Nonlinear Waves: Emmanuel Kengne Wuming Liu

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 525

Emmanuel Kengne

WuMing Liu

Nonlinear
Waves
From Dissipative Solitons to Magnetic
Solitons
Nonlinear Waves
Emmanuel Kengne · WuMing Liu

Nonlinear Waves
From Dissipative Solitons to Magnetic
Solitons
Emmanuel Kengne WuMing Liu
Zhejiang Normal University Institute of Physics
Jinhua, China Chinese Academy of Sciences
Beijing, China

ISBN 978-981-19-6743-6 ISBN 978-981-19-6744-3 (eBook)


https://doi.org/10.1007/978-981-19-6744-3

© The Editor(s) (if applicable) and The Author(s), under exclusive license to Springer Nature
Singapore Pte Ltd. 2022
This work is subject to copyright. All rights are solely and exclusively licensed by the Publisher, whether
the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse
of illustrations, recitation, broadcasting, reproduction on microfilms or in any other physical way, and
transmission or information storage and retrieval, electronic adaptation, computer software, or by similar
or dissimilar methodology now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication
does not imply, even in the absence of a specific statement, that such names are exempt from the relevant
protective laws and regulations and therefore free for general use.
The publisher, the authors, and the editors are safe to assume that the advice and information in this book
are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or
the editors give a warranty, expressed or implied, with respect to the material contained herein or for any
errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional
claims in published maps and institutional affiliations.

This Springer imprint is published by the registered company Springer Nature Singapore Pte Ltd.
The registered company address is: 152 Beach Road, #21-01/04 Gateway East, Singapore 189721,
Singapore
One of the authors, Emmanuel Kengne,
dedicates this book to the memory of his
parents, Papa François Mtopi and KUA
Marie Djomou, who passed away,
respectively, on October 1980 and September
2021. For if he had not believed that they
would have wished him to give such help as
he could toward making their life’s work of
service to mankind, he should never have
been led to co-author this book.
Preface

Mathematical modeling is a main key for understanding and investigating diverse


phenomena in both linear and nonlinear systems. When nonlinear phenomena are
governed by nonlinear partial differential equations (PDEs), it is important to obtain
their exact analytical solutions if they exist, otherwise, to obtain the numerical solu-
tions. Therefore, explorations of exact solutions of nonlinear PDEs are turned out
to be a charming and challenging area of research for mathematicians, theoretical
physicists, and research communities for many years because they are widely playing
a significant role in the study of nonlinear physical phenomena in applied physics and
mathematical studies with essential applications in several areas of engineering and
natural science including fluid mechanics, chemistry, thermodynamic, physics, elec-
tromagnetism, bio-mathematics, mathematical physics, and vice versa. This is why
it is very important to develop different analytical and numerical methods for finding
either exact analytical or numerical solutions of nonlinear PDEs which describe
nonlinear phenomena. This book focusses both on the mathematical modeling of
nonlinear phenomena in nonlinear electric transmission networks and on the study
of dynamics of both nonlinear modulated waves and matter waves in nonlinear
systems. One of its main aims is, through exact analytical solutions of nonlinear
PDEs, to present methodically several methods to engineer dissipative and magnetic
nonlinear waves such as dissipative and magnetic solitons (MSs) propagating in
nonlinear systems. In soliton theory, the development of dissipative, non-dissipative
solitons as well as soliton signals might be a consequence of the nonlinear, dispersive,
and diffusive results, whereas a soliton is also called solitary or soliton-like wave
that maintains its shapes, velocity, and amplitude with the impact of others solitons.
The book consists of two parts and thirteen chapters. Part I, titled “Engineering
Nonlinear Modulated Waves in Nonlinear Transmission Networks”, consists of five
chapters and investigates the generation and the transmission of nonlinear modulated
waves, in general, and solitary/solitonlike waves in particular in lossless and in lossy
electric nonlinear transmission networks (NLTNs). Part I of the book is composed of
two parts, which deal with, respectively, one-dimensional (1D) and two-dimensional
(2D) NLTNs. For both 1D and 2D NLTNs, we consider both lossless and dissipative

vii
viii Preface

NLTNs. The respective models are based on either 1D or 2D conservative and dissi-
pative nonlinear Schrödinger (NLS) equations, derivative NLS equations cubic and
cubic-quintic complex Ginzburg-Landau (CGL) equations, Kundu–Eckhaus equa-
tion, Chen-Liu and generalized Chen–Lee–Liu equations and cubic-quintic CGL
equations with derivative terms. Chapter 2 deals with the NLS models for solitons
propagation in 1D lossless nonlinear transmission networks, while Chap. 3 treats
the generation of dissipative solitons in one-dimensional NLTNs. In Chaps. 4 and 5,
we focus our attention, respectively, to the dynamics of rogue waves in 1D NLTNs
and to the transmission of nonlinear modulated waves in multi-component nonlinear
transmission networks.
Part II of this book titled “Dynamics of Matter-Wave and Magnetic-Wave Soli-
tons” is composed of eight chapters. In this part of the book, we develop basic theo-
retical results for the dynamics matter-wave and magnetic-wave solitons of nonlinear
systems in general and of Bose–Einstein condensates (BECs) in particular, as well
as of optical fields trapped in external potentials, combined with the time-modulated
nonlinearity. The respective models are based on one-, two-, and three-component
non-autonomous Gross–Pitaevskii equations or NLS equations with external poten-
tials. Chapter 7 addresses the dynamics of one-component condensates with time
modulation of the scattering length and trapping potential. Here, the approach
outlined above leads to the construction and management of non-autonomous soli-
tons of 1D cubic self-defocusing NLS equations with spatiotemporally modulated
coefficients that may be transformed into the classical integrable NLS equation. In
this setting, matter-wave soliton solutions are constructed in an analytical form, and
it is shown that the instability of those solitons, if any, may be delayed or completely
eliminated by varying the nonlinearity’s strength in time. In this Chap. 7, we also engi-
neer non-autonomous matter-wave solitons in BECs with spatially modulated local
nonlinearity and a time-dependent harmonic-oscillator (HO) potential. The modula-
tional instability in that setting is considered too. Chapter 8 deals with the dynamics of
matter rogue waves in BECs trapped in time-dependent external potential. Chapter 9
treats the soliton generation in multi-component Bose–Einstein condensates. In
Chap. 10, we proceed to more general non-integrable models, which are treated
by means of the semi-analytical variational approximation and direct numerical
simulations. In this case, we address the dynamics of two- and three-dimensional
condensates with the nonlinearity strength containing constant and harmonically
varying parts, which can be implemented with the help of AC magnetic field tuned
to FR. In particular, the spatially uniform temporal modulation of the nonlinearity
may readily play the role of an effective trap that confines the condensate and some-
times enforces its collapse. Chapter 11 deals with the investigation of soliton states
in a model based on a set of three coupled GP equations modeling the dynamics
of the spinor BEC. Both integrable and non-integrable versions of the system are
considered, and exact soliton solutions are demonstrated. The stability of the solu-
tions was checked, in most cases, by direct simulations, and, in some cases, it was
investigated in a more rigorous form, based on linearized Bogoliubov–de Gennes
(BdG) equations for small perturbations. The motion of bright and dark matter-wave
solitons in 1D BEC in the presence of spin-orbit coupling (SOC) is also treated in
Preface ix

Chap. 11. Here, we demonstrate that the spin dynamics of the SOC solitons are
governed by a nonlinear Bloch equation and affects the orbital motion of the soli-
tons, leading to SOC effects in the dynamics of macroscopic quantum objects. The
models addressed in Chap. 11, as well as the applied methods and the obtained
results, are quite similar to those produced by the engineering. In particular, the
macroscopic SOC phenomenology is explained by the fact that an effective time-
periodic force produced by rotation of the soliton’s (pseudo-) spin plays the role of
temporal management which affects the motion of the same soliton. Chapters 12
and 13 deal with the dynamics of magnetization in ferromagnet with or without
spin-transfer torque.
This research book, which accurately treats the mathematical modeling of
nonlinear phenomena and presents methodically several methods for engineering
the propagation of nonlinear waves in nonlinear systems, is suitable for physicists,
mathematicians, engineers, as well as for graduate and postgraduate students from
schools of mathematics, physics, network, and information engineering.

Jinhua, China Emmanuel Kengne


Beijing, China WuMing Liu
Acknowledgements

Nobody has been more important to Prof. Emmanuel Kengne of the Zhejiang
Normal University (China), one of the authors, in the pursuit of the multiple projects
leading to this book than the members of his family. He wishes to thank his loving and
supportive wife, Eleonore Nkuojip Sado, and his three wonderful children, Kengneson
Delma Djomo, Kengneson Weierstrass Owan Wambo, and Kengneson Cris-Carelle
Djike all at Gatineau-Quebec-Canada, whose love and guidance are with him in
whatever he pursues. They are the ultimate role models.
This book was supported by the Chinese Academy of Sciences PIFI under the
grant No. 2023VMA0019, National Key R&D Program of China under grants No.
2021YFA1400900, 2021YFA0718300, 2021YFA1402100, and NSFC under grants
Nos. 11835011, 61835013, 12234012, Space Application System of China Manned
Space Program.

xi
Contents

Part I Engineering Nonlinear Modulated Waves in Nonlinear


Transmission Networks
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2 Nonlinear Schrödinger Models for Solitons Propagation in 1D
Lossless Nonlinear Transmission Networks (NLTNs) . . . . . . . . . . . . . . 7
2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.2 Standard Nonlinear Schrödinger Equations for Modulated
Waves Propagation in a Lossless Electrical Transmission
the Network: Effects of the Dispersive Element C S . . . . . . . . . . . . 8
2.2.1 Model Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.2.2 Bright and Dark Single-Solitary Waves
Propagating in the 1D Lossless Network of Fig. 2.1 . . . . 13
2.3 Modulational Instability and Transmission of Chirped
Femtosecond Signals Through the Lossless Electrical
Network of Fig. 2.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.3.2 Modulational Instability and Evolution of Chirped
Femtosecond Solitary Signals Embedded
on a Non-vanishing CW Background . . . . . . . . . . . . . . . . 23
2.3.3 Sister Femtosecond Nonlinear Modulated Waves
with Two Nonlinear Chirp Terms . . . . . . . . . . . . . . . . . . . . 30
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
3 Transmission of Dissipative Solitonlike Signals Through
One-Dimensional Transmission Networks . . . . . . . . . . . . . . . . . . . . . . . 47
3.1 Chirped Lambert W-Kink Waves Propagation in a Lossy
Electrical Transmission Network . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
3.1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
3.1.2 Amplitude Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49

xiii
xiv Contents

3.1.3 Baseband Modulational Instability Analysis . . . . . . . . . . 54


3.1.4 Evolution of Chirped Lambert W-Kink Pulses
in the Network of Fig. 3.1 . . . . . . . . . . . . . . . . . . . . . . . . . . 58
3.2 Spatiotemporal Modulation of Damped Solitonlike Wave
Signals in a Nonlinear RLC Transmission Network . . . . . . . . . . . . 67
3.2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
3.2.2 Linear Dispersion Relation and Spatial Decreasing
Rate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
3.2.3 Amplitude Equation and Modulated Damped . . . . . . . . . 72
3.2.4 Linear Stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
3.2.5 Spatiotemporal Modulated Signals Propagating
Through the Network of Fig. 3.8 . . . . . . . . . . . . . . . . . . . . 80
3.2.6 Transmission of Spatiotemporal Modulated
Damped Envelope Signals Through a Lossy
Network . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
3.3 Modulated Wavetrains in a Dissipative Bi-Inductance
Transmission Network . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
3.3.1 Introduction and Circuit Equations . . . . . . . . . . . . . . . . . . 93
3.3.2 Generalized Cubic-Quintic Complex
Ginzburg-Landau Equation for a Lossy Network . . . . . . 94
3.3.3 Linear Analysis and Modulational Instability . . . . . . . . . 96
3.3.4 Coherent Structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
4 Emission of Rogue Wave Signals in Nonlinear Electrical
Transmission Networks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
4.1 Emission of Rogue Wave Signals Through the Modified
Noguchi Electric Transmission Network . . . . . . . . . . . . . . . . . . . . . 109
4.1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
4.1.2 Modulated Waves and Linear Analysis . . . . . . . . . . . . . . . 110
4.1.3 Construction of Rogue Waves in the Lossless
Network Under the Condition P Q > 0 . . . . . . . . . . . . . . 113
4.1.4 Emission of Rogue Wave Signals Through
the Network of Fig. 2.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118
4.1.5 Conclusion and Discussions . . . . . . . . . . . . . . . . . . . . . . . . 124
4.2 Generation of Network Modulated Rogue Waves Under
the Action of the Quintic Nonlinearity . . . . . . . . . . . . . . . . . . . . . . . 124
4.2.1 The Kundu–Eckhaus Model for Non-autonomous
Modulated Rogue Waves in a Lossless Electric
Network . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
4.2.2 Generation of First-Order Non-autonomous
Modulated Rogue Waves (Alias Peregrine
Solitons) for a Lossless Electric Network . . . . . . . . . . . . . 127
4.2.3 Conclusion and Discussions . . . . . . . . . . . . . . . . . . . . . . . . 136
Contents xv

4.3 Chirped Super Rogue Waves Propagating Along a Lossless


Nonlinear Electrical Network . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138
4.3.1 Generalized Nonlinear Schrödinger Equation
for a Lossless Electric Network . . . . . . . . . . . . . . . . . . . . . 139
4.3.2 Phase Engineering Chirped Super Rogue Waves
for a Lossless Electric Network . . . . . . . . . . . . . . . . . . . . . 145
4.3.3 Computational/Numerical Simulations . . . . . . . . . . . . . . . 153
4.3.4 Conclusion and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . 156
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159
5 Emission of Nonlinear Modulated Waves in Multi-coupled
Nonlinear Transmission Networks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163
5.1 Transmission of Solitonlike Wave Signals
in a Two-Dimensional Lossless Dispersive Nonlinear
Transmission Network . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163
5.1.1 Description of the and Network Equations . . . . . . . . . . . . 163
5.1.2 Amplitude Equation for the Dynamics
of Modulated Waves in the Two-Dimensional
Lossless Electric Network . . . . . . . . . . . . . . . . . . . . . . . . . . 165
5.1.3 Computational Simulations . . . . . . . . . . . . . . . . . . . . . . . . . 167
5.2 Coherent Structures for a Multi-coupled Nonlinear
Transmission Network with Dissipative Elements . . . . . . . . . . . . . 169
5.2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 170
5.2.2 Amplitude Equation for a Two-Dimensional
Lossy Network . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172
5.2.3 Modulational Instability in a Two-Dimensional
Lossy Transmission Network . . . . . . . . . . . . . . . . . . . . . . . 173
5.2.4 Coherent Structures for a Two-Dimensional Lossy
Nonlinear Transmission Network . . . . . . . . . . . . . . . . . . . 177
5.3 Soliton Signals in an Alternate Right-Handed
and Left-Handed Multi-coupled Lossy Nonlinear
Transmission Network . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181
5.3.1 Description of the Model and Basic Equations . . . . . . . . 181
5.3.2 Complex Ginzburg-Landau Equation
for the Dynamics of Modulated Waves
in a Multi-coupled Lossy Network . . . . . . . . . . . . . . . . . . 183
5.3.3 Linear Stability Analysis: Modulational Instability
of Stokes Waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189
5.3.4 Dissipative Effects on Electrical Modulated Wave
Propagation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 191
5.3.5 Conclusion and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . 209
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 210
xvi Contents

Part II Dynamics of Matter-Wave and Magnetic-Wave Solitons


6 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 215
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 218
7 Dynamics of One-Dimensional Condensates with Time
Modulation of the Scattering Length and Trapping Potential . . . . . . 223
7.1 Non-autonomous Solitons in Bose-Einstein Condensates
with a Spatially Modulated Scattering length . . . . . . . . . . . . . . . . . 223
7.1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 224
7.1.2 Cubic Inhomogeneous Nonlinear Schrödinger
Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 224
7.1.3 Matter-Wave Solitons in an Inhomogeneous
Nonlinear Schrödinger Equation
with the Spatiotemporal HO Potential . . . . . . . . . . . . . . . . 235
7.1.4 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 240
7.2 Soliton Management (SM) in One-Dimensional
Bose-Einstein Condensates with Two- and Three-Body
Inter-atomic Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 241
7.2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 242
7.2.2 Modulational Instability
in Bose-Einstein Condensates Trapped
in a Spatiotemporal-Dependent Dissipative
Potential . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 244
7.2.3 Soliton Management in Bose-Einstein Condensate
Systems with Two- and Three-Body Interactions
Trapped in a Spatiotemporal-Dependent
Dissipative Potential . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 249
7.2.4 Conclusion and Discussions . . . . . . . . . . . . . . . . . . . . . . . . 263
7.3 Chirped Solitons and Chirped Double-Kink Solitons
in Bose-Einstein Condensates with Time-Varying Atomic
Scattering Length in a Complicated External Potential . . . . . . . . . 264
7.3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 265
7.3.2 Derivation of the Higher-Order Nonlinear
Schrödinger Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 266
7.3.3 Dynamics of Chirped Femtosecond
Solitons and Chirped Double-Kink
Solitons in Bose-Einstein Condensates
with Time-Dependent Atomic Scattering Length
in a Complex Potential . . . . . . . . . . . . . . . . . . . . . . . . . . . . 276
7.3.4 Conclusion and Discussions . . . . . . . . . . . . . . . . . . . . . . . . 283
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 283
Contents xvii

8 Rogue Matter Waves in Bose-Einstein Condensates Trapped


in Time-Varying External Potentials . . . . . . . . . . . . . . . . . . . . . . . . . . . . 289
8.1 Non-autonomous Rogue Matter Waves in Bose-Einstein
Condensates With Dissipative External Potential . . . . . . . . . . . . . . 289
8.1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 289
8.1.2 Model and Analytical Exact First-
and Second-Order Rational Solutions
of the Gross-Pitaevskii Equation . . . . . . . . . . . . . . . . . . . . 290
8.1.3 Management of Dissipative Rogue Matter Waves
in Bose-Einstein Condensates with Complicated
Potential . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 292
8.1.4 Conclusion and Discussions . . . . . . . . . . . . . . . . . . . . . . . . 302
8.2 Chirped Rogue Matter Waves in Bose-Einstein
Condensates with a Variable Scattering Length
in an Expulsive Potential . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 302
8.2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 302
8.2.2 Phase Engineering and Chirped Wave Solutions . . . . . . . 303
8.2.3 Evolution of Chirped Rogue Waves Under
the Action of the Time-Dependent Atomic
Scattering Length and Parabolic Potential . . . . . . . . . . . . 309
8.2.4 Conclusion and Discussions . . . . . . . . . . . . . . . . . . . . . . . . 320
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 322
9 Dynamics of Matter-Wave Solitons in Multi-component
Bose-Einstein Condensates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 329
9.1 Soliton Management in a Binary Bose-Einstein Condensate . . . . 329
9.1.1 The Model and Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . 329
9.1.2 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 333
9.1.3 Conclusion and Discussions . . . . . . . . . . . . . . . . . . . . . . . . 337
9.2 Soliton Stability in Binary Bose-Einstein Condensate
Under Temporal Modulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 338
9.2.1 The Physical Model and the Lax Pair . . . . . . . . . . . . . . . . 339
9.2.2 Analytical and Numerical Results
for Two-Component Bright Solitons
in the Integrable System . . . . . . . . . . . . . . . . . . . . . . . . . . . 341
9.2.3 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 347
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 348
10 Dynamics of Higher-Dimensional Condensates with Time
Modulated Nonlinearity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 349
10.1 Dynamics of Two- and Three-Dimensional Bose-Einstein
Condensates with Time Modulated Nonlinearities . . . . . . . . . . . . . 349
10.1.1 The Model and Variational Approximation (VA) . . . . . . 350
10.1.2 Two-Dimensional Case . . . . . . . . . . . . . . . . . . . . . . . . . . . . 351
10.1.3 The Three-Dimensional Gross-Pitaevskii Model . . . . . . . 359
10.1.4 Conclusion and Discussions . . . . . . . . . . . . . . . . . . . . . . . . 363
xviii Contents

10.2 Stable Vortex Modes in Two-Dimensional Bose-Einstein


Condensates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 363
10.2.1 Model Description and Main Transformation . . . . . . . . . 364
10.2.2 Exact Vortex-Soliton Solutions for the Attractive
Nonlinearity (g0 < 0) when E = 0 . . . . . . . . . . . . . . . . . . 365
10.2.3 Exact Analytical Vortex-Soliton Solutions
for the Repulsive Nonlinearity (g0 > 0) . . . . . . . . . . . . . . 367
10.2.4 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 370
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 371
11 Engineering Matter-Wave Solitons in Spinor Bose-Einstein
Condensates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 373
11.1 Formulation of the Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 373
11.2 Exact Analytical One-, Two-, and Three-Component
Soliton Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 377
11.2.1 Single-Component FM Solitons . . . . . . . . . . . . . . . . . . . . . 377
11.2.2 Single-Component Polar Solitons . . . . . . . . . . . . . . . . . . . 378
11.2.3 Two-Component Polar Solitons . . . . . . . . . . . . . . . . . . . . . 380
11.2.4 Three-Component Polar Solitons . . . . . . . . . . . . . . . . . . . . 382
11.2.5 Multistability of Solitons . . . . . . . . . . . . . . . . . . . . . . . . . . . 384
11.2.6 Finite-Background Solitons . . . . . . . . . . . . . . . . . . . . . . . . 385
11.3 The Darboux Transform and Nonlinear Development
of Modulational Instability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 387
11.4 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 393
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 393
12 Engineering Magnetic Solitons in Nonlinear Systems . . . . . . . . . . . . . 395
12.1 Dynamics of Dissipative Magnetic Matter-Wave Solitons
in a Spinor Polariton Bose-Einstein Condensate . . . . . . . . . . . . . . . 395
12.1.1 Introduction and Model Equations . . . . . . . . . . . . . . . . . . . 396
12.1.2 Exact Analytical Solutions and Dissipative
Magnetic Polariton Solitons in a Spinor Polariton
Bose-Einstein Condensate . . . . . . . . . . . . . . . . . . . . . . . . . 398
12.1.3 Conclusion and Discussions . . . . . . . . . . . . . . . . . . . . . . . . 403
12.2 Nonlinear Magnetization Dynamics of a Classical
Ferromagnet . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 403
12.2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 403
12.2.2 Macroscopic Description and Equation of Motion . . . . . 404
12.2.3 Soliton Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 412
12.2.4 Asymptotic Behavior of Multisoliton Solutions . . . . . . . 428
12.2.5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 430
12.3 Nonlinear Magnetization Dynamics in the Presence
of Spin-Polarized Current . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 431
12.3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 431
12.3.2 Long Ferromagnetic Nanowire with a Uniform
Cross Section (High-Q Model, Q > 1) . . . . . . . . . . . . . . . 432
Contents xix

12.3.3 Short Ferromagnetic Nanowire with a Uniform


Cross Section (Low-Q Model, Q < 1) . . . . . . . . . . . . . . . 438
12.3.4 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 441
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 441
13 Current Driven Dynamics of Magnetization in Ferromagnet
with Spin Transfer Torque . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 447
13.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 447
13.2 Current Driven Dynamics of Domain Wall in Ferrimagnets . . . . . 448
13.2.1 Domain-Wall Resonance Induced
by Spin-Polarized Current . . . . . . . . . . . . . . . . . . . . . . . . . . 448
13.2.2 Screw-Pitch Effect and Velocity Oscillation
of Domain Wall . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 453
13.2.3 Non-autonomous Helical Motion of Magnetization
in Ferromagnetic Nanowire . . . . . . . . . . . . . . . . . . . . . . . . 458
13.3 Current Driven Dynamics of Soliton . . . . . . . . . . . . . . . . . . . . . . . . 468
13.3.1 Dark Solitons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 469
13.3.2 Bright Soliton . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 473
13.3.3 Current Driven Interaction of Spin Wave
and Soliton . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 474
13.3.4 Dynamics of Magnetic Rogue Wave . . . . . . . . . . . . . . . . . 484
13.4 Ferromagnetic Resonance in Magnetic Trilayers . . . . . . . . . . . . . . 490
13.4.1 Ferromagnetic Resonance
in a Perpendicular-Analyzer Magnetic
Trilayer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 491
13.4.2 Stability Analysis of Perpendicular Magnetic
Trilayers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 498
13.5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 505
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 506

Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 511
About the Authors

Emmanuel Kengne obtained both his Master’s degrees in mathematics and physics
and his Ph.D. in Physicomathematical Sciences from the School of Mathematics
and Mechanical Engineering at Kharkov State University (now Kharkov National
University), Ukraine, on, respectively, July 1991 and March 1994. He also obtained
a Master of Education with concentration in Teaching, Learning and Evaluation at
the University of Ottawa (Canada) on April 2012. Emmanuel Kengne is an applied
mathematician and theoretical physicist, full professor at the School of Physics and
Electronic Information Engineering of Zhejiang Normal University (China), adjunct
professor at the Department of Computer Science and Engineering of the Univer-
sity of Quebec at Outaouais (Canada), and adjunct researcher at the Institute of
Physics, Chinese Academy of Sciences (China). A multidisciplinary researcher, his
research plan mainly encompasses four main themes that lead to interdisciplinary
collaborations among different fields such as applied mathematics and pure mathe-
matics, nonlinear physics, mathematical physics, electrical and telecommunication
engineering, biomedical engineering, computational sciences, nonlinear theory and
its applications, and biological sciences. Emmanuel Kengne has made major contri-
butions to a vast number of fields, including the theory of well-posedness boundary
value problems for partial differential equations, wave propagation on nonlinear
transmission networks, optical and heat solitons, nonlinear dynamical lattices,
Ginzburg-Landau models, Boson–Fermion models, bio-thermal physics, light prop-
agation, thermal therapy for tumors, as well as many other physicomathematical
fields.

WuMing Liu obtained his Ph.D. degree from the Institute of Metal Research, Chinese
Academy of Sciences, Shenyang, China, in June 1994. He became an associate
professor at the Institute of Theoretical Physics, Chinese Academy of Sciences,
Beijing, China, in 1996, and is a full professor at the Institute of Physics at the
same Academy since 2002. He has served as an editorial board member for several
international journals, including Scientific Reports and Journals of Physics.

xxi
Abbreviations and Acronyms

1D One dimensional
2D Two dimensional
3D Three dimensional
AC Alternating Current
AM Ansatz method
AMI Azimuthal modulational instability
BC Boundary condition
BdG equation Bogoliubov–de Gennes equation
BEC Bose–Einstein condensate
C.C. Complex Conjugation
CGL equation Complex Ginzburg-Landau equation
CLL equation Chen–Lee–Liu equation
CQ Qubic-quintic
CW Continuous wave
DC Direct Current
DM Direct method
DM Dispersion management
DNLS Derivative nonlinear Schrödinger equation
DT Darboux transform
DW Domain wall
EODE Elliptic ordinary differential equation
FP Fixed point
FR Feshbach resonance
GCQ Generalized cubic-quintic
GCQD-NLS Generalized cubic-quintic derivative nonlinear Schrödinger
GI equation Gerdjikov–Ivanov equation
GL equation Ginzburg-Landau equation
GLM equation Gel’fand-Levitan-Marchenko equation
GP equation Gross–Pitaevskii equation
GVD Group velocity dispersion
HBM Hirota bilinear method

xxiii
xxiv Abbreviations and Acronyms

HO Harmonic oscillator
HO Higher order
IST Inverse scattering transformation
KdV Korteweg–de Vries
KE equation Kundu–Eckhaus equation
KN equation Kaup–Newell equation
LH Left handed
lhs Left-hand side
LL equation Landau–Lifshitz equation
LLG equation Landau–Lifshitz–Gilbert equation
LLGS equation Landau–Lifshitz–Gilbert–Slonczewski equation
MI Modulational instability
MRW Magnetic rogue wave
MS Magnetic soliton
NLS equation Nonlinear Schrödinger equation
NLTN Nonlinear transmission network
NM Nonlinearity management
NMDs Nonlinear magnetization dynamics
NPDE Nonlinear partial differential equation
NSW Nonlinear spin wave
ODE Ordinary differential equation
PDE Partial differential equation
PIT Phase engineering transformation
PT Parity time
RH Right handed
rhs Right-hand side
RKMK method Runge–Kutta–Munthe–Kaas method
RW Rogue wave
SM Soliton management
SOC Spin-orbit coupling
SPM Self-phase modulation
SRW Super rogue wave
SW Spin wave
TS wave Tollmien–Schlichting wave
VA Variational approximation
VS Vortex soliton
XPM Cross-phase modulation
Part I
Engineering Nonlinear Modulated Waves
in Nonlinear Transmission Networks
Chapter 1
Introduction

Abstract This Chapter introduces the first part of the Book. It presents nonlinear
transmission networks in general and, lossless/lossy nonlinear electric transmission
networks, in particular. Also, we present here scientific motivations for studying non-
linear excitations in both lossless and dissipative nonlinear transmission networks.

While various nonlinear partial differential equations have been explored in detail,
used to model the real systems in different scientific domains, and many solutions
of these equations have been found, there are only a few systems where solitons are
easily and directly observed in controlled laboratory experiments. Nonlinear trans-
mission networks (NLTNs)—most often, electric transmission lines—are relevant
examples of such experimentally available setups [1, 2]. Being discrete systems,
they approximate continuum media quite well, as shown in this part of our book.
Thus, NLTNs provide a reliable platform for studying the dynamics of excitations in
nonlinear media [2].
An electric NLTN is built as a transmission line with periodically inserted var-
actors, in which the capacitance nonlinearity arises from a variable depletion-layer
width, that depends both on the dc bias voltage and ac voltage of the propagating
waves. Usually, diodes with reverse-biased capacitance are employed as nonlinear
capacitors [3]. NLTNs constructed according to this scheme are very convenient
tools to study the wave propagation in one-dimensional (1D) nonlinear dispersive
media [4]. A growing interest to the work with NLTNs has been drawn since the
pioneering works by Hirota and Suzuki and by Nagashima and Amagishi on a single
electrical network simulating the Toda lattice [1, 5–10]. The multiplicative process
in NLTNs is understood as a direct consequence of the propagation of modulated
waves in this medium. Qualitatively, the origin of nonlinear waves, such as solitons,
in NLTNs is explained by the balance between the effect of dispersion, due to the
periodic location of capacitors, and nonlinearity due to the voltage dependence of
the capacitance [11].
Discrete electrical transmission networks have been used in the design of thin-
film and diffusion resistors, capacitors, and conductors, as well as for evaluation of
© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2022 3
E. Kengne and W. Liu, Nonlinear Waves,
https://doi.org/10.1007/978-981-19-6744-3_1
4 1 Introduction

undesirable interaction between different components of integrated circuits. These


networks can be successfully used to simulate characteristics of some active micro-
circuit elements, such as the field-effect transistors [12, 13]. In particular, dissipative
transmission networks with a nonlinear load are very important for the design of
high-speed electronic circuits [14]. To properly consider increasing frequencies and
bit rates, one has to take transmission network effects into account. In particular,
these are the signal delay, refection, attenuation, and crosstalk [15]. In this context,
nonlinear transmission networks have proven to be of great practical use for the
focusing and shaping of extremely wide-band signals (with frequencies from dc to
100 GHz), i.e., controllably changing basic features of incoming signals, such as the
frequency content, pulse width, and amplitude, which is usually a hard problem [16].
Further, topoelectric (i.e., topological electric [17], also called “topoelectric" [18])
two-dimensional (2D) circuit networks can be employed to emulate various topo-
logical states of matter, including surface and corner modes in topological insula-
tors with a spatially periodic [18–20] and quasiperiodic [21] structure. Interest has
also been recently drawn to the creation of non-Hermitian topologically structured
electric networks incorporating parity-time (PT ) symmetric or chiral anti-PT -
symmetric elements [22]. In biomedical engineering, the RLC networks are used
for modeling blood flow in biological tissues [23, 24]. In the linear regime, NLTNs
can be used as phase shifters in phased antenna arrays, where the time delay can
be controlled by means of a dc bias applied to Schottky diodes acting as variable
reactance [25]. Under large-signal conditions, NLTNs can serve as impulse com-
pressors or frequency multipliers. Efforts have been made to study, both analytically
and numerically, nonlinear excitations in NLTNs. In particular, it has been shown
experimentally that modulated waves can exist in such systems [16]. A decade ago,
the backward-wave propagation and formation of discrete bright and dark solitons
in a nonlinear electrical lattice were addressed [26]. Still earlier, the existence of
stable discrete solitons in an electric transmission network emulating an ac-driven
lossy Toda lattice was experimentally demonstrated by Kuusela [27]. Those results
confirm the stable motion of the solitons at resonant velocities, as predicted in [28]
and [29].
The aim of this part of the book is to survey recent and most essential advances in
the application of nonlinear PDEs for the generation of nonlinear modulated waves
in real NLTNs. We present mathematical models of nonlinear phenomena occur-
ring along NLTNs and produce a study of dynamics of nonlinear modulated waves
propagating in these networks, aiming to make it sufficiently complete. For this pur-
pose, essential elements of the derivation of nonlinear PDEs are included, aiming to
make the presentation accessible to the general readership. To put the new findings
in the context and clarify core application of nonlinear PDEs to NLTNs, we are par-
ticularly concerned with topics such as the modulational instability, the generation
of envelope and/or hole solitons (dissipative or non dissipative), the evolution and
decay of chirped solitons, and dynamics of rogue waves (RWs). It is important to
note that research on modulation instability and dynamical behavior of the localized
References 5

wave solutions (as for examples solitons, breathers, and rogue waves) for the nonlin-
ear systems is a fascinating subject in the field of contemporary nonlinear science,
which has attracted widespread attention from many experts and scholars. In nonlin-
ear dispersive media, MI is a nonlinear process, in which a continuous plane wave
generates amplitude and frequency self-modulation through a nonlinear dispersive
medium, resulting in exponential growth of small perturbations superimposed on a
plane wave.
This part of the book is composed of two parts, which deal with, respectively,
single-(1D) and multi-component (2D) NLTNs. In each part, we consider both loss-
less and the dissipative NLTNs. The respective models are based on either 1D or
2D conservative and dissipative nonlinear Schrödinger (NLS) equations, derivative
NLS (DNLS) equation, cubic and cubic-quintic complex Ginzburg-Landau (CGL)
equations, Kundu–Eckhaus (KE) equation [30–34], Chen-Liu [35] and generalized
Chen-Lee-Liu equation [36], and a cubic-quintic (CQ) CGL equation with derivative
terms. The first part of this book, which addresses the dynamics of modulated waves
in single (1D) and two-dimensional (alias multi-component) NLTNs, is composed
of four Chapters. This chapter deals with the NLS models for solitons propagation in
one-dimensional lossless nonlinear transmission networks, while Chap. 2 treats the
generation of dissipative solitons in one-dimensional NLTNs. Chapters 3 and 4 deal
with respectively the dynamics of rogue waves in 1D NLTNs and the transmission
of modulated waves in multi-component nonlinear transmission networks.

References

1. R. Hirota, K. Suzuki, Theoretical and experimental studies of lattice solitons in nonlinear


lumped networks. Proc. IEEE 61, 1483 (1973)
2. K. Lonngren, Solitons in Action. eds. by K. Lonngren, A. Scott (New York: Academic, 1978)
3. R. Marquié, J.M. Bilbault, M. Remoissenet, Nonlinear Schrödinger models and modulational
instability in real electrical lattices. Phys. D 87, 371–374 (1995)
4. A.C. Newell, J.V. Moloney, Nonlinear Optics (Addison-Wesley, 1991)
5. R. Hirota, K. Suzuki, Studies on lattice solitons by using electrical networks. J. Phys. Soc.
Japan 28, 1366 (1970)
6. H. Nagashima, Y. Amagishi, Experiment on the Toda lattice using nonlinear transmission lines.
J. Phys. Soc. Japan 45, 680 (1978)
7. M. Toda, Wave propagation in an harmonic lattices. J. Phys. Soc. Japan 23, 501 (1967)
8. P. Marquié, J.M. Bilbault, M. Remoissenet, Generation of envelope and hole solitons in an
experimental transmission line. Phys. Rev. E 49, 828 (1994)
9. E. Kengne, A. Lakhssassi, W.M. Liu, Nonlinear Schamel-Korteweg de Vries equation for a
modified Noguchi nonlinear electric transmission network: analytical circuit modeling. Chaos,
Solitons Fractals 140, 110229 (2020)
10. E. Kengne, B.A. Malomed, S.T. Chui, W.M. Liu, Solitary signals in electrical nonlinear trans-
mission line. J. Math. Phys. 48
11. A.C. Scoot, F.Y.F. Chu, W. Mclaughlin, The soliton: a new concept in applied science. Proc.
IEEE 61, 1443 (1973)
12. R.S.C. Cobbold, Theory and Application of Field-Effect Transistors (Wiley, New York, 1970)
13. V.P. Popov, T.A. Bickart, RC transmission line with nonlinear. resistance: large-signal response
computation. IEEE Trans. Circuits Syst. CAS 21, 666 (1974)
6 1 Introduction

14. A.R. Djordjevic, T.K. Sarkar, R.F. Harrington, Analysis of Lossy transmission lines with arbi-
trary nonlinear terminal networks. IEEE Trans. Microwave Theory Tech. MTT 34, 660 (1986)
15. Q. Gu, Y.E. Yang, J.A. Kong, Transient analysis of frequency-dependent transmission line
systems terminated with nonlinear loads. J. Electromagn. Waves Appl. 3, 183 (1989)
16. E. Afshari, A. Hajimiri, Nonlinear transmission lines for pulse shaping in silicon. IEEE J.
Solid-State Circuits 40, 744 (2005)
17. C.H. Lee, S. Imhof, C. Berger, F. Bayer, J. Brehm, L.W. Molenkamp, T. Kiessling, R. Thomale,
Topoelectrical circuits. Commun. Phys. 1, 39 (2018)
18. J.C. Bao, D.Y. Zou, W.X. Zhang, W.J. He, H.J. Sun, X.D. Zhang, Topoelectrical circuit octupole
insulator with topologically protected corner states. Phys. Rev. B 100, 201406 (2019)
19. S. Imhof, C. Berger, F. Bayer, J. Brehm, L.W. Molenkamp, T. Kiessling, F. Schindler, C.H.
Lee, M. Greiter, T. Neupert, R. Thomale, Topolectrical-circuit realization of topological corner
modes. Nat. Phys. 14, 925–929 (2018)
20. X. Ni, Z.C. Xiao, A.B. Khanikaev, A. Alu, Robust multiplexing with topoelectrical higher-order
Chern insulators. Phys. Rev. Appl. 13, 064031 (2020)
21. B. Lv, R. Chen, R. Li, C. Guan, B. Zhou, G. Dong, C. Zhao, Y. Li, Y. Wang, H. Tao, J. Shi, D.-
H. Xu, Realization of quasicrystalline quadrupole topological insulators in electrical circuits.
Commun. Phys. 4, 108 (2021)
22. A. Stegmaier, S. Imhof, T. Helbig, T. Hofmann, C.H. Lee, M. Kremer, A. Fritzsche, T. Feichtner,
S. Klembt, S. Höfling, I. Boettcher, I.C. Fulga, L. Ma, O.G. Schmidt, M. Greiter, T. Kiessling,
A. Szameit, R. Thomale, Topological defect engineering and PT symmetry in non-Hermitian
electrical circuits, Phys. Rev. Lett. 126
23. R. Gowrishankar, D.A. Stewart, G.T. Martin, J.C. Weaver, Transport lattice models of heat
transport in skin with spatially heterogeneous, temperature-dependent perfusion. Biomed. Eng.
Online 3, 42 (2004)
24. I. Kokalari, T. Karaja, M. Guerrisi, Review on lumped parameter method for modeling the
blood flow in systemic arteries. J. Biomed. Sci. Eng. 6, 92–99 (2013)
25. R. Liu, X. An, H. Zheng, M. Wang, Z. Gao, E. Li, Neutralization line decoupling tri-band
multiple-input multiple-output antenna design. IEEE Access 8, 27018–27026 (2020)
26. L.Q. English, S.G. Wheeler, Y. Shen, G.P. Veldes, N. Whitaker, P.G. Kevrekidis, D.J.
Frantzeskakis, Backward-wave propagation and discrete solitons in a left-handed electrical
lattice. Phys. Lett. A 375, 1242 (2011)
27. T. Kuusela, Soliton experiments in a damped ac-driven nonlinear electrical transmission line.
Phys. Lett. A 167, 54–59 (1992)
28. B.A. Malomed, Propagating solitons in ac-driven chains. Phys. Rev. A 45, 4097–4101 (1992)
29. T. Kuusela, J. Hietarinta, B.A. Malomed, Numerical study of solitons in the damped ac-driven
Toda lattice. J. Phys. A 26, L21–L26 (1993)
30. A. Kundu, Landau-Lifshitz and higher-order nonlinear systems gauge generated from nonlinear
Schrödinger-type equations. J. Math. Phys. 25, 3433 (1984)
31. D. Qiu, J. He, Y. Zhang, K. Porsezian, The Darboux transformation of the Kundu-Eckhaus
equation. Proc. R. Soc. A 471, 20150236 (2015)
32. F. Yuan, D. Qiu, W. Liu, K. Porsezian, J. He, On the evolution of a rogue wave along the
orthogonal direction of the (x, t)-plane. Commun. Nonlinear Sci. Numer. Simul. 44, 445–457
(2017)
33. S. Xu, J. He, L. Wang, The Darboux transformation of the derivative nonlinear Schrödinger
equation. J. Phys. A 44, 305203 (2011)
34. Y. Zhang, L. Guo, S. Xu, Z. Wu, J. He, The hierarchy of higher order solutions of the derivative
nonlinear Schrödinger equation. Commun. Nonlinear Sci. Numer. Simul. 19, 1706–1722 (2014)
35. H.-H. Chen, C.-H. Liu, Solitons in nonuniform media. Phys. Rev. Lett. 37, 693–697 (1976)
36. H.H. Chen, Y.C. Lee, C.S. Liu, Integrability of nonlinear Hamiltonian systems by inverse
scattering method. Phys Scr. 20, 490 (1979)
Chapter 2
Nonlinear Schrödinger Models for
Solitons Propagation in 1D Lossless
Nonlinear Transmission Networks
(NLTNs)

Abstract In this Chapter, we consider the lossless 1D discrete dispersive NLTN


sketched in Fig. 2.1. When lattice effects are considered, the reductive perturbation
method in the semi-discrete limit can be used to show that the dynamics of modu-
lated waves can be modeled by PDEs of Schrödinger type, which describes the MI
and propagation of modulated waves on a continuous-wave (CW) background (also
referred to as the constant-amplitude state). Analytical solutions of these PDEs are
used to investigate the transmission of modulated nonlinear waves in the network
under consideration. Here, we also investigate the effects of the linear dispersive
element C S on the dynamics of modulated nonlinear waves. We show that the linear
dispersive parameter C S can be used to manipulate the motion of bright, dark, and
kink solitons in the network.

2.1 Introduction

Nonlinear Schrödinger type equations are models emerging from a wide variety of
fields and playing a significant role in many fields of science such as fluids, nonlin-
ear optics, the theory of deep water waves, plasma physics, nonlinear transmission
networks, nuclear physics, Bose–Einstein condensates, condensed matter physics,
and so on. The present chapter deals with the study of a modified NLTN of the
Noguchi type (which was introduced in [1]) depicted in Fig. 2.1. Such a model is
composed of N identical cells, each of which contains a linear inductor L1 shunted
by a linear capacitor C S in the series branch, and in the transverse branch, a non-
linear capacitor C shunted by another linear inductor L2 . The nonlinear capacitor
is voltage-dependent and is realized as a reverse-biased diode. This physical model
differs from the Hirota-Suzuki model [2] by the presence of the linear dispersive
capacitance C S . Built by Noguchi to experimentally study the propagation of the
first-order Korteweg-de Vries (KdV) solitons [1], the linear dispersive capacitance
C S was absent in its first version.
The physical model shown in Fig. 2.1 has been used by Ichikawa et al. [3] to predict
the realization of KdV-like solitons in it; it has been generalized by Yoshinaga et al.
[4] to experimentally study properties of solitons of the KdV type. Most recently,
© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2022 7
E. Kengne and W. Liu, Nonlinear Waves,
https://doi.org/10.1007/978-981-19-6744-3_2
8 2 Nonlinear Schrödinger Models for Solitons Propagation …

Fig. 2.1 Schematic representation of an elementary cell of the lossless dispersive nonlinear trans-
mission network with the linear inductor in the series branch, and the nonlinear capacitor in the
parallel branch. A linear capacitor (dispersive element) is connected in parallel to the linear inductor.
The line is composed of N identical cells. Reprint from Ref. [8], Copyright 2022, with permission
from American Physical Society

Pelap et al. [5] addressed the existence of solitary waves in this network system.
Some years ago, Marklund and Shukla [6] investigated by means of Kengne and Liu
model [7] the MI of partially coherent signals in the transmission lines shown in Fig.
2.1. The main purpose of this Chapter is to present various PDEs of Schrödinger
type that are used to model and to investigate the dynamics of nonlinear modulated
waves propagating in the network of Fig. 2.1. A particular attention will be paid to
the effects of linear dispersive capacitance C S on the soliton propagation.

2.2 Standard Nonlinear Schrödinger Equations for


Modulated Waves Propagation in a Lossless Electrical
Transmission the Network: Effects of the Dispersive
Element C S

The present Section deals with the study of dynamics of solitons in the modified
Noguchi lossless electrical transmission network displayed in Fig. 2.1. Employing
the reductive perturbation method in the semi-discrete limit, we establish that the
dynamics of modulated waves are described by one-dimensional NLS equations.
Based on these equations, the analytical results predict either two or four frequency
regions with different behavior concerning the MI. Applying the Darboux transform
to the found network equations, we produce exact bright and dark solitons embedded
in the CW background. Our investigation reveal that the physical system under con-
sideration may support the propagation of bright, dark, and kink soliton signals. Also,
we show that when a wave packet travels along the lattice, it experiences velocity
fluctuations which become more and more pronounced with the increase of lattice
2.2 Standard Nonlinear Schrödinger Equations… 9

effects. Through analytical exact solutions of the network equations, we establish


that the dispersion parameter C S can be used to control the motion of bright, dark,
and kink solitary waves propagating in the network. The methodology presented in
this Section offers an efficient tool for systematically investigating the dynamics of
solitons in 1D nonlinear transmission networks.

2.2.1 Model Equations

In order to study the effects of the dispersive element C S on nonlinear modu-


lated waves propagating in the network system of Fig. 2.1, we adopt the voltage-
capacitance dependence as
 
Q n (Vn ) = C0 Vn − αVn2 + βVn3 ,

C0 being the characteristic capacitance, and α and β are positive nonlinear coeffi-
cients. Applying the Kirchhoff’s laws to our network system yields the following
system of equations

 
d2   1 1 CS d2
Vn − αVn
2
+ βVn
3
+ Vn − + (Vn−1 − 2Vn + Vn+1 ) = 0, (2.1)
dt 2 C0 L 2 C0 L 1 C0 dt 2

n = 1, 2, ..., N . In the linear limit Vn ∼ exp [i (kn − ωt)], we derive the following
linear dispersion relation, similar to that of a typical passband filter:
     
2 k 1 4 2 k
ω C0 + 4C S sin
2
− + sin = 0. (2.2)
2 L2 L1 2

Throughout this Section, we focus ourselves to the network system of Fig. 2.1 with
linear dispersion element C S that satisfies the condition C S < (L 2 /L 1 ) C0 so that
the group velocity υg = dω/dk remains positive in the Brilloum zone 0 ≤ k ≤ π .
During the numerical simulations, the following network parameters will be used [9]:

L 1 = 28 μH, L 2 = 14 μH, C0 = 540 pF, V0 = 1.5 V, α = 0.16 V−1 , β = 0.0197 V−2 , (2.3)

with C S ∈ [0, C0 /2] pF.


