Skip to main content
Log in

The Focusing NLS Equation with Step-Like Oscillating Background: The Genus 3 Sector

  • Published:
Communications in Mathematical Physics Aims and scope Submit manuscript

Abstract

We consider the Cauchy problem for the focusing nonlinear Schrödinger equation with initial data approaching different plane waves \(A_j\mathrm {e}^{\mathrm {i}\phi _j}\mathrm {e}^{-2\mathrm {i}B_jx}\), \(j=1,2\) as \(x\rightarrow \pm \infty \). The goal is to determine the long-time asymptotics of the solution, according to the value of \(\xi =x/t\). The general situation is analyzed in a recent paper where we develop the Riemann–Hilbert approach and detect different asymptotic scenarios, depending on the relationships between the parameters \(A_1\), \(A_2\), \(B_1\), and \(B_2\). In particular, in the shock case \(B_1<B_2\), some scenarios include genus 3 sectors, i.e., ranges of values of \(\xi \) where the leading term of the asymptotics is given in terms of hyperelliptic functions attached to a Riemann surface \(M(\xi )\) of genus three. The present paper is devoted to the complete asymptotic analysis in such a sector.

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
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15
Fig. 16
Fig. 17
Fig. 18
Fig. 19

Similar content being viewed by others

References

  1. Andreiev, K., Egorova, I., Lange, T.L., Teschl, G.: Rarefaction waves of the Korteweg-de Vries equation via nonlinear steepest descent. J. Differ. Equ. 261(10), 5371–5410 (2016)

    Article  ADS  MathSciNet  Google Scholar 

  2. Bertola, M., Minakov, A.: Laguerre polynomials and transitional asymptotics of the modified Korteweg-de Vries equation for step-like initial data. Anal. Math. Phys. 9(4), 1761–1818 (2019)

    Article  MathSciNet  Google Scholar 

  3. Bikbaev, R.F.: Structure of a shock wave in the theory of the Korteweg-de Vries equation. Phys. Lett. A 141(5–6), 289–293 (1989)

    Article  ADS  MathSciNet  Google Scholar 

  4. Biondini, G.: Riemann problems and dispersive shocks in self-focusing media. Phys. Rev. E 98(5), 052220 (2018)

    Article  ADS  Google Scholar 

  5. Biondini, G., Fagerstrom, E., Prinari, B.: Inverse scattering transform for the defocusing nonlinear Schrödinger equation with fully asymmetric non-zero boundary conditions. Phys. D 333, 117–136 (2016)

    Article  MathSciNet  Google Scholar 

  6. Biondini, G., Kovačič, G.: Inverse scattering transform for the focusing nonlinear Schrödinger equation with nonzero boundary conditions. J. Math. Phys. 55(3), 031506 (2014)

    Article  ADS  MathSciNet  Google Scholar 

  7. Biondini, G., Mantzavinos, D.: Long-time asymptotics for the focusing nonlinear Schrödinger equation with nonzero boundary conditions at infinity and asymptotic stage of modulational instability. Commun. Pure Appl. Math. 70(12), 2300–2365 (2017)

    Article  Google Scholar 

  8. Biondini, G., Prinari, B.: On the spectrum of the Dirac operator and the existence of discrete eigenvalues for the defocusing nonlinear Schrödinger equation. Stud. Appl. Math. 132(2), 138–159 (2014)

    Article  MathSciNet  Google Scholar 

  9. Boutet de Monvel, A., Kotlyarov, V.P., Shepelsky, D.: Focusing NLS equation: long-time dynamics of step-like initial data. Int. Math. Res. Not. IMRN 7, 1613–1653 (2011)

  10. Boutet de Monvel, A., Lenells, J., Shepelsky, D.: The focusing NLS equation with step-like oscillating background: scenarios of long-time asymptotics. Commun. Math. Phys. 383(2), 893–952 (2021)

    Article  ADS  MathSciNet  Google Scholar 

  11. Boutet de Monvel, A., Lenells, J., Shepelsky, D.: The focusing NLS equation with step-like oscillating background: asymptotics in a transition zone. arXiv:2006.01137 (2020)

  12. Buckingham, R., Venakides, S.: Long-time asymptotics of the nonlinear Schrödinger equation shock problem. Commun. Pure Appl. Math. 60(9), 1349–1414 (2007)

    Article  Google Scholar 

  13. Buslaev, V., Fomin, V.: An inverse scattering problem for the one-dimensional Schrödinger equation on the entire axis, Russian, with English summary. Vestnik Leningrad. Univ. 17(1), 56–64 (1962)

    MathSciNet  Google Scholar 

  14. Cohen, A., Kappeler, T.: Scattering and inverse scattering for steplike potentials in the Schrödinger equation. Indiana Univ. Math. J. 34(1), 127–180 (1985)

    Article  MathSciNet  Google Scholar 

  15. Deift, P., Kamvissis, S., Kriecherbauer, T., Zhou, X.: The Toda rarefaction problem. Commun. Pure Appl. Math. 49(1), 35–83 (1996)

    Article  MathSciNet  Google Scholar 

  16. Deift, P., Venakides, S., Zhou, X.: The collisionless shock region for the long-time behavior of solutions of the KdV equation. Commun. Pure Appl. Math. 47(2), 199–206 (1994)

    Article  MathSciNet  Google Scholar 

  17. Deift, P., Venakides, S., Zhou, X.: New results in small dispersion KdV by an extension of the steepest descent method for Riemann-Hilbert problems. Int. Math. Res. Notices 6, 286–299 (1997)

    Article  MathSciNet  Google Scholar 

  18. Deift, P., Zhou, X.: A steepest descent method for oscillatory Riemann-Hilbert problems. Asymptotics for the MKdV equation. Ann. Math. (2) 137(2), 295–368 (1993)

    Article  MathSciNet  Google Scholar 

  19. Demontis, F., Prinari, B., van der Mee, C., Vitale, F.: The inverse scattering transform for the defocusing nonlinear Schrödinger equations with nonzero boundary conditions. Stud. Appl. Math. 131(1), 1–40 (2013)

    Article  MathSciNet  Google Scholar 

  20. Demontis, F., Prinari, B., van der Mee, C., Vitale, F.: The inverse scattering transform for the focusing nonlinear Schrödinger equation with asymmetric boundary conditions. J. Math. Phys. 55(10), 101505 (2014)

    Article  ADS  MathSciNet  Google Scholar 

  21. Egorova, I., Gladka, Z., Kotlyarov, V., Teschl, G.: Long-time asymptotics for the Korteweg-de Vries equation with step-like initial data. Nonlinearity 26(7), 1839–1864 (2013)

