Skip to main content
Log in

On the Bifurcation Diagram of the Capillary–Gravity Whitham Equation

  • Original Article
  • Published:
Water Waves Aims and scope Submit manuscript

Abstract

We study the bifurcation of periodic travelling waves of the capillary–gravity Whitham equation. This is a nonlinear pseudo-differential equation that combines the canonical shallow water nonlinearity with the exact (unidirectional) dispersion for finite-depth capillary–gravity waves. Starting from the line of zero solutions, we give a complete description of all small periodic solutions, unimodal as well bimodal, using simple and double bifurcation via Lyapunov–Schmidt reductions. Included in this study is the resonant case when one wavenumber divides another. Some bifurcation formulas are studied, enabling us, in almost all cases, to continue the unimodal bifurcation curves into global curves. By characterizing the range of the surface tension parameter for which the integral kernel corresponding to the linear dispersion operator is completely monotone (and, therefore, positive and convex; the threshold value for this to happen turns out to be \(T = \frac{4}{\pi ^2}\), not the critical Bond number \(\frac{1}{3}\)), we are able to say something about the nodal properties of solutions, even in the presence of surface tension. Finally, we present a few general results for the equation and discuss, in detail, the complete bifurcation diagram as far as it is known from analytical and numerical evidence. Interestingly, we find, analytically, secondary bifurcation curves connecting different branches of solutions and, numerically, that all supercritical waves preserve their basic nodal structure, converging asymptotically in \(L^2(\mathbb {S})\) (but not in \(L^\infty \)) towards one of the two constant solution curves.

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

Similar content being viewed by others

Notes

  1. See also the discussion in Sect. 5 concerning related results for the Euler equations.

  2. That is, \((A \bigtriangleup B)=\left( A\cap B^c\right) \cup \left( B\cap A^c\right) \).

  3. Throughout, we use the notation that \(\mathbb {N}_0:=\mathbb {N}\cup \{0\}\).

  4. Note that the function \(T_*(\cdot ;\cdot )\) can be extended to the cases \(n=0\) and \(k=0\) through continuity.

  5. We lack a proof of non-existence of the \(k_2\)-modal waves in the resonant case of Theorem 4.1, but these waves do not seem to exist numerically.

  6. It is possible that the periodised kernel is positive even when the original kernel is not, depending on the period, but we have not investigated that here.

References

  1. Aasen, A., Varholm, K.: Traveling gravity water waves with critical layers. J. Math. Fluid Mech. 20, 161–187 (2018)

    Article  MathSciNet  Google Scholar 

  2. Akers, B.F., Ambrose, D.M., Sulon, D.W.: Periodic travelling interfacial hydroelastic waves with or without mass II: Multiple bifurcations and ripples. Eur. J. Appl. Math. 30, 756–790 (2019)

    Article  MathSciNet  Google Scholar 

  3. Akers, B.F., Ambrose, D.M., Wright, J.D.: Gravity perturbed Crapper waves. Proc. R. Soc. Lond. Ser. A Math. Phys. Eng. Sci 470, 20130526 (2014)

    Article  MathSciNet  Google Scholar 

  4. Ambrose, D.M., Strauss, W.A., Wright, J.D.: Global bifurcation theory for periodic traveling interfacial gravity–capillary waves. Ann. Inst. H. Poincaré Anal. Non Linéaire 33, 1081–1101 (2016)

    Article  MathSciNet  Google Scholar 

  5. Amick, C.J., Turner, R.E.L.: A global theory of internal solitary waves in two-fluid systems. Trans. Am. Math. Soc. 298, 431–484 (1986)

    Article  MathSciNet  Google Scholar 

  6. Arnesen, M.N.: A non-local approach to waves of maximal height for the Degasperis-Procesi equation. arXiv:1808.08057

  7. Arnesen, M.N.: Existence of solitary-wave solutions to nonlocal equations. Discret. Contin. Dyn. Syst. 36, 3483–3510 (2016)

    Article  MathSciNet  Google Scholar 

  8. Aston, P.: Local and global aspects of the (1, n) mode interaction for capillary–gravity waves. Phys. D 52, 415–428 (1991)

    Article  MathSciNet  Google Scholar 

  9. Bhatia, R.: Positive Definite Matrices. Princeton Series in Applied Mathematics. Princeton University Press, Princeton (2007)

    Google Scholar 

  10. Buffoni, B., Toland, J.: Analytic Theory of Global Bifurcation—An Introduction, Princeton Series in Applied Mathematics. Princeton University Press, Princeton (2003)

