Skip to main content
Log in

On short-wave instability of the stratified Kolmogorov flow

  • Original Article
  • Published:
Theoretical and Computational Fluid Dynamics Aims and scope Submit manuscript

Abstract

The problem of the linear stability of the stratified Kolmogorov flow driven by a sinusoidal in space force in a viscous and diffusive Boussinesq fluid is re-visited using the Floquet theory, Galerkin approximations and the method of (generalized) continued fractions. Numerical and analytical arguments are provided in favor of a conjecture that an ideal stratified Kolmogorov flow is prone to short-wave instability for Richardson numbers markedly greater than the critical Richardson number Ri \(=\) ¼ that appears in the Miles–Howard theorem. The short-wave instability of the stratified Kolmogorov flow is conjectured to be due to a resonance amplification of the Doppler-shifted internal gravity wave modes, in the presence of critical levels of the main flow that are ignored in the proof of the Miles–Howard theorem, but it is emphasized that the complete resolution of the above paradox is a task for future research.

Graphical abstract

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

Access this article

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

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6

Similar content being viewed by others

Notes

  1. Miles [9] proved this theorem for a special class of monotonic velocity profiles, while Howard [10] provided the general proof which is reproduced in “Appendix A” for a particular case of the stratified Kolmogorov flow.

  2. Balmforth and Young [7] noted that solutions (8a) and (8b) are not regular functions and require interpretation when \(\cos z<0\). It was conjectured (ibid) that \(\psi _{1} \) is a limiting neutral mode of a branch of unstable modes and it should be taken as \(\psi _{1} =C_{1} \left| {\cos z} \right| ^{\sqrt{1-k^{2}} }\).

  3. For the conditions in the Earth’s middle atmosphere with \(N=2\times 10^{-2}\hbox {s}^{-1}\), \(\Lambda _{z} ={2\uppi }/q_{0}=3\times 10^{3}\hbox {m}\) this corresponds to the turbulent viscosity and thermal diffusivity \(\sim 1.1\hbox {m}^{2}\hbox {s}^{-1}\).

  4. An account of double, triple products, etc. of \(\gamma _{i} \) greatly complicates the formal analysis, but leads to sums that assume finite values and are therefore small compared to the logarithmically divergent sums that appear in the linear approximation (see below).

References

  1. Fritts, D., Alexander, M.: Gravity wave dynamics and effects in the middle atmosphere. Rev. Geophys. 41, 1003 (2003). https://doi.org/10.1029/2001RG000106

    Article  Google Scholar 

  2. Sutherland, B.R.: Internal Gravity Waves. Cambridge University Press, Cambridge (2014)

    MATH  Google Scholar 

  3. Andrews, D.G., Holton, J.R., Leovy, C.B.: Middle Atmosphere Dynamics. Academic Press, San Diego (1987)

    Google Scholar 

  4. Caillol, P.: Nonlinear internal waves in the upper atmosphere. Geophys. Astrophys. Fluid Dyn. 99, 271–308 (2005). https://doi.org/10.1080/03091920500213585

    Article  MathSciNet  MATH  Google Scholar 

  5. Kurgansky, M.V.: Instability of internal gravity waves propagating at small angles to the vertical. Izv. Atmos. Ocean. Phys. 16, 758–764 (1980)

    Google Scholar 

  6. Lombard, P.N., Riley, J.J.: Instability and breakdown of internal gravity waves. I. Linear stability analysis. Phys. Fluids 8, 3271–3287 (1996). https://doi.org/10.1063/1.869117

    Article  MATH  Google Scholar 

  7. Balmforth, N.J., Young, Y.-N.: Stratified Kolmogorov flow. J. Fluid Mech. 450, 131–167 (2002). https://doi.org/10.1017/S0022111002006371

    Article  MathSciNet  MATH  Google Scholar 

  8. Balmforth, N.J., Young, Y.-N.: Stratified Kolmogorov flow. Part. 2. J. Fluid Mech. 528, 23–42 (2005). https://doi.org/10.1017/S0022111002006371

    Article  MathSciNet  MATH  Google Scholar 

  9. Miles, J.W.: On the stability of heterogeneous shear flows. J. Fluid Mech. 10, 496–508 (1961). https://doi.org/10.1017/S0022112061000305

    Article  MathSciNet  MATH  Google Scholar 

  10. Howard, L.N.: Note on a paper of John W Miles. J. Fluid Mech. 10, 509–512 (1961). https://doi.org/10.1017/S0022112061000317

    Article  MathSciNet  MATH  Google Scholar 

  11. Drazin, P.G., Reid, W.H.: Hydrodynamic Stability, 2nd edn. Cambridge University Press, Cambridge (2004)

    Book  Google Scholar 

  12. Kurgansky, M.V.: On the instability of internal gravity waves propagating at small but finite angles to the vertical. Geophys. Astrophys. Fluid Dyn. 112, 265–276 (2018). https://doi.org/10.1080/03091929.2018.1510920

    Article  MathSciNet  MATH  Google Scholar 

  13. Meshalkin, L.D., Sinai, Y.G.: Investigation of the stability of a stationary of equations for the plane movement of an incompressible viscous fluid. J. Appl. Math. Mech. (Prikl. Mat. Mech.) 25, 1700–1705 (1961). https://doi.org/10.1016/0021-8928(62)90149-1

    Article  MATH  Google Scholar 

  14. Obukhov, A.M.: Laboratory modeling of Kolmogorov flow. Russ. Math. Surv. (Usp. Mat. Nauk) 38, 101–111 (1983)

    Article  Google Scholar 

  15. Gledzer, E.B., Dolzhansky, F.V., Obukhov, A.M.: Systems of Hydrodynamic Type and their Applications. Nauka, Moscow (1981). (in Russian)

  16. Kalashnik, M., Kurgansky, M.: Nonlinear dynamics of long-wave perturbations of the Kolmogorov flow for large Reynolds numbers. Ocean Dyn. 68, 1001–1012 (2018). https://doi.org/10.1007/s10236-018-1180-6

    Article  Google Scholar 

  17. Arnol’d, V.I.: Ordinary Differential Equations. MIT (1973) (translated from Russian)

  18. Arnol’d, V.I.: Conditions for nonlinear stability of stationary plane curvilinear flows of an ideal fluid. Dokl. Akad. Nauk SSSR 162, 975–978 (1965) (English translation: Sov. Math. 6, 773–777 (1965))

  19. Abarbanel, H.D.I., Holm, D.D., Marsden, J.E., Ratiu, T.S.: Nonlinear stability analysis of stratified fluid equilibria. Philos. Trans. R. Soc. Lond. A318, 349–409 (1986). https://doi.org/10.1098/rsta.1986.0078

    Article  MathSciNet  MATH  Google Scholar 

  20. Libin, A., Sivashinsky, G., Levich, E.: Long-wave instability of periodic flows at large Reynolds numbers. Phys. Fluids 30, 2984–2986 (1987). https://doi.org/10.1063/1.866076

    Article  MATH  Google Scholar 

  21. Kurgansky, M.V.: Inertial instability of the Kolmogorov flow in a rotating stratified fluid. Fluid Dyn. Res. 53, 035502 (2021). https://doi.org/10.1088/1873-7005/abfaf0

    Article  MathSciNet  Google Scholar 

  22. Howard, L.N., Maslowe, S.A.: Stability of stratified shear flows. Bound. Layer Meteorol. 4, 511–523 (1973). https://doi.org/10.1007/BF02265252

    Article  Google Scholar 

  23. Taylor, G.I.: Effect of variation in density on the stability of superposed streams of fluid. Proc. R. Soc. Lond. A 132, 499–523 (1931). https://doi.org/10.1098/rspa.1931.0115

    Article  MATH  Google Scholar 

  24. Batchaev, A.M., Dovzhenko, V.A., Kurgansky, M.V.: Simulation of shear flow in a stratified fluid. Izv. Atmos. Ocean. Phys. 20, 439–442 (1984)

    Google Scholar 

  25. Harlander, U., Kurgansky, M.V.: Two-dimensional internal gravity wave beam instability. Linear theory and subcritical instability. Geophys. Astrophys. Fluid Dyn. 115(5–6), 612–647 (2021). https://doi.org/10.1080/03091929.2021.1943379

    Article  MathSciNet  MATH  Google Scholar 

  26. Morrison, P.J., Kotschenreuther, M.: The free energy principle, negative energy modes, and stability. In: Bar’yakhtar, V.G., Chernousenko, V.M., Erokhin, N.S., Sitenko, A.B., Zakharov, V.E. (eds.) Nonlinear World: IV International Workshop on Nonlinear and Turbulent Processes in Physics, vol. 2, pp. 910–932. World Scientific, Singapore (1990)

    MATH  Google Scholar 

  27. Le Gal, P., Harlander, U., Borcia, I.D., Le Dizès, S., Chen, J., Favier, B.: Instability of vertically stratified horizontal plane Poiseuille flow. J. Fluid Mech. 907, R1 (2021). https://doi.org/10.1017/jfm.2020.917

