Skip to main content
Log in

Unstable kink and anti-kink profile for the sine-Gordon equation on a \({\mathcal {Y}}\)-junction graph

  • Published:
Mathematische Zeitschrift Aims and scope Submit manuscript

Abstract

The sine-Gordon equation on a metric graph with a structure represented by a \({\mathcal {Y}}\)-junction, is considered. The model is endowed with boundary conditions at the graph-vertex of \(\delta '\)-interaction type, expressing continuity of the derivatives of the wave functions plus a Kirchhoff-type rule for the self-induced magnetic flux. It is shown that particular stationary, kink and kink/anti-kink soliton profile solutions to the model are linearly (and nonlinearly) unstable. To that end, a recently developed linear instability criterion for evolution models on metric graphs by Angulo and Cavalcante (2020), which provides the sufficient conditions on the linearized operator around the wave to have a pair of real positive/negative eigenvalues, is applied. This leads to the spectral study to the linearized operator and of its Morse index. The analysis is based on analytic perturbation theory, Sturm-Liouville oscillation results and the extension theory of symmetric operators. The methods presented in this manuscript have prospect for the study of the dynamics of solutions for the sine-Gordon model on metric graphs with finite bounds or on metric tree graphs and/or loop graphs.

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

References

  1. Ablowitz, M.J., Kaup, D.J., Newell, A.C., Segur, H.: Method for solving the sine-Gordon equation. Phys. Rev. Lett. 30, 1262–1264 (1973)

    Article  MathSciNet  Google Scholar 

  2. Ablowitz, M.J., Kaup, D.J., Newell, A.C., Segur, H.: The inverse scattering transform-Fourier analysis for nonlinear problems. Stud. Appl. Math. 53(4), 249–315 (1974)

    Article  MathSciNet  Google Scholar 

  3. Angulo Pava, J., Cavalcante, M.: Linear instability criterion for the Korteweg-de Vries equation on metric star graphs. Nonlinearity 34(5), 3373–3410 (2021)

    Article  MathSciNet  Google Scholar 

  4. Angulo Pava, J., Cavalcante, M.: Nonlinear Dispersive Equations on Star Graphs, Lecture Notes, Advanced Course, 32o Colóquio Brasileiro de Matemática, IMPA, IMPA, (2019)

  5. Angulo Pava, J., Goloshchapova, N.: Extension theory approach in the stability of the standing waves for the NLS equation with point interactions on a star graph. Adv. Differ. Equ. 23(11–12), 793–846 (2018)

    MathSciNet  MATH  Google Scholar 

  6. Angulo Pava, J., Goloshchapova, N.: On the orbital instability of excited states for the NLS equation with the \(\delta \)-interaction on a star graph. Discrete Contin. Dyn. Syst. 38(10), 5039–5066 (2018)

    Article  MathSciNet  Google Scholar 

  7. Angulo Pava, J., Lopes, O., Neves, A.: Instability of travelling waves for weakly coupled KdV systems. Nonlinear Anal. 69(5–6), 1870–1887 (2008)

    Article  MathSciNet  Google Scholar 

  8. Angulo Pava, J., Natali, F.: On the instability of periodic waves for dispersive equations. Differ. Integral Equ. 29(9–10), 837–874 (2016)

    MathSciNet  MATH  Google Scholar 

  9. Angulo Pava, J., Plaza, R. G.: Unstable kink-soliton profiles for the sine-Gordon equation on a \({\cal{Y}}\)-junction graph with \(\delta \)-interaction. J. Nonlinear Sci. 31, no. 3, art. 50, 1-32 (2021)

  10. Barone, A., Esposito, F., Magee, C.J., Scott, A.C.: Theory and applications of the sine-Gordon equation. Rivista del Nuovo Cimento 1(2), 227–267 (1971)

    Article  Google Scholar 

  11. Barone, A., Paternó, G.: Physics and applications of the Josephson effect. Wiley, New York (1982)

    Book  Google Scholar 

  12. Berezin, F.A., Shubin, M.A.: The Schrödinger equation, vol. 66. Kluwer Acad. Publ, Dordrecht (1991)

    Book  Google Scholar 

  13. Berkolaiko, R. G.: An elementary introduction to quantum graphs, in Geometric and computational spectral theory, A. Girouard, D. Jakobson, M. Levitin, N. Nigam, I. Polterovich, and F. Rochon, eds., vol. 700 of Contemp. Math., Amer. Math. Soc., Providence, RI, pp. 41–72 (2017)

  14. Blank, J., Exner, P., Havlíček, M.: Hilbert space operators in quantum physics, Theoretical and Mathematical Physics, Springer, New York; AIP Press, New York, second ed., (2008)