    Book  Google Scholar 

  11. Córdoba, D., Enciso, A., Grubic, N.: On the existence of stationary splash singularities for the Euler equations. Adv. Math. 288, 922–941 (2016)

    Article  MathSciNet  Google Scholar 

  12. Dancer, E.: Bifurcation theory for analytic operators. Proc. Lond. Math. Soc. XXVI, 359–384 (1973)

    Article  MathSciNet  Google Scholar 

  13. Dancer, E.: Global structure of the solutions set of non-linear real-analytic eigenvalue problems. Proc. Lond. Math. Soc. XXVI, 747–765 (1973)

    Article  Google Scholar 

  14. Ehrnström, M., Escher, J., Wahlén, E.: Steady water waves with multiple critical layers. SIAM J. Math. Anal. 43, 1436–1456 (2011)

    Article  MathSciNet  Google Scholar 

  15. Ehrnström, M., Holden, H., Raynaud, X.: Symmetric waves are traveling waves. Int. Math. Res. Not. 24, 4578–4596 (2009)

    MathSciNet  MATH  Google Scholar 

  16. Ehrnström, M., Johnson, M.A., Claassen, K.M.: Existence of a highest wave in a fully dispersive two-way shallow water model. Arch. Ration. Mech. Anal. (2018). https://doi.org/10.1007/s00205-018-1306-5

    Article  MATH  Google Scholar 

  17. Ehrnström, M., Kalisch, H.: Traveling waves for the Whitham equation. Differ. Integral Equ. 22, 1193–1210 (2009)

    MathSciNet  MATH  Google Scholar 

  18. Ehrnström, M., Kalisch, H.: Global bifurcation for the Whitham equation. Math. Model. Nat. Phenom. 8, 13–30 (2013)

    Article  MathSciNet  Google Scholar 

  19. Ehrnström, M., Wahlèn, E.: Trimodal steady water waves. Arch. Rational Mech. Anal. 216, 449–471 (2015)

    Article  MathSciNet  Google Scholar 

  20. Ehrnström, M., Wahlèn, E.: On Whitham’s conjecture of a highest cusped wave for a nonlocal dispersive equation. Ann. Inst. H. Poincaré Anal. Non Linéaire 36, 1603–1637 (2019)

    Article  MathSciNet  Google Scholar 

  21. Grafakos, L.: Modern Fourier Analysis. Graduate Texts in Mathematics, 3rd edn. Springer, New York (2014)

    Google Scholar 

  22. Hunter, J.K., Vanden-Broeck, J.-M.A.: Solitary and periodic gravity–capillary waves of finite amplitude. J. Fluid Mech. 134, 205–219 (1983)

    Article  MathSciNet  Google Scholar 

  23. Hur, V.M., Johnson, M.A.: Modulational instability in the Whitham equation with surface tension and vorticity. Nonlinear Anal. 129, 104–118 (2015)

    Article  MathSciNet  Google Scholar 

  24. Johnson, M.A., Wright, J.D.: Generalized solitary waves in the gravity-capillary Whitham equation. arXiv:1807.11469

  25. Katznelson, Y.: An Introduction to Harmonic Analysis. Cambridge Mathematical Library, 3rd edn. Cambridge Univerisity Press, Cambridge (2004)

    Book  Google Scholar 

  26. Kielhöfer, H.: Bifurcation Theory. Applied Mathematical Sciences, vol. 156. Springer, New York (2004)

    Book  Google Scholar 

  27. Kozlov, V., Kuznetsov, N.: Steady free-surface vortical flows parallel to the horizontal bottom. Q. J. Mech. Appl. Math. 64, 371–399 (2011)

    Article  MathSciNet  Google Scholar 

  28. Kozlov, V., Lokharu, E.: N-Modal steady water waves with vorticity. J. Math. Fluid Mech. 2017, 1–15 (2017)

    MATH  Google Scholar 

  29. Lannes, D.: The Water Waves Problem, Vol. 188 of Mathematical Surveys and Monographs. Mathematical Analysis and Asymptotics. American Mathematical Society, Providence (2013)

    Google Scholar 

  30. Martin, C., Matioc, B.-V.: Existence of Wilton ripples for water waves with constant vorticity and capillary effects. SIAM J. Appl. Math. 73, 1582–1595 (2013)