    Article  Google Scholar 

  28. Kamke, E.: Differentialgleichungen: Losungsmethoden und Losungen, I, Gewöhnliche Differentialgleichungen. B. G. Teubner, Leipzig (1977)

    MATH  Google Scholar 

  29. Morrison, P.J.: Hamiltonian description of the ideal fluid. Rev. Mod. Phys. 70(2), 467–521 (1998). https://doi.org/10.1103/RevModPhys.70.467

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

I am sincerely grateful to Uwe Harlander for his valuable comments on this work. The author thanks two anonymous reviewers and the Handling Editor whose critical comments and helpful suggestions greatly contributed to improving the content and style of this article.

Funding

This work was supported by the Russian Science Foundation (Grant No. 22-27-00039).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to M. V. Kurgansky.

Ethics declarations

Author contributions

The work reported in this paper has been done solely by the author.

Data availability

Not applicable.

Code availability

Not applicable.

Conflict of interest

The author declares that he has no known competing financial interests or personal relationships that could have influenced the work reported in this paper.

Ethics approval

Not applicable.

Consent to participate

Not applicable.

Consent for publication

Not applicable.

Additional information

Communicated by William Dewar.

Publisher's Note

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

Appendices

Appendix A: comments on the Miles–Howard theorem application

When the solution of (5) is sought in the form \(\left( {\psi ,\xi } \right) =\left( {\hat{{\psi }}\left( z \right) ,\hat{{\xi }}\left( z \right) } \right) \exp \left\{ {\text{ i }k\left( {x-ct} \right) } \right\} \), then Eqs. (5a, 5b) read

$$\begin{aligned} \left( {\varepsilon \cos z+c} \right) \left( {\frac{\text{ d}^{2}\hat{{\psi }}}{\text{ d }z^{2}}-k^{2}\hat{{\psi }}} \right) +\varepsilon \cos z{}{}\hat{{\psi }}=-\hat{{\xi }},\quad \left( {\varepsilon \cos z+c} \right) \hat{{\xi }}{}=\hat{{\psi }}, \end{aligned}$$
(A.1)

By eliminating \(\hat{{\xi }}\) between (A.1) it follows that

$$\begin{aligned} \left( {\varepsilon \cos z+c} \right) \left( {\frac{\text{ d}^{2}\hat{{\psi }}}{\text{ d }z^{2}}-k^{2}\hat{{\psi }}} \right) +\left( {\varepsilon \cos z+\frac{1}{\varepsilon \cos z+c}} \right) \hat{{\psi }}=0. \end{aligned}$$
(A.2)

Equation (A.2) is the Taylor–Goldstein equation. We denote \(W=\varepsilon \cos z+c\) and by dividing both sides of (A.2) by W rewrite (A.2) as

$$\begin{aligned} \frac{\text{ d}^{2}\hat{{\psi }}}{\text{ d }z^{2}}-k^{2}\hat{{\psi }}-\frac{{W}''}{W}\hat{{\psi }}+\frac{1}{W^{2}}\hat{{\psi }}=0. \end{aligned}$$
(A.3)

Primes denote differentiation with respect to z. We note that passing from (A.1) to (A.3), by dividing by W, we disregard the continuous spectrum of eigenvalues, which correspond to the critical levels \(\varepsilon \cos z+c=0\) in the problem (see more below). Therefore, the subsequent stability analysis due to [10] may be incomplete in this regard.

Following [10], assume that c has a nonzero imaginary part \(c_{i} \) and introduce a new dependent variable \(G={\hat{{\psi }}} / {\sqrt{W} }\). After some algebra, we obtain

$$\begin{aligned} \left( {W{G}'} \right) ^{\prime }-\left[ {\textstyle {1 \over 2}{W}''+k^{2}W+W^{-1}\left( {\textstyle {1 \over 4}{W}'^{2}-1} \right) } \right] G=0. \end{aligned}$$
(A.4)

Then, multiplying both sides of (A.4) by \(\bar{{G}}\), where the overbar denotes the complex conjugate, integrating over z, integrating \(\left( {W{G}'} \right) ^{\prime }\) by parts and invoking the periodic boundary conditions, we obtain

$$\begin{aligned} \int {\left\{ {W\left| {{G}'} \right| ^{2}+k^{2}W\left| G \right| ^{2}+\textstyle {1 \over 2}{W}''\left| G \right| ^{2}+W^{-1}\left( {\textstyle {1 \over 4}{W}'^{2}-1} \right) \left| G \right| ^{2}} \right\} } {}\text{ d }z=0. \end{aligned}$$
(A.5)