  15. Bona, J.L., Cascaval, R.C.: Nonlinear dispersive waves on trees. Can. Appl. Math. Q. 16(1), 1–18 (2008)

    MathSciNet  MATH  Google Scholar 

  16. Cadoni, M., Franzin, E., Masella, F., Tuveri, M.: A solution-generating method in Einstein-scalar gravity. Acta Appl. Math. 162, 33–45 (2019)

    Article  MathSciNet  Google Scholar 

  17. Caputo, J.-G., Dutykh, D.: Nonlinear waves in networks: Model reduction for the sine-Gordon equation. Phys. Rev. E 90, 022912 (2014)

    Article  Google Scholar 

  18. Derks, G., Gaeta, G.: A minimal model of DNA dynamics in interaction with RNA-polymerase. Phys. D 240(22), 1805–1817 (2011)

    Article  MathSciNet  Google Scholar 

  19. Drazin, P.G.: Solitons. London Mathematical Society Lecture Note Series, vol. 85. Cambridge University Press, Cambridge (1983)

  20. Dutykh, D., Caputo, J.-G.: Wave dynamics on networks: method and application to the sine-Gordon equation. Appl. Numer. Math. 131, 54–71 (2018)

    Article  MathSciNet  Google Scholar 

  21. Englander, S.W., Kallenbach, N.R., Heeger, A.J., Krumhansl, J.A., Litwin, S.: Nature of the open state in long polynucleotide double helices: possibility of soliton excitations. Proc. Natl. Acad. Sci. USA 77(12), 7222–7226 (1980)

    Article  Google Scholar 

  22. Franzin, E., Cadoni, M., Tuveri, M.: Sine-Gordon solitonic scalar stars and black holes. Phys. Rev. D 97(12), 124018, 7 (2018)

  23. Frenkel, J., Kontorova, T.: On the theory of plastic deformation and twinning. Acad. Sci. U.S.S.R. J. Phys. 1, 137–149 (1939)

    MathSciNet  MATH  Google Scholar 

  24. Grillakis, M., Shatah, J., Strauss, W.: Stability theory of solitary waves in the presence of symmetry. II. J. Funct. Anal. 94(2), 308–348 (1990)

    Article  MathSciNet  Google Scholar 

  25. Grunnet-Jepsen, A., Fahrendorf, F., Hattel, S., Grønbech-Jensen, N., Samuelsen, M.: Fluxons in three long coupled Josephson junctions. Phys. Lett. A 175(2), 116–120 (1993)

    Article  Google Scholar 

  26. Henry, D.B., Perez, J.F., Wreszinski, W.F.: Stability theory for solitary-wave solutions of scalar field equations. Comm. Math. Phys. 85(3), 351–361 (1982)

    Article  MathSciNet  Google Scholar 

  27. Ivancevic, V.G., Ivancevic, T.T.: Sine-Gordon solitons, kinks and breathers as physical models of nonlinear excitations in living cellular structures. J. Geom. Symmetry Phys. 31, 1–56 (2013)

    MathSciNet  MATH  Google Scholar 

  28. Josephson, B.D.: Supercurrents through barriers. Adv. Phys. 14(56), 419–451 (1965)

    Article  Google Scholar 

  29. Kato, T.: Perturbation Theory for Linear Operators, Classics in Mathematics, Springer-Verlag, Berlin, Second ed., (1980)

  30. Knobel, R.: An introduction to the mathematical theory of waves, vol. 3 of Student Mathematical Library, American Mathematical Society, Providence, RI; Institute for Advanced Study (IAS), Princeton, NJ, (2000). IAS/Park City Mathematical Subseries

  31. Kogan, V.G., Clem, J.R., Kirtley, J.R.: Josephson vortices at tricrystal boundaries. Phys. Rev. B 61, 9122–9129 (2000)

    Article  Google Scholar 

  32. Krasnosel’skiĭ, M. A.: Positive solutions of operator equations, Translated from the Russian by Richard E. Flaherty; edited by Leo F. Boron, P. Noordhoff Ltd. Groningen, (1964)

  33. Lopes, O.: A linearized instability result for solitary waves. Discrete Contin. Dyn. Syst. 8(1), 115–119 (2002)

    Article  MathSciNet  Google Scholar 

  34. Mugnolo, D., Noja, D., Seifert, C.: Airy-type evolution equations on star graphs. Anal. PDE 11(7), 1625–1652 (2018)

    Article  MathSciNet  Google Scholar 

  35. Naimark, M. A.: Linear differential operators. Part I: Elementary theory of linear differential operators, Frederick Ungar Publishing Co., New York, (1967)

  36. Naimark, M. A.: Linear differential operators. Part II: Linear differential operators in Hilbert space, Frederick Ungar Publishing Co., New York, (1968)

