Log in

Multi solitary waves to stochastic nonlinear Schrödinger equations

  • Published:
Probability Theory and Related Fields Aims and scope Submit manuscript

Abstract

In this paper, we present a pathwise construction of multi-soliton solutions for focusing stochastic nonlinear Schrödinger equations with linear multiplicative noise, in both the \(L^{2}\)-critical and subcritical cases. The constructed multi-solitons behave asymptotically as a sum of K solitary waves, where K is any given finite number. Moreover, the convergence rate of the remainders can be of either exponential or polynomial type, which reflects the effects of the noise in the system on the asymptotical behavior of the solutions. The major difficulty in our construction of stochastic multi-solitons is the absence of pseudo-conformal invariance. Unlike in the deterministic case (Merle in Commun Math Phys 129:223–240, 1990; Röckner et al. in Multi-bubble Bourgain–Wang solutions to nonlinear Schrödinger equation, ar**v: 2110.04107, 2021), the existence of stochastic multi-solitons cannot be obtained from that of stochastic multi-bubble blow-up solutions in Röckner et al. (Multi-bubble Bourgain–Wang solutions to nonlinear Schrödinger equation, ar**v:2110.04107, 2021), Su and Zhang (On the multi-bubble blow-up solutions to rough nonlinear Schrödinger equations, ar**v:2012.14037v1, 2020). Our proof is mainly based on the rescaling approach in Herr et al. (Commun Math Phys 368:843–884, 2019), relying on two types of Doss–Sussman transforms, and on the modulation method in Côte and Friederich (Commun Partial Differ Equ 46:2325–2385, 2021), Martel and Merle (Ann Inst H Poincaré Anal Non Linéaire 23:849–864, 2006), in which the crucial ingredient is the monotonicity of the Lyapunov type functional constructed by Martel et al. (Duke Math J 133:405-466, 2006). In our stochastic case, this functional depends on the Brownian paths in the noise.

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

Access this article

Subscribe and save

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

Buy Now

Price includes VAT (Germany)

Instant access to the full article PDF.

Similar content being viewed by others

Data Availability

Data sharing not applicable to this article as no datasets were generated or analysed during the current study.

References

  1. Albeverio, S., Brzeźniak, Z., Daletskii, A.: Stochastic Camassa–Holm equation with convection type noise. J. Differ. Equ. 276, 404–432 (2021)

    MathSciNet  MATH  Google Scholar 

  2. Bang, O., Christiansen, P.L., If, F., Rasmussen, K.O.: Temperature effects in a nonlinear model of monolayer Scheibe aggregates. Phys. Rev. E 49, 4627–4636 (1994)

    Google Scholar 

  3. Bang, O., Christiansen, P.L., If, F., Rasmussen, K.O., Gaididei, Y.B.: White noise in the two-dimensional nonlinear Schrödinger equation. Appl. Anal. 57(1–2), 3–15 (1995)

    MathSciNet  MATH  Google Scholar 

  4. Barbu, V., Da Prato, G., Röckner, M.: Stochastic porous media equations and self-organized criticality. Commun. Math. Phys. 285(3), 901–923 (2009)

    MathSciNet  MATH  Google Scholar 

  5. Barbu, V., Röckner, M., Zhang, D.: The stochastic nonlinear Schrödinger equation with multiplicative noise: the rescaling approach. J. Nonlinear Sci. 24, 383–409 (2014)

    MathSciNet  MATH  Google Scholar 

  6. Barbu, V., Röckner, M., Zhang, D.: Stochastic nonlinear Schrödinger equations. Nonlinear Anal. 136, 168–194 (2016)

    MathSciNet  MATH  Google Scholar 

  7. Barbu, V., Röckner, M., Zhang, D.: The stochastic logarithmic Schrödinger equation. J. Math. Pures Appl. (9) 107(2), 123–149 (2017)

    MathSciNet  MATH  Google Scholar 

  8. Barbu, V., Röckner, M., Zhang, D.: Optimal bilinear control of nonlinear stochastic Schrödinger equations driven by linear multiplicative noise. Ann. Probab. 46(4), 1957–1999 (2018)

    MathSciNet  MATH  Google Scholar 

  9. Barchielli, A., Gregoratti, M.: Quantum Trajectories and Measurements in Continuous Case. The Diffusive Case. Lecture Notes Physics, vol. 782. Springer Verlag, Berlin (2009)

    MATH  Google Scholar 

  10. Berestycki, H., Lions, P.: Nonlinear scalar field equations, I. Existence of a ground state. Arch. Ration. Mech. Anal 82(4), 313–345 (1983)

    MathSciNet  MATH  Google Scholar 

  11. Brzeźniak, Z., Millet, A.: On the stochastic Strichartz estimates and the stochastic nonlinear Schrödinger equation on a compact Riemannian manifold. Potential Anal. 41(2), 269–315 (2014)

    MathSciNet  MATH  Google Scholar 

  12. Cao, D., Su, Y., Zhang, D.: On uniqueness of multi-bubble blow-up solutions and multi-solitons to \(L^2\)-critical nonlinear Schrödinger equations. Arch. Ration. Mech. Anal. 247, 4 (2023)

    MATH  Google Scholar 

  13. Cazenave, T.: Semilinear Schrödinger Equations. Courant Lecture Notes in Mathematics, 10. New York University, Courant Institute of Mathematical Sciences, New York; American Mathematical Society, Providence, RI, p. xiv+323 (2003)

  14. Chen, C., Hong, J., Ji, L., Kong, L.: A compact scheme for coupled stochastic nonlinear Schrödinger equations. Commun. Comput. Phys. 21(1), 93–125 (2017)

    MathSciNet  MATH  Google Scholar 

  15. Combet, V.: Multi-soliton solutions for the supercritical gKdV equations. Commun. Partial Differ. Equ. 36(3), 380–419 (2011)

    MathSciNet  MATH  Google Scholar 

  16. Côte, R., Friederich, X.: On smoothness and uniqueness of multi-solitons of the non-linear Schrödinger equations. Commun. Partial Differ. Equ. 46(12), 2325–2385 (2021)

    MATH  Google Scholar 

  17. Côte, R., Martel, Y., Merle, F.: Construction of multi-soliton solutions for the \(L^2\)-supercritical gKdV and NLS equations. Rev. Mat. Iberoam. 27(1), 273–302 (2011)

    MathSciNet  MATH  Google Scholar 

  18. Côte, R., Muñoz, C.: Multi-solitons for nonlinear Klein–Gordon equations. Forum Math. Sigma 2(e15), 38 (2014)

    MathSciNet  MATH  Google Scholar 

  19. de Bouard, A., Debusche, A.: The stochastic nonlinear Schrödinger equation in \(H^1\). Stoch. Anal. Appl. 21, 97–126 (2003)

    Google Scholar 

  20. de Bouard, A., Debussche, A.: Random modulation of solitons for the stochastic Korteweg-de Vries equation. Ann. Inst. H. Poincaré Anal. Non Linéaire 24(2), 251–278 (2007)

    MathSciNet  MATH  Google Scholar 

  21. de Bouard, A., Debussche, A.: Soliton dynamics for the Korteweg-de Vries equation with multiplicative homogeneous noise. Electron. J. Probab. 14(58), 1727–1744 (2009)

    MathSciNet  MATH  Google Scholar 

  22. de Bouard, A., Debussche, A., Menza, L.D.: Theoretical and numerical aspects of stochastic nonlinear Schrödinger equations, Journées "Équations aux Dérivées Partielles" (Plestin-les-Grèves, 2001), Exp. No. III, p 13 , Univ. Nantes, Nantes, (2001)

  23. de Bouard, A., Fukuizumi, R.: Modulation analysis for a stochastic NLS equation arising in Bose–Einstein condensation. Asymptot. Anal. 63(4), 189–235 (2009)

    MathSciNet  MATH  Google Scholar 

  24. de Bouard, A., Fukuizumi, R.: Representation formula for stochastic Schrödinger evolution equations and applications. Nonlinearity 25(11), 2993–3022 (2012)

    MathSciNet  MATH  Google Scholar 

  25. de Bouard, A., Gautier, E.: Exit problems related to the persistence of solitons for the Korteweg-de Vries equation with small noise. Discrete Contin. Dyn. Syst. 26(3), 857–871 (2010)

    MathSciNet  MATH  Google Scholar 

  26. Debussche, A., Gautier, E.: Small noise asymptotic of the timing jitter in soliton transmission. Ann. Appl. Probab. 18(1), 178–208 (2008)

    MathSciNet  MATH  Google Scholar 

  27. Debussche, A., Menza, L.D.: Numerical simulation of focusing stochastic nonlinear Schrödinger equations. Phys. D 162(3–4), 131–154 (2002)

    MathSciNet  MATH  Google Scholar 

  28. Dodson, B.: Global well-posedness and scattering for the defocusing, \(L^2\)-critical nonlinear Schrödinger equation when \(d\ge 3\). J. Am. Math. Soc. 25(2), 429–463 (2012)