    Article  MathSciNet  Google Scholar 

  31. Okamoto, H., Shōji, M.: The Mathematical Theory of Permanent Progressive Water-Waves. Advanced Series in Nonlinear Dynamics, vol. 20. World Scientific Publishing Co., Inc., River Edge (2001)

    Book  Google Scholar 

  32. Paley, R.E.A.C., Wiener, N.: Fourier Transforms in the Complex Domain, Vol. 19 of American Mathematical Society Colloquium Publications. American Mathematical Society, Providence (1987). (Reprint of the 1934 original)

    Google Scholar 

  33. Reeder, J., Shinbrot, M.: On Wilton ripples, II: Rigorous results. Arch. Ration. Mech. Anal. 77, 321–347 (1981)

    Article  MathSciNet  Google Scholar 

  34. Remonato, F., Kalisch, H.: Numerical Bifurcation for the Capillary Whitham Equation. Phys. D 343, 51–62 (2017)

    Article  MathSciNet  Google Scholar 

  35. Schilling, R.L., Song, R., Vondraĉek, Z.: Bernstain Functions. de Gruyter Studies in Mathematics, vol. 37, 2nd edn. Walter de Gruyter & Co., Berlin (2012)

    Google Scholar 

  36. Taylor, M.E.: Partial Differential Equations III. Nonlinear Equations, Applied Mathematical Sciences, 2nd edn. Springer, Berlin (2011)

    Book  Google Scholar 

  37. Toland, J.F., Jones, M.C.W.: The bifurcation and secondary bifurcation of capillary–gravity waves. Proc. R. Soc. Lond. Ser. A 399, 391–417 (1985)

    Article  MathSciNet  Google Scholar 

  38. Trichtchenko, O., Deconinck, B., Wilkening, J.: The instability of Wilton ripples. Wave Motion 66, 147–155 (2016)

    Article  MathSciNet  Google Scholar 

  39. Turner, R.E.L., Vanden-Broeck, J.-M.: The limiting configuration of interfacial gravity waves. Phys. Fluids 29, 372–375 (1986)

    Article  MathSciNet  Google Scholar 

  40. Vanden-Broeck, J.-M.: Nonlinear gravity–capillary standing waves in water of arbitrary uniform depth. J. Fluid Mech. 139, 97–104 (1984)

    Article  MathSciNet  Google Scholar 

  41. Whitham, G.: Variational methods and applications to water waves. Proc. R. Soc. Lond. A 299, 6–25 (1967)

    Article  Google Scholar 

Download references

Acknowledgements

The authors would like to acknowledge valuable input from the referees. Their comments helped improve both the exposition and the mathematical precision of the paper.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Mats Ehrnström.

Additional information

Publisher's Note

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

All authors acknowledge the support by Grant nos. 231668 and 250070 from the Research Council of Norway. M.J. was supported by the National Science Foundation under Grant DMS-1614785.

Appendix A: Bifurcation formulas

Appendix A: Bifurcation formulas

This appendix contains higher order expansions of the quantities in Theroem 3.1 and Theorem 4.1. We start with the first- and second-order terms in the expansion for the speed c(t) in the one-dimensional bifurcation case, which is required by the proof of the global extension in Theorem 3.10. We then proceed to study the first-order terms for the expansions of the functions r and p in the two-dimensional bifurcation case.

1.1 One-Dimensional Bifurcation Case

We begin by determining the derivatives \(\dot{c}(0)\) and \(\ddot{c}(0)\) associated to the bifurcation curve constructed in Theorem 3.1. This can be done either directly using the Lyapunov–Schmidt reduction carried out in the proof of Theorem 3.1 or by the means of bifurcation formulas given for example in [26]. The latter requires an identification between the bifurcation function \(\phi (u,c) = \Pi F(u + \psi (u,c), c)\) used in [26] and the functions v and r used in the proof of Theorem 3.1. This relation is given by \(v(t) = \psi (t \cos (k x), c(t))\).