Taking the imaginary part of (A.5), we have

$$\begin{aligned} c_{i} \int {\left\{ {\left| {{G}'} \right| ^{2}+k^{2}\left| G \right| ^{2}+\left| W \right| ^{-2}\left( {1-\textstyle {1 \over 4}{W}'^{2}} \right) \left| G \right| ^{2}} \right\} } {}\text{ d }z=0. \end{aligned}$$

Because \(c_{i} \ne 0\), this equality is possible only when \(\textstyle {1 \over 4}{W}'^{2}>1\). Recalling that \({W}'^{2}=\varepsilon ^{2}\sin ^{2}z\), it shows that this necessary condition for instability is fulfilled when \(\varepsilon >2\), that is, \(\text{ Ri }<\textstyle {1 \over 4}\) [10].

Alternatively, by eliminating \(\hat{{\psi }}\) between (A.1), we obtain, see also [9]

$$\begin{aligned} \frac{\text{ d }}{\text{ d }z}\left( {W^{2}\frac{\text{ d }\hat{{\xi }}}{\text{ d }z}} \right) -k^{2}W^{2}\hat{{\xi }}=-\hat{{\xi }}. \end{aligned}$$
(A.6)

Then, multiplying (A.6) by \(\bar{{\hat{{\xi }}}}\) and integrating over z with periodic boundary conditions, we have

$$\begin{aligned} \int {W^{2}\left( {\left| {{\hat{{\xi }}}'} \right| ^{2}+k^{2}\left| {\hat{{\xi }}} \right| ^{2}} \right) {}{}} {}\text{ d }z=\int {\left| {\hat{{\xi }}} \right| ^{2}} \text{ d }z. \end{aligned}$$
(A.7)

Taking the imaginary part of (A.7), we obtain (cf. [9])

$$\begin{aligned} c_{i} \int {\left( {\varepsilon \cos z+c_{r} } \right) \left( {\left| {{\hat{{\xi }}}'} \right| ^{2}+k^{2}\left| {\hat{{\xi }}} \right| ^{2}} \right) {}{}} {}\text{ d }z=0. \end{aligned}$$
(A.8)

If \(c_{i} \ne 0\) then it follows from (A.8) that \(c_{r} \in \left[ {-\varepsilon ,\,\varepsilon } \right] \). Due to symmetry with respect to mirror reflections about the z-axis, we have

$$\begin{aligned} \int {\left( {\varepsilon \cos z+c_{r} } \right) \left( {\left| {{\hat{{\xi }}}'} \right| ^{2}+k^{2}\left| {\hat{{\xi }}} \right| ^{2}} \right) {}{}} {}\text{ d }z=\int {\left( {\varepsilon \cos z-c_{r} } \right) \left( {\left| {{\hat{{\xi }}}'} \right| ^{2}+k^{2}\left| {\hat{{\xi }}} \right| ^{2}} \right) {}{}} {}\text{ d }z, \end{aligned}$$
(A.9)

and (A.9) shows that \(c_{r} =0\), that is, the principle of exchange of stabilities is in operation. Now, the equation \(\varepsilon \cos z+c_{r} =0\), which generally determines the continuum spectrum of eigenvalues in the problem, shows that \(\cos z_{*} =0\) and \(z_{*} ={{\pi }} /2+n{\pi }\), \(n=0,\,\pm 1,\,\pm 2,\ldots \).Therefore, the spectrum quantization occurs and it reduces to the eigenvalue \(c_{r} =0\), corresponding to a countable infinite set of critical levels responsible for instability via the resonance amplification mechanism (see the main text). Next, taking the real part of (A.7) we have

$$\begin{aligned} \int {\left( {\varepsilon ^{2}\cos ^{2}z-c_{i}^{2} } \right) \left( {\left| {{\hat{{\xi }}}'} \right| ^{2}+k^{2}\left| {\hat{{\xi }}} \right| ^{2}} \right) {}{}} {}\text{ d }z=\int {\left| {\hat{{\xi }}} \right| ^{2}} \text{ d }z>0, \end{aligned}$$
(A.10)

that is, \(c_{i} \) cannot exceed \(\varepsilon \).This is the Howard’s [10] semicircle theorem as applied to the stratified Kolmogorov flow. This theorem does not exclude the possibility of instability at \(\varepsilon <2\).