    MATH  Google Scholar 

  29. Dodson, B.: Global well-posedness and scattering for the mass critical nonlinear Schrödinger equation with mass below the mass of the ground state. Adv. Math. 285, 1589–1618 (2015)

    MathSciNet  MATH  Google Scholar 

  30. Dodson, B.: Global well-posedness and scattering for the defocusing, L2 critical, nonlinear Schrödinger equation when \(d=1\). Am. J. Math. 138(2), 531–569 (2016)

    MATH  Google Scholar 

  31. Dodson, B.: Global well-posedness and scattering for the defocusing, L2-critical, nonlinear Schrödinger equation when d=2. Duke Math. J. 165(18), 3435–3516 (2016)

    MathSciNet  MATH  Google Scholar 

  32. Dodson, B.: A determination of the blowup solutions to the focusing NLS with mass equal to the mass of the soliton \(d=1\). ar**v:2104.11690

  33. Dodson, B.: A determination of the blowup solutions to the focusing NLS with mass equal to the mass of the soliton. ar**v:2106.02723

  34. Fan, C.J., Su, Y., Zhang, D.: A note on log-log blow up solutions for stochastic nonlinear Schrödinger equations. Stoch. Partial Differ. Equ. Anal. Comput. 10(4), 1500–1514 (2022)

    MathSciNet  MATH  Google Scholar 

  35. Fan, C.J., Xu, W.J.: Global well-posedness for the defocusing mass-critical stochastic nonlinear Schrödinger equation on \({\mathbb{R} }\) at \(L^2\) regularity. Anal. PDE 14(8), 2561–2594 (2021)

    MathSciNet  MATH  Google Scholar 

  36. Fan, C.J., Xu, W.J.: Subcritical approximations to stochastic defocusing mass-critical nonlinear Schrödinger equation on \({\mathbb{R} }\). J. Differ. Equ. 268(1), 160–185 (2019)

    MathSciNet  MATH  Google Scholar 

  37. Fan, C.J., Xu, W.J., Zhao, Z.H.: Long time behavior of stochastic NLS with a small multiplicative noise, ar**v: 2111.07212v1

  38. Fan, C.J., Zhao, Z.H.: On long time behavior for stochastic nonlinear Schrödinger equations with a multiplicative noise, ar**v: 2010.11045v1

  39. Friz, P., Hairer, M.: A Course on Rough Paths. With an Introduction to Regularity Structures Universitext, p. xiv+251. Springer, Cham (2014)

    MATH  Google Scholar 

  40. Gubinelli, M.: Controlling rough paths. J. Funct. Anal. 216(1), 86–140 (2004)

    MathSciNet  MATH  Google Scholar 

  41. Herr, S., Röckner, M., Zhang, D.: Scattering for stochastic nonlinear Schrödinger equations. Commun. Math. Phys. 368(2), 843–884 (2019)

    MATH  Google Scholar 

  42. Karatzas, I., Shreve, S.E.: Brownian Motion and Stochastic Calculus. Graduate Texts in Mathematics, vol. 113, 2nd edn., p. xxiv+470. Springer-Verlag, New York (1991)

    MATH  Google Scholar 

  43. Krieger, J., Martel, Y., Raphaël, P.: Two-soliton solutions to the three-dimensional gravitational Hartree equation. Commun. Pure Appl. Math. 62(11), 1501–1550 (2009)

    MathSciNet  MATH  Google Scholar 

  44. Lan, Y.: On asymptotic dynamics for \(L^2\) critical generalized KdV equations with a saturated perturbation. Anal. PDE 12(1), 43–112 (2019)

    MathSciNet  MATH  Google Scholar 

  45. Martel, Y.: Asymptotic N-soliton-like solutions of the subcritical and critical generalized Korteweg-de Vries equations. Am. J. Math. 127(5), 1103–1140 (2005)

    MathSciNet  MATH  Google Scholar 

  46. Martel, Y., Merle, F.: Multi solitary waves for nonlinear Schrödinger equations. Ann. Inst. H. Poincaré Anal. Non Linéaire 23(6), 849–864 (2006)

    MathSciNet  MATH  Google Scholar 

  47. Martel, Y., Merle, F., Tsai, T.P.: Stability in \(H^1\) of the sum of \(K\) solitary waves for some nonlinear Schrödinger equations. Duke Math. J. 133(3), 405–466 (2006)

    MathSciNet  MATH  Google Scholar 

  48. Marzuola, J., Metcalfe, J., Tataru, D.: Strichartz estimates and local smoothing estimates for asymptotically flat Schrödinger equations. J. Funct. Anal. 255(6), 1479–1553 (2008)

    MATH  Google Scholar 

  49. Merle, F.: Construction of solutions with exactly k blow-up points for the Schrödinger equation with critical nonlinearity. Commun. Math. Phys. 129(2), 223–240 (1990)

    MathSciNet  MATH  Google Scholar 

  50. Merle, F.: Determination of blow-up solutions with minimal mass for nonlinear Schrödinger equations with critical power. Duke Math. J. 69(2), 427–454 (1993)

    MathSciNet  MATH  Google Scholar 

  51. Merle, F., Raphaël, P.: Profiles and quantization of the blow up mass for critical nonlinear Schrödinger equation. Commun. Math. Phys. 253(3), 675–704 (2005)

    MATH  Google Scholar 

  52. Millet, A., Rodriguez, A.D., Roudenko, S., Yang, K.: Behavior of solutions to the 1d focusing stochastic nonlinear Schrödinger equation with spatially correlated noise. Stoch. Partial Differ. Equ. Anal. Comput. 9(4), 1031–1080 (2021)

    MathSciNet  MATH  Google Scholar 

  53. Millet, A., Roudenko, S., Yang, K.: Behavior of solutions to the 1d focusing stochastic \(L^2\)-critical and supercritical nonlinear Schrödinger equation with space-time white noise. IMA J. Appl. Math. 86(6), 1349–1396 (2021)

    MathSciNet  MATH  Google Scholar 

  54. Ming, M., Rousset, F., Tzvetkov, N.: Multi-solitons and related solutions for the water-waves system. SIAM J. Math. Anal. 47(1), 897–954 (2015)

    MathSciNet  MATH  Google Scholar 

  55. Rasmussen, K.O., Gaididei, Y.B., Bang, O., Chrisiansen, P.L.: The influence of noise on critical collapse in the nonlinear Schrödinger equation. Phys. Lett. A 204, 121–127 (1995)

    Google Scholar 

  56. Röckner, M., Su, Y., Zhang, D.: Multi-bubble Bourgain–Wang solutions to nonlinear Schrödinger equation (2021). ar**v: 2110.04107

  57. Röckner, M., Zhu, R.C., Zhu, X.C.: A remark on global solutions to random 3D vorticity equations for small initial data. Discrete Contin. Dyn. Syst. Ser. B 24(8), 4021–4030 (2019)

    MathSciNet  MATH  Google Scholar 

  58. Su, Y., Zhang, D.: Minimal mass blow-up solutions to rough nonlinear Schrödinger equations. J. Funct. Anal. 284, 109796 (2023). https://doi.org/10.1016/j.jfa.2022.109796

    Article  MATH  Google Scholar 

  59. Su, Y., Zhang, D.: On the multi-bubble blow-up solutions to rough nonlinear Schrödinger equations (2020). ar**v: 2012.14037v1

  60. Tao, T., Visan, M., Zhang, X.Y.: The nonlinear Schrödinger equation with combined power-type nonlinearities. Commun. Partial Differ. Equ. 32(7–9), 1281–1343 (2007)

    MATH  Google Scholar 

  61. Weinstein, M.: Nonlinear Schrödinger equations and sharp interpolation estimates. Commun. Math. Phys. 87(4), 567–576 (1982/83)

  62. Zhang, D.: Optimal bilinear control of stochastic nonlinear Schrödinger equations: mass-(sub)critical case. Probab. Theory Relat. Fields 178(1–2), 69–120 (2020)

    MATH  Google Scholar 

  63. Zhang, D.: Strichartz and local smoothing estimates for stochastic dispersive equations with linear multiplicative noise. SIAM J. Math. Anal. 54(6), 5981–6017 (2022)

    MathSciNet  MATH  Google Scholar 

  64. Zhang, D.: Stochastic nonlinear Schrödinger equations in the defocusing mass and energy critical cases. ar**v:1811.00167v2, to appear in Ann. Appl. Probab

Download references

Acknowledgements

The authors would like to thank the anonymous referees for the careful reading and useful comments which helped to improve the paper. The authors also thank Daomin Cao for valuable discussions. M. Röckner and D. Zhang thank for the financial support by the Deutsche Forschungsgemeinschaft (DFG, German Science Foundation) through SFB 1283/2 2021-317210226 at Bielefeld University. D. Zhang is also grateful for the support by NSFC (No. 12271352) and Shanghai Rising-Star Program 21QA1404500.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Michael Röckner.

Additional information

Publisher's Note

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

Appendix

Appendix

1.1 Linearized operator

Let \(L=(L_+,L_-)\) be the linearized operator around the ground state state defined by

$$\begin{aligned} L_{+}:= -\Delta + I -(1+p)Q^{p}, \ \ L_{-}:= -\Delta +I -Q^{p}. \end{aligned}$$
(6.1)

For any complex valued \(H^1\) function, set \(f := f_1 + i f_2\) in terms of the real and imaginary parts and

$$\begin{aligned} (Lf,f) :=\int f_1L_+f_1dx+\int f_2L_-f_2dx. \end{aligned}$$
(6.2)

The crucial coercivity property of linearized operators in the subcritical and critical case are summarized below.

Lemma 6.1

([61], see also [47, Lemma 2.2]) Let \(1< p< 1+\frac{4}{d}\). Then, there exists \(C>0\) such that

$$\begin{aligned} (Lf,f)\ge C\Vert f\Vert _{H^1}^2 -\frac{1}{C}\left( \left( \int Qf_1dx\right) ^2+\left( \int Qf_2dx\right) ^2+\left( \int \nabla Qf_1dx\right) ^2\right) , \end{aligned}$$
(6.3)

where \(f=f_1+if_2\).

Lemma 6.2

( [16, Proposition 3.17]) Let \(p=1 + \frac{4}{d}\). Then, there exists \(C>0\) such that

$$\begin{aligned} (Lf,f)&\ge C\Vert f\Vert _{H^1}^2-\frac{1}{C}\nonumber \\&\quad \times \left( \left( \int Qf_1dx\right) ^2+\left( \int Qf_2dx\right) ^2+\left( \int \nabla Qf_1dx\right) ^2+\left( \int x\cdot \nabla Qf_1dx\right) ^2\right) , \end{aligned}$$
(6.4)

where \(f=f_1+if_2\).

1.2 Decoupling lemma

Lemma 6.3

(Decoupling lemma) Let \(\delta _0\) be as in (1.5). For every \(1\le k\le K\), let

$$\begin{aligned} G_{i,k}(t,x) = w_k^{-\frac{2}{p-1}} g_{i}(\frac{x-v_k t - {\alpha }_k}{w_k}),\ \ i=1,2, \end{aligned}$$
(6.5)

where \(1\le p\le 1+ \frac{4}{d}\), \(g_i \in C_b^2\) decays exponentially fast at infinity, i.e., for some \(C_1>0\),

$$\begin{aligned} |g_i(y)| \le C_1 e^{-\delta _0|y|},\ \ y\in {{\mathbb {R}}}^d,\ i=1,2, \end{aligned}$$
(6.6)

the parameters \(w_k>0\), \(v_k, {\alpha }_k\in {{\mathbb {R}}}^d\), satisfying that

$$\begin{aligned} w_k^{-1}+ w_k+|v_k|+|{\alpha }_k| \le C_2. \end{aligned}$$
(6.7)

Then, if \(v_j \not = v_k\), \(j\not =k\), we have that for any \(p_1, p_2 >0\),

$$\begin{aligned} \int | G_{1,j}(t)|^{p_1}|G_{2,k}(t)|^{p_2}dx \le Ce^{-\delta |v_j-v_k| t}, \end{aligned}$$
(6.8)

where \(C,\delta _2>0\) depend on \(\delta _0, C_i,p_i\), \(i=1,2\).

Proof

We use (6.5) and the change of variables to compute

$$\begin{aligned}&\int | G_{1,j}(t)|^{p_1}|G_{2,k}(t)|^{p_2}dx\nonumber \\&\quad = w_j^{d-\frac{2p_1}{p-1}} w_k^{-\frac{2p_2}{p-1}} \int |g_1|^{p_1}(y) |g_2|^{p_2}\left( \frac{w_j y + (v_j-v_k) t +({\alpha }_j- {\alpha }_k)}{w_k}\right) dy \nonumber \\&\quad = w_j^{d-\frac{2p_1}{p-1}} w_k^{-\frac{2p_2}{p-1}} \left( \int _{\Omega } + \int _{\Omega ^c}\right) |g_1|^{p_1}(y) |g_2|^{p_2} \left( \frac{w_j y + (v_j-v_k) t +({\alpha }_j- {\alpha }_k)}{w_k}\right) dy \nonumber \\&\quad =: I_1 + I_2. \end{aligned}$$
(6.9)

where \(\Omega := \{y \in {{\mathbb {R}}}^d: |y|\le \frac{1}{2w_j} |v_j - v_k|t\}\) and \(\Omega ^c= {{\mathbb {R}}}^d\setminus \Omega \). On one hand, by (6.7), for t large enough,

$$\begin{aligned} \big | {w_j y + (v_j-v_k) t +({\alpha }_j- {\alpha }_k)} \big | \ge \frac{1}{2} |v_j -v_k|t - |{\alpha }_j- {\alpha }_k| \ge \frac{1}{4} |v_j - v_k| t, \ \ y\in \Omega , \end{aligned}$$

which along with the exponential decay (6.6) and \(w_j, w_k \ge C_2^{-1}>0\) yields that

$$\begin{aligned} I_1 \le C e^{-\frac{\delta _1 p_2}{4w_k} |v_j-v_k|t} \int g_1^{p_1}(y) dy \le C e^{-\delta ' |v_j - v_k|t}, \end{aligned}$$
(6.10)

where \(C,\delta '>0\) depend on \(\delta _0, C_i, p_i\), \(i=1,2\).

On the other hand, using (6.6) again we infer that

$$\begin{aligned} |g_1(y)| \le C_1 e^{-\frac{\delta _0}{2w_j}|v_j-v_k|t} , \ \ y\in \Omega ^c, \end{aligned}$$

and thus

$$\begin{aligned} I_2 \le C e^{-\frac{\delta _0 p_1}{2w_j} |v_j-v_k|t} \int _{\Omega ^c} g_2^{p_2}\left( \frac{w_j y + (v_j-v_k) t +({\alpha }_j- {\alpha }_k)}{w_k}\right) dy \le C e^{-\delta '' |v_j - v_k|t},\nonumber \\ \end{aligned}$$
(6.11)

where where \(C,\delta _2>0\) depend on \(\delta _0, C_i, p_i\), \(i=1,2\).

Therefore, plugging (6.10) and (6.11) into (6.9) we obtain (6.8) and finish the proof. \(\square \)

1.3 Proof of Proposition 3.1

Below we present the proof of the geometrical decomposition in Proposition 3.1 in a fashion close to that of [12]. Given any \(L>0\), \(w_k^0\in {\mathbb {R}}^+\), \(x_k^0, v_k\in {\mathbb {R}}^d\), \({\theta }_k^0\in {\mathbb {R}}\), \(1\le k\le K\), set

$$\begin{aligned} R_L(x) : =\sum _{k=1}^{K}R_{k,L}(x) =\sum _{k=1}^{K}(w_k^0)^{-\frac{2}{p-1}}Q\left( \frac{x-v_kL-x_k^0}{w_{k}^0}\right) e^{i\left( \frac{1}{2}v_k\cdot x-\frac{1}{4}|v_k|^2L+(w_{k}^0)^{-2}L+{\theta }_{k}^0\right) }. \end{aligned}$$
(6.12)

Note that, if \(L=t\), then \(R_{k,L}=R_k\) with \(R_k\) given by (1.12).

Lemma 6.4

There exists a universal small constant \(\delta _*>0\) such that the following holds. For any \(0<r, L^{-1} <\delta _*\) and for any \(u\in H^{-1}({\mathbb {R}}^d)\) satisfying \(\Vert u-R_L\Vert _{H^{-1}}\le r\), there exist unique \(C^1\) functions \({\mathcal {P}}(u)=(\widetilde{\alpha }, \widetilde{\theta }, \widetilde{w}): H^{-1}\rightarrow {{\mathbb {X}}}^K\) such that u admits the decomposition

$$\begin{aligned} u= & {} \sum _{k=1}^{K}(\widetilde{w}_k w_k^0)^{-\frac{2}{p-1}}Q\left( \frac{x-v_kL-x_k^0-\widetilde{\alpha }_k}{\widetilde{w}_k w_{k}^0}\right) e^{i\left( \frac{1}{2}v_k\cdot x-\frac{1}{4}|v_k|^2L+(w_{k}^0)^{-2}L+{\theta }_{k}^0+\widetilde{\theta }_k\right) }\nonumber \\{} & {} \quad + {\varepsilon }_L \ \ (=: \sum _{k=1}^{K}\widetilde{R}_{k,L}+{\varepsilon }_L), \end{aligned}$$
(6.13)

and the following orthogonality conditions hold: for \(1\le k\le K\),

$$\begin{aligned} \begin{aligned}&\textrm{Re}\ _{H^{1}}\langle \nabla \widetilde{R}_{k,L}, {\varepsilon }_L\rangle _{H^{-1}}=0,\ \ \textrm{Im}\ _{H^{1}}\langle \widetilde{R}_{k,L}, {\varepsilon }_L \rangle _{H^{-1}}=0,\\&\textrm{Re}\ _{H^{1}}\langle \frac{d}{2}\widetilde{R}_{k,L}+y_k\cdot \nabla \widetilde{R}_{k,L} - \frac{i}{2}v_k\cdot y_k \widetilde{R}_{k,L}, {\varepsilon }_L \rangle _{H^{-1}}=0, \end{aligned} \end{aligned}$$
(6.14)

where \(y_k :=x-v_kL-x_k^0-\widetilde{\alpha }_k\). Moreover, there exists a universal constant \(C>0\) such that,

$$\begin{aligned} \Vert \varepsilon _L\Vert _{H^{-1}}+\sum _{k=1}^{K}(|\widetilde{\alpha }_k|+|\widetilde{\theta }_k|+|\widetilde{w}_k-1|)\le C\Vert u-R_L\Vert _{H^{-1}}. \end{aligned}$$
(6.15)

Proof

The proof proceeds in four steps.