Here, start from the Lyapunov–Schmidt representation

$$\begin{aligned} \begin{aligned} 0 =&F(t\cos (kx) + v(t) , c_0 + r(t))\\ =&t\cos (kx) + v(t)\\&+ L \left[ (t\cos (kx) + v(t))^2 -(c_0 + r(t)) (t\cos (kx) + v(t)) \right] . \end{aligned} \end{aligned}$$
(A.1)

Here, it is understood that for each t small the function v(t) is a \(2\pi /k\)-periodic function of x. Differentiating (A.1) once with respect to t, evaluating at \(t=0\) and using that \(v(0) = \dot{v}(0) = r(0) = 0\) yields the equation

$$\begin{aligned} \left( 1-c_0L\right) \cos (kx)=0, \end{aligned}$$

which holds by our choice of \(c_0\). Similarly, differentiating (A.1) twice with respect to t and evaluating at \(t=0\) yields

$$\begin{aligned} \begin{aligned} (1-c_0L)\ddot{v}(0)&=2\dot{r}(0)L\cos (kx)-2L\cos ^2(kx)\\&=2\dot{r}(0)l(k)\cos (kx)-\left( 1+l(2k)\cos (2kx)\right) . \end{aligned} \end{aligned}$$
(A.2)

Since \(\int _{-\pi }^\pi v(t)\cos (kx)dx=0\) for all \(|t|\ll 1\), the above implies that \(\dot{r}(0)=0\). Returning to (A.2), it now follows that

$$\begin{aligned} \ddot{v}(0)=\frac{1}{c_0-1}+\frac{l(2k)\cos (2kx)}{c_0l(2k)-1}. \end{aligned}$$
(A.3)

Continuing, we observe that taking the third derivative of (A.1) with respect to t and evaluating at \(t=0\) yield

$$\begin{aligned} \left( 1-c_0L\right) \dddot{v}(0)=3\ddot{r}(0)L\cos (kx)-6L\left( \ddot{v}(0)\cos (kx)\right) . \end{aligned}$$

Using (A.3), we compute that

$$\begin{aligned} L\left( \ddot{v}(0)\cos (kx)\right) =\frac{l(k)\cos (kx)}{c_0-1}+\frac{l(2k)\left( l(k)\cos (kx)+l(3k)\cos (3kx)\right) }{2(c_0l(2k)-1)}. \end{aligned}$$

Using again that \(\int _{-\pi }^\pi v(t)\cos (kx)\,\,{\mathrm {d}}x=0\) for all \(|t|\ll 1\), it follows that

$$\begin{aligned} \ddot{r}(0)=\frac{3}{c_0-1}+\frac{l(2k)}{c_0l(2k)-1}=\frac{3c_0l(2k)-l(2k)-2}{(c_0-1)(c_0l(2k)-1)}, \end{aligned}$$

which is the expression (3.10) for \(\ddot{c}(0)\) given in Theorem 3.6. Note that the above procedure could be continued to obtain asymptotic expansions of r(t) and v(t) to arbitrarily high order in t. We also note that the above result is consistent with the asymptotic formulas in [23].

1.2 Two-Dimensional Bifurcation Case

We now consider the case of a two-dimensional bifurcation as considered in Section 4 above. Recall that the solutions constructed in Theorem 4.1 can be written as

$$\begin{aligned} u(t_1, t_2)&= t_1 \cos (k_1 x) + t_2 \cos (k_2 x) + v(t_1, t_2), \\ c(t_1, t_2)&= c_0 + r(t_1, t_2), \\ \kappa (t_1, t_2)&= \kappa _0 + p(t_1, t_2), \end{aligned}$$

with v of order \(\mathcal {O}(|(t_1, t_2)|^2)\) and r, p of order \(\mathcal {O}(|(t_1,t_2)|)\). We now characterize the order of vanishing of the functions r and p at the origin.

Proposition A.1

Let the functions r and p be as in Theorem 4.1. If \(k_2/k_1 \notin \mathbb {N}_0\), then

$$\begin{aligned} \nabla r(0,0) = 0, \qquad \nabla p(0,0) = 0 \end{aligned}$$

so that, in particular, r and p are of order \(\mathcal {O}(|(t_1,t_2)|^2)\) near the origin. If instead \(k_2/k_1 \in \mathbb {N}_0\), then for any \(\delta >0\) small we have that, in polar coordinates,

$$\begin{aligned} r_\varrho \left( 0,\vartheta \right) = 0, \qquad p_\varrho \left( 0,\vartheta \right) = 0 \end{aligned}$$

if and only if either \(k_2 \notin \{0, 2k_1 \}\) or \((k_2,\vartheta )=\left( 2k_1,\frac{\pi }{2}\right) \).