  37. Nakajima, K., Onodera, Y.: Logic design of Josephson network. II. J. Appl. Phys. 49(5), 2958–2963 (1978)

    Article  Google Scholar 

  38. Nakajima, K., Onodera, Y., Ogawa, Y.: Logic design of Josephson network. J. Appl. Phys. 47(4), 1620–1627 (1976)

    Article  Google Scholar 

  39. Pazy, A.: Semigroups of linear operators and applications to partial differential equations. Applied Mathematical Sciences, vol. 44. Springer-Verlag, New York (1983)

  40. Reed, M., Simon, B.: Methods of modern mathematical physics. Academic Press, New York - London, I. Functional analysis (1972)

  41. Reed, M., Simon, B.: Methods of modern mathematical physics. II. Fourier analysis, self-adjointness, Academic Press – Harcourt Brace Jovanovich, Publishers, New York - London, (1975)

  42. Reed, M., Simon, B.: Methods of modern mathematical physics. Academic Press - Harcourt Brace Jovanovich, Publishers, New York - London, IV. Analysis of operators (1978)

  43. Sabirov, K., Rakhmanov, S., Matrasulov, D., Susanto, H.: The stationary sine-Gordon equation on metric graphs: exact analytical solutions for simple topologies. Phys. Lett. A 382(16), 1092–1099 (2018)

    Article  MathSciNet  Google Scholar 

  44. Scott, A.C.: Nonlinear science, Emergence and dynamics of coherent structures. Oxford Texts in Applied and Engineering Mathematics, vol. 8, 2nd edn. Oxford University Press, Oxford (2003)

  45. Scott, A.C., Chu, F.Y.F., McLaughlin, D.W.: The soliton: a new concept in applied science. Proc. IEEE 61(10), 1443–1483 (1973)

    Article  MathSciNet  Google Scholar 

  46. Scott, A.C., Chu, F.Y.F., Reible, S.A.: Magnetic-flux propagation on a Josephson transmission line. J. Appl. Phys. 47(7), 3272–3286 (1976)

    Article  Google Scholar 

  47. Shatah, J., Strauss, W.: Spectral condition for instability, in Nonlinear PDE’s, dynamics and continuum physics (South Hadley, MA, 1998), J. Bona, K. Saxton, and R. Saxton, eds., vol. 255 of Contemp. Math., Amer. Math. Soc., Providence, RI, pp. 189–198 (2000)

  48. Susanto, H., Karjanto, N., Zulkarnain, Nusantara, T., Widjanarko, T.: Soliton and breather splitting on star graphs from tricrystal Josephson junctions, Symmetry 11, pp. 271–280 (2019)

  49. Susanto, H., van Gils, S.: Existence and stability analysis of solitary waves in a tricrystal junction. Phys. Lett. A 338(3), 239–246 (2005)

    Article  Google Scholar 

  50. Tahtadžjan, L. A., Faddeev, L. D.: The Hamiltonian system connected with the equation \(u_{ }{}_{\eta }+{\rm sin}\ u=0\). Trudy Mat. Inst. Steklov 142, 254–266, 271 (1976)

  51. Tsuei, C.C., Kirtley, J.R.: Phase-sensitive evidence for \({d}\)-wave pairing symmetry in electron-doped cuprate superconductors. Phys. Rev. Lett. 85, 182–185 (2000)

    Article  Google Scholar 

  52. Tsuei, C.C., Kirtley, J.R., Chi, C.C., Yu-Jahnes, L.S., Gupta, A., Shaw, T., Sun, J.Z., Ketchen, M.B.: Pairing symmetry and flux quantization in a tricrystal superconducting ring of YBa\(_2\)Cu\(_3\)O\(_{7-\delta }\). Phys. Rev. Lett. 73, 593–596 (1994)

    Article  Google Scholar 

Download references

Acknowledgements

J. Angulo was supported in part by CNPq/Brazil Grant and Math-amsud program CAPES/Brazil (Nonlinear and fractional evolutions equations: dispersion, dynamics, well-posedness and F.A. tools). The work of R. G. Plaza was partially supported by DGAPA-UNAM, program PAPIIT, grant IN-104922.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jaime Angulo Pava.

Additional information

Publisher's Note

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

Appendix

Appendix

1.1 Extension Theory

For the sake of completeness, in this part of the appendix we develop the extension theory of symmetric operators suitable for our needs. For further information on the subject the reader is referred to the monographs by Naimark [35, 36]. The following classical result, known as the von-Neumann decomposition theorem, can be found in [41], p. 138.

Theorem A.1

Let A be a closed, symmetric operator, then

$$\begin{aligned} D(A^*)=D(A)\oplus {\mathcal {N}}_{-i} \oplus {\mathcal {N}}_{+i}. \end{aligned}$$
(A.1)

with \({\mathcal {N}}_{\pm i}= \ker (A^*{\mp } iI)\). Therefore, for \(u\in D(A^*)\) and \(u=x+y+z\in D(A)\oplus {\mathcal {N}}_{-i} \oplus {\mathcal {N}}_{+i}\),

$$\begin{aligned} A^*u=Ax+(-i)y+iz. \end{aligned}$$
(A.2)

Remark A.2

The direct sum in (A.1) is not necessarily orthogonal.

The following propositions provide a strategy for estimating the Morse-index of the self-adjoint extensions and can be found in Naimark [36] (Theorem 16, p. 44) and Reed and Simon, vol. 2, [41] (see Theorem X.2, p. 140).

Proposition A.3

Let A be a densely defined lower semi-bounded symmetric operator (that is, \(A\ge mI\)) with finite deficiency indices, \(n_{\pm }(A)=k<\infty \), in the Hilbert space \({\mathcal {H}}\), and let \({\widehat{A}}\) be a self-adjoint extension of A. Then the spectrum of \({\widehat{A}}\) in \((-\infty , m)\) is discrete and consists of, at most, k eigenvalues counting multiplicities.

Proposition A.4

Let A be a densely defined, closed, symmetric operator in some Hilbert space H with deficiency indices equal \(n_{\pm }(A)=1\). All self-adjoint extensions \(A_\theta \) of A may be parametrized by a real parameter \(\theta \in [0,2\pi )\) where

$$\begin{aligned} \begin{aligned} D(A_\theta )&=\{x+c\phi _+ + \zeta e^{i\theta }\phi _{-}: x\in D(A), \zeta \in {\mathbb {C}}\},\\ A_\theta (x + \zeta \phi _+ + \zeta e^{i\theta }\phi _{-})&= Ax+i \zeta \phi _+ - i \zeta e^{i\theta }\phi _{-}, \end{aligned} \end{aligned}$$

with \(A^*\phi _{\pm }=\pm i \phi _{\pm }\), and \(\Vert \phi _+\Vert =\Vert \phi _-\Vert \).

Next Proposition can be found in Naimark [36] (see Theorem 9, p. 38).

Proposition A.5

All self-adjoint extensions of a closed, symmetric operator which has equal and finite deficiency indices have one and the same continuous spectrum.

The following result was used in the proof of Proposition  3.10.

Proposition A.6

Let \({\mathcal {Y}}\) be a \({\mathcal {Y}}\)-junction of type I. Consider the following closed symmetric operator, \(({\mathcal {H}}, D({\mathcal {H}}))\), densely defined on \(L^2({\mathcal {Y}})\) by

$$\begin{aligned} \begin{aligned} {\mathcal {H}}&=\Big (\Big (-c_j^2\frac{d^2}{dx^2}\Big )\delta _{j,k} \Big ),\;\;1\leqq j, k\leqq 3,\\ D({\mathcal {H}})&= \Big \{(v_j)_{j=1}^3\in H^2({\mathcal {Y}}): c_1v_1'(0-)\\&=c_2v'_2(0+)=c_3v_3'(0+)=0,\;\; \sum _{j=2}^3 c_jv_j(0+)-c_1v_1(0-)=0 \Big \}, \end{aligned} \end{aligned}$$
(A.3)

with \(c_j>0\), \(1\leqq j\leqq 3\), and \(\delta _{j,k}\) being the Kronecker symbol. Then, the deficiency indices are \(n_{\pm }( {\mathcal {H}})=1\). Therefore, we have that all the self-adjoint extensions of \(( {\mathcal {H}}, D( {\mathcal {H}}))\) are given by the one-parameter family \(({\mathcal {H}}_\lambda , D({\mathcal {H}}_\lambda ))\), \(\lambda \in {\mathbb {R}}\), with \({\mathcal {H}}_\lambda \equiv {\mathcal {H}}\) and \(D({\mathcal {H}}_\lambda )\) defined by

$$\begin{aligned} D({\mathcal {H}}_\lambda )= & {} \Big \{(u_j)_{j=1}^3\in H^2({\mathcal {Y}}): c_1u'_1(0-)=c_2u'_2(0+)\nonumber \\= & {} c_3u'_3(0+),\;\; \sum _{j=2}^3 c_ju_j(0+)-c_1u_1(0-)=\lambda c_1u'_1(0-) \Big \}. \end{aligned}$$
(A.4)

Proof

We show initially that the adjoint operator \(( {\mathcal {H}}^*, D( {\mathcal {H}}^*))\) of \(( {\mathcal {H}}, D( {\mathcal {H}}))\) is given by

$$\begin{aligned} {\mathcal {H}}^*={\mathcal {H}}, \quad D( {\mathcal {H}}^*)=\{(u_j)_{j=1}^ 3\in H^2({\mathcal {Y}}) : c_1u'_1(0-)=c_2u'_2(0+)=c_3u'_3(0+)\}. \end{aligned}$$
(A.5)