Finally, it is worth mentioning that solution (11), (12), which supports the deductions of the Miles–Howard theorem, does not exhaust all the solutions of (6) that are relevant to the neutral stability conditions. When \(c=0\), then the substitution \(\eta =\tan z\) transforms (A.6) to

$$\begin{aligned} \left( {\eta ^{2}+1} \right) ^{2}\frac{\text{ d}^{2}\hat{{\xi }}}{\text{ d }\eta ^{2}}-k^{2}\hat{{\xi }}+\frac{1}{\varepsilon ^{2}}\left( {\eta ^{2}+1} \right) =0. \end{aligned}$$
(A.11)

For \(k^{2}=3/ 4\) and \(\varepsilon <2\) this equation has the exact solution (see, e.g., [28], eq. 2.385)

$$\begin{aligned} \hat{{\xi }}=\left( {\eta ^{2}+1} \right) ^{1 / 4}\left( {C_{1} \cos I+C_{2} \sin I} \right) ,\quad I=\textstyle {1 \over 2}\sqrt{4 / {\varepsilon ^{2}}-1} \,\,\log \left( {\eta +\sqrt{\eta ^{2}+1} } \right) , \end{aligned}$$
(A.12)

Here, \(C_{1} \) and \(C_{2} \) are arbitrary constants. For \(\varepsilon >2\) it is necessary to replace \(4 / {\varepsilon ^{2}}-1\), \(\cos \), \(\sin \) with, respectively, \(1-4 /{\varepsilon ^{2}}\), \(\cosh \), \(\sinh \). Interesting is the case of \(\varepsilon =2\), that is, \(\text{ Ri }=\textstyle {1 \over 4}\). Now the solution of (A.11) is \(\hat{{\xi }}=C_{1} \left( {\eta ^{2}+1} \right) ^{1 / 4}\), which exactly matches (11) and (12). However, in the case of \(\varepsilon <2\), that is, \(\text{ Ri }>\textstyle {1 \over 4}\) relevant to this study, we have the continuum of solutions belonging to candidate neutral stability curves, and the arguments of Sect. 6 virtually distinguish that solution which corresponds to \(\varepsilon =2\sqrt{2 /3} \), that is, \(\text{ Ri }=\textstyle {3 \over 8}>\textstyle {1 \over 4}\). There is evidence based on Maplesoft application that for \(k^{2}\ne 3 / 4\), including \(k^{2}>1\), the solutions of (A.11) for \(\varepsilon <2\) are given by a complex analytical expression containing associated Legendre functions of the first and second kind.

Appendix B: comparison with Balmforth and Young [7]

For the comparison with [7] it is convenient to change to the Richardson number \(\text{ Ri }={Y^{2}k^{2}} / 4\) and write \(\gamma _{n} =2-{4\text{ Ri }} / {\left( {n^{2}+k^{2}-1} \right) }\). Now, solvability conditions (14) and (15), the latter taken with the plus sign on its right-hand side, have a ”trivial” solution \(\text{ Ri }=\text{0 }\). Indeed, in this case the resulting infinite continued fraction

$$\begin{aligned} S_{\infty } =2-\frac{1}{2-\frac{1}{2-\cdots }} \end{aligned}$$

is summed up, \(S_{\infty } =1\), see the main text. The only “non-trivial” solution to be compared with curve (7) originating from [7] is supplied by Eq. (16) with the minus sign on its right-hand side. At very small \(k^{2}\)-values we have approximately from (16) by setting \(\text{ Ri }=\text{0 }\) in the second term on the left-hand side of (16)

$$\begin{aligned} 2-\frac{4\text{ Ri }}{k^{2}}-1\cong -1. \end{aligned}$$

This equation has a solution \(\text{ Ri }\cong {k^{2}}/ 2\) consistently with (11). To find a correction to this asymptotic solution we proceed as follows. We start from (18), make a replacement \(Y^{2}k^{2}=4\text{ Ri }\), and find that expressions (19) assume the form

$$\begin{aligned} f\left( M \right) =1-4\text{ Ri }\sum \limits _{n=2}^M {\frac{\left( {n-1} \right) }{\left( {2n+1} \right) ^{2}+k^{2}-1}} ,\quad g\left( M \right) =1-4\text{ Ri }\sum \limits _{n=1}^M {\frac{\,n}{\left( {2n+1} \right) ^{2}+k^{2}-1}} . \end{aligned}$$