Proof

We begin the non-resonant case, \(k_2/k_1 \notin \mathbb {N}_0\). From the proof of Theorem 4.1, we know for all \(0<|(t_1,t_2)|\ll 1\) the functions r and p satisfy

$$\begin{aligned} \Psi _i (t_1, t_2, r(t_1, t_2), p(t_1, t_2)) = 0\quad {\mathrm{for}}~~ i=1,2, \end{aligned}$$

where the \(\Psi _i\) are defined in (4.9) and (4.13). Fixing \(j\in \{1,2\}\) we find that differentiating the above with respect to \(t_j\) and evaluating at \((t_1,t_2)=(0,0)\) gives the system of equations

$$\begin{aligned} \left( \begin{array}{cc}\Psi _{1,r}(\mathbf{0}) &{} \Psi _{1,p}(\mathbf{0})\\ \Psi _{2,r}(\mathbf{0}) &{} \Psi _{2,p}(\mathbf{0})\end{array}\right) \left( \begin{array}{c}r_{t_j}(0,0)\\ p_{t_j}(0,0)\end{array}\right) = -\left( \begin{array}{c}\Psi _{1,t_j}(\mathbf{0})\\ \Psi _{2,t_j}(\mathbf{0})\end{array}\right) , \end{aligned}$$
(A.4)

where here \(\mathbf{0}\) denotes the origin in \(\mathbb {R}^4\). Since the above system matrix is invertible by (4.15), it remains to determine the values of \(\Psi _{i,t_j}(\mathbf{0})\) for \(i=1,2\). This can be accomplished by recalling (4.9) and (4.13) and noting that (4.7) implies that

$$\begin{aligned} \frac{\partial ^2 Q_i}{\partial t_j^2}(\mathbf{0}) = -\frac{2}{\pi }\,l(\kappa _0 k_i) \int _{-\pi }^\pi \cos ^3(k_i x) \;{\mathrm {d}}x \end{aligned}$$

and

$$\begin{aligned} \frac{\partial ^2 Q_i}{\partial t_1 \partial t_2}(\mathbf{0}) = \left\{ \begin{aligned} -\frac{2}{\pi } l(\kappa _0 k_2) \int _{-\pi }^\pi \cos ^2(k_1 x) \cos (k_2 x) \;{\mathrm {d}}x,\quad i=1,\\ -\frac{2}{\pi } l(\kappa _0 k_1) \int _{-\pi }^\pi \cos ^2(k_2 x) \cos (k_1 x) \;{\mathrm {d}}x,\quad i=2 \end{aligned}\right. \end{aligned}$$

Consequently, since \(k_2/k_1\notin \mathbb {N}_0\) it follows that \(\Psi _{i,t_j}(\vec {0})=0\) for \(i=1,2\) and hence (A.4) implies that \(r_{t_j}(0,0)=p_{t_j}(0,0)=0\) as claimed. Since \(j\in \{1,2\}\) was arbitrary, this proves the proposition in the non-resonant case.

Now, consider the resonant case when \(k_2/k_1\in \mathbb {N}_0\) and fix \(\delta >0\) small. In this case, for each \(\delta<|\vartheta |<\pi -\delta \) and \(0<\varrho \ll 1\) the functions \(r(\varrho ,\vartheta )\) and \(p(\varrho ,\vartheta )\) satisfy the system

$$\begin{aligned} {\widetilde{\Psi }}_i\left( \varrho ,\vartheta ,r(\varrho ,\vartheta ),p(\varrho ,\vartheta )\right) =0\quad {\mathrm{for}}~~i=1,2, \end{aligned}$$

where here the \({\widetilde{\Psi }}_i\) are as in (4.20) and (4.18). Differentiating this system with respect to \(\varrho \) at \(\varrho =0\) gives the system of equations

$$\begin{aligned} \left( \begin{array}{cc}{\widetilde{\Psi }}_{1,r}(0,\vartheta ,0,0) &{} {\widetilde{\Psi }}_{1,p}(0,\vartheta ,0,0)\\ {\widetilde{\Psi }}_{2,r}(0,\vartheta ,0,0) &{} {\widetilde{\Psi }}_{2,p}(0,\vartheta ,0,0)\end{array}\right) \left( \begin{array}{c}r_\varrho (0,\vartheta )\\ p_\varrho (0,\vartheta )\end{array}\right) = -\left( \begin{array}{c}{\widetilde{\Psi }}_{1,\varrho }(0,\vartheta ,0,0)\\ {\widetilde{\Psi }}_{2,\varrho }(0,\vartheta ,0,0)\end{array}\right) . \end{aligned}$$
(A.5)