Indeed, formally for \(\mathbf{u} =(u_1, u_2, u_3), \mathbf{v} =(v_1, v_2, v_3)\in H^2({\mathcal {Y}})\) we have

$$\begin{aligned} \begin{aligned} \langle {\mathcal {H}}{} \mathbf{u} , \mathbf{v} \rangle&=c_1^2u_1(0-)v_1' (0-)- c_1^2u'_1(0-)v_1 (0-) \\&\quad + \sum _{j=2}^ 3c_j^2[u'_{j}(0+)v_j(0+) -u_{j}(0+)v'_j(0+)]+\langle \mathbf{u} ,{\mathcal {H}} \mathbf{v} \rangle \\&\equiv R+ \langle \mathbf{u} ,{\mathcal {H}} \mathbf{v} \rangle . \end{aligned} \end{aligned}$$
(A.6)

Hence, for \(\mathbf{u} , \mathbf{v} \in D({\mathcal {H}})\) we obtain \(R=0\) in (A.6) and so the symmetric property of \({\mathcal {H}}\). Next, let us denote \(D_0^*:=\{(u_j)_{j=1}^ 3\in H^2({\mathcal {G}}) : c_1u'_1(0-)=c_2u'_2(0+)=c_3u'_3(0+)\}\). We shall show that \(D_0^*=D( {\mathcal {H}}^*)\). Indeed, from (A.6) we obtain for \(\mathbf{v} \in D_0^*\) and \(\mathbf{u} \in D( {\mathcal {H}})\) that \(R=0\) and so \(\mathbf{v} \in D ({\mathcal {H}}^*)\) with \({\mathcal {H}}^*\mathbf{v} = {\mathcal {H}}{} \mathbf{v} \). Let us show the inclusion \(D_0^*\supseteq D({\mathcal {H}}^*)\). Take \(\mathbf{u} =(u_1, u_2, u_3)\in D({\mathcal {H}})\), then for any \(\mathbf{v} =(v_1, v_2, v_3)\in D({\mathcal {H}}^*)\) we have from (A.6)

$$\begin{aligned} \langle {\mathcal {H}} \mathbf{u} , \mathbf{v} \rangle = R+\langle \mathbf{u} ,{\mathcal {H}}^* \mathbf{v} \rangle =\langle \mathbf{u} ,{\mathcal {H}} \mathbf{v} \rangle . \end{aligned}$$
(A.7)

Thus, we arrive for any \(\mathbf{u} \in D({\mathcal {H}})\) at the equality

$$\begin{aligned} \sum _{j=2}^ 3c_j^2v'_{j}(0+)u_j(0+) - c_1^2v_1' (0-)u_1(0-)=0. \end{aligned}$$
(A.8)

Next, let us consider \(\mathbf{u} =(u_1, u_2, 0)\in D({\mathcal {H}})\). Then \(c_1 u_1'(0-)=c_2 u_2'(0+)=0\) and from Eqs. (A.3) and (A.8), we obtain \( [c_2v'_2(0+)-c_1v'_1(0-)]c_1 u_1(0-)=0\). Thus, by choosing \(u_1(0-)\ne 0\) follows \(c_1v'_1(0-)=c_2v'_2(0+)\). Similarly, we have \(c_1v'_1(0-)=c_3v'_3(0+)\). Therefore, \(\mathbf{v} \in D_0^*\). This shows the relations in (A.5).

Now, from (A.5) we obtain that the deficiency indices for \(( {\mathcal {H}}, D( {\mathcal {H}}))\) are \(n_{\pm }( {\mathcal {H}})=1\). Indeed, \(\ker ({\mathcal {H}}^*\pm iI)=[\Phi _{\pm }]\) with \(\Phi _{\pm }\) defined by

$$\begin{aligned} \Phi _{\pm }= \left( \underset{x<0}{e^{\frac{ ik_{\mp }}{c_1}x}}, \underset{x>0}{-e^{\frac{- ik_{\mp }}{c_2}x}}, \underset{x>0}{-e^{\frac{ -ik_{\mp }}{c_3}x}}\right) , \end{aligned}$$
(A.9)

\(k^2_{\mp }={\mp } i\), \(\mathrm {Im}\,(k_{-})<0\) and \(\mathrm {Im}\,(k_{+})<0\). Moreover, \(\Vert \Psi _{-}\Vert _{L^ 2({\mathcal {Y}})}=\Vert \Psi _{+}\Vert _{L^ 2({\mathcal {Y}})}\).