    Article  ADS  MathSciNet  Google Scholar 

  22. Egorova, I., Michor, J., Teschl, G.: Long-time asymptotics for the Toda shock problem: non-overlapping spectra. Zh. Mat. Fiz. Anal. Geom. 14(4), 406–451 (2018)

    Article  MathSciNet  Google Scholar 

  23. Farkas, H.M., Kra, I.: Riemann Surfaces, 2nd edn., Graduate Texts in Mathematics, vol. 71, p. xvi+363. Springer, New York (1992)

  24. Fay, J.D.: Theta functions on Riemann surfaces, Lecture Notes in Mathematics, vol. 352, p. iv+137. Springer, Berlin (1973)

  25. Grava, T., Minakov, A.: On the long-time asymptotic behavior of the modified Korteweg-de Vries equation with step-like initial data. SIAM J. Math. Anal. 52(6), 5892–5993 (2020)

    Article  MathSciNet  Google Scholar 

  26. Gurevich, A.V., Pitaevskiĭ, L.P.: Decay of initial discontinuity in the Korteweg-de Vries equation. JETP Lett. 17, 193–195 (1973)

    ADS  Google Scholar 

  27. Gurevich, A.V., Pitaevskiĭ, L.P.: Nonstationary structure of a collisionless shock wave. Soviet Phys. JETP 38, 291–297 (1974)

    ADS  Google Scholar 

  28. Its, A.R.: Asymptotic behavior of the solutions to the nonlinear Schrödinger equation, and isomonodromic deformations of systems of linear differential equations (Russian). Dokl. Akad. Nauk SSSR 261(1), 14–18: Soviet Math. Dokl. 24, 452–456 (1981)

  29. Khruslov, E Ya.: Decay of initial steplike discontinuity in the Korteweg-de Vries equation. JETP Lett. 21, 217–218 (1975)

    ADS  Google Scholar 

  30. Khruslov, E.Ya.: Asymptotic behavior of the solution of the Cauchy problem for the Korteweg-de Vries equation with steplike initial data (Russian). Math. USSR-Sb. 28(2), 229–248 (1976)

  31. Kotlyarov, V., Minakov, A.: Dispersive shock wave, generalized Laguerre polynomials, and asymptotic solitons of the focusing nonlinear Schrödinger equation. J. Math. Phys. 60(12), 123501 (2019)

    Article  ADS  MathSciNet  Google Scholar 

  32. Leach, J.A., Needham, D.J.: The large-time development of the solution to an initial-value problem for the Korteweg-de Vries equation. I. Initial data has a discontinuous expansive step. Nonlinearity 21(10), 2391–2408 (2008)

    Article  ADS  MathSciNet  Google Scholar 

  33. Leach, J.A., Needham, D.J.: The large-time development of the solution to an initial-value problem for the Korteweg-de Vries equation. II. Initial data has a discontinuous compressive step. Mathematika 60(2), 391–414 (2014)

    Article  MathSciNet  Google Scholar 

  34. Lenells, J.: The Nonlinear Steepest Descent Method for Riemann-Hilbert Problems of Low Regularity. Indiana Math. J. 66(4), 1287–1332 (2017)

    Article  MathSciNet  Google Scholar 

  35. Lenells, J.: Matrix Riemann-Hilbert problems with jumps across Carleson contours. Monatsh. Math. 186(1), 111–152 (2018)

    Article  MathSciNet  Google Scholar 

  36. Minakov, A.: Asymptotics of step-like solutions for the Camassa-Holm equation. J. Differ. Equ. 261(11), 6055–6098 (2016)

    Article  ADS  MathSciNet  Google Scholar 

  37. Muskhelishvili, N.I.: Singular integral equations, Boundary problems of function theory and their application to mathematical physics; Translated from the second (1946) Russian edition and with a preface by J. R. M. Radok; Corrected reprint of the 1953 English translation, Dover Publications, Inc., New York, 447 (1992)

  38. Novokshenov, V.Yu.: Time asymptotics for soliton equations in problems with step initial conditions. J. Math. Sci. (N.Y.) 125(5), 717–749 (2005)

  39. Olver, F.W.J., Lozier, D.W., Boisvert, R.F., Clark, C.W. (eds.): NIST Handbook of Mathematical Functions, p. xvi+951. Cambridge University Press, Cambridge (2010)

  40. Venakides, S.: Long time asymptotics of the Korteweg-de Vries equation. Trans. Am. Math. Soc. 293(1), 411–419 (1986)

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

J. Lenells acknowledges support from the Göran Gustafsson Foundation, the Ruth and Nils-Erik Stenbäck Foundation, the Swedish Research Council, Grant No. 2015-05430, and the European Research Council, Grant Agreement No. 682537.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Anne Boutet de Monvel.

Additional information

Communicated by H.-T.Yau.

Publisher's Note

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

Appendices

Appendix A. Exact Solution in Terms of Parabolic Cylinder Functions

Let X denote the cross \(X:=X_1\cup \dots \cup X_4\subset \mathbb {C}\), where the rays

$$\begin{aligned}&X_1:=\lbrace s\mathrm {e}^{\frac{\mathrm {i}\pi }{4}}\mid 0\le s<\infty \rbrace ,&\quad&X_2:=\lbrace s\mathrm {e}^{\frac{3\mathrm {i}\pi }{4}}\mid 0\le s<\infty \rbrace ,\nonumber \\&X_3:=\lbrace s\mathrm {e}^{-\frac{3\mathrm {i}\pi }{4}}\mid 0\le s<\infty \rbrace ,&X_4:=\lbrace s\mathrm {e}^{-\frac{\mathrm {i}\pi }{4}}\mid 0\le s<\infty \rbrace \end{aligned}$$
(A.1)

are oriented toward the origin as in Fig. 20.

Fig. 20
figure 20

The contour \(X=X_1\cup \dots \cup X_4\)

Define the function \(\nu :\mathbb {C}\rightarrow [0,\infty )\) by \(\nu (q):=\frac{1}{2\pi }\ln (1+|q|^2)\). Let \(\rho (q,z):=\mathrm {e}^{\mathrm {i}\nu (q)\ln _{-\pi /2}z}\), i.e., \(\rho (q,z)=z^{\mathrm {i}\nu (q)}\) with the branch cut along the negative imaginary axis.

We consider the following family of RH problems parametrized by \(q\in \mathbb {C}\):

$$\begin{aligned} {\left\{ \begin{array}{ll} m^X(q,\,\cdot \,)\in I+\dot{E}^2(\mathbb {C}\setminus X),&{}\\ m_+^X(q,z)=m_-^X(q,z)v^X(q,z)&{}\text {for a.e. }z\in X, \end{array}\right. } \end{aligned}$$
(A.2)

where the jump matrix \(v^X(q,z)\) is defined by

$$\begin{aligned} v^X(q,z):={\left\{ \begin{array}{ll} \begin{pmatrix}1&{}0\\ q\rho (q,z)^{-2}\mathrm {e}^{2\mathrm {i}z^2}&{}1\end{pmatrix},&{}k\in X_1,\\ \begin{pmatrix}1&{}{{\bar{q}}}\rho (q,z)^2\mathrm {e}^{-2\mathrm {i}z^2}\\ 0&{}1\end{pmatrix},&{}k\in X_2,\\ \begin{pmatrix}1&{}0\\ -\frac{q}{1+|q|^2}\rho (q,z)^{-2}\mathrm {e}^{2\mathrm {i}z^2}&{}1\end{pmatrix},&{}k\in X_3,\\ \begin{pmatrix}1&{}-\frac{{{\bar{q}}}}{1+|q|^2}\rho (q,z)^2\mathrm {e}^{-2\mathrm {i}z^2}\\ 0&{}1\end{pmatrix},&k\in X_4. \end{array}\right. } \end{aligned}$$
(A.3)

The matrix \(v^X\) has entries that oscillate rapidly as \(z\rightarrow 0\) and \(v^X\) is not continuous at \(z=0\); however \(v^X(q,\,\cdot \,)-I\in L^2(X)\cap L^{\infty }(X)\). The RH problem (A.2) can be solved explicitly in terms of parabolic cylinder functions [28].

Theorem A.1

The RH problem (A.2) has a unique solution \(m^X(q,z)\) for each \(q\in \mathbb {C}\). This solution satisfies

$$\begin{aligned} m^X(q,z)=I+\frac{\mathrm {i}}{z} \begin{pmatrix} 0&{}-\mathrm {e}^{-\pi \nu }\beta ^X(q)\\ \mathrm {e}^{\pi \nu }\overline{\beta ^X(q)}&{}0\end{pmatrix}+\mathrm {O}\left( \frac{1}{z^2}\right) ,\quad z\rightarrow \infty ,\quad q\in \mathbb {C}, \end{aligned}$$
(A.4)

where the error term is uniform with respect to \(\arg z\in [0,2\pi ]\) and q in compact subsets of \(\mathbb {C}\), and the function \(\beta ^X(q)\) is defined by

$$\begin{aligned} \beta ^X(q):=\frac{\sqrt{\nu (q)}}{2}\mathrm {e}^{\mathrm {i}\left( -\frac{3\pi }{4}-2\nu (q)\ln 2-\arg q+\arg \Gamma (\mathrm {i}\nu (q))\right) },\quad q\in \mathbb {C}. \end{aligned}$$
(A.5)

Moreover, for each compact subset \(K\subset \mathbb {C}\),

$$\begin{aligned} \sup _{q\in K}\sup _{z\in \mathbb {C}\setminus X}|m^X(q,z)|<\infty . \end{aligned}$$
(A.6)

Proof

Uniqueness follows because \(\det v^X=1\). Define \(m^{(X1)}\) by

$$\begin{aligned} m^{(X1)}(q,z):=m^X(q,z)\rho (q,z)^{\sigma _3}\mathrm {e}^{-\mathrm {i}z^2\sigma _3}. \end{aligned}$$
(A.7)

Using that \(\rho (q,\,\cdot \,)\) has a branch cut along the negative imaginary axis, we see that \(m^X\) satisfies the jump condition in (A.2) iff \(m^{(X1)}\) satisfies

$$\begin{aligned} m_+^{(X1)}(q,z)=m_-^{(X1)}(q,z)v^{(X1)}(q,z)\quad \text {for a.e. }z\in X \cup \mathrm {i}\,\mathbb {R}_-, \end{aligned}$$

where

$$\begin{aligned} v^{(X1)}:={\left\{ \begin{array}{ll} \begin{pmatrix}1&{}0\\ q&{}1\end{pmatrix},&{}z\in X_1,\\ \begin{pmatrix}1&{}{{\bar{q}}}\\ 0&{}1\end{pmatrix},&{}z\in X_2,\\ \begin{pmatrix}1&{}0\\ -\frac{q}{1+|q|^2}&{}1\end{pmatrix},&{}z\in X_3,\\ \left( \frac{\rho _+(q,z)}{\rho _-(q,z)}\right) ^{\sigma _3}=\begin{pmatrix}\frac{1}{1+|q|^2}&{}0\\ 0&{}1+|q|^2\end{pmatrix},&{}z\in \mathrm {i}\,\mathbb {R}_-,\\ \begin{pmatrix}1&{}-\frac{{{\bar{q}}}}{1+|q|^2}\\ 0&{}1\end{pmatrix},&z\in X_4, \end{array}\right. } \end{aligned}$$

with all contours oriented toward the origin. We next merge the contours \(X_1\) and \(X_2\) along \(\mathrm {i}\,\mathbb {R}_+\) and the contours \(X_3\) and \(X_4\) along \(\mathrm {i}\,\mathbb {R}_-\). Thus we let

$$\begin{aligned} \psi (q,z)=m^{(X1)}(q,z)B(q,z), \end{aligned}$$
(A.8)

where

$$\begin{aligned} B(z):={\left\{ \begin{array}{ll} \begin{pmatrix}1&{}0\\ q&{}1\end{pmatrix},&{}\arg z\in \left( \frac{\pi }{4},\frac{\pi }{2}\right) ,\\ \begin{pmatrix}1&{}-{{\bar{q}}}\\ 0&{}1\end{pmatrix},&{}\arg z\in \left( \frac{\pi }{2},\frac{3\pi }{4}\right) ,\\ \begin{pmatrix}1&{}0\\ -\frac{q}{1+|q|^2}&{}1\end{pmatrix},&{}\arg z\in \left( \frac{5\pi }{4},\frac{3\pi }{2}\right) ,\\ \begin{pmatrix}1&{}\frac{{{\bar{q}}}}{1+|q|^2}\\ 0&{}1\end{pmatrix},&{}\arg z\in \left( \frac{3\pi }{2},\frac{7\pi }{4}\right) ,\\ \,I,&{}\text {else}. \end{array}\right. } \end{aligned}$$

Then \(m^X\) satisfies the jump condition in (A.2) iff \(\psi \) satisfies

$$\begin{aligned} \psi _+(q,z)=\psi _-(q,z)v^{\psi }(q)\quad \text {for a.e. }z\in \mathrm {i}\,\mathbb {R}, \end{aligned}$$
(A.9)

where the constant matrix \(v^{\psi }\) is defined by

$$\begin{aligned} v^{\psi }(q):=\begin{pmatrix}1+|q|^2&{}\bar{q}\\ q&{}1\end{pmatrix},\quad k\in \mathrm {i}\,\mathbb {R}, \end{aligned}$$
(A.10)

and the contour \(\mathrm {i}\,\mathbb {R}\) is oriented downward.

Let

$$\begin{aligned} \beta ^X:=-\frac{\mathrm {e}^{\frac{\pi \mathrm {i}}{4}}2^{-2\mathrm {i}\nu }\mathrm {e}^{\frac{\pi \nu }{2}}\sqrt{\pi }}{q\sqrt{2}\,\Gamma (-\mathrm {i}\nu )}. \end{aligned}$$
(A.11)

Then \(\beta ^X\) can be written as in (A.5). Define a sectionally analytic function \(\psi (q,z)\) by

$$\begin{aligned} \psi (q,z):=\begin{pmatrix} \psi _{11}(q,z)&{}\frac{\left( \frac{\mathrm {d}}{\mathrm {d}z}-2\mathrm {i}z\right) \psi _{22}(q,z)}{4\mathrm {e}^{\pi \nu }\overline{\beta ^X(q)}}\\ \frac{\left( \frac{\mathrm {d}}{\mathrm {d}z}+2\mathrm {i}z\right) \psi _{11}(q,z)}{4\mathrm {e}^{-\pi \nu }\beta ^X(q)}&{}\psi _{22}(q,z)\end{pmatrix},\quad q\in \mathbb {C},\quad z\in \mathbb {C}\setminus \mathrm {i}\,\mathbb {R}, \end{aligned}$$
(A.12)

where the functions \(\psi _{11}\) and \(\psi _{22}\) are defined by

$$\begin{aligned} \psi _{11}(q,z)&:={\left\{ \begin{array}{ll} 2^{-\mathrm {i}\nu }\mathrm {e}^{-\frac{3\pi \nu }{4}}D_{\mathrm {i}\nu }(2\mathrm {e}^{-\frac{3\pi \mathrm {i}}{4}}z),&{}{\text {Re}}z<0,\\ 2^{-\mathrm {i}\nu }\mathrm {e}^{\frac{\pi \nu }{4}}D_{\mathrm {i}\nu }(2\mathrm {e}^{\frac{\pi \mathrm {i}}{4}}z),&{}{\text {Re}}z>0, \end{array}\right. } \end{aligned}$$
(A.13a)
$$\begin{aligned} \psi _{22}(q,z)&:={\left\{ \begin{array}{ll} 2^{\mathrm {i}\nu }\mathrm {e}^{\frac{5\pi \nu }{4}}D_{-\mathrm {i}\nu }(2\mathrm {e}^{\frac{3\pi \mathrm {i}}{4}}z),&{}{\text {Re}}z<0,\\ 2^{\mathrm {i}\nu }\mathrm {e}^{\frac{\pi \nu }{4}}D_{-\mathrm {i}\nu }(2\mathrm {e}^{-\frac{\pi \mathrm {i}}{4}}z),&{}{\text {Re}}z>0, \end{array}\right. } \end{aligned}$$
(A.13b)

and \(D_a(z)\) denotes the parabolic cylinder function.

Since \(D_a(z)\) is an entire function of both a and z, \(\psi (q,z)\) is analytic in the left and right halves of the complex z-plane with a jump across the imaginary axis. The function \(\psi \) satisfies

$$\begin{aligned} \partial _z\psi +2\mathrm {i}z\sigma _3\psi =4\begin{pmatrix}0&{}\beta ^X\mathrm {e}^{-\pi \nu }\\ {{\bar{\beta }}}^X\mathrm {e}^{\pi \nu }&{}0\end{pmatrix}\psi ,\quad q\in \mathbb {C},\quad z\in \mathbb {C}\setminus \mathrm {i}\,\mathbb {R}. \end{aligned}$$

Since \(\psi _+\) and \(\psi _-\) solve the same second order ODE, there exists a function \(v^{\psi }\) independent of z such that (A.9) holds. Setting \(z=0\) and using that

$$\begin{aligned} D_{\mathrm {i}\nu }(0)=\frac{\sqrt{\pi }\,2^{\frac{\mathrm {i}\nu }{2}}}{\Gamma \left( \frac{1}{2}(1-\mathrm {i}\nu )\right) },\quad D_{\mathrm {i}\nu }'(0)=-\frac{\sqrt{\pi }\,2^{\frac{1}{2}(1+\mathrm {i}\nu )}}{\Gamma \left( -\frac{\mathrm {i}\nu }{2}\right) }\,, \end{aligned}$$

we find

$$\begin{aligned} v^{\psi }(q)=\psi _-(q,0)^{-1}\psi _+(q,0)=\begin{pmatrix}1+|q|^2&{}\bar{q}\\ q&{}1\end{pmatrix}. \end{aligned}$$

Hence \(\psi \) satisfies (A.9). This shows that \(m^X\) satisfies the jump condition in (A.2).

For each \(\delta >0\), the parabolic cylinder function satisfies the asymptotic formula [39]

$$\begin{aligned} D_a(z)&=z^a\mathrm {e}^{-\frac{z^2}{4}}\left( 1-\frac{a(a-1)}{2z^2}+\mathrm {O}(z^{-4})\right) \\&\quad -\frac{\sqrt{2\pi }\mathrm {e}^{\frac{z^2}{4}}z^{-a-1}}{\Gamma (-a)}\left( 1+\frac{(a+1)(a+2)}{2z^2}+\mathrm {O}(z^{-4})\right) \\&\qquad \times {\left\{ \begin{array}{ll} 0,&{}\arg z\in [-\frac{3\pi }{4}+\delta ,\frac{3\pi }{4}-\delta ],\\ \mathrm {e}^{\mathrm {i}\pi a},&{}\arg z\in [\frac{\pi }{4}+\delta ,\frac{5\pi }{4}-\delta ],\\ \mathrm {e}^{-\mathrm {i}\pi a},&{}\arg z\in [-\frac{5\pi }{4}+\delta ,-\frac{\pi }{4}-\delta ], \end{array}\right. }\quad z\rightarrow \infty ,\quad a\in \mathbb {C}, \end{aligned}$$

where the error terms are uniform with respect to a in compact subsets and \(\arg z\) in the given ranges. Using this formula and the identity

$$\begin{aligned} \frac{\mathrm {d}}{\mathrm {d}z}D_a(z)=\frac{z}{2}D_a(z)-D_{a+1}(z), \end{aligned}$$

the asymptotic equation (A.4) follows from a tedious but straightforward computation. The boundedness in (A.6) is a consequence of (A.4) and the definition (A.12) of \(\psi (q,z)\). \(\square \)