It is readily seen that

$$\begin{aligned} f\left( M \right) \equiv 1-4\text{ Ri }\sum \limits _{n=1}^M {\frac{\left( {n-1} \right) }{\left( {2n+1} \right) ^{2}+k^{2}-1}} =g\left( M \right) +4\text{ Ri }\sum \limits _{n=1}^M {\frac{1}{\left( {2n+1} \right) ^{2}+k^{2}-1}} \end{aligned}$$

and therefore

$$\begin{aligned} \frac{f\left( M \right) }{g\left( M \right) }=1+4\text{ Ri }\sum \limits _{n=1}^M {\frac{1}{\left( {2n+1} \right) ^{2}+k^{2}-1}} \cdot \left( {1-4\text{ Ri }\sum \limits _{n=1}^M {\frac{\,n}{\left( {2n+1} \right) ^{2}+k^{2}-1}} } \right) ^{-1}. \end{aligned}$$
(B.1)

By assuming \(\text{ Ri }<<1\), \(k^{2}<<1\), thus neglecting the contribution from \(k^{2}\) everywhere in (B.1) and taking the limit \(M\rightarrow \infty \), such that the resulting infinite set before the parentheses is readily summed up, we obtain

$$\begin{aligned} \lim \limits _{M\rightarrow \infty } \frac{f\left( M \right) }{g\left( M \right) }\cong 1+4\text{ Ri }\cdot \frac{1}{4}\left( {1-4\text{ Ri }\cdot \lim \limits _{M\rightarrow \infty } \sum \limits _{n=1}^M {\frac{\,n}{\left( {2n+1} \right) ^{2}-1}} } \right) ^{-1}. \end{aligned}$$
(B.2)

We reached a somewhat paradoxical situation. If the Richardson number \(\text{ Ri }\) in (B.2) assumes a very small but finite value, then the second logarithmically divergent term in parentheses dominates (cf. Sect. 6) and we obtain from (B.2)

$$\begin{aligned} \lim \limits _{M\rightarrow \infty } \frac{f\left( M \right) }{g\left( M \right) }\cong 1. \end{aligned}$$
(B.3)

On the other hand, if we rigorously set \(\text{ Ri }=\text{0 }\) in parentheses in (B.2), then

$$\begin{aligned} \lim \limits _{M\rightarrow \infty } \frac{f\left( M \right) }{g\left( M \right) }\cong 1+4\text{ Ri }\cdot \frac{1}{4}. \end{aligned}$$
(B.4)

From (B.4) we would have, cf. (18) in Sect. 6,

$$\begin{aligned} 2-\frac{4\text{ Ri }}{k^{2}}-4\text{ Ri }\cdot \frac{1}{4}\cong 0, \end{aligned}$$

which in the first approximation yields \(\text{ Ri }\cong {k^{2}} / 2-{k^{4}} / 8\). This coincides exactly with the two first terms of the binominal expansion of the equation \(\text{ Ri }=\sqrt{1-k^{2}} -\left( {1-k^{2}} \right) \cong {k^{2}} / 2-{k^{4}} / 8+\cdots \) for the neutral stability curve in [7], cf. Sect. 3. This coincidence may not be accidental, but correlate with a singular character of solution (8a, 8b) in [7]. The two neutral stability curves, (i) \(\text{ Ri }=\sqrt{1-k^{2}} -\left( {1-k^{2}} \right) \) of [7] and (ii) \(\text{ Ri }={k^{2}} / 2\) obtained from (A.1) and (A.3), diverge significantly afterward. In particular, this can be seen from the fact that while for the first curve \(\text{ Ri }=0\) at \(k^{2}=1\) it is not the case for the second curve, as for example follows from (A.1), the arguments of Sect. 6 show explicitly that \(\text{ Ri }\approx \textstyle {1 \over 2}\) for \(k^{2}=1\).

At \(k^{2}\rightarrow 0\) irregular solution (12) \(\psi _{1} =C_{1} \left( {\cos z} \right) ^{\sqrt{1-k^{2}} }\) reduces to a regular, spatially periodic solution \(\psi _{1} =C_{1} \cos z\) that fits the Galerkin first approximation in [5, 12]. It can be hypothesized, although cannot be proved at this point, that \(k^{2}=0\) represents a bifurcation point in the parameter space at which two independent solutions branch off from the solution \(\psi _{1} =C_{1} \cos z\). The first solution corresponds to \(\psi _{1} =C_{1} \left( {\cos z} \right) ^{\sqrt{1-k^{2}} }\), see [7], and the second one meets our solution obtained by the method of continued fractions, while the Miles−Howard theorem is relevant only to the first of them.