Next, from extension theory for symmetric operators we have that every self-adjoint extension \((\widehat{{\mathcal {H}}}, D(\widehat{{\mathcal {H}}}))\) of \(({\mathcal {H}}, D({\mathcal {H}}))\) is characterized by \(D(\widehat{{\mathcal {H}}})\subset D({\mathcal {H}}^*)\), and \(\mathbf{u} =(u_1,u_2,u_3)\in D(\widehat{{\mathcal {H}}})\) if and only if \( \sum _{j=2}^3 c_ju_j(0+)-c_1u_1(0-)=\lambda c_1u'_{1}(0-), \) where, for \(\theta \in [0, 2\pi )-\{\pi /2\}\),

$$\begin{aligned} \lambda = -\Big (\sum _{j=1}^3 c_j \Big )\frac{1+e^{i\theta }}{e^{-i\frac{\pi }{4}}+e^{i(\theta +\frac{\pi }{4})}} \in {\mathbb {R}}. \end{aligned}$$

Thus, the set of self-adjoint extensions \((\widehat{{\mathcal {H}}}, D(\widehat{{\mathcal {H}}}))\) of the symmetric operator \(({\mathcal {H}}, D({\mathcal {H}}))\) can be seen as one-parametrized family \(({\mathcal {H}}_\lambda , D({\mathcal {H}}_\lambda ))\) defined by \({\mathcal {H}}_\lambda ={\mathcal {H}}\) and \(D({\mathcal {H}}_\lambda )\) by (A.4). This finishes the proof. \(\square \)

1.2 Morse index calculations

The following result gives a precise value for the Morse-index of the operator \( {\mathcal {L}}_{\lambda } \) in (3.31) associated to the kink/anti-kink profile \((\phi _j)\) obtained in Sect. 3.2.1, when we consider the domain \({\mathcal {C}}_1\cap D ({\mathcal {L}}_{\lambda })\), \({\mathcal {C}}_1=\{(u_j)_{j=1}^3\in L^2({\mathcal {Y}}): u_2=u_3\}\). We establish this result in order to possible future study related to the stability properties of kink/anti-kink profiles. We conjecture that these profiles are in fact unstable (see Remark 3.18 and [3]).

Proposition A.7

Let \(\lambda \in (-\infty , -\frac{\pi }{2})\) and consider \({\mathcal {B}}_2={\mathcal {C}}_1 \cap D ({\mathcal {L}}_{\lambda })\). Then \({\mathcal {L}}_{\lambda } : {\mathcal {B}}_2\rightarrow {\mathcal {C}}_1\) is well defined and \(n( {\mathcal {L}}_{\lambda } |_{{\mathcal {B}}_2} )=2\).

Proof

The proof is based on analytic perturbation theory. By Proposition 3.21 we have for \(\Phi '_{\lambda _0}=(2\phi '_1, \phi '_2, \phi '_2)\in {\mathcal {C}}_1\), \(\phi _i\equiv \phi _{i, \lambda _0}\), that \({{\,\mathrm{span}\,}}\{\Phi '_{\lambda _0}\}=\ker ({\mathcal {L}}_{\lambda _0} |_{{\mathcal {B}}_2} )\) (\(\lambda _0\equiv -\frac{\pi }{2}\)). Moreover, by the proof of Proposition 3.22 we have for \(\Lambda _1=(\phi '_1, \phi '_2, \phi '_2)\in {\mathcal {C}}_1\cap H^1({\mathcal {Y}})\) that the quadratic form \({\mathcal {Q}}\) in (3.36) satisfies \({\mathcal {Q}}(\Lambda _1)<0\) thus \(n({\mathcal {L}}_{\lambda _0} |_{{\mathcal {B}}_2} )=1\). Let us denote by \(\chi _{_{\lambda _0}}\) the unique negative eigenvalue for \({\mathcal {L}}_{\lambda _0}|_{{\mathcal {B}}_2}\). We note in this point of the analysis that by using the classical perturbation theory and the convergence \({\mathcal {L}}_{\lambda } \rightarrow {\mathcal {L}}_{\lambda _0}\) as \(\lambda \rightarrow \lambda _0\) in the generalized sense (Kato [29]) we obtain that given a closed curve \(\Gamma \) such that \(\sigma _0=\{\chi _{_{\lambda _0}}, 0\}\) belongs to the inner domain of \(\Gamma \), then we can only conclude that \( n({\mathcal {L}}_{\lambda } |_{{\mathcal {B}}_2} )\geqq 1\) (by Proposition 3.21) for \(\lambda \approx \lambda _0\). So we will need to determine exactly how the eigenvalue zero will move, either to right or to left . In this form, we divide our analysis into several steps:

  1. i)