As in the non-resonant case, the above system matrix is invertible, this time thanks to (4.21), and hence it remains to determine the values of \({\widetilde{\Psi }}_{i,\varrho }(0,\vartheta ,0,0)\) for \(i=1,2\). Let us begin by determining the value in the case \(i=1\). From (4.20) and the preceding discussion, we know we can write

$$\begin{aligned} {\widetilde{\Psi }}_1(\varrho ,\vartheta ,0,0)=\int _0^1\frac{\partial {\widetilde{Q}}_1}{\partial \varrho }(z\varrho ,\vartheta ,0,0)\,\,{\mathrm {d}}z \end{aligned}$$

where, using (4.7), we have explicitly

$$\begin{aligned} \widetilde{Q}_1(\varrho ,\vartheta ,0,0)&=Q_1(\varrho \cos (\vartheta ),\varrho \sin (\vartheta ),0,0)\\&=-\frac{2\varrho ^2 l(k_0k_1)\cos (\vartheta )\sin (\vartheta )}{\pi }\int _{-\pi }^\pi \cos ^2(k_1 x)\cos (k_2x)\,\,{\mathrm {d}}x. \end{aligned}$$

Clearly then, \(\widetilde{Q}_{2,\varrho \varrho }(0,\vartheta ,0,0)\) is equal to zero if and only if either \(\vartheta =\frac{\pi }{2}\) or \(k_2\notin \{0,2k_1\}\). Since

$$\begin{aligned} {\widetilde{\Psi }}_{1,\varrho }(0,\vartheta ,0,0)=\frac{1}{2}\frac{\partial ^2\widetilde{Q}_1}{\partial \varrho ^2}(0,\vartheta ,0,0) \end{aligned}$$

by above, we have shown that \({\widetilde{\Psi }}_{1,\varrho }(0,\vartheta ,0,0)=0\) if and only if either of the conditions \(\vartheta =\frac{\pi }{2}\) or \(k_2\notin \{0,2k_1\}\) hold.

Similarly, we have

$$\begin{aligned} {\widetilde{\Psi }}_{2,\varrho }(0,\vartheta ,0,0)=\frac{1}{2}\frac{\partial ^2\widetilde{Q}_2}{\partial \varrho ^2}(0,\vartheta ,0,0) \end{aligned}$$

where, using (4.16), we have

$$\begin{aligned} \widetilde{Q}_2(\varrho ,\vartheta ,0,0)=&-\frac{\varrho ^2 l(k_0k_2)}{\pi }\\&\times \int _{-\pi }^\pi \cos (k_2 z)\left[ \cos ^2(\vartheta )\cos ^2(k_1 x)+\sin ^2(\vartheta )\cos ^2(k_2 x)\right] \,{\mathrm {d}}x. \end{aligned}$$

Clearly, \(\widetilde{Q}_{2,\varrho \varrho }(0,\vartheta ,0,0)\) vanishes whenever \(k_2\notin \{0,2k_1\}\). When \(k_2=0\), \(\widetilde{Q}_{2,\varrho \varrho }(0,\vartheta ,0,0)\) does not vanish for any \(\vartheta \), and when \(k_2=2k_1\) it only vanishes when \(\vartheta =\frac{\pi }{2}\). Consequently, \({\widetilde{\Psi }}_{2,\varrho }(0,\vartheta ,0,0)\) vanishes only when either \(k_2\notin \{0,2k_1\}\) or \(\left( k_2,\vartheta \right) =(2k_1,\frac{\pi }{2})\). Together with the results concerning \({\widetilde{\Psi }}_{1,\varrho }\), this completes the proof. \(\square \)

Remark A.2

The special case \(k_2 = 2k_1\) has been found also in the Euler equations (with gravity and vorticity) by the authors of [1]. The special case \(k_2 = 0\) is instead due to the transcritical double bifurcation allowed by the capillary–gravity Whitham equation.

Remark A.3

An explicit example where \(r_\varrho (0,\vartheta ) \ne 0\) can be seen in [34, Figure 6], where the branch of nontrivial solutions has a non-vertical tangent at the bifurcation point in the speed–height plane.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ehrnström, M., Johnson, M.A., Maehlen, O.I.H. et al. On the Bifurcation Diagram of the Capillary–Gravity Whitham Equation. Water Waves 1, 275–313 (2019). https://doi.org/10.1007/s42286-019-00019-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s42286-019-00019-4

Keywords

Mathematics Subject Classification

Navigation