Appendix C: on the Lyapunov stability of the stratified Kolmogorov flow

We will show that (14) corresponds to what follows from Arnol’d’s energy-Casimir method [18, 19]. We start from the nonlinear equations of ideal fluid motion derived from Eq. (1), taken in a dimensionless form using a transition to dimensionless variables as described in the main text

$$\begin{aligned} \frac{\partial }{\partial t}\nabla ^{2}\psi +J\left( {\psi ,\nabla ^{2}\psi } \right) =\frac{\partial \xi }{\partial x},\quad \frac{\partial }{\partial t}\xi +J\left( {\psi ,\xi } \right) =-\frac{\partial \psi }{\partial x}. \end{aligned}$$
(C.1)

The asterisks on the dimensionless variables are omitted. The energy conservation law

$$\begin{aligned} \frac{\text{ d }}{\text{ d }t}E=0,\quad E=\frac{1}{2}\int \!\!\!\int {\left\{ {\left( {\nabla \psi } \right) ^{2}+\xi ^{2}} \right\} } {}{}\text{ d }x\text{ d }z, \end{aligned}$$

and the invariance of the Casimir functional (cf. [19])

$$\begin{aligned} \frac{\text{ d }}{\text{ d }t}F=0,\quad F=\int \!\!\!\int {\left\{ {\Phi \left( {z+\xi } \right) +\Gamma \left( {x+\xi } \right) \nabla ^{2}\psi } \right\} } {}{}\text{ d }x\text{ d }z, \end{aligned}$$

with \(\Phi \) and \(\Gamma \) as arbitrary differentiable functions, follow from (C.1) given the periodic boundary conditions. We study the nonlinear Lyapunov stability of the solution \(\psi _{0} =\varepsilon \sin z\) (a non-essential additive constant is omitted), \(\xi _{0} =0\), with respect to small but finite variations \(\psi =\psi _{0} +\delta \psi \), \(\xi =\delta \xi \). We introduce the functional \(H=E+F\) and expand \(H\left[ {\psi ,\xi } \right] -H\left[ {\psi _{0} ,\xi _{0} } \right] \) into a series of variations of increasing order \(H\left[ {\psi ,\xi } \right] -H\left[ {\psi _{0} ,\xi _{0} } \right] =\delta H+\textstyle {1 \over 2}\delta ^{2}H+\cdots \). We have

$$\begin{aligned} \begin{array}{l} \delta H=\int \!\!\!\int {\left\{ {\nabla \psi _{0} \cdot \nabla \delta \psi +{\Phi }'\delta \xi +{\Gamma }'\delta \xi {}{}\nabla ^{2}\psi _{0} +\Gamma \nabla ^{2}\delta \psi } \right\} } {}{}\text{ d }x\text{ d }z \\ \quad \;\;\,=\int \!\!\!\int {\left\{ {\left( {-\psi _{0} +\Gamma } \right) \nabla ^{2}\delta \psi +\left( {{\Phi }'+{\Gamma }'\nabla ^{2}\psi _{0} } \right) \delta \xi {}{}} \right\} } {}{}\text{ d }x\text{ d }z \\ \end{array}, \end{aligned}$$

Here, the prime denotes differentiation. The conditions for vanishing of the first variation read as \(\Gamma \left( z \right) =\varepsilon \sin z\), \({\Phi }'\left( z \right) =\varepsilon ^{2}\sin z\cos z\equiv \textstyle {1 \over 2}\varepsilon ^{2}\sin 2z\). Next, we calculate the second variation

$$\begin{aligned} \delta ^{2}H=\int \!\!\!\int {\left\{ {\left( {\nabla \delta \psi } \right) ^{2}+\left( {1+{\Phi }''+{\Gamma }''\nabla ^{2}\psi _{0} } \right) \left( {\delta \xi } \right) ^{2}{}+2{\Gamma }'\nabla ^{2}\delta \psi \delta \xi {}{}} \right\} } {}{}\text{ d }x\text{ d }z. \end{aligned}$$

Because, \({\Phi }''\left( z \right) =\varepsilon ^{2}\cos 2z\equiv \varepsilon ^{2}\left( {\cos ^{2}z-\sin ^{2}z} \right) \), \({\Gamma }'=\varepsilon \cos z\) and \({\Gamma }''\nabla ^{2}\psi _{0} =\varepsilon ^{2}\sin ^{2}z\), we obtain

$$\begin{aligned} \delta ^{2}H=\int \!\!\!\int {\left\{ {\left( {\nabla \delta \psi } \right) ^{2}+\left( {1+\varepsilon ^{2}\cos ^{2}z} \right) \left( {\delta \xi } \right) ^{2}{}+2\varepsilon \cos z{}{}\nabla ^{2}\delta \psi \delta \xi } \right\} } {}{}\text{ d }x\text{ d }z, \end{aligned}$$
(C.2)