    The mapping \(\lambda \in (-\infty , \infty )\rightarrow \Psi _\lambda = (\phi _{i, a_i(\lambda )})\) is real-analytic and we have the convergence \( \Psi _\lambda - \Psi _{\lambda _0}\rightarrow 0\), as \(\lambda \rightarrow \lambda _0\), in \(H^1({\mathcal {Y}})\). Then, it follows that \({\mathcal {L}}_{\lambda }\) converges to \({\mathcal {L}}_{\lambda _0}\) as \(\lambda \rightarrow \lambda _0\) in the generalized sense (see proof of Proposition 3.24). Moreover, the family \(\{{\mathcal {L}}_\lambda \}_{\lambda \in I}\) represents a real-analytic family of self-adjoint operators of type (B) in the sense of Kato (see [29]) on \({\mathcal {C}}_1\). Therefore, from Theorem IV-3.16 from Kato [29] we obtain \(\Gamma \subset \rho ({\mathcal {L}}_{\lambda })\) for \(|\lambda -\lambda _0|\) sufficiently small, and \(\sigma ({\mathcal {L}}_{\lambda })\) is likewise separated by \(\Gamma \) into two parts, such that the part of \(\sigma ({\mathcal {L}}_{\lambda })\) inside \(\Gamma \) consists of a finite number of eigenvalues with total multiplicity (algebraic) equal to two. Then, it follows from Kato-Rellich’s theorem (see [40]) the existence of two analytic functions, \(\Omega \) and \(\Pi \), defined in a neighborhood of \(\lambda _0\) with \(\Omega : (\lambda _0-\delta , \lambda _0+\delta )\rightarrow {\mathbb {R}}\) and \(\Pi : (\lambda _0-\delta , \lambda _0+\delta )\rightarrow {\mathcal {C}}_1\) such that \(\Omega (\lambda _0)=0\) and \(\Pi (\lambda _0)=\Phi '_{\lambda _0}\). For all \(\lambda \in (\lambda _0-\delta , \lambda _0+\delta )\), \(\Omega (\lambda )\) is the simple isolated second eigenvalue of \({\mathcal {L}}_{\lambda }\), and \(\Pi (\lambda )\) is the associated eigenvector for \(\Omega (\lambda )\). Moreover, \(\delta \) can be chosen small enough to ensure that, for \(\lambda \in (\lambda _0-\delta , \lambda _0+\delta )\), the spectrum of \({\mathcal {L}}_{\lambda }\) in \({\mathcal {C}}_1\) is positive, except for, at most, the first two eigenvalues.

  2. ii)

    Using Taylor’s theorem, we obtain for sufficiently small \(\delta \) the following expansions:

    $$\begin{aligned} \Omega (\lambda )=\beta (\lambda -\lambda _0)+ O(|\lambda -\lambda _0|^2), \;\; \Pi (\lambda )=\Phi '_{\lambda _0} + \Pi '(\lambda _0) (\lambda -\lambda _0) + \mathbf{O} (|\lambda -\lambda _0|^2),\nonumber \\ \end{aligned}$$
    (A.10)

    where \(\beta = \Omega '(\lambda _0)\). Moreover, by analyticity we also obtain the expansion

    $$\begin{aligned} \Psi _\lambda -\Psi _{\lambda _0}=(\lambda -\lambda _0) \partial _\lambda \Psi _\lambda |_{\lambda =\lambda _0} + \mathbf{O} (|\lambda -\lambda _0|^2). \end{aligned}$$
    (A.11)

    Next, we determine the sign of \(\beta \). Thus, we compute \(\langle {\mathcal {L}}_{\lambda } \Pi (\lambda ), \Phi '_{\lambda _0} \rangle \) in two different ways. On one hand, using \({\mathcal {L}}_{\lambda } \Pi (\lambda )= \Omega (\lambda )\Pi (\lambda )\), (A.10) leads to

    $$\begin{aligned} \langle {\mathcal {L}}_{\lambda } \Pi (\lambda ), \Phi '_{\lambda _0} \rangle = \beta (\lambda -\lambda _0)\Vert \Phi '_{\lambda _0}\Vert ^2 + O(|\lambda -\lambda _0|^2). \end{aligned}$$
    (A.12)

    On the other hand, since \(\Phi '_{\lambda _0}\in D({\mathcal {L}}_{\lambda } )\cap {\mathcal {C}}_1\) for all \(\lambda \ne \lambda _0\) and \({\mathcal {L}}_{\lambda } \) is self-adjoint, we get \(\langle {\mathcal {L}}_{\lambda } \Pi (\lambda ), \Phi '_{\lambda _0} \rangle =\langle \Pi (\lambda ), {\mathcal {L}}_{\lambda } \Phi '_{\lambda _0} \rangle \). Now, using the notation \(\phi _{i,\lambda }=\phi _{i,a_i(\lambda )}\), we have from (3.13) the relation \( {\mathcal {L}}_{\lambda } \Phi '_{\lambda _0} = {\mathcal {L}}_{\lambda _0} \Phi '_{\lambda _0} +{\mathcal {A}} \Phi '_{\lambda _0} ={\mathcal {A}} \Phi '_{\lambda _0}\) where \({\mathcal {A}}\) is the diagonal-matrix \( {\mathcal {A}}=\Big ((\cos (\phi _{i,\lambda })-\cos (\phi _{i,\lambda _0}))\delta _{i,k}\Big )\), \( 1\leqq i,k\leqq 3\). Next, for \(G_i(\lambda )=\cos (\phi _{i,\lambda })\) we have the expansion

    $$\begin{aligned} G_i(\lambda )-G_i(\lambda _0)=-\sin (\phi _{i,\lambda _0})\Big ( \partial _\lambda \phi _{i,\lambda }|_{_{\lambda =\lambda _0}}\Big )(\lambda -\lambda _0)+ O(|\lambda -\lambda _0|^2). \end{aligned}$$