Step 1. Set \(\widetilde{\mathcal {P}}_{0,k}: =(0,0,1) \in {\mathbb {X}}\) and \(\widetilde{\mathcal {P}}_0 = (\widetilde{\mathcal {P}}_{0,1}, \cdots , \widetilde{\mathcal {P}}_{0,K})\in {\mathbb {X}}^K\). Similarly, let \(\widetilde{\mathcal {P}}_k:= (\widetilde{\alpha }_k, \widetilde{\theta }_k,\widetilde{w}_k) \in {\mathbb {X}}\), \(\widetilde{\mathcal {P}}:= (\widetilde{\mathcal {P}}_1, \cdots , \widetilde{\mathcal {P}}_K) \in {\mathbb {X}}^K\). Let

$$\begin{aligned} \alpha _k : =\widetilde{\alpha }_k+x_k^0,\quad {\theta }_k:=\widetilde{\theta }_k+{\theta }_k^0,\quad w_k:=\widetilde{w}_kw_k^0. \end{aligned}$$
(6.16)

For any \(u_0\in H^1\), let \(B_\delta (u_0, \widetilde{\mathcal {P}}_0)\) denote the closed ball centered at \((u_0, \widetilde{\mathcal {P}}_0)\) of radius \(\delta \), i.e.,

$$\begin{aligned} B_\delta (u_0, \widetilde{\mathcal {P}}_0) :=\{(u, \widetilde{\mathcal {P}}) \in H^{-1}\times {\mathbb {X}}^K:\ \Vert u-u_0\Vert _{H^{-1}}\le \delta ,\ \ |\widetilde{\mathcal {P}}- \widetilde{\mathcal {P}}_0|\le \delta \}, \end{aligned}$$
(6.17)

where \(\delta \) is a small constant to be chosen later, and

$$\begin{aligned} |\widetilde{\mathcal {P}}-\widetilde{\mathcal {P}}_0|:= \sum \limits _{k=1}^K |\widetilde{\mathcal {P}}_k- \widetilde{\mathcal {P}}_{0,k}| =\sum _{k=1}^{K}(|\widetilde{\alpha }_k|+|\widetilde{\theta }_k|+|\widetilde{w}_k-1|). \end{aligned}$$
(6.18)

For \(1\le k\le K\), let

$$\begin{aligned}&f_{1,j}^k(u, \widetilde{\mathcal {P}}) :=\textrm{Re}\ _{H^{1}}\langle \partial _j \widetilde{R}_{k,L}, {\varepsilon }_L \rangle _{H^{-1}}, \ \ 1\le j\le d, \\&f_{2}^k(u,\widetilde{\mathcal {P}}) :=\textrm{Im}\ _{H^{1}} \langle \widetilde{R}_{k,L}, {\varepsilon }_L, \rangle _{H^{-1}}, \\&f_{3}^k(u, \widetilde{\mathcal {P}}) :=\textrm{Re}\ _{H^{1}}\langle \frac{2}{p-1}\widetilde{R}_{k,L}+y_k\cdot \nabla \widetilde{R}_{k,L} - \frac{i}{2}v_k\cdot y_k \widetilde{R}_{k,L}, {\varepsilon }_L \rangle _{H^{-1}}, \end{aligned}$$

where \(y_k\) is as in (6.14). Let \(F^k:= (f^k_{1,1}, \cdots , f^k_{1,d}, f^k_2, f^k_{3})\) and \(\frac{\partial F^k}{\partial \widetilde{\mathcal {P}}_j}\) denote the Jacobian matrix

$$\begin{aligned} \frac{\partial F^k}{\partial \widetilde{\mathcal {P}}_j}:=\left( \begin{array}{ccc} \frac{\partial f^k_{1,1}}{\partial \widetilde{\alpha }_{j,1}} &{}\cdots \ \frac{\partial f^k_{1,1}}{\partial \widetilde{\alpha }_{j,d}},\ \frac{\partial f^k_{1,1}}{\partial \widetilde{\theta }_j}, &{} \frac{\partial f^k_{1,1}}{\partial \widetilde{w}_j} \\ \vdots &{} &{} \vdots \\ \frac{\partial f^k_{3}}{\partial \widetilde{\alpha }_{j,1}} &{} \cdots \ \frac{\partial f^k_{3}}{\partial \widetilde{\alpha }_{j,d}},\ \frac{\partial f^k_{3}}{\partial \widetilde{\theta }_j},&{} \frac{\partial f^k_{3}}{\partial \widetilde{w}_j} \\ \end{array} \right) , \ \ 1\le j, k\le K, \end{aligned}$$
(6.19)

where \(\widetilde{\alpha }_j:=(\widetilde{\alpha }_{j,l}, 1\le l\le d)\in {{\mathbb {R}}}^d\). Similarly, let \(F:=(F^1, \cdots , F^K)\) and \(\frac{\partial F}{\partial \widetilde{\mathcal {P}}} := (\frac{\partial F^k}{\partial \widetilde{\mathcal {P}}_j})_{1\le j,k\le K}\).

Note that, by the definition (6.12) of \(R_L\), \( F^k(R_L, \widetilde{\mathcal {P}}_0)=0\), \(1\le k\le K\). Moreover, for any \((u, \widetilde{\mathcal {P}})\in B_\delta (R_L, \widetilde{\mathcal {P}}_0)\), we have that, if \(\widetilde{R}_L:= \sum _{k=1}^K \widetilde{R}_{k,L}\),

$$\begin{aligned} \Vert \varepsilon _L\Vert _{H^{-1}} \le \Vert u-R_L\Vert _{H^{-1}} + \Vert R_L - \widetilde{R}_L\Vert _{H^{-1}}. \end{aligned}$$
(6.20)

By the explicit expressions of \(R_L\) and \(\widetilde{R}_L\) in (6.12) and (6.13), respectively,

$$\begin{aligned} \Vert R_L - \widetilde{R}_L\Vert _{H^{-1}} \le \Vert R_L - \widetilde{R}_L\Vert _{L^2} \le C\sum _{k=1}^{K}(|\widetilde{\alpha }_k|+|\widetilde{\theta }_k|+|\widetilde{w}_k-1|)\le C |\widetilde{\mathcal {P}}- \widetilde{\mathcal {P}}_0|, \end{aligned}$$
(6.21)

where \(C>0\). Thus, we get that for a universal constant \(\widetilde{C}>0\),

$$\begin{aligned} \Vert \varepsilon _L\Vert _{H^{-1}}\le \widetilde{C} (\Vert u-R_L\Vert _{H^{-1}} + |\widetilde{\mathcal {P}}- \widetilde{\mathcal {P}}_0|) \le 2 \widetilde{C} \delta ,\ \ \forall (u, \widetilde{\mathcal {P}})\in B_\delta (R_L, \widetilde{\mathcal {P}}_0).\nonumber \\ \end{aligned}$$
(6.22)

Step 2. We claim that, there exist small constants \(\delta _*, c_1, c_2>0\) such that for any \(0<\delta , L^{-1} \le \delta _*\),

$$\begin{aligned} 0<c_1\le \bigg |\det \frac{\partial F}{\partial \widetilde{\mathcal {P}}}(u, \widetilde{\mathcal {P}}) \bigg | \le c_2<{\infty }, \ \ \forall (u, \widetilde{\mathcal {P}})\in B_{\delta }(R_L, \widetilde{\mathcal {P}}_0). \end{aligned}$$
(6.23)

To this end, we compute that for \(1\le j,k\le d\),

$$\begin{aligned} \begin{aligned}&\partial _{\widetilde{\alpha }_{k,j}} f^k_{1,j} = - w_k^{-2} \Vert \partial _j Q_{w_k}\Vert _{L^2}^2+{\mathcal {O}}(\Vert \varepsilon _L\Vert _{H^{-1}}), \\&\partial _{\widetilde{\theta }_k} f_{1,j}^k = - \frac{ v_{k,j}}{2}\Vert Q_{w_k}\Vert _{L^2}^2+{\mathcal {O}}(\Vert \varepsilon _L\Vert _{H^{-1}}) , \\&\partial _{\widetilde{\theta }_k} f^k_{2} = \Vert Q_{w_k}\Vert _{L^2}^2+{\mathcal {O}}(\Vert \varepsilon _L\Vert _{H^{-1}}), \ \ \partial _{\widetilde{w}_k} f_3^k = w_k ^{-1} \Vert \Lambda Q_{w_k}\Vert _{L^2}^2+{\mathcal {O}}(\Vert \varepsilon _L\Vert _{H^{-1}}). \end{aligned} \end{aligned}$$
(6.24)

Moreover, by the exponential decay of Q, we infer that, there exists \(\delta >0\) such that the other terms in the Jacobian matrices are of the order \({\mathcal {O}}(\Vert {\varepsilon }\Vert _{H^{-1}} + e^{-\delta L})\). This yields that

$$\begin{aligned} \bigg |\det \left( \frac{\partial F}{\partial \widetilde{\mathcal {P}}} \right) \bigg |&= \prod \limits _{k=1}^K \left( (w_k^0)^{-2d} \widetilde{w}_k^{-2d-1} \Vert Q_{w_k}\Vert _{L^2}^2 \Vert \Lambda Q_{w_k}\Vert _{L^2}^2 \prod \limits _{j=1}^d \Vert \partial _j Q_{w_k}\Vert _{L^2}^2 \right) \nonumber \\&\quad + {\mathcal {O}}\left( \Vert {\varepsilon }\Vert _{H^{-1}} + e^{-\delta L}\right) . \end{aligned}$$
(6.25)

Taking into account \(|\widetilde{\mathcal {P}}- \widetilde{\mathcal {P}}_0| \le \delta \) we obtain (6.23), as claimed.

Step 3. In this step, we claim that there exists a universal constant \(C_*(\ge 1)\) such that, for any \(0<\delta , L^{-1} \le \delta _*\) and any \((u_1, \widetilde{\mathcal {P}}(u_1)), (u_2, \widetilde{\mathcal {P}}(u_2))\in B_\delta (R_L, \widetilde{\mathcal {P}}_0 )\), if \(F(u_1, \widetilde{\mathcal {P}}(u_1))= F(u_2, \widetilde{\mathcal {P}}(u_2))=0\), then

$$\begin{aligned} |\widetilde{\mathcal {P}}(u_1)-\widetilde{\mathcal {P}}(u_2)|\le C_* \Vert u_1-u_2\Vert _{H^{-1}}. \end{aligned}$$
(6.26)

To this end, we infer that

$$\begin{aligned} F(u_1, \widetilde{\mathcal {P}}(u_1))-F(u_1, \widetilde{\mathcal {P}}(u_2))= F(u_2, \widetilde{\mathcal {P}}(u_2) ) - F(u_1, \widetilde{\mathcal {P}}(u_2)). \end{aligned}$$
(6.27)

By the differential mean value theorem,

$$\begin{aligned} \left( \frac{\partial F}{\partial \widetilde{\mathcal {P}}}(u_1, \widetilde{\mathcal {P}}_{r})\right) (\widetilde{\mathcal {P}}(u_1)-\widetilde{\mathcal {P}}(u_2))^t = (F(u_2, \widetilde{\mathcal {P}}(u_2))-F(u_1,\widetilde{\mathcal {P}}(u_2)))^t, \end{aligned}$$
(6.28)

where \(\widetilde{\mathcal {P}}_r=r\widetilde{\mathcal {P}}(u_1)+(1-r)\widetilde{\mathcal {P}}(u_2)\) for some \(0<r<1\), and the superscript t means the transpose of matrices. Since the Jacobian matrix \(\frac{\partial F}{\partial \widetilde{\mathcal {P}}}(u_1, \widetilde{\mathcal {P}}_r)\) is invertible by (6.23), this leads to

$$\begin{aligned} (\widetilde{\mathcal {P}}(u_1)-\widetilde{\mathcal {P}}(u_2))^t= \left( \frac{\partial F}{\partial \widetilde{\mathcal {P}}}(u_1,\widetilde{\mathcal {P}}_r) \right) ^{-1} (F(u_2, \widetilde{\mathcal {P}}(u_2)) - F(v_1, \widetilde{\mathcal {P}}(u_2)))^t. \end{aligned}$$
(6.29)

Note that, by (6.24), there exists a universal constant \(C>0\) such that

$$\begin{aligned} \bigg \Vert \bigg (\frac{\partial F}{\partial \widetilde{\mathcal {P}}}(u_1, \widetilde{\mathcal {P}}_r)\bigg )^{-1} \bigg \Vert \le C, \end{aligned}$$
(6.30)

where \(\Vert \cdot \Vert \) denotes the Hilbert-Schmidt norm of matrices. Moreover, by (1.5),

$$\begin{aligned}&|F(u_2, \widetilde{\mathcal {P}}(u_2)) - F(u_1, \widetilde{\mathcal {P}}(u_2))|\le C \Vert u_2-u_1\Vert _{H^{-1}}. \end{aligned}$$
(6.31)

Thus, we infer from (6.29), (6.30) and (6.31) that (6.26) holds, as claimed.

Step 4. Let \(\delta _*, C_*\) be the universal constants as in Step 1 and Step 2, respectively, and set

$$\begin{aligned} B:=\{ v\in B_{\frac{\delta _*}{C_*}}(R_L): \exists \widetilde{\mathcal {P}}\in B_{\delta _*}(\widetilde{\mathcal {P}}_0), \ such\ that\ F(v,\widetilde{\mathcal {P}}) =0\}. \end{aligned}$$
(6.32)

Since \(B_{\frac{\delta _*}{C_*}}(R_L)\) is connected and \(R_L \in B\), in order to prove that

$$\begin{aligned} B = B_{\frac{\delta _*}{C_*}}(R_L). \end{aligned}$$
(6.33)

we only need to show that B is both open and closed in \(B_{\frac{\delta _*}{C_*}}(R_L)\).

To this end, For any \(u\in B\), by definition there exists \(\widetilde{\mathcal {P}}(u) \in B_{\delta _*}(\widetilde{\mathcal {P}}_0)\) such that \(F(u,\widetilde{\mathcal {P}}(u))=0\). Taking into account the non-degeneracy of the Jacobian matrix at \((u, \widetilde{\mathcal {P}}(u))\) due to (6.23), we can apply the implicit function theorem to get a small open neighborhood \({\mathcal {U}}(u)\) of u in \(B_{\frac{\delta _*}{C_*}}(R_L)\) such that \({\mathcal {U}}(u) \subseteq B\). This yield that B is open in \(B_{\frac{\delta _*}{C_*}}(R_L)\).

Moreover, for any sequence \(\{u_n\} \subseteq B\) such that \(u_n \rightarrow u_*\) in \(H^{-1}\) for some \(u_*\in B_{\frac{\delta _*}{C_*}}(R_L)\), by definition there exist modulation parameters \(\widetilde{\mathcal {P}}(u_n)\in B_{\delta _*}(\widetilde{\mathcal {P}}_0)\) such that \(F(u_n, \widetilde{\mathcal {P}}(u_n))=0\), \(n \ge 1\). In particular, \(\{\widetilde{\mathcal {P}}(v_n)\} \subseteq {\mathbb {X}}^K\) is uniformly bounded and so, along a subsequence (still denoted by \(\{n\}\)), \(\widetilde{\mathcal {P}}(v_{n}) \rightarrow \widetilde{\mathcal {P}}_*\ (\in B_{\delta _*}(\widetilde{\mathcal {P}}_0))\) for some \(\widetilde{\mathcal {P}}_* \in {\mathbb {X}}^K\).

Then, let \(\widetilde{R}_{k,L, \widetilde{\mathcal {P}}(u_n)}\) and \(\widetilde{R}_{k,L,\widetilde{\mathcal {P}}_*}\) be the k-th soliton profiles corresponding to \(\widetilde{\mathcal {P}}(u_n)\) and \(\widetilde{\mathcal {P}}_*\), respectively. By the above convergence of \(u_n\) and \(\widetilde{\mathcal {P}}(u_n)\) we infer that \(u_{n} - \sum _{k=1}^K \widetilde{R}_{k,L, \widetilde{\mathcal {P}}(u_{n})} \rightarrow u_* - \sum _{k=1}^K \widetilde{R}_{k,L,\widetilde{\mathcal {P}}_*}\) in \(H^{-1}\). Taking \(n\rightarrow {\infty }\) and using the fact that \(F(u_{n}, \widetilde{\mathcal {P}}(u_{{n}})) =0\) we obtain \(F(u_*, \widetilde{\mathcal {P}}_*) = 0\), and so \(u_*\in B\). Hence, B is also closed in \(B_{\frac{\delta _*}{C_*}}(R_L)\).

Therefore, (6.33) is verified. The geometrical decomposition (6.13) and the orthogonality conditions in (6.14) hold. Moreover, estimate (6.15) follows from (6.22) and (6.26) by taking \(u_1= u\) and \(u_2 = R_L\). The proof of Lemma 6.4 is complete. \(\square \)

Proof of Proposition 3.1

Since \(u(T)=R(T)\), by the local wellposedness theory, there exists \(T^*\) close to T, such that \(u(t)\in C^1([T^*,T]; H^{-1}) \bigcap C([T^*,T];H^1)\) and \(\Vert u(t)-R(T)\Vert _{H^{1}} \in B_{\delta }(u(T))\) for all \(t\in [T^*,T]\), where \(\delta >0\) is as in Lemma 6.4.

Hence, applying Lemma 6.4 to \(\{u(t)\}\) we obtain that for T large enough, there exist unique \(C^1\) functions \((\alpha _k,{\theta }_k,\omega _k) \in C^1([T^*,T]; {\mathbb {X}}^K)\), \(1\le k\le K\), such that for any \(t\in [T^*,T]\), u(t) admits the decomposition (6.13) and the orthogonality conditions in (6.14) hold with t replacing T.

Then, taking into account \(u(t)\in H^1\) and (6.13), the remainder \({\varepsilon }(t)\) is indeed in the space \(H^1\). Thus, the parings between \(H^{-1}\) and \(H^1\) in (6.14) are exactly the \(L^2\) inner products, which yields the orthogonality conditions in (3.5) for any \([T^*,T]\). Therefore, the proof is complete. \(\square \)

1.4 Proof of (4.45)

We set \(\widetilde{S}_k:= \sum _{j=k}^K \widetilde{R}_j\), \(1\le k\le K\). Then,

$$\begin{aligned} \widetilde{S}_k = \widetilde{R}_k + \widetilde{S}_{k+1},\ \ 1\le k\le K-1. \end{aligned}$$
(6.34)

Lemma 6.5

Let \(0<q<{\infty }\), we have

$$\begin{aligned} \left| |\widetilde{S}_k |^q-|\widetilde{R}_k|^q\right| \le C h(\widetilde{S}_{k+1}), \end{aligned}$$
(6.35)

where \(C>0\), \(h(\widetilde{S}_{k+1}) = |\widetilde{S}_{k+1}|^q\) if \(0<q<1\), and \(h(\widetilde{S}_{k+1}) = |\widetilde{S}_{k+1}|\) if \(1\le q<{\infty }\).

Proof

The case where \(0<q<1\) follows from the inequality

$$\begin{aligned} (a+b)^{q}\le a^q+b^q, \ \ a,b \ge 0, \end{aligned}$$

while the case \(1\le q<{\infty }\) follows from the inequality

$$\begin{aligned} ||\widetilde{S}_k |^q-|\widetilde{R}_k|^q | \le C(|\widetilde{S}_{k+1}|^{q-1} + |\widetilde{R}_{k}|^{q-1}) |\widetilde{S}_{k+1}| \end{aligned}$$

and the uniform boundedness of \(\widetilde{R}_j\), \(1\le j\le K\). \(\square \)

Lemma 6.6

There exist constants \(C, \delta _2>0\) depending on \(w_k^0\), \(x_k^0\), \(v_k\) and \(\delta _0\) such that

$$\begin{aligned} |\int |\widetilde{S}_{k}|^{p+1}-|\widetilde{R}_k|^{p+1}-|\widetilde{S}_{k+1}|^{p+1}dx|\le Ce^{-\delta _2 t}. \end{aligned}$$
(6.36)

Proof

Using the expansion

$$\begin{aligned} |\widetilde{S}_k|^2 = |\widetilde{R}_k|^2 + |\widetilde{S}_{k+1}|^2 + 2 \textrm{Re} (\widetilde{R}_k \widetilde{S}_{k+1}), \end{aligned}$$

and Lemmas 6.3 and 6.5 we have

$$\begin{aligned}&\big | \int |\widetilde{S}_k|^{p+1}-|\widetilde{R}_k|^{p+1}-|\widetilde{S}_{k+1}|^{p+1}dx \big | \\&\le \int \big ||\widetilde{S}_k|^{p-1}-|\widetilde{R}_k|^{p-1}\big | |\widetilde{R}_k|^2 +\big ||\widetilde{S}_k|^{p-1}-|\widetilde{S}_{k+1}|^{p-1}\big ||\widetilde{S}_{k+1}|^2 +2|\widetilde{S}_{k}|^{p-1}|\widetilde{R}_k\widetilde{S}_{k+1}|dx \\&\le C \int h(\widetilde{S}_{k+1}) |\widetilde{R}_k|^2 + h(\widetilde{R}_k) |\widetilde{S}_{k+1}|^2 +2|\widetilde{S}_{k}|^{p-1} |\widetilde{R}_k \widetilde{S}_{k+1}|dx\\&\le Ce^{-\delta _2 t}, \end{aligned}$$

which yields (6.36). \(\square \)

Lemma 6.7

There exist constants \(C, \delta _2>0\) depending on \(w_k^0\), \(x_k^0\), \(v_k\) and \(\delta _0\) such that

$$\begin{aligned} |\int (| \widetilde{S}_{k}|^{p-1} \widetilde{S}_{k} - | \widetilde{R}_k|^{p-1} \widetilde{R}_k- | \widetilde{S}_{k+1}|^{p-1} \widetilde{S}_{k+1}) \overline{{\varepsilon }}dx| \le Ce^{-\delta _2 t} \Vert {\varepsilon }\Vert _{L^2}. \end{aligned}$$
(6.37)

Proof

By the expansion (6.34), Lemmas 6.3 and 6.5 and Hölder’s inequality,

$$\begin{aligned}&\big |\int (| \widetilde{S}_{k}|^{p-1} \widetilde{S}_{k} -| \widetilde{R}_k|^{p-1} \widetilde{R}_k-| \widetilde{S}_{k+1}|^{p-1} \widetilde{S}_{k+1})\overline{{\varepsilon }}dx\big |\\&\le \int \left( \big || \widetilde{S}_{k}|^{p-1} -| \widetilde{R}_k|^{p-1} \big ||\widetilde{R}_k| +\big || \widetilde{S}_{k}|^{p-1} -| \widetilde{S}_{k+1}|^{p-1} \big || \widetilde{S}_{k+1}|\right) |{\varepsilon }|dx\\&\le C (\Vert h(\widetilde{S}_{k+1}) \widetilde{R}_k\Vert _{L^2}+ \Vert h(\widetilde{R}_k) \widetilde{S}_{k+1}\Vert _{L^2}) \Vert {\varepsilon }\Vert _{L^2}\\&\le Ce^{-\delta t} \Vert {\varepsilon }\Vert _{L^2}, \end{aligned}$$

which yields (6.37). \(\square \)

Lemma 6.8

There exist constants \(C, \delta _2>0\) depending on \(w_k^0\), \(x_k^0\), \(v_k\) and \(\delta _0\) such that

$$\begin{aligned} |\int (| \widetilde{S}_{k}|^{p-1} -| \widetilde{R}_k|^{p-1}-| \widetilde{S}_{k+1}|^{p-1} ) |{\varepsilon }|^2dx|\le Ce^{-\delta _2 t} \Vert {\varepsilon }\Vert _{L^2}^2. \end{aligned}$$
(6.38)

Proof

Let \(\Omega _k:=\{ x: |x-v_kt|\le \frac{1}{2} \min _{j\not =k} |v_k-v_j|t \}\). By Lemma 6.5,

$$\begin{aligned}&\bigg |\int (| \widetilde{S}_k|^{p-1} -| \widetilde{R}_k|^{p-1}-| \widetilde{S}_{k+1}|^{p-1} ) |{\varepsilon }|^2dx\bigg | \nonumber \\&\le 2\int _{\Omega _k} (h(\widetilde{S}_{k+1}) + |\widetilde{S}_{k+1}|^{p-1}) |{\varepsilon }|^2dx + 2\int _{\Omega _k^c} (h(\widetilde{R}_k) + |\widetilde{R}_k|^{p-1}) |{\varepsilon }|^2dx \nonumber \\&\le C \Vert h(\widetilde{S}_{k+1}) + |\widetilde{S}_{k+1}|^{p-1} \Vert _{L^{\infty }(\Omega _k)} \Vert {\varepsilon }\Vert _{L^2}^2 + C \Vert h(\widetilde{R}_k) + |\widetilde{R}_k|^{p-1}\Vert _{L^{\infty }(\Omega _k^c)} \Vert {\varepsilon }\Vert _{L^2}^2. \end{aligned}$$
(6.39)

Note that, for \(x\in \Omega _k\), for any \(j\not =k\),

$$\begin{aligned} |x-v_j t-{\alpha }_j| \ge |v_j-v_k|t - |x- v_kt| - |{\alpha }_j| \ge \frac{1}{4} |v_j - v_k|t, \end{aligned}$$

and thus by the exponential decay of Q,

$$\begin{aligned} \Vert h(\widetilde{S}_{k+1}) + |\widetilde{S}_{k+1}|^{p-1}\Vert _{L^{\infty }(\Omega _k^c)} \le C e^{-\delta _2 t}. \end{aligned}$$
(6.40)

Similarly, for \(x\in \Omega _k^c\), there exists \(c>0\) such that for t large enough,

$$\begin{aligned} |x-v_kt - {\alpha }_k| \ge \frac{1}{2} \min _{j\not =k}\{|v_j-v_k|t\} - |{\alpha }_k| \ge ct, \end{aligned}$$

and thus

$$\begin{aligned} \Vert h(\widetilde{R}_k) + |\widetilde{R}_k|^{p-1}\Vert _{L^{\infty }(\Omega _k^c)} \le C e^{-\delta _2 t}. \end{aligned}$$
(6.41)

Therefore, plugging (6.40) and (6.41) into (6.39) we obtain (6.38) and finish the proof. \(\square \)

Lemma 6.9

There exist constants \(C, \delta _2>0\) depending on \(w_k^0\), \(x_k^0\), \(v_k\) and \(\delta _0\) such that

$$\begin{aligned} |\int (|\widetilde{S}_k|^{p-3}\widetilde{S}_k^2 -| \widetilde{R}_k|^{p-3}\widetilde{R}_k^2 -| \widetilde{S}_{k+1}|^{p-3}\widetilde{S}_{k+1}^2) \overline{{\varepsilon }}^2dx|\le Ce^{-\delta _2 t}. \end{aligned}$$
(6.42)

Proof

Since

$$\begin{aligned} |\widetilde{S}_k|^{p-1} \frac{\widetilde{S}_{k}^2}{|\widetilde{S}_k^2|} = |\widetilde{R}_k|^{p-1} \frac{\widetilde{S}_{k}^2}{|\widetilde{S}_k^2|} + |\widetilde{S}_{k+1}|^{p-1} \frac{\widetilde{S}_{k}^2}{|\widetilde{S}_k^2|} + {\mathcal {O}}( | |\widetilde{S}_k|^{p-1} - |\widetilde{R}_k|^{p-1} - |\widetilde{S}_{k+1}|^{p-1} |), \end{aligned}$$

we have

$$\begin{aligned}&\big |\int (|\widetilde{S}_k|^{p-3}\widetilde{S}_k^2 -| \widetilde{R}_k|^{p-3}\widetilde{R}_k^2 -| \widetilde{S}_{k+1}|^{p-3}\widetilde{S}_{k+1}^2) \overline{{\varepsilon }}^2dx\big | \nonumber \\ \le&\int \bigg ||\widetilde{S}_k|^{p-1}\frac{\widetilde{S}_k^2}{|\widetilde{S}_k|^2} -| \widetilde{R}_k|^{p-1}\frac{\widetilde{R}_k^2}{|\widetilde{R}_k|^2} -|\widetilde{S}_{k+1}|^{p-1}\frac{\widetilde{S}_{k+1}^2}{|\widetilde{S}_{k+1}|^2}\bigg | |{\varepsilon }|^2dx \nonumber \\ \le&\int | \widetilde{R}_k|^{p-1}\bigg |\frac{\widetilde{S}_k^2}{|\widetilde{S}_k|^2} -\frac{\widetilde{R}_k^2}{|\widetilde{R}_k|^2} \bigg | |{\varepsilon }|^2dx +\int | \widetilde{S}_{k+1}|^{p-1}\bigg |\frac{\widetilde{S}_k^2}{|\widetilde{S}_k|^2} -\frac{\widetilde{S}_{k+1}^2}{|\widetilde{S}_{k+1}|^2} \bigg | |{\varepsilon }|^2dx \nonumber \\&+ {\mathcal {O}}\left( \int \big ||\widetilde{S}_k|^{p-1}-| \widetilde{S}_k|^{p-1}-| \widetilde{R}_k|^{p-1}\big | |{\varepsilon }|^2dx\right) \nonumber \\ =&\int | \widetilde{R}_k|^{p-1}\bigg |\frac{\widetilde{S}_{k}^2}{|\widetilde{S}_k|^2} -\frac{\widetilde{R}_k^2}{|\widetilde{R}_k|^2} \bigg | |{\varepsilon }|^2dx +\int | \widetilde{S}_{k+1}|^{p-1}\bigg |\frac{\widetilde{S}_k^2}{|\widetilde{S}_k|^2} -\frac{\widetilde{S}_{k+1}^2}{|\widetilde{S}_{k+1} |^2} \bigg | |{\varepsilon }|^2dx+ {\mathcal {O}}(e^{-\delta _2 t}) \nonumber \\ =:&J_1 + J_2 + {\mathcal {O}}(e^{-\delta _2 t}). \end{aligned}$$
(6.43)

where the last step is due to Lemma 6.8.

Below we estimate \(J_1\) and \(J_2\) separately. For this purpose, let us set \(d_*:= \min _{k\le j\not =l\le K}\{|v_jt + {\alpha }_j - v_lt-{\alpha }_l|\}\). Similarly, let \(w_*:=\min _{k\le j\le K} w_j\), \(w^*:=\max _{k\le j\le K} w_j\). For every \(k\le j\le K\), set

$$\begin{aligned} \Omega _j:= \bigg \{x\in {{\mathbb {R}}}^d: |x-v_jt -{\alpha }_j| \le {\varepsilon }d_* \bigg \}, \end{aligned}$$

where \({\varepsilon }\) is a small constant to be specified below.

(i) Estimate of \(J_1\). We decompose

$$\begin{aligned} J_1&= \int _{\Omega _k^c} | \widetilde{R}_k|^{p-1}\bigg |\frac{\widetilde{S}_k^2}{|\widetilde{S}_k|^2} -\frac{\widetilde{R}_k^2}{|\widetilde{R}_k|^2} \bigg | |{\varepsilon }|^2dx + \int _{\Omega _k} | \widetilde{R}_k|^{p-1}\bigg |\frac{\widetilde{S}_k^2}{|\widetilde{S}_k|^2} -\frac{\widetilde{R}_k^2}{|\widetilde{R}_k|^2} \bigg | |{\varepsilon }|^2 dx\nonumber \\&:=J_{11}+J_{12}. \end{aligned}$$
(6.44)

Note that, for \(x\in \Omega _k^c\), since

$$\begin{aligned} |x-v_kt -{\alpha }_k| \ge&{\varepsilon }d_* > \frac{c}{2} t \end{aligned}$$
(6.45)

for t large enough, where \(c>0\), by (1.5), there exist \(C,\delta _2>0\) such that

$$\begin{aligned} J_{11}\le C\Vert R_k\Vert ^{p-1}_{L^{\infty }(\Omega ^c_k)}\Vert {\varepsilon }\Vert _{L^2}^2\le Ce^{-\delta _2 t}\Vert {\varepsilon }\Vert _{L^2}^2 . \end{aligned}$$
(6.46)

Concerning the first term \(J_{12}\) in (6.44), since \(Q(x)\sim e^{-\delta _0 |x|}\) (see [10]), we infer that

$$\begin{aligned} |\widetilde{R}_{k}(t,x)| \ge C e^{-\delta _0 \frac{ {\varepsilon }d_* }{w_k} }\ge C e^{-\delta _0 \frac{ {\varepsilon }d_* }{w_*} },\ \ x\in \Omega _k. \end{aligned}$$
(6.47)

On the other hand, for \(x\in \Omega _k\) and any \(j\not =k\),

$$\begin{aligned} |x-v_jt -{\alpha }_j| \ge |(v_k-v_j)t + {\alpha }_k-{\alpha }_j| - |x-v_kt-{\alpha }_k| \ge (1-{\varepsilon })d_*, \end{aligned}$$

which yields that

$$\begin{aligned} |\widetilde{S}_{k+1}(t,x)| \le C \sum \limits _{j=k+1}^K e^{- \delta _0 \frac{(1-{\varepsilon })d_*}{w_j}} \le C e^{-\delta _0 \frac{(1-{\varepsilon })d_*}{w^*}},\ \ x\in \Omega _k. \end{aligned}$$
(6.48)

Hence, we obtain from (6.47) and (6.48) that, for \({\varepsilon }\) small enough such that

$$\begin{aligned} {\varepsilon }<\frac{w_*}{w^*+w_*}, \end{aligned}$$

there exist \(C,\delta _2>0\) such that

$$\begin{aligned} \bigg |\frac{\widetilde{S}_{k+1}(t,x)}{\widetilde{R}_k(t,x)}\bigg | \le C e^{-\delta _0 d_* (\frac{(1-{\varepsilon })}{w^*}-\frac{{\varepsilon }}{w_*})} \le C e^{-\delta _2 t},\ \ x\in \Omega _k. \end{aligned}$$
(6.49)

Taking into account

$$\begin{aligned} \frac{\widetilde{S}_k^2}{|\widetilde{S}_k^2|} - \frac{\widetilde{R}_k^2}{|\widetilde{R}_k^2|} =&\frac{\widetilde{S}_k^2 |\widetilde{R}_k|^2 -| \widetilde{S}_k|^2 \widetilde{R}_k^2 + 2 \widetilde{R}_k \widetilde{S}_{k+1} |\widetilde{R}_k|^2 - 2 \textrm{Re}(\widetilde{R}_k \widetilde{S}_{k+1}) \widetilde{R}_k^2 }{|\widetilde{R}_k + \widetilde{S}_{k+1}|^2 |\widetilde{R}_k|^2} \nonumber \\ \le&\frac{|\widetilde{R}_k^{-1} \widetilde{S}_{k+1} |^2 + |\widetilde{R}_k^{-1} \widetilde{S}_{k+1}|}{|1+ \widetilde{R}_k^{-1} \widetilde{S}_{k+1}|^2} \end{aligned}$$

we thus lead to

$$\begin{aligned} \bigg | \frac{\widetilde{S}_k^2}{|\widetilde{S}_k^2|} - \frac{\widetilde{R}_k^2}{|\widetilde{R}_k^2|} \bigg | \le C e^{-\delta _2 t}, \ \ x\in \Omega _k, \end{aligned}$$
(6.50)

which yields that

$$\begin{aligned} J_{12} \le Ce^{-\delta _2 t}\Vert {\varepsilon }\Vert ^2_{L^2}. \end{aligned}$$
(6.51)

Thus, plugging (6.46) and (6.51) into (6.44) we obtain

$$\begin{aligned} J_2 \le Ce^{-\delta _2 t}. \end{aligned}$$
(6.52)

(ii) Estimate of \(J_2\). Set

$$\begin{aligned} \Omega = \bigcup \limits _{j=k+1}^K \Omega _j = \bigcup \limits _{j=k+1}^K \bigg \{x\in {{\mathbb {R}}}^d: |x-v_jt -{\alpha }_j| \le {\varepsilon }d_* \bigg \} \end{aligned}$$
(6.53)

and decompose

$$\begin{aligned} J_2&= \int \limits _{\Omega } | \widetilde{S}_{k+1}|^{p-1}\bigg |\frac{\widetilde{S}_k^2}{|\widetilde{S}_k|^2} -\frac{\widetilde{S}_{k+1}^2}{|\widetilde{S}_{k+1} |^2} \bigg | |{\varepsilon }|^2dx + \int \limits _{\Omega ^c} | \widetilde{S}_{k+1}|^{p-1}\bigg |\frac{\widetilde{S}_k^2}{|\widetilde{S}_k|^2} -\frac{\widetilde{S}_{k+1}^2}{|\widetilde{S}_{k+1} |^2} \bigg | |{\varepsilon }|^2dx \nonumber \\&= J_{21} + J_{22}. \end{aligned}$$
(6.54)

Note that, for every \(k+1\le j\le K\), since \(Q(x)\sim e^{-\delta _0|x|}\),

$$\begin{aligned} |\widetilde{R}_j(t,x)| \ge C e^{-\delta _0 \frac{{\varepsilon }d_*}{w_j}} \ge C e^{-\delta _0 \frac{{\varepsilon }d_*}{w_*}},\ \ x\in \Omega _j. \end{aligned}$$
(6.55)

Moreover, for \(x\in \Omega /\Omega _j\), there exists \(j'\not = j\) such that \(x\in \Omega _{j'}\) and so

$$\begin{aligned} |x-v_jt-{\alpha }_j| \ge |v_{j'}t +{\alpha }_{j'} - v_jt -{\alpha }_j| - |x-v_{j'}t-{\alpha }_{j'}| \ge (1-{\varepsilon })d_*. \end{aligned}$$

This yields that

$$\begin{aligned} |\widetilde{R}_j (t,x)| \le C e^{-\delta _0\frac{(1-{\varepsilon })d_*}{w_j}} \le C e^{-\delta _0 \frac{(1-{\varepsilon })d_*}{w^*}}, \ \ x\in \Omega /\Omega _j. \end{aligned}$$

Hence, for \({\varepsilon }\) very small such that

$$\begin{aligned} {\varepsilon }<\frac{w_*}{w_*+w^*}, \end{aligned}$$

we obtain that

$$\begin{aligned} |\widetilde{S}_{k+1}|\ge C e^{-\delta _0 \frac{{\varepsilon }d_*}{w_*}} - C' e^{-\delta _0 \frac{(1-{\varepsilon })d_*}{w^*}} \ge \frac{1}{2} C e^{-\delta _0 \frac{{\varepsilon }d_*}{w_*}}, \ \ x\in \Omega ,\ k+1\le j\le K, \end{aligned}$$
(6.56)

which yields that there exist \(C,\delta >0\) such that

$$\begin{aligned} |\widetilde{S}_{k+1}| \ge C e^{-\delta _0 \frac{{\varepsilon }d_*}{w_*}}, \ \ x\in \Omega . \end{aligned}$$
(6.57)

Moreover, for any \(x\in \Omega \), there exists \(k+1\le j\le K\) such that \(x\in \Omega _j\) and so

$$\begin{aligned} |x-v_kt -{\alpha }_k| \ge |(v_j-v_k)t -({\alpha }_j-{\alpha }_k)| - |x-v_jt -{\alpha }_j| \ge (1-{\varepsilon }) d_*, \end{aligned}$$
(6.58)

which yields that

$$\begin{aligned} |\widetilde{R}_k(t,x)| \le e^{-\delta _0\frac{(1-{\varepsilon })d_*}{w_k}}\le e^{-\delta _0\frac{(1-{\varepsilon })d_*}{w^*}}, \ \ x\in \Omega . \end{aligned}$$
(6.59)

Thus, we infer that for \({\varepsilon }\) possibly smaller such that

$$\begin{aligned} {\varepsilon }< \frac{w_*}{w_*+w^*}, \end{aligned}$$

then for \(x\in \Omega \),

$$\begin{aligned} \bigg |\frac{\widetilde{R}_k(t,x)}{\widetilde{S}_{k+1}(t,x)}\bigg | \le C e^{-\delta _0d_*(\frac{(1-{\varepsilon })}{w^*} - \frac{{\varepsilon }}{w_*} )} \le C e^{-\delta _2 t},\ \ x\in \Omega . \end{aligned}$$
(6.60)

Then, similar to (6.50), we have

$$\begin{aligned} \bigg | \frac{\widetilde{S}_k^2}{|\widetilde{S}_k^2|} - \frac{\widetilde{S}_{k+1}^2}{|\widetilde{S}_{k+1}^2|} \bigg | \le C \frac{|\widetilde{S}_{k+1}^{-1}\widetilde{R}_k |^2 + |\widetilde{S}_{k+1}^{-1}\widetilde{R}_k |}{|1+\widetilde{S}_{k+1}^{-1} \widetilde{R}_k |^2} \le C e^{-\delta _2 t}, \ \ x\in \Omega , \end{aligned}$$
(6.61)

which yields that

$$\begin{aligned} J_{21} \le C e^{-\delta _2 t} \Vert {\varepsilon }\Vert ^2_{L^2}. \end{aligned}$$
(6.62)

Concerning \(J_{22}\), we see that for \(x\in \Omega ^c\), for \(k+1\le j\le K\),

$$\begin{aligned} |x-v_jt -{\alpha }_j| \ge {\varepsilon }d_*, \end{aligned}$$

and so

$$\begin{aligned} |\widetilde{R}_j(t,x)| \le C e^{-\delta _0 \frac{{\varepsilon }d_*}{w_j}},\ \ x\in \Omega ^c. \end{aligned}$$
(6.63)

This yields that there exist \(C,\delta >0\) such that

$$\begin{aligned} |\widetilde{S}_{k+1}| \le C \sum \limits _{j=k+1}^K |\widetilde{R}_j| \le C e^{-\delta _2 t},\ \ x\in \Omega ^c, \end{aligned}$$
(6.64)

and thus

$$\begin{aligned} J_{22} \le C e^{-\delta _2 t} \Vert {\varepsilon }\Vert ^2_{L^2}. \end{aligned}$$
(6.65)

Thus, we obtain from (6.54), (6.62) and (6.65) that

$$\begin{aligned} J_2 \le Ce^{-\delta _2 t}. \end{aligned}$$
(6.66)

Therefore, plugging (6.52) and (6.66) into (6.43) we prove (6.42) and thus finish the proof. \(\square \)

Now, estimate (4.45) follows from Lemmas 6.6, 6.7, 6.8 and 6.9.

Rights and permissions

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

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Röckner, M., Su, Y. & Zhang, D. Multi solitary waves to stochastic nonlinear Schrödinger equations. Probab. Theory Relat. Fields 186, 813–876 (2023). https://doi.org/10.1007/s00440-023-01201-z

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00440-023-01201-z

Keywords

Mathematics Subject Classification

Navigation