Remark A.2

The definition (A.12) of \(\psi \) can be motivated as follows. Since \((\partial _z+2\mathrm {i}z\sigma _3)\psi \) and \(\psi \) have the same jump across \(\mathrm {i}\,\mathbb {R}\), the function \((\partial _z\psi +2\mathrm {i}z\sigma _3\psi )\psi ^{-1}\) is entire. Suppose \(m^X=1+m_1^X(q)z^{-1}+\mathrm {O}(z^{-2})\) as \(z\rightarrow \infty \), where the matrix \(m_1^X(q)\) is independent of z. Then we expect

$$\begin{aligned} \psi =\left( I+\frac{m_1^X}{z}+\mathrm {O}(z^{-2})\right) \rho ^{\sigma _3}\mathrm {e}^{-\mathrm {i}z^2\sigma _3},\quad z\rightarrow \infty , \end{aligned}$$
(A.14)

which suggests that

$$\begin{aligned} (\partial _z\psi +2\mathrm {i}z\sigma _3\psi )\psi ^{-1}=2\mathrm {i}[\sigma _3,m_1^X]+\mathrm {O}(z^{-1}),\quad z\rightarrow \infty . \end{aligned}$$
(A.15)

Strictly speaking, due to the factor B(qz) in (A.8), equation (A.14) is not valid for z close to \(\mathrm {i}\,\mathbb {R}\). Equation (A.15) implies that \((\partial _z\psi +2\mathrm {i}z\sigma _3\psi )\psi ^{-1}\) is bounded; hence a constant. Thus

$$\begin{aligned} \partial _z\psi +2\mathrm {i}z\sigma _3\psi =2\mathrm {i}[\sigma _3,m_1^X]\psi . \end{aligned}$$

Letting

$$\begin{aligned} \beta _{12}=4\mathrm {i}(m_1^X)_{12},\qquad \beta _{21}=-4\mathrm {i}(m_1^X)_{21}, \end{aligned}$$

we find that the (11) and (22) entries of \(\psi \) satisfy the equations

$$\begin{aligned} {\left\{ \begin{array}{ll} \partial _z^2\psi _{11}+(4z^2+2\mathrm {i}-\beta _{12}\beta _{21})\psi _{11}=0,&{}\\ \partial _z^2\psi _{22}+(4z^2-2\mathrm {i}-\beta _{12}\beta _{21})\psi _{22}=0,&{} \end{array}\right. }z\in \mathbb {C}\setminus \mathrm {i}\,\mathbb {R}. \end{aligned}$$

Introducing the new variable \(\zeta \) by \(\zeta :=2\mathrm {e}^{-\frac{3\pi \mathrm {i}}{4}}z\), we find that \(f(\zeta ):=\psi _{11}(z)\) satisfies the parabolic cylinder equation

$$\begin{aligned} \partial _{\zeta }^2f+\left( \frac{1}{2}-\frac{\zeta ^2}{4}+a\right) f=0 \end{aligned}$$

for \(a=\frac{\mathrm {i}}{4}\beta _{12}\beta _{21}\). This means that

$$\begin{aligned} \psi _{11}(z)={\left\{ \begin{array}{ll} c_1D_a(2\mathrm {e}^{-\frac{3\pi \mathrm {i}}{4}}z)+c_2D_a(2\mathrm {e}^{\frac{\pi \mathrm {i}}{4}}z),&{}{\text {Re}}z>0,\\ c_3D_a(2\mathrm {e}^{-\frac{3\pi \mathrm {i}}{4}}z)+c_4D_a(2\mathrm {e}^{\frac{\pi \mathrm {i}}{4}}z),&{}{\text {Re}}z<0, \end{array}\right. } \end{aligned}$$

for some constants \(\lbrace c_j\rbrace _1^4\). Since

$$\begin{aligned} D_a(\zeta )=\zeta ^a\mathrm {e}^{-\frac{\zeta ^2}{4}}\left( 1+\mathrm {O}(\zeta ^{-2})\right) ,\quad \zeta \rightarrow \infty ,\quad |\arg \zeta |<\frac{3\pi }{4}-\delta , \end{aligned}$$

the sought-after asymptotics (A.14) of \(\psi \) can be obtained by choosing

$$\begin{aligned} a=\mathrm {i}\nu ,\quad c_1=0,\quad c_2=2^{-\mathrm {i}\nu }\mathrm {e}^{\frac{\pi \nu }{4}},\quad c_3=2^{-\mathrm {i}\nu }\mathrm {e}^{-\frac{3\pi \nu }{4}},\quad c_4=0. \end{aligned}$$

This yields the expression (A.13a) for \(\psi _{11}\); the expression (A.13b) for \(\psi _{22}\) is derived in a similar way. The functions \(\psi _{12}\) and \(\psi _{21}\) can be obtained from the equations

$$\begin{aligned} \psi _{21}=\beta _{12}^{-1}(\psi _{11}'(z)+2\mathrm {i}z\psi _{11}(z)),\qquad \psi _{12}=\beta _{21}^{-1}(\psi _{22}'(z)-2\mathrm {i}z\psi _{22}(z)). \end{aligned}$$

Using that \(\det \psi =1\), the (21) and (12) entries of the jump condition \((\psi _-)^{-1}\psi _+=v^{\psi }\) evaluated at \(z=0\) then yield

$$\begin{aligned} v_{21}^{\psi }(q)=-\frac{\sqrt{2\pi }\mathrm {e}^{\frac{\pi \mathrm {i}}{4}} 2^{-2\mathrm {i}\nu +1}\mathrm {e}^{-\frac{\pi \nu }{2}}}{\beta _{12}\Gamma (-\mathrm {i}\nu )},\qquad v_{12}^{\psi }(q)=-\frac{\sqrt{2\pi }\mathrm {e}^{-\frac{\pi \mathrm {i}}{4}} 2^{2\mathrm {i}\nu +1}\mathrm {e}^{\frac{3\pi \nu }{2}}}{\beta _{21}\Gamma (\mathrm {i}\nu )}. \end{aligned}$$

Thus \(v^{\psi }\) has the form (A.10) provided that \(\beta _{12}=4\mathrm {e}^{-\pi \nu }\beta ^X\) and \(\beta _{21}=4\mathrm {e}^{\pi \nu }{{\bar{\beta }}}^X\) with \(\beta ^X\) given by (A.11). This motivates the form of equation (A.12).