which up to “1/2” matches (14). If the integrand in (C.2) were positive definite for some \(\varepsilon \) value, then the stratified Kolmogorov flow would be stable according to Lyapunov for this \(\varepsilon \) value. However, this is impossible in principle due to the last term in (C.2) which describes the destabilizing effect of short-wave perturbations (cf. [19]). The indefiniteness of such a second variation implies either the existence of negative-energy modes (which by definition are neutrally stable, but can be unstable (a) in the presence of dissipation or (b) with the inclusion of nonlinearity) or linear instability, and there is no way in general of distinguishing from the expressions for \(\delta ^{2}H\) alone which is the case [29]. It is conjectured in [26] that (a) and (b) are generic; i.e., although there exist cases where the second variation is indefinite and the system is stable, and there exist special types of dissipation that do not result in instability, these are conjectured to be exceptional. The authors of [26] conclude that if we accept the conjecture and ignore these possibilities because of their rarity, we obtain a sense in which the second variation provides a "necessary" and sufficient condition for stability. It is worthwhile mentioning in this context how it is stated in [19] that “Consistent with indefiniteness of the second variation we conjecture that all stratified, incompressible, ideal fluid flows for sufficiently high wavenumbers of the density variation are nonlinearly unstable” (italics is as in [19]). We also have a feeling that the second variation \(\delta ^{2}H\) provides a "necessary" and sufficient condition for stability of the stratified Kolmogorov flow, but because \(\delta ^{2}H\) exactly matches invariant (14) of linearized equations, we attribute indefiniteness of \(\delta ^{2}H\) to the linear instability through resonance amplification of internal gravity wave modes in the presence of critical levels of the main flow, although in view of the short-wave nature of the emerging instability, the role of dissipative factors may be crucial.

Furthermore, following [19] to assume the upper cutoff for the spectrum of short-wave disturbances, we have

$$\begin{aligned} \int \!\!\!\int {\left( {\nabla ^{2}\delta \psi } \right) ^{2}} {}{}\text{ d }x\text{ d }z<K^{2}\int \!\!\!\int {\left( {\nabla \delta \psi } \right) ^{2}} {}{}\text{ d }x\text{ d }z, \end{aligned}$$

where \(K^{2}\) is a (large) positive constant having the meaning of dimensionless wavenumber squared. Now, second variation (C.2) admits a lower estimate

$$\begin{aligned} \delta ^{2}H>\delta ^{2}\tilde{{H}}=\int \!\!\!\int {\left\{ {K^{-2}\left( {\nabla ^{2}\delta \psi } \right) ^{2}+\left( {1+\varepsilon ^{2}\cos ^{2}z} \right) \left( {\delta \xi } \right) ^{2}{}+2\varepsilon \cos z{}{}\nabla ^{2}\delta \psi \delta \xi } \right\} } {}{}\text{ d }x\text{ d }z \end{aligned}$$
(C.3)

and the binary form in (C.3) is positive definite, that is, the stratified Kolmogorov flow is stable according to Lyapunov, if \(\varepsilon ^{2}\cos ^{2}z<\left( {K^{2}-1} \right) ^{-1}\) everywhere in the flow, that is, if \(\varepsilon ^{2}<\left( {K^{2}-1} \right) ^{-1}\). At \(K^{2}\rightarrow \infty \) the zone of guaranteed Lyapunov stability shrinks to zero.

Appendix D: Band-like structure of instability

Table 1 The band-like structure of the possible instability zones as revealed by application of the Galerkin approximations corresponding to \(k=3\) and \(M=1,2,\ldots ,7\) in the non-viscous and non-diffusive case \(R=S=\infty \)
Table 2 The band-like structure of the possible instability zones as revealed by application of the Galerkin approximations corresponding to \(k=3\) and \(M=1,\ldots ,7\) for a slightly viscous and diffusive case when \(R=S=4000\)
Table 3 The roots of the equation \(S_{M} =-1\), see (17), calculated for \(k=3\) and \(M=1,\ldots ,41\). Underlined are the numbers, which coincide exactly with those calculated for the Galerkin first, third, fifth and seventh approximations (see Table 1)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kurgansky, M.V. On short-wave instability of the stratified Kolmogorov flow. Theor. Comput. Fluid Dyn. 36, 575–595 (2022). https://doi.org/10.1007/s00162-022-00613-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00162-022-00613-2

Keywords

Navigation