For the network parameters 2.3, the linear dispersion curves of a typical passband
filter are represented in Fig. 2.2 for different values of coefficient C S . These dispersion
curves show (a) the propagating frequency f = ω/2π and (b) dispersion coefficient
Pl = d 2 ω/dk 2 as functions of the wavenumber k. It is clearly seen from the plots of
Fig. 2.2a that the frequency decreases when C S increases. Figure 2.2b reveals that for
any dispersion element C S ∈ [0, C0 /2], equation Pl (k) = 0 admits one and only one
solution k z ∈ [0, π ] that corresponds to the frequency f z = f (k z ), and, k z → 1/2
as C S → C0 /2 − 0.
10 2 Nonlinear Schrödinger Models for Solitons Propagation …

Fig. 2.2 Linear dispersion curves showing frequency f = ω/2π (a) and dispersion coefficient
Pl = d 2 ω/dk 2 (b) versus wavenumber k for different values of dispersive parameter Cs . The
network parameters used in these plots are given by Eq. (2.3). Reprint from Ref. [8], Copyright
2022, with permission from American Physical Society

In order to focus ourselves on waves with a slowly varying envelope in time


and space with regard to a given carrier wave with angular frequency ω = 2π f and
wavenumber k and use the reductive perturbation method, we introduce two slow
envelope variables as [10]

ξ = ε(n − υg t), τ = ε2 t, (2.4)

with a small parameter ε  1. Here, υg = dω/dk, ω and k being related by the linear
dispersion relation (2.2). Then, we seek solutions for system (2.1) in the general form

Vn (t) = εV01 (ξ, τ ) exp(iθ ) + ε2 [V02 (ξ, τ ) + V12 (ξ, τ ) exp(2iθ )] + c.c., (2.5)

where θ = kn − ωt is the rapidly varying phase and c.c. stands for the complex
conjugation. In the general form (2.5), the term proportional to ε2 is included in the
dc part to take into account the asymmetry of the charge-voltage relation.
Substituting ansatz (2.5) into the nonlinear discrete system (2.1) and neglecting
terms of higher order in ε, one arrives, after keeping up to the second-order derivatives
of Vn (t) (this is done to secure the balance of the dispersion and nonlinearity), to the
following results

V02 (ξ, τ ) = − α02 |V01 |2 , (2.6a)


αω2
V12 (ξ, τ ) = 12 (V01 ) ,
2
(2.6b)
i ∂∂τ + P ∂∂ξV201 + Q |V01 |2 V01 = 0.
2
V01
(2.6c)
2.2 Standard Nonlinear Schrödinger Equations… 11

Here,
 
2L 1 C0 α 2 ωυg2 1 4L 1 C S ω2 − 1 sin2 k
02 = , 12 = ω − 2
+ , (2.7a)
1 − L 1 C0 υg2 4L 2 C0 L 1 C0
 
υg2 4C S k 1 CS CS
P =− 1+ sin2 + − ω cos k − 2 υg sin k,
2ω C0 2 2L 1 C0 ω 2C0 C0
(2.7b)
3β α 2 ω3
Q = ω + 02 − . (2.7c)
2 12

As we can see from Eqs. (2.6a), (2.6b), (2.7a), and (2.7c), the two last terms in the
expression for nonlinear coefficient Q of the NLS equation (2.6c) come from the ε2
terms in the general expression (2.5) for Vn (t). From expression (2.7b), it is seen that
the ε2 terms does not affect the dispersion coefficient P. This means that, assuming
a solution in the form of (2.5) without ε2 terms, one arrives at another NLS equation
similar to Eq. (2.6c), except for the nonlinear coefficient which is

Q = 3βω/2. (2.8)

The above methodology shows that one may obtain two NLS equations with the
same dispersion coefficient P, but different nonlinear coefficients Q, allowing the
investigation of the dynamics of solitons in the network of Fig. 2.1. If we apply now
the MI criterion for the NLS equation [11], we can conclude that a uniform wavetrain
propagating along the network of Fig. 2.1 is unstable under modulation for P Q > 0,
and remains stable for P Q < 0. In the situation when the nonlinearity parameter Q of
the NLS equation is given by Eq. (2.8), product P Q and P have the same sign as soon
as ω is always positive; this behavior can be seen from Fig. 2.3b. Now, considering
that the nonlinear coefficient Q is given by Eq. (2.7c), one of the three scenarios
shown in Fig. 2.3 occurs, depending on the value of dispersive parameter C S . Indeed,
max Q(k) = Q max (C S ) is an increasing continuous function of C S , with Q max (0) <
[0,π]
0 < Q max (C0 /2). Therefore, there exists a unique value, C S = C S0 ∈ [0, C0 /2] that
satisfies the relationship Q max (C S0 ) = 0. For every C S < C S0 , one has Q (k) < 0,
while for every C S > C S0 , equation Q(k) = 0 has two solutions, kq1 and kq2 , in
interval [0, π ]. Different shapes of the region of the MI by the wave solution of the
NLS equation (2.6c) with the nonlinear coefficient (2.7c) are displayed in Fig. 2.3.
In Fig. 2.3d–f, f k shows the value of the frequency for the wavenumber k = k z that
satisfies the relationship P(k z ) = 0, that is, f z = f (k z ); in these same figures, f q1 and
f q2 are frequencies associated with Q(k) = 0. Figure 2.3d shows two MI regions,
the one with f ∈ [ f min , f z ] corresponding to P Q < 0, hence the modulationally
stable CW and hole (dark) solitons, and region f ∈ [ f z , f max ] corresponding to
P Q > 0, that is, MI and bright envelope solitons. In Fig. 2.3e, three regions of
the modulational instability can be observed, the one f ∈ [ f min , f z ] corresponding
to P Q < 0 (modulational stability and dark solitons), and two regions f ∈ [ f z ,
f q1 ] ∪ [ f q1 , f max ] corresponding to P Q > 0 (MI and envelope solitons). From Fig.
12 2 Nonlinear Schrödinger Models for Solitons Propagation …

Fig. 2.3 Linear dispersion curves (bottom) and plots of the product P Q (top) versus wavenumber
k, showing different scenarios of the MI of solutions of the NLS equation (2.6c) with nonlinear
coefficient ( 2.7c). a, d Product P Q and frequency f = ω/2π for C S = C0 /8 < C S0 ; b, e Product
P Q and frequency f = ω/2π for C S = C S0 ≈ C0 /5.87; c, f: Product P Q and frequency f =
ω/(2π ) for C S = C0 /5 > C S0 . To generate these plots, parameters (2.3) are used. Reprint from Ref.
[8], Copyright 2022, with permission from American Physical Society

2.3f, we can clearly see four regions of MI: two regions of f ∈ [ f min , f z ] and f ∈
[ f q1 , f q2 ], corresponding to P Q < 0 (modulational stability and hole solitons), and
two regions f ∈ [ f z , f q1 ] and f ∈ [ f q2 , f max ] corresponding to P Q > 0 (once again,
this implies the presence of the MI and bright solitons).
In Fig. 2.4, we depict the product P Q versus the wavenumber k for different
values of dispersion parameter C S . Plots of Fig. 2.4a correspond to the NLS equation
(2.6c) with nonlinear coefficient (2.8), while Fig. 2.4b and c are obtained with the
nonlinear coefficient Q given by Eq. (2.7c) for C S < C S0 and C S > C S0 , respectively. It
is clearly seen from Figs. 2.2a and 2.4a, that the region of negative P Q (modulational
stability) for NLS equation (2.6c) with nonlinear coefficient (2.8) enhances when the
dispersion parameter C S increases. From Figs. 2.2a and 2.4b, we can see that the
region of negative P Q (i.e., the region of the modulational stability) of NLS equation
(2.6c) with nonlinear coefficient (2.7c) shrinks when the dispersion parameter C S
increases, being capped by C S < C S0 . As one can seen from Figs. 2.2a and 2.4c, using
nonlinear coefficient (2.7c), the modulational stability region (P Q < 0) shrinks for
low frequencies and enhances for
higher frequencies,
when C S > C S0 increases. In
other words, the size of interval f min , f z decreases ( f z is getting smaller) while
that of interval [ f q1 , f q2 ] increases ( f q1 decreases and f q2 increases) when C S > C S0
increases.
2.2 Standard Nonlinear Schrödinger Equations… 13

Fig. 2.4 Plots of product P Q showing the regions of positive and negative values for different
values of the dispersion parameter C S . a Plots of product P Q with nonlinear coefficient Q given
by Eq. (2.8) for different C S . b The plot of product P Q generated with nonlinear coefficient (2.7c)
for different parameters C S < C S0 ≈ C0 /5.87. c The plot of product P Q generated with nonlinear
coefficient (2.7c) for different parameters C S > C S0 ≈ C0 /5.87. These plots are generated with
parameters (2.3). Reprint from Ref. [8], Copyright 2022, with permission from American Physical
Society

2.2.2 Bright and Dark Single-Solitary Waves Propagating in


the 1D Lossless Network of Fig. 2.1

With the use of the NLS equations (2.6c) with the nonlinear coefficient Q given by
Eq. (2.7c) and by Eq. (2.8), we here aim to study analytically the propagation of
bright and dark solitary waves through the network of Fig. 2.1. For this purpose, we
first find analytical soliton like solutions of the NLS equation (2.6c).

2.2.2.1 Analytical Solitonlike Solutions of the NLS Equation

Here, we focus ourselves on solutions on top of a CW background,



V01c (ξ, τ ) = (iφcb ) , if P Q > 0
Acb exp
−Acd − 2P
Q
exp (iφcd ) , if P Q < 0,

with

Q Q 2  A2 Q 2 + 2P 2 kcd
2
φcb = kcb ξ+ 2 Acb − kcb
2
τ, φcd = kcd ξ − cd τ + δ0 ;
2P 2 2P
14 2 Nonlinear Schrödinger Models for Solitons Propagation …

here, Acb , Acd , kcb , kcd , and δ0 are arbitrary real constants. If we apply the Darboux
transformation [12] for the NLS equation (2.6c) to the CW background V01c (ξ, τ ),
we obtain the following antidark and dark soliton solutions

d1 cosh θ + cos ϕ d2 sinh θ + d3 sin ϕ
V01 (ξ, τ ) = Acb + Asb + i Asb
cosh θ + d1 cos ϕ cosh θ + d1 cos ϕ
× exp (iφcb ) , (2.9a)

Q
V10 (ξ, τ ) = − − [Acd + i Asd tanh ζ ] exp (icd ) , (2.9b)
2P

respectively. It is important to notice that “antidark” are solitons of the bright type
sitting on top of a flat background. Different parameters appearing in solutions (2.9a)
and (2.9b) are given as

Q [(kcb + ksb ) M R − Asb M I ]
θ = MR ξ− Qτ − θ0 ,
2P 2

Q [(kcb + ksb ) M I + Asb M R ]
ϕ = MI ξ− Qτ − ϕ0 ,
2P 2
 
Q Asd Acd Q 2 − 4P 2 kcd2
ζ = Asd ξ+ τ + ζ0 ,
2P  2P

Q 2 A2sd + A2cd + 2kcd 2
P2
cd = kcd ξ − τ + δ0 ,
2P
Acb (M R − Asb ) Acb (ksb − kcb + M I ) D − 2 A2cb
d1 = , d2 = , d3 = ,
D D D
(M R − Asb )2 + (ksb − kcb + M I )2
D = A2cb + ,
4
M= (−Asb + iksb − ikcb )2 − 4 A2cb = M R + i M I .

Here, Asb , Asb , θ0 , ϕ0 , ksb , ζ0 , and ksd are arbitrary real parameters. The antidark
soliton solution (2.9a) and the dark-soliton solution (2.9b) carry over into the CW
background V01c (ξ, τ ) when Asb = 0 and when Asd = 0, respectively (the modula-
tional stability of the background is considered below). On the other hand, for the
vanishing CW background, i.e., when Acb = 0, and when Asd = 0, they respectively,
turn into the following bright soliton solution and the kink solution
  
Q  2 
Asb exp i ksb 2P ξ − ksb − A2sb Qτ
2 − ϕ 0
V01 (ξ, τ ) =    , (2.10a)
Q
cosh Asb 2P ξ − ksb Qτ − ξ0
    
Q Q Asd A2 Q 2
V10 (ξ, τ ) = −i Asd − tanh ξ − ξ0 exp i − sd τ − τ0 , (2.10b)
2P 2P 2P
2.2 Standard Nonlinear Schrödinger Equations… 15

ξ0 and τ0 being arbitrary real constants. This means that solutions (2.9a) and (2.9b)
produce solitary waves embedded into the CW background.
Substituting expressions (2.9a) and (2.9b) into Eqs. (2.6a) and (2.6b) leads to
analytical expressions for V02 (ξ, τ ) and V12 (ξ, τ ). If we then insert the expressions
of V10 (ξ, τ ), V02 (ξ, τ ) and V12 (ξ, τ ) in the general form (2.5) for voltage Vn (t) and
going back to the original coordinate n and time t by means of Eq. (2.4), we will
obtain the analytical expressions for Vn (t). These results for Vn (t) can be used to
investigate analytically the dynamics of solitons in the network system of Fig. 2.1.
Next, we separately discuss the dynamics of solitons in the MI region (envelope
solitons given by Eq. (2.9a)) and in the region of the modulational stability (hole
solitons defined by Eq. (2.9b)).

2.2.2.2 Dynamics of Solitons in the Regions of Positive P Q


(Modulational Instability)

Using the zeros of functions P = P(k) and Q = Q(k), the regions of the MI for
the NLS equation (2.6c) are displayed in Fig. 2.3 for a given set of the network
parameters, with nonlinear coefficient (2.7c) or (2.8).
In the regions of MI, that is, the regions of positive P Q, the solitons are enve-
lope modes propagating on top of the CW background, defined by expression (2.9a).
Without the loss of generality, we consider here the NLS equation (2.6c) with disper-
sion coefficient (2.7b). First of all, we notice that the situation with 4 A2cb − A2sb = 0
leads to a constant solution. In what follows, we focus on three special cases, viz.,
the bright soliton with a vanishing CW background (Acb = 0), the case of ksb = kcb
and A2sb − 4 A2cb < 0, and, finally, the one with ksb = kcb and A2sb − 4 A2cb > 0.
Case Acb = 0 By taking Acb = 0, solution (2.9a) of NLS equation (2.6c) for positive
P Q (i.e., solution (2.10a)) can be written as follows:

Asb
V10 (ξ, τ ) =    exp (i) , (2.11)
cosh Asb Q
2P
ξ − ksb Qτ − ξ0

where   2 
Q ksb − A2sb Q
(ξ, τ ) = ksb ξ− τ − ϕ0 ,
2P 2

ϕ0 and ξ0 being two arbitrary real constants. Expression (2.11) is an envelope


soliton solution of the NLS equation (2.6c) when P Q > 0. Analyzing solution
(2.11), it
√ appears that the coordinate of the center of the envelope soliton is
η(τ ) = 2P/Q (ksb Qτ + ξ0 ) and the width of the envelope soliton is inversely
proportional to (Q/2P)1/2 , meaning that one can use Eq. (2.11) to describe the com-
pression of envelope solitons when (Q/2P)1/2 increases with the variation of fre-
quency f = ω/2π . It is obvious that for P(k)Q(k) > 0, one has (Q/2P)1/2 → +∞
as k → k z ; therefore, for a given set of parameters, the width of the envelope soliton
16 2 Nonlinear Schrödinger Models for Solitons Propagation …

Fig. 2.5 The wave propagation in the region of low frequencies (top) and higher frequencies
(bottom), showing the extension and compression of envelope solitons, respectively. These plots are
obtained for parameters (2.3) with C S = C0 /5 and ε = 10−3 , Asb = 10, ksb = −0.6, ξ0 = −10,
ϕ0 = 0. The frequencies are a f = 2.018 MHz, b f = 2.047 MHz, c f = 2.077 MHz, d f = 2.247
MHz, e f = 2.265 MHz, f f = 2.281 MHz. Reprint from Ref. [8], Copyright 2022, with permission
from American Physical Society

shrinks as f (k) is approaching f z = f (k z ) for P Q > 0. Also, it is seen that the


center of the soliton moves at a constant velocity, dη/dτ = ksb Q (2P/Q)1/2 .
To demonstrate the dynamics of envelope soliton propagating in the network of
Fig. 2.1, we use parameters (2.3) with C S = C0 /5. In this case, the region of positive
P Q (that is, the region of the existence of envelope solitons) are (in MHz) intervals
of low frequencies, f ∈ [1.999, 2.1125], and of higher frequencies, f ∈ [2.231,
2.3631]. Function (Q/2P)1/2 decreases for f ∈ [1.999, 2.1125] and increases for
f ∈ [2.231, 2.3631]. Using these parameters, we depict in Fig. 2.5 the evolution of
voltage Vn (t) at cell n = 1 for different frequencies. Figure 2.5a–c and d–f show the
pulses propagating, respectively, with low and high frequencies. As we can see from
Fig. 2.5a–c, the width of the envelope solitons increases with the frequency; Fig.
2.5d–f reveal that the width decreases as the frequency increases. These findings are
explained by the fact that quantity (Q/2P)1/2 (which determines the compression
or expansion of the solitons) is a decreasing/increasing function of the propagating
frequency f in the region of low/high frequencies.
Case ksb = kcb and A2sb − 4 A2cb < 0 In the present case, we can rewrite solution
(2.9a) in the following form:
⎛ ⎞
−Asb cosh θ + 2 Acb cos ϕ + i 4 A2cb − A2sb sinh θ
V01 (ξ, τ ) = ⎝ Acb + Asb ⎠
2 Acb cosh θ − Asb cos ϕ

   
Q Q 2 
× exp i kcb ξ+ 2 Acb − kcb
2
τ , (2.12)
2P 2
2.2 Standard Nonlinear Schrödinger Equations… 17
√
with θ = (1/2)Asb 4 A2cb − A2sb Qτ − θ0 and ϕ = 4 A2cb − A2sb Q/2Pξ −
kcb Qτ ) − ϕ0 ; here, θ0 and ϕ0 are arbitrary real parameters.From the expressions
for θ and ϕ and from transformation (2.4), we can see that solution (2.5) associated
with solution (2.12) is periodic with the period

2π 2P
N =  
ε Q 4 A2cb − A2sb

in the spatial coordinate n and aperiodic in time t, if the wavenumber of the CW back-
ground is chosen as kcb =  k 4 A2cb − A2sb , where k is an integer. Any wavenumber

kcb =  k 4 A2cb − A2sb will lead to an aperiodic solution in both time and space.
We display in Fig. 2.6 the evolution of the MI corresponding to solution (2.12)
for the network parameters (2.3) with C S = C0 /5. Figure 2.6a–c and d–f show the
transmission of the soliton-like signals through the network at frequencies f = 2.018
MHz and f = 2.256 MHz, respectively. Therefore, the envelope solitons in Fig.
2.6a–c propagate at the frequency taken from the MI region [ f min , f z ], while the
pulses of Fig. 2.6g, h are associated with the frequency taken from the MI region,
[ f q2 , f max ]. Figure 2.6g, h show solitons propagating at different frequencies taken
from the same MI region; here, it is clearly seen that the CW background becomes
unstable. From Fig. 2.6, we can conclude that the choice of the propagating frequency
in the MI region affects the soliton propagation. Figure 2.6 also reveals that, with the
best choice of solution parameters, the solution is periodic in time, with the period
increasing with the underlying frequency. This can be clearly seen in Fig. 2.6c, f, as
well as in Fig. 2.6g, h. Moreover, Fig. 2.6g, h reveal that the envelope soliton features
different behavior at frequencies taken from the low- and high-frequency regions of
the MI. We conclude from Fig. 2.6g and h that the temporal period of the wave is
related to the propagating frequency f of the envelope soliton: In the low-frequency
MI region, the period increases with f , while in the high-frequency region the period
decreases.
Case ksb = kcb and A2sb − 4 A2cb > 0 In the special case when ksb = kcb and A2sb −
4 A2cb > 0, one can compute coefficients of solution (2.9a) and obtain the solution in
the form
⎛  ⎞
Asb − 4 Acb
2 2
Asb − 4 Acb cos ϕ − i Asb sin ϕ
2 2
⎜ ⎟
V01 (ξ, τ ) = ⎜
⎝ −A cb + ⎟

Asb cosh θ − 2 Acb cos ϕ

   
Q Q 2 
× exp i kcb ξ+ 2 Acb − kcb
2
τ , (2.13)
2P 2
18 2 Nonlinear Schrödinger Models for Solitons Propagation …

Fig. 2.6 The evolution of the MI associated with Eq. (2.12) for parameters (2.3) with dispersion
coefficient C S = C0 /5 and the frequency taken from the lower region, f = 2.018 MHz (top), and
from the higher region, f = 2.273 MHz (middle). a, d: The spatiotemporal evolution of the voltage
signal; b, e: The spatial profile of the signal at t = 50; c, f The temporal evolution of the voltage
at cell n = 350; g: The evolution of the voltage (at cell n = 350) with frequencies taken from the
MI region [ f min , f z ] ( f = 2.018 MHz ⇐⇒ k = 0.90 and f = 2.047 MHz ⇐⇒ k = 1.05); h The
evolution of the voltage (at cell n = 350) with frequencies taken from the MI region [ f g2 , f max ]
( f = 2.256 MHz ⇐⇒ k = 1.90 and f = 2.273 MHz ⇐⇒ k = 2.0). The curves are generated
with parameters ε = 10−3 , Asb = 10−4 , Acb = Asb /1.8, kcb = ksb = 2 4 A2cb − A2sb , ϕ0 = 0,
and θ0 = −10−8 . Reprint from Ref. [8], Copyright 2022, with permission from American Physical
Society


with θ = A2sb − 4 A2cb Q
2P
ξ − ksb Qτ − θ0 and ϕ = − 21 Q Asb A2sb − 4 A2cb τ −
ϕ0 , where θ0 and ϕ0 are two arbitrary real constants. Solution (2.13) is aperiodic in
both space and time. Going back to the original variables n and t, one concludes
that the motion of the envelope soliton is aperiodic in both n and t, as shown in
Fig. 2.7. It is seen from Fig. 2.7c and f that the envelope solitons feature different
behavior when propagating at frequencies taken from the low- and high-frequency
MI regions: For the solitons propagating in the low-frequency region, waves with a
smaller frequency reach their peak earlier than ones with a higher frequency. The
opposite situation is observed for the solitons in high-frequency MI region.
2.2 Standard Nonlinear Schrödinger Equations… 19

Fig. 2.7 The evolution of envelope solitons on top of the CW background given by Eq. (2.13)
−3 −4, A = A /3, k =
the same network as in Fig. 2.6 with parameters ε = 10 , Asb = 10
for cb sb cb
2 Asb − 4 Acb , ϕ0 = 0, θ0 = −4. a, d Spatial profile of bright solitons at time t = 50 s propagating
2 2

at frequencies f = 2.018 MHz (a) and f = 2.256 MHz (d); b, e The evolution of the envelope
soliton at cell n = 350 with frequencies f = 2.018 MHz (b) and f = 2.256 MHz (e); c Voltages at
n = 350 as function of time t for different frequencies taken from the MI region of low frequencies,
[ f min , f z ] (c), and from the MI region of high frequencies, [ f g2 , f max ] (f). Reprint from Ref. [8],
Copyright 2022, with permission from American Physical Society

2.2.2.3 Dynamics of Solitons in the Region of the Negative PQ


(Modulational Stability Region)

Employing now solution (2.9b), we address here the dynamics of solitons propagating
in the network at frequencies taken from the regions of the modulational stability. As
we have already mentioned, the analytical solution (2.9b) describes the propagation of
dark (hole) solitons on the non-vanishing CW background V01c (ξ, τ ). √ The analytical
expression (2.9b) shows that the soliton amplitude is proportional to −Q/(2P).
Firstly, it is important to notice that for vanishing CW background, i.e., when
Acd = 0, solution (2.9b) is associated with the propagation of a kink soliton through
the network, as it is shown in Fig. 2.8, which is produced by the NLS equation (2.6c)
with the nonlinearity coefficient (2.7c). Figure 2.8 shows the temporal profile of
stable kink solitary waves at the 350-th cell at low and high frequencies (left and right
panels, respectively). Plots of the top, middle, and bottom panels correspond to the
network with the dispersion parameter C S = C0 /5, C S = C0 /4.5 and C S = C0 /4.1,
respectively. The left and right panels of Fig. 2.8 show that, independently of the
value of C S , the soliton’s speed decreases as the propagating frequency increases.
Secondly, in the non-vanishing CW background (i.e., at Acb = 0), solution (2.9b)
leads to the propagation of dark solitons embedded in the CW background. Thus,
expression (2.9b), may lead to the dark soliton solution of Eq. (2.1) which corre-
sponds to waves propagating on a CW background. An example of the propagation
of the dark soliton is shown in Fig. 2.9. The top and the bottom panels of Fig.
2.9 show respectively dark solitons propagating at different frequencies, taken from
20 2 Nonlinear Schrödinger Models for Solitons Propagation …

Fig. 2.8 The propagation of stable kink solitons through the network for the system’s parameters
(2.3) with ε = 10−3 and solution parameters ksd = π/5, Asd = −10−2 , ζ0 = 2, and δ0 = 0. The
top panel: C S = C0 /5; the middle panel: C S = C0 /4.5; the bottom panel: C S = C0 /4.1. Reprint
from Ref. [8], Copyright 2022, with permission from American Physical Society

the low-frequency region [ f min , f z ] of the modulational stability, and dark solitons
propagating in the high-frequency region of the modulational stability, i.e., [ f q1 ,
f q2 ]. Different plots of Fig. 2.9 are generated with the use of the solution parameters
kcd = 0.1, Acd = −10−4 , ksd = 0.35, Asd = 3 × 10−4 , δ0 = 0, and ζ0 = −0.532.

2.3 Modulational Instability and Transmission of Chirped


Femtosecond Signals Through the Lossless Electrical
Network of Fig. 2.1

In this Section, we employ a cubic-quintic nonlinear Schrödinger equation with self-


steepening and self-frequency shift terms to study the MI phenomenon and to show
2.3 Modulational Instability and Transmission… 21

Fig. 2.9 The signal voltage (measured in volt) at cell n = 350 as a function of time t (in ms).
The temporal profile of dark solitons propagating at different frequencies (in MHz) for parameters
(2.3) with C S = C0 /5 and ε = 10−3 . a f = 1.837, b f = 1.854, c f = 1.882, d f = 2.119, e
f = 2.146, f f = 2.171. Reprint from Ref. [8], Copyright 2022, with permission from American
Physical Society

that the competing cubic–quintic nonlinearity induces propagating chirped femtosec-


ond pulses, signals of first-order rogue wave type, and two sister modulated nonlinear
wave signals along the modified Noguchi electric network of Fig. 2.1. We show the
relationship between the nonlinear chirp and the intensity of the corresponding sig-
nal. We introduce the concept of two sister signals and show how the corresponding
chirps are related. Our findings reveal that chirping’s features and behavior depend on
both the self-steepening and self-frequency shift, while its amplitude can be managed
by varying the parameters of the group velocity dispersion and cubic nonlinearity.

2.3.1 Introduction

Femtosecond pulses have been extensively investigated for the past few decades, due
to their wide applications in many different physical systems such as ultrahigh-bit-
rate optical communication systems [13], ultrafast physical processes [14], infrared
time-resolved spectroscopy [15], and optical sampling systems [16]. Nonlinear
partial differential equations with higher-order nonlinear and dispersive terms of
Schrödinger types have been used to explain a variety of effects in the propagation
of pulses in such systems. These equations in general and the NLS models in partic-
ular are generally not completely integrable and cannot be solved exactly by using
traditional methods [17]. Therefore, these models do not have soliton solutions in
general. However, they can admit solitary-wave solutions. The generation and prop-
agation of solitons in nonlinear systems result, as we well-know, generally from the
phenomenon of the MI [18, 19]. The phenomenon of the modulational instability
22 2 Nonlinear Schrödinger Models for Solitons Propagation …

appears in many nonlinear dispersive systems and indicates that due to the interplay
between nonlinearity and the dispersive effects, a small perturbation on the envelope
of a plane wave may induce an exponential growth of the wave amplitude, resulting
in the carrier-wave breakup into a train of localized waves [20].
The main purpose of this Section is the analytical transmission of chirped fem-
tosecond soliton signals and periodic pulse trains through the modified Noguchi
nonlinear electric transmission network system of Fig. 2.1. Some years ago, Liu
and Kengne [10] presented a model for wave propagation on the discrete electrical
transmission network of Fig. 2.1 based on the generalized cubic-quintic derivative
nonlinear Schrödinger (GCQD-NLS) equation, derived in the small amplitude and
long wavelength limit using the standard reductive perturbation technique and com-
plex expansion on the governing nonlinear equations. This GCQD-NLS equation,
also referred to as the cubic-quintic nonlinear Schrödinger equation with derivative
terms, reads

∂ψ ∂ 2ψ ∂ψ ∗ ∂ψ
i + P 2 + Q 3 |ψ|2 ψ + Q 5 |ψ|4 ψ + iχ1 ψ 2 + iχ2 |ψ|2 = 0.
∂t ∂x ∂x ∂x
(2.14)
Here, ψ(x, t) is the complex envelope of the electric field, t and x are respectively
the temporal and the spatial coordinates, and P, Q 3 , Q 5 , χ1 , and χ2 coefficients
expressed in terms of the network parameters. Throughout this Section, we will
mostly use the following equivalent form of Eq. (2.14)

∂ψ ∂2ψ ∂  2  ∂  2
i + P 2 + Q 3 |ψ|2 ψ + Q 5 |ψ|4 ψ + i Q 13 |ψ| ψ + i Q 14 ψ |ψ| = 0,
∂t ∂x ∂x ∂x
(2.15)
where Q 13 = χ2 − χ1 and Q 14 = 2χ1 − χ2 . In the context of NLS models, P is
the group velocity dispersion (GVD) parameter, Q 3 and Q 5 represent respectively
the cubic or Kerr nonlinearity and the quintic or non-Kerr nonlinearity, while Q 13
and Q 14 are respectively the self-steepening and the self-frequency shift coefficients
which arise from stimulated Raman scattering. In cases of self-focusing (Q 3 > 0),
self-defocusing (Q 3 < 0), and zero Kerr nonlinearities, cubic nonlinearity Q 3 can
be normalized to 1, −1, and 0. The nonlinearity that arises because of the fifth-order
susceptibility can be obtained in many optical materials such as semiconductor-doped
glasses, semiconductors, polydiacetylene toluene sulfonate, AlGaAs, chalcogenide
glasses, and some transparent organic materials [21]. The cubic term |ψ|2 ψ in the
context of fiber optics is often referred to as self-phase modulation and its coefficient
Q 3 can be scaled out to the GVD term, which will be termed normal (anomalous)
dispersion depending on whether Q 3 < 0 (Q 3 > 0) [21]. Arising from an intensity-
dependent group velocity, the self-steepening of an optical fiber pulse produces a
temporal pulse distortion and an asymmetry in the pulse spectrum [22, 23]. The last
term of Eq. (2.15) related to Q 14 incorporates the intrapulse Raman scattering and
originates from the delayed Raman response [note that for model (2.15), the intrapulse
Raman scattering coefficient is Q 13 + Q 14 = χ1 ], which causes a self-frequency
shift. The self-frequency shift is the main driving mechanism of several wavelength
2.3 Modulational Instability and Transmission… 23

conversion and supercontinuum generation processes [24]. The self-frequency shift


enhances the effects introduced by self-steepening and also shifts the frequency of
the pulse. In the mathematical point of view, the chirping, δω(x, t), of a wave (wave
solution of the NLS equation (2.15)) of amplitude ρ(x, t) and phase φ(x, t) is the
negation of the spatial derivative of phase φ, that is, δω(x, t) = −∂φ(x, t)/∂ x.
In the following, we employ Eq. (2.15) which combines the cubic, quintic and
cubic derivative nonlinearities in the description of the propagation of chirped fem-
tosecond pulses along the modified Noguchi electric network of Fig. 2.1. Here, we
intend to develop via Eq. (2.15) an analytical framework to show that the lossless
network of Fig. 2.1 supports the propagation of (1) chirped solitonlike signals and
chirped first-order rogue wave signals embedded on non-vanishing CW background
and (2) special solitonlike and periodic nonlinear modulated wave pulses named
“two sister modulated nonlinear waves”, with chirping that varies simultaneously as
directly and inversely proportional to the intensity of the wave with nonzero bound-
ary conditions. Further, we show that the amplitude of chirping can be managed by
varying the self-steeping and self-frequency shift terms.

2.3.2 Modulational Instability and Evolution of Chirped


Femtosecond Solitary Signals Embedded on a
Non-vanishing CW Background

Here, we focus ourselves on the MI and the transmission through the network of
Fig. 2.1 of femtosecond solitonlike pulses, propagating on a non-vanishing CW
background with nonlinear chirp. In order to build such pulses and investigate their
MI, we start by performing the Madelung transformation [25, 26]

ψ(x, t) = φ(x, t) exp [iθ (x, t)] , (2.16)

where φ(x, t) is the new unknown complex amplitude and θ (x, t) is a real function,
relating with φ(x, t) as follows
∂θ(x, t)
= α10 |φ(x, t)|2 , (2.17a)
∂x
 ∗
∂θ(x, t) ∂φ(x, t) ∂φ (x, t)
= iα12 φ ∗ (x, t) − φ(x, t) + α13 |φ(x, t)|4 ; (2.17b)
∂t ∂x ∂x

in Eq. (2.17b), φ ∗ (x, t) is a the complex conjugate of φ(x, t) and α10 , α12 , and α13
are three real parameters given in terms of the network parameters as

3Q 13 + 2Q 14 Q 13 + 2Q 14 16P Q 5 + 3Q 213 − 4Q 14 (Q 13 + Q 14 )
α10 = − , α12 = − , α13 = .
4P 4 16P
(2.17c)
Replacing Eq. (2.16) into Eq. (2.15) and employing the relationships (2.17a), (2.17b)
and (2.17c), we obtain the following cubic NLS equation for φ :
24 2 Nonlinear Schrödinger Models for Solitons Propagation …

∂φ ∂ 2φ
i + P 2 + Q 3 |φ|2 φ = 0. (2.18a)
∂t ∂x
If we set  
2P 2P
φ= ± u= κ0 u and T = 2κ0 Pt,
Q3 Q3

with κ0 = sign[P Q 3 ] , Eq. (3.63) under the condition P Q 3 = 0 will take the form

∂u κ0 ∂ 2 u
i + + |u|2 u = 0. (2.19a)
∂T 2 ∂x2
Equation (2.18a) (or its equivalent form (2.19a)) is the standard NLS equation.
As we well-know, Eq. (2.19a) admits envelope (hole) solitons if P Q 3 > 0, i.e., κ0 =
+1, (P Q 3 < 0, i.e, κ0 = −1) [27]. Its soliton solutions obtained in the vanishing
boundary conditions are known. In order to obtain our results in terms of the network
parameters, our preference of the equation to be used here is Eq. (2.18a) (remember
that the coefficients of Eq. (2.15) are expressed, according to Ref. [10], in terms of
the network parameters). To study the phenomenon of the MI, we need to find exact
solutions of Eq. (2.18a) on the non-vanishing CW background
 2 
φcw (x, t) = Acw exp [i cw ] , cw = kcw x − Pkcw − Q 3 A2cw t, (2.20)

where Acw and kcw are two arbitrary real constants. For seeking such exact solutions,
we distinguish two cases, P Q 3 > 0 and P Q 3 < 0.

(i) Case of Envelope Solitons: P Q 3 > 0

In the case of positive P Q 3 , we use solution (2.20) as the seed solution and seek for
the solution solutions of the NLS equation (2.18a) in the form [28]
⎛ ⎞
d cosh [ϑ] + cos [ϕ]
A
⎜ cw + A
cosh [ϑ] + d cos [ϕ] ⎟
sol
φ(x, t) = ⎜

⎟ exp [i cw ] . (2.21)
b sinh [ϑ] + c sin [ϕ] ⎠
+i Asol
cosh [ϑ] + d cos [ϕ]

Here, cw is given by Eq. (2.20), d, b, and c are real parameters to be determined,


and ϑ = ϑ(x, t) and ϕ(x, t) are two real functions to be determined. Substituting
Eq. (2.21) into Eq. (2.18a) and removing exponential terms lead to an algebraic-
differential system having as solutions the following

Q3 M R Q 3 [(kcw + ksol ) M R − Asol M I ]


ϑ(x, t) = x− t − ϑ0 ,
2P 2

Q3 MI Q 3 [(kcw + ksol ) M I + Asol M R ]


ϕ(x, t) = x− t − ϕ0 ,
2P 2
2.3 Modulational Instability and Transmission… 25

Fig. 2.10 Spatiotemporal plots showing the propagation of bright solitary pulses on cw background,
according to the exact soliton solution given by Eq. (2.24b) and the corresponding frequency chirp
defined in Eq. (2.25a) for P = 0.5, Q 3 = 1, Q 13 = −0.5, Q 14 = 1/8, with Asol = 1, ksol = 0.2,
and ϑ0 = ϕ0 = 0. a, d Acw = 0, kcw = 0; b, e Acw = 2, kcw = 0.2; c, f Acw = 0.25, kcw = 0.2.
Reprint from Ref. [19], Copyright 2022, with permission from Elsevier

Acw (ksol − kcw + M I ) D − 2 A2cw Acw (M R − Asol )


b= ,c = ,d = ,
D D D

4 A2cw + (M R − Asol )2 + (ksol − kcw + M I )2


D = , (2.22)
4
 2
A2sol − 4 A2cw − (ksol − kcw )2 + (ksol − kcw )2 + 4 A2cw − A2sol + 4 A2sol (kcw − ksol )2
2
MR = ,
2

 2
(ksol − kcw )2 + 4 A2cw − A2sol + (ksol − kcw )2 + 4 A2cw − A2sol + 4 A2sol (kcw − ksol )2
M I2 = ;
2

here, ϑ0 and ϕ0 are two arbitrary real constants.

(ii) Case of Hole Solitons: P Q 3 < 0

When P Q 3 is negative, the NLS equation (2.18a) admit the exact solution

φ(x, t) = (Acw + i Asol tanh [ξ(x, t)]) exp [i dr ] , (2.23a)


    
Q3 Q3 Q3
ξ(x, t) = ± − x + 2P Asol − k1 + Acw − t, (2.23b)
2P 2P 2P
 2 
dr = k1 x − Pkcw − Q 3 A2cw t, (2.23c)

where k1 is an arbitrary real constant.


26 2 Nonlinear Schrödinger Models for Solitons Propagation …

Replacing now Eqs. (2.21) and (2.23a) in Eq. (2.16) yields respectively the fol-
lowing envelope and hole soliton solutions of Eq. (2.15) on the non-vanishing CW
background:
⎛ ⎞
d cosh [ϑ] + cos [ϕ]
A
⎜ cw + A
cosh [ϑ] + d cos [ϕ] ⎟
sol
ψ(x, t) = ⎜ ⎝
⎟ (2.24a)
b sinh [ϑ] + c sin [ϕ] ⎠
+i Asol
cosh [ϑ] + d cos [ϕ]
× exp [i cw + iθ (x, t)] ,
ψ(x, t) = (Acw + i Asol tanh [ξ(x, t)]) exp [i dr + iθ (x, t)] , (2.24b)

cw and dr being given respectively by Eq. (2.20) and Eq. (2.23c). The exact
soliton solutions (2.24b) and (2.24b) describe respectively envelope and hole soliton
signals embedded on a CW background.
Using now Eqs. (2.17a), (2.17c), and (2.20), we compute the chirping correspond-
ing to Eqs. (2.24b) and (2.24b) and find

∂ Q 3 kcw 3Q 13 + 2Q 14
δω(x, t) = − ( cw + θ ) = − +
∂x 2P 4P
⎡ ⎤
d cosh [ϑ] + cos [ϕ] 2
⎢ A + A
cosh [ϑ] + d cos [ϕ] ⎥
cw sol
⎢ ⎥
×⎢  2 ⎥ ,
⎣ b sinh [ϑ] + c sin [ϕ] ⎦
+A2sol
cosh [ϑ] + d cos [ϕ]
(2.25a)

δω(x, t) = − ( cw + θ )
∂x
3Q 13 + 2Q 14  2 
= −k1 + Acw + A2sol tanh2 [ξ(x, t)] . (2.25b)
4P
Here, b, c, d, ϕ and ϑ and ξ are given respectively by Eqs. (2.22) and (2.23b), and Acw ,
kcw , Acw , and Asol are arbitrary real constants. It is well seen from Eqs. (2.25a) and
(2.25b) that the nonlinear part of the frequency chirp strongly depends of the GVD
parameter P and both the self-steepening and the self-frequency shift coefficients
Q 13 and Q 14 through the nonlinear chirp parameter μ = −α10 , α10 being given by
Eq. (2.17c); on the other hand, the pulse intensity |ψ(x, t)|2 is independent of Q 13
and Q 14 . This means that we can control the chirping amplitude by varying the self-
steeping and self-frequency shift terms of Eq. (2.15). It is important to notice that
solutions (2.24b) and (2.24b) and the corresponding chirping (2.25a) and (2.25b)
are different from those obtained in Ref. [29] and can be used to investigate the
dynamics of chirped femtosecond bright and dark solitons in the network of Fig. 2.1.
The nonlinear chirp parameter μ and the constant chirp parameters − Q2P 3 kcw
or −k1
are similar to those obtained in Ref. [29].
2.3 Modulational Instability and Transmission… 27

Now, we turn to the study of the dynamics of chirped bright/dark solitons embed-
ded on a CW background and the corresponding chirping. For this aim, we distinguish
three special cases, namely, the case Acw = 0 corresponding to soliton signals prop-
agating on the vanishing CW background, the case kcw = ksol and 4 A2cw − A2sol > 0,
and the case kcw = ksol and 4 A2cw − A2sol < 0 corresponding respectively to

M R2 = 0, M I2 = 4 A2cw − A2sol , and M R2 = A2sol − 4 A2cw , M I2 = 0.

When kcw = ksol and 4 A2cw − A2sol > 0, we have



Q 3 Asol 4 A2cw − A2sol
ϑ(x, t) = ± t − ϑ0 ,
2

Q 3 4 A2cw − A2sol  1
and ϕ(x, t) = ± x − 2ksol t − ϕ0
2 P

so that the amplitude of the bright soliton is periodic in the x− coordinate with period

4π P
Tspat =
Q 3 4 A2cw − A2sol

and aperiodic in the t−coordinate. In the special case when kcw = ksol and 4 A2cw −
A2sol < 0, our computations give

Q 3 A2sol − 4 A2cw  1
ϑ(x, t) = x − 2ksol t − ϑ0 ,
2 P

Q 3 Asol A2sol − 4 A2cw
ϕ(x, t) = ∓ t − ϕ0 ;
2
this means that the wave amplitude is periodic in the t−coordinate with period


Ttemp = ,
Q 3 Asol A2sol − 4 A2cw

and aperiodic in the x−coordinate.


In Figs. 2.10 and 2.11, we show the dynamics of respectively bright soliton sig-
nals and dark soliton signals with the corresponding chirping propagating along the
network of Fig. 2.1. Generated with the nonlinear chirp parameter μ = −0.625, Fig.
2.10a and d, b and e, and c and f show the propagation of the bright soliton signals
and the corresponding chirping when Acw = 0, kcw = ksol and 4 A2cw − A2sol > 0, and
kcw = ksol and 4 A2cw − A2sol < 0, respectively. As we can see from these figures, the
chirping associated with the bright pulse in each of the cases Acw = 0 and kcw = ksol
28 2 Nonlinear Schrödinger Models for Solitons Propagation …

Fig. 2.11 Transmission of the dark solitary pulse through the network according to the exact
solution given by Eq. (2.24b) and the corresponding chirping (2.25b) for P = 0.5, Q 3 = −1, Q 14 =
0.5, with Asol = 0.5 and k1 = −2. a, d Acw = 0, Q 13 = −1.8; b, e Acw = 0.2, Q 13 = 1.8; c, f:
Acw = 0.2, Q 13 = −1.8. Reprint from Ref. [19], Copyright 2022, with permission from Elsevier

and 4 A2cw − A2sol < 0 is a dark soliton (see the first and last columns of Fig. 2.10),
while that of bright pulse in the case kcw = ksol and 4 A2cw − A2sol > 0 is a bright soli-
tary wave, as we can see from the plots of the middle column of Fig. 2.10. As we
can clearly see from plots of the middle (right) columns, the pulse amplitude and the
corresponding chirping are periodic (aperiodic) in the x− coordinate and aperiodic
(periodic) in the t− coordinate when kcw = ksol and 4 A2cw − A2sol > 0 (kcw = ksol and
4 A2cw − A2sol < 0). In Fig. 2.11a and d, we show the evolution of the dark solitary
signal propagating on a vanishing CW background and the corresponding chirping
when the nonlinear chirping parameter μ is negative. From the plots of the middle
and left panels of Fig. 2.11, we can observe the propagation of dark pulses embedded
on a non-vanishing CW background for respectively positive and negative nonlinear
chirping parameter μ. From the plots of the left and right panels of Fig. 2.11, we can
conclude that the chirping associated with dark solitary signals corresponding to a
negative nonlinear chirping parameter μ are bright pulses, while that with positive
nonlinear chirping parameter μ corresponding to the dark solitary pulse is of dark
soliton type, as we can well see from Fig. 2.11b and e. We thus conclude from the
plots of Figs. 2.10 and 2.11 that the chirping profile of a bright (Fig. 2.10) or a dark
(Fig. 2.11) solitary signal can be either of bright type or of dark soliton type.
Remark: It is important to notice that when kcw = ksol , Acw << Asol and 4 A2cw −
A2sol < 0, the background becomes very small within the existence of chirped bright
soliton, and the bright soliton given by Eq. (2.24b) will be stable under modulation.
As well as we know, a small perturbation for Eq. (2.24b) may lead to the modulation
instability in the situation when 4A2cw − A2sol > 0 with kcw = ksol [17, 30]. In order to
understand such a modulational instability process, we compute, under the conditions
kcw = ksol and 4 A2cw − A2sol > 0, the initial value of solution (2.24b) for ϑ0 = 0 as
follows
2.3 Modulational Instability and Transmission… 29

Fig. 2.12 The dynamics of (left) bright solitary signals given by # Eq. (2.24b) # and (middle and
right) the corresponding chirping defined by Eq. (2.25a) for small #4 A2cw − A2sol # with the nonlinear
chirping parameter μ of different sign, ϑ0 = ϕ0 = 0, kcw = ksol = 0.2, P = 0.5, and Q 3 = 1.
a, b Acw = 2, Asol = 3.9, Q 13 = 0.5, and Q 14 = −1/8 (meaning that μ = 0.625); c Acw = 2,
Asol = 3.9, Q 13 = 0.5, and Q 14 = −1/8 (meaning that μ = −0.625); d, e Acw = 0.95, Asol = 2,
Q 13 = 0.5, and Q 14 = −1/8 (meaning that μ = 0.625); f Acw = 0.95, Asol = 2, Q 13 = −0.5,
and Q 14 = 1/8 (meaning that μ = −0.625). Plots of the top (bottom) panels show first-order rogue
wave signals which are periodically loaded along the x− (t−) direction. Reprint from Ref. [19],
Copyright 2022, with permission from Elsevier
# #2
# #
# #
# A2 − 4 A2cw 1 #
|ψ(x, 0)|2 = ## Acw + sol  √  ## . (2.26)
# 2 Acw
1− Asol
cos 3 2Pcw sol x + ϕ0 ##
Q 4 A2 −A2
# 2 Acw

 √ 
Q3 4 A2cw −A2sol
It follows from conditions 4 A2cw −A2sol > 0 and 1 − Asol
2 Acw
cos 2P
x + ϕ0 >
0 for all x that the initial value of solution (2.24b) for ϑ0 = 0 given by Eq. (2.26)
is defined for all real x. Therefore, |ψ(x, 0)|2 diverges as A2sol → 4 A2cw − 0. This
means that a small periodic perturbation in the CW background may lead to the
modulational instability.
It is interesting to study the behavior of the chirped bright soliton
# given by#Eq.
(2.24b) and the corresponding chirping (2.25a) when kcw = ksol and #4 A2cw − A2sol # →
0. In such a situation, Tspat → +∞ and Ttemp → ∞. This means that the bright soli-
ton given by Eq. (2.24b) and the corresponding frequency chirp (2.25a) under the
parametric condition

0 < 4 A2cw − A2sol << 1 (0 < A2sol − 4 A2cw << 1)

behave like first-order rogue waves appearing periodically along the x− (t−) coordi-
nate and propagating on a CW background. For a better understanding, we depict in
Fig. 2.12 the three-dimensional evolution
# of bright
# solitary signals embedded in non-
vanishing CW background when #4 A2cw − A2sol # is small enough. In Fig. 2.12b, c and
30 2 Nonlinear Schrödinger Models for Solitons Propagation …

e, f, we have displayed the chirping corresponding to the bright soliton respectively


of Fig. 2.12a and d for μ = 0.625 and μ = −0.625, respectively. As it is clearly seen
from Fig. 2.12b and e, the chirping δω corresponding to bright first-order rogue wave
is a bright first-order rogue wave; Fig. 2.12c and f generating with μ = −0.625 and
depicting the chirping corresponding to bright first-order rogue waves of respectively
Fig. 2.12a and d show periodic dark first-order rogue waves.

2.3.3 Sister Femtosecond Nonlinear Modulated Waves with


Two Nonlinear Chirp Terms

Here, we are interested in finding chirped modulated nonlinear waves with special
properties that propagate along the network of Fig. 2.1. For this purpose, we use ansatz
(2.16) under the restrictions (2.17a) and (2.17b). One of the particularities of such
nonlinear modulated waves is that their chirping has two nonlinear parts and varies
as directly and inversely proportional to the intensity of the wave. To derive such
nonlinear modulated waves, we seek the complex envelope traveling-wave solutions
of Eq. (2.18a) as
φ(x, t) = ρ(z) exp [i {λ(z) − υt}] , (2.27)

where z = x − υt is the traveling coordinate and ρ(z) and λ(z) are real functions of
z. Inserting Eq. (2.27) in ansatz (2.16) yields

ψ(x, t) = ρ(z) exp [i {λ(z) − υt + θ (x, t)}] .

The chirping is then computed as

∂ dλ(z) 3Q 13 + 2Q 14 2
δω(x, t) = − [λ(z) − υt + θ (x, t)] = − + ρ (z).
∂x dz 4P

Replacing Eq. (2.27) in Eq. (2.18a) and separating out the real and imaginary parts
of the resulting equation yield
 2
d 2ρ dλ dλ
P + υρ − Pρ + υρ + Q 3 ρ 3 = 0, (2.28a)
dz 2 dz dz
 2
dρ d λ dλ dρ
−υ + P ρ 2 +2 = 0. (2.28b)
dz dz dz dz

A first integral of Eq. (2.28b) reads

dλ υ Q1 1
= + , (2.28c)
dz 2P P ρ2
2.3 Modulational Instability and Transmission… 31

where Q 1 is a real constant of integration. Hence, chirping is given as

υ 3Q 13 + 2Q 14 2 Q 1 1
δω(z) = − + ρ − . (2.28d)
2P 4P P ρ2

It is seen from Eq. (2.28d) that the chirping will be directly and inversely proportional
to the wave intensity as soon as Q 1 = 0. Equation (2.28d) means that the value of the
chirp parameter contains a free parameter Q 1 and depends on different coefficients of
the higher-order NLS equation (2.15) such as diffraction, self-steepening, and self-
frequency shift. Therefore, the amplitude of chirping can be managed by varying
these four parameters. Now, if we insert Eq. (2.28c) into Eq. (2.28a), we obtain the
elliptic ordinary differential equation (EODE)
 2
du def.
= 4βu 3 + 6γ u 2 + 4δu + ε = F(u),
dz
(2.29)
Q3 υ + 4Pυ
2
4Q 2
u = ρ2, β = − ,γ = − ,ε = − 21 , δ ∈ R,
2P 6P 2 P
which is known to have a variety of solutions such as solitonlike solutions and periodic
solutions that can be expressed in a generic form by means of the Weierstrass elliptic
function ℘ [31, 32]. The classification of the behavior of solutions u(z) of Eq.
(2.29) can be made as follows: First, we introduce the so-called invariants g2 and g3
of function ℘ [33]: g2 = −4βδ + 3γ 2 and g3 = 2βδγ − γ 3 − εβ 2 , which are then
used to introduce the discriminant of ℘ and F as = g23 − 27g32 . Using these three
parameters, we can classify the behavior of solutions u(z) of Eq. (2.29) as follows:
If either = 0 or = 0 and g2 > 0 and g3 > 0, then u(z) is periodic; if = 0,
g2 ≥ 0 and g3 ≤ 0, then u(z) is solitary wavelike [33]. Conditions under which Eq.
(2.29) admits real and bounded solutions u(z) can be obtained by considering the
phase diagrams of F(u) [33]. In the following, we focus ourselves to special cases
of Eq. (2.29) leading to solitonlike and periodic solutions, that leads respectively to
chirped solitary pulses and chirped periodic pulses propagating along the network
of Fig. 2.1.
Definition: We call two wave solutions u 1 (z) and u 2 (z) of Eq. (2.29) two sister wave
solutions if there exist two real constants B0 = 0 and C0 = 0 so that

u 1 (z) = A0 + B0 y(z), u 2 (z) = A0 + C0 w(z), C0 = B0 , (2.30a)

where A0 = 0 is any real solution of equation


 
ε + 2 A0 2δ + 3γ A0 + 2β A20 = 0, (2.30b)

and y(z) and w(z) are solutions of the nonlinear ordinary differential equations
32 2 Nonlinear Schrödinger Models for Solitons Propagation …
 2
dy δ + 3γ A0 + 3β A20
= 4β B0 y 3 + 6 (γ + 2β A0 ) y 2 + 4 y, (2.30c)
dz B0
 2
dw δ + 3γ A0 + 3β A20
= 4βC0 w 3 + 6 (γ + 2β A0 ) w 2 + 4 w, (2.30d)
dz C0

respectively. In what follows, any two wave solutions of Eq. (2.15) obtained with
the use of two sister wave solutions of Eq. (2.29) will be referred to as two sister
wave solutions of the NLS equation (2.15). The corresponding electric signals will
be referred to as two sister electric signals.
Sister Solitary Wave Solutions of the Higher-Order NLS Equation (2.15)
In order that Eqs. (2.30c) and (2.30d) admit simultaneously solitary solutions, param-
eters β, γ , δ, and A0 must satisfy the condition
 
 2 16  
= 36β 2 δ + 3γ A0 + 3β A20 (γ + 2β A0 )2 − β δ + 3γ A0 + 3β A20 = 0,
9

g2 ≥ 0 and g3 ≤ 0; here [34]


 
g2 = 3 (γ + 2β A0 )2 − 16β δ + 3γ A0 + 3β A20 ,

 
g3 = (γ + 2β A0 ) 8β δ + 3γ A0 + 3β A20 − (γ + 2β A0 )2 .

In what follows, we limit ourselves to the situation when

δ + 3γ A0 + 3β A20 = 0.

If we insert ε = −4Q 21 /P 2 into Eq. (2.30b), we obtain an equation for finding


the free parameters Q 1 and A0 = 0. The strategy here is to consider A0 = 0 as
an arbitrary real parameter and solve Eq. (2.30b) in Q 1 and then, replace δ by its
expression obtained from the condition = 0. For example, in the situation when
δ + 3γ A0 + 3β A20 = 0, we obtain, after using Eq. (2.29), the following result:
A0  2 2
Q 21 = − υ + 4Pυ + 2P Q 3 A0 , A0 P Q 3 < 0. (2.31)
32P Q 3

Integrating Eqs. (2.30c) and (2.30d) and taking into account the condition (2.31), we
obtain the following sister solitary wave solutions of Eq. (2.29)
    
3 υ 2 +4Pυ+6P Q 3 A0 υ 2 +4Pυ+6P Q 3 A0
3A0 + 2 A0 − 8P Q 3 B0 sinh2 4P 2
z
ρ 2 (z) = u(z) =   ,
υ 2 +4Pυ+6P Q 3 A0
3 + 2 sinh2 4P 2
z

(2.32a)
2.3 Modulational Instability and Transmission… 33
    
3 υ 2 +4Pυ+6P Q 3 A0 υ 2 +4Pυ+6P Q 3 A0
3A0 + 2 A0 − 8P Q 3 C0 sinh2 4P 2
z
ρ 2 (z) = u(z) =   ;
υ 2 +4Pυ+6P Q 3 A0
3 + 2 sinh2 4P 2
z

(2.32b)

here, A0 , B0 , C0 , and υ must be chosen from conditions B0 = C0

A0 > 0, P Q 3 < 0, υ 2 + 4Pυ + 6P Q 3 A0 > 0,


(2.33)
1
B0
> 16P Q 3 A0
, 1
3(υ 2 +4Pυ+6P Q 3 A0 ) C0
> 16P Q 3 A0
3(υ 2 +4Pυ+6P Q 3 A0 )
.

It is important to note that each pair of B0 and C0 satisfying conditions (2.33) leads
to two sister solitonlike solutions (2.32a)-(2.32b) of Eq. (2.29). As we can see from
Eqs. (2.32a) and (2.32b), two sister solitonlike solutions (2.32a)-(2.32b) will take
the same value, A0 at z = 0.
Now, if we use Eqs. (2.32b), (2.32a), (2.27), and (2.16), we obtain, under condi-
tions (2.33), the following sister solitonlike solutions of Eq. (2.15)

$  
%   
%
% 3A0 + 2 A0 − 3 υ +4Pυ+6P Q 3 A0 sinh2
2 υ 2 +4Pυ+6P Q 3 A0
% (x − υt)
8P Q 3 D0 4P 2
%
ψ(x, t) = ±%  
%
& υ 2 +4Pυ+6P Q 3 A0
3 + 2 sinh2 2 (x − υt)
4P

× exp [i {λ(z) − υt + θ (x, t)}] . (2.34)

In Eq. (2.34), one should take D0 = B0 for Eq. (2.32a) and D0 = C0 for Eq. (2.32b),
θ (x, t) being the real function defined by Eqs. (2.17a) and (2.17b), and λ(z) is
the solution of Eq. (2.28c) with ρ 2 (z) given by Eq. Eq. (2.32a) for D0 = B0 and
Eq. (2.32b) for D0 = C0 . Using now Eq. (2.28d), we find the corresponding sister
chirping as follows:
υ 3Q 13 + 2Q 14
δω(x, t) = − +
2P 4P
    
3 υ 2 +4Pυ+6P Q 3 A0 υ 2 +4Pυ+6P Q 3 A0
3A0 + 2 A0 − 8P Q 3 D0 sinh2 4P 2 (x − υt)
×  
υ 2 +4Pυ+6P Q A
3 + 2 sinh2 4P 2
3 0
(x − υt)

 
υ 2 +4Pυ+6P Q 3 A0
3 + 2 sinh2 4P 2
(x − υt)
Q1
−    .
3A0 + 2 A0 − (
P 3 υ 2 +4Pυ+6P Q 3 A0 ) υ 2 +4Pυ+6P Q 3 A0
8P Q 3 D0 sinh2 4P 2 (x − υt)

(2.35)
34 2 Nonlinear Schrödinger Models for Solitons Propagation …

Fig. 2.13 a and c: The transmission of two sister bright-dark solitary pulses along the network of
Fig. 2.1, generated with the use of the wave solution given by Eq. (2.34) for B0 = 2Q3P
3
, C0 = − 4Q
3P
3
,
respectively. b and d Frequency chirp given by Eq. (2.35) corresponding to two sister bright-dark
solitary signals of plots (a) and (c), respectively. Different parameters used in generating these plots
are given in the text. Reprint from Ref. [19], Copyright 2022, with permission from Elsevier

It is important to notice that depending on the choice of D0 = B0 or/and D0 = C0


and the coefficients of the NLS equation (2.15), the sister solitonlike solutions (2.34)
with the corresponding chirp have diverse behaviors, as we can see from the examples
below where we use
  
3A0 Q 3
υ = υ± = −2P 1 ± 2 − ,
2P

so that
υ 2 + 4Pυ + 6P Q 3 A0 = 4P 2 .

Transmission of two sister bright-dark solitonlike wave signals along the net-
work
To experience the transmission of two sister bright-dark solitonlike wave signals
along the network under consideration, we consider the following set of parameters
for our first example:

3P 3P  √ 
A0 = 1, B0 = , C0 = − , 3Q 13 + 2Q 14 = 4P, Q 3 = −3 + 2 2 P.
2Q 3 4Q 3
2.3 Modulational Instability and Transmission… 35

Fig. 2.14 Spatiotemporal evolution of two sister dark-dark solitary pulses obtained with the use of
the analytical solution defined by Eq. (2.34) for a B0 = − 5Q3P
3
and b C0 = − 2QP
3
. The frequency
chirp δω(x, t) given by Eq. (2.35) are shown in plots (b) and (d) and correspond respectively
to B0 = − 5Q3P
3
and C0 = − 2Q P
3
. Reprint from Ref. [19], Copyright 2022, with permission from
Elsevier

From Eq. (2.31), we then compute Q 1 = −2P. For P = − 21 and υ = υ+ . For these
parameters, we depict in Fig. 2.13 the spatiotemporal evolution of two sister bright-
dark solitonlike wave pulses and the corresponding chirping along the network of
Fig. 2.1. Figure 2.13a, c show the evolution of bright and dark solitonlike wave
pulses, respectively; in Fig. 2.13b and d, we show respectively the evolution of a
dark solitonlike signal and a dark-bright-dark solitonlike signal. Thus, the frequency
chirp Fig. 2.13b of the bright solitonlike wave (a) is of a dark solitonlike wave type,
while the dark solitonlike wave signal plotted in Fig. 2.13c leads to a dark-bright-dark
solitonlike wave (dark two-soliton).
Evolution of two sister dark-dark solitonlike pulses through the network
To demonstrate the evolution of two sister dark-dark solitonlike pulses through the
network of Fig. 2.1, we consider the same set of parameters as in the previous
example, but with
3P P
B0 = − and C0 = − .
5Q 3 2Q 3

For those parameters, we display in Fig. 2.14 the evolution of two sister dark-dark
solitonlike pulses and the corresponding chirping. As we can see from different plots
36 2 Nonlinear Schrödinger Models for Solitons Propagation …

Fig. 2.15 Transmission of two sister bright-bright solitary signals through the lossless network of
Fig. 2.1 generated with the help of exact the solution given by Eq. (2.34) for a B0 = − 5Q3P
3
and c
C0 = − 2Q
P
3
with P = − 21 , and the corresponding frequency chirp given by Eq. (2.35). b and d are
obtained with B0 = − 5Q3P
3
and C0 = − 2QP
3
, respectively. Other parameters are given in the text.
Reprint from Ref. [19], Copyright 2022, with permission from Elsevier

of Fig. 2.14, the frequency chirp corresponding to the two sister dark-dark solitonlike
signals of Fig. 2.14a and c are dark-bright-dark pulses, as it is well seen from Fig.
2.14b and d.
Propagation of Two Sister Bright-Bright Solitonlike Signals Along the Network
Two sister bright-bright solitonlike signals can be generated with the following set
of parameters

P 3P  √ 
A0 = 2, B0 = , C0 = , 3Q 13 + 2Q 14 = 4P, Q 3 = −P, υ = υ+ = −2 1 + 5 P.
Q3 2Q 3

From Eq. (2.31), we then obtain Q 1 = −3P. For P = − 21 , we depict in Fig. 2.15
the evolution of the two sister bright-bright solitonlike pulses (Fig. 2.15a and c); the
corresponding chirping are depicted in Fig. 2.15b and d, respectively. As we can see
from Fig. 2.15b and d, the corresponding chirping for the two sister bright-bright
solitonlike signals shown in Figs. 2.15a and c are dark-bright-dark solitonlike waves.
2.3 Modulational Instability and Transmission… 37

Fig. 2.16 a and c The evolution of two sister solitary wave pulses along the network of Fig. 2.1 for
respectively B0 = − QP3 and C0 = − 4Q 3P
3
; b and d Plots of the chirping corresponding to the two
sister dark-dark solitary signals plotted on (a) and (c), respectively. Other parameters are given in
the text. Reprint from Ref. [19], Copyright 2022, with permission from Elsevier

Engineering Two Sister Solitonlike Pulses With Bright-Dark-Bright Chirping


We now turn to engineer two sister solitonlike pulses with bright-dark-bright chirping.
For this end, we consider the following set of parameters:

1 P 3P  √ 
A0 = , B0 = − , C0 = − , Q 3 = −2 15 + 4 14 P,
2 Q3 4Q 3


υ = −P 2 − 98 + 24 14 , 3Q 13 + 2Q 14 = −12P, Q 1 = 2P = 1.

For this set of parameters, we display in Fig. 2.16a and c two sister dark-dark soliton-
like signals. The corresponding chirping are displayed in Fig. 2.16b for B0 = − QP3
and 2.16d for C0 = − 4Q 3P
3
. As one can seen from Fig. 2.16b and d, the chirping
of the two sister dark-dark solitonlike waves plotted in Fig. 2.16a and c show a
bright-dark-bright solitonlike (bright two-soliton solitonlike) behavior.
Two Sister Periodic Wave Solutions of the Higher-Order NLS Equation (2.15)
To find periodic solutions of Eqs. (2.30c) and (2.30d), we consider only the case
when = 0 [34]. Inserting Eq. (2.29) into Eq. (2.30b) and solving the resulting
equation in δ yield
38 2 Nonlinear Schrödinger Models for Solitons Propagation …

Fig. 2.17 Top: Amplitude profile of the two sister dn cnoidal signals obtained with the use of
analytical solutions (2.43a) (solid line) and (2.43b) (dashed line) for two different values of QP3 : a
Q 3 = −0.1, and b Q 3 = −0.245. Bottom: Effects of the GVD parameter P and cubic nonlinearity
P P

Q 3 on the femtosecond cnoidal pulses propagating through the electric network of Fig. 2.1 and
modeled by the higher-order NLS Eq. (2.15). Plots (c) and (d) are generated with the help of
respectively the dn periodic solution (2.43a) and the nd periodic solution (2.43a), for QP3 = −0.1
(solid line) and QP3 = −0.245 (dashed line). Other parameters are given in the text. Reprint from
Ref. [19], Copyright 2022, with permission from Elsevier

 
Q 21 2P Q 3 A20 + υ 2 + 4Pυ A0
δ= + . (2.36a)
P 2 A0 4P 2

First of all, we notice that if B0 and C0 satisfy the relationship


 2 
δ + 3γ A0 + 3β A20 υ + 4Pυ + 4P Q 3 A0 A20 − 4Q 21
B0 C0 = = (2.36b)
β 2P Q 3 A0

and y(z) is a periodic wave solution of Eq. (2.30c), then w(z) = 1/y(z) will be a
periodic wave solution of Eq. (2.30d); therefore,

u 1 (z) = A0 + B0 y(z),
 2 
υ + 4Pυ + 4P Q 3 A0 A20 − 4Q 21 1
u 2 (z) = A0 +
2P Q 3 A0 B0 y(z)
2.3 Modulational Instability and Transmission… 39

Fig. 2.18 Chirping profile (2.28d) for the two sister dn-nd periodic pulses (top) in Fig. 1.26a and
+2Q 14 +2Q 14
(bottom) in Fig. 1.26b for three different values of 3Q 134P . Top panels: a 3Q 134P = −5,
3Q 13 +2Q 14 3Q 13 +2Q 14 3Q 13 +2Q 14 +2Q 14
b 4P = −3.5, and c 4P = 5; Bottom panels: d 4P = −8, e 3Q 134P =
+2Q 14
−12, and f 3Q 134P = −16. Plots in solid line correspond to solution (2.43a), while those in
dashed line are obtained with the use of solution (2.43b). Other parameters are given in the text.
Reprint from Ref. [19], Copyright 2022, with permission from Elsevier

will be two sister periodic wave solutions of Eq. (2.29). Henceforth, we consider
A0 = 0 as an arbitrary real parameter so that δ could be expressed in terms of A0 as
showed in Eq. (2.36a).
If we use Eq. (2.36a) and introduce α1 , α2 , and α3 as follow,
 
2Q 3 υ 2 + 4Pυ + 6P Q 3 A0 4Q 21 − υ 2 + 4Pυ + 4P Q 3 A0 A20
α1 = − D0 , α2 = − , α3 =
P P2 P 2 D0 A 0
(2.37a)
for D0 ∈ {B0 , C0 }, equations. (2.30c) and (2.30d) will take the common form
 2
dy
= α1 y 3 + α2 y 2 + α3 y. (2.38)
dz

In the following subsections, we describe two classes of two sister periodic wave
solutions of Eq. (2.38) with the help of which one can build two sister periodic
solutions of the higher-order NLS equation (2.15) that model the propagation of two
sister periodic wave pulses through the network of Fig. 2.1.
Two sister Cnoidal wave solutions
Solutions y(z) of this class is sought as a quadratic function of sn [μz, m] , namely,

y(z) = A + Bsn2 [μz, m] , (2.39a)

where sn [μz, m] is the Jacobi elliptic sine function with the elliptic modulus m ∈
[0, 1], μ being an arbitrary real parameter. Asking that function (2.39a) satisfies Eq.
40 2 Nonlinear Schrödinger Models for Solitons Propagation …

(2.38) leads, under the conditions



α22 − 4α1 α3 > 0, α2 ± α22 − 4α1 α3 > 0, (2.39b)

to the periodic Cnoidal wave solution


⎡  ⎤
 
⎢ α2 + 4μ −α2 ± α2 − 4α1 α3 ⎥
2 2
⎢− ⎥
2α3 ⎢ ⎢ ⎥
y(z) = − 4α1 α3 ⎥; (2.39c)

α2 + 4μ2 ⎢    ⎥
1 α1 α3 ⎥
⎣ +sn μz,
2
−   ⎦
μ 2 α2 + 4μ2

here, μ is any real root of the algebraic equation



32μ4 + 4 α2 ± 3 α22 − 4α1 α3 μ2 − 4α1 α3 − α22 ± 3α2 α22 − 4α1 α3 = 0
(2.40a)
satisfying the condition
α1 α3
−1≤   < 0. (2.40b)
2 α2 + 4μ2 μ2

Two Sister dn-nd Wave Solutions


When α1 , α2 , and α3 satisfy condition α22 − 4α1 α3 > 0 and one of the conditions

 
α2 + 4μ2 −α2 ± α22 − 4α1 α3
− α2 ± α22 − 4α1 α3 > 0 and > 1,
4α1 α3
(2.40c)
solution (2.39c) will satisfy the condition y(z) = 0 for all z, so that w(z) = 1/y(z)
will also be a Cnoidal wave solution of Eq. (2.30d) in the special situation when
B0 C0 is defined by equation (2.36b). Expressing sn2 in terms of dn2 , we derive from
Eq. (2.39c), under the condition

α22 > 4α1 α3 > 0 and α2 > 0, (2.40d)

the following dn periodic solution of Eq. (2.30c)


⎡ $ $ ⎤
% %
% α + α 2 − 4α α %
α2 + α22 − 4α1 α3 ⎢ 1 & 2 1 3 % 2 α22 − 4α1 α3 ⎥
dn2 ⎢ z, % ⎥.
2
y(z) = − ⎣2 & ⎦
2α1 2 α + α 2 − 4α α
2 2 1 3

(2.41a)
2.3 Modulational Instability and Transmission… 41

Fig. 2.19 Effects of the GVD parameter P and the cubic nonlinearity parameter Q 3 on the frequency
chirp for the two sister dn-nd periodic signals given by Eqs. (2.43a) and (2.43b) for the same set
+2Q 14
of parameters as in Fig. 2.18 for 3Q 134P = −16 and three values of QP3 , QP3 = −0.22 (solid
line), QP3 = −0.23 (dashed line), and QP3 = −0.24 (dot-dashed line). a Chirping profile for the dn
periodic wave pulse (2.43a); b Chirping profile for the nd periodic wave signal (2.43b). Reprint
from Ref. [19], Copyright 2022, with permission from Elsevier

It is important to note that Eq. (2.41a) and Eq. (2.39c) are two distinct solutions of
Eq. (2.30c); indeed, they are obtained under different conditions of α1 , α2 , and α3 .
Because y(z) = 0 for all real z and B0 and C0 satisfy the Eq. (2.36b), solution (2.41a)
gives birth to the solution

⎡ $ $ ⎤
% %
% α + α 2 − 4α α %
2α1 ⎢ 1 & 2 1 3 % 2 α 2 − 4α α
1 3 ⎥
nd2 ⎢ z, % ⎥,
2 2
w(z) = − ⎣ & ⎦
α2 + α22 − 4α1 α3 2 2 α2 + α22 − 4α1 α3
(2.41b)
of Eq. (2.30d), where nd(x, m) = 1/dn (x, m).
Two Sister Periodic Wave Solutions of Eq. (2.29)
By using the above two sister periodic wave solutions (2.39c) and (2.41a) of Eq.
(2.38), we can, with the use of Eq. (2.30a), write down the following two sister
periodic wave solutions of Eq. (2.29),

ρ 2 (z) = u 1 (z) = A0 + B0 y B0 (z), ρ 2 (z) = u 2 (z) = A0 + C0 yC0 (z), (2.42a)

where y B0 (z) and yC0 (z) are obtained from Eqs. (2.39c) and (2.41a) by replacing
respectively D0 by B0 and C0 in α1 and α3 , as one can see from Eq. (2.37a). For
example, in the special case when B0 and C0 satisfy Eq. (2.36b), Eqs. (2.41a) and
(2.41b) lead to the following two sister dn-nd periodic wave solutions of Eq. (2.29)

P α2 + α2 − 4α1 α3
2

ρ 2 (z) = u 1 (z) = A0 + dn2


4Q 3
42 2 Nonlinear Schrödinger Models for Solitons Propagation …
⎡ $ $ ⎤
% %
% α + α 2 − 4α α % 2 α 2 − 4α α
⎢1 & 2 1 3 % 1 3 ⎥
×⎢
2 %
2
⎥,
⎣2 2
z, & ⎦ (2.43a)
α + α 2 − 4α α
2 2 1 3

 
2 υ 2 + 4Pυ + 4P Q 3 A0 A20 − 4Q 21
ρ (z) = u 2 (z) = A0 +
2

P 2 A0 α2 + α22 − 4α1 α3

⎡ $ $ ⎤
% %
% α + α 2 − 4α α % 2 α 2 − 4α α
⎢ 1 & 2 1 3 % 1 3 ⎥
×nd2 ⎢ z, % ⎥ . (2.43b)
2 2
⎣2 & ⎦
2 α + α 2 − 4α α
2 2 1 3

Here, α1 , α2 , and α3 in which D0 = B0 satisfy conditions (2.40d). Henceforth, solu-


tions (2.43a) and (2.43b) will be referred to as dn and nd type solutions of Eq.
(2.29). From the equality u(z) = ρ 2 (z), various free parameters A0 = 0, B0 = 0,
Q 1 = 0, and υ must be chosen so that u 1 (z) and u 2 (z) should be positive. Figures
2.17 and 2.18 display the amplitude profile of typical two sister dn wave (2.43a) and
nd wave (2.43b) and the corresponding chirping for the following numerical values
of parameters
  
10 199 199 1 Q3
α2 = , α1 α3 = , υ = P −2 + −15 000 A0 − 199 ,
2500 625 50 P
 
P 1 P Q3
Q1 = − α1 α3 A0 − 2 A0 + α2 A20 , A20 = 1, m = 0.99.
2 2 Q3 P

For the above numerical values of parameters, we show in Fig. 2.17 the typical
periodic solutions generated with the use of Eqs. (2.43a) and (2.43b); different plots
of this figure behave like periodic dark solitonlike waves. As we can clearly see
from plots of Fig. 2.17, the amplitude of the nd wave given by Eq. (2.43b) is higher
than that of the dn wave given by Eq. (2.43a). Also, Fig. # 2.17c
# and d show that the
#P#
amplitude and the width of the wave increase when # Q 3 # increases. We can then
conclude that parameter P of the GVD and parameter Q 3 of the cubic nonlinearity
can be used to control two sister nonlinear modulated waves of the higher NLS
equation (2.15). Figure 2.18 show how parameters Q 13 and Q 14 of self-steeping and
self-frequency shift can be used to manage the motion of the chirping. Depending
on the choice of these two parameters, the chirping corresponding to the two sister
dn-nd periodic wave solutions (2.43a) and (2.43b) can behave like (i) periodic bright
solitons as it is seen from Fig. 2.18a, (ii) periodic bright two-solitons as one can see
from dashed plot of Fig. 2.18b, (iii) periodic dark solitons as shown in Fig. 2.18c,
or (iv) periodic bright-dark-bright solitons, as we can see from dashed plots of Fig.
2.18d–f. Comparing plots of the top panels of Fig. 2.18 obtained with P/Q 3 = −1
2.3 Modulational Instability and Transmission… 43

Fig. 2.20 Top: Spatiotemporal evolution of the amplitude |ψ(x, t)|2 showing the evolu-
tion of two sister pulses of type dn (a) and b periodic waves according to the wave
solutions given by Eqs. (2.44a) and (2.44b), respectively, for P = 21 , Q 3 = −5, 3Q 13 +

Q3
2Q 14 = −15P, α2 = 10 2500 , α1 α3 = 625 , υ = P −2 + 50 −15 000 P A0 − 199 , Q 1 =
199 199 1

  
Q3
2 − 2 Q 3 α1 α3 A0 − 2 P A0 + α2 A0 , and A0 = 1. Bottom: The spatiotemporal chirping
P 1 P 2
#
υ +2Q 14 2 #
δω(x, t) = − 2P + 3Q 134P ρ (z) − QP1 ρ 21(z) # for (c) the dn periodic modulated wave
z=x−υt
plotted in (a) with ρ 2 (z) given by Eq. (2.43a) and (d) the nd periodic modulated wave plotted in
(b) with ρ 2 (z) given by Eq. (2.43b). Reprint from Ref. [19], Copyright 2022, with permission from
Elsevier
with those of the bottom panels generated with P/Q 3 = −0.245, we conclude that
the behavior of the chirping also depends on parameter P of the GVD and parameter
Q 3 of the cubic nonlinearity. As we can clearly see from plots of Fig. 2.18, the
amplitude and the width of the chirping increase as |P/Q 3 | increases. For a better
understanding, we display in Fig. 2.19 the chirping profiles for both the dn and nd
periodic wave solutions (2.43a) and (2.43b) for different values of P/Q 3 .
In order to obtain two sister periodic wave solutions of Eq. (2.15), we go
back to ψ(x, t) by substituting Eq. (2.42a) into Eq. (2.27) and then, inserting the
resulting equation into Eq. (2.16). For example, two sister dn and nd periodic
wave solutions of the higher-order NLS equation (2.15) generated from Eqs. (2.43a)
and (2.43b) are found to be

ψ(x, t)
$  ⎡ $ $ ⎤
% % %
% % α + α 2 − 4α α %
% P α2 + α22 − 4α1 α3 & 2 % 2 α22 − 4α1 α3
% ⎢ 1 1 3 ⎥
dn2 ⎢ (x − υt) , % ⎥
2
=%& A0 + ⎣2 & ⎦
4Q 3 2 α + α 2 − 4α α
2 2 1 3

× exp [i {λ(z) − υt + θ(x, t)}] , (2.44a)


44 2 Nonlinear Schrödinger Models for Solitons Propagation …

ψ(x, t)
⎧ ⎡ $
⎪ %


  %
⎢ 1 & α2 + α2 − 4α1 α3
2
2 υ 2 + 4Pυ + 4P Q 3 A0 A20 − 4Q 21
= A0 +  2
nd ⎣⎢ (x − υt) ,

⎪ 2 2
⎩ P 2 A0 α2 + α22 − 4α1 α3

$ ⎤⎫ 2 1
% ⎪
%
% 2 α22 − 4α1 α3 ⎥⎪

× %
& ⎥
⎦ exp [i {λ(z) − υt + θ(x, t)}] (2.44b)

α2 + α22 − 4α1 α3 ⎪

where α1 , α2 , and α3 are given by Eq. (2.37a) with which D0 = B0 and satisfy
conditions (2.40d), θ (x, t) is any real function satisfying conditions (2.17a) and
(2.17b), λ(z) is any solution of Eq. (2.28c), with z = x − υt. The free parameters
A0 , Q 1 = 0, υ, δ, and B0 must be suitably chosen from Eq. (2.36a), that is, they must
be chosen so that the conditions (2.40d) are fulfilled and the main square root in Eq.
(2.44a) is defined. Chirping δω(x, t) associated with solutions (2.44a) and (2.44b)
are obtained by inserting respectively Eqs. (2.43a) and (2.43b) with z = x − υt
into Eq. (2.28d). In Fig. 2.20a and b, we depict the spatiotemporal evolution of the
two sister dn and nd periodic nonlinear modulated waves obtained with the help of
respectively solutions (2.44a) and (2.44b) of the higher-order NLS equation (2.15).
The evolution of the corresponding chirping is shown in Fig. 2.20c and d, respectively.
From Fig. 2.20a and b, it is clearly seen that the two sister dn-nd modulated wave
corresponding to solutions (2.44a) and (2.44b) behave like periodic dark solitons.
The chirping shown in Fig. 2.20c has the behavior of periodic bright solitons. It is
clearly seen from Fig. 2.20d that the chirping for the nd modulated wave plotted in
Fig. 2.20b behaves like periodic bright two-solitons.
In conclusion, we have showed that the propagation of ultrashort pulses along the
lossless network of Fig. 2.1 are modeled by a higher-order nonlinear Schrödinger
equation with derivative terms. With the help of this NLS equation, the phenomenon
of the modulational instability in the network under consideration has been inves-
tigated and our results reveal that the competing higher-order nonlinearity induces
propagating on CW background solitonlike bright (dark) solitary wave and rogue
wave signals in the lossless network of Fig. 2.1. The concept of two sister nonlin-
ear modulated waves was introduced and their properties were explored; through
bright solitonlike waves and periodic waves, the propagation of such nonlinear mod-
ulated waves in the network under consideration is analyzed. We have shown that the
nonlinear chirp associated with each of electric solitary pulses embedded on a CW
background is directly proportional to the intensity of the pulse, while chirping asso-
ciated with each two sister modulated nonlinear waves is formed of two nonlinear
parts, one of which is directly proportional to the intensity of the wave, while another
part is inversely proportional to the intensity of the wave. Also, we have obtained
that the amplitude of the chirping can be managed by varying the parameters of the
GVD and cubic nonlinearity.
References 45

Our investigations also revealed the amplification [the compression] of two sister
chirped femtosecond modulated signals can be enhanced by either increasing |P|
or decreasing |Q 3 | [decreasing |P| or increasing |Q 3 |], and the wave structure is
not destroyed during the process. It is also established that these two parameters P
and Q 3 have important effects of the feature of the chirping. They also have some
effect on the unlimited transmission of the compressed amplified chirping in an
inhomogeneous media. Through our results, one can conclude that the self-steeping
and self-frequency parameters Q 13 and Q 14 have a universal influence on chirping
dynamics.

References

1. A. Noguchi, Solitons in a nonlinear transmission line. Electr. Commun. Japan 57A, 9 (1974)
2. R. Hirota, K. Suzuki, Studies on lattice solitons by using electrical networks. J. Phys. Soc.
Jpn. 28, 1366–1367 (1970)
3. Y.H. Ichikawa, T. Mitsuhaski, K. Konno, Contribution of higher order terms in the reductive
perturbation theory. I. A case of weakly dispersive wave. J. Phys. Soc. Jpn. 41, 1382–1386
(1976)
4. T. Yoshinaga, T. Kakutani, Second order K-dV soliton on a nonlinear transmission line. J.
Phys. Soc. Jpn. 53, 85–92 (1984)
5. F.B. Pelap, M.M. Faye, Solitonlike excitations in a one-dimensional electrical transmission
line. J. Math. Phys. 46, 033502 (2005)
6. M. Marklund, P.K. Shukla, Modulational instability of partially coherent signals in electrical
transmission lines. Phys. Rev. E 73, 057601 (2006)
7. E. Kengne, W.M. Liu, Exact solutions of the derivative nonlinear Schrödinger equation for a
nonlinear transmission line. Phys. Rev. E 73, 026603 (2006)
8. E. Kengne, A. Lakhssassi, Analytical study of dynamics of matter-wave solitons in lossless
nonlinear discrete bi-inductance transmission lines. Phys. Rev. E 91, 032907 (2015)
9. M. Remoissenet, Waves Called Solitons, 3rd edn. (Springer, Berlin, 1999)
10. W.M. Liu, E. Kengne, Schrödinger Equations in Nonlinear Systems (Springer, Singapore,
2019)
11. T.B. Benjamin, J.E. Feir, The disintegration of wavetrains on deep water. Part 1. J. Fluid Mech.
27, 417–430 (1967)
12. B. Li, X.F. Zhang, Y.Q. Li, Y. Chen, W.M. Liu, Solitons in Bose-Einstein condensates with
time-dependent atomic scattering length in an expulsive parabolic and complex potential.
Phys. Rev. A 78, 023608 (2008)
13. C.T.A. Brown, M.A. Cataluna, A.A. Lagatsky, E.U. Rafailov, M.B. Agate, C.G. Leburn, W.
Sibbett, Compact laser-diode-based femtosecond sources. New J. Phys. 6, 175 (2004)
14. A. Margiolakis, G.D. Tsibidis, K.M. Dani, G.P. Tsironis, Ultrafast dynamics and subwave-
length periodic structure formation following irradiation of GaAs with femtosecond laser
pulses. Phys. Rev. B 98, 224103 (2018)
15. M.L. Groot, R. Van Grondelle, Femtosecond time-resolved infrared spectroscopy. in Biophysi-
cal Techniques in Photosynthesis, eds. by Aartsma T.J., Matysik J. Advances in Photosynthesis
and Respiration, vol. 26 (Springer, 2008)
16. Lu. Qiming, Qi. Shen, Jianyu Guan, Min Li, Jiupeng Chen, Shengkai Liao, Qiang Zhang,
Chengzhi Peng, Sensitive linear optical sampling system with femtosecond precision. Rev.
Sci. Instrum. 91, 035113 (2020)
17. M.J. Ablowitz, P.A. Clarkson, Soliton, Nonlinear Evolution Equations and Inverse Scattering
(Cambridge University Press, Cambridge, 1991)
18. G.P. Agrawal, Nonlinear Fiber Optics (Academic, San Diego, 2001)
46 2 Nonlinear Schrödinger Models for Solitons Propagation …

19. E. Kengne, W.M. Liu, Modulational instability and sister chirped femtosecond modulated
waves in a nonlinear Schrödinger equation with self-steepening and self-frequency shift.
Commun. Nonlinear Sci. Numer. Simul. 108, 106240 (2022)
20. A. Mohamadou, E. Wamba, S.Y. Doka, T.B. Ekogo, T.C. Kofane, Generation of matter-wave
solitons of the Gross-Pitaevskii equation with a time-dependent complicated potential. Phys.
Rev. A 84, 023602 (2011)
21. G.P. Agrawal, Nonlinear Fiber Optics, 4th ed. (Academic, 2007)
22. J.R. de Oliveira, Marco A. de Moura, J. Miguel Hickmann, A.S.L. Gomes, Self-steepening
of optical pulses in dispersive media. J. Opt. Soc. Am. B 9, 2025 (1992)
23. S.-H. Han, Effect of self-steepening on optical solitons in a continuous wave background.
Phys. Rev. E 83, 066601 (2011)
24. J.K. Lucek, K.J. Blow, Soliton self-frequency shift in telecommunications fiber. Phys. Rev.
A 45, 6666 (1992)
25. E. Madelung, Eine anschauliche Deutung der Gleichung von Schrödinger. Naturwis-
senschaften 14, 1004 (1926)
26. V.R. Kumar, R. Radha, M. Wadati, Phase engineering and solitons of Bose–Einstein conden-
sates with two- and three-body interactions. J. Phys. Soc. Jpn. 79, 074005 (2010)
27. R. Marquié, J.M. Bilbault, M. Remoissenet, Nonlinear Schrödinger models and modulational
instability in real electrical lattices. Phys. D 87, 371 (1995)
28. Z.X. Liang, Z.D. Zhang, W.M. Liu, Dynamics of a bright soliton in Bose–Einstein condensates
with time-dependent atomic scattering length in an expulsive parabolic potential. Phys. Rev.
Lett. 94, 050402 (2005)
29. Alka A. Goyal, R. Gupta, C.N. Kumar, Chirped femtosecond solitons and double-kink solitons
in the cubic-quintic nonlinear Schrödinger equation with self-steepening and self-frequency
shift. Phys. Rev. A 84, 063830 (2011)
30. Z.Y. Xu, L. Li, Z. Li, G. Zhou, Modulation instability and solitons on a CW background in
an optical fiber with higher-order effects. Phys. Rev. E. 67, 026603 (2003)
31. K. Weierstrass, Mathematische Werke V (New York, Johnson, 1915), pp.4–16
32. E.T. Whittaker, G.N. Watson, A Course of Modern Analysis (Cambridge University Press,
Cambridge, 1927), p. 454
33. K. Chandrasekharan, Elliptic Functions (Springer, Berlin, 1985), p.44
34. H.W. Schürmann, V.S. Serov, Traveling wave solutions of a generalized modified Kadomtsev-
Petviashvili equation. J. Math. Phys. 45, 2181 (2004)
Chapter 3
Transmission of Dissipative Solitonlike
Signals Through One-Dimensional
Transmission Networks

Abstract This chapter presents analytical results for the dynamics of nonlin-
ear modulated waves in dissipative electrical nonlinear transmission networks. We
employ the reductive perturbation method in the semi-discrete limit to derive a dis-
sipative NLS equation that models the propagation of nonlinear modulated waves
in the given network. Based on both the direct method and the method of Weier-
strass elliptic functions, we present classes of bright, kink, dark, or chirped Lambert
W-kink solitonlike solutions of the amplitude equations. These solitonlike solutions
are then used to investigate analytically the transmission of dissipative solitonlike
signals through the networks under consideration. Effects of the dissipation are also
investigated. Naturally, the wave’s amplitude decreases and its width increases under
the action of the dissipation. We show how the dissipative elements of the networks
can be used to control the motion of solitonlike signals during their propagation along
the dissipative network system under consideration.

3.1 Chirped Lambert W-Kink Waves Propagation


in a Lossy Electrical Transmission Network

3.1.1 Introduction

By means of both dissipative CQ NLS equations and complex CQ CGL equations,


the dynamics of nonlinear modulated waves in lossy electrical nonlinear transmission
networks have been studied in details [1–3] and it has been proven that both the cubic
CGL equations as well as the cubic-quintic ones support a class of localized solu-
tions such as stationary solitons, sources, sinks, Lambert W-kink solitons, moving
solitons and fronts with fixed velocity [4, 5]. Also, these models admit erupting soli-
ton solutions, which periodically exhibit explosive instability [6, 7]. Such erupting
solitons were found analytically, numerically, and experimentally in ultrafast fiber
lasers with normal dispersion and in passively mode-locked lasers [8, 9].
The Lambert function W(x), also called the product logarithm, defined as one
inverse to x = W exp(W) [10–12] finds many applications in physics, applied math-
ematics, and computer science [13, 14]. In the present Section, we mainly aim to
© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2022 47
E. Kengne and W. Liu, Nonlinear Waves,
https://doi.org/10.1007/978-981-19-6744-3_3
48 3 Transmission of Dissipative Solitonlike Signals …

demonstrate through a generalized CQ CGL equation [15] how the Lambert function
W function can be used to investigate the propagation of chirped pulses along a lossy
nonlinear electric transmission network.
The model under consideration is the discrete nonlinear transmission network
with dissipative elements made of N identical blocks, illustrated in Fig. 3.1 [1]. Each
single block contains a linear inductor with inductance L1 in the series branch and
a linear inductor with inductance L2 in parallel with the bias-dependent nonlinear
capacitor C(V ) in the shunt branches. To take into consideration the dissipative effects
of the network, two linear conductances G 1 and G 2 are connected in parallel with,
respectively, L1 and L2 . In this model, the nonlinear capacitance C(V ), responsible
for nonlinearity of the system, consists of a reversed-biased diode with differential
capacitance which is a function of voltage Vn across the n-th capacitor, biased by a
constant voltage Vb . For low voltage, it is of the form [1]

dQn  
C(Vb + Vn ) =  C0 1 − 2αVn + 3βVn2 ; (3.1a)
dVn

here, C0 = C(Vb ) is the capacitance of the nonlinear diode at the dc bias-voltage Vb ,


and α and β are nonlinear parameters of the stored electrical charge Qn . Here, α is
assumed to be positive [16].
If we apply the Kirchhoff’s laws to the network in Fig. 3.1, we arrive to the
following difference differential system for the propagation of modulated waves in
the model
 2   
d d d
+ ω 2
0 + 2u 
σ
0 2 Vn + u0
2
+ 2u σ
0 1 (2Vn − Vn−1 − Vn+1 )
dt 2 dt dt
d2  
= 2 αVn2 − βVn3 , (3.2)
dt

with n = 1, 2, . . . , N − 1; here, u02 = (C0 L1 )−1 and ω02 = (L2 C0 )−1 are the dimen-
sionless capacitance and squared characteristic frequency of the network, respec-
tively, while σ1 and  σ2 are dimensionless conductances
 
L1 L1
σ1 = G1, 
σ2 = G2. (3.3)
4C0 4C0

Throughout this Section, numerical results will be displayed for the network
parameters

L1 = 220 µH, L2 = 470µH, C0 = 370 pF, α = 0.21 V−1 , β = 0.0197 V−2 . (3.4)

For the numerical simulation, the total number N of cells in the network must be
chosen so to eliminate wave reflection at the edge of the network.
3.1 Chirped Lambert W-Kink Waves Propagation in a Lossy Electrical … 49

3.1.2 Amplitude Equation

In order to derive the governing equation (amplitude equation) that models the prop-
agation of slowly modulated waves in the lossy network of Fig. 3.1, we employ the
reductive perturbation method in the semidiscrete limit. In other words, we introduce
the continuum approximation for the envelope and keep a discrete description for the
carrier wave, whose wavenumber kp ranges over the Brilloum zone (0 ≤ kp ≤ π ),
with the corresponding frequency ωp = 2π fp . For this aim, two slow time scales are
introduced as T1 = εt and T2 = ε2 t, in addition to the original one, T0 ≡ t. Also,
one slow spatial coordinate X1 is introduced as X1 = εx, in addition to original one,
x ≡ n. When studying the dissipative effects on the dynamics of modulated waves
propagating through the network system of Fig. 3.1, Yemélé et al. [1] have found that
the most important one comes from conductance G 1 in the series branch of the net-
work. Conductance G 2 in the shunt branch may then be disregarded when addressing
the linear propagation, while effects originating from G 2 and the nonlinearity appear
in the same perturbation equation; therefore we set  σ2 ≡ ε2 σ2 . Next, the solution
Vn (t) of the difference-differential system (3.2) is sought in the general form [2]
√        
Vn (t) = εu exp i θ + ε u20 + u22 exp 2i
θ + ε3/2 u30 + u33 exp 3i θ
    
+ε u40 + u42 exp 2i
2
θ + u44 exp 4i
θ (3.5)
      7/2 
+ε 5/2
u50 + u53 exp 3i
θ + u55 exp 5iθ + c.c. + O ε ,

where  θ = (k + iχ ) n − ωt =  kn − ωt is the rapidly varying complex phase, c.c.


stands for the complex conjugation, while u and ulm are complex amplitudes of
variables X1 , T1 , T2 (l = 2, 3, 4, 5, m = 0, 2, 3, 4, 5); here, ω is the real angular
frequency, while k = kp and χ are respectively the wavenumber and the spatial linear
parameter. The spatial linear parameter χ is introduced to take into account dissipa-
tive effects on the linear oscillations originating from conductance G 1 . Expanding
Vn±1 (t) into Taylor series and inserting Eq. (3.5) in (3.2) yields a power series in
ε1/2 and exp (iθ ), where θ = kn − ωt. Equating to zero coefficients of like powers
of ε1/2 , exp (iθ ) yields a series of equations for u and ulm .

Fig. 3.1 The schematic representation of the dissipative nonlinear discrete electrical transmission
network with a linear inductor placed in the series branch, and a nonlinear capacitor in the parallel
branch. Dissipative elements are connected in parallel with two linear inductors. This network is
built of N identical blocks. Reprint from Ref. [17], Copyright 2022, with permission from Springer
50 3 Transmission of Dissipative Solitonlike Signals …
 
At order ε1/2 , exp (iθ ) , we obtain that ω, k, and χ must satisfy the complex
equation
     
ω2 + 4iωσ1 u0 1 − cosh i
k − ω02 − 2u02 1 − cosh i
k = 0. (3.6)

Expanding Eq. (3.6) leads respectively to the following linear dispersion relation and
the spatial linear parameter
 
1 ω2 8σ12 − 1 + 2u02 + ω02
cos k =   (3.7a)
μ(ω) 2 u02 + 4ω2 σ12

and

χ (ω) = ln μ(ω) + μ2 (ω) − 1 , (3.7b)

where
⎡ ⎤
 ⎛ ⎞2
  2   2
 ⎢ 4σ12 ω2 ω2 − ω02 + u02 E22  E22 ⎥
2 2
4σ1 ω ω − ω0 + u0 E2
2 2 2 2
1 ⎢ ⎜ ⎟ ⎥
μ(ω) = 
2 ⎢1 + + ⎝1 + ⎠ − 2 ⎥,
 ⎣ 4u02 E12 2 2
4u0 E1 E1 ⎦

(3.7c)
with  
E1 = u02 + 4ω2 σ12 , E2 = ω02 + 2u02 − ω2 1 − 8σ12 . (3.7d)

To investigate the effects of the dissipative parameter σ1 on both the propagating


frequency fp = ω/2π and the linear dissipation parameter χ , we depict in Fig. 3.2a,
b the linear dispersion relation (3.7a) and the variation of the spatial linear dissipation
parameter χ given by Eq. (3.7b) for different values of σ1 . It is seen from Fig. 3.2a
that, at the first order with respect to σ1 (e.g., σ1 ∼ 10−2 ), the linear dispersion curve
of the network (3.7a) can be approximated by the well-known dispersion relation of
the typical bandpass filter,

ω2 − ω02 − 4u02 sin2 (k/2) = 0, (3.8)

obtained from Eq. (3.6) by setting σ1 = 0. Figure 3.2 shows that the spatial linear
dissipation parameter χ is an increasing function of the angular frequency ω in the
operating range of the network, meaning that the amplitude of the propagating wave
decays more rapidly when
 the frequency of input wave increases.
At order ε3/2 , eiθ , one gets

∂u ∂u
+ υg = iN22 e−2χn |u|2 u, (3.9)
∂T1 ∂X1
3.1 Chirped Lambert W-Kink Waves Propagation in a Lossy Electrical … 51

Fig. 3.2 a The linear dispersion curve of the network defined by Eq. (3.7a) for different values
of the dimensionless conductance σ1 : σ1 = 0 (the solid line), σ1 = 0.01 (the dashed-dotted line),
and σ1 = 0.03 (the dashed line). b The variation of the spatial linear dissipation parameter χ as
a function of frequency ω, according to Eq. (3.7b) for three values of σ1 : σ1 = 0.01 (the solid
line), σ1 = 0.015 (the dashed-dotted line), and σ1 = 0.02 (the dashed line). Network parameters
used to generate different plots are given in the text. Reprint from Ref. [17], Copyright 2022, with
permission from Springer

with
 
1 ω2 (3β − 2α (2α + N12 )) dω u0 (2σ1 ω + iu0 ) sinh i
k
N22 =    , υg = =−    ,
2 ω + 2iσ1 u0 1 − cosh i
k d
k ω + 2iσ1 u0 1 − cosh i k
(3.10)
where N12 given by

4αω2
N12 = . (3.11)
4ω2 − ω02 − 2u0 (u0 − 4iσ1 ω) {1 − cosh [2 (χ − ik)]}

Denoting by 
kc the critical wavenumber satisfying condition N22 (
kc ) = 0 and setting
  
k = k − kc we obtain that
     
N22 
kc + k − N22 
kc dN22
lim =
k→0 k d
k k=k c

so that  
    dN22
   
N22 k = N22 kc + k ≈ k .
d
k k=kc
52 3 Transmission of Dissipative Solitonlike Signals …
 
Assuming that   = 
 with Q
k = O (ε) and letting N22 ≈ εQ k dN22
d
k 
, we follow
k=
kc
 
 3/2  [18] and focus ourselves on a vicinity of kc , setting N22 = εQ
Kakutani and Michihiro
so that at order O ε , Eq. (3.9) becomes

∂u ∂u
+ υg = 0. (3.12)
∂T1 ∂X1

Equation (3.12) meansthat u is a function of two variables ξ = X1 − υg T1 and T2 , that


is, u(X1 , T1 , T2 ) = ψ ξ = X1 − υg T1 , T2 . We then conclude that in the reference
frame moving with group velocity υg = υgr + iυgi , the amplitude u of the signal
remains constant with respect to ξ = X1 − υg T1 . Thus, the rhs of Eq. (3.9) will be
shifted to the corresponding nonsecular condition at order ε5/2 , eiθ .
Equations of order ε5/2 eiθ lead to the following amplitude equation, a generalized
CQ CGL equation for the dynamics of modulated waves in the network system of
Fig. 3.1

∂ψ ∂ 2ψ ∂ψ ∂ψ ∗
i + P 2 + iRψ = Q1 |ψ|2 ψ + Q2 |ψ|4 ψ + iQ3 |ψ|2 + iQ4 ψ 2 ;
∂T2 ∂ξ ∂ξ ∂ξ
(3.13a)
here,
     
P=P  k = Rr + iRi , Q1 = Q1 
k = Pr + iPi , R = R  k = Q1r + iQ1i ,
     
Q2 = Q2 
k = Q2r + iQ2i , Q3 = Q3 
k = Q3r + iQ3i , and Q4 = Q4 k = Q4r + iQ4i

are complex functions of complex variable 


k = k + iχ , given by
  
υg υg + 4σ1 u0 sinh i k
P=−      , (3.14a)
2 ω + 2iσ1 u0 1 − cosh i k
σ2 u0 ω
R=    , (3.14b)
ω + 2iσ1 u0 1 − cosh ik
 
  dN22
Q1 = e−2χn k −kc , (3.14c)
d
k k=kc


 

ω2 e−4χn 2αN12 N33 − 16α NN4341 − 3β 4α 2 + N33 + 4αN12 + 2 |N12 |2
Q2 =      ,
2 ω + 2iσ1 u0 1 − cosh i k
(3.14d)
ωυg N41 [3β − 2α (2α + N12 )] − 2iαω2 N42
Q3 = 2e−2χn      , (3.14e)
N41 ω + 2iσ1 u0 1 − cosh ik
ωυg∗ [3β − 2α (2α + N12 )]
Q4 = e−2χn    . (3.14f)
ω + 2iσ1 u0 1 − cosh ik
3.1 Chirped Lambert W-Kink Waves Propagation in a Lossy Electrical … 53

Complex amplitudes ujk used in expressions (3.5) are given in terms of ψ by

u20 = 2αe−2χ n |ψ|2 , u22 = N12 ψ 2 , u30 = u40 = u50 = 0,


u33 = N33 ψ 3 , u44 = N44 ψ 4 , u55 = N55 ψ 5 , (3.15)
N42 ∂ψ N43 N ∂ψ N
u42 = 4 ψ − 8e−2χ n |ψ|2 ψ 2 , u53 = 6 53 ψ 2 − e−2χ n 52 |ψ|2 ψ 3 ,
N41 ∂X1 N41 N51 ∂X1 N51
(3.16)

where N12 and N22 are given, respectively by Eqs. (3.10) and (3.11), and other Njk
are

9ω2 (β − 2αN12 )
N33 = ,
ω0 − 9ω + 2u0 (u0 − 6iωσ1 ) (1 − cosh [3 (χ − ik)])
2 2
  
N41 = ω02 − 4ω2 + 2u0 (u0 − 4iωσ1 ) 1 − cosh 2i
k ,
       
N42 = u02 N12 − 4iωσ1 u0 N12 sinh 2i
k + 2υg σ1 u0 N12 1 − cosh 2i
k + iωα − iωN12 ,
(3.17)
     
N43 = N22 iω (α − N12 ) + σ1 u0 N12 1 − cosh 2ik + ω2 2α 2 N12 + αN33 − 3αβ − 3βN12 ,
  2 
16ω2 3βN12 − α N12 + 2N33
N44 = 2    ,
ω0 − 16ω2 + 2u0 (u0 − 8iωσ1 ) 1 − cosh 4i k


  
N51 = ω02 − 9ω2 + 2u0 (u0 − 6σ1 iω) 1 − cosh 3i
k ,
 
N52 = 36α 2 ω2 N33 + 18αω2 N44 − 27βω2 4αN12 + N12
2 − 54βω2 N
33
N43    
− 144αω2 + 6N22 3iω (2αN12 − N33 − β) + 2σ1 u0 N33 1 − cosh 3i
k ,
N41
(3.18)
   
N53 = υg 3iω (2αN12 − N33 − β) + 2σ1 u0 N33 1 − cosh 3i
k
N42  
− 12αω2 − u0 (u0 − 6iωσ1 ) N33 sinh 3ik .
N41

Because k is a function of the angular frequency ω as seen from Eq. (3.7a),


one concludes that coefficients (3.14a)–(3.14f) of Eq. (3.13a) are functions of the
propagating frequency f = fp = ω/2π and spatial linear parameter χ . It is important
to remember that coefficient Q1 of the cubic nonlinearity is produced by mismatch

k − kc (≡  k which is assumed to be ∼ O (ε)), where N22 ( kc ) = 0, with N22 given
by Eq. (3.10). The CQ-CGL equation (3.13a) is a generic amplitude equation that
governs the nonlinear evolution of the patterns near the criticality.
A number of works have been done for chirped pulses in special cases of the
CQ-CGL equation (3.13a) with terms such as the self-steepening, the self-frequency
shift, and the third-order dispersion. They are relevant to the design of solitary-wave-
based communications links, fiber-optical amplifiers, and optical pulse compressors
54 3 Transmission of Dissipative Solitonlike Signals …

[19, 20]. In the following, we consider the general case of Eq. (3.13a) and focus our
attention on its chirped Lambert W-kink solutions.

3.1.3 Baseband Modulational Instability Analysis

By mean of the CQ-CGL equation (3.13a), we establish in this subsection the con-
ditions under which a uniform wavetrain in the lossy network of Fig. 3.1 is unstable
against small perturbations with infinitesimally small wavenumbers. The MI of this
type is referred to as the baseband MI [21]. First of all, we note that Eq. (3.13a)
admits Stokes’ waves of the form

ψ(ξ, T2 ) = ρ0 exp (iK0 ξ − i0 T2 ) , (3.19)

where ρ0 is a constant real amplitude, and K0 and 0 are adjustable real wavenumber
and frequency, given by
 
0 = Pr K02 + (Q4r − Q3r ) ρ02 K0 + Q1r + Q2r ρ02 ρ02 + Ri , (3.20a)

with algebraic equation



Q2i ρ04 + (Q4i − Q3i ) K0 + Q1i ρ02 + Pi K02 − Rr = 0 (3.20b)

for ρ02 . Next, we focus on the situation when the “global existence condition (GEC)”

4Pi Q2i − (Q3i − Q4i )2 > 0 (3.21)

is satisfied. This avoids the situation when ρ02 → +∞ as K0 → +∞. In the limit
K0 → 0, Eq. (3.20b) leads to

−Q1i ± 2
Q1i + 4Rr Q2i
ρ02 → ,
2Q2i

which imposes the following restrictions called “local existence conditions (LECs)”
 
2
Q1i + 4Rr Q2i > 0andQ2i −Q1i ± 2
Q1i + 4Rr Q2i > 0. (3.22)

Using the network parameters (3.4), we show in Fig. 3.3d the domain (in term of
frequency ω) in which GEC (3.21) and LECs (3.22) are satisfied simultaneously; in
this figure, we have used the sign “+” in (3.22).
3.1 Chirped Lambert W-Kink Waves Propagation in a Lossy Electrical … 55

Fig. 3.3 (Color online) Domain of the existence of the Stokes wave solution (3.19) when Q2i =
0. a–c show the plots of, respectively, curve GEC = 4Pi Q2i − (Q3i − Q4i )2 associated with the
“global existence condition” (3.21), curve LEC1 = Q1i2 + 4R Q associated with the first of the
 r 2i 
“local existence conditions” (3.22), and curve LEC2 = Q2i −Q1i + 2 + 4R Q
Q1i r 2i associated

with the second of the “local existence conditions” (3.22) versus the propagating angular frequency
ω. Panel (d) simultaneously shows curves GEC and two LECs; here, the sign “+” refers to the
domain where both GEC (3.21) and two LECs (3.22) are satisfied simultaneously, while sign
“−” refers to the frequency range for which at least one of GEC (3.21) and two LECs (3.22) is
not satisfied. Different plots are produced for network parameters (3.4) with σ1 = 0.461 × 10−2 ,
σ2 = 0.01, kc = 0.1, and χc = 10−4 ; for the LECs (3.22), sign “+” is used. Reprint from Ref.
[17], Copyright 2022, with permission from Springer

Under the above GEC and LECs, one has

K0 min ≤ K0 ≤ K0 max ,

where K0 min and K0 max are given as


 2 
Q1i (Q4i − Q3i ) − 2 Pi Q2i Q1i + 4Rr Q2i − Rr Q2i (Q3i − Q4i )2
K0 min = < 0,
4Pi Q2i − (Q3i − Q4i )2
 2 
Q1i (Q4i − Q3i ) + 2 Pi Q2i Q1i + 4Rr Q2i − Rr Q2i (Q3i − Q4i )2
K0 max = > 0.
4Pi Q2i − (Q3i − Q4i )2
56 3 Transmission of Dissipative Solitonlike Signals …

It is seen from the dispersion relation (3.20a) that plane wave (3.19) is nonlinear,
hence the superposition principle cannot be applied to it. To investigate the MI of
the carrier wave under the GEC (3.21), we consider a small perturbation of the form

ψ(ξ, T2 ) = (ρ0 + δψ(ξ, T2 )) exp (iK0 ξ − i0 T2 ) , (3.23)

where δψ(ξ, T2 ) is a small perturbation, solution of the equation

∂δψ ∂ 2 δψ  
2 ∂δψ 2 ∂δψ

i +P + i 2PK0 − Q3 ρ − iQ4 ρ
∂T2 ∂ξ 2 0
∂ξ 0
∂ξ
 2 ∗

+ (Q3 − Q4 ) K0 − Q1 − 2Q2 ρ0 ρ0 δψ + δψ = 0.
2
(3.24)

Next, we seek the perturbation δψ(ξ, T2 ) in the form


 
δψ(ξ, T2 ) = b1 exp (iKξ + T2 ) + b∗2 exp −iKξ + ∗ T2 , (3.25)

K and  being the wavenumber and complex frequency of the modulation waves,
respectively. Following Chen et al. [21], we limit our consideration to perturbations
with infinitesimally small K. Substituting Eq. (3.25) in Eq. (3.24) yields the following
dispersion law for the perturbations:
 √
 X+ X2 + Y2
 = ± = Q1i + (Q4i − Q3i ) K0 + 2Q2i ρ02 ρ02 + Pi K 2 ±
2
⎡  √ ⎤
  −X + X 2 + Y2
+i ⎣ Q3r ρ02 − 2Pr K0 K ± ⎦, (3.26)
2

where X and Y are real quantities given in work [17]. Following the discussion carried
out in Ref. [17], one obtains, under GEC (3.21) and LECs (3.22), the following
baseband-MI condition for the Stokes waves (3.19):

Q1i + (Q4i − Q3i ) K0 + 2Q2i ρ02 > 0, (3.27)

where ρ02 is any solution of Eq. (3.20b) with Q2i = 0. Under the MI criterion (3.27),
the wavenumber K of the perturbation in the case when Pi < 0 satisfies the conditions

2
Q1i + 4Rr Q2i ! "
0≤K < 2
Q1i − 2
Q1i + 4Rr Q2i . (3.28)
2Pi Q2i

Relation (3.28) means that Q1i > 0 and Rr Q2i < 0. The propagating frequency f =
fp = ω/2π , as well as dimensionless conductances σ1 and σ2 significantly modify
the instability domain. In fact, different values of either fp or σ1 and σ2 correspond to
3.1 Chirped Lambert W-Kink Waves Propagation in a Lossy Electrical … 57

different instability diagrams. Each of these three parameters enhances the instability
of the system. This behavior is shown in Fig. 3.4 through the MI growth rate:
 √

X+ X2 + Y2
Re − (K, K0 , ρ0 ) = Q1i + (Q4i − Q3i ) K0 + 2Q2i ρ02 ρ02 + Pi K 2 − ,
2
(3.29)
when K0 = 0. Figure 3.4a–c show that the MI growth rate increases when each
of fp , σ1 , and σ2 increases. From the plots of Fig. 3.4, we can also conclude that
new ranges of wavenumber K may become available to the MI, causing expansion
of the MI domain. Figure 3.4a confirms the behavior shown in Fig. 3.2a, that is,
the propagating frequency f = fp increases when the dimensionless conductance σ1
increases.

Fig. 3.4 The instability growth rate as per Eq. (3.29) in the limit of K0 = 0. a The instability
growth rate for σ1 = 0.00461, σ2 = 0.01, and three values of the frequency: fp = 445.634 kHz
(the solid line), fp = 461.549 kHz (the dashed line), fp = 477.465 kHz (the dashed-dotted line). b
The instability growth rate for fp = 477.465 kHz, σ2 = 0.01, and three values of the dimensionless
conductance: σ1 = 0.00461 (the solid line), σ1 = 0.00561 (the dashed line), σ1 = 0.00661 (the
dashed-dotted line). c The instability growth rate for fp = 477.465 kHz, σ1 = 0.00461, and three
values of the dimensionless conductance: σ2 = 0.01 (the solid line), σ2 = 0.015 (the dashed line),
σ2 = 0.02 (the dashed-dotted line). Here, kc = 0.1, χc = 10−4 , and the network parameters are
taken as per Eq. (3.4). Reprint from Ref. [17], Copyright 2022, with permission from Springer
58 3 Transmission of Dissipative Solitonlike Signals …

3.1.4 Evolution of Chirped Lambert W-Kink Pulses


in the Network of Fig. 3.1

Here, we seek for analytical solutions for chirped Lambert W-kink signals in the
lossy network of Fig. 3.1. For this aim, we employ the ansatz

ψ(ξ, T2 ) = ρ(ζ ) exp [i (ϕ(ζ ) − ωk T2 )] , (3.30a)

where ζ = α0 ξ − β0 T2 is the travelling coordinate, ρ and ϕ are real functions of ζ , ωk


is a real constant, and ϕ(ζ ) − ωk T2 is a phase function. For ansatz (3.30a), υp = β0 /α0
and 1/α0 are, respectively, the group velocity and width of the pulse’s envelope,
while the corresponding intensity of the propagating pulse is |ψ(ξ, T2 )|2 = ρ 2 (ζ ).
With ansatz (3.30a), the chirping δph (ξ, T2 ) across the pulse is given as δph (ξ, T2 ) =
− (∂/∂ξ ) (ϕ(ζ ) − ωk T2 ) = −α0 d ϕ(ζ )/d ζ , which is time-dependent, unless ϕ is a
linear function of variable ζ .
Assuming that the frequency chirp depends on the wave’s amplitude ρ through
relation d ϕ(ζ )/d ζ = A2 ρ 2 + A0 and inserting Eq. (3.30a) into Eq. (3.13a), we obtain
that ρ must simultaneously satisfy the following two equations
! "
d 2ρ α0 Q3i + α0 Q4i − 4A2 α02 Pi 2 2A0 α02 Pi d ρ
+ ρ −
dζ2 Pr α02 Pr α02 dζ
ωk − Ri + A0 β0 − A20 α02 Pr
+ ρ
Pr α02
A2 β0 − Q1r − 2A0 A2 α02 Pr + A0 (α0 Q3r − α0 Q4r ) 3
+ ρ
Pr α02
A2 (α0 Q3r − α0 Q4r ) − A22 α02 Pr − Q2r 5
+ ρ = 0, (3.31)
Pr α02

# $
d 2ρ 4A2 α02 Pr − α0 Q4r − α0 Q3r 2 2A0 α02 Pr − β0 d ρ Rr − A20 α02 Pi
+ ρ + + ρ
dζ2 Pi α02 Pi α02 dζ Pi α02
A0 (α0 Q3i − α0 Q4i ) − 2A0 A2 α02 Pi − Q1i
+ ρ3
Pi α02
A2 (α0 Q3i − α0 Q4i ) − A22 α02 Pi − Q2i
+ ρ 5 = 0. (3.32)
Pi α02

Asking that Eq. (3.31) coincides with Eq. (3.32), we obtain

Pr β0
A0 = , (3.33)
2 |P| α02
2

Pr (Q3r + Q4r ) + Pi (Q3i + Q4i )


A2 = , (3.34)
4α0 |P|2
3.1 Chirped Lambert W-Kink Waves Propagation in a Lossy Electrical … 59

Pr Rr + Pi Ri Pr β02
ωk = − , (3.35)
Pi 2 |P|2 α02
β0 4 |P|2 (Pi Q1r − Q1i Pr )
= ,
α0 2Pr (Pi (Q3r − Q4r ) − Pr (Q3i − Q4i )) + Pi Pr (Q3r + Q4r ) + Pi2 (Q3i + Q4i )
(3.36)

under the condition that coefficients of the amplitude equation (3.13a) satisfy the
condition

(Pr (Q3r + Q4r ) + Pi (Q3i + Q4i )) (Pi (Q3r − Q4r ) − Pr (Q3i − Q4i )) + 4 |P|2 (Pr Q2i − Pi Q2r ) = 0.
(3.37a)
Under assumptions (3.33)–(3.37a), we arrive to the following second order ordinary
differential equation for ρ

d 2ρ   dρ
+ M2 ρ 2 + M0 + N1 ρ + N3 ρ 3 + N5 ρ 5 = 0, (3.38)
dζ2 dζ

where M0 , M2 , N1 , N3 , and N5 are given as


 
Pi β0 Pr (Q3i +Q4i )−Pi (Q3r +Q4r ) P2 β 2
M0 = − |P| 2 2 , M2 =
α
, N1 = α12 Rr
− 4|P|r 4 α02 ,
 2 2|P| α0 0
2 Pi
0 0

− PQα1i2 + 2|P|Pr2 α2 ( r i ) 3i 2P 4i|P|2 r i 3r 4r


2P +P (Q +Q )−P P (Q +Q ) β0
N3 = α0
, (3.39)
i 0 0 i
α0 (Q3r −Q4r )A2 −α02 Pr A22 −Q2r
N5 = Pr α02
.

Equation (3.38) admits a variety of solutions, such as periodic, kink, and soliton-
like ones. When the network parameters are all fixed, except linear conductance G 1 ,
Eq. (3.37a) establishes relationship g0 (ω, σ1 ) = 0 between frequency fp = ω/2π
and spatial linear parameter χ , g0 (ω, σ1 ) being the function that coincides with the
lhs of Eq. (3.37a) times exp (4χ n), that is,

g0 (ω, σ1 ) = exp (4χ n) (Pr (Q3r + Q4r ) + Pi (Q3i + Q4i ))

× (Pi (Q3r − Q4r ) − Pr (Q3i − Q4i )) + 4 |P|2 (Pr Q2i − Pi Q2r ) .

To prove that equation g0 (ω, σ1 ) = 0 admits real solutions in σ1 , we depict in Fig. 3.5
the dimensionless conductance σ1 versus frequency ω, in an implicit form defined
by g0 (ω, σ1 ) = 0 [any set of values (ω, σ1 ) on the curve of Fig. 3.5 satisfies the
equation (3.37a)]. It is seen from Fig. 3.5 that for given propagating frequency ω,
there exist more than one value of dimensionless conductance σ1 for which the set
(ω, σ1 ) satisfies relationship (3.37a).
To derive chirped kink solitonlike solutions of the CQ-CGL equation (3.13a), we
impose to the first derivative d ρ/d ζ to be a polynomial of the third degree in ρ. The
chirped Lambert W-kink solution of Eq. (3.13a) is expressed via the Lambert W-kink
60 3 Transmission of Dissipative Solitonlike Signals …

Fig. 3.5 The variation of


dimensionless conductance
σ1 as an implicit function
defined by equation
g0 (ω, σ1 ) = 0, showing
(ω, σ1 ) that satisfy condition
(3.37a). The network
parameters used to generate
different plots are given in
the text. Reprint from Ref.
[17], Copyright 2022, with
permission from Springer

solitonlike solution of Eq. (3.38). To this end, we seek the solution to Eq. (3.38) in
the form r
ρ(ζ ) = q + , (3.40a)
1 + f [g(ζ )]

where q and r are two real constants, and f and g are two real functions of variable
ζ satisfying relationship

df f0 + f1 (q + r) + (f0 + f1 q) f (ζ )
= , (3.40b)
dg g(ζ ) (1 + f (ζ ))

f0 and f1 being two real parameters. If we differentiate Eq. (3.40a) in respect with ζ
and replace f by f = (r + q − ρ) / (ρ − q), we obtain

dρ 1 1 dg
= − (ρ − q)2 (f0 + f1 ρ) , (3.40c)
dζ r g dζ

where f0 and f1 are two real parameters. Asking to the rhs of Eq. (3.40c) to be a
third-degree polynomial in ρ, we find that g −1 d g/d ζ must be a constant and f1 = 0.
Hence, we arrive to g(ζ ) = λ0 exp (μ0 ζ ), where 
λ0 is a real constant. Replacing now
g by this expression in Eq. (3.40b) and integrating the resulting equation, we find
that f = f (ζ ) = 0 must be any solution of the algebraic equation
  ! " # $
rf1 f0 + qf1 0 − (f 0 + qf 1 ) 2
1+ +f exp − f = exp K μ0 ζ , (3.41a)
f0 + qf1 rf1 rf1
3.1 Chirped Lambert W-Kink Waves Propagation in a Lossy Electrical … 61

0 being a constant of integration. Restricting ourselves to the special case when


K
f1 = −1/r and f0 = (r + q) /r, Eq. (3.41a) gives
 
0 ,
f (ζ ) = W exp μ0 ζ + K (3.42)

where W = W (z) is the Lambert W function [10]. Inserting Eq. (3.42) in Eq. (3.40a)
leads to r
ρ(ζ ) = q +   . (3.43)
1 + W exp μ0 ζ + K 0

Inserting now g = 
λ0 exp (μ0 ζ ) into Eq. (3.40c), we find, after some algebraic manip-
ulations, that %   &
5M0 5M0 M0
(r, q, μ0 ) = ± ,∓ ,− , (3.44)
3M2 3M2 3

if N1 = 29 M02 , N3 = − 15
4
M0 M2 , N5 = 2
M 2,
25 2
and M0 M2 > 0,
%   &
M0 M0
(r, q, μ0 ) = ∓3 − , ± − , 3M0 , (3.45)
M2 M2

if M0 M2 is negative and N1 = N3 = N5 = 0, and


%   &
5M0 5M0
(r, q, μ0 ) = ∓2 − ,± − , 4M0 , (3.46)
M2 M2

if
M0 M2 < 0, N1 = 2M02 , N3 = (4/5)M0 M2 , andN5 = (2/25)M22 .

Here, M0 , M2 , N1 , N3 , and N5 are given by Eq. (3.39).


Hence, by taking μ0 , q, and r from Eqs. (3.44)–(3.46), equation (3.43) will be a
solution of Eq. (3.38). Now, if we insert Eq. (3.43) together with (r, q, μ0 ) in Eq.
(3.30a), we arrive to the following (rather cumbersome) chirped Lambert W-kink
solution of the CQ-CGL equation (3.13a):
% &
r 
ψ(ξ, T2 ) = q +   exp i (ϕ(ξ, T2 ) − ωk T2 ) ,

1 + W exp μ0 (α0 ξ − β0 T2 ) + K0
(3.47)
with
62 3 Transmission of Dissipative Solitonlike Signals …

  2qrA2    
ϕ(ξ, T2 ) = A0 + A2 q2 ζ + ln W exp μ0 ζ + K0
μ
⎡  0  ⎤'
μ0 ζ +K0 '
A2 r 2 W e '
+ ln ⎣  
 ''
⎦ ; (3.48)
μ0 1+W e μ 0 ζ +K 0
'
ζ =α0 ξ −β0 T2

the corresponding chirp reads


⎡ % &2 ⎤
r
δph (ξ, T2 ) = −α0 ⎣A0 + A2 q +   ⎦. (3.49)
1 + W eμ0 (α0 ξ −β0 T2 )+K0

Here, K0 is an arbitrary real constant, α0 = 0 and β0 = 0 are two arbitrary real
parameters satisfying condition (3.36), A0 , A2 , and ωk are given by Eqs. (3.33)–
(3.35), respectively, ϕ(ξ, T2 ) = ϕ(ζ )|ζ =α0 ξ −β0 T2 , and
⎡ % &2 ⎤
( (
  r
ϕ(ζ ) = A0 + A2 ρ 2 d ζ = ⎣A0 + A2 q +   ⎦ dζ
1 + W eμ0 ζ +K0
(

= A0 ζ + A2 q ζ + 2qrA2
2
 
1 + W eμ0 ζ +K0
(

+r 2 A2    2
1 + W eμ0 ζ +K0
( (
2qrA2 dx A2 r 2 dx
= A0 ζ + A2 q2 ζ + + ,
μ0 x [1 + W (x)] μ0 x [1 + W (x)]2

with x = eμ0 ζ +K0 .

Setting w = W (x), one has x = wew and dx = (1 + w) ew dw, which leads to Eq.
(3.48). Equations (3.47) and (3.49) show that both the chirped Lambert W-kink
solution of Eq. (3.13a) and the corresponding chirping are expressed in terms of the
Lambert W function [10].
It is important to note that Eq. (3.38) also admits kink solitonlike solution, solution
of the following first-order ODE for ρ,


= λ1 ρ + λ3 ρ 3 , (3.50)

where λ1 and λ3 = 0 are real parameters. A differentiation of Eq. (3.50) yields

d 2ρ
− 3λ23 ρ 5 − 4λ1 λ3 ρ 3 − λ21 ρ = 0. (3.51)
dζ2
3.1 Chirped Lambert W-Kink Waves Propagation in a Lossy Electrical … 63

Inserting now Eq. (3.50) in Eq. (3.38) and comparing the resultant equation with Eq.
(3.51) determine the wave parameters,

−M0 ± M02 − 4N1 −M2 ± M22 − 12N5


λ1 = , λ3 = , (3.52)
2 6
under the conditions that

M02 − 4N1 ≥ 0, M22 − 12N5)≥ 0,


  
6N3 − 2M0 M2 ± M2 M0 − 4N1 ∓ M0 M22 − 12N5 + 2 M02 − 4N1 M22 − 12N5 = 0.
2

(3.53)
Under condition λ1 λ3 < 0, we integrate Eq. (3.50) to obtain the kink soliton solution

−λ3 λ1 ζ
λ1 λ0 λ1
e − e−λ1 ζ
ρ(ζ ) = ± − [1 + tanhλ (λ1 ζ )], with tanhλ (λ1 ζ ) ≡ −λ3 λ1 ζ
,
2λ3 + e−λ1 ζ
λ0 λ1
e
(3.54)
λ0 > 0 being a real constant. By inserting Eq. (3.54) into Eq. (3.30a), we arrive to
the following chirped kink solution of Eq. (3.13a):

λ1
ψ(ξ, T2 ) = ε0 − (1 + tanhλ [λ1 (α0 ξ − β0 T2 )]) exp [i {ϕ(ξ, T2 ) − ωk T2 }] ,
2λ3
(3.55)
where ε0 is a real number satisfying condition |ε0 | = 1 and
! "
A2 λ3 2λ1 (α0 ξ −β0 T2 )
ϕ(ξ, T2 ) = A0 (α0 ξ − β0 T2 ) + ϕ0 − ln 1 − e ; (3.56)
2λ3 λ0 λ1

the corresponding chirp being


! "
A2 λ1
δph (ξ, T2 ) = −α0 A0 − (1 + tanhλ [λ1 (α0 ξ − β0 T2 )]) . (3.57)
2λ3

Here, A0 , A2 , and ωk are given by Eqs. (3.33)–(3.35), respectively, λ1 and λ3 are


given by Eq. (3.52), α0 = 0 and β0 = 0 being real parameters satisfying condition
(3.36).
Now, we turn to the transmission of chirped Lambert W-kink pulses in the lossy
network of Fig. 3.1. For this aim, we go back to dimensional parameters and original
variables and use the decomposition (3.5) of voltage Vn (t). In expansion
' ' (3.5), we
neglect all harmonics, in comparison with the fundamental one u, 'ujl '  |u| for all j
and l, so that the shape of the electrical pulse will be well approximated by the shape
of the fundamental harmonic, the expression for which is
64 3 Transmission of Dissipative Solitonlike Signals …

Fig. 3.6 The real and imaginary parts of complex group velocity υg versus frequency ω for different
values of dissipative parameter σ1 . (a): The variation of real part υgr of the group velocity for
σ1 = 0.461 × 10−2 (the solid line), σ1 = 0.04 (the dashed line), and σ1 = 0.08 (the dashed-dotted
line). (b): The variation of imaginary part υgi of the group velocity for σ1 = 0.461 × 10−2 (the
solid line), σ1 = 0.561 × 10−2 (the dashed line), and σ1 = 0.661 × 10−2 (the dashed-dotted line).
Different plots are produced for network parameters (3.4). Reprint from Ref. [17], Copyright 2022,
with permission from Springer

!   "
√ ε β0 Pr (Q3r + Q4r ) + Pi (Q3i + Q4i ) 2
Vn (t) = 2 ε exp −χ n + υ gi P r + q t
2 |P|2 α0 2
⎛ ⎞
⎜ ⎟
⎜ r ⎟
×⎜
⎜q + %    & ⎟

  β
⎝ εα0 μ0 n− υgr +ε α0 t+φ0W ⎠
1 + W cos εα0 μ0 υgi t × e 0

× cos [L (n, t) + NL (n, t)] . (3.58)

In this equation (3.58), L = L (n, t) and NL = NL (n, t) are given respectively
by
      
L = k + εα0 A0 + A2 q2 n − ω + ωk ε2 + ε A0 + A2 q2 α0 υgr + εβ0 t,
! 
  "
2qrA2   εα μ n− υgr +ε αβ0 t+φ0W
NL = ln W cos εα0 μ0 υgi t e 0 0 0 (3.59)
μ0
⎡ 
   ⎤
  εα0 μ0 n− υgr +ε αβ0 t+φ0W
W cos εα μ υ t e 0
A2 r 2 ⎢ ⎥
0 0 gi
+ ln ⎢
⎣ 
   ⎥ ,

μ0   β0
εα μ n− υgr +ε α t+φ0W
1 + W cos εα0 μ0 υgi t e 0 0 0

where k and ω are the wavenumber and angular frequency of the carrier wave, χ is the
spatial linear parameter given by Eq. (3.7b), α0 = 0 and β0 = 0 are real parameters
satisfying equation (3.36), φ0W and φ0K are two arbitrary real parameters, 0 < ε  1
is a small dimensionless parameter, υgr and υgi are, respectively, the real and the
imaginary parts of complex group velocity υg defined in Eq. (3.10), A0 , A2 , and ωk
are given, respectively, by Eqs. (3.33), (3.34), and (3.35), and (μ0 , q, r), taken from
3.1 Chirped Lambert W-Kink Waves Propagation in a Lossy Electrical … 65

Eqs. (3.44)–(3.46), are given by Eqs. (3.44)–(3.46). In general, υgr is positive, while
υgi is negative. This behavior is seen in Fig. 3.6, in which υgr and υgi are plotted
versus the angular frequency ω for different values of dimensionless conductance σ1 .
As we can see from Fig. 3.6a, b, υgr and vgi severally increases and decreases with
the increase of σ1 . The presence of factor exp [−χ n] in Eq. (3.58) indicates that the
dissipative effects, introduced by conductance G 1 in the series branch of the network
of Fig. 3.1, lead to a decay of the amplitude as the Lambert W-kink pulses propagate
in the lattice. Depending on the sign of the parameter λ given by
 
β0 Pr (Q3r + Q4r ) + Pi (Q3i + Q4i ) 2

λ = υgi Pr + q ,
α0 2


ε
λ
the presence of factor exp t
in Eq. (3.58) shows that dissipative effects intro-
2|P|2
duced by G 1 lead to either decrease (for  λ < 0) or increases (when  λ > 0) of the
amplitude in the course of the propagation of the Lambert W-kink signals through the
network. Because G 1 accounts for the loss in the network of Fig. 3.1, 
λ must be nega-
tive so that the wave amplitude decreases. One then concludes that the self-steepening
term Q3 and self-frequency shift Q4 can be used to modulate the amplitude of the
Lambert W-kink signals propagating in the dissipative network. As a consequence
of the presence of the dissipation, the wave’s phase is a superposition of linear and
nonlinear phases L (n, t) and NL (n, t), given by Eq. (3.59). Equation (3.58) reveals
that the center of the chirped Lambert W-kink pulses moves with a constant speed,
υc = υgr + εβ0 /α0 , which is different from the group velocity. Equation (3.58) also
indicates that the width of the W-kink pulses is 1/ (εα0 μ0 ), and, according to the
expression for μ0 [see Eqs. (3.44)–(3.46)], it is proportional to α0 /β0 . Therefore,
the width of the W-kink pulse remains constant during its propagation. With the
use of network parameters (3.4) with σ1 = 0.766 × 10−2 and σ2 = 1.152 × 10−2 ,
and using (r, q, μ0 ) given by Eq. (3.45), a relevant set of values of the solution
parameters is chosen to depict, in Fig. 3.7, the temporal and spatial (left and right
panels, respectively) profiles of the chirped Lambert W-kink signals, propagating at
frequency fp = 1171.38 kHz. For these data, condition (3.21) is violated, meaning
that the propagating frequency fp = 1174.05 kHz does not belong to the domain of
the baseband MI. It is important to notice that for the signals shown in Fig. 3.7d–f,
the center of the pulse is located (approximately) at n = 800, n = 809, and n = 813,
respectively, meaning that the pulse moves with a positive speed. Thus, Fig. 3.7
demonstrates that the dynamics of the Lambert W-kink signal is different from that
of the well-known kink soliton, cf. Ref. [22].
In conclusion, we have studied analytically in this Section the results for the
baseband MI and the transmission of chirped Lambert W-kink signals through a
lossy electrical network. First, we have derived a generalized CQ CGL equation
with self-steepening and self-frequency shift that governs the dynamics of slowly
modulated waves propagating in our network system. The spatial-decay rate χ is
found in the analytical form, including its dependence on the wave’s frequency. The
66 3 Transmission of Dissipative Solitonlike Signals …

Fig. 3.7 The propagation of chirped Lambert W-kink signals in the network at frequency fp =
1174.05 kHz. Plots in the left panels show the evolution of the chirped Lambert W-kink signal at
different cells of the network for φ0W = 40: a n = 0, b n = 5, and c n = 10. Plots (d)–(f) in the right
panels show the spatial shape of the chirped Lambert W-kink signal with φ0W = −800 at different
times, t = 0.0 μs, t = 0.18 μs, and t = 0.26 μs, respectively. Different curves
 are produced for
voltage (3.58) and parameters ε = 10−3 , kc = 2.75, χc = 10−4 (note that 2π fp , σ1 satisfies Eq.
(3.37a)). Other parameters are given in the text. Reprint from Ref. [17], Copyright 2022, with
permission from Springer

derived amplitude equation predicts the linear stability of slowly modulated Stokes’
waves, showing how the dissipative elements of the network affect the baseband
MI of the system. Further, our investigations have revealed that the model under
the consideration supports the propagation of a variety of chirped pulses including
the chirped Lambert W-kink signals. Analytical expressions for the chirped phase,
ϕ(ξ ) − ωk t, and nonlinear frequency change δph (x, t) across the pulse are obtained
as functions of the intensity of the signal, while its amplitude can be modulated by
varying parameters Q3 and Q4 of the self-steepening and self-frequency shifts. The
derived chirped Lambert W-kink solution of the amplitude equation (3.13a) is then
used to investigate analytically the effects of the dissipation on the amplitude of
3.2 Spatiotemporal Modulation of Damped Solitonlike Wave … 67

the Lambert W-kink soliton; we have obtained that its width and velocity remain
constant as the wave propagates in the network. Our results show that the effects of
dissipative losses in the series branch of the network are stronger than those resulting
from the dissipative losses in the shunt branch. Results of the present Section can
be useful for exploring the dynamics of Lambert W-kink pulses in other complex
systems, which, in the first approximation, may be modelled by equations of the
CQ-CGL type, including the self-steepening and self-frequency shift.

3.2 Spatiotemporal Modulation of Damped Solitonlike


Wave Signals in a Nonlinear RLC Transmission
Network

3.2.1 Introduction

Focusing our ourselves on lossy nonlinear transmission networks, we present in this


Section a method to investigate analytically spatiotemporal nonlinear modulated
waves with damped amplitude (in both time and space) in dissipative NLTNs .
Although device and transmission network losses can degrade the performance of
NLTNs [23] and cannot be discarded in realistic circuit simulations, we believe that
in real nonlinear dispersive media, the coexistence of the dissipative effect with the
nonlinear and dispersive effects may play some roles in a wave generation as well
as in its propagation.
Thus, the model studied in this Section is a distributed model of a lossy NLTN
which can be viewed as a multiple combination of small RLC circuit segments shown
in Fig. 3.8 [24–26]. In this model, the series inductance is due to magnetic field effects,
while the capacitance is due to electric field coupling between the lines. The losses in
this model are depicted by the series and the shunt resistors which represent the finite
conductivity of the conductors and the dielectric insulator between the conductors,
respectively. The linear resistances R1 and R2 that account for the network losses, the
linear inductance L, and the voltage-dependent capacitance C are the relevant network
parameters. In this study, we assume that the model under consideration is formed of
N identical cells, each of which consists of a linear inductive element of inductance
L and a linear resistor with resistance R1 in the series branch constituting the linear
dispersive element, and a nonlinear capacitor of voltage-dependent capacitance C
and a linear resistor with resistance R2 in the shunt branch. The nonlinear element
of this model is the voltage-dependent capacitor which is assumed to be biased by
an equilibrium constant voltage V0 (alias bias voltage of the capacitor) and depends
on the voltage Vn across the n-th capacitor.
A complex cubic GL equation governing slowly modulated wave propagation
through our model is derived. Considering linear wave propagating in the network
68 3 Transmission of Dissipative Solitonlike Signals …

Fig. 3.8 Schematic representation of the one cell of the lossy NLTL with the linear inductor in
series branch and the nonlinear capacitor in parallel branch. Conductance (dissipative elements) are
connected in series on the two elements. Reprint from Ref. [27], Copyright 2022, with permission
from Springer

under consideration, we derive in terms of the propagating frequency the spatial


decreasing rate alias linear dissipation parameter and show that its must important
contribution comes from the dissipative element of the shunt branch. The MI crite-
rion of modulated Stokes wave propagating in the network is investigated and the
analytical expression of the MI gain is derived. We show that in the case of weak
dissipation, there are no significant changes for the bandwidth frequency where the
network may exhibit MI. Exact and approximative envelope solitonlike solutions of
the amplitude equation are presented and used to investigate analytically the dynam-
ics of spatiotemporal modulated damped signals along our network system. We also
show that the solution parameters can be used for managing the evolution of the
envelope soliton signals along the network. Our investigations show that the sig-
nal amplitude decays in both space and time, while the velocity remains constant
when the envelope soliton signal propagates along the dissipative network under
consideration.
In order to investigate the dynamics of spatiotemporal modulated damped signals
in the network model of Fig. 3.8, we assume, for low voltage at cell n, the voltage-
dependent capacitance of the form [25, 27–34]

dQn  
C(V0 + Vn ) = ≈ C0 1 − 2αVn + 3βVn2 , (3.60)
dVn

where C0 = C(V0 ), and α and β are nonlinear coefficients that determine the electric
charge Qn stored in the n- th capacitor in the line, and Vn is the voltage across the n- th
capacitor. Here, the subscript n stands for the cell number in the network. According
to the sign of α and β, Eq. (3.60) can be the second order curve fitting for the diode
characteristics or the MOS varactor characteristics; for the diode, both α and β are
positive, while for the MOS, they are negative [16]. For the network system under
consideration, we assume α and β to be positive. During computation, we follow
Sekulic et al. [24] and use the below values of the network parameters
3.2 Spatiotemporal Modulation of Damped Solitonlike Wave … 69

Fig. 3.9 Linear dispersion curve of the network showing the angular frequency ω as a function of
the wavenumber k defined implicitly by Eq. (3.66) for different sets of the dimensionless resistances.
a σ1 = σ2 = 0 (solid line) and σ1 = 1.152 × 10−2 and σ2 = 0.461 × 10−2 (dashed line); b σ1 =
σ2 = 0 (solid line), σ1 = 0.461 × 10−2 and σ2 = 0 (dashed line), and σ1 = 0.15 and σ2 = 0 (dotted
line); c σ1 = σ2 = 0 (solid line), σ1 = 0 and σ2 = 0.461 × 10−2 (dashed line), and σ1 = 0 and
σ2 = 00.15 (dotted line); d σ1 = σ2 = 0 (solid line), σ1 = 0.461 × 10−2 and σ2 = 0.461 × 10−2
(dashed line), and σ1 = 0.15 and σ2 = 00.15 (dotted line). Reprint from Ref. [27], Copyright 2022,
with permission from Springer

L = 10µH, α = 0.21V−1 , β = 0.0197V−2 , V0 = 2V, C0 = 10pF. (3.61)

Assuming voltage across the capacitor to be Vn + V0 and applying the Kirchhoff


laws, we arrive to the following difference-differential system
70 3 Transmission of Dissipative Solitonlike Signals …

d 2 Vn dVn d
+ u02 (2Vn − Vn−1 − Vn+1 ) + 2u0 σ1 + 2u0 σ2 (2Vn − Vn−1 − Vn+1 )
dt 2 dt dt
d2   d  2 
= 2 αVn2 − βVn3 + 2u0 σ1 αVn − βVn3
dt dt
d   2   
+ 2u0 σ2 α 2Vn − Vn−1 − Vn+1
2 2
− β 2Vn3 − Vn−1 3
− Vn+1
3
. (3.62)
dt

Here, u0 = /1 LC0 is the characteristic angular frequency of the network, and σ1
and σ2 are two dimensionless resistances, related to resistances R1 and R2 as follows

R1 R2
= 2u0 σ1 , = 2u0 σ2 . (3.63)
L L

3.2.2 Linear Dispersion Relation and Spatial Decreasing


Rate

Following Yemélé et al. [32], the expression of propagating linear waves of initial
amplitude Vm and angular frequency ω can be written as
  
Vn (t) = Vm exp i 
kn − ωt + c.c., (3.64)

where c.c. stands for the complex conjugation and  k = k + iχ , k and χ being respec-
tively the real wavenumber and the real spatial decreasing rate (alias linear dissipation
parameter) to be determined. Inserting Eq. (3.64) into Eq. (3.62) and ignoring non-
linear terms yield the linear algebraic system
⎛ ⎞
u02 −σ2 u0 ω    2 1 2 
⎝ ⎠ cos [k] cosh [χ ] = u0 − 2 ω
. (3.65)
sin [k] sinh [χ ] ω (σ1 + 2σ2 )
ωσ2 u0

It is obvious that system (3.65) has a mathematical sense when k = 0 only if


ω(0) = χ (0) = 0; we also notice that it is incompatible when k = π . Therefore,
it is reasonable for us to work with the wavenumber k taken from the zone
0 ≤ k ≤ π0 < π . Solving system (3.65) in the region 0 < k ≤ π0 for cos [k] cosh [χ ]
and sin [k] sinh [χ ] yields respectively the following linear dispersion relation and
spatial decreasing rate

cos2 [2k] + z1 (ω) cos [2k] + z2 (ω) = 0 (3.66)

and ! "
χ = ln z3 (ω, k) + 1 + z32 (ω, k) , (3.67)
3.2 Spatiotemporal Modulation of Damped Solitonlike Wave … 71

Fig. 3.10 Evolution plot of the spatial decreasing rate χ (cell−1 ) versus the angular frequency ω for
different values of the dimensionless resistances σ1 and σ2 : a σ1 = 0 and σ2 = 0.01; b σ1 = 0.01
and σ2 = 0; and c σ1 = 1.152 × 10−2 and σ2 = 0.461 × 10−2 . Reprint from Ref. [27], Copyright
2022, with permission from Springer

where
 2   2
u02 2ω2 σ2 (σ1 + 2σ2 ) − ω2 + 2u02 + 2ωu02 (σ1 + 2σ2 ) + ωσ2 ω2 − 2u02
z1 (ω) = −  2 ,
2u02 u02 + ω2 σ22
!    
z2 (ω) =
1 2 2 2 2 2 2 2 2 2 2 2
 2 2u0 2ω σ2 (σ1 + 2σ2 ) − ω + 2u0 − 2u0 u0 + ω σ2 − u0
2u02 u02 + ω2 σ22
(3.68)
 2   2 "
× 2ω2 σ2 (σ1 + 2σ2 ) − ω2 + 2u02 − 2ωu02 (σ1 + 2σ2 ) + ωσ2 ω2 − 2u02 ,
 
2 2 2
ω 2u0 (σ1 + 2σ2 ) + σ2 ω − 2u0
z3 (ω, k) =   .
sin [k] 2u u2 + ω2 σ 2
0 0 2

Because of the non-negativity of z3 (ω, k) for k ∈ [0, π0 ], the linear dissipation param-
eter χ (k, ω(k)) is positive for all k ∈]0 , π0 ]. Therefore, the amplitude of linear
waves defined by Eq. (3.64) is an exponentially decreasing function of the network
cell number n (this is clearly seen if we replace in Eq. (3.64)  k by its expression

k = k + iχ ).
In the special case of a lossless network (σ1 = σ2 = 0), the linear dissipation
parameter (3.67) gives χ (ω) = 0, while the linear spectrum (3.66), under the con-
dition that ω(0) = 0, reduces to the well-known dispersion relation of a band-pass
filter ! "
2 k
ω = 4u0 sin
2 2
, (3.69)
2
72 3 Transmission of Dissipative Solitonlike Signals …

which possesses a gap ω0 = 0 at k = 0 and is also limited by the cut-off angular
frequency ωc = 2u0 sin π20 when k = π0 . Therefore, relation (3.66) establishes the
relation between the wavenumber k and the angular frequency ω in the general case
of dissipative network of Fig. 3.8, and the propagation of plane waves and modulated
waves in our dissipative system will take place in the frequency range of the allowed
band, that is, fp ∈ [f0 , fc [, where f0 = 0 and fc = ωc /(2π ). Throughout this Section,
we use, without loss of generality, the value π0 = π/2 so that the propagating fre-
quency fp = ω/2π will be an increasing function of wavenumber k in [0, π0 ].
As one can see from plots of Fig. 3.9, to different sets of the dimensionless
resistances σ1 and σ2 correspond different linear dispersion curves. It is seen from
Fig. 3.9a, that the linear dispersion curve of the network associated with Eq. (3.66)
can be well approximated by Eq. (3.69) for small enough σ1 and σ2 . We can see
from plots of Fig. 3.9b that the introduction of the series dissipative elements in our
system decreases the propagation frequency fp = ω/2π . Figure 3.9c reveals that the
introduction of the dissipative elements in the shunt branches decreases [increases]
the propagating frequency associating with low [high] wavenumber k.
In order to analyze the contribution of the two resistances in the resulting mag-
nitude of spatial decreasing rate χ , we have displayed, for the network parameters
(3.61) and different values of σ1 and σ2 , χ given by Eq. (3.67) in Fig. 3.10 as a func-
tion of the angular frequency ω. It is seen from different plots of Fig. 3.10 that the
spatial decreasing rate χ is more manifested for frequencies close  to f c = ωc /2π and
is an increasing function of the frequency in the operating range f0 , fc = [0, ωc /2π ]
of the network. Therefore, the wave amplitude will decrease more rapidly when the
frequency of input wave increases. When comparing Fig. 3.11a, b, we observe that
the linear dissipative parameter χ associated to σ2 only (Fig. 3.10a) is greater than
that induced by σ1 only (Fig. 3.10b). Hence, the must important contribution to χ
comes from the resistance R2 in the shunt branch of the network and consequently,
this resistance R2 cannot be neglected in the analytical study of wave propagation in
the network under consideration.
 Plots of Fig. 3.10 also reveal that for frequencies in
the operating range f0 , fc of the network, the spatial decreasing rate χ is very small
(0 ≤ χ ≤ 0.03) so that its contribution to the linear waves (3.64) can be neglected.
As we will see in the next Section, its presence can be crucial in the derivation of the
model equations governing the dynamics of nonlinear modulated waves propagating
in the dissipative nonlinear electrical network under consideration.

3.2.3 Amplitude Equation and Modulated Damped

In this subsection, we first derive the amplitude equation that governs the dynamics
of spatiotemporal modulated damped waves propagating in the network system of
Fig. 3.8. Then we investigate the linear stability of plane wave with time-varying
amplitude propagating in our network system.
3.2 Spatiotemporal Modulation of Damped Solitonlike Wave … 73

3.2.3.1 Mathematical Considerations

Now, we focus our attention to modulated damped waves with a slowly varying
envelope in time and space with regard to a given carrier wave with angular fre-
quency ω = ωp = 2π fp and wavenumber k = kp . To derive the amplitude equation,
that is, the nonlinear partial differential equation that governs spatiotemporal mod-
ulated damped waves in the network of Fig. 3.8, we use the results of the preceding
subsection that lead to the following statements:

1. Following Marquié et al. [30], we introduce, through a perturbative small param-


eter ε that measures the smallness of the modulation frequency and the amplitude
of the input waves, two envelope
 variables
 x and τ beside the space variable n and
the time t, namely x = ε n −  υg t , and τ = ε2 t. These two envelope variables
will help us to apply the reductive perturbation method in the semi-discrete limit
[35].
2. In order to take into account the dissipation effects of our network system dur-
ing applying the reductive perturbation method, we will use, instead of a real
wavenumber k = kp , the complex wavenumber  k = kp + iχ , χ being the spatial
dissipation parameter defined in Eq. (3.67). Therefore 
k and ω satisfy the complex
linear dispersion relation
  
2u02 − ω2 − 2iu0 (σ1 + 2σ2 ) ω − 2 u02 − 2iu0 σ2 ω cosh i
k = 0. (3.70)

The following complex group velocity is derived from Eq. (3.70)


  
dω 2u0 σ2 ω + iu02 sinh ik

υg ≡ υgr + iυgi = =−  , (3.71)

dk ω + iu0 (σ1 + 2σ2 ) − 2iu0 σ2 cosh i
k

where
u0 Dr
υgr =  2  2 ,
ω − 2u0 σ2 sinh [χ ] sin kp + u02 σ1 + 2σ2 − 2σ2 cosh [χ ] cos kp
(3.72a)
u0 Di
υgi =   2   2 ,
ω − 2u0 σ2 sinh [χ ] sin kp + u02 2σ2 cosh [χ ] cos kp − σ1 − 2σ2
(3.72b)

with
74 3 Transmission of Dissipative Solitonlike Signals …
  
Dr = ω − 2u0 σ2 sinh [χ ] sin kp
   
× 2σ2 ω sinh [χ ] cos kp + u0 cosh [χ ] sin kp
   
+ u0 2σ2 ω cosh [χ ] sin kp − u0 sinh [χ ] cos kp
  
× 2σ2 cosh [χ ] cos kp − σ1 − 2σ2 , (3.73a)
   
Di = 2σ2 ω cosh [χ ] sin kp − u0 sinh [χ ] cos kp
  
× 2u0 σ2 sinh [χ ] sin kp − ω
   
+ u0 2σ2 ω sinh [χ ] cos kp + u0 cosh [χ ] sin kp
  
× 2σ2 cosh [χ ] cos kp − σ1 − 2σ2 . (3.73b)

3. The solution of the difference-differential system (3.62) is assumed to have the


following general form

Vn (t) = εψ(x, τ ) exp [iθ ] + ε2 ψ1 (x, τ ) + ψ2 (x, τ ) exp [2iθ ] + c.c., (3.74)

where θ =  kn − ωt is the rapidly varying phase and ψ1 and ψ2 are respectively


the dc term and the second-harmonic term added to the fundamental term ψ(x, τ )
in order to take into account the asymmetry of the charge-voltage relation given
by Eq. (3.60).

3.2.3.2 The Amplitude Equation [36] for the Dynamics of Modulated


Waves

Under assumptions (a)–(c), we present here the nonlinear partial differential equation
that governs the dynamics of modulated waves propagating in the network of Fig. 3.8.
Substituting Eq. (3.74) into Eq. (3.62) and taking into account conditions (a), (b),
and (c), we obtain the dc term and the second-harmonic terms ψ1 and ψ2 as,

ψ1 (x, τ ) = 2α |ψ|2 ,

 
 
4αω ω + iu0 (σ1 + 2σ2 ) − iu0 σ2 e−2k + e2k
ψ2 (x, τ ) =   
 ψ 2,
4ω − 2u0 [u0 − 2iω (σ1 + 2σ2 )] + u0 (u0 − 4iσ2 ω) e
2 −2 k +e 2 k

and the following amplitude equation

∂ψ ∂ 2ψ
i + (Pr + iPi ) 2 + (Qr + iQi ) |ψ|2 ψ = 0. (3.75)
∂τ ∂x
 
In Eq. (3.75), Pr = Re[P], Pi = Im[P] , Qr = Re Q , and Qi = Im Q , P and Q
being given by in terms of the complex wavenumber 
k and angular frequency ω as
3.2 Spatiotemporal Modulation of Damped Solitonlike Wave … 75


υg2
P=   , (3.76a)
2 −ω − iu0 (σ1 + 2σ2 ) + 2iu0 σ2 cosh  k
 
2α 2 ω ω + 2iu0 (σ1 + 2σ2 ) − 4iu0 σ2 cosh  k
Q= 
−ω − iu0 (σ1 + 2σ2 ) + 2iu0 σ2 cosh k 
 
3βω 1 + 2u0 (σ1 + 2σ2 ) − 4u0 σ2 cosh  k
−  
2 −ω − iu0 (σ1 + 2σ2 ) + 2iu0 σ2 cosh k 
 
4α 2 ω2 ω + 2iu0 (σ1 + 2σ2 ) − 4iu0 σ2 cosh  k
+ 
−ω − iu0 (σ1 + 2σ2 ) + 2iu0 σ2 cosh k

ω + iu0 (σ1 + 2σ2 ) − 2iu0 σ2 cosh 2 k
×  . (3.76b)
4ω − 2u0 [u0 − 2iω (σ1 + 2σ2 )] + 2u0 (u0 − 4iσ2 ω) cosh 2
2 k

Here,  υg is the complex group velocity given by Eq. (3.71) and  k = kp + iχ , χ


being the dissipation coefficient given by Eq. (3.67). The nonlinear partial differential
equation (3.75) is a complex cubic Ginzburg-Landau (GL) equation that governs the
propagation of modulated damped waves in the network of Fig. 3.8. In Eq. (3.75),
the complex parameters P and Q are respectively the dispersion and the nonlinear
coefficients of the amplitude equation. It is seen from expressions (3.76a) and (3.76b)
that Pr , Pi , Qr , and Qi depend on the dissipation coefficient χ and can take any sign.
When Pi and Qi are respectively negative and positive, they designate the dispersion
dissipation coefficient and the nonlinear dissipation coefficient, respectively. In the
regime when Pi is positive and Qi is negative, the resistance R1 and R2 act as if
the driven field Pi and Qi are taken separately. It should be noted that in certain
frequency range, it can occur the situation when Pi Qi > 0. In this situation, the
effect of resistances R1 and R2 in the dynamics of modulated waves depends on the
balance between the two coefficients [32].

3.2.4 Linear Stability

We now turn to the analysis of the uniform wave train propagating along the lossy
nonlinear electrical transmission network under small perturbations. We start by
seeking exact plane wave solutions of the GL equation (3.75) of the form

ψ(x, τ ) = a0 (τ ) exp [i0 x + iϕ0 (τ )] , (3.77)

where 0 is a real number and a0 (τ ) and ϕ0 (τ ) are two real functions of τ . For
equation (3.77) to satisfy the complex GL equation (3.75), a0 (τ ) and ϕ0 (τ ) must
satisfy the equation
da0 d ϕ0
i − a0 − 20 Pa0 + Qa03 = 0,
dτ dτ
76 3 Transmission of Dissipative Solitonlike Signals …

which leads to the following differential system of ordinary differential equations


for a0 (τ ) and ϕ0 (τ ) :

da0 d ϕ0
= Pi 20 a0 − Qi a03 , = Qr a02 − Pr 20 . (3.78)
dτ dτ
Integrating the first equation of system (3.78) yields

a02 (0)
a02 (τ ) = , ifPi = 0, (3.79a)
1 + 2Qi a02 (0)τ
a0 (0)
a0 (τ ) = )   , ifPi = 0. (3.79b)
Qi a0 (0)
2
Qi a02 (0) 
P 2
+ 1 − P 2
exp −2P 2
i 0 τ
i 0 i 0

Another important special solution of system (3.78) under the condition Pi Qi > 0 is

Pi Qr − Pr Qi 2 Qi 2
a0 (τ ) = a00 , ϕ0 (τ ) = a00 τ, with 20 = a , (3.79c)
Pi Pi 00

where a00 = 0 is any real constant. Solution (3.79c) of system (3.78) corresponds to
a plane wave solution (3.77) with a constant real amplitude.
To investigate the linear stability of the plane wave solution (3.77), we consider
the ansatz of the form

ψ(x, τ ) = [a0 (τ ) + δψ] exp [i0 x + iϕ0 (τ )] , (3.80)

where δψ is a small perturbation on the wave amplitude a0 . Substituting Eq. (3.80)


into the GL equation (3.75) and linearizing the resulting equation with respect in δψ
yield the following equation that describes the dynamics of the perturbation

∂δψ ∂ 2 δψ ∂δψ  
i +P + 2iP0 + Qr a02 + 2iQi a02 − iPi 20 δψ + Qa02 δψ ∗ = 0.
∂τ ∂x 2 ∂x
(3.81)
Let us denote respectively by K and  the real wavenumber and the complex fre-
quency of the modulation waves and seek nonzero solutions of Eq. (3.81) in the form
⎡ ⎤ ⎡ ⎤
(τ (τ
δψ = b1 exp ⎣iKx − i (ξ )d ξ ⎦ + b∗2 exp ⎣−iKx + i ∗ (ξ )d ξ ⎦ , (3.82)
0 0

where b1 and b2 are two complex constants satisfying the condition |b1 | + |b2 | > 0.
Inserting Eq. (3.82) into Eq. (3.81) and asking that |b1 | + |b2 | > 0, we obtain the
linear modulation dispersion relation
3.2 Spatiotemporal Modulation of Damped Solitonlike Wave … 77

    2  
 − 2Pr 0 K + i 2Qi a02 − Pi 20 + K 2 + 4iPi 0 K Qr a02 − Pr K 2
 
−Pr2 K 4 + 2 Pr Qr a02 + 2Pi2 20 K 2 + Qi2 a04 = 0, (3.83)

leading to
  
(τ ) = ± (τ ) = 2Pr 0 K ± ξ + i Pi 20 + K 2 − 2Qi a02 ± ζ ; (3.84a)

here,
 √  √
X+ X2 + Y2 −X + X 2 + Y 2
ξ = ,ζ = , (3.84b)
2 2
   
X = Pr2 K 4 − 2 Pr Qr a02 + 2Pi2 20 K 2 − Qi2 a04 , Y = 4KPi 0 K 2 Pr − a02 Qr .
(3.84c)

For the phenomenon of the modulational instability to occur, at least one of the fre-
quencies of modulation ± must possess a positive imaginary part [37–39], meaning
that   
Im [] = Im ± = Pi 20 − 2Qi a02 + Pi K 2 ± ζ > 0, (3.85)

 of −2 and + . When K → +0,2 the quantities Pi 20 − 2Q 2


for at2 least one  i a0 and
Pi 0 − 2Qi a0 + Pi K have the same sign. If Pi 0 − 2Qi a0 > 0, then Im + will
2 2

be positive as K → +0, and the corresponding perturbation (3.82) will increase expo-
nentially when τ → +∞, meaning that the system remains  unstable under modu-
lation. In the case where Pi 20 − 2Qi a02 < 0, we have Im − < 0 as K → +0, and
the modulational instability will occur only when
  
Im + = Pi 20 − 2Qi a02 + Pi K 2 + ζ > 0.

By expanding Im + as
 √
 
   1 4Pi2 20 K 2 + Qi2 a04 + X 2 + Y 2
Im + = Pi 20 − 2Qi a02 + Pi K 2 + Pr Qr a02 − Pr2 K 2 K 2 +
2 2
)
 2  1
> Pi 0 − 2Qi a0 + Pi K + K Pr Qr a02 − Pr2 K 2 ,
2 2
2

we can see that the condition Im + > 0 is satisfied as soon as

1
Pr Qr a02 − Pr2 K 2 > 0 and
) 2
 2  1
Pi 0 − 2Qi a0 + Pi K + K Pr Qr a02 − Pr2 K 2 > 0.
2 2
2

These last inequalities lead to


78 3 Transmission of Dissipative Solitonlike Signals …

Fig. 3.11 Plots of a: Real part of the dispersion coefficient Pr as a function of the angular frequency
ω ; b Imaginary part of the dispersion coefficient Pi as a function of the angular frequency ω; c Real
part of the nonlinear coefficient Qr as a function of the angular frequency ω ; d Imaginary part of
the nonlinear coefficient Qr as a function of the angular frequency ω; e Product Pr Qr as a function
of the angular frequency ω; f Product Pi Qi as a function of the angular frequency ω. Plots (b1 ) and
(f1 ) show the behavior of respectively Pi and Pi Qi far from ω = 0, while plots (e1 ) and (f2 ) show
how behave the products Pr Qr and Pi Qi near ω = 0, respectively. Plot (f3 ) shows the behavior of
the imaginary parts of the dispersion coefficient Pi and nonlinear coefficient Qi far from ω = 0 and
ω = ωc . Different plots are generated with the network parameters (3.61) with the dimensionless
resistances σ1 = 25 × 10−4 and σ2 = 5 × 10−3 associated with R1 = 5 and R2 = 10 (value
close to the numerical one [24]). Reprint from Ref. [27], Copyright 2022, with permission from
Springer

 2
2Qi a02 − Pi 20 − Pi K 2 Pr2 K 2
Pr Qr > + > 0.
a02 K 2 2a02

Therefore, for the perturbation (3.82) to increase exponentially as τ → +∞, it nec-


essary that Pr Qr > 0 and sufficient that 0 ≤ K 2 < 2 QPrr a02 . The condition of the mod-
ulational instability of the system under consideration is thus

Pr Qr > 0; (3.86)

the corresponding local growth rate (gain) of MI reads


 √
   −X + X2 + Y2
Im [] = Im + = Pi 20 + K 2 − 2Qi a02 + , 0 ≤ K < Kmax
2

Qr
= |a0 | 2 , (3.87)
Pr

where X and Y are given by Eq. (3.84c).


Relation (3.86) is the modulational instability criterion for the plane wave with
time-varying amplitude in the dissipative nonlinear transmission network of Fig. 3.8.
3.2 Spatiotemporal Modulation of Damped Solitonlike Wave … 79

Using Eq. (3.75) with a nonzero right-hand side as the basic model, Lange and
Newell [40] studied the modulational instability of harmonic waves by means of
the cumulative momentum method. They showed that plane waves with constant
amplitude are unstable (stable) under modulation for

Pr Qr + Pi Qi > 0 (Pr Qr + Pi Qi < 0);

in the physics literature, the relation

Pr Qr + Pi Qi > 0

is known as the Lange and Newell’s criterion of the MI [40]. When examining the
MI of slowly modulated waves in physical systems described by a higher-order
GL equation, Pelap and Faye [41] generalized the Lange and Newell’s criterion
with a corrective term which depends on the characteristic parameters of the carrier
wave. The condition (3.86) of the MI of Stokes waves with time-varying amplitude
thus differs from the Lange and Newell’s criterion for plane waves with constant
amplitude. It is important to note that MI criterion (3.86) is in agreement with the
Benjamin–Feir criterion for Stokes waves of NLS equation [42–45].
With the use of the network parameters given in Eq. (3.61), we display in Fig. 3.11
the behavior of coefficients P and Q through their real and imaginary parts, as well as
products Pr Qr and Pi Qi as functions of angular frequency ω. For generating different
plots, we use the dimensionless resistances σ1 = 25 × 10−4 and σ2 = 5 × 10−3 that
correspond respectively to R1 = 5 and R2 = 10 (values close to the numerical
one [24]). As it is seen from Fig. 3.12a, b and d, Pr , Pi , and Qi change their sign
in the whole frequency range of the allowed band; as we can see from Fig. 3.11c,
the real part Qr of the nonlinear coefficient Q is always negative for the network
parameter (3.61). It is also seen from Fig. 3.11f that product Pi Qi changes its sign
when ω varying in the segment [0, ωc ]. Plots (b), (d), and (f3 ) reveal that different
scenarios described after Eqs. (3.76a) and (3.76b) occur for the network parameters
(3.61) and the dimensionless parameters σ1 = 25 × 10−4 and σ2 = 5 × 10−3 . Espe-
cially, in some given range of angular frequency, Pi is negative while Qi is positive,
as it is clearly seen from Fig. 3.11 (f3 ); this means that in such a region of vari-
ation of the angular frequency ω, Pi and Qi designate respectively the dispersion
dissipation coefficient and the nonlinear dissipation coefficient. Also, it is seen from
Fig. 3.11 (f3 ) that the situation where either Pi > 0 or Qi > 0, the resistances act
as if the driven field Pi and Qi are taken separately. The situations where Pi < 0
and Qi < 0 and Pi > 0 and Qi > 0 can be also observed in Fig. 3.11 (f3 ); in the
corresponding frequency range, the effect of resistances in the dynamics of modu-
lated waves depends on the balance between the two coefficients. Lastly, Fig. 3.11e
reveals that equation Pr Qr = 0 admits only one solution, ωnull = 156106.5, which
corresponds to the propagating frequency fnull = 24. 85 kHz. According to condition
(3.86) of the modulational instability, Fig. 3.11e establishes the existence of two
regions concerning the MI of plane wave with time-varying amplitude and possi-
ble soliton solutions of the complex Ginzburg-Landau equation (3.75): (i) Region
80 3 Transmission of Dissipative Solitonlike Signals …

associating with fp ∈ [0, fnull [ for which Pr > 0 and Qr < 0, leading to Pr Qr < 0
and corresponding the modulational stability and hole soliton, and (ii) region for
fp ∈ [fnull , fc [ for which Pr < 0 and Qr < 0, leading to Pr Qr > 0 and corresponding
to the modulational instability and envelope soliton.
Because Pr , Pi , Qr , and Qi depend on the dissipative elements R1 and R2 , the
instability growth rate (3.87) and the domain of the variability of the wavenumber
K of the modulation depend on R1 and R2 . To show how R1 and R2 affect the MI
growth rate , we have depicted in Fig. 3.12 the MI growth rate according to Eq. (3.87)
for the carrier wave with parameters (3.79c) for a00 = 1 and for different values of
resistances R1 and R2 . Figure 3.12a shows the MI gain when only R2 varies, while
Fig. 3.12(b) shows the MI growth rate when only R1 varies. It is seen from these
plots that the dissipative elements R1 and R2 enhance the instability region. For the
wavenumber of modulation K near zero (small K = 0), the MI growth rate decreases
[increases] when R2 increases [R1 increases], while for wavenumber of modulation K
near Kmax , the MI growth rate increases [decreases] when R2 increases [R1 increases];
this situation is well seen in Fig. 3.12c. Also, we can see from Fig. 3.12 that Kmax
increases with R2 as we can seen from plots (a), while increasing R1 decreases
the value of Kmax , as we can well observe from plots (b). By comparing plots of
Fig. 3.12a obtained with a fixed value of R1 and those of Fig. 3.12b obtained with
a fixed value of R2 , it appears that the values of the growth rate of the MI showed
in Fig. 3.13a is higher that those showed in Fig. 3.12b. This means that the must
important contribution to the MI gain comes from the resistance R2 in the shunt
branch of the network. In conclusion, the effects of dissipative losses in shunt branch
are more manifested than those resulting from the dissipative losses in the series
branch.
Remark It is important to notice that in the case of damping modulated plane waves,
the amplitude of the propagating wave may decrease rapidly to zero and the waves
vanish after few cells. In such situations, the phenomena of the MI would not be
experimentally observed if the processing of instability takes place after many cells.
Consequently, it should not be experimentally possible to predict the domain in which
the electrical network may support the propagation of envelope solitons.

3.2.5 Spatiotemporal Modulated Signals Propagating


Through the Network of Fig. 3.8

We start this subsection by presenting exact and approximative solitonlike solutions


of the complex GL equation (3.75). Then, these exact and approximate solutions are
used to investigate analytically the transmission of spatiotemporal damped envelope
soliton signals though the network system of Fig. 3.8. Without loss of generality, we
limit our study to the exact and approximative bright solitonlike solutions.
When Pi = Qi = 0, Eq. (3.75) coincides with the standard nonlinear Schrödinger
equation and admits, under the condition Pr Qr > 0, the bright soliton solutions of
the form
3.2 Spatiotemporal Modulation of Damped Solitonlike Wave … 81

Fig. 3.12 Modulational instability growth rate Im[] according to Eq. ( 3.87) for the carrier wave
with parameters (3.79c) for different values of resistances R1 and R2 . a MI gain for R1 = 5 and
three values of resistance R2 , R2 = 5 (solid line), R2 = 15 (dash-dotted line), and R2 = 23
(dash line); b Mi growth rate for R2 = 5 and three values of R1 , R1 = 5 (solid line), R1 = 15
(dash-dotted line), and R1 = 23 (dash line); c MI gain for R1 = 5 and R2 = 23 (solid line)
and for R1 = 23 and R2 = 5 (dashing line line). In different plots (a)–(c), the dotted lines
show the variation of the MI growth rate in the absence of the dissipative elements, that is, when
R1 = R2 = 0. Different plots are generated with the network parameters (3.61) and are associated
with the linear wave propagating at frequency fp = 19.10 MHz for which the conditions Pi Qi > 0
and Pi 20 − 2Qi a02 < 0 are satisfied. Reprint from Ref. [27], Copyright 2022, with permission from
Springer

!  "
As 2Pr 20 − Qr As
ψ(x, τ ) =
exp i 0 x − τ , (3.88)
Qr
cosh 2P A (x − υ τ ) 2
r
s 0

where As is the wave amplitude and υ0 = 2Pr 0 is the wave speed. In the situation
when |Pi | + |Qi | > 0, Eq. (3.75) is not integrable and does not admit envelope soliton
solutions of form (3.88). In this situation, we can build either exact or approximative
82 3 Transmission of Dissipative Solitonlike Signals …

Fig. 3.13 a Plot of the dimensionless resistance σ2 versus the angular frequency ω obtained as an
implicit function defined form equation Pr Qi − Pi Qr = 0 for the network parameters (3.61) with
σ1 = 25 × 10−4 . b Three-dimensional plot of the product Pi Qi . Reprint from Ref. [27], Copyright
2022, with permission from Springer

bright solitonlike solutions with time-varying parameters when Pr Qr > 0. In what


follows, we focus ourselves to the exact and approximative envelope solitonlike
solutions of Eq. (3.75) under the condition that coefficients P and Q of the GL Eq.
(3.75) satisfy the condition (3.86) of the MI and |Pi | + |Qi | > 0.

3.2.5.1 Bright Solitonlike Solution with Time-Varying Parameters

To find analytical bright solitonlike solutions with time-varying parameters of the


GL equation ( 3.75), we introduce the ansatz

2 exp [i {2 x + h4 (τ )}]


ψ(x, τ ) = , (3.89a)
coshN [1 x + h3 (τ )]

where 1 > 0 and 2 are two free real parameters called solution parameters, h3 (τ )
and h4 (τ ) are two real functions of variable τ , and coshN [X ] = N exp[X ]+exp[−X
2
]
,N
being any positive parameter. Imposing to (3.89a) to satisfy the complex cubic GL
equation (3.75) yields
3.2 Spatiotemporal Modulation of Damped Solitonlike Wave … 83
  
h3 (τ ) = Pi 22 − 21 − 2Pr 1 2 τ + 3 , (3.89b)
  
h4 (τ ) = Pr 21 − 22 − 2Pi 1 2 τ + 4 , (3.89c)
1 Qr
N = 2 , Pr Qi − Pi Qr = 0, (3.89d)
21 Pr

where 3 and 4 are real parameters. The solution parameters 1 > 0 and 2 , as we
will see in the following, play an important role in the management of spatiotemporal
envelope signals propagating through the network of Fig. 3.8. The positivity of N in
Eq. (3.89b) follows from that of product Pr Qr . It is important to notice that the exact
solitonlike solution (3.89a) with parameters (3.89b)–(3.89d) is valid only when P
and Q satisfy the additional condition Pr Qi − Pi Qr = 0; for Qi = 0, condition (3.86)
of the MI under the condition Pr Qi − Pi Qr = 0 becomes Pi Qi > 0.
In the practice, condition Pr Qi − Pi Qr = 0 can be used as follows: For the given
network parameters (3.61) and, let us say, for a given dimensionless resistance σ1 , one
can use equation Pr Qi − Pi Qr = 0 to express σ2 in terms of the angular frequency
ω (here, taking into account the fact the most contribution to the spatial decreasing
rate comes from R2 , we have preferred to solve Pr Qi − Pi Qr = 0 in σ2 ); in other
words, for a given propagating frequency fp = ω/2π , one can find the appropriate
dimensionless resistance σ2 = σ2 (ω) that leads to Pr Qi − Pi Qr = 0. If condition
Pi Qi > 0 is satisfied for such σ2 (ω), then Eq. (3.89a) with parameters (3.89b)–(3.89d)
will give an envelope solitonlike solution of the complex GL equation (3.75) for the
given network parameters (3.61) with dimensionless resistances σ1 and σ2 = σ2 (ω).
For a better understanding, we show in Fig. 3.13a the plot of the dimensionless
resistance σ2 versus the angular frequency ω as an implicit function derived from the
equation Pr Qi − Pi Qr = 0 for the network parameters (3.61) and the dimensionless
resistance σ1 = 25 × 10−4 (this value of σ1 corresponds to the numerical value R1 =
5  [24]). As we can see from Fig. 3.13b, product Pi Qi is positive for σ2 around the
numerical value σ2 = 5 × 10−3 . It is seen from Fig. 3.13a that the numerical value
σ2 = 5 × 10−3 corresponds to the propagating frequency fp = 10.135 MHz.
Now, if we insert Eq. (3.89a) in the decomposition (3.74) and ignore the dc term
ψ1 and the second harmonics term ψ2 , we will obtain that the shape of the network
soliton signal can be well approximated by the shape of the first harmonic whose
expression is given as

cos [z2 ] cos ε1 υgi t coshN [z1 ]

 − sin ε1 υgi t sin [z2 ] sinhN [z1 ]
Vn (t) = 4ε exp 2 ευgi t − χ n  ; (3.90a)
sinh2N [z1 ] + N cos2 ε1 υgi t

here, z1 , z2 , and coshN z1,2 as defined as
84 3 Transmission of Dissipative Solitonlike Signals …

Fig. 3.14 Evolution plots of the real part υgr (a) and the imaginary part υgi (b) of the complex group
velocity υg as functions of the angular frequency ω defined by Eqs. (3.72a) and ( 3.72b), respectively.
Plot (a) shows that the group velocity υgr is nonnegative, while plot (b) shows that υgi changes its
sign in the whole range of the propagating frequency. Different plots are generated with the network
parameters (3.61) with the dimensionless parameters σ1 = 25 × 10−4 and σ2 = 5 × 10−3 . Reprint
from Ref. [27], Copyright 2022, with permission from Springer


   
z1 = z1 (n, t) = 1 εn + ε ε Pi 22 − 21 − 2Pr 1 2 − 1 υgr t + 3 ,
 
   
z2 = z2 (n, t) = kp + 2 ε n + ε ε Pr 21 − 22 − 2Pi 1 2 − 2 υgr − ω t + 4 ,
(3.91)
N exp [z1 ] + exp [−z1 ] N exp [z1 ] − exp [−z1 ]
coshN [z1 ] = , sinhN [z1 ] = .
2 2

In general, the real part υgr of group velocity υg is a nonnegative quantity, while
its imaginary part υgi can take any sign, negative, null, or positive. For a better
understanding, we depict in Fig. 3.14 the variation of υgr and υgi versus the angular
frequency ω for the network parameters (3.61); here, we use the dimensionless resis-
tances σ1 = 25 × 10−4 and σ2 = 5 × 10−3 corresponding respectively to R1 = 5 
and R2 = 10  (values close to the numerical one [24]). As we can see from dif-
ferent plots of Fig. 3.14, υgr is nonnegative (Fig. 3.14a) and υgi changes its sign
(Fig. 3.14b).
 From Eq. (3.90a),
we can see that the wave amplitude is proportional to exp
ε2 υgi t − χ n so that if the free parameter 2 is taken from the condition 2 υgi < 0,
the envelope soliton signal associated with Eq. (3.90a) will be damped in both time
and space. That is, the signal amplitude will decrease in both the time t and the cell
number n (since χ is a positive quantity). The expression of z1 (n, t) given in Eq.
(3.91) reveals that the soliton width is ωdth = 1/ε1 , while its speed is defined as
  
υsol = υgr + ε Pi 1 − −1
1 2 + 2Pr 2
2

and coincides with the group velocity if the free parameters 1 and 2 are chosen
from the condition (Pr ± |P|) 1 − Pi 2 = 0. Therefore, the situation when 1 and 2
satisfy the relationship (Pr ± |P|) 1 − Pi 2 = 0 corresponds to the case where the
electrical network is excited by a soliton signal with initial velocity corresponding to
3.2 Spatiotemporal Modulation of Damped Solitonlike Wave … 85

the group velocity υgr of the wave packet. The centre of the envelope soliton in such
a situation will move with the constant speed υsol = υgr , while its amplitude will
vary proportionally to exp 2 ευgi t − χ n . If 1 and 2 are taken from the condition
(Pr ± |P|) 1 − Pi 2 = 0, we obtain a situation in which our network system is
excited by a bright soliton with a speed υsol different from the group  velocity υgr
of the wave packet, and its amplitude varies proportionally to exp 2 ευgi t − χ n .
It should be to noticed that in the case when N = 1, both 1 and 2 can be used to
control the wave amplitude in the propagating regime where υgi = 0. It follows from
the expression ωdth = 1/ε1 of the wave width, that parameter 1 can also be used to
manage the soliton width. Hence, the two solution parameters 1 and 2 can be used
to manage the spatiotemporal bright soliton signals propagating through the network
of Fig. 3.8.

3.2.5.2 Approximative Bright Solitonlike Solutions of the GL Equation


(3.75)

In the following, we assume that P and Q satisfy the condition of the MI (3.86) and
intend to find the approximative bright solitonlike solution of Eq. (3.75). For this
aim, we consider the ansatz

ψ(x, τ ) = (τ )ρ(ξ ) exp [i (λx + (τ ))] , ξ = Ax + B(τ ), A = 0, (3.92)

where A and λ are two real constants, and (τ ), ρ(ξ ), (τ ), and B(τ ) are real
functions to be determined. Inserting Eq. (3.92) into Eq. (3.75) and setting the real
and imaginary parts of the resulting equation equal to zero yield the differential
system of two equations in ρ, , B, , and  :
 
d 2ρ dρ d
Pr A2− 2P i λA − Pr λ 2
+ ρ + Qr 2 ρ 3 = 0,
dξ2 dξ dτ
(3.93a)
   
d ρ
2
dB dρ d
Pi A2 2 +  + 2Pr λA + − Pi λ2  ρ + Qi 3 ρ 3 = 0.
dξ dτ dξ dτ
(3.93b)

To find the approximative bright solitonlike solution of Eq. (3.75), we ask that A
and B satisfy the equation
dB
+ 2Pr λA = 0. (3.94a)

For simplicity, we assume that ρ(ξ ) and its various order derivatives vanish at |ξ | =
+∞, and limit ourselves to the localized solutions of the field equation.
If we multiply both sides of Eq. (3.93a) by ρd ξ and integrate the resulting equation
with respect to ξ from ξ = −∞ to ξ = +∞, we arrive to
86 3 Transmission of Dissipative Solitonlike Signals …

(+∞   (+∞ (+∞ 2


d dρ
Qr 
2
ρ d ξ − Pr λ +
4 2
ρ d ξ − Pr A
2 2
d ξ = 0. (3.94b)
dτ dξ
−∞ −∞ −∞


Multiplying now the both sides of Eq. (3.93b) by dξ
and integrating the resulting
equation under the condition (3.94a) yield
 2  
dρ d 1
Pi A2 + − Pi λ2  ρ 2 + Qi 3 ρ 4 = 0. (3.94c)
dξ dτ 2

Let μ0 = μ0 (τ ) and ρ0 = ρ0 (τ ) be two real functions of τ that satisfy the relationship

1 Qi 2 2
μ20 = − ρ (implying that Pi Qi < 0). (3.95)
2 Pi A2 0

Asking that  satisfies the equation

d 1
− Pi λ2  + Qi ρ02 3 = 0, (3.96)
dτ 2
it is easily seen that
ρ0 (τ )
ρ (ξ ) = (3.97)
cosh [μ0 (τ )ξ ]

is a special solution of Eq. (3.94c). Using now Eqs. (3.97) and (3.94b), we obtain
that
d 4Pi Qr + Pr Qi 2 2
− ρ0  + Pr λ2 = 0. (3.98)
dτ 6Pi

Gathering Eqs. (3.94a), (3.96), and (3.98) yields the following ordinary differen-
tial system for the determination of the solution parameters B(τ ), (ξ ), and (τ ),
assuming that λ and A are two arbitrary real constants

dB
+ 2Pr λA = 0, (3.99a)

d 1
− Pi λ2  + Qi ρ02 3 = 0, (3.99b)
dτ 2
d 4Pi Qr + Pr Qi 2 2
− ρ0  + Pr λ2 = 0. (3.99c)
dτ 6Pi

We can easily see that solutions of system (3.99a)–(3.99c) depend on the choice of
the functional parameter ρ0 (τ ) from Eq. (3.95). In the special case when ρ02 = 2, Eq.
(3.99b) coincides with the first equation in Eq. (3.78) when λ2 is replaced by 20 and
 is replaced by a0 . In this special case, a special solution of system (3.99a)–(3.99c)
is given as
3.2 Spatiotemporal Modulation of Damped Solitonlike Wave … 87

B(τ ) = A (−2Pr λτ + B0 ) ,
0
(τ ) =    , (3.100)
2
Qi 0 Q 2
i 0 
P 2 + 1− exp −2Pi λ τ
2
i λ 2Pi λ
⎡ ⎤
2λ2 (2Pi Qr − Qi Pr ) Pr Qi + 4Pi Qr ⎣ Pi λ2 20
(τ ) = 0 + τ+ ln   ⎦, if λ = 0,
3Qi 6Pi Qi Pi λ2 − Qi 20 2 (τ )

B = AB0 ,
20
2 = , (3.101)
220 Qi τ + 1
(4Pi Qr + Pr Qi ) 
= ln 1 + 220 Qi τ + 0 , if λ = 0.
6Pi Qi

Here, B0 = B(0)/A, 0 = (0) and 0 = (0) are three real constants of integra-
tion. To make (τ ) and (τ ) defined as τ → +∞, Pi must be negative (this leads,
from condition Pi Qi < 0, to Qi > 0). Inserting Eqs. (3.97) and (3.100) into Eq. (3.92),
we arrive to the following approximative bright solitonlike solution of the complex
GL Eq. (3.75)

as (τ )
ψ(x, τ ) =
ϕ (x, τ )] ,
exp [i
Qi
cosh − 2P as (τ ) (x − 2Pr λτ + B0 )
i

20
as (τ ) =    , (3.102)
Qi 0
2 2
Qi 0 
Pi λ 2 + 1−
Pi λ 2
exp −2Pi λ2 τ
⎡ ⎤
2λ2 (2Pi Qr − Qi Pr ) Pr Qi + 4Pi Qr ⎣ 2Pi λ2 20

ϕ (x, τ ) = λx + τ+ ln   ⎦ + 0 .
3Qi 6Pi Qi Pi λ2 − Qi 20 as2 (τ )

Here, λ = 0, 0 = 0 and 0 are three arbitrary real constants, as (τ ) and ϕ(x, τ )


are respectively the soliton amplitude and the soliton phase. Henceforth, the free
parameter λ will be referred to as “the solution parameter” for the approximative
bright solitonlike solution of the GL equation (3.75). Inserting now Eqs. (3.97) and
(3.101) into Eq. (3.92), we obtain, for λ = 0, the following approximative bright
solitonlike solution of Eq. (3.75)

  
√ exp i λx + 0 + 4Pi Q6Pr +P r Qi
ln 1 + 2Q 2
τ
20 Q i 0
ψ(x, τ ) = ! "
i i
,
2Qi 0 τ + 1
2
cosh √ 0 Qi
− Pi (x + B0 )
2
2Qi 0 τ +1
(3.103a)
0 = 0 and 0 being two arbitrary real constants.
88 3 Transmission of Dissipative Solitonlike Signals …

It is important to note that Eqs. (3.102) and (3.103a) are solutions in the averaged
sense. They can be obtained with the help of the adiabatic perturbation theory devel-
oped for bright solitons [46, 47]. It is seen from Eqs. (3.102) and (3.103a) that the
soliton amplitude decays in time τ as the wave progresses.
In the special case when λ = 0, solution (3.102) corresponds to the situation where
the electrical network is excited by a soliton signal with initial velocity corresponding
to the group velocity υgr of the wave packet. The centre of the envelope soliton in such
a situation moves with the speed dx/d τ = 0, and this means that the soliton speed
remains constant while its amplitude as (τ ) varies according to the second equation in
Eq. (3.102). Therefore, the network electrical bright soliton propagates in the lossy
network of Fig. 3.8 with a constant speed which coincides with the group velocity
of the carrier wave. In the situation when λ = 0, the initial amplitude as (τ ) = 0 and
speed υs = 2Pr λ = 0 of the bright soliton satisfy the relationship

± 20
as (τ ) = )  
,
4Pr2 Qi 20 4Pr2 Qi 20 Pi υs2
Pi υs2
+ 1− Pi υs2
exp − 2P2 τ
r

and solution (3.102) in this case λ = 0 corresponds to the situation when the network
under consideration is excited by an envelope solitons with a velocity υs different to
the group velocity of the wave packet.
To use the above approximate soliton solution for the transmission of bright soli-
tonlike signal through our network system, we proceed as in the case of the exact
soliton solution of the amplitude equation. We use the decomposition (3.74) of the
network voltage Vn (t) and the approximative envelope solitonlike solutions of Eq.
(3.75), and return back to dimensioned parameters variables. We then obtain, after
ignoring the dc and the second harmonics, that the shape of the network envelope
soliton signal can be well approximated by the shape of the first harmonic whose
expression is given by either

 cosh [z] cos [φ] cos [ϕ] − sinh [z] sin [φ] sin [ϕ]
Vn (t) = 2εas (t) exp ελυgi t − χ n ,
sinh2 [z] + cos2 [ϕ]
(3.104)
√ 
20 Qi
as (t) =    , ϕ = ϕ(t) = ευgi t − as (t), λ = 0,
2 2  2Pi
Qi 0 Qi 0
Pi λ 2 + 1− exp −2Pi λ2 ε 2 t
Pi λ 2
 ! "
Qi   B0
z = z(n, t) = ε − as (t) n − υgr + 2Pr λε t + , (3.105)
2Pi ε
% &
  2ε 2 λ2 (Qi Pr − 2Pi Qr )
φ = φ(n, t) = kp + λε n − ω + λευgr + t
3Qi
⎡ ⎤
Pr Qi + 4Pi Qr ⎣ 2Pi λ2 20
+0 + ln   ⎦.
6Pi Qi P λ2 − Q 2 a2 (t)
i i 0 s
3.2 Spatiotemporal Modulation of Damped Solitonlike Wave … 89

Fig. 3.15 Evolution of the damped envelope soliton signal in the network at frequency fp = 10.135
MHz for the network parameters (3.61) with R1 = 5 , R2 = 10 , and ε = 2 × 10−3 . Different
plots are generated with the help of the shape of the first harmonic (3.90a) for (a): 1 = 0.2, 2 = 3,
3 = 10, and 4 = 0; b 1 = 0.2, 2 = 0.1, 3 = −2.2, and 4 = 0. Plots of panels (a) show that the
soliton amplitude decreases as the cell number n increases, while those of column (b)  show that the
wave amplitude decays as time t increases. c, d show the wave amplitude factor exp 2 ευgi t − χn
as function of time t for n = 500 with the same data as in plots (a) and function of the cell number
n for t = 4 μs with the same data as in plots (b), respectively. Reprint from Ref. [27], Copyright
2022, with permission from Springer

for the approximative solution (3.102) with λ = 0, or

0 cosh [z] cos [φ] cos [ϕ] − sin [φ] sinh [z] sin [ϕ]
Vn (t) = 2ε exp [−χn] ,
ε2 Qi 20 t + 1 sinh2 [z] + cos2 [ϕ]
2
(3.106)
ε0   ευgi 0
z = z(t) = n − υgr t + ε−1 B0 , ϕ = ϕ(t) = t,
Pi Pi
−2ε2 Pi 20 t − Qi −Qi
− 2ε2 Pi 20 t
(3.107)
4Pi Qr + Pr Qi

φ = φ(n, t) = kp n − ωt + 0 + ln 1 + 2ε2 Qi 20 t , (3.108)
6Pi Qi

for the approximative solution (3.103a) with λ = 0.


90 3 Transmission of Dissipative Solitonlike Signals …

It follows from Eqs. (3.104) and (3.105) that: 


(i) The soliton amplitude is proportional to 2ε exp λευgi t − χ n , meaning that the
soliton amplitude decreases as the cell number n increases;
(ii) Because Pi < 0, the soliton amplitude as (t) decreases as the wave propagates
[lim as (t) = 0 as t → +∞];
(iii) The soliton speed υ = υgr + 2εPr λ is constant and depends on the free parameter
λ; (iv) The soliton width is

−1 2Pi 1
ωdth = ε −
Qi |as (t)|

and tends to +∞ as t → +∞;


(v) As we can see from the expression of φ [see Eq. (3.105 )], the angular frequency
of the dissipative bright soliton voltage (3.104 ) is formed of two terms, a linear term

2ε2 λ2 (Qi Pr − 2Pi Qr )


ωp = ω + λευgr + ,
3Qi

and a nonlinear term


Pr Qi + 4Pi Qr 1 das
ωn = ,
3Pi Qi as (t) dt

while the wavenumber is k = kp + λε and depends on the free parameter λ. The


above analysis can be done for the approximate shape of the network soliton (3.106)–
(3.108). As we will see in the following, the free parameter λ plays an important role
in the transmission of spatiotemporal damped envelope soliton signals through the
network of Fig. 3.8.

3.2.6 Transmission of Spatiotemporal Modulated Damped


Envelope Signals Through a Lossy Network

Our purpose now is to investigate analytically the transmission of spatiotemporal


modulated envelope signals through the RLC network of Fig. 3.8. For this aim, we
use the above shapes of the first harmonic (3.90a), (3.104), and (3.106) obtained with
the help of exact and approximate solutions of the amplitude equation (3.75).
Our first example is obtained with the use of the shape of the first harmonic (3.90a)
corresponding to the exact bright solitonlike solution (3.89a). Here, the network
parameters (3.61) with σ1 = 25 × 10−4 corresponding to R1 = 5  are used; using
a value of σ2 that corresponds to the numerical value R2 = 10  used in Ref. [24],
we find the propagating frequency to be fp = 10.135 MHz. Computing υgr and υgi ,
we find that υgr is positive, while υgi is negative (this can be seen from Fig. 3.15a, b
when ω ≈ 6.368 × 107 ). We then use these data to depict in Fig. 3.15 the evolution of
spatiotemporal damped bright soliton signals propagating at frequency fp = 10.135
3.2 Spatiotemporal Modulation of Damped Solitonlike Wave … 91

Fig. 3.16 Plot of the voltage Vn (t) as a function of the cell number n, at given times showing the
evolution of damped soliton signals in the network of Fig. 1 at frequency fp = 7.96 MHz. Plots of
the left and right columns show the soliton voltage signals obtained with the approximative shape
of the network soliton (3.104) for respectively λ = −5 and λ = 5, while the plots of the middle
column show the evolution of the soliton voltage signals obtained with the help of the approximative
shape of the network soliton (3.106) associated with λ = 0. Different plots are generated with the
network parameters (3.61) for ε = 4 × 10−3 and the dimensionless resistances σ1 = 25 × 10−5 and
σ2 = 5 × 10−3 . Other parameters appearing in solutions (3.104) and (3.106) are B0 = 0, 0 = 0,
and 0 = 70. It is important to notice that the propagating frequency fp = 7.96 MHz and the above
dimensionless resistances σ1 and σ2 lead to a positive υgi . Reprint from Ref. [27], Copyright 2022,
with permission from Springer

MHz through our network system. From plots of Fig. 3.15, we can see that the
soliton amplitude decays as either the cell number increases [column (a)] or the time
t increases [column (b)]. Also, plots of Fig. 3.15 reveal that the centre of the bright
soliton moves with both the cell number n [column (a)]  and time t [column
(b)].
Fig. 3.15c and 8d show how the exponential factor exp 2 ευgi t − χ n of the wave
amplitude maximum decays respectively versus time t and versus cell number n.
Our second example is generated with the shape of the first harmonic (3.104)
and (3.106) associated with the approximative envelope solitonlike solutions of the
complex GL equation (3.75). This example demonstrates how the solution parameter
λ affects the transmission of the spatiotemporal damped envelope soliton signals
propagating through the lossy network of Fig. 3.8. In order to examine the effects
parameter λ on the soliton signals, we have depicted in Fig. 3.16 the evolution plots
92 3 Transmission of Dissipative Solitonlike Signals …

of damped bright soliton signals for three values of λ. The data used in Fig. 3.16 lead
to a positive υgi so that λυgi > 0 for positive λ and λυgi < 0 for negative λ. From
plots of Fig. 3.16, we conclude that
1. in the situation where the electrical network is excited by a soliton with initial
velocity corresponding to the group velocity υg = υgr of the wave packet (that
is, λ = 0), the soliton amplitude is higher than in the general situation where
the electrical network is excited by the envelope solitons with initial velocity
different to the group velocity of the wave packet (that is, λ = 0);

2. the soliton amplitude for λ > 0 is higher than that for λ < 0 (this is well seen
when comparing plots of the left column with those of the right column, espe-
cially for large time);

3. the soliton speed decreases as λ increases; this behavior is a consequence of


the fact that the soliton speed, as we can see from Eqs. (3.105) and (3.107),
is υ = υgr + 2εPr λ and Pr < 0 for the data used in Fig. 3.16. We can then
conclude that parameter λ of the solution can be used to modulate the damped
soliton amplitude and velocity.
In conclusion, we have investigated in this Section the effects of losses of a RLC
nonlinear electrical transmission network on the slowly modulated wave dynamics.
We have derived the spatial decreasing rate χ in terms of the propagating frequency
and showed that the amplitude of linear waves is an exponentially decreasing function
of the network cell number n. Our studies reveal that the introduction of the losses in
the network increases the propagating frequency in the operating frequency range of
the network. We have showed that the dynamics of modulated waves in our network
model are governed by a complex cubic GL equation whose coefficients depend
on the spatial decreasing rate χ . We have investigated the linear stability of slowly
modulated plane wave with time-varying amplitude and established the criterion of
their MI. We found that for the modulated damped plane waves, the phenomena of
the MI would not be experimentally observed if the processing of instability takes
place after many cells.
Parametric analytical exact and approximate bright solitonlike solutions of the
amplitude equation (3.75) are presented and used to investigate the effects of losses
on the electrical envelope soliton propagation through the network. Our investigation
showed that the amplitude of the envelope soliton decreases in both time t and cell
number n, the width of the soliton depends on the spatial decreasing rate χ as the
soliton propagates along the system. More interestingly, we found that both the
soliton speed in the moving frame at the group velocity of carrier wave and the group
velocity of carrier wave remain constant during the soliton propagation. Comparing
the contribution of dissipative losses in the series branch to that in the shunt branch
of the network on the modulational instability, we found that in the domain where the
network may support the envelope soliton propagation, effects of dissipative losses
in shunt branch are more manifested than those resulting from the dissipative losses
3.3 Modulated Wavetrains in a Dissipative Bi-Inductance Transmission Network 93

in the series branch. Also, our study has showed that the solution parameters 1 and
2 or λ can be used to manage the soliton amplitude and speed.

3.3 Modulated Wavetrains in a Dissipative Bi-Inductance


Transmission Network

3.3.1 Introduction and Circuit Equations

We consider in this Section a lossy nonlinear bi-inductance transmission network


with dispersive linear elements. The envelope modulation is reduced to a general-
ized cubic-quintic Ginzburg-Landau (CQ-CGL) equation. Using this equation, we
analyze the MI phenomenon and derive a generalized Lange-Newell criterion [48].
The nonlinear coherent shapes are demonstrated by formulating the analytical solu-
tions for the envelope equation via appropriate methods. To make possible the study
of coherent structures, the found CQ-CGL equation is first reduced to a third-order
ordinary differential system.
The electrical network considered in this Section is the discrete nonlinear-
dispersive bi-inductance transmission network with losses shown in Fig. 3.17. Apply-
ing the Kirchhoff’s laws to this network, we arrive to the following set of network
equations

∂I1,n ∂I1,n+1
L2 = Vn−1 − Vn , L1 = Vn − Vn+1 ,
∂t ∂t
∂ ∂Qn 1
In − I1,n = CS (Vn−1 − Vn ) , + Vn = In − In+1 , (3.109)
∂t ∂t G

where Qn = Qn (Vn ) is the charge carried by the nonlinear capacitance C(Vn ), Vn =


Vn (t) is the voltage on it, I1,n is the current passing through the n-th linear inductor L2 .

Fig. 3.17 One section of the discrete nonlinear-dispersive dissipative bi-inductance transmission
network composed of N such identical sections. Reprint from Ref. [2], Copyright 2022, with
permission from Springer
94 3 Transmission of Dissipative Solitonlike Signals …

Eliminating the currents from system (3.109) and assuming a voltage-dependence


nonlinear capacitor of the form
 
V
C(V ) ≈ C0 1 − (3.110)
V0

yield the following difference-differential equations for the voltage (network equa-
tions)

d 2 Vn 2 d 2V 2
C0 2
−CS dtd 2 (Vn−1 − 2Vn + Vn+1 ) + G dVdt
n
− C0 b dt 2n
dt
= L12 (Vn−1 − Vn ) − L11 (Vn − Vn+1 ) , (3.111)

where n = 1, 2, . . . , N (N being the total number of cells in our network system),


C0 and V0 are the capacitance and voltage scales, and V  V .

3.3.2 Generalized Cubic-Quintic Complex Ginzburg-Landau


Equation for a Lossy Network

To analyze short-wavelength modulated nonlinear waves propagating in our network


system, a semi-discrete approximation is used [49, 50]. We introduce, for analyzing
the propagation of the group of waves centered around wavenumber k (0 ≤ k ≤ π )
and frequency ω, two slow time scales T1 = t and T2 =  2 t , in addition to the
original one T0 = t, along with the spatial long scale, and one slow spatial variable
X1 = x, in addition to the original x = n, where   1 is a small dimensionless
parameter. Following Taniuti and Yajima [51], the voltage Vn (t), solution of system
(3.111) is sought in the general form

Vn (t) =  1/2 v11 eiθ + v22 e2iθ +  3/2 v33 e3iθ



+ 2 v40 + v42 e2iθ + v44 e4iθ + c.c., (3.112)

where θ = kn − ωt, and vjk = vjk (X1 , T1 , T2 ) are complex amplitudes.


 Substituting
 (3.112) into (3.111) and equating  coefficients
 of like powers of
, eiθ , one obtains many equations. At order  1/2 , eiθ , the following linear dis-
persion relation is derived
   
2 2k 2 k sin k
sin − C0 + 4CS sin ω − i Gω −
2
= 0, (3.113)
L 2 2 L0

with
1 1 1 1 1 1
= + , = − .
L L1 L2 L0 L2 L1
3.3 Modulated Wavetrains in a Dissipative Bi-Inductance Transmission Network 95

Separating the linear dispersion relation (3.113) into real and imaginary parts and
solving the resulting equations in 1/G = R(k) and ω = ω(k) yield
 
1 L0 2 k 2
≡R=   , ω = sin  ,
G 2 cos k2 L C0 + 4CS sin2 k2 2 L C0 + 4CS sin2 k2
(3.114)
with 0 ≤ k ≤ π.
Note that, for C0 < 4CS , relation R = R(k) exhibits one minimum for 0 ≤ k ≤
π/2, which corresponds to the leading unstable mode, as we can see from Fig. 3.18 for
the given network parameters. The minimization of the dispersion relation R = R(k)
leads to a numerical problem for the critical wavenumber k = kC , all parameters
evaluated at kC being referred to as the “critical network parameters”. In particular,
to the critical network parameters

C0C = CSC = 540 pF, L1C = 2L2C = 28 μH, bC = 0.16 V−1 (Lp)

there correspond the critical wavenumber k = kC ≈ 1.318, angular frequency ω =


ωc ≈ 4.45 × 106 , and RC = R(kC ) ≈ RC = 315. In Fig. 3.18a, the horizontal dotted
 critical
line is obtained with the  value R = RC ≈ 315.
Equation of order  5/2 , eiθ gives the following amplitude equation that governs
the evolution of the wave packet

∂v11 ∂ 2 v11 ∂ |v11 |2 ∂v11


i +P + Q |v 11 |2
v 11 + Q3 |v 11 |4
v 11 + γ v 11 + Q1 v 11 + Q2 |v11 |2 = 0.
∂t ∂x2 ∂x ∂x
(3.115)
Here, P = Pr + iPi , Q = Qr + iQi , γ = γr + iγi , and Qj = Qjr + iQji , (j = 1, 2, 3)
depend on the wavenumber k and are given in terms of the network parameters as

e−ik
 eik
 
CS ω2 − 4CS Aω sin [k] − −
+ A2 C0 − 4CS sin2 k2
2L2 2L1
P =−  ,
2C0 ω + 8CS ω sin k2 + iG 2


4 2 C0 bω2 Re(Q)
Q =− ,
2C0 ω + 8CS ω sin 2 + iG
2 k

4C0 bω2 Re(Q)*


γ =−  ,
2C0 ω + 8CS ω sin 2 + iG
2 k
   

C B (1 + A) C0 − 4CS sin2 k2 − 4BCS ω sin [k] + 4iC0 bAω
Q1 =− ,
2C0 ω + 8CS ω sin2 k2 + iG
(3.116)
2C0 bω2 (A3 + AA2 )
Q2 = −  + Q1 ,
2C0 ω + 8CS ω sin2 k2 + iG
2 2 C0 bA2 B
Cω2
Q3 =  − 2
2C0 ω + 8CS ω sin2 k2 + iG
96 3 Transmission of Dissipative Solitonlike Signals …

Fig. 3.18 Dependency a R = R(k) and b ω = ω(k) given by Eq. (3.114) for critical values C0C =
CSC = 540 pF and L1C = 2L2C = 28 μH. The unique minimum of R = R(k) corresponds to the
most unstable mode. Reprint from Ref. [2], Copyright 2022, with permission from Springer

' '2  2     ∗ 
C ' B C0 − 4CS sin2 k2 + 4iC0 bBω + 2C0 bω2 
' C D + A1
×  .
2C0 ω + 8CS ω sin2 k2 + iG

The nonlinear partial differential equation (3.115) is often called the modified or
generalized CQ CGL equation. Its equivalent form reads

i
∂v11
+P
∂ 2 v11 2 ∂v11 + (Q + Q ) |v |2 ∂v11 = 0
+ Q |v11 |2 v11 + Q3 |v11 |4 v11 + γ v11 + Q1 v11 2 1 11
∂t ∂x2 ∂x ∂x
(3.117)

where v11 is the complex conjugate of v11 . Equation (3.117) is most general and is
usually called a generalized CQ CGL equation with derivative terms |v11 |2 ∂v11 /∂x

2
and v11 ∂v11 /∂x [52]. Deissler and Brand [16] showed numerically that these two
additional terms can significantly reduce the speed of pulses and break the symmetry
of their shape [31].

3.3.3 Linear Analysis and Modulational Instability

Here, we address the MI of the Stokes-wave solution of the generalized CQ CGL


equation (3.115). We first note that Eq. (3.115) admits Stokes-wave solution

v11 (x, t) = a0 exp (ik0 x − iω0 t) , (3.118)

where a0 , k0 and ω0 are related by

ω0 − Pk02 + γ + Q |a0 |2 + ik0 |a0 |2 Q2 + Q3 |a0 |4 = 0. (3.119)

Splitting equation (3.119) into real and imaginary parts yields


3.3 Modulated Wavetrains in a Dissipative Bi-Inductance Transmission Network 97
+
ω0 = Pr k02 − γr + (k0 Q2i − Qr ) |a0 |2 − Qr |a0 |4 ,
(3.120)
Q3i |a0 |4 + (Qi + k0 Q2r ) |a0 |2 − Pi k02 + γi = 0.

Solving the second equation of system (3.120) in |a0 |2 yields


 2 
−Qi − Q2r k0 ± Q2r + 4Pi Q3i k02 + 2Qi Q2r k0 + Qi2 − 4γi Q3i
|a0 |2 = .
2Q3i
(3.121)
It follows from Eq. (3.121) that the amplitude |a0 | will be bounded as a func-
tion of k0 only when the global existence condition, Q2r2
+ 4Pi Q3i < 0, is adopted.
Moreover, the square root in Eq. (3.121) will make sense only under the condition
Qi2 − 4γi Q3i > 0 so that the above Stokes-wave solution exists for k0 in a bounded
interval (k0 )lower , (k0 )upper , where

−Qi Q2r + 2 γi Q3i Q2r


2
+ 4Pi γi Q3i
2
− Pi Qi2 Q3i
(k0 )lower = ,
2
Q2r + 4Pi Q3i
−Qi Q2r − 2 γi Q3i Q2r
2
+ 4Pi γi Q3i
2
− Pi Qi2 Q3i
(k0 )upper = .
2
Q2r + 4Pi Q3i

The CW solutions (3.118)–(3.119) are parameterized by the wavenumber k0 . Now


we study the MI of the carrier CW solution (3.118) by letting

v11 = (a0 + δa) exp [i (k0 x − ω0 t)] , (3.122)

where δa is a small perturbation. Substituting (3.122) in (3.115), linearizing it with


respect to the perturbation δa and using (3.119), we obtain

∂δa ∂ 2 δa   ∂δa
i + P 2 + 2ik0 P + |a0 |2 (Q1 + Q2 )
∂t ∂x ∂x
∂δa ∗  
+ a02 Q1 + |a0 |2 Q + ik0 Q2 + 2 |a0 |2 Q3 δa
 ∂x 
+ a02 Q + ik0 Q2 + 2 |a0 |2 Q3 δa∗ = 0. (3.123)

Non trivial solutions of Eq. (3.123) is then looked in the form of


  
a = b1 exp [i (Kx + t)] + b∗2 exp −i Kx + ∗ t , (3.124)

where K and  are the modulation wavenumber and frequency, b1 and b2 are two
complex constants. Substituting (3.124) into (3.123) and using the fact that |b1 | +
|b2 | > 0 yield
( − αr − iαi )2 = β = βr + iβi , (3.125)
98 3 Transmission of Dissipative Solitonlike Signals …

where αi and βr are as


 
αi = |a0 |2 Qi + k0 Q2r + 2 |a0 |2 Q3i − Pi K 2 ,
⎡ 2 2 ⎤
4k0 |P| − |a0 |4 |Q1 |2
⎢ ⎥
⎢ − 4k0 |a0 |2 (Pi Q1i − Pr Q1r + Pi Q2r − Pr Q2r ) ⎥ 2
βr = |P|2 K 4 − ⎢
⎢ + 2 |a |2 (P Q + P Q + k (P Q2i) − P Q )⎥ K

⎣ 0 r r i i 0 r i 2r ⎦

+ |a0 |4 |Q1 + Q2 |2 + 4 |a0 |4 (Pr Q3r + Pi Q3i )


⎡ ⎤
4k0 (Pi Q2r − Pr Q2i )
⎢ ⎥
2⎢
⎢ − 2 |a0 |2 (Qi (Q1r + Q2r ) − Qr (Q1i + Q2i )) ⎥
− |a0 | ⎢ ⎥K

⎣ − 2k0 |a0 | (Q2r (Q1r + Q2r ) + Q2i (Q1i + Q2i ))⎦
2

− 4 |a0 |4 (Q3i (Q1r + Q2r ) − Q3r (Q1i + Q2i ))


 2
+ 2k0 Pr + |a0 |2 (Q1i + Q2i ) K 2
   2
− |a0 |2 Qi + k0 Q2r + 2 |a0 |2 Q3i − Pi K 2
⎡ ⎤
− 2k0 (Q2i Qr − Q2r Qi )
⎢ ⎥
⎢ |a |2
4 ⎢ − 4k0 0 (Q2i Q3r − Q2r Q3i )⎥

− |a0 | ⎢ ⎥.
⎣ + 2k0 (Qi Q2r − Qr Q2i ) ⎦
− 4k0 |a0 | (Q3i Q2r − Q3r Q2i )
2

Solving Eq. (3.125) in  yields


) % ) &
βr + |β| −βr + |β|
 = αr ± + i αi ± . (3.126)
2 2

The MI sets in provided that the imaginary part of the complex angular frequency 
is negative, that is, )
−βr + |β|
Im []± = αi ± < 0. (3.127)
2

For condition (3.127) to be satisfied, it is necessary that αi should be negative:


 
|a0 |2 Qi + k0 Q2r + 2 |a0 |2 Q3i − Pi K 2 < 0. (3.128)

Under condition (3.128), we arrive to



Im []− = αi − (−βr + |β|) /2 < 0.

For Im[]+ to be negative under the condition (3.128), it is sufficient that


3.3 Modulated Wavetrains in a Dissipative Bi-Inductance Transmission Network 99

σ
Pr Qr + Pi Qi + 2 |a0 |2 (Pr Q3r + Pi Q3i ) + < 0, (3.129)
2 |a0 |2 K 2

where σ is the real number given by


# $
− |a0 |4 |Q1 |2 − 4k0 |a0 |2 (Pi Q1i − Pr Q1r + Pi Q2r − Pr Q2r )
σ = K2
+ 2 |a0 |2 k0 (Pr Q2i − Pi Q2r )
⎡ ⎤
4k0 (Pi Q2r − Pr Q2i )
⎢ ⎥
⎢ − 2 |a0 |2 (Qi (Q1r + Q2r ) − Qr (Q1i + Q2i )) ⎥
+ |a0 |2 ⎢
⎢ − 2k |a |2 (Q (Q + Q ) + Q (Q + Q ))⎥ K

⎣ 0 0 2r 1r 2r 2i 1i 2i ⎦

− 4 |a0 |4 (Q3i (Q1r + Q2r ) − Q3r (Q1i + Q2i ))


+4 |a0 |4 k0 (Q2r Qi − Q2i Qr )
 2
− 2k0 Pr + |a0 |2 (Q1i + Q2i ) K 2 − 2αi2 − |P|2 K 4 .

Relation (3.129), in which K verifies condition (3.128), represents the MI criterion


associated to the generalized CQ CGL equation (3.115) or its equivalent form (3.117).
This result generalizes the well-known Lange-Newell criterionfor the stability
 of
Stokes waves [48] by the presence of the additional term σ/ 2 |a0 |2 K 2 , which
 
depends on the amplitude and wavenumber of the carrier k0 , |a0 |2 , as well as on
perturbation wavenumber K.

3.3.4 Coherent Structures

Now, we turn our attention to the coherent structures in our network system. Through-
out this subsection, ω and k are not the same angular frequency and wavenumber
used in the derivation of the amplitude equation; they have the same meaning as
ω0 and k0 used in the Stokes-wave solution of the amplitude equation. The tempo-
ral evolution of coherent structures in the modified/generalized CQ-CGL equation
amounts to the uniform propagation with velocity υ and overall phase oscillations
with frequency ω :

v11 (x, t) = a(z) exp [i(z) − iωt] , z = x − υt. (3.130)

This ansatz covers all possible traveling wave solutions. The substitution of Eq.
(3.130) in the modified/generalized CQ-CGL equation (3.115) gives rise to an ordi-
nary differential system for a and :
100 3 Transmission of Dissipative Solitonlike Signals …
⎧!  2 "

⎪ d d d 2

⎪ ω + υ − P r − P i + γr a − 2Pi ddz da d 2a
dz + (Q2r + 2Q1r ) a dz + Pr dz 2
2 da

⎪ 
dz

dz dz 2


⎪ d
⎨ + Qr − Q2i dz a + Q3r a = 0
3 5

⎪ !  2 "  



⎪ γ + P d 2
− P d
a + 2Pr ddz − υ da 2
+ Pi ddz2a + (Q2i + 2Q1i ) a2 da

⎪ i r dz 2 i dz dz dz

⎪  

⎩ + Qi + Q2r ddz a3 + Q3i a5 = 0.
(3.131)

In the special case of the plane wave, that is, when a(ξ ) = a0 and (z) = kz,
system (3.131) yields, under the conditions Q2r2
+ 4Pi Q3i ≥ 0, Q2r
2
+ 4Pi Q3i < 0,
and Qi − 4γi Q3i > 0 :
2

ω = Pr k 2 − υk − γr + (Q2i − Qr ) a02 − Q3r a04 ,


 2 
−Qi − Q2r k ± Q2r + 4Pi Q3i k 2 + 2Qi Q2r k + Qi2 − 4γi Q3i
a0 =
2
.
2Q3i

We have thus found that plane waves exist for k taken from a bounded interval
(k0 )lower , (k0 )upper , where

−Qi Q2r + 2 γi Q3i Q2r


2
+ 4Pi γi Q3i
2
− Pi Qi2 Q3i
(k0 )lower = ,
2
Q2r + 4Pi Q3i
−Qi Q2r − 2 γi Q3i Q2r
2
+ 4Pi γi Q3i
2
− Pi Qi2 Q3i
(k0 )upper = .
2
Q2r + 4Pi Q3i
0
Seeking (z) in the form  = y(z)dz, we obtains, after some manipulations,
the following three-dimensional dynamical system:

da
= q, (3.132a)
dz ⎧ ⎡  ⎤
1 ⎨ ⎣ Pr y − υy − ω − γr a + 2Pi yq
2
dq ⎦
= Pr
dz |P|2 ⎩ − (Q2r + 2Q1r ) a2 q + (Q2i y − Qr ) a3 − Q3r a5
# $1
(Pi y − γi ) a + (υ − 2Pr y) q
+Pi , (3.132b)
− (Q2i + 2Q1i ) a2 q − (Qi + Q2r y) a3 − Q3i a5
dy 1 q
2y

= P r (υ − 2P r y) − 2P i
dz |P|2 a
1 2

+ 2 Pr Pi y − γi − (Q2i + 2Q1i ) aq − (Qi + Q2r y) a2 − Q3i a4
|P|

3
−Pi Pr y2 − υy − ω − γr − (Q2r + 2Q1r ) aq + (Q2i y − Qr ) a2 − Q3r a4 .
(3.132c)
3.3 Modulated Wavetrains in a Dissipative Bi-Inductance Transmission Network 101

It is clear that a = 0 is singular point for system (3.132a)–(3.132c). To overcome


this difficulty, we can use the so-called ”blowup” transform, or the “σ -process" [53],
by setting
x = q/a. (3.133)

The singularized (a, q, y) system (3.132a)–(3.132c) under transformation (3.133)


reduces to the following desingularized (a, x, y) system

a = ax, (3.134a)
⎧ # $⎫

⎪ Pr y2 − υy − ω − γr − (Q2r + 2Q1r ) a2 x ⎪ ⎪

⎪ P ⎪


⎨ r + (Q y − Qr ) a2 − Q a4 ⎪

 2 1 2i 3r
x = −x + # $ , (3.134b)
|P|2 ⎪

⎪ Pi y − γi + υx − (Q2i + 2Q1i ) a2 x ⎪


⎪ ⎪
⎩ + Pi
⎪ ⎪

− (Qi + Q2r y) a2 − Q3i a4



⎪ P (υ − 2Pr y) − 2Pi2 y x ⎪

⎪ r
⎪ ⎪


⎪ # $ ⎪


⎪ P y − γ − (Q + 2Q ) a 2 x ⎪

1 ⎨ + Pr
i i 2i 1i ⎬

y = 2 4 . (3.134c)
⎪ − (Q + Q y) a − Q a ⎪
|P| ⎪
2
⎪ #
i 2r 3i
$ ⎪


⎪ ⎪

⎪ Pr y2 − υy − ω − γr − (Q2r + 2Q1r ) a2 x ⎪ ⎪


⎩ − Pi ⎪

2
+ (Q2i y − Qr ) a − Q3r a 4

The desingularized (a, x, y) system (3.134a)–(3.134c) has the invariant plane a = 0.


Thus, the σ -process has transformed the former singular point into an invariant plane.
On the invariant plane a = 0, the system (3.134a)–(3.134c) reduces to

Pr2 2 Pi υ P 2 − Pr υ Pr (ω + γr ) + Pi γi
x = −x2 + y + x+ i y− , (3.135a)
|P| 2
|P| 2
|P|2 |P|2
Pi Pr Pr υ Pi (Pr + υ) Pi (ω + γr ) − Pr γi
y = − 2 y2 − 2xy + x+ y+ .
(3.135b)
|P| |P| 2
|P| 2
|P|2

It is clear that system (3.135a)–(3.135b) does not depend on Qi , Qr , Q1r , Q1i , Q2r ,
Q2i , Q3r , Q3i . Following the results of Ref. [54], we can easily show that system
(3.135a)–(3.135b) characterizes the behavior of solutions of the singularized system
(a, q, y) system at a → 0.
In the following, system (3.134a)–(3.134c) is addressed in the case when

ω = −γr and γi = 0. (3.136)



We should remember that γi = Im (γ ) and γr = Re γ are free parameters so that
one may indeed set γi = 0 and choose ω as a function of the independent free
parameter γr . Under these conditions, the desingularized (a, x, y) system (3.134a)–
(3.134c) becomes a dynamical system in the standard form. This system gives rise
102 3 Transmission of Dissipative Solitonlike Signals …

Fig. 3.19 The phase portrait of system (3.134a)–(3.134c) for variables (a, x, y) in the invari-
ant plane a = 0 when Pi (Pr + 2υ) ≥ 0 : (a1 ) υ = −2.210 1 × 107 < υmin (unstable focus); (a2 )
υ = 2.2101 × 107 > υmax (unstable focus); (a3 ) υ = 91313 > 0 ∈ (υmin , υmax ) (unstable node);
(a4 ) υ = 2.2101 × 106 > υmax with Pi (Pr + 2υ) = 0 (center); (a5 ) υ = −99541 ∈ (υmin , υmax )
(unstable node); (a6 ) υ = 0 ∈ (υmin , υmax ) (unstable node); (a7 ) υ = υmin = −99542 (unstable
node); (a8 ) υ = υmax = 91317 (unstable node). Reprint from Ref. [2], Copyright 2022, with
permission from Springer

Fig. 3.20 The phase portrait of system (3.134a)–(3.134c) for variables (a, z) in the invariant plane
a = 0 with z = x − υt, when Pi (Pr + 2υ) < 0 and υ = 2.210 1 × 107 > υmax (a stable focus).
Reprint from Ref. [2], Copyright 2022, with permission from Springer
3.3 Modulated Wavetrains in a Dissipative Bi-Inductance Transmission Network 103

Fig. 3.21 Behavior of |v11 (x, t)| = |a(z = x − υt)| corresponding to (a1 ) , (a2 ) , (a3 ) , (a4 ) , (a5 ) ,
(a6 ) , (a7 ) , and (a8 ), respectively, from Fig. 26. Plots show the dependence of v11 on x and t. Waves
shown in panels (b1 ), (b5 ) and (b7 ) travel to the left, while other waves travel to the right. Reprint
from Ref. [2], Copyright 2022, with permission from Springer

Fig. 3.22 Behavior of |v11 (x, t)| = |a(z = x − υt)| corresponding to Fig. 3.21. This figure shows
the dependence of v11 on x and t for the wave traveling to the right. Reprint from Ref. [2], Copyright
2022, with permission from Springer
104 3 Transmission of Dissipative Solitonlike Signals …

to a number of critical points, (0, 0, 0) being one of them, which corresponds to the
zero amplitude, a = 0. If the coefficients of the amplitude equation (3.115) satisfy
the conditions

Qr Q3i − Qi Q3r = 0and (Pr Qi − Pi Qr ) (Pi Q3r − Pr Q3i ) > 0,

then  
± (Pr Qi − Pi Qr ) / (Pi Q3r − Pr Q3i ), 0, 0

are also critical points of system (3.134a)–(3.134c) under conditions (3.136). Using
the linearization in the (a, x, y) space, the stability of the fixed point (0, 0, 0) is
addressed here. In similar way, we can address the stability of the other critical
points.
The result is that point (0, 0, 0) has a two-dimensional unstable manifold when
Pi (2υ + Pr ) ≥ 0, and a one-dimensional one, when Pi (2υ + Pr ) < 0. The respec-
tive linearized system has real eigenvalues for υ ∈ [υmin , υmax ] , and complex eigen-
values when υ ∈ / [υmin , υmax ] , where
 
Pi2 Pr + Pi Pr |P| Pi2 Pr − Pi Pr |P|
υmin = min , ,
2Pr2 2Pr2
 2 
Pi Pr + Pi Pr |P| Pi2 Pr − Pi Pr |P|
υmax = max , .
2Pr2 2Pr2

The results are summarized by means of phase portraits shown in Fig. 3.19 for
Pi (2υ + Pr ) ≥ 0, and in Fig. 3.20, when Pi (2υ + Pr ) < 0, for the critical-line
parameters Lp . For Pi (2υ + Pr ) ≥ 0, we depict |v11 (x, t)| = |a(z)|, with z =
x − υt in Fig. 3.21 for Pi (2υ + Pr ) ≥ 0, and in Fig. 3.22 for Pi (2υ + Pr ) < 0.
To analyze the propagation of the wave, we here take  = 10−2 . As initial condi-
tions, we use (a0 , x0 , y0 ) = (a(0), x(0), y(0)) = (0.01, 0.001, −0.001). We can see
from panels (b1 ), (b5 ) and (b7 ) of Fig. 3.21 that waves are absorbed somewhere to
the left of z = 0 (and to the right in that of Fig. 3.22). It is also seen from panels Fig.
3.21 (b2 ) , (b3 ) , (b6 ) and (b8 ) that waves are emitted from somewhere to the left of
z = 0. The wave shown in panel (b4 ) of Fig. 3.21 has a uniform profile.
In conclusion, modulated wavetrains in the discrete nonlinear-dispersive dissipa-
tive transmission network sketched in Fig. 3.17 have been addressed in this Section.
With the use of the semi-discrete approximation and the reductive perturbation
method, it is proven that the evolution of nonlinear excitations in the network is
governed by the modified/generalized CQ CGL equation with the cubic derivative
terms. The MI analysis of the Stokes waves is included, along with the MI criterion.
Also, a real-valued system of ordinary differential equations was derived to describe
coherent structures of the amplitude equation.
References 105

References

1. D. Yemélé, P.K. Talla, T.C. Kofané, Dynamics of modulated waves in a nonlinear discrete LC
transmission line: dissipative effects. J. Phys. D: Appl. Phys. 36, 1429 (2003)
2. E. Kengne, C.N. Bame, Dynamics of modulated wave trains in a discrete nonlinear-dispersive
dissipative bi-inductance transmission line. Physica Scripta 71, 423–432 (2005)
3. S.M. Ngounou, F.B. Pelap, Nonlinear continuous bi-inductance electrical line with dissipative
elements: dynamics of the low frequency modulated waves. Chin. Phys. B 29, 040502 (2020)
4. J. Soto-Crespo, N. Akhmediev, Exploding soliton and front solutions of the complex cubic-
quintic Ginzburg-Landau equation. Math. Comput. Simul. 69, 526 (2005)
5. Nisha, N. Maan, A. Goyal, T.S. Raju, C.N. Kumar, Chirped Lambert W-kink solitons of the
complex cubic-quintic Ginzburg-Landau equation with intrapulse Raman scattering. Phys. Lett.
A 384, 126675 (2020)
6. J. Soto-Crespo, N. Akhmediev, K.S. Chiang, Phys. Lett. A 291, 115 (2001)
7. S.V. Gurevich, C. Schelte, J. Javaloyes, Impact of high-order effects on soliton explosions in
the complex cubic-quintic Ginzburg-Landau equation. Phys. Rev. A 99, 061803(R) (2019)
8. Y. Du, X. Shu, Dynamics of soliton explosions in ultrafast fiber lasers at normal-dispersion.
Opt. Exp. 26, 5564 (2018)
9. J. Peng, S. Boscolo, Z. Zhao, H. Zeng, Breathing dissipative solitons in mode-locked fiber
lasers. Sci. Adv. 5, eaax1110
10. J.H. Lambert, Observationes Variae in Mathesin Puram. Acta Helvetica Physico-Mathematico-
Anatomico-Bota-nico-Medica 3, 128-168 (1758)
11. J.H. Lambert, in Nouveaux mémoires de l’Académie royale des sciences et belles-lettres, voI.
1, for 1770 (Berlin 1772)
12. R.M. Corless, G.H. Gonnet, D.E.G. Hare, D.J. Jeffrey, D.E. Knuth, On the LambertW function.
Adv. Comput. Math. 5, 329 (1996)
13. J.D. Anderson, Introduction to Flight, 3rd edn. (McGraw-Hill, New York, 1989)
14. D. Veberič, PSEM approximations for both branches of lambert function with applications.
Comput. Phys. Commun. 183, 2622 (1012)
15. J. Duan, P. Holmes, Generalized Ginzburg-Landau equation. Proc. Edinburgh Math. Soc. 38,
77 (1995)
16. S.I. Mostafa, Analytical study for the ability of nonlinear transmission lines to generate solitons.
Chaos Solit. Fract. 39, 2125–2132 (2009)
17. E. Kengne, Engineering chirped Lambert W-kink signals in a nonlinear electrical transmission
network with dissipative elements. Eur. Phys. J. Plus 136, 266 (2021)
18. T. Kakutani, K. Michihiro, Marginal state of modulational instability—Note on Benjamin-Feir
instability. J. Phys. Soc. Jpn. 52, 4129 (1983)
19. Alka, A. Goyal, R. Gupta, C.N. Kumar, Chirped femtosecond solitons and double-kink solitons
in the cubic-quintic nonlinear Schrödinger equation with self-steepening and self-frequency
shift. Phys. Rev. A 84, 063830 (2011)
20. R. Pal, A. Goyal, S. Loomba, T.S. Raju, C.N. Kumar, Compression of optical similaritons
induced by cubic-quintic nonlinear media in a graded-index waveguide. J. Nonlinear Opt.
Phys. Mater. 25, 1650033 (2016)
21. S. Chen, F. Baronio, J.M. Soto-Crespo, P. Grelu, D. Mihalache, Versatile rogue waves in scalar,
vector, and multidimensional nonlinear systems. J. Phys. A: Math. Theor. 50, 463001 (2017)
22. E. Kengne, A. Lakhssassi, W.M. Liu, Dynamics of modulated waves in a lossy modified
Noguchi electrical transmission line. Phys. Rev. E 91, 062915 (2015)
23. J.A. Giannini, R.I. Joseph, The propagation of bright and dark solitons in lossy optical fibers.
IEEE J. Quant. Electron. 26, 2109–2114 (1990)
24. D.L. Sekulic, M.V. Sataric, M.B. Zivanov, J.S. Bajic, Soliton-like pulses along electrical non-
linear transmission line. Electron. Electr. Eng. 5, 53–58 (2012)
25. E. Kengne, R. Vaillancourt, Propagation of solitary waves on lossy nonlinear transmission
lines. Int. J. Mod. Phys. B 23, 1–19 (2009)
106 3 Transmission of Dissipative Solitonlike Signals …

26. S.D. Yamigno, Propagation of dark solitary waves in the Korteveg-Devries-Burgers equation
describing the nonlinear RLC transmission. J. Mod. Phys. 5, 394 (2014)
27. E. Kengne, E.B. Ngompe Nkouankam, A. Lakhssassi, Dynamics of spatiotemporal modulated
damped signals in a nonlinear RLC transmission network. Nonlinear Dyn. 104, 4181–4201
(2021)
28. M. Remoissenet, Waves Called Solitons, 3rd edn. (Springer, Berlin, 1999)
29. E. Kengne, A. Lakhssassi, W.M. Liu, Modeling of matter-wave solitons in a nonlinear inductor-
capacitor network through a Gross-Pitaevskii equation with time-dependent linear potential.
Phys. Rev. E 96, 022221 (2017)
30. P. Marquié, J.M. Bilbault, M. Remoissenet, Nonlinear Schrödinger models and modulational
instability in real electrical lattices. Physica D 87, 371–374 (1995)
31. P. Marquié, J.M. Bilbault, M. Remoissenet, Generation of envelope and hole solitons in an
experimental transmission line. Phys. Rev. E 49, 828 (1994)
32. D. Yemélé, P.K. Talla, T.C. Kofané, Dynamics of modulated waves in a nonlinear discrete LC
transmission line: dissipative effects. J. Phys. D: Appl. Phys. 36, 1429–1437 (2003)
33. E. Kengne, W.M. Liu, Transmission of rogue wave signals through a modified Noguchi elec-
trical transmission network. Phys. Rev. E 96, 062222 (2019)
34. E. Kengne, W.M. Liu, Engineering rogue waves with quintic nonlinearity and nonlinear dis-
persion effects in a modified Nogochi nonlinear electric transmission network. Phys. Rev. E
102, 012203 (2020)
35. T. Taniuti, N. Yajima, Perturbation method for a nonlinear wave modulation I. J. Math. Phys.
10, 1369–1372 (1969)
36. I.S. Aranson, L. Kramer, The world of the complex Ginzburg-Landau equation. Rev. Mod.
Phys. 74, 99 (2002)
37. E. Kengne, S.T. Chui, W.M. Liu, Modulational instability criteria for coupled nonlinear trans-
mission lines with dispersive element. Phys. Rev. E 74, 036614 (2006)
38. E. Kengne, A. Lakhssassi, W.M. Liu, R. Vaillancourt, Phase engineering, modulational insta-
bility, and solitons of Gross–Pitaevskii-type equations in 1 + 1 dimensions, Phys. Rev. E 87
39. E. Kengne, W.M. Liu, B.A. Malomed, Spatiotemporal engineering of matter-wave solitons in
Bose-Einstein condensates. Phys. Rep. 899, 1–62 (2021)
40. C. Lange, A.C. Newell, A stability criterion for envelope equations. SIAM J. Appl. Math. 27,
441–456 (1974)
41. F.B. Pelap, M.M. Faye, A modified stability criterion for envelope equations. Phys. Scr. 71,
238 (2005)
42. T.B. Benjamin, J.E. Feir, The disintegration of wave trains on deep water Part 1. Theory. J.
Fluid Mech. 27, 417 (1967)
43. S. Amiranashvili, E. Tobisch, Extended criterion for the modulation instability. New J. Phys.
21, 033029 (2019)
44. S.G. Sajjadi, A note on Benjamin-Feir instability for water waves. Adv. Appl. Fluid Mech. 17,
17 (2015)
45. V.E. Zakharov, L.A. Ostrovsky, Modulation instability: the beginning. Physica D 238, 540
(2009)
46. A. Hasegawa, Y. Kodama, Solitons in Optical Communications (Oxford University Press,
Oxford, 1995)
47. Y.S. Kivshar, B.A. Malomed, Dynamics of solitons in nearly integrable systems. Rev. Mod.
Phys. 61, 763 (1989)
48. C.G. Lange, A.C. Newell, A stability criterion for envelope equations. SIAM J. Appl. Math.
27, 441–456 (1974)
49. W.M. Liu, E. Kengne, Schrödinger Equations in Nonlinear Systems (Springer, Singapore, 2019)
50. R. Marquié, J.M. Bilbault, M. Remoissenet, Nonlinear Schrödinger models and modulational
instability in real electrical lattices. Physica D 87, 371–374 (1995)
51. T. Taniuti, N. Yajima, Perturbation method for a nonlinear wave modulation II. J. Maths. Phys.
10, 1369 (1969)
References 107

52. H.R. Brand, P.S. Lomdahl, A.C. Newell, Benjamin-Feir turbulence in convective binary fluid
mixtures. Physica (Amsterdam) 23D, 345 (1986)
53. V.I. Arnold, Mathematical Methods of Classical Mechanics (Springer, Berlin, 1978)
54. C.K.R.T. Jones, T. Kapitula, J. Powell, Nearly real fronts in a Ginzburg-Landau equation. Proc.
Roy. Soc. Edinburgh A 116, 193–206 (1990)
55. A. Noguchi, Solitons in a nonlinear transmission line. Electron. Commun. Jpn. 57A, 9 (1974)
Chapter 4
Emission of Rogue Wave Signals
in Nonlinear Electrical Transmission
Networks

Abstract In this chapter, we analytical discuss the transmission of modulated rogue


waves through electric nonlinear transmission networks when the losses are either
neglected or taken into consideration. For a given electrical model of the nonlinear
transmission network, we employ the reductive perturbation method in the semi-
discrete limit to reduce the dynamics of modulated waves propagating along the
network to nonlinear partial differential equations (NPDEs) such as equations of NLS
type either without or with a dissipative term or complex cubic/quintic Ginzburg-
Landau equation with derivative terms. Using the derived NPDE of a given physical
system, we investigate the phenomenon of baseband modulational instability for the
physical model to be studied. For each physical model to be studied, the integrable
conditions under which rational polynomial solutions of the derived NPDE exist
are obtained and first-order or/and second-order rational polynomial solutions are
built. Using these rational polynomial solutions, we study in details the emission
of first-order or/and second-order modulated rogue waves along the given electrical
network.

4.1 Emission of Rogue Wave Signals Through the Modified


Noguchi Electric Transmission Network

4.1.1 Introduction

Firstly observed in the ocean [1–3], rogue waves (alias killer, giant, or extreme storm
waves) are waves that are about twice or thrice higher than expected for the ocean
state. Rogue waves can appear out of nowhere [4]. Mathematically, rogue waves
are rational polynomial solutions of a class of NPDEs such as equations of NLS
type [4–8]. In the physical context, rogue waves are located both in space and time,
propagate on a nonzero continuous wave background, and concentrate the energy of
the CW background into a small region [3, 6, 9–17]. As physical and mathematical
phenomena, rogue waves occur in many fields of nonlinear science including optical
fibers [18], super fluids [19], quantum mechanics [20], BECs [21], NLTNs [22],
atmospherics [23], plasmas [24], etc.
© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2022 109
E. Kengne and W. Liu, Nonlinear Waves,
https://doi.org/10.1007/978-981-19-6744-3_4
110 4 Emission of Rogue Wave Signals in Nonlinear Electrical Transmission Networks

The present Section deals with the propagation of RWs in the modified Noguchi
electrical NLTN depicted in Fig. 2.1. These RWs are produced as solutions of a
NLS equation with an additional time-varying linear-potential term [25]. The RWs
propagate at bandwidth frequencies at which the network may exhibit baseband MI.
Effects of network parameters on characteristics of the RW parameters are consid-
ered. We shown how to use parameters and the propagation frequency to maintain
or suppress the RWs in a network such as that of Fig. 2.1.

4.1.2 Modulated Waves and Linear Analysis

4.1.2.1 Network Equations

The physical model to be studied in this Section is the 1D modified Noguchi electrical
NLTN shown in Fig. 2.1 made of N identical cells. Assuming the capacitance-voltage
(C − V ) characteristics to be of the form

d Qn  
C (Vb + Vn ) =  C0 1 − 2αVn + 3βVn2 , (4.1)
d Vn

we employ the Kirchhoff’s laws and obtain the following network system

d2  
2
Vn + λ (2Vn − Vn+1 − Vn−1 ) − αVn2 + βVn3
dt
+ ω02 Vn + u 20 (2Vn − Vn+1 − Vn−1 ) = 0, (4.2)

where n = 1, 2, ..., N , Vb and C0 = C(Vb ) are respectively the equilibrium volt-


age and the capacitance scale, α and β are nonlinearity coefficients that determine
the electric charge Q n stored in the n-th capacitor in the network, Vn is the voltage
across the n-th capacitor, u 20 = (L 1 C0 )−1 and ω02 = (L 2 C0 )−1 are the characteristic
frequencies of the network, and λ = C S /C0 is the dimensionless dispersive param-
eter.
In the following, we focus our attention on nonlinear modulated waves centered at
(ω, k), where ω = ω p = 2π f p and k = k p are respectively the angular frequency and
the wavenumber. We assume that the perturbation voltage |Vn /Vb |  1 and apply
the reductive perturbation method in the semi-discrete limit [26]. We then introduce
one slow space and one slow time variables, x = ε(n − υg t) and τ = ε2 t, where
0 < ε  1 is a small ordering parameter, υg is the group velocity defined as
 2 
dω u 0 − λω02 sin k
υg = =  2 , (4.3)
dk ω 1 + 4λ sin2 (k/2)
4.1 Emission of Rogue Wave Signals Through the Modified … 111

and the angular frequency ω is related to the wavenumber k through the following
linear dispersion relation
   
ω2 1 + 4λ sin2 (k/2) − ω02 + 4u 20 sin2 (k/2) = 0. (4.4)

Following Marquié et al. [27], we seek the solution of system (4.2) in the general
form [27, 28]
 
Vn (t) = εψ(x, τ ) exp (iθ ) + ε2 ψ10 (x, τ ) + ψ20 (x, τ ) exp (2iθ ) + c.c., (4.5)

where θ = (kn − ωt) is the rapidly varying phase, ψ(x, τ ), ψ10 (x, τ ), and ψ20 (x, τ )
are respectively the fundamental, dc and second-harmonic terms. During the numer-
ical computations, the wavenumber k will be taken from the first Brilloum zone
(0 ≤ k ≤ π ). In the following, we assume the group velocity (4.3) to be nonnegative
in the first Brilloum zone, which is possible only when the dispersive parameter λ
satisfies the following restriction [29]:

L2 L2
0≤λ< , i.e., 0 ≤ C S < C0 . (4.6)
L1 L1

Under restriction (4.6), the propagating frequency f p = ω/2π will be an increasing


function of the wavenumber k in the first Brilloum zone, limited by the cut-off
frequencies f 0 = ω0 / (2π ) and f c = ω(π )/ (2π ) ; therefore, f 0 ≤ f p ≤ f c [27].
Under all the above assumptions, we find that [30],

2αωυg2
ψ10 (x, τ ) = |ψ|2 + χ0 (τ )x, (4.7a)
υg2 − u 20
4αω2
ψ20 (x, τ ) =   ψ 2, (4.7b)
4ω2 − ω02 + 4 4λω2 − u 20 sin2 k

∂ψ ∂ 2ψ
i + P 2 + Q |ψ|2 ψ + χ xψ = 0, (4.8)
∂τ ∂x

where χ0 (τ ) is an arbitrary real function and


   
υg2 u 20
k λω
P = P(k) = − 1 + 4λ sin2 − + cos k − 2λυg sin k, (4.9a)
2ω 2ω2 2

3β 4α 2 ω2 2α 2 υg2
Q = Q(k) = ω −   − 2 , (4.9b)
2 4ω2 − ω02 + 4 4λω2 − u 20 sin2 k υg − u 20
αω
χ = χ (τ ; k) = − χ0 (τ ). (4.9c)
1 + 4λ sin2 (k/2)
112 4 Emission of Rogue Wave Signals in Nonlinear Electrical Transmission Networks

In the special case when χ = 0, Eq. (4.8) coincides with the standard NLS equation.
When χ = 0, it turns to a NLS equation with a linear density profile (alias the GP
equation in the BEC context). It is important to note that Eq. (4.8) can be reduced
to the standard NLS equation [31] by means of transformations [30]


ψ (x, τ ) ≡ ψ̃ (x̃, τ ) exp i(τ )x − i P dτ 2 (τ ) , (4.10)

and
x̃ ≡ x − 2P dτ (τ ), (τ ) ≡ λ(τ )dτ. (4.11)

4.1.2.2 Linear Stability of Plane Waves with a Time-Varying Amplitude

The NLS equation (4.8) admits the following plane wave solution
⎡  2 ⎤
T
1 1 x + σx
ψ(x, τ ) = √ ψ0 exp ⎣−i  (y)dy + i ⎦, (4.12)
0 (τ + τ0 ) 4P τ + τ0
T0

where

1 P dσ 20 2 ψ02 Q
T =− , (T ) =
= 4P (τ + τ0 ) χ (τ ),  σ + ;
2 (τ + t ) dτ 0 T
0 0 16P 2

here, ψ0 , τ0 , and 0 = 0 are real constants satisfying condition 0 (τ + τ0 ) > 0. The


phenomenon of the MI for network system of Fig. 2.1 is investigated by perturbing
the plane wave solution (4.12) as follows,
⎡  2 ⎤
T
ψ0 + δψ 1 x + σx
ψ(x, τ ) = √ exp ⎣−i  (y)dy + i ⎦, (4.13)
0 (τ + τ0 ) 4P τ + τ0
T0

where δψ is a small perturbation. Using the linear stability analysis and denoting
by respectively K and  the modulation wavenumber and the complex modulation
angular, we arrive to the following time-dependent dispersion relation:
 2

2 σ K 2Q ψ02
(T ) − 0 − K 2 40 K 2 − (τ + τ0 ) = 0. (4.14)
P 2 P 0

The modulational instability may set in provided that τ0 < 0 and Im[] = 0, whence
the following necessary condition for the MI
4.1 Emission of Rogue Wave Signals Through the Modified … 113

Q 2 τ + τ0
K2 − 2 ψ < 0. (4.15)
P 0 0

Since 0 (τ + τ0 ) > 0, for condition (4.15) to be satisfied, it is necessary that P Q >


0 : this MI criterion of the plane wave solution (4.12) coincides with the MI criterion
P Q > 0 for plane-wave solutions of the standard NLS equation. The corresponding
local MI growth rate (gain) reads

Q 2 τ + τ0
|Im ()| = 20 K 2 2 ψ − K 2. (4.16)
P 0 0

It is important to note that the CW solution (4.12) will be stable under modulation as
soon as P Q < 0. Although parameter χ (τ ) of the linear potential does not affect the
MI criterion, it affects implicitly the local MI growth rate. Indeed, the dependence
of the growth rate (4.16) on “time” τ is a consequence of the presence of χ (τ ) in the
amplitude equation (4.8) of the network.
Employing the network parameters [30]

L 1 = 220 µH,L 2 = 470 µH,C0 = 370 pF,C S = 56 pF, α = 0.21V−1 ,


β = 0.0197V−2 , (4.17)

we display, respectively, in Fig. 4.1a–c linear dispersion curves showing the propagat-
ing frequency f = ω/ (2π ), the dispersion coefficient P, the nonlinearity coefficient
Q, and the product P Q versus the wavenumber k.
Figure 4.1a shows four different regions concerning the MI and possible soliton
solutions (regions of positive P Q correspond to envelope solitons, while region of
negative P Q lead to hole solitons):

• Region (1): P > 0 and Q < 0 (here, PQ < 0);


• Region (2): P > 0 and Q > 0 (here, PQ > 0);
• Region (3): P < 0 and Q > 0 (here, PQ < 0);
• Region (4): P < 0 and Q < 0 (here, PQ > 0).

4.1.3 Construction of Rogue Waves in the Lossless Network


Under the Condition P Q > 0

We now turn to the construction of RWs in the lossless network of Fig. 2.1. For this
purpose, we first need to find rogue wave solutions of the amplitude equation (4.8).
114 4 Emission of Rogue Wave Signals in Nonlinear Electrical Transmission Networks

Fig. 4.1 a The linear-dispersion curve showing frequency f = ω/(2π ) as a function of wavenum-
ber k. The allowed band [ f 0 , f c ] is divided in four regions concerning the instability of the system,
depending on the sign of P Q. b Plots of the dispersion and nonlinearity coefficients, in the form of
0.2P × 10−6 and Q × 10−5 , versus wavenumber k. c Product P Q as a function of k. The network
parameters used in plots (a)–(c) are given in the text. Reprint from Ref. [25], Copyright 2022, with
permission from American Physical Society

4.1.3.1 First-Order and Second-Order RW Solutions of Eq. (4.8)

To build the first- and the second-order RW solutions of Eq. (4.8), we use the following
seed solutions  
ψ(x, τ ) = ± exp i (β0 x + γ0 + Qτ ) , (4.18)

where β0 and γ0 satisfy the differential system

dβ0 dγ0
− χ = 0, + Pβ02 = 0. (4.19)
dτ dτ
Using the seed solution (4.18) and applying the direct method, we seek for the first-
order and second-order RW solutions of Eq. (4.8) in the form


F(ξ, τ )  
ψ1,2 (x, τ ) = ±1 + exp i (β0 x + γ0 + Qτ ) , (4.20)
G(ξ, τ )

where F(ξ, τ ) is a complex polynomial and G(ξ, τ ) is a real polynomial satisfying


the conditions deg F < deg G and G(ξ, τ ) = 0 for all ξ and τ , and ψ1 and ψ2 stand
for respectively the first- and second-order rational polynomial solution of Eq. (4.8).
For the first-order and the second-order RWs, we ask that deg G = 2 and deg G = 6,
respectively. Under these assumptions, we insert Eq. (4.20) into Eq. (4.8) and arrive
to


4 + 8i Q/2Pξ  
ψ1 (ξ, τ ) = −1 + exp i (β0 x + γ0 + Qτ ) ,(4.21)
1 + 4Q τ + 2(Q/P)ξ
2 2 2


g2 (ξ, τ ) + i h 2 (ξ, τ )  
ψ2 (ξ, τ ) = 1 + exp i (β0 x + γ0 + Qτ ) , (4.22)
d2 (ξ, τ )
4.1 Emission of Rogue Wave Signals Through the Modified … 115

where
  
3 Q 2 3 Q 2 3
g2 (ξ, τ ) = − + ξ + 5Q 2 τ 2 + ξ + Q2τ 2 + , (4.23a)
4 2P 4 2P 4
  2
15 3Q 2 Q 2
h 2 (ξ, τ ) = Qτ − + Q τ − 2 2
ξ +2 ξ +Q τ
2 2
, (4.23b)
8 2P 2P
   
3 6Q 2 1 Q 2 2
d2 (ξ, τ ) = 1 + 44Q τ +
2 2
ξ + 3Q τ −
2 2
ξ
64 P 4 2P
 
1 Q 2 3
+ Q2τ 2 + ξ , (4.23c)
3 2P
d X1
ξ = x + X 1 (τ ), + 2Pβ0 = 0. (4.23d)

4.1.3.2 Propagation of RWs Through the Lossless Network

In the following, we employ the above first- and second-order rational polyno-
mial solutions of the amplitude equation (4.8) to investigate the dynamics of first-
and second-order modulated rogue waves propagating along the network system of
Fig. 2.1. Such investigation is carried out via typical physical example, based on the
form of the magnitude χ (τ ) of the linear potential term of the amplitude equation
(4.8).
The case of constant χ (τ ) (time-independent linear density profile) As the first
example, we consider the time-independent linear density profile and set χ (τ ) = χ =
constant. Solving then Eqs. (4.19) and (4.23d) yields

β0 = χ τ + β00 , γ0 (τ ) = −(3χ )−1 P (χ τ + β00 )3 + γ00 ,

and
ξ(τ ) = x − (P/χ ) (χ τ + β00 )2 + X 10

if χ = 0, and β0 = γ0 = 0 and ξ = x, if χ = 0, where β00 , γ00 , and X 10 are three


free real constants.
For the numerical value k = 0.75 rad/cell of the wavenumber and for the net-
work parameters (4.17), we display in Fig. 4.2 the typical structure of the first-order
(left panels) and second-order (right panels) modulated rogue waves obtained with
respectively the rational polynomial solutions (4.21) and (4.22). As we can see from
plots of Fig. 4.2, the first-order [second-order] modulated rogue wave is formed of
one hump [four humps] located around the centre (0, 0).
116 4 Emission of Rogue Wave Signals in Nonlinear Electrical Transmission Networks

Fig. 4.2 RWs (rogue waves) |ψ(x, τ )| in the network of Fig. 2.1 for a time-independent linear-
density profile with χ(τ ) = 1. The structure of the a first-order and b second-order RWs is provided
by Eqs. (4.21) and (4.22), respectively. Contour plot of the c first-order and d second-order RWs,
plotted as per Eqs. (4.21) and (4.22). respectively. Different plots are generated with wavenumber
k = 0.75 rad/cell and network parameters (4.17). Solution parameters used in these plots are β00 =
X 10 = 0. Reprint from Ref. [25], Copyright 2022, with permission from American Physical Society

The case of the temporal periodic modulation of the linear density profile Our
second example corresponds to a temporal periodic modulation of the linear potential
whose magnitude is taken as
 
χ (τ ) = (2P)−1 2 + 452 cos (τ ) ,

with the constant frequency . From Eqs. (4.19) and (4.23d), we obtain for this
example that

β0 (τ ) = (2P)−1 (2τ + 45 sin (τ ) + β00 ) ,


ξ(τ ) = x − τ 2 − 2β00 τ + 45 cos (τ ) + X 10 ,

where β00 and X 10 are two constants of integration.


A typical example of the evolution of the first-order (left panels) and second-
order (right panels) modulated rogue waves, obtained with the use of the above
found rational polynomial solution (4.21) and (4.22), and propagating through the
network of Fig. 2.1. is depicted in Fig. 4.3 for the network parameters (4.17). Different
4.1 Emission of Rogue Wave Signals Through the Modified … 117

plots of Fig. 4.3 show RWs moving along an oscillating trajectory. Also, plots of
Fig. 4.3 reveal that the first-order [second-order] RW features only one extremum
[two extrema].
The case of an exponentially decaying/growing linear density profile Our last
example is obtained when the strength of the linear-potential undergoes a temporal
modulation, that is, χ (τ ) = χ0 exp (χ1 τ ), χ0 and χ1 being nonzero constants. In this
special case, we obtain from Eqs. (4.19) and (4.23d) that

β0 (τ ) = (χ0 /χ1 ) exp (Pτ ) + β00 ,


  
ξ(τ ) = x − 2P χ0 /χ12 exp (χ1 τ ) + β00 τ + X 10 ,

where β00 and X 10 are two constants of integration.


For the present example, we show respectively in Figs. 4.4 and 4.5 the dynamics
of the first-order and the second-order modulated RWs generated with the rational
polynomial solutions (4.21) and (4.22) and showing the evolution of modulated rogue
waves along the network system of Fig. 2.1. As we can see from plots of Fig. 4.4,

Fig. 4.3 The dynamics of the first-order and second-order RWs, |ψ(x, τ )|, of the net-
work of Fig. 2.1 for a temporal
 periodic modulation of the linear density profile, χ(τ ) =
(2P)−1 2 + 452 cos (τ ) with  = 104 for wavenumber k = 0.75 rad/cell and network param-
eters (4.17). The evolution plot are shown for a the first-order |ψ1 (x, τ )| and b the second-order
|ψ2 (x, τ )| RWs produced by Eqs. (4.21) and (4.22), respectively. Contour plot are shown for the c
first-order and d second-order RWs described by Eqs. (4.21) and (4.22) respectively. The solution
parameters used in these plots are β00 = 0 and X 10 = −100. Reprint from Ref. [25], Copyright
2022, with permission from American Physical Society
118 4 Emission of Rogue Wave Signals in Nonlinear Electrical Transmission Networks

Fig. 4.4 The dynamics of the first-order RW |ψ(x, τ )| of the network of Fig. 2.1 described by
Eq. (4.21) for an exponentially decaying or growing linear density profile, χ(τ ) = χ0 exp (χ1 τ )
with k = 0.75 rad/cell and the network parameters (4.17). a χ0 = 1, χ1 = −P < 0; b χ0 = 1,
χ1 = P > 0; c χ0 = −1, χ1 = −P < 0; d χ0 = −1, χ1 = P > 0. The solution parameters used
in these plots are β00 = 0 and X 10 = 0. Reprint from Ref. [25], Copyright 2022, with permission
from American Physical Society

the first-order rogue wave is reduced to a dark-bright-dark soliton. Plots of Fig. 4.5
show a second-order RW with four peaks. Plots of Figs. 4.4 and 4.5 also show how
much the solution parameters χ0 and χ1 affect the evolution of the modulated RWs
propagating through the network of Fig. 2.1. Although these two parameters have
no effect on the wave shape, they affect the wave behavior. Plots of Figs. 4.4 and 4.5
reveal that modulated RWs of nonlinear electric transmission networks may either
take birth from dark [bright] solitonlike waves and “disappear without a trace”, or
“appear from nowhere” and disappear as dark [bright] solitonlike waves [4].

4.1.4 Emission of Rogue Wave Signals Through the Network


of Fig. 2.1

Based on the exact rational polynomial solutions (4.21) and (4.22) of the amplitude
equation (4.8), we turn to the emission of RW signals through the network of Fig. 2.1.
For this purpose, it is convenient to go back to the original variables n and t and use
expansion (4.5):
4.1 Emission of Rogue Wave Signals Through the Modified … 119

Fig. 4.5 The dynamics of the second-order RE |ψ(x, τ )| of the network of Fig. 2.1 described by
Eq. (4.22) for an exponentially decaying or growing linear density profile, χ(τ ) = χ0 exp (χ1 τ )
with k = 0.75 rad/cell and the network parameters (4.17). a χ0 = 1, χ1 = −P < 0; b χ0 = 1,
χ1 = P > 0; c χ0 = −1, χ1 = −P < 0; d χ0 = −1, χ1 = P > 0. The solution parameters used
in these plots are β00 = 0 and X 10 = 0. Reprint from Ref. [25], Copyright 2022, with permission
from American Physical Society

 
Vn (t) = 2εRe ψ(x, τ ) exp [i (kn − ωt)] + ε {ψ10 (x, τ ) + ψ20 (x, τ ) exp [2i (kn − ωt)]} ,
(4.24)
where x = ε(n − υg t), τ = ε2 t, and ψ(x, τ ) = ψ1,2 (ξ(x, τ ), τ ). For the simplicity,
we focus ourselves to the case when the linear potential is absent in Eq. (4.8), that
is, χ (t) = 0; in this case, we take β0 = 0, γ0 = 0, and X 1 = 0 so that ξ = x. In the
following, we aim to address effects of various parameters such as the wavenumber
k, the linear dispersive element C S , the linear capacitance C0 , as well as the linear
inductances L 1 and L 2 on the evolution of first- and second-order modulated RWs
propagating along our network system. For the computational purpose, we use the
numerical value of ε = 10−3 .

4.1.4.1 Effects of Network and Equation Parameters on the Waveform

To study the effect of the wavenumber k on the modulated RWs propagating along our
network system, we show, for different values of the wavenumber k, the waveform
at time t = 0 of the first- and second-order RWs in Fig. 4.6. It is seen from different
plots of Fig. 4.6 that the amplitudes and the widths of both the RW and its background
decreases with the increase in the values of the wavenumber k. Since the propagating
120 4 Emission of Rogue Wave Signals in Nonlinear Electrical Transmission Networks

Fig. 4.6 The waveforms of the first-order (a–f) and the second-order (g–l) RW voltage with differ-
ent values of the wavenumber k for C S = 0.9999 (L 2 /L 1 ) C0 and other network parameters given
by Eq. (4.17). a, g k = 0.35 rad/cell; b, h k = 0.365 rad/ cell; c, i k = 0.38 rad/cell; d, j k = 2.85
rad/cell; e, k k = 2.95 rad/cell; f, l k = 3.10 rad/cell. Reprint from Ref. [25], Copyright 2022, with
permission from American Physical Society

frequency f p is an increasing function of the wavenumber k, we conclude that the


wave amplitude and the wave width decrease as the propagating frequency increases.
To observe how much the linear dispersive parameter C S and the capacitance
scale C0 affect the motion of modulated RWs during their propagation along our
network system, we have displayed in respectively Figs. 4.7 and 4.8 the waveforms
of the first-order (top panels) and the second-order (bottom panels) RW voltages, for
different numerical values of C S and C0 . Plots of Fig. 4.7 (Fig. 4.8) show that the
wave amplitude and the wave width decreases (increases) with the increase in the
values of parameter C S (C0 ). This means that linear capacitances C S and C0 have
opposite effects on the wave amplitude and width.
Lastly, we investigate the effects of the linear inductances L 1 and L 2 on the mod-
ulated RWs moving along the network of Fig. 2.1. For this aim, we have, for different
values of L 2 and L 1 , displayed in respectively Figs. 4.9 and 4.10 the waveforms of
the first-order (top panels) and second-order (bottom panels) modulated RWs propa-
gating along our network system. Plots of Fig. 4.9 (Fig. 4.10) reveal that any increase
in the values of L 2 (of L 1 ) leads to an increase (decrease) in the wave amplitude.
Observing different plots of Figs. 4.9 and 4.10, it is clearly seen that the linear induc-
tances L 1 and L 2 have no effect on the wave amplitude. Therefore, L 1 and L 2 have
opposite effects on the wave width.
4.1 Emission of Rogue Wave Signals Through the Modified … 121

Fig. 4.7 The waveforms of the first-order (top panels) and second-order (bottom panels) RW
voltage for different values of the dispersive element C S for the network parameters (4.17)
with k = 0.365 rad/cell. a, d C S = 0.98 (L 2 /L 1 ) C0 , b, e C S = 0.99 (L 2 /L 1 ) C0 , c, f: C S =
0.999(L 2 /L 1 )C0 . Reprint from Ref. [25], Copyright 2022, with permission from American
Physical Society

Fig. 4.8 The waveforms of the first-order (top panels) and second-order (bottom panels) RW
voltage with different values of the linear capacitance C0 for the network parameters (4.17)
with C S = 0.999 (L 2 /L 1 ) C0 and k = 0.365 rad/cell. a, d C0 = 370 pF, b, e C0 = 373.7 pF,
c, f C0 = 377.4 pF. Reprint from Ref. [25], Copyright 2022, with permission from American
Physical Society
122 4 Emission of Rogue Wave Signals in Nonlinear Electrical Transmission Networks

Fig. 4.9 The waveforms of the first-order (top panels) and second-order (bottom panels) RW
voltage with different values of the linear inductance L 2 for the network parameters (4.17) with
C S = 0.98 (L 2 /L 1 ) C0 and k = 3.1 rad/cell. a, d L 2 = 470 mH; b, e L 2 = 517 mH; c, f L 2 = 564
mH. Reprint from Ref. [25], Copyright 2022, with permission from American Physical Society

Fig. 4.10 The waveforms of the first-order (top panels) and second-order (bottom panels) RW
voltage with different values of the inductance of linear element L 1 , for network parameters (4.17)
with C S = 0.98 (L 2 /L 1 ) C0 and k = 3.1 rad/cell. a, d L 1 = 176 mH; b, e L 1 = 198 mH; c, f
L 1 = 220 mH. Reprint from Ref. [25], Copyright 2022, with permission from American Physical
Society

4.1.4.2 Analytical Emission of RWs Voltages Via Exact RW Solutions


(4.21) and (4.22)

Using now the exact rational polynomial solutions (4.21) and (4.22) of the ampli-
tude equation (4.8), we can experience the emission electrical modulated RWs in
the network of Fig. 2.1. The typical spatial and temporal evolution of the first- and
second-order modulated RWs are showed in respectively Figs. 4.11 and 4.12. Plots
of Fig. 4.11 (Fig. 4.12) show the spatial [temporal] evolution of the first-order (top
panels) and second-order (bottom panels) modulated RWs propagating at frequency
f p = 382.074 kHz along the network system of Fig. 2.1. Different plots reveal that
4.1 Emission of Rogue Wave Signals Through the Modified … 123

Fig. 4.11 Voltage Vn as a function of the cell number n, at given times (t0 , t1 , t2 , and t3 ), showing
the propagation of the first-order (top panels) and second-order (bottom panels) RWs in the network
at frequency f p = 930.524 kHz for parameters (4.17). a, e t0 = 0; b, f t1 = 0.033; c, g t2 = 0.066;
d, h t3 = 0.1. Reprint from Ref. [25], Copyright 2022, with permission from American Physical
Society

Fig. 4.12 The propagation of the first-order (top panels) and second-order (bottom panels) RWs in
the network at frequency f p = 930.524 kHz for parameters (4.17). a, e n = 0; b, f n = 1000; c, g
n = 2000; d, h n = 3000. Reprint from Ref. [25], Copyright 2022, with permission from American
Physical Society

second-order [first-order] modulated RWs can be emitted without essential distor-


tions [are emitted with distortions].

4.1.4.3 Effect of the Carrier Phase on the RW Shape

Although the carrier phase ϕ(x, τ ) = β0 (τ )x + γ0 (τ ) + Qτ has no affect on the


wave amplitude, it can affect the shape of the RW peaks and holes. For a better
understanding, we consider a carrier phase with a nonzero parameter β0 (t). Integrat-
ing Eq. (4.19) yields
τ
β0 (τ ) = χ (s)ds + β0 (0).
0
124 4 Emission of Rogue Wave Signals in Nonlinear Electrical Transmission Networks

Therefore, studying the effect of β0 (τ ) on the motion of the modulated RWs through
the network of Fig. 2.1 is tantamount to studying the effect of parameter χ (t) on
the RWs. On the basis of Figs. 4.6, 4.7, 4.8, 4.9, 4.10, 4.11 and 4.12, which pertain
to different values of k, can be interpreted as the effect of the carrier phase on the
modulated RWs propagating along our the network system. To investigate the effect
of the carrier phase originating from parameter χ (t) only, one may vary χ , keeping
all the network parameters and wavenumber k constant.

4.1.5 Conclusion and Discussions

We have considered in this Section the lossless one-dimensional modified Noguchi


nonlinear NLTN with dispersive elements, displayed in Fig. 2.1. The dynamics of
nonlinear modulated wave of this network are reduced to a NLS equation with an
external linear potential. The criterion of the MI is presented and exact first- and
second-order rational polynomial solutions of the amplitude equation are built. Using
these solutions, we have, for various physical linear potential, studied the dynamics of
modulated RWs propagating through our network system. The effects of the network
parameters as well as those of the solution parameters and the carrier wavenumber
are investigated too. We found that the linear capacitances C S and C0 and linear
inductances L 1 and L 2 have opposite effects on the wave amplitude/width. More
interestingly, we found that suitably choosing the linear capacitances C S and C0
and the propagating frequency f p may lead to either the wave amplification or the
wave suppression. We have also found that the propagating frequency f p , the linear
dispersive element C S , the linear inductance L 1 , the linear capacitance C0 , and the
linear inductance L 2 can be chosen adequately for the wave compression.
Our studies also revealed that modulated RWs in electric nonlinear transmission
network can either appear from nowhere and disappear without a trace, or take
birth from dark/bright solitonlike waves and disappear without a trace, or appear
from nowhere and disappear as a dark/bright solitonlike waves. The results found in
the present Section may be useful to realize the RW dynamics in various physical
systems, in addition to NLTNs, such as nonlinear fibers, BECs, plasmas, etc.

4.2 Generation of Network Modulated Rogue Waves Under


the Action of the Quintic Nonlinearity

The present Section deals with the transmission of non-autonomous rogue waves
along the network system with dispersive elements illustrated in Fig. 2.1. Here, we
use the same capacitance-voltage dependence (4.1) as in the previous Section. We
show that the amplitude equation is reduced to a cubic-quintic NLS equation with
cubic derivative terms. One-parameter first-order rational polynomial solutions of
4.2 Generation of Network Modulated Rogue Waves Under … 125

this equation are presented and used to investigate analytically the generation of first-
order non-autonomous modulated RWs propagating in the network system of Fig. 2.1.
We study the effects of the quintic nonlinearity and nonlinear-dispersion parameters
on the first-order non-autonomous modulated RWs during their transmission through
our network system.

4.2.1 The Kundu–Eckhaus Model for Non-autonomous


Modulated Rogue Waves in a Lossless Electric Network

Applying the Kirchhoff’s laws to the network in Fig. 2.1 yields the system of non-
linear discrete equations (4.2). Asking to the linear waves propagating along the
network of Fig. 2.1 to be proportional to the expression exp [i (kn − ωt)] yields the
linear dispersion relation (4.4) of a typical passband filter. In the following, we use
the following network parameter for the computational purpose [22, 25]:

L 1 = 28µH, L 2 = 14µH, C0 = 540 pF,C S ∈ [0, C0 /2[pF,α ≤ 0.21 V−1 ,

β ≤ 0.0197 V−2 , V0 = 1.5 V. (4.25)

In order to carry out efficient investigation of the dynamics of non-autonomous


modulated RWs propagating along the network of Fig. 2.1, it is reasonable to first
derive the amplitude equation. To accomplish this, we employ the reductive pertur-
bation method in the semidiscrete approximation, considering the interaction of any
wave packet centered at (ω, k), where ω = 2π f is the carrier frequency and k is the
carrier wavenumber. Following Marquie et al. [25], we introduce the following slow
variables
ξ = ε(n − υg t) and τ = ε2 t, (4.26)

where ε  1 is a small parameter,


 2 
dω u − λω02 sin k
υg = =  0 2 (4.27)
dk ω 1 + 4λ sin2 (k/2)

is the group velocity, and the frequency ω is related to the wavenumber k via the
following linear dispersion relation
 
ω02 + 4u 20 sin2 k2
ω2 =   . (4.28)
1 + 4λ sin2 2k

Next, we seek for the solution of the difference-differential system (4.2) in the general
form [25]
126 4 Emission of Rogue Wave Signals in Nonlinear Electrical Transmission Networks
 
Vn (t) = εψ(ξ, τ ) exp [iθ − i(ξ, τ )] + ε2 V02 (ξ, τ ) + V12 (ξ, τ ) exp (2iθ ) + c.c.,
(4.29)
where c.c. stands for the complex conjugation, θ = kn − ωt is the rapidly varying
phase, (ξ, τ ) is an additional imprinted phase, and ψ, V02 , and V12 are respectively
the fundamental, dc, and second harmonics.
Under all the above assumptions, we find that

2α 2 ωυg2
V02 (ξ, τ ) =   |ψ(ξ, τ )|2 , (4.30a)
α υg2 − u 20
4αω2  2
V12 (ξ, τ ) =   2 ψ(ξ, τ ) exp [−i(ξ, τ )] ,
4ω − ω0 + 4 4λω − u 0 sin k
2 2 2 2

(4.30b)
∂
= β0 |ψ(ξ, τ )|2 , (4.30c)
∂ξ
 
∂ ∂ψ ∗ ∂ψ
= −iβ0 P ψ − ψ∗ + 2β02 P |ψ(ξ, τ )|4 (4.30d)
∂τ ∂ξ ∂ξ
 
∂ψ ∂ 2ψ ∂ |ψ|2
i + P 2 + Q |ψ|2 ψ + β02 P |ψ|4 ψ − 2iβ0 P ψ = 0, (4.30e)
∂τ ∂ξ ∂ξ

where β0 is an arbitrary real constant, P and Q are real parameters, function of the
wavenumber k, defined as


   2 
υg2 2 k u0 λ
P=− 1 + 4λ sin + − ω cos k − 2λυg sin k, (4.31)
2ω 2 2ω 2
3β 2α 2
ωυ 2
g 4α ω
2 3
Q= ω+ 2 −   . (4.32)
2 u 0 − υg2 4ω2 − ω02 + 4 4λω2 − u 20 sin2 k

Letting   = Q/P, and ψ


τ = Pτ, Q (ξ, 
τ ) = ψ(ξ, τ ), Eq. (4.30e) takes the simpler
form,
  2 
 ∂ 2ψ
∂ψ   2  4 
∂ ψ
i + +Q  ψ
 ψ  ψ
 + β02 ψ  − 2iβ0 =0
ψ (4.33)
∂
τ ∂ξ 2 ∂ξ

and contains only two real parameters, Q and β0 .


The cubic-quintic derivative NSL equation (4.30e), or its simpler form (4.33)
is known as an equation of Kundu–Eckhaus (KE) [32–37] type, and, when β0 =
0, coincides with the standard cubic NLS equation. It governs the dynamics of
modulated waves in the network of Fig. 2.1. In Eq. (4.30e), parameter β02 = 0 is
the relative magnitude of the quintic nonlinearity, while the last term represents the
nonlinear dispersion, with relative magnitude β0 . Using now the phase-engineering
transformation (PIT) [38]
4.2 Generation of Network Modulated Rogue Waves Under … 127

∂ϕ 4Pβ0 + 2 2
ψ = q exp (iϕ) , = |q| ,
∂ξ 3P
 
∂q 1 + 2Pβ0 ∂q ∗ ∂q 4 (1 + 2Pβ0 )2 4
= −i q − q∗ − |q| + Q |q|2
∂τ 3 ∂ξ ∂ξ 9P

and results from [39], it is seen that the KE equation (4.30e) conserves the integral
power,
+∞
P= |ψ(ξ, τ )|2 dξ.
−∞

The amplitude equation (4.30e) admits the CW solution


   
ψ0 (ξ, τ ) = q0 exp ik0 ξ − i k02 P − Qq02 − β02 Pq04 τ

with the real constant amplitude q0 and wavenumber k0 (of the carrier wave). Carrying
out the linear stability analysis and denoting by K and  the wavenumber and
complex angular frequency of modulation, we arrive to the following MI growth rate

 02 − K 2 ,
|Im ()| = |P K | 2 Qq (4.34)

 = Q/P > 0. Therefore, the uniform wave propagating along the lossless net-
if Q
work of Fig. 2.1 is stable against the modulation if P Q < 0, and unstable under
modulation if P Q > 0.

4.2.2 Generation of First-Order Non-autonomous Modulated


Rogue Waves (Alias Peregrine Solitons) for a Lossless
Electric Network

 = Q/P > 0, exact first-order ratio-


Here, we first present, under the condition Q
nal polynomial solutions of the amplitude equation (4.30e) which are then used
for investigating the generation and management of the first-order non-autonomous
modulated RWs (alias Peregrine solitons [40, 41]) in the lossless electric network of
Fig. 2.1.
128 4 Emission of Rogue Wave Signals in Nonlinear Electrical Transmission Networks

4.2.2.1 One-Parameter First-Order Rational Polynomial Solutions of


the Amplitude Equation

To obtain exact first-order rational polynomial solutions of the amplitude equation


(4.30e), we introduce the ansatz

ψ(ξ, τ ) = r0 u(X, T ) exp [i (ξ − Pτ )] , X = r0 (ξ − 2Pτ ) , T = r02 τ, (4.35)

under which the KE equation (4.30e) take a similar form,

∂u ∂ 2u ∂ |u|2
i +P + Q |u| 2
u + r β
2 2
P |u| 4
u − 2ir 0 β0 Pu = 0, (4.36)
∂T ∂ X2 0 0
∂X
which admits the special CW solution
   
u 1 (X, T ; q1 , k1 ) = q1 exp ik1 X − i k12 P − Qq12 − r02 β02 Pq14 T ;

here, r0 = 0 is an arbitrary real constant, referred to as the “solution parameter”,


and q1 and k1 are real constants. As in the case of the KE equation (4.30e), equation
(4.36) conserves the integral power:

+∞ +∞
−1
P= |u(X, T )| d X = |r0 |
2
|ψ(ξ, τ )|2 dξ.
−∞ −∞

Using the CW solution u 1 (X, T ; 1, 0) as the seed solution and following the method
elaborated in Refs. [41, 42], we arrive to the following one parameter first-order ratio-
nal polynomial solution (with parameter r0 = 0) of the amplitude equation (4.30e):
⎡ ⎤
⎢   ⎥
⎢ 4 1 + 2i Qr02 τ ⎥
ψ(ξ, τ ) = r0 ⎢−1 +   2 ⎥
⎣   −1/2 3 ⎦
4 2 2 Q 
1 + 4r0 Q τ + 4 1 + r0 2 (ξ − 2Pτ ) + 2Q 2 Q r 0 β0 τ

   
1 − r04 β02
× exp i ξ + r02 P −
Q τ (4.37)
r02

⎫⎤
   ⎪


 −1/2 Qτ ⎪
⎬⎥
8r0 β0 1 + r0 Q2 (ξ − 2Pτ ) + 2r03 β0 2 Q ⎥
+   2 ⎪⎥ .
2Q   −1/2 ⎪⎦
Q
1 + 4r0 Q τ + 4 1 + r0 2 (ξ − 2Pτ ) + 2Qr0 β0 2 Q
4 2 2 3  τ ⎭ ⎪
4.2 Generation of Network Modulated Rogue Waves Under … 129

Next, we employ the first-order rational polynomial solution (4.37) to evaluate the
wave phase:
 
1 − r04 β02 ϕ N (X, T )
ϕ(ξ, τ ) = ξ + r02Q− P τ+
2
r0 ϕ D (X, T )

2 (G − R)
+ arccos  , (4.38)
4 (G − R)2 + Q 2 T 2 H 2

with
X = r0 (ξ − 2Pτ ) , T = r02 τ.

4.2.2.2 Results and Discussions

Since
lim ψ(ξ, τ ) = r0 ,
|ξ |2 +|τ |2 →∞

we conclude that the one-parameter Peregrine soliton solution (4.37) of the KE equa-
tion (4.30e) has the form of a single-peak structure that merges into the plane wave
at |ξ | + |τ | → +∞. It is evident that various parameters such as the quintic non-
linearity and nonlinear-dispersion parameter β0 , solution parameter r0 , the carrier
wavenumber k, coefficients P and Q of the amplitude equation (4.30e), as well as
on the network parameters impact the transmission of first-order modulated wave
through the network system of Fig. 2.1. These parameters determine the wave char-
acteristics. As well as we know, the ideal initial perturbation that corresponds to
the exact Peregrine soliton solution of the amplitude equation will evolve into a
modulated RWs [43, 44].
It follows from solution (4.37) that the maximal value of the wave intensity is
reached at τ = 0 :
|ψ(ξ, τ )|max = |ψ(ξ, 0)|max = 3 |r0 | .

This means that the wave amplitude and that of its CW background increase with
the increase as |r0 | increases; moreover, |ψ(ξ, τ )|max is independent of the network
system’s parameters.
Effects of parameter r0 and β0 on the Peregrine solitons propagating through the
lossless network Now, we use the exact rational polynomial solution (4.37) to study
the effects of parameters r0 and β0 on the first-order modulated rogue waves propa-
gating along the network system of Fig. 2.1, we have displayed in respectively Figs.
4.13 and 4.14 intensity |ψ| for different values of respectively r0 and β0 . Smoothly
increasing the numerical values of r0 from an optimal value, we can visualize, as
shown plots of Fig. 4.13, the formation of crests and troughs with an increase in
the amplitude wave, large amplitude wave corresponding to large values of r0 . Plots
of Fig. 4.14 reveal that the sign and the magnitude of parameter β0 as well as the
wavenumber k affect the wave speed and have no effect on the wave amplitude. Plots
130 4 Emission of Rogue Wave Signals in Nonlinear Electrical Transmission Networks

Fig. 4.13 Plots of amplitude |ψ(ξ, τ )| produced by rational solution (4.37), showing the formation
of the first-order non-autonomous RW in the lossless nonlinear electric network of Fig. 2.1 with
parameters (4.25) for α = 0.21 V−1 and β = 0.0197 V−2 . The parameters are λ = C S /C0 = 0.3,
k = 0.9, β0 = 0.2 and a r0 = 0.3; b r0 = 0.5; c r0 = 0.7; and d r0 = 1.2. Reprint from Ref. [22],
Copyright 2022, with permission from American Physical Society

of Fig. 4.14 also reveal that the carrier wavenumber k has no effect on the wave
shape during the wave propagation through our network system; however, the wave
shape may be affected by the variation of parameter β0 . This fact is clearly seen
from plots of Figs. 4.15 and 4.16 showing the wave evolution for different values
of parameter β0 . When β0 is taken around some critical value β0c , the shape of the
RW remains unchange during the wave propagation, as one can see from plots of
Fig. 4.15. It is seen from plots of Fig. 4.16 that when parameter β0 increases from
some critical value β0c of parameter β0 (the same results are obtained for β0 < β0c ),
Peregrine solitons get more delocalized in space. Therefore, parameter β0 can be
used to control the Peregrine solitons of the network of Fig. 2.1.
Effects of the dispersive element C S on the Peregrine solitons To show the effects
of the network parameter C S on the network Peregrine solitons, we depict in Figs.
4.17 and 4.18 the evolution of Peregrine solitons for different values of the dimen-
sionless dispersive parameter λ = C S /C0 and respectively k = 0.3 and k = 0.9. We
can clearly see from plots of Figs. 4.17 and 4.18 how much the dispersive element
C S through parameter λ affects both the wave’s speed and the localization of its
main peak. Also, plots of Figs. 4.17 and 4.18 show that, depending on the value of
the carrier wavenumber k, the wave speed increases [decreases] and the Peregrine
4.2 Generation of Network Modulated Rogue Waves Under … 131

Fig. 4.14 Density plots of the first-order non-autonomous RWs associated with rational solution
(4.37) for r0 = 1.2 and λ = C S /C0 = 0.3. Relative magnitude β0 of the nonlinear dispersion takes
the following values: a, d β0 = −0.3; b, e β0 = 0.7; c, f β0 = 1.2. Plots (a)–(c) are generated with
wavenumber k = 0.9, while plots (d)–(f) are obtained with k = 0.8. Other parameters are shown
in the text. It is seen from these plots that the density distribution is not affected by the variation
of β0 , which, however, affects the RW’s velocity. Reprint from Ref. [22], Copyright 2022, with
permission from American Physical Society
132 4 Emission of Rogue Wave Signals in Nonlinear Electrical Transmission Networks

Fig. 4.15 The dynamics of the first-order non-autonomous RW |ψ(ξ, τ )|, associated with rational
solution (4.37), with the same parameters as in Fig. 4.14a–c. Values of parameter β0 are a β0 = −0.3;
b β0 = 0.7; c β0 = 1.2. The RWs shown in plots (a)–(c) have the same maximum value of the main
peak, |ψ(ξ, τ )|max = 3r0 . Reprint from Ref. [22], Copyright 2022, with permission from American
Physical Society

soliton gets more [less] localized with the increase in the values of λ as shown in
Fig. 4.17 [Fig. 4.18]. For better understanding, we display in Fig. 4.19 the amplitude
profile |ψ(ξ, τ )| at time τ = 0 for two values of the carrier wavenumber and different
values of parameter λ.
Effects of nonlinear parameters α and β and wavenumber k on the Peregrine
solitons For different values of α, β, and k, we depict respectively in Figs. 4.20, 4.21,
and 4.22 the magnitude of amplitude profile |ψ(ξ, 0)| at time τ = 0 for different
values of respectively α, β, and k. It follows from plots of Figs. 4.20, 4.21, and 4.22
that the wave speed decreases [increases] and the Peregrine soliton gets less√[more]
localized with the increase of each α, β and k taken from the region where P/2Q
is an increasing [a decreasing] function of the corresponding α, β and k. Comparing
plots of Fig. 4.19 to those of Figs. 4.20, 4.21, and 4.22, we conclude that parameter
λ and parameters α, β, and k produce opposite effects, while parameters α, β, and k
have the same effect on the Peregrine solitons. Therefore, these four parameters are
useful for controlling the motion of Peregrine solitons along the electrical network
of Fig. 2.1.
4.2 Generation of Network Modulated Rogue Waves Under … 133

Fig. 4.16 Plots of evolution (top) and density (bottom) of the first-order non-autonomous RWs
|ψ(ξ, τ )| produced by rational solution (4.37) for network parameters (4.25). The constants are
λ = C S /C0 = 0.3 and r0 = 1.2. Parameter β0 of the nonlinear dispersion takes values a, d β0 = 0.7;
b, e β0 = 3; c, f β0 = 9. Other parameters are given in the text. Reprint from Ref. [22], Copyright
2022, with permission from American Physical Society

Effects parameters β0 , λ, α, and β on the phase profile of the Peregrine solitons


Although the wave phase does not affect the wave amplitude, the steepness of the
first-order modulated RWs is strongly dependent on the wave phase profile [45].
For better understanding, the phase evolution of the Peregrine soliton is displayed
in Fig. 4.23 for different values of β0 and in Fig. 4.24 for different values of λ (Fig.
4.24 a–c), α (Fig. 4.24d–f) and β (Fig. 4.24g–i). We can see from Fig. 4.23 that the
symmetric shape of the Peregrine soliton is broken by parameter β0 and the phase
distributions are almost opposite for β0 and −β0 . Plots of Fig. 4.24 reveal that the
symmetry of the Peregrine soliton is maintained under the variation of parameter λ,
134 4 Emission of Rogue Wave Signals in Nonlinear Electrical Transmission Networks

Fig. 4.17 The density (top) and evolution (bottom) plots of the first-order non-autonomous RWs
corresponding to solution (4.37) for network parameters (4.25) with α = 0.16 V−1 and wavenumber
k = 0.3 of the voltage signal for different values of dispersive parameter λ = C S /C0 , namely, a, d
λ = 0.46; b, e λ = 0.47; c, f λ = 0.49. Other parameters are given in the text. Reprint from Ref.
[22], Copyright 2022, with permission from American Physical Society
4.2 Generation of Network Modulated Rogue Waves Under … 135

Fig. 4.18 First-order non-autonomous RWs associated with solution (4.37) for network parameters
(4.25) with α = 0.21 V−1 and wavenumber k = 0.9 of the voltage signal for different dispersion
parameter λ = C S /C0 , with values a, d λ = 0.27; b, e λ = 0.33; c, f λ = 0.40. Other parameters
are shown in the text. Reprint from Ref. [22], Copyright 2022, with permission from American
Physical Society
136 4 Emission of Rogue Wave Signals in Nonlinear Electrical Transmission Networks

Fig. 4.19 Magnitude of the rational solution, |ψ(ξ, 0)|, at time τ = 0, as given by Eq. (4.37) with
r0 = 1.2 and β0 = 0.695 for network parameters (4.25) with β = 0.0197. At τ = 0, the amplitude
of the rational solution (4.37) is real and its maximum value is |ψ(ξ, 0)|max = 3 |r0 |. Plot (a) shows

P/2Q as a function of parameter λ = C S /C0 of the linear dispersive element C S . Plots (b)
are generated with α = 0.16 , k = 0.3 and different values of λ, viz., λ = 0.46 [(the solid line),
λ = 0.47 (the dashed line), and λ = 0.49 (the dotted line). Plots (c) are generated for α = 0.21,
k = 0.9 and λ = 0.27 (the solid line), λ = 0.33 (the dashed line), and λ = 0.40 (the dotted line).
Other parameters are the same as in Figs. 4.17 and 4.18. Plots (b) and (c) are produced for values of
λ taken from Region (I) and Region (II), respectively. It is seen that λ affects both the wave’s shape
and velocity. Reprint from Ref. [22], Copyright 2022, with permission from American Physical
Society

α, and β. These three parameters just causes a shift of the phase evolution, as we can
see from plots of Fig. 4.25 showing the phase profile at time τ = 0 for different values
of λ, α, or β. It follows from different plots of Figs. 4.24 and 4.25 that parameter λ
and parameter β produce the same effect on the wave phase, while λ and α, and α
and β demonstrate opposite effects.

4.2.3 Conclusion and Discussions

In this Section, we have considered the electric nonlinear transmission network of


Fig. 2.1 and reduced the amplitude equation to a one-parameter KE equation. The
criterion of the MI of Stokes waves is obtained and a one-parameter first-order rational
4.2 Generation of Network Modulated Rogue Waves Under … 137

Fig. 4.20 The absolute value of the rational solution, |ψ(ξ, 0)|, at time τ = 0 as given by Eq. (4.37)
with r0 = 1.2 and β0 = 0.7 for √ network parameters (4.25) with λ = C S /C0 = 0.29 and α = 0.16.
Plot (a) shows the variation of P/2Q as a function of nonlinearity parameter α. Wavenumber k
and α are varied as follows. b k = 0.8 and α = 0.1 (the solid line), α = 0.15 (the dashed line), and
α = 0.18 (the dotted line); c k = 0.9 and α = 0.196 (the solid line), α = 0.2 (the dashed line), and
α = 0.21 (the dotted lines). Plots (b) and (c) are generated with nonlinearity parameter β taken from
Region (I) and Region (II), respectively. Reprint from Ref. [22], Copyright 2022, with permission
from American Physical Society

polynomial solution is built. Using this solution, we have investigated analytically


the dynamics and the transmission of Peregrine soliton through our network system.
The effects of various parameters including the network parameters on the dynamics
of the Peregrine soliton of our network system are also investigated. We found that
the built rational polynomial solution produces large-amplitude waves, localized in
both space and time, predicting that the wave energy may be concentrated into a
small region. We also found that the solution parameter can be used either to amplify
or to suppress modulated RWs in the network of Fig. 2.1. Our results also reveal
that the network parameters are useful for controlling the motion of the Peregrine
solitons in our network system.
138 4 Emission of Rogue Wave Signals in Nonlinear Electrical Transmission Networks

Fig. 4.21 The absolute value of the rational solution, |ψ(ξ, 0)|, at time τ = 0 as given by Eq. (4.37)
with r0 = 1.2 and β0 = 0.7 for√network parameters (4.25) with λ = C S /C0 = 0.3 and α = 0.16.
Plot (a) shows the variation of P/2Q as a function of the nonlinearity parameter β. Wavenumber
k and parameter β are varied as follows. b k = 2.0 and β = 0 (the solid line), β = 0.005 (the dashed
line), and β = 0.01 (the dotted lines); c k = 0.8 and β = 0.015 (the solid line), β = 0.017 (dashed
line), and β = 0.019 (the dotted lines). Plots (b) and (c) are generated with nonlinear parameter β
taken from Region (I) and Region (II), respectively. Reprint from Ref. [22], Copyright 2022, with
permission from American Physical Society

4.3 Chirped Super Rogue Waves Propagating Along a


Lossless Nonlinear Electrical Network

This Section focuses on the investigation of the chirped version of higher-order


rogue waves with nonlinear chirping, termed super chirped rogue waves (SRWs)
for their super high peak amplitude, within the framework of the nonlinear electric
transmission network shown in Fig. 2.1. Such an investigation is carried out with the
help of a generalized cubic-quintic NLS equation with a self-steepening term [46–
48]. The effects of the network parameters such as the linear dispersive parameter
on the chirped SRWs will be analyzed.
4.3 Chirped Super Rogue Waves Propagating … 139

Fig. 4.22 The absolute value of rational solution, |ψ(ξ, 0)|, at time τ = 0, as given by Eq. (4.37)
with r0 = 1.2 and β0 = 0.6 for network parameters
√ (4.25) with λ = C S /C0 = 0.3, α = 0.21, and
β = 0.0197. Panel (a) shows the variation of P/2Q as a function of wavenumber k. It is varied
as follows. b k = 1.0 (the solid line), k = 1.05 (the dashed line), and k = 1.1 (the dotted lines); c
k = 2.1 (the solid line), k = 2.3 (the dashed line), and k = 3.0 (the dotted lines). Plots (b) and (c)
are generated with k taken from Region (I) and Region (II), respectively. Reprint from Ref. [22],
Copyright 2022, with permission from American Physical Society

4.3.1 Generalized Nonlinear Schrödinger Equation


for a Lossless Electric Network

The model to be studied in the Section is the lossless electric transmission network
depicted in Fig. 2.1 made of N identical cells. For this system, we use the same
voltage-capacitance
 dependence
 as in the previous two Sections, that is, Q n (Vn ) =
C0 Vn − αVn2 + βVn3 , where C0 is the characteristic capacitance, and α and β are
two nonlinear coefficients which are assumed to be positive. Applying Kirchhoff’s
laws to the network system of Fig. 2.1, we arrive to the following network equations:
 
d 2 Vn d2 d2  2 
+ u 2
0 + λ (2V n − Vn+1 − Vn−1 ) + ω0
2
Vn = αVn − βV n
3
,
dt 2 dt 2 dt 2
√ √ (4.39)
where n = 1, 2, . . . , N , u 0 = 1/ L 1 C0 and ω0 = 1/ L 2 C0 are the characteristic
frequencies of the network, and λ = CC0S is the dimensionless dispersive parameter.
140 4 Emission of Rogue Wave Signals in Nonlinear Electrical Transmission Networks

Fig. 4.23 The evolution of the RW’s phase according to Eq. (4.38) for network parameters (4.25)
with α = 0.21, β = 0.0197, λ = C S /C0 = 0.3, r0 = 1.2, and different nonlinear dispersion param-
eter β0 . a The evolution of the first-order non-autonomous RW’s phase with β0 = 0. b The evolution
of the first-order non-autonomous RW’s phase with β0 = −0.7. c The evolution of the first-order
non-autonomous RW’s phase with β0 = 0.7. d The phase curve of the first-order non-autonomous
RW at time t = 0 for different values of parameter β0 , viz., β0 = 0 (the solid line), β0 = −0.7
(the dashed line), and β0 = 0.7 (the dotted line). Different plots are generated with wavenumber
k = 0.9 of the voltage signal propagating in the network. It is seen that the high-order effects break
the symmetry of the first-order non-autonomous RW. Reprint from Ref. [22], Copyright 2022, with
permission from American Physical Society

In the linear approximation Vn ∼ exp [i (kn − ωt)], Eq. (4.39) leads to the fol-
lowing linear dispersion relation of a typical band-pass filter

 

k k
ω2 1 + 4λ sin2 − ω02 + 4u 20 sin2 = 0. (4.40)
2 2

For values of wavenumber k chosen in the first Brilloum zone (0 ≤ k ≤ π ), the


angular frequency ω defined
 by Eq. (4.40) is limited by the cutoff angular frequencies
ω2 +4u 2
ω0 at k = 0 and ωc = 0
1+4λ
0
at k = π : ω0 ≤ ω ≤ ωc . Therefore, the propagating
frequency f p will be limited by the lower cut-off frequency f 0 = ω0 /2π and the
upper cut-off frequency f c = ωc /2π .
4.3 Chirped Super Rogue Waves Propagating … 141

Fig. 4.24 The evolution of the RW’s phase according to Eq. (4.38) for network parameters (4.25),
with r0 = 1.2 and β0 = 0.7. a–c The evolution of the phase profile with different values of linear
dispersive parameter λ = C S /C0 . viz., a λ = 0.3, b λ = 0.31, and c λ = 0.32. d–f The evolution of
the phase profile for different values of the nonlinear parameter α, viz., d α = 0.205, e α = 0.2075,
and f α = 0.21. g–i The evolution of the phase profile for different values of the nonlinearity
parameter β, viz., g β = 0.0157, h β = 0.0177, i β = 0.0197. Different plots are generated with
wavenumber k = 0.9 of the voltage signal propagating in the network of Fig. 2.1. It is seen that linear
dispersive parameter λ = C S /C0 and nonlinear parameters α and β do not affect the symmetry of
the first-order non-autonomous RW phase evolution. Reprint from Ref. [22], Copyright 2022, with
permission from American Physical Society

Next, our attention is focused to wave packet centered around (ω, k) for the angular
frequency ω = ω p = 2π f p and wavenumber k = k p . To derive the amplitude equa-
tion that governs the slowly modulated waves propagating along our network system,
we follow Taniuti and Yajima [49] and employ the reductive perturbation method in
the semidiscrete
  limit. For this aim, we introduce the following two slow variables,
x = ε n − υg t and τ = ε2 t, where the small ε  1 measures the smallness of the
modulation frequency and the amplitude of the input waves, and υg = dω/dk is the
group velocity of the wave packet. Then, we seek the solution Vn (t) of Eq. (4.39) in
the general form [49]
142 4 Emission of Rogue Wave Signals in Nonlinear Electrical Transmission Networks

Fig. 4.25 The evolution of the RW’s phase at time τ = 0 for a different values of the linear
dispersion parameter λ = C S /C0 , λ = 0.3 (the solid line), λ = 0.31 (the dashed line), and λ = 0.32
(the dotted line); or b different values of the nonlinearity parameter α, viz., α = 0.205 (the solid line),
α = 0.2075 (the dashed line), and α = 0.21 (the dotted line); or c different values of the nonlinearity
parameter β, viz., β = 0.0157 (the solid line), β = 0.0177 (the dashed line), and β = 0.0197 (the
dotted line). Other parameters are the same as those used in Fig. 4.24. It is seen that λ and β have
the same effects on the phase evolution, while λ and β, or α and β have opposite effects. Reprint
from Ref. [22], Copyright 2022, with permission from American Physical Society

1  
Vn (t) = ε 2 u(x, τ )eiθ + ε u 20 (x, τ ) + u 22 (x, τ )e2iθ
3  
+ε 2 u 30 (x, τ ) + u 33 (x, τ )e3iθ
 
+ε2 u 40 (x, τ ) + u 42 (x, τ )e2iθ + u 44 (x, τ )e4iθ (4.41)
5    7
+ε 2 u 50 (x, τ ) + u 53 (x, τ )e3iθ + u 55 (x, τ )e5iθ + c.c. + O ε 2 ,

where θ = kn − ωt is the rapidly varying phase, and c.c. stands for the complex
conjugation.
Under all the above assumptions, we find that the fundamental harmonic u(x, τ )
of ansatz (4.41) must satisfy the following generalized NLS equation with cubic
derivatives terms:

∂u ∂ 2u ∂ |u|2 u ∂ |u|2
i + P 2 + Q 1 |u|2 u + Q 2 |u|4 u + iγ1 + i (μ1 − 2γ1 ) u = 0,
∂τ ∂x ∂x ∂x
(4.42)
4.3 Chirped Super Rogue Waves Propagating … 143

where



1  2 k
P= u 0 + λω02 cos [k] − 4λυg2 sin2 − 4ωυg sin [k] , (4.43)
2ω 2
3β − 2αℵ0
Q 1 = υg − 2αυg ℵ1
2

 u 20 + 4λω2
+ ℵ1 2υg 1 + 4λ cos [k] −
2 2
sin [2k]
ω
    2 
4u 20 ωα 2 υg υg sin [k] u 0 − λω2 cos [k]
+ υg − + 6λ   +    , (4.44)
2 ω 1 + 4λ cos2 2k ω 1 + 4λ sin2 2k
υg2 − u 20
   
6αβυg2 (ℵ0 + ℵ1 ) + u 20 − υg2 2α (ℵ2 + ℵ4 + ℵ1 ℵ3 ) − 3β ℵ3 + 2ℵ21
Q2 = ω  , (4.45)
2 υg2 − u 20
 !   "
2αω 2αωυg + ℵ1 u 20 + 4λω2 sin [k] − 2ωυg 1 + 4λ cos2 [k]
γ1 = 2υg (αℵ0 − β) −   ,
ω02 + 4u 20 − 4ω2 − 4 u 20 + 4λω2 cos2 [k]
(4.46)
 !   "
2αω 2αωυg + ℵ1 u 20 + 4λω2 sin [k] − 2ωυg 1 + 4λ cos2 [k]
μ1 = 4υg (αℵ0 − β) −   .
ω02 + 4u 20 − 4ω2 − 4 u 20 + 4λω2 cos2 [k]
(4.47)

Here, ω is the angular frequency given by the linear dispersion relation (4.40) and

2αυg2 4αω2
ℵ0 = −    ,
υg2 − u 20 ω02 + 4u 20 − 9ω2 − 4 u 20 + 9λω2 cos2 3k2
4αω2
ℵ1 = −    ,
ω02 + −
4u 20 − 4 u 20 + 9λω2 cos2 3k2
9ω2
  
αυg2 4α 2 υg4 2β 4αυg2
ℵ2 = 2 2ℵ1 + 
2
2 − + 3ℵ1 ,
υg − u 20 υg2 − u 20 α υg2 − u 20
9ω2
ℵ3 =    
ω02 + 4u 20 − 9ω2 − 4 u 20 + 9λω2 cos2 3k2
 
8α 2 ω2
β+ 2   ,
ω0 + 4u 20 − 4ω2 − 4 u 20 + 4λω2 cos2 [k]
  
8ω2 υg2 − u 20 (3βℵ1 − αℵ3 ) + αυg2 (3β − 2αℵ1 )
ℵ4 =  2    ,
υg − u 20 ω02 + 4u 20 − 4ω2 − 4 u 20 + 4λω2 cos2 [k]
 2 
u 0 − λω2 sin [k]
υg =    . (4.48)
ω 1 + 4λ sin2 k2

Equation (4.42) is the amplitude equation that describes the dynamics of modu-
lated waves propagating through the network system of Fig. 2.1. It is a generalized
144 4 Emission of Rogue Wave Signals in Nonlinear Electrical Transmission Networks

Fig. 4.26 a Linear dispersive curve showing the evolution of the frequency f p = ω/2π as function
of the wavenumber k for given three values of the dispersive parameter λ; b Behavior of the group-
velocity dispersion P as function of the wavenumber k for given three values of λ; c Plot of the
Kerr nonlinearity Q 1 versus the wavenumber k for given three values of λ: λ = 0 (solid line),
λ = 0.03 (dot-dashed line), and λ = 0.06 (dashed line). Different plots are generated with the
network parameters (4.49). Reprint from Ref. [48], Copyright 2022, with permission from John
Wiley and Sons

cubic-quintic NLS equation with cubic derivative terms that can be used to various
context of nonlinear phenomena [47, 50–59] . Its coefficients have different mean-
ings, depending on the physical context of its use. In the context of fiber optics
for example, the u(x, τ ) designates the complex envelope of an optical pulse, vari-
ables τ and x are respectively the distance and retarded time, P, Q 1 , and Q 2 are
respectively the GVD, the Kerr nonlinearity, and the quintic nonlinearity, γ1 and μ1
account respectively for the pulse self-steepening effect and the self-frequency shift
coefficient.
For the numerical computation purpose, we will use the following network param-
eters

Vb = 2V, C0 = 320 pF, L 1 = 220 mH, L 2 = 470 mH, α = 0.21 V−1 ,


β = b = 0.0197 V−2 , (4.49)

and choose the linear dispersive element C S from the condition λ < LL 21 , so that the
group velocity υg remains nonnegative in the first Brilloum zone 0 <= k <= π . .
It is important to note that for a given set of the network parameters, parameter
P of the GVD as well as parameter Q of the cubic nonlinearity, as we can see from
plots of Fig. 4.26 showing the evolution of the propagating frequency f p , the GVD
parameter P and the kerr nonlinearity Q 1 versus k, may take, depending on the
value of the carrier wavenumber k, any sign. Plots of Fig. 4.26 show how much the
dispersive element of the network of Fig. 1.1 may affect the propagating frequency as
well as parameters P and Q 1 . For example, the propagating frequency [parameters
P and Q 1 ] decreases [increase] with the increase in the parameter λ = C S /C0 .
4.3 Chirped Super Rogue Waves Propagating … 145

4.3.2 Phase Engineering Chirped Super Rogue Waves for a


Lossless Electric Network

The main purpose of this subsection is to show that the network system of Fig. 2.1 may
support the propagation of chirped SRWs. For this aim, it is sufficient to show that the
amplitude equation (4.42) admits chirped super rogue wave solutions, that is, higher-
order (HO) rational polynomial solutions of nonlinear chirping. In the following, we
limit our study to the special case of positive GVD P(k) and present such HO rational
solutions. For such parameter P(k) of the GVD, the Kerr nonlinearity Q 1 is positive,
as we can see from Fig. 2.26ac.

4.3.2.1 Phase Imprint Transformation (PIT)

To find HO rational polynomial solutions of the amplitude equation (4.42), we suit-


ably reduce this equation to either an integrable or a nearly integrable model whose
solutions are known. For this aim, use the PIT and introduce the following ansatz

u(x, τ ) = ψ(ξ, τ ) exp [iϑ(ξ, τ )] (4.50a)


 ∗ 
x ∂ϑ ∂ϑ ∂ψ ∂ψ
ξ=√ , = α0 |ψ|2 , = iα1 ψ − ψ∗ + α2 |ψ|4 ,
2P ∂ξ ∂τ ∂ξ ∂ξ
(4.50b)

where ψ and ϑ are respectively the new order parameter and the phase-imprint,
α0 , α1 , and α2 the “phase-imprint parameters”. Transformation (4.50a)–(4.50b) then
reduces the amplitude equation (4.42) to the following form:

∂ψ 1 ∂ 2ψ γ + 2Pα0 + 2α1 ∂ψ |ψ|2


i + + Q 1 |ψ| 2
ψ + i √
∂τ 2 ∂ξ 2 2P ∂ξ
μ − 2γ − Pα0 − 3α1 ∂ |ψ| 2
+i √ ψ
2P ∂ξ
 
+ Q 2 − γ α0 − α2 − Pα02 |ψ|4 ψ = 0. (4.51)

To obtain a known integrable or nearly integrable equation, we ask that



γ1 − 2μ1 + (2μ − γ ) 2P
α0 = ,
4P √
2μ1 − 3γ1 + (3γ − 2μ) 2P
α1 = , (4.52)
4
1
α2 = Q 2 − γ1 α0 − Pα02 + (γ − μ) (μ − 2γ ) .
2
146 4 Emission of Rogue Wave Signals in Nonlinear Electrical Transmission Networks

For these values α0 , α1 , and α2 , Eq. (4.52) becomes

∂ψ 1 ∂ 2ψ ∂ψ |ψ|2 ∂ |ψ|2
i + + Q 1 |ψ| 2
ψ + iγ + i (μ − 2γ ) ψ
∂τ 2 ∂ξ 2 ∂ξ ∂ξ
1
+ (μ − γ ) (μ − 2γ ) |ψ| ψ = 0,
4
(4.53)
2
which contains only three parameters, namely, the Kerr nonlinearity Q 1 and two free
real parameters γ and μ. Equation (4.53) contains a number of known integrable
equations that admit HO rational polynomial solutions with nonlinear chirping such
as the Kundu–Eckhaus (KE) equation (γ = 0) [60], the Kaup–Newell type NLS
(KN–NLS) equation (μ = 2γ ) [61], the Gerdjikov–Ivanov (GI) equation (μ = 0)
[62], the Chen–Lee–Liu type NLS (CLL–NLS) equation (μ = γ = 0) [63], as well as
the standard NLS equation (μ = γ = 0). In this last section of this chapter, we apply
for the first time, to the best of our knowledge, HO rational polynomial solutions of
Eq. (4.53) to engineer chirped super rogue waves in nonlinear transmission networks.

4.3.2.2 Baseband Modulational Instability

Before presenting HO rational polynomial solutions of Eq. (4.53), it is reasonable to


start by investigating the problem of the baseband MI, a mechanism responsible of
rogue wave formation in nonlinear media [64].
One can easily verify the that Eq. (4.53) admits the below Stokes constant ampli-
tude wave solution

k02 1
ψ0 (ξ, τ ) = a0 exp [i (k0 ξ − 0 τ )] , 0 = + (γ k0 − Q 1 ) a02 + (γ − μ) (μ − 2γ ) a04 .
2 2
(4.54)
Its baseband MI can be studies by introducing the following perturbation,

ψ p (ξ, τ ) = (a0 + δa0 (ξ, τ )) exp [i (k0 ξ − 0 τ )] ,

where δa0 (ξ, τ ) is a small complex perturbation. Following the idea developed in
Ref. [64], we obtain the following criterion of the baseband MI of the CW (4.54):
γ  
Q 1 > 0 and k0 + a02 (μ − γ ) < 1. (4.55)
Q1

Hence, Eq. (4.55) gives the parameter conditions under which equation (4.53) admits
rational polynomial solutions.

4.3.2.3 Dynamics of Chirped Super Rogue Waves Propagating


Through the Lossless Electrical Network

Now, we turn to the dynamics of chirped super rogue waves propagating along the
network system of Fig. 2.1. For any a0 and k0 for which conditions (4.55) of the
4.3 Chirped Super Rogue Waves Propagating … 147

baseband MI are satisfied, we can use the CW solution (4.54) as the seed solution.
Following then Chen et al. [47], we obtain the following first-order and second-order
rational polynomial solutions ψ [I ] (ξ, τ ) and ψ [I I ] (ξ, τ ) of Eq. (4.53):
⎡    ⎤
1 + 2i β 2 −γ a02 a02 μ+ω
γξ + τ
⎢ a02 a02 ⎥
ψ [I ] (ξ, τ ) = a0 ⎢
⎣1 −


M +iN



2 (γ − μ) N
× exp i k0 ξ − 0 τ + arctan , (4.56a)
γ M
with
 
 
1 β2   2
M= + + γ 2 a02 ξ 2 − 2 k0 + μa02 ξ τ + 2 2
β + k0 + μa0 τ2 ,
4a02 a02
(4.56b)
#  $
N = γ ξ − a02 (γ + μ) + k0 τ ,
 
γ # $
β = a0 Q 1 1 − k0 + a02 (μ − γ ) , (4.56c)
Q1

and


G 2 + i H2
ψ [I I ]
(ξ, τ ) = a0 1 −
G 1 − i H1

  

μ H1
exp i k0 ξ − 0 τ + 2 − 1 arctan , (4.57a)
γ G1

with G 1 , G 2 , H1 , and H2 being polynomials of τ and ζ = ξ − (a0 μ + k0 ) τ,


defined as
 3 2 2 3  # 
G 1 = 64 a04 γ 2 + β 2 β τ + ζ 2 + a04 γ 2 + β 2 48β 2 4a08 γ 4 + 17a04 β 2 γ 2 + 9β 4 τ 4
 
+384a06 β 2 γ 3 ζ τ 3 − 288 2a04 γ 2 + β 2 a04 γ 2 + β 2 ζ 2 τ 2
 $
+384a06 γ 3 ζ 3 τ − 48 3a04 γ 2 − β 2 ζ 4
 
+36 28a08 γ 4 + 35a04 β 2 γ 2 + 11β 4 τ 2 − 288a06 γ 3 ζ τ + 36 7a04 γ 2 + 3β 2 ζ 2 + 9,
 2 2 2 2 2
H1 = 192a02 γ a04 γ 2 + β 2 β τ + ζ2 a0 γ τ − ζ
#  
+96a0 γ a0 γ 6a0 γ + 13a0 β γ + 9β 4 τ 3 + 6a08 γ 4 + 15a04 β 2 γ 2 + 3β 4 ζ τ 2
2 2 8 4 4 2 2
  $ 
+3a02 γ a04 γ 2 − β 2 ζ 2 τ − 3a04 γ 2 + β 2 ζ 3 + 36a02 γ 11a02 γ τ − 3ζ , (4.58)
 2  
G 2 = 192 a04 γ 2 + β 2 β 2 τ 2 + ζ 2 4a02 γ ζ τ + 5β 2 τ 2 + ζ 2 − 1152a04 β 2 γ 2 τ 2
 # $
+288 3a04 γ 2 + β 2 2a04 γ 2 + 3β 2 τ 2 − 2a02 γ ζ τ + ζ 2 − 36,
 2 2 2 2 2 #
H2 = 384 a04 γ 2 + β 2 β τ + ζ2 a0 γ ζ + β 2 τ + 192 4a08 γ 4 + 3a04 β 2 γ 2 + β 4 β 2 τ 3
148 4 Emission of Rogue Wave Signals in Nonlinear Electrical Transmission Networks
  
−3a02 γ a04 γ 2 + 3β 2 2a04 γ 2 + β 2 ζ τ 2 − 3 2a08 γ 4 + a04 β 2 γ 2 + β 4 ζ 2 τ
 $ # $
+a02 γ a04 γ 2 − β 2 ζ 3 − 72 12a04 γ 2 + 5β 2 τ + a02 γ ζ .

In Eqs. (4.56a)–(4.58), a0 , k0 , and 0 are three real parameters satisfying the rela-
tionship showed in Eq. (4.54) as well as the baseband MI conditions (4.55).
Going back to variables x and τ, we obtain respectively the following first- and
second-order SRW solutions of the amplitude equation (4.42)
⎡    ⎤
β 2 −γ a02 a02 μ+k0
1
+ 2i γ ξ + τ
⎢ a02 a02 ⎥
u [I ] (x, τ ) = a0 ⎢
⎣1 −


M + iN



 
2 (γ − μ) N 
× exp i k0 ξ − 0 τ + ϑ(ξ, τ ) + arctan  , (4.59)
γ M 
ξ = √x
2P




G 2 + i H2 2 (μ − γ ) H1
u [I I ] (x, τ ) = a0 1 − exp i k0 ξ − 0 τ + ϑ(ξ, τ ) + arctan ,
G 1 − i H1 γ G1
(4.60)

with ξ = √x2P , ζ = √x2P − (a0 μ + k0 ) τ , and ϑ(ξ, τ ) beings the phase-imprint


defined by Eq. (4.50b). Equations (4.59) and (4.60) show that the intensity of each
of the super first- and second-order RWs is proportional to ρ0 = |a0 |; henceforth,
parameter ρ0 will be referred to as the amplitude parameter.
The chirping across the wave is given by

∂  [I,I I ] 
δω(x, τ ) = −  ,
∂x

where [I ] and [I I ] stand for respectively the first- and the second-order rational
polynomial solutions (4.59) and (4.60), and are defined respectively as follows:


2 (γ − μ) N
[I ] = k0 ξ − 0 τ + ϑ + arctan
γ M
    
N − 2M a02 γ ξ + β 2 − γ a02 a02 μ + ω τ
+ arctan 2  2   2     , (4.61a)
a0 M + N 2 − M − 2N a0 γ ξ + β 2 − γ a02 a02 μ + ω τ


2 (μ − γ ) H1 G 1 H2 + H1 G 2
[I I ] = k0 ξ − 0 τ + ϑ + arctan − arctan 2 .
γ G1 G 1 + H12 + H1 H2 − G 1 G 2
(4.61b)

The first and the last terms of phase [I ] and those of phase [I I ] being intrinsic
respectively to all first-order RW categories [47, 65] and to all second-order RW
categories [47, 66], we will ignore their contribution in the evolution of the fre-
quency chirp. Nonlinear chirping associated with the first- and second-order rational
solutions (4.59) and (4.60) are then found to be respectively
4.3 Chirped Super Rogue Waves Propagating … 149

⎡     2
 β 2 − γ a02 a02 μ + k0
1 ⎣  1 
δω[I ] (x, τ ) = − √   a 2
α
0 0  M + i N − − 2i γ ξ + τ 
2P M 2 + N 2 a02 a02 
  
2 (γ − μ) 2 ∂ N 
+ M  (4.62a)
γ ∂ξ M ξ = √x
2P

 
1 ! " 2 (μ − γ ) 2 ∂ H1
δω[I I ] (x, τ ) = − √  2  a 2
0 α0 (G 1 − G 2 )2
+ (H1 + H 2 )2
+ G 1 ,
2P G 1 + H12 γ ∂ζ G 1
(4.62b)

with ζ = √x2P − (a0 μ + k0 ) τ .


Before discussing on the first- and second-order rational solutions (4.59) and
(4.60) and the corresponding chirping, we demonstrate in Fig. 4.27, as an illustrative
example, the evolution of the first-order (left panels) and second-order (right panels)
chirped SRWs with the corresponding chirping, obtained with different values of
parameters γ and μ. From plots of Fig. 4.27, we can see that both waves and the
corresponding chirping are spatiotemporal localized, the first-order [second-order]
SRW displays 3-fold [5-fold] peak amplitude formed of one hump and two valleys
[one hump and four valleys] around the center; depending of the numerical values
of γ and μ, frequency chirp associated with the first- and second-order SRWs may
exhibit either a bright or dark doubly localized structure. Therefore, the behavior of
the chirping strongly depends on the choice of parameters γ and μ.
Surely, the structure and the behavior of SRWs propagating along the network
system of Fig. 2.1 may be affected by the network parameters as well as by various
parameters of the amplitude equation. In the following, we restrict our investigation
on the effects of parameter μ of the nonlinearity dispersion, the dispersive parame-
ter λ, or the propagating frequency f p = ω/2π on the dynamics of chirped SRWs
propagating thought the network system of Fig. 2.1.
Effects of parameter μ of the nonlinearity dispersion on the chirped super rogue
waves To study the effects of parameter μ of the nonlinearity dispersion on the SRW
motion, we set γ = 1 and display the evolution of the first-order and the second-
order SRW ans the corresponding chirping in respectively Fig. 4.28 and Fig. 4.29
for different values of the nonlinearity dispersion parameter μ. Different plots are
generated with the network parameters (4.49). The first, the second, and the third
columns in each of these figures show the wave evolution corresponding respectively
to the GI equation, the CLL–NLS equation, and the KN-NLS equation. It follows
from plots of these figures that the first-order [second-order] super rogue waves for
each of the three equations displays a 3-fold [5-fold] peak amplitude located on
the origin. Plots of Figs. 4.28 and 4.29 also reveal that all SRWs have an extended
spatiotemporal distribution as parameter μ increases. Also, these figures show that
depending on the GI, CLL, or KN models used, the chirping corresponding to the
first-order SRW exhibit a dark single or doubly localized structure, while that cor-
responding to the second-order SRW exhibits, independently of the equation model,
either a dark doubly localized structure or a bright doubly localized structure. As we
150 4 Emission of Rogue Wave Signals in Nonlinear Electrical Transmission Networks

Fig. 4.27 Evolution plots of the chirped SRWs |u(x, τ )| and the frequency chirp δω(x, τ ) obtained
as the result of the rational solutions (4.59) and (4.60) of Eq. (4.42) with the corresponding frequency
chirps (4.62a) and (4.62b). a, d First- and second-order SRWs obtained with γ = 0.1 and μ = 3.5;
b, e Frequency chirps corresponding to the first- and second-order SRWs of plot (a) and (b),
respectively; c and f Frequency chirps given by Eqs. (4.62a) and (4.62b) obtained with γ = 3.5
and μ = 0.1. Other parameters used for generating different plots are the network parameters
(4.49), λ = 0.175, k = π/8, k0 = π/3, and a0 = 1. Reprint from Ref. [48], Copyright 2022, with
permission from John Wiley and Sons

can observe from plots of Figs. 4.28 and 4.29, the nonlinear dispersion parameter
μ produces an important skew angle relative to the ridge of the rogue wave in the
clockwise direction as μ increases.
Effects of the dispersive element C S on the evolution of chirped SRWs Via the
dispersive parameter λ = C S /C0 , we here study the effects of the dispersive element
C S of the network on the dynamics of modulated chirped SRWs propagating along
the network system of Fig. 2.1. With the use of the network parameters (4.49) with
different C S (therefore, with different λ), we show in Fig. 4.30 the evolution of the
first-order (Fig. 4.30a–f) and second-order (Fig. 4.30g–l) SRWs with the correspond-
ing frequency chirp in the general situation when γ μ = 0 (here, we have used γ = 1
and μ = 1.5). It is clearly seen from plots of this figure that the first-order [second-
order] SRW has a 3-fold [a 5-fold] peak amplitude whose width increases with the
4.3 Chirped Super Rogue Waves Propagating … 151

Fig. 4.28 Super first-order rogue waves in the anomalous dispersion regime. a, d, g, j Super first-
RW and the corresponding frequency chirp δω[I ] (x, τ ) of the GI equation (γ = 1, μ = 0) in the
anomalous dispersion regime; b, e, h, k Super first-RW and the corresponding frequency chirp
δω[I ] (x, τ ) of the CLL–NLS equation ( μ = γ = 1) in the anomalous dispersion regime; c, f, i,
l Super first-RW and the corresponding frequency chirp δω[I ] (x, τ ) of the KN-NLS equation (
μ = 2γ = 2) in the anomalous dispersion regime. To generate different plots, we have used the
network parameters (4.49) with λ = 0.175, the wavenumber k ≈ 1.27 × 10−6 corresponding to the
propagating frequency f p ≈ ω0 /2π, and the plane wave parameters a0 = 1 and k0 = −1. Reprint
from Ref. [48], Copyright 2022, with permission from John Wiley and Sons

increase in parameter λ. Also, this figure reveals that the chirping associated with
the first-order [second-order] SRW exhibits a two dark [one bright] doubly localized
structure whose width increases with increasing in the values of parameter λ.
Effects of parameters γ , μ, k0 and a0 on the SRWs propagating through a lossless
network system Here, we intend to show that the choice in parameters γ and μ may
affect the wave structure. For this purpose, we use the same network parameters as
in Fig. 4.30 and the same carrier wavenumber, but different values of parameters γ
and μ (here we have used γ = 0.1 and μ = 3.5) to show in Fig. 4.31 the structure of
the first- and second-order SRWs with the corresponding chirping. Comparing plots
of Fig. 4.30 with the corresponding plots in Fig. 4.31, it is clearly seen that they have
different structures. We can visualize the destruction of chirped SRWs in Fig. 4.31
when varying the dispersion parameter λ in the same way as in Fig. 4.30. More
152 4 Emission of Rogue Wave Signals in Nonlinear Electrical Transmission Networks

Fig. 4.29 Super second-order rogue waves in the anomalous dispersion regime. a, d Super second-
RW of the GI equation (γ = 1, μ = 0) in the anomalous dispersion regime; g, j: Frequency chirp
u [I I ] (x, τ ) associated with the super second-RW of the GI equation (γ = 1, μ = 0) in the anomalous
dispersion regime; b, e: Super second-RW of the CLL–NLS equation ( μ = γ = 1) in the anomalous
dispersion regime; h, k Frequency chirp u [I I ] (x, τ ) corresponding to the super second-RW of the
CLL–NLS equation ( μ = γ = 1) in the anomalous dispersion regime; c, f Super second-RW
of the KN-NLS equation ( μ = 2γ = 2) in the anomalous dispersion regime; i, l Frequency chirp
u [I I ] (x, τ ) associated with the super second-rogue wave of the KN-NLS equation ( μ = 2γ = 2) in
the anomalous dispersion regime. Different plots are generated with the network parameters (4.49)
and λ = 0.175 for the wavenumber k ≈ 1.27 × 10−6 associating to the propagating frequency
f p ≈ ω0 /2π, and for the plane wave parameters a0 = 1 and k0 = −1. Reprint from Ref. [48],
Copyright 2022, with permission from John Wiley and Sons

interestingly, plots of Fig. 4.31 reveal that with the increasing in λ, the structure of
the SRWs changes progressively, and after a certain critical value λc of λ, the SRW
and the corresponding chirping splits into two giant bright pulses (two-soliton) whose
amplitude and width can be modulated by readjusting the propagating frequency f p ,
as we can clearly observe from plots of Figs. 4.32 and 4.33.
Our investigations have also showed that for a given set of the network param-
eters and for given parameters k0 , a0 , γ , and μ, there does not exist a critical
frequency f pc before/after which the above chirped SRWs may split into two-
solitons. For a better understanding, we depict in Fig. 4.34 the density plot of the
wave intensity |u(x, τ )| for the network parameters (4.49) with (λ, γ , μ, a0 , k0 ) =
4.3 Chirped Super Rogue Waves Propagating … 153

Fig. 4.30 Effects of the dispersive element C S on the chirp SRWs for the network parameters
(4.49) and different values of λ = C S /C0 . a–c and g–i Spatiotemporal evolution of the super first-
order RW and the corresponding frequency chirp δ (I ) ω(x, τ ); 5d–f and j–l Evolution plots of the
super second-order RWs with the corresponding frequency chirp δ (I ) ω(x, τ ). Different plots are
generated with λ = L 2 /20L 1 (left column), λ = 0.5 L 2 /L 1 (middle column), and λ = 0.9 L 2 /L 1
second, and third column (right column). Other parameters are the same as in Figs. 4.28 and 4.29
with γ = 1 and μ = 1.5. Reprint from Ref. [48], Copyright 2022, with permission from John Wiley
and Sons

(0.9L 2 /L 1 , 0.1, 3.5, 1, −1) for different propagating frequencies, namely, f p =


412.01 kHz, f p = 415.455 kHz, and f p = 428.062 kHz. We can conclude from
plots of Fig. 4.34 that for a given set of the network parameters and for a given set of
parameters γ , μ, a0 , and k0 , the frequency region of the baseband MI can be divided
into a finite number of subregions such that for all frequencies taken from the same
subregion, each of the rational polynomial solutions (4.59) and (4.60) either leads
to only the chirped SRWs, or leads to only the one-solitons, or leads to only the
two-solitons.

4.3.3 Computational/Numerical Simulations

Before ending this Section, we carry out computational/numerical simulations to


confirm the propagation of chirped SRWs through the network system of Fig. 2.1. The
numerical study is done with the use of the network equations (4.39), assuming the
154 4 Emission of Rogue Wave Signals in Nonlinear Electrical Transmission Networks

Fig. 4.31 Destruction of chirped SRWs obtained from Eqs. (4.59) and (4.60) under the manipu-
lation of the dispersive parameter λ. a–c, d–f Evolution plots of the first-order SRW according to
Eq. (4.59) and the corresponding frequency chirp given by Eq. (4.62a) for different values of the
dispersive parameter λ; g–i and j–l Spatiotemporal evolution of the second-order SRW according
to Eq. (4.60) and the corresponding frequency chirp given by Eq. (4.62b) for different values of the
dispersive parameter λ. Plots of the left, middle, and right columns are obtained with respectively
L2 9L 2
λ = L 2 /20L 1 , λ = 2L 1
, and λ = 10L 1
. Different plots are obtained for the network parameters
(4.49) with γ = 0.1, μ = 3.5, other parameters being the same as in Fig. 4.30. Reprint from Ref.
[48], Copyright 2022, with permission from John Wiley and Sons

left extremity of the network to be excited with one of the above rational polynomial
solutions of the amplitude equation (4.42). Next, we set to zero the voltage across
the other extremity and runs the experiment for a sufficiently long time [67]. For
simplicity, we focus our attention to the situation when α0 = 0, corresponding to a
zero phase-imprint. We then have, for the input signal,

  
γ1 − 2μ1 1 γ1 − 2μ1 2 (γ1 − 2μ1 )
ϑ = 0, α1 = α2 = 0, γ = √ + 2μ, Q 2 − √ +μ 3μ − √ = 0.
2P 2 2P 2P
(4.63)
In the numerical simulations, parameters μ is considered as a free parameter, and
can be used to modulate the propagating frequency.
Going back to original variables n and t and considering only the first term in the
expansion (4.41), we find that the shape of the electrical chirped SRWs can be well
4.3 Chirped Super Rogue Waves Propagating … 155

Fig. 4.32 Wave profile (a, b) and the corresponding chirp profile c, d showing respectively the
wave propagation and the evolution of the corresponding frequency chirp at time τ = 0 for different
values of the dispersive element λ. Different plots are generated with the following values of the
L2
dispersive element λ = 20L 1
(solid line), λ = 0.5L 9L 2
L 1 (dot-dashed line), and λ = 10L 1 (dashed line).
2

Other parameters are the same as those used in Fig. 4.31. Reprint from Ref. [48], Copyright 2022,
with permission from John Wiley and Sons

approximated by the shape of u(x, τ ):


1    
Vn (t) = 2ε 2 Re u [I ] (ε n − υg t , ε2 t) exp [i (kn − ωt)] , (4.64a)

for the first-order SRWs, and


1    
Vn (t) = 2ε 2 Re u [I I ] (ε n − υg t , ε2 t) exp [i (kn − ωt)] , (4.64b)

for the super second-order rogue wave.


With the use of the network parameters (4.49) and different values of the dis-
persive parameter λ, the CW background parameters a0 = 1/4 and k0 = −1, and
ε = 10−3 , we solve the network equations (4.39) by means of the fourth order
Runge–Kutta schemes with normalized constant time step t = 4. 6616 × 10−3 =
1.33 × 10−12 u 0 . Numerical values of different parameters involving in the compu-
tational simulation are give in Table 4.1.
156 4 Emission of Rogue Wave Signals in Nonlinear Electrical Transmission Networks

Fig. 4.33 Effects of the wavenumber k (i.e., of the propagating frequency f p = ω(k)/2π ) on the
two-soliton of the network system of Fig. 2.1. Plots (a) and (b) show the profile at time τ = 0
of respectively the first- and second-order SRW, while plots (c) and (d) show the profile of the
corresponding frequency chirp at the same time τ . To generate different plots, we have used the
dispersive element λ = 9L 2 /10L 1 and the same parameters as in Fig. 4.31 except the values of the
wavenumber k which have been taken as shown in the legend. In other words, the solid line, the dot-
dashed line, and the dashed line show the curve obtained with the wavenumber k = 1.26 × 10−6 ,
k = 5.26 × 10−6 , and k = 8.26 × 10−6 , respectively. Reprint from Ref. [48], Copyright 2022,
with permission from John Wiley and Sons

Examples of the obtained results are displayed in Figs. 4.35 and 4.36 showing
the propagation of respectively the first-order and second-order chirped super rogue
wave voltages along the network of Fig. 2.1. Plots of these two figures show that the
propagating speed increases [decreases] with the increase [decrease] in the propa-
gating frequency f p [in the dispersive parameter λ].

4.3.4 Conclusion and Discussion

In this last Section of this Chapter, we have considered the one-dimensional lossless
nonlinear electric transmission network with dispersive element showed in Fig. 2.1.
The amplitude equation is reduced to a cubic-quintic NLS equation with cubic
derivative terms which generalized the Chen–Lee–Liu type NLS equation, the Kaup–
Newell type NLS equation, the Gerdjikov–Ivanov equation, and the Kundu–Eckhaus
equation. The criterion of the baseband MI is presented. HO rational polynomial
solutions with nonlinear frequency chirp are built and used to experience the trans-
mission of chirped super rogue wave through the network system of Fig. 2.1. The
4.3 Chirped Super Rogue Waves Propagating … 157

Fig. 4.34 Evolution density plot of (a)–(c) the chirped super first-order rogue wave solution (4.59)
and (d)–(f) the chirped super second-order rogue wave solution (4.60) for three values of the
propagating frequency f p , a, d f p = 412.01 kHz, b, e f p = 415.455 kHz, and c, f f p = 428.062
kHz. Different parameters used in different plots are given in the text. Reprint from Ref. [48],
Copyright 2022, with permission from John Wiley and Sons

Table 4.1 Parameters used in the computational/numerical simulations


λ μ k f p [kHz] γ
0.45 LL 21 15 0.106811 413.099 0.0021
0.45 LL 21 6 0.145288 415.339 45.7513
0.45 LL 21 17 0.173257 417.348 47.7513
0.45 LL 21 15.9999 0.145285 415.339 45.7508
0.5 LL 21 14.4412 0.152735 415.339 41.9316
0.55 LL 21 13.0636 0.161462 415.339 38.3722

effects of various parameters such as the network parameters, the carrier wavenum-
ber, parameters of the CW background, as well as the solution parameters on the
emission of chirped SRWs through our network system are investigated. Our theo-
retical investigations have been confirmed by a numerical simulations from which
we have obtained that the dispersive linear capacitance C S of the network impacts
the SRWs speed during their propagation.
158 4 Emission of Rogue Wave Signals in Nonlinear Electrical Transmission Networks

Fig. 4.35 Propagation of the chirped super first-order rogue wave in the network for the line
parameters (4.49) and the CW background parameters a0 = 1 and k0 = π/3. a–c Evolution of
the chirped first-order SRW in the network at different frequencies, a f p = 413.099 kHz, b f p =
415.339 kHz, and c f p = 417.348 kHz. d–f Chirped first-order SRW propagating in the network at
45L 2
the frequency f p = 415.339 kHz for different values of the dispersive parameter λ, d λ = 100L 1
,e
5L 2 55L 2
λ = 10L 1
, and (f) λ = 100L 1
. Other parameters are given in the text and in Table 4.1. Reprint from
Ref. [48], Copyright 2022, with permission from John Wiley and Sons

Fig. 4.36 Spatial evolution of the chirped super first-second rogue wave voltage in the network for
the line parameters (4.49) and the CW background parameters a0 = 1 and k0 = π/3. Plots of panels
(a)–(c) show the chirped first-second SRW propagating in the network at the different frequencies,
a f p = 413.099 kHz, b f p = 415.339 kHz, and c f p = 417.348 kHz, while plots d–f show the
chirped first-second SRWs propagating at frequency f p = 415.339 kHz in the network for different
45L 2 5L 2 55L 2
values of the dispersive parameter λ, d λ = 100L 1
, e λ = 10L 1
, and (f) λ = 100L 1
. Other parameters
are given in Table 4.1 and it the text. Reprint from Ref. [48] , Copyright 2022, with permission from
John Wiley and Sons
References 159

References

1. E. Pelinovsky, C. Kharif, Extreme Ocean Waves (Springer, New York, 2008)


2. C. Kharif, E. Pelinovsky, A. Slunyaev, Rogue Waves in the Ocean (Springer, New York, 2009)
3. P. Müller, C. Garrett, A. Osborne, Rogue waves. Oceanography 18, 66 (2005)
4. N. Akhmediev, A. Ankiewicz, J.M. Soto-Crespo, Rogue waves and rational solutions of the
nonlinear Schrödinger equation. Phys. Rev. E 80, 026601 (2009)
5. A. Ankiewicz, N. Devine, N. Akhmediev, Phys. Lett. A 373, 3997 (2009)
6. D.H. Peregrine, J. Aust. Math. Soc. Ser. B 25, 16 (1983)
7. B. Kibler, J. Fatome, C. Finot, G. Millot, F. Dias, G. Genty, N. Akhmediev, J.M. Dudley, The
Peregrine soliton in nonlinear fibre optics. Nat. Phys. 6, 790 (2010)
8. N. Akhmediev, A. Ankiewicz, M. Taki, Phys. Lett. A 373, 675 (2009)
9. J.S. He, H.R. Zhang, L.H. Wang, K. Porsezian, A.S. Fokas, Generating mechanism for higher
order rogue waves. Phys. Rev. E 87, 052914 (2013)
10. L. Wang, J. He, X. Hui, J. Wang, K. Porsezian, Generation of higher-order rogue waves from
multibreathers by double degeneracy in an optical fiber. Phys. Rev. E 95, 042217 (2017)
11. L.W.C. Yang, J. Wang, J. He, The height of an nth-order fundamental rogue wave for the
nonlinear Schrödinger equation. Phys. Lett. A 381, 1714–1718 (2017)
12. J. He, S. Xu, K. Porsezian, P. Tchofo Dinda, D. Mihalache, B.A. Malomed, E. Ding, Handling
shocks and rogue waves in optical fibers. Rom. J. Phys. 62, 203 (2017)
13. T.B. Benjamin, J.E. Feir, The disintegration of wavetrains on deep water. Part 1. J. Fluid Mech.
27, 417-430 (1967)
14. A. Ankiewicz, N. Devine, N. Akhmediev, Are rogue waves robust against perturbations? Phys.
Lett. A 373, 3997 (2009)
15. A. Chabchoub, N. Hoffmann, N. Akhmediev, Rogue wave observation in a water wave tank.
Phys. Rev. Lett. 106, 204502 (2011)
16. J. Soto-Crespo, N. Devine, N. Akhmediev, Integrable turbulence and rogue waves: breathers
or solitons? Phys. Rev. Lett. 116, 103901 (2016)
17. B. Kibler, A. Chabchoub, A. Gelash, N. Akhmediev, V.E. Zakharo, Superregular breathers
in optics and hydrodynamics: omnipresent modulation instability beyond simple periodicity.
Phys. Rev. X 5, 041026 (2015)
18. D.R. Solli, C. Ropers, P. Koonath, B. Jalali, Optical rogue waves. Nature 450(7172), 1054
(2007)
19. A.N. Ganshin, V.B. Efimov, G.V. Kolmakov, L.P. Mezhov- Deglin, P.V.E. McClintock, Obser-
vation of an inverse energy cascade in developed acoustic turbulence in superfluid helium.
Phys. Rev. Lett. 101, 065303 (2008)
20. J. Kempe, Quantum random walks: an introductory overview. Contemp. Phys. 44, 307 (2003)
21. Y.V. Bludov, V.V. Konotop, N. Akhmediev, Matter rogue waves. Phys. Rev. A. 80, 033610
(2009)
22. E. Kengne, W.M. Liu, Engineering rogue waves with quintic nonlinearity and nonlinear dis-
persion effects in a modified Noguchi nonlinear electric transmission network. Phys. Rev. E
102, 012203 (2020)
23. L. Stenflo, M. Marklund, Rogue waves in the atmosphere. J. Plasma Phys. 76, 293 (2010)
24. E. Kengne, W.M. Liu, Dissipative ion-acoustic solitons in ion-beam plasma obeying a kappa-
distribution. AIP Adv. 10, 045218 (2020)
25. E. Kengne, W.M. Liu, Transmission of rogue wave signals through a modified Noguchi elec-
trical transmission network. Phys. Rev. E 99, 062222 (2019)
26. T. Taniuti, N. Yajima, Perturbation method for a nonlinear wave modulation. II. J. Maths. Phys.
10, 1369 (1969)
27. R. Marquié, J.M. Bilbault, M. Remoissenet, Nonlinear Schrödinger models and modulational
instability in real electrical lattices. Physica D 87, 371 (1995)
28. Y.S. Kivshar, M. Peyrard, Modulational instabilities in discrete lattices. Phys. Rev. A 46, 3198
(1992)
160 4 Emission of Rogue Wave Signals in Nonlinear Electrical Transmission Networks

29. F.B. Pelap, J.H. Kamga, S.B. Yamgoue, S.M. Ngounou, J.M. Ndecfo, Dynamics and properties
of waves in a modified Noguchi electrical transmission line. Phys. Rev. E 91, 022925 (2015)
30. E. Kengne, A. Lakhssassi, W.M. Liu, Modeling of matter-wave solitons in a nonlinear inductor-
capacitor network through a Gross-Pitaevskii equation with time-dependent linear potential.
Phys. Rev. E 96, 022221 (2017)
31. H.-H. Chen, C.-H. Liu, Solitons in nonuniform media. Phys. Rev. Lett. 37, 693–697 (1976)
32. A. Kundu, Landau-Lifshitz and higher-order nonlinear systems gauge generated from nonlinear
Schrödinger-type equations. J. Math. Phys. 25, 3433 (1984)
33. Y. Zhang, L. Guo, A. Chabchoub, J. He, Higher-order rogue wave dynamics for a derivative
nonlinear Schrödinger equation. Rom. J. Phys. 62, 102 (2017)
34. Y.S. Zhang, L.J. Guo, J.S. He, Z.X. Zhou, Darboux transformation of the second-type derivative
nonlinear Schrödinger equation. Lett. Math. Phys. 105, 853–891 (2015)
35. L. Guo, Y. Zhang, S. Xu, Z. wu, J. He, The higher order Rogue Wave solutions of the Gerdjikov-
Ivanov equation. Phys. Scr. 89, 035501 (2014)
36. Y. Kodama, