t)) + O(e^{-b_1 t}), \end{aligned}\end{aligned}$$
(F.13)

where \(b_1 = \min (b, \beta _1)\), b and \(\beta _1\) are given in the assumption (F.2) and in Lemma F.1, respectively.

Proof

The proof is based on that of [10, Theorem 2.4.2]. Let \(({\textbf{x}}(t), {\textbf{y}}(t))\) be a solution of (F.3). Since the zero solution of (F.6) is Lyapunov stable, solutions \({\textbf{u}}(t)\) of (F.6) are Lyapunov stable if \({\textbf{u}}(0)\) is sufficiently small. Let \({\textbf{u}}(t)\) be a solution of (F.6) with \({\textbf{u}}(0)\) sufficiently small. Let \({\textbf{z}}(t) = {\textbf{y}}(t) - h({\textbf{x}}(t))\), \(\varvec{\phi }(t) = {\textbf{x}}(t) - {\textbf{u}}(t)\). Then

$$\begin{aligned} \dot{\textbf{z}}&= {\mathcal {B}}{\textbf{z}} + {\mathcal {R}}(\varvec{\phi } + {\textbf{u}},\textbf{z},\dot{\varvec{\phi }} + \dot{\textbf{u}},\dot{\textbf{z}}), \end{aligned}$$
(F.14a)
$$\begin{aligned} \dot{\varvec{\phi }}&= {\mathcal {A}}\varvec{\phi } + {\mathcal {V}}(\varvec{\phi },\textbf{z},\dot{\varvec{\phi }},\dot{\textbf{z}}), \end{aligned}$$
(F.14b)

where \({\mathcal {R}}\) is defined in (F.8) and

$$\begin{aligned} \begin{aligned} {\mathcal {V}}(\varvec{\phi },{\textbf {z}},\dot{\varvec{\phi }},\dot{{\textbf {z}}})&= f({\textbf {u}} {+} \varvec{\phi }, {{\textbf {z}}} + h({{\textbf {u}}} {+} \varvec{\phi }), \dot{{\textbf {u}}} {+} \dot{\varvec{\phi }}, \dot{{\textbf {z}}} + h'({\textbf {u}} {+} \varvec{\phi }) (\dot{{\textbf {u}}} {+} \dot{\varvec{\phi }}))\\ {}&\quad - f({{\textbf {u}}}, h({{\textbf {u}}}), \dot{{\textbf {u}}}, h'({{\textbf {u}}})\dot{{\textbf {u}}}). \end{aligned} \end{aligned}$$

We now formulate (F.14a)–(F.14b) as a fixed point problem. Let \({\mathscr {X}}\) be the set of continuous differentiable functions \(\varvec{\phi }:[0,\infty )\rightarrow X\) of (F.14b) with \(\left\| \varvec{\phi }(t) e^{at} \right\| \le 1\) and \(\left\| \dot{\varvec{\phi }}(t) e^{at} \right\| \le a\) for all \(t\ge 0\), where \(a=b/2\) in which b is defined in (F.2). We define the norm \(\left\| \varvec{\phi } \right\| _{{\mathscr {X}}} = \sup \{\left\| \varvec{\phi }(t) e^{at} \right\| + \left\| \dot{\varvec{\phi }}(t) e^{at} \right\| : t\ge 0\}\). By the assumption on the operator \({\mathcal {A}}\), we can decompose \({\mathcal {A}}\) into \({\mathcal {A}} = {\mathcal {A}}_1 + {\mathcal {A}}_2\) where

$$\begin{aligned} \begin{aligned} \left\| e^{{\mathcal {A}}_1 t}{\textbf{x}} \right\| = \left\| {\textbf{x}} \right\| ,\qquad \left\| {\mathcal {A}}_2{\textbf{x}} \right\| \le (b/4) \left\| {\textbf{x}} \right\| , \end{aligned}\end{aligned}$$
(F.15)

where b is defined by (F.2). Then (F.14b) can be written as

$$\begin{aligned} \dot{\varvec{\phi }} = {\mathcal {A}}_1\varvec{\phi } + \left[ {\mathcal {A}}_2\varvec{\phi } + {\mathcal {V}}(\varvec{\phi },\textbf{z},\dot{\varvec{\phi }},\dot{\textbf{z}}) \right] , \end{aligned}$$

and \(\varvec{\phi }(\infty ) = 0\) for \(\varvec{\phi }\in {\mathscr {X}}\). Let \({\textbf{z}}(t)\) be a given solution of (F.14a). By Duhamel’s formula, a solution \(\varvec{\phi }\in {\mathscr {X}}\) of (F.14b) must satisfy

$$\begin{aligned} \varvec{\phi }(t) = -\int _t^\infty e^{{\mathcal {A}}_1(t-s)} \left[ {\mathcal {A}}_2\varvec{\phi }(s) + {\mathcal {V}}(\varvec{\phi }(s),\textbf{z}(s),\dot{\varvec{\phi }}(s),\dot{\textbf{z}}(s)) \right] \textrm{d}s. \end{aligned}$$

Thus, a solution \(\varvec{\phi }\in {\mathscr {X}}\) of (F.14b) is a fixed point of the map** T that defined by

$$\begin{aligned} \begin{aligned} (T\varvec{\phi })(t) = -\int _t^\infty e^{{\mathcal {A}}_1(t-s)} \left[ {\mathcal {A}}_2\varvec{\phi }(s) + {\mathcal {V}}(\varvec{\phi }(s),\textbf{z}(s),\dot{\varvec{\phi }}(s),\dot{\textbf{z}}(s)) \right] \textrm{d}s. \end{aligned}\end{aligned}$$
(F.16)

Using the bounds on f, g, h, and the fact that \({\mathcal {R}}({\textbf{x}},{\textbf{0}},\dot{\textbf{x}},{\textbf{0}}) = {\textbf{0}}\) and \({\mathcal {V}}({\textbf{0}}, {\textbf{0}}, {\textbf{0}}, {\textbf{0}}) = {\textbf{0}}\), there is a continuous function \(k(\varepsilon )\) with \(k(0)=0\) such that if \(\varvec{\phi }_1,\varvec{\phi }_2\in X\), \({\textbf{z}}_1, \textbf{z}_2\in Y\), and \(\dot{\varvec{\phi }}_1,\dot{\varvec{\phi }}_2, \dot{\textbf{z}}_1, \dot{\textbf{z}}_2\in Z\) with \(\left\| (\varvec{\phi }_i, {\textbf{z}}_i) \right\| \), \(\left\| (\dot{\varvec{\phi }}_i, \dot{\textbf{z}}_i) \right\| < \varepsilon \), \(i=1,2\), then

$$\begin{aligned} \begin{aligned}&\left\| {\mathcal {R}}(\varvec{\phi }_1,\textbf{z}_1,\dot{\varvec{\phi }}_1,\dot{\textbf{z}}_1) - {\mathcal {R}}(\varvec{\phi }_2,\textbf{z}_2,\dot{\varvec{\phi }}_2,\dot{\textbf{z}}_2) \right\| \\&\quad \le k(\varepsilon ) \left[ \left\| ({\textbf{z}}_1, \dot{\textbf{z}}_1) \right\| \left( \left\| \varvec{\phi }_1 - \varvec{\phi }_2 \right\| + \left\| \dot{\varvec{\phi }}_1 - \dot{\varvec{\phi }}_2 \right\| \right) + \left\| {\textbf{z}}_1 - {\textbf{z}}_2 \right\| + \left\| \dot{\textbf{z}}_1 - \dot{\textbf{z}}_2 \right\| \right] ,\\&\left\| {\mathcal {V}}(\varvec{\phi }_1,\textbf{z}_1,\dot{\varvec{\phi }}_1,\dot{\textbf{z}}_1) - {\mathcal {V}}(\varvec{\phi }_2,\textbf{z}_2,\dot{\varvec{\phi }}_2,\dot{\textbf{z}}_2) \right\| \le k(\varepsilon ) \left[ \left\| {\textbf{z}}_1 - {\textbf{z}}_2 \right\| + \left\| \dot{\textbf{z}}_1 - \dot{\textbf{z}}_2 \right\| + \left\| \varvec{\phi }_1 - \varvec{\phi }_2 \right\| + \left\| \dot{\varvec{\phi }}_1 - \dot{\varvec{\phi }}_2 \right\| \right] . \end{aligned}\end{aligned}$$
(F.17)

By the same argument as in the proof of Lemma F.1, one has

$$\begin{aligned} \begin{aligned} \left\| {\textbf{z}}(t) \right\| \le C_1 \left\| {\textbf{z}}(0) \right\| e^{-\beta _1 t}, \end{aligned}\end{aligned}$$
(F.18)

where

$$\begin{aligned} \beta _1 = b - ck(\varepsilon ) \left( 1 + C_0 \right) ,\qquad C_0 = \left( 1-k(\varepsilon ) \right) ^{-1} \left( \left\| {\mathcal {B}} \right\| + k(\varepsilon ) \right) . \end{aligned}$$

Using (F.15), we obtain, from (F.16), that

$$\begin{aligned} \begin{aligned} \left\| T\varvec{\phi }(t) \right\|&\le \int _t^\infty \frac{b}{4} |\varvec{\phi }(s)|\, \textrm{d}s + \int _t^\infty \left\| {\mathcal {V}}(\varvec{\phi }(s),\textbf{z}(s),\dot{\varvec{\phi }}(s),\dot{\textbf{z}}(s)) \right\| \textrm{d}s\\&= \frac{a}{4} \int _t^\infty \left\| \varvec{\phi }(s) \right\| \textrm{d}s + \int _t^\infty \left\| {\mathcal {V}}(\varvec{\phi }(s),\textbf{z}(s),\dot{\varvec{\phi }}(s),\dot{\textbf{z}}(s)) - {\mathcal {V}}({\textbf{0}}, {\textbf{0}}, {\textbf{0}}, {\textbf{0}}) \right\| \textrm{d}s\\&\le \frac{e^{-at}}{2} + k(\varepsilon ) \int _t^\infty \left[ \left\| \varvec{\phi }(s) \right\| + \left\| {\textbf{z}}(s) \right\| + \left\| \dot{\varvec{\phi }}(s) \right\| + \left\| \dot{\textbf{z}}(s) \right\| \right] \textrm{d}s, \end{aligned}\nonumber \\\end{aligned}$$
(F.19)

where we’ve used \({\mathcal {V}}({\textbf{0}}, {\textbf{0}}, {\textbf{0}}, {\textbf{0}}) = \textbf{0}\) in the second equation and (F.17) in the last inequality. In view of (F.11) in the proof of Lemma F.1 and (F.18),

$$\begin{aligned} \begin{aligned} \left\| {\textbf{z}}(s) \right\| + \left\| \dot{\textbf{z}}(s) \right\| \le \left( 1 + C_0 \right) \left\| {\textbf{z}}(s) \right\| \le \left( 1 + C_0 \right) C_1 \left\| {\textbf{z}}(0) \right\| e^{-\beta _1 s} =: C_2\, e^{-\beta _1 s}. \end{aligned}\nonumber \\\end{aligned}$$
(F.20)

Together with the hypothesis \(\varvec{\phi }\in {\mathscr {X}}\), we derive

$$\begin{aligned} \left\| T\varvec{\phi }(t) \right\| \le \frac{e^{-at}}{2} + k(\varepsilon ) \int _t^\infty \left[ (1+a) e^{-as} + C_2 e^{-\beta _1 s} \right] \textrm{d}s \le e^{-at} \end{aligned}$$

for \(\varepsilon \) sufficiently small such that \(\beta _1 = b - ck(\varepsilon ) \left( 1 + C_0 \right) \ge b/2 = a\) and \(k(\varepsilon )\le \min \{(1+a)^{-1}, C_2^{-1}\}/ 2\). To estimate \(\frac{d}{dt} T\varvec{\phi }\), we compute

$$\begin{aligned} \begin{aligned} \left\| \frac{d}{dt} T\varvec{\phi }(t) \right\|&= \left\| A_2\varvec{\phi }(t) + V(\varvec{\phi }(t),\textbf{z}(t),\dot{\varvec{\phi }}(t),\dot{\textbf{z}}(t)) \right\| \\&\le \frac{a}{2} \left\| \varvec{\phi }(t) \right\| + k(\varepsilon ) \left( \left\| \varvec{\phi }(t) \right\| + \left\| {\textbf{z}}(t) \right\| + \left\| \dot{\varvec{\phi }}(t) \right\| + \left\| \dot{\textbf{z}}(t) \right\| \right) \\&\le \left( \frac{a}{2} + k(\varepsilon ) (1+a) \right) e^{-at} + k(\varepsilon ) C_2 e^{-\beta _1t} \le ae^{-at} \end{aligned} \end{aligned}$$

for \(\varepsilon \) sufficiently small. This proves T maps \({\mathscr {X}}\) into \({\mathscr {X}}\).

We now show that T is a contraction on \({\mathscr {X}}\). Let \(\varvec{\phi }_1, \varvec{\phi }_2\in {\mathscr {X}}\) and let \({\textbf{z}}_1, {\textbf{z}}_2\) be the corresponding solutions of (F.14a) with \({\textbf{z}}_i(0) = {\textbf{z}}_0\), \(i=1,2\). We first estimate \({\textbf{v}}(t) = {\textbf{z}}_1(t) - {\textbf{z}}_2(t)\). From (F.14a) and (F.17),

$$\begin{aligned} \begin{aligned} \left\| {\textbf{v}}(t) \right\|&\le ck(\varepsilon ) \int _0^t e^{-b(t-s)} \left[ \left\| (\textbf{z}_1(s), \dot{\textbf{z}}_1(s)) \right\| \left( \left\| \varvec{\phi }_1(s) - \varvec{\phi }_2(s) \right\| + \left\| \dot{\varvec{\phi }}_1(s) - \dot{\varvec{\phi }}_2(s) \right\| \right) + \left\| {\textbf{v}}(s) \right\| + \left\| \dot{\textbf{v}}(s) \right\| \right] \textrm{d}s. \end{aligned}\nonumber \\\end{aligned}$$
(F.21)

Since

$$\begin{aligned} \dot{\textbf{v}} = {\mathcal {B}}{\textbf{v}} + {\mathcal {R}}(\varvec{\phi }_1 + {\textbf{u}},\textbf{z}_1,\dot{\varvec{\phi }}_1 + \dot{\textbf{u}},\dot{\textbf{z}}_1) - {\mathcal {R}}(\varvec{\phi }_2 + {\textbf{u}},\textbf{z}_2,\dot{\varvec{\phi }}_2 + \dot{\textbf{u}},\dot{\textbf{z}}_2), \end{aligned}$$

using (F.17), we get

$$\begin{aligned} \left\| \dot{\textbf{v}} \right\| \le \left\| {\mathcal {B}} \right\| \left\| v \right\| + k(\varepsilon ) \left[ \left\| ({\textbf{z}}_1, \dot{\textbf{z}}_1) \right\| \left( \left\| \varvec{\phi }_1 - \varvec{\phi }_2 \right\| + \left\| \dot{\varvec{\phi }}_1 - \dot{\varvec{\phi }}_2 \right\| \right) + \left\| {\textbf{v}} \right\| + \left\| \dot{\textbf{v}} \right\| \right] . \end{aligned}$$

So

$$\begin{aligned} \left\| \dot{\textbf{v}} \right\| \le (1 - k(\varepsilon ))^{-1} \left[ (\left\| {\mathcal {B}} \right\| + k(\varepsilon )) \left\| \textbf{v} \right\| + k(\varepsilon ) \left\| ({\textbf{z}}_1, \dot{\textbf{z}}_1) \right\| \left( \left\| \varvec{\phi }_1 - \varvec{\phi }_2 \right\| + \left\| \dot{\varvec{\phi }}_1 - \dot{\varvec{\phi }}_2 \right\| \right) \right] , \end{aligned}$$

implying

$$\begin{aligned} \left\| \dot{\textbf{v}}(s) \right\| \le C_3 \left\| {\textbf{v}}(s) \right\| + k_1(\varepsilon ) \left\| ({\textbf{z}}_1, \dot{\textbf{z}}_1) \right\| \left\| \varvec{\phi }_ 1 - \varvec{\phi }_2 \right\| _{{\mathscr {X}}} e^{-as},\qquad k_1(0) = 0, \end{aligned}$$

for some \(C_3>0\). Together with \(\left\| ({\textbf{z}}_1(s), \dot{\textbf{z}}_1(s)) \right\| \le C_2 e^{-\beta _1s}\), which is followed by (F.20), (F.21) implies

$$\begin{aligned} \begin{aligned} \left\| {\textbf{v}}(t) \right\|&\le ck(\varepsilon ) \int _0^t e^{-b(t-s)} \left[ \left( 1 + k_1(\varepsilon ) \right) C_2 e^{-\beta _1s} \left\| \varvec{\phi }_ 1 - \varvec{\phi }_2 \right\| _{{\mathscr {X}}} e^{-as} + (1+C_3) \left\| \textbf{v}(s) \right\| \right] ds\\&\le C_4 k(\varepsilon ) \left\| \varvec{\phi }_ 1 - \varvec{\phi }_2 \right\| _{{\mathscr {X}}} e^{-bs} + c k(\varepsilon )(1+C_3) \int _0^t e^{-b(t-s)} \left\| {\textbf{v}}(s) \right\| \textrm{d}s \end{aligned} \end{aligned}$$

for \(\varepsilon \) sufficiently small, where \(C_4>0\) is a constant. By Gronwall’s lemma,

$$\begin{aligned} \begin{aligned} \left\| {\textbf{v}}(t) \right\| \le C_4 k(\varepsilon ) \left\| \varvec{\phi }_ 1 - \varvec{\phi }_2 \right\| _{{\mathscr {X}}} e^{-\beta _2 t},\qquad \beta _2 = b - c k(\varepsilon )(1+C_3). \end{aligned}\end{aligned}$$
(F.22)

Using (F.15) and (F.22),

$$\begin{aligned} \begin{aligned}&\left\| T\varvec{\phi }_1(t) - T\varvec{\phi }_2(t) \right\| \\&\ \ \le \int _t^\infty \left[ \frac{a}{2} \left\| \varvec{\phi }_1(s) - \varvec{\phi }_2(s) \right\| + \left\| {\mathcal {R}}(\varvec{\phi }_1(s),{\textbf{z}}_1(s), \dot{\varvec{\phi }}_1(s), \dot{\textbf{z}}_1(s)) - {\mathcal {R}}(\varvec{\phi }_2(s),{\textbf{z}}_2(s), \dot{\varvec{\phi }}_2(s), \dot{\textbf{z}}_2(s) ) \right\| \right] \\&\ \ \le \frac{a}{2} \int _t^\infty \left\| \varvec{\phi }_ 1 - \varvec{\phi }_2 \right\| _{{\mathscr {X}}} e^{-as}\, \textrm{d}s\\&\ \ \quad + k(\varepsilon ) \int _t^\infty \left[ \left\| ({\textbf{z}}_1(s), \dot{\textbf{z}}_1(s)) \right\| \left( \left\| \varvec{\phi }_1(s) - \varvec{\phi }_2(s) \right\| + \left\| \dot{\varvec{\phi }}_1(s) - \dot{\varvec{\phi }}_2(s) \right\| \right) + \left\| {\textbf{v}}(s) \right\| + \left\| \dot{\textbf{v}}(s) \right\| \right] \textrm{d}s\\&\ \ \le \frac{1}{2} \left\| \varvec{\phi }_ 1 - \varvec{\phi }_2 \right\| _{{\mathscr {X}}}\\&\ \ \quad + k(\varepsilon ) \int _t^\infty \left[ (1+k_1(\varepsilon ))C_2 e^{-\beta _1s} \left\| \varvec{\phi }_ 1 - \varvec{\phi }_2 \right\| _{{\mathscr {X}}} e^{-as} + (1+C_3)C_4 k(\varepsilon ) \left\| \varvec{\phi }_ 1 - \varvec{\phi }_2 \right\| _{{\mathscr {X}}} e^{-\beta _2 s} \right] \textrm{d}s\\&\ \ \le \alpha \left\| \varvec{\phi }_ 1 - \varvec{\phi }_2 \right\| _{{\mathscr {X}}},\qquad \alpha <1, \end{aligned} \end{aligned}$$

for \(\alpha \) sufficiently small. This shows that T is a contraction and, hence, has a unique fixed point.

Note that \(T = T_{\textbf{z}}\) and the above analysis proves that \(T_\textbf{z}\) has a unique fixed point in \({\mathscr {X}}\) provided \({\textbf{z}}\) and \(\dot{\textbf{z}}\) are sufficiently small. Denote \(\varvec{\phi }\in {\mathscr {X}}\) the fixed point of the contraction \(T_{\textbf{z}}\), where \({\textbf{z}} = {\textbf{y}} - h({\textbf{x}})\). Let \({\textbf{u}}(0) = {\textbf{x}}(0) - \varvec{\phi }(0)\). Then let \(\textbf{u}(t)\) be the solution of (F.6) evolving from \(\textbf{u}(0)\). Then \(({\textbf{x}}(t), {\textbf{y}}(t))\) can be decomposed as in \({\textbf{x}}(t) = {\textbf{u}}(t) + \varvec{\phi }(t)\) and \({\textbf{y}} = h({\textbf{x}}(t)) + {\textbf{z}}(t)\). In view of the fact that \(\varvec{\phi }\in {\mathscr {X}}\) and Lemma F.1, the asymptotic limits in (F.13) follow, which completing the proof of Proposition F.2. \(\quad \square \)

Appendix G: Asymptotic Expansion of the Local Center Manifold

In this appendix, we check the expression of local center manifold in Lemma 9.6 by asymptotic expansion.

If we substitute \({\textbf{y}} = h({\textbf{x}})\) into (9.32b), we see that the center manifold can be obtained by solving

$$\begin{aligned} h'({\textbf{x}}) \left[ Q_1 {\mathcal {N}}({\textbf{x}} + h({\textbf{x}}), \dot{\textbf{x}} + h'({\textbf{x}})\dot{\textbf{x}}) \right] = {\mathcal {L}} h({\textbf{x}}) + Q_2 {\mathcal {N}}({\textbf{x}} + h({\mathbf{x)}}, \dot{\textbf{x}} + h'(\textbf{x})\dot{\textbf{x}}). \end{aligned}$$

We want to show the manifold of equilibria \({\mathcal {M}}_*\) is a center manifold by showing it satisfies the above equation. On the manifold of equilibria \({\mathcal {M}}_*\), \(\dot{\textbf{x}} = 0\). So we need to check

$$\begin{aligned} \begin{aligned} h'({\textbf{x}}) \left[ Q_1 {\mathcal {N}}({\textbf{x}} + h({\textbf{x}}), {\textbf{0}}) \right] = {\mathcal {L}} h({\textbf{x}}) + Q_2 {\mathcal {N}}({\textbf{x}} + h({\mathbf{x)}}, {\textbf{0}}). \end{aligned}\end{aligned}$$
(G.1)

Note that

$$\begin{aligned} {\mathcal {N}}({\textbf{x}} + h({\textbf{x}}), {\textbf{0}})= & {} {\mathcal {N}}^0(\textbf{x} + h({\textbf{x}})) = \begin{bmatrix} 0\\ 0\\ -\frac{{\mathscr {R}}_gT_\infty }{\rho _lR_*}\, H^0({\textbf{x}} + h({\textbf{x}}))\\ 0\\ 0\\ \vdots \end{bmatrix}\\= & {} \begin{bmatrix} 0\\ 0\\ -\frac{2\sigma }{\rho _lR_*^3} \frac{(\alpha +R_{**}(\alpha )-R_*)^2}{\alpha +R_{**}(\alpha )}\\ 0\\ 0\\ \vdots \end{bmatrix}. \end{aligned}$$

So \(Q_1{\mathcal {N}}({\textbf{x}} + h({\textbf{x}}), {\textbf{0}}) = {\textbf{0}}\) and \({\mathcal {N}}({\textbf{x}} + h({\textbf{x}}), {\textbf{0}}) = {\mathcal {N}}(\textbf{x} + h({\textbf{x}}), {\textbf{0}})\). It suffices to check \({\mathcal {L}} h({\textbf{x}}) + Q_2 {\mathcal {N}}({\textbf{x}} + h({\mathbf{x)}}, {\textbf{0}}) = \textbf{0}\) in (G.1). Taylor expand \(h({\textbf{x}}) = h(\alpha {\textbf{b}})\) at \(\alpha =0\), one gets

$$\begin{aligned} h({\textbf{x}}) = \begin{bmatrix} \rho _{**}(\alpha ) - \rho _*\\ R_{**}(\alpha ) - R_*\\ 0\\ 0\\ 0\\ \vdots \end{bmatrix} = \begin{bmatrix} \rho _{**,1} \alpha + \rho _{**,2} \alpha ^2 + \rho _{**,3} \alpha ^3 + \cdots \\ R_{**,1} \alpha + R_{**,2} \alpha ^2 + R_{**,3} \alpha ^3 + \cdots \\ 0\\ 0\\ 0\\ \vdots \end{bmatrix}, \end{aligned}$$

where \(\rho _{**,m} = (m!)^{-1} \rho _{**}^{(m)}(0)\) and \(R_{**,m} = (m!)^{-1} R_{**}^{(m)}(0)\) in which \(\rho _{**}^{(m)}(0)\) and \(R_{**}^{(m)}(0)\) are the mth derivatives of \(\rho _{**}\) and \(R_{**}\) at \(\alpha =0\), respectively. Then \({\mathcal {L}} h({\textbf{x}}) + {\mathcal {N}}({\textbf{x}} + h({\mathbf{x)}}, {\textbf{0}}) = {\textbf{0}}\) is equivalent to

$$\begin{aligned} \begin{aligned} \sum _{m=1}^\infty \left( \frac{{\mathscr {R}}_gT_\infty }{\rho _lR_*} \rho _{**,m} + \frac{2\sigma }{\rho _lR_*^3} R_{**,m} \right) \alpha ^m - \frac{2\sigma }{\rho _lR_*^3} \frac{\left( \alpha +R_{**}(\alpha )-R_* \right) ^2}{\alpha +R_{**}(\alpha )} = 0. \end{aligned}\nonumber \\\end{aligned}$$
(G.2)

For \(m=1\), we differentiate (9.34) with respect to \(\alpha \), evaluate at \(\alpha =0\) and use \(R_{**}(0) = R_*\), we get \({\mathscr {R}}_gT_\infty \rho _{**}'(0) = -(2\sigma /R_*^2) R_{**}'(0)\). So we have

$$\begin{aligned} \frac{{\mathscr {R}}_gT_\infty }{\rho _lR_*} \rho _{**,1} + \frac{2\sigma }{\rho _lR_*^3} R_{**,1} = 0. \end{aligned}$$

For \(m=2\), we first expand

$$\begin{aligned} \frac{\left( \alpha +R_{**}(\alpha )-R_* \right) ^2}{\alpha +R_{**}(\alpha )} = \frac{( 1+R_{**}'(0) )^2}{R_*} \alpha ^2 + \cdots . \end{aligned}$$

Then the coefficient of \(\alpha ^2\) in (G.2) is

$$\begin{aligned} \begin{aligned} \frac{{\mathscr {R}}_gT_\infty }{\rho _lR_*} \rho _{**,2} + \frac{2\sigma }{\rho _lR_*^3} R_{**,2} - \frac{2\sigma }{\rho _lR_*^4} \left( 1+R_{**,1} \right) ^2 . \end{aligned}\end{aligned}$$
(G.3)

By differentiating (9.34) with respect to \(\alpha \) twice, evaluating at \(\alpha =0\) and using \(R_{**}(0) = R_*\), we have \({\mathscr {R}}_gT_\infty \rho _{**}''(0) = (4\sigma /R_*^3) (R_{**}'(0))^2 - (2\sigma /R_*^2) R_{**}''(0)\). So

$$\begin{aligned} \frac{{\mathscr {R}}_gT_\infty }{\rho _lR_*} \rho _{**,2} + \frac{2\sigma }{\rho _lR_*^3} R_{**,2} = (2!)^{-1} \frac{4\sigma }{\rho _l R_*^4} (R_{**,1})^2. \end{aligned}$$

In order to have the term (G.3) for \(m=2\) vanishing, we require

$$\begin{aligned} 0 = \frac{2\sigma }{\rho _lR_*^4} R_{**,1}^2 - \frac{2\sigma }{\rho _lR_*^4} (1 + R_{**,1})^2, \end{aligned}$$

whose solution is \(R_{**,1} = -1/2\).

For \(m=3\), we further expand, using \(R_{**,1} = -1/2\), that

$$\begin{aligned} \frac{\left( \alpha +R_{**}(\alpha )-R_* \right) ^2}{\alpha +R_{**}(\alpha )} = \frac{( 1+R_{**,1} )^2}{R_*} \alpha ^2 + \left( \frac{1}{2R_*} R_{**}''(0) - \frac{1}{8R_*^2} \right) \alpha ^3 + \cdots \end{aligned}$$

Then the coefficient of \(\alpha ^3\) in (G.2) is

$$\begin{aligned} \begin{aligned} \frac{{\mathscr {R}}_gT_\infty }{\rho _lR_*} \rho _{**,3} + \frac{2\sigma }{\rho _lR_*^3} R_{**,3} - \frac{2\sigma }{\rho _lR_*^3} \left( \frac{1}{2R_*} R_{**}''(0) - \frac{1}{8R_*^2} \right) . \end{aligned}\end{aligned}$$
(G.4)

By differentiating (9.34) with respect to \(\alpha \) three times, evaluating at \(\alpha =0\) and using \(R_{**}(0) = R_*\), we get \({\mathscr {R}}_gT_\infty \rho _{**}'''(0) = -(12\sigma /R_*^4)(R_{**}'(0))^3 + (12\sigma /R_*^3)R_{**}'(0)R_{**}''(0) -(2\sigma /R_*^2) R_{**}'''(0)\). So we have

$$\begin{aligned} \frac{{\mathscr {R}}_gT_\infty }{\rho _lR_*} \rho _{**,3} = -\frac{2\sigma }{\rho _lR_*^5} R_{**,1}^3 + \frac{4\sigma }{\rho _lR_*^4} R_{**,1} R_{**,2} - \frac{2\sigma }{\rho _lR_*^3} R_{**,3}. \end{aligned}$$

In order to have the term (G.4) for \(m=3\) vanishing, we require

$$\begin{aligned} 0= & {} -\frac{2\sigma }{\rho _lR_*^5} R_{**,1}^3 + \frac{4\sigma }{\rho _lR_*^4} R_{**,1} R_{**,2} - \frac{2\sigma }{\rho _lR_*^3} R_{**,3} + \frac{2\sigma }{\rho _lR_*^3} R_{**,3} \\{} & {} - \frac{\sigma }{\rho _lR_*^3} \left( \frac{1}{R_*} R_{**}''(0) - \frac{1}{4R_*^2} \right) , \end{aligned}$$

where the third term cancels the fourth. Using \(R_{**,1}=-1/2\) and \(R_{**}''(0) = 2R_{**,2}\),

$$\begin{aligned} 0 = \frac{\sigma }{4\rho _lR_*^5} - \frac{2\sigma }{\rho _lR_*^4} R_{**,2} - \frac{2\sigma }{\rho _lR_*^4} R_{**,2} + \frac{\sigma }{4\rho _lR_*^5} \end{aligned}$$

for which the solution is \(R_{**,2} = 1/(8R_*)\). Thus,

$$\begin{aligned} R_{**}(\alpha ) = R_* - \frac{1}{2} \alpha + \frac{1}{8R_*} \alpha ^2 + \cdots = \frac{-\alpha + \sqrt{\alpha ^2 + 4R_{*}^2}}{2}, \end{aligned}$$

which coincides with the center manifold expression in (9.34).

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Lai, CC., Weinstein, M.I. Free Boundary Problem for a Gas Bubble in a Liquid, and Exponential Stability of the Manifold of Spherically Symmetric Equilibria. Arch Rational Mech Anal 247, 100 (2023). https://doi.org/10.1007/s00205-023-01927-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00205-023-01927-z

Navigation