Appendix B. Exact Solution in Terms of Airy Functions

Let \(Y:=Y_1\cup \dots \cup Y_4\subset \mathbb {C}\) denote the union of the four rays

$$\begin{aligned}&Y_1:=\lbrace s\mid 0\le s<\infty \rbrace ,&\qquad&Y_2:=\lbrace s\mathrm {e}^{\frac{2\mathrm {i}\pi }{3}}\mid 0\le s<\infty \rbrace ,\nonumber \\&Y_3:=\lbrace -s\mid 0\le s<\infty \rbrace ,&Y_4:=\lbrace s\mathrm {e}^{-\frac{2\mathrm {i}\pi }{3}}\mid 0\le s<\infty \rbrace , \end{aligned}$$
(B.1)

oriented toward the origin as in Fig. 21.

Fig. 21
figure 21

The sectors \(S_j\), \(j=1,\dots ,4\)

Define the open sectors \(\lbrace S_j\rbrace _1^4\) by

$$\begin{aligned}&S_1:=\lbrace \arg k\in (0,\tfrac{2\pi }{3})\rbrace ,&\qquad&S_2:=\lbrace \arg k\in (\tfrac{2\pi }{3},\pi )\rbrace ,\\&S_3:=\lbrace \arg k\in (\pi ,\tfrac{4\pi }{3})\rbrace ,&S_4:=\lbrace \arg k\in (\tfrac{4\pi }{3},2\pi )\rbrace . \end{aligned}$$

Let \(\omega :=\mathrm {e}^{\frac{2\pi \mathrm {i}}{3}}\). Define the function \(m^{\mathrm {Ai}}(\zeta )\) for \(\zeta \in \mathbb {C}\setminus Y\) by

$$\begin{aligned} m^{\mathrm {Ai}}(\zeta ):=\Psi (\zeta )\times {\left\{ \begin{array}{ll} \,\mathrm {e}^{\frac{2}{3}\zeta ^{3/2}\sigma _3},&{}\zeta \in S_1\cup S_4,\\ \begin{pmatrix}1&{}0\\ -1&{}1\end{pmatrix}\mathrm {e}^{\frac{2}{3}\zeta ^{3/2}\sigma _3},&{}\zeta \in S_2,\\ \begin{pmatrix}1&{}0\\ 1&{}1\end{pmatrix}\mathrm {e}^{\frac{2}{3}\zeta ^{3/2}\sigma _3},&\zeta \in S_3, \end{array}\right. } \end{aligned}$$
(B.2)

where

$$\begin{aligned} \Psi (\zeta ):={\left\{ \begin{array}{ll} \begin{pmatrix}\mathrm {Ai}(\zeta )&{}\mathrm {Ai}(\omega ^2\zeta )\\ \mathrm {Ai}'(\zeta )&{}\omega ^2\mathrm {Ai}'(\omega ^2\zeta )\end{pmatrix}\mathrm {e}^{-\frac{\pi \mathrm {i}}{6}\sigma _3},&{}\zeta \in \mathbb {C}^+,\\ \begin{pmatrix}\mathrm {Ai}(\zeta )&{}-\omega ^2\mathrm {Ai}(\omega \zeta )\\ \mathrm {Ai}'(\zeta )&{}-\mathrm {Ai}'(\omega \zeta )\end{pmatrix}\mathrm {e}^{-\frac{\pi \mathrm {i}}{6}\sigma _3},&\zeta \in \mathbb {C}^-. \end{array}\right. } \end{aligned}$$

Note that \(\det m^{\mathrm {Ai}}(\zeta )=\mathrm {e}^{\pi \mathrm {i}/6}/(2\pi )\) is constant and nonzero. For each integer \(N\ge 0\), define the asymptotic approximations \(m_{\mathrm {as},N}^{\mathrm {Ai}}(\zeta )\) and \(m_{\mathrm {as},N}^{\mathrm {Ai},\mathrm {inv}}(\zeta )\) of \(m^{\mathrm {Ai}}(\zeta )\) and \(m^{\mathrm {Ai}}(\zeta )^{-1}\), respectively, by

$$\begin{aligned} m_{\mathrm {as},N}^{\mathrm {Ai}}(\zeta )&:=\frac{\mathrm {e}^{\frac{\mathrm {i}\pi }{12}}}{2\sqrt{\pi }}\sum _{k=0}^N\frac{1}{\left( \frac{2}{3}\zeta ^{3/2}\right) ^k}\,\zeta ^{-\frac{1}{4}\sigma _3}\!\begin{pmatrix}(-1)^ku_k&{}u_k\\ -(-1)^k\nu _k&{}\nu _k\end{pmatrix}\mathrm {e}^{-\frac{\pi \mathrm {i}}{4}\sigma _3},&\quad&\zeta \in \mathbb {C}\setminus Y, \end{aligned}$$
(B.3a)
$$\begin{aligned} m_{\mathrm {as},N}^{\mathrm {Ai},\mathrm {inv}}(\zeta )&:=\sqrt{\pi }\mathrm {e}^{-\frac{\mathrm {i}\pi }{12}}\sum _{k=0}^N\frac{1}{\left( \frac{2}{3}\zeta ^{3/2}\right) ^k}\mathrm {e}^{\frac{\pi \mathrm {i}}{4}\sigma _3}\begin{pmatrix}\nu _k &{} -u_k\\ (-1)^k\nu _k &{}(-1)^k u_k\end{pmatrix}\zeta ^{\frac{1}{4}\sigma _3},&\zeta \in \mathbb {C}\setminus Y, \end{aligned}$$
(B.3b)

where the real constants \(\lbrace u_j,\nu _j\rbrace _0^{\infty }\) are defined by \(u_0:=\nu _0:=1\) and

$$\begin{aligned} u_k:=\frac{(2k+1)(2k+3)\dots (6k-1)}{(216)^kk!}\,,\quad \nu _k:=\frac{6k+1}{1-6k}u_k,\quad k=1,2,\dots \end{aligned}$$

For \(N=0\), we have

$$\begin{aligned} m_{\mathrm {as},0}^{\mathrm {Ai}}(\zeta )=m_{\mathrm {as},0}^{\mathrm {Ai},\mathrm {inv}}(\zeta )^{-1}=\frac{\mathrm {e}^{\frac{\mathrm {i}\pi }{12}}}{2\sqrt{\pi }}\begin{pmatrix}\zeta ^{-\frac{1}{4}}&{}0\\ 0&{}\zeta ^{\frac{1}{4}}\end{pmatrix}\begin{pmatrix}1&{}1\\ -1&{}1\end{pmatrix}\mathrm {e}^{-\frac{\pi \mathrm {i}}{4}\sigma _3}. \end{aligned}$$

Theorem B.1

(a) The function \(m^{\mathrm {Ai}}(\zeta )\) defined in (B.2) is analytic for \(\zeta \in \mathbb {C}\setminus Y\) and satisfies the jump condition

$$\begin{aligned} m_+^{\mathrm {Ai}}(\zeta )=m_-^{\mathrm {Ai}}(\zeta )v^{\mathrm {Ai}}(\zeta ),\quad \zeta \in Y\setminus \lbrace 0\rbrace , \end{aligned}$$
(B.4)

where the jump matrix \(v^{\mathrm {Ai}}\) is defined by

$$\begin{aligned} v^{\mathrm {Ai}}(\zeta ):={\left\{ \begin{array}{ll} \begin{pmatrix}1&{}-\mathrm {e}^{-\frac{4}{3}\zeta ^{3/2}}\\ 0&{}1\end{pmatrix},&{}\zeta \in Y_1,\\ \begin{pmatrix}1&{}0\\ \mathrm {e}^{\frac{4}{3}\zeta ^{3/2}}&{}1\end{pmatrix},&{}\zeta \in Y_2\cup Y_4,\\ \begin{pmatrix}0&{}1\\ -1&{}0\end{pmatrix},&\zeta \in Y_3. \end{array}\right. } \end{aligned}$$

(b) For each integer \(N\ge 0\), the functions \(m_{\mathrm {as},N}^{\mathrm {Ai}}(\zeta )\) and \(m_{\mathrm {as},N}^{\mathrm {Ai},\mathrm {inv}}(\zeta )\) are analytic for \(\zeta \in \mathbb {C}\setminus (-\infty ,0]\) and satisfy the following jump relations on the negative real axis:

$$\begin{aligned} m_{\mathrm {as},N+}^{\mathrm {Ai}}(\zeta )&=m_{\mathrm {as},N-}^{\mathrm {Ai}}(\zeta )\begin{pmatrix}0&{}1\\ -1&{}0\end{pmatrix},&\quad&\zeta <0, \end{aligned}$$
(B.5a)
$$\begin{aligned} m_{\mathrm {as},N+}^{\mathrm {Ai},\mathrm {inv}}(\zeta )&=\begin{pmatrix}0&{}-1\\ 1&{}0\end{pmatrix}m_{\mathrm {as},N-}^{\mathrm {Ai},\mathrm {inv}}(\zeta ),&\zeta <0. \end{aligned}$$
(B.5b)

(c) The functions \(m_{\mathrm {as},N}^{\mathrm {Ai}}\) and \(m_{\mathrm {as},N}^{\mathrm {Ai},\mathrm {inv}}\) approximate \(m^{\mathrm {Ai}}\) and its inverse as \(\zeta \rightarrow \infty \) in the sense that

$$\begin{aligned} m^{\mathrm {Ai}}(\zeta )^{-1} m_{\mathrm {as},N}^{\mathrm {Ai}}(\zeta )&=I+\mathrm {O}(\zeta ^{-\frac{3(N+1)}{2}}),&\quad&\zeta \rightarrow \infty , \end{aligned}$$
(B.6a)
$$\begin{aligned} m_{\mathrm {as},N}^{\mathrm {Ai},\mathrm {inv}}(\zeta )m^{\mathrm {Ai}}(\zeta )&=I+\mathrm {O}(\zeta ^{-\frac{3(N+1)}{2}}),&\zeta \rightarrow \infty , \end{aligned}$$
(B.6b)

where the error terms are uniform with respect to \(\arg \zeta \in [0,2\pi ]\).

Proof

The analyticity of \(m^{\mathrm {Ai}}\) is a direct consequence of the Airy function \(\mathrm {Ai}(\zeta )\) being entire. The jump condition (B.4) can be verified by means of the identity

$$\begin{aligned} \mathrm {Ai}(\zeta )+\omega \mathrm {Ai}(\omega \zeta )+\omega ^2\mathrm {Ai}(\omega ^2\zeta )=0,\quad \zeta \in \mathbb {C}. \end{aligned}$$

On the other hand, the analyticity of \(m_{\mathrm {as},N}^{\mathrm {Ai}}(\zeta )\) and \(m_{\mathrm {as},N}^{\mathrm {Ai},\mathrm {inv}}(\zeta )\) is immediate from (B.3). Using that \((\zeta ^{3/2})_{\pm }=\mp \mathrm {i}|\zeta |^{3/2}\) and \((\zeta ^{1/4})_{\pm }=\mathrm {e}^{\pm \frac{\pi \mathrm {i}}{4}}|\zeta |^{1/4}\) for \(\zeta \in (-\infty ,0)\), a straightforward computation gives (B.5). This proves (a) and (b).

In order to prove (c), we note that, for each small number \(\delta >0\), the Airy function satisfies the following asymptotic expansions uniformly in the stated sectors, see [39]:

$$\begin{aligned} {\left\{ \begin{array}{ll} \mathrm {Ai}(z)=\frac{\mathrm {e}^{-\frac{2}{3}z^{3/2}}}{2\sqrt{\pi }z^{1/4}}\sum \limits _{k=0}^{\infty }\frac{(-1)^ku_k}{\left( \frac{2}{3}z^{3/2}\right) ^k},&{}\\ \mathrm {Ai}'(z)=-\frac{z^{1/4}\mathrm {e}^{-\frac{2}{3}z^{3/2}}}{2\sqrt{\pi }}\sum \limits _{k=0}^{\infty }\frac{(-1)^k(6k+1)u_k}{(1-6k)\left( \frac{2}{3}z^{3/2}\right) ^k},&{} \end{array}\right. }\qquad z\rightarrow \infty ,\quad |\arg z|\le \pi -\delta , \end{aligned}$$
(B.7)

and

$$\begin{aligned}&{\left\{ \begin{array}{ll} \mathrm {Ai}(z)=\frac{(-z)^{-1/4}}{\sqrt{\pi }}\sum \limits _{k=0}^{\infty }(-1)^k\left( \frac{u_{2k}\cos \left( \frac{2}{3}(-z)^{3/2}-\frac{\pi }{4}\right) }{\left( \frac{2}{3}(-z)^{3/2}\right) ^{2k}}+\frac{u_{2k+1}\sin \left( \frac{2}{3}(-z)^{3/2}-\frac{\pi }{4}\right) }{\left( \frac{2}{3}(-z)^{3/2}\right) ^{2k+1}}\right) ,&{}\\ \mathrm {Ai}'(z)=\frac{(-z)^{1/4}}{\sqrt{\pi }}\sum \limits _{k=0}^{\infty }(-1)^k\left( \frac{\nu _{2k}\sin \left( \frac{2}{3}(-z)^{3/2}-\frac{\pi }{4}\right) }{\left( \frac{2}{3}(-z)^{3/2}\right) ^{2k}}-\frac{\nu _{2k+1}\cos \left( \frac{2}{3}(-z)^{3/2}-\frac{\pi }{4}\right) }{\left( \frac{2}{3}(-z)^{3/2}\right) ^{2k+1}}\right) ,&{} \end{array}\right. }\nonumber \\& z\rightarrow \infty ,\quad \frac{\pi }{3}+\delta \le \arg z\le \frac{5\pi }{3}-\delta . \end{aligned}$$
(B.8)

The function \(m_{\mathrm {as},N}^{\mathrm {Ai}}(\zeta )\) is defined by the expression obtained by substituting the asymptotic sums in (B.7) from \(k=0\) to \(k=N\) into the definition of \(m^{\mathrm {Ai}}(\zeta )\) for \(\zeta \in S_1\). Similarly, \(m_{\mathrm {as},N}^{\mathrm {Ai},\mathrm {inv}}(\zeta )\) is defined by the expression obtained by substituting the asymptotic sums in (B.7) from \(k=0\) to \(k=N\) into the following expression for \(m^{\mathrm {Ai}}(\zeta )^{-1}\) which is valid for \(\zeta \in S_1\):

$$\begin{aligned} m^{\mathrm {Ai}}(\zeta )^{-1}=2\pi \mathrm {i}\omega ^2\mathrm {e}^{-\frac{2}{3}\zeta ^{3/2} \sigma _3}\mathrm {e}^{\frac{\pi \mathrm {i}}{6}\sigma _3} \begin{pmatrix}\omega ^2 \mathrm {Ai}'(\omega ^2\zeta )&{}-\mathrm {Ai}(\omega ^2\zeta )\\ -\mathrm {Ai}'(\zeta ) &{} \mathrm {Ai}(\zeta )\end{pmatrix}. \end{aligned}$$

It follows that the estimates in (B.6) hold as \(\zeta \rightarrow \infty \) in \({{\bar{S}}}_1\). Long but straightforward computations using (B.7) and (B.8) show that the estimates in (B.6) hold as \(\zeta \rightarrow \infty \) in \(\bar{S}_2\cup {{\bar{S}}}_3\cup {{\bar{S}}}_4\) as well. \(\square \)

Appendix C. Endpoint Behavior of a Cauchy Integral

Let \(\gamma :[\alpha ,\beta ]\rightarrow \mathbb {C}\) be a smooth simple contour from \(a=\gamma (\alpha )\) to \(b=\gamma (\beta )\) with \(a\ne b\). By a slight abuse of notation, we let \(\gamma \) denote both the map \([\alpha ,\beta ]\rightarrow \mathbb {C}\) and its image \(\gamma :=\gamma ([\alpha ,\beta ])\) as a subset of \(\mathbb {C}\).

Lemma C.1

Let \(f:\gamma \rightarrow \mathbb {C}\) be such that \(t\mapsto f(\gamma (t))\) is \(C^1\) on \([\alpha ,\beta ]\). Suppose g(k) is a \(C^1\)-function of \(k\in \mathbb {C}\) and define the function h(k) by

$$\begin{aligned} h(k):=\frac{g(k)}{2\pi \mathrm {i}}\int _{\gamma }\frac{f(s)\mathrm {d}s}{s-k},\quad k\in \mathbb {C}\setminus \gamma . \end{aligned}$$

Then the functions \(\mathrm {e}^{\mathrm {i}h(k)}\) and \(\mathrm {e}^{-\mathrm {i}h(k)}\) are bounded as \(k\in \mathbb {C}\setminus \gamma \) approaches b if and only if the product g(b)f(b) is purely imaginary.

Proof

We know (see [37, Appendix A2]) that the endpoint behavior of the Cauchy integral (here near \(k=b\)) is given by

$$\begin{aligned} \frac{1}{2\pi \mathrm {i}}\int _a^b\frac{f(s)}{s-k}\,\mathrm {d}s=\frac{f(b)}{2\pi \mathrm {i}}\ln (k-b)+\Phi _b(k), \end{aligned}$$

where \(\Phi _b(k)\) is bounded near \(k=b\). We thus have \(h(k)=\frac{g(k)}{2\pi \mathrm {i}}f(b)\ln (k-b)+\Psi _b(k)\) where \(\Psi _b(k)\) is bounded near \(k=b\). Moreover,

$$\begin{aligned} \frac{g(k)}{2\pi \mathrm {i}}f(b)\ln (k-b)=\frac{g(k)-g(b)}{2\pi \mathrm {i}}f(b)\ln (k-b)+\frac{g(b)}{2\pi \mathrm {i}}f(b)\ln (k-b), \end{aligned}$$

where the first term on the right-hand side is clearly bounded as \(k \rightarrow b\). The term \(\frac{g(b)}{2\pi \mathrm {i}}f(b)\ln (k-b)\) is unbounded as \(k\rightarrow b\), but its contribution to \(\mathrm {e}^{\pm \mathrm {i}h(k)}\) is bounded iff the product g(b)f(b) is purely imaginary. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Monvel, A.B.d., Lenells, J. & Shepelsky, D. The Focusing NLS Equation with Step-Like Oscillating Background: The Genus 3 Sector. Commun. Math. Phys. 390, 1081–1148 (2022). https://doi.org/10.1007/s00220-021-04288-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00220-021-04288-4

Navigation