    Then \( {\mathcal {A}}=\Big ((-\sin (\phi _{i,\lambda _0})\partial _\lambda \phi _{i,\lambda }|_{_{\lambda =\lambda _0}}(\lambda -\lambda _0))\delta _{i,k}\Big ) + \mathbf{O} (|\lambda -\lambda _0|^2)\), \(1\leqq i,k\leqq 3\). Thus,

    $$\begin{aligned} \begin{aligned} \langle {\mathcal {L}}_{\lambda } \Pi (\lambda ), \Phi '_{\lambda _0} \rangle&=\langle \Pi (\lambda ), {\mathcal {L}}_{\lambda } \Phi '_{\lambda _0} \rangle =\langle \Phi '_{\lambda _0} + \Pi '(\lambda _0) (\lambda -\lambda _0) + \mathbf{O} (|\lambda -\lambda _0|^2), {\mathcal {A}} \Phi '_{\lambda _0} \rangle \\&=\Big \langle \Phi '_{\lambda _0}, \Big ((-\sin (\phi _{i,\lambda _0})\partial _\lambda \phi _{i,\lambda }|_{_{\lambda =\lambda _0}}(\lambda -\lambda _0))\delta _{i,k}\Big )\Phi '_{\lambda _0} \Big \rangle + O(|\lambda -\lambda _0|^2)\\&\equiv (\lambda -\lambda _0) \eta + O(|\lambda -\lambda _0|^2), \end{aligned}\nonumber \\ \end{aligned}$$
    (A.13)

    with \( \eta \equiv 4a_1'(\lambda _0) \int _{-\infty }^0 (\phi '_{1,\lambda _0})^3\sin (\phi _{1,\lambda _0})\, dx+2a_1'(\lambda _0) \int _0^{\infty } (\phi '_{2,\lambda _0})^3\sin (-\phi _{2,\lambda _0}) \, dx\). Next, we obtain the sign of \(\eta \). Indeed, from the qualitative properties of the anti-kink profile \(\phi _{i,\lambda _0}\) we get immediately that the two integrals above are positive. Next, from the relation in (3.28) we obtain that \(a_1'(\lambda _0)=\frac{1}{3}\) (indeed we always have \(a_1'(\lambda )>0\)). Then \(\eta >0\). Next, from (A.12) and (A.13) there follows \( \beta =\frac{\eta }{\Vert \Phi '_{\lambda _0}\Vert ^2} + O(|\lambda -\lambda _0|)\). Therefore, from (A.10) there exists \(\delta _0>0\) sufficiently small such that \(\Omega (\lambda )>0\) for any \(\lambda \in (\lambda _0, \lambda _0+\delta _0)\), and \(\Omega (\lambda )<0\) for any \(\lambda \in (\lambda _0-\delta _0, \lambda _0)\). Thus, in the space \( {\mathcal {C}}_1\), the Morse index \(n({\mathcal {L}}_\lambda |_{{\mathcal {B}}_2})=1\) for \(\lambda >\lambda _0\) and \(\lambda \approx \lambda _0\) (equality that is used in Proposition 3.22), and \(n({\mathcal {L}}_\lambda |_{{\mathcal {B}}_2})=2\) for \(\lambda <\lambda _0\) and \(\lambda \approx \lambda _0\).

  3. iii)

    Next, since \(\ker ({\mathcal {L}}_\lambda )=\{{0}\}\) for \(\lambda \ne -\frac{\pi }{2}\), we obtain via a continuation argument based on the Riesz-projection operator (see proof of Proposition 3.24) that for any \(\lambda \in (-\infty , -\frac{\pi }{2})\) follows \(n({\mathcal {L}}_\lambda |_{{\mathcal {B}}_2})=2\) . This finishes the proof.

\(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Angulo Pava, J., Plaza, R.G. Unstable kink and anti-kink profile for the sine-Gordon equation on a \({\mathcal {Y}}\)-junction graph. Math. Z. 300, 2885–2915 (2022). https://doi.org/10.1007/s00209-021-02899-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00209-021-02899-0

Keywords

Mathematics Subject Classification

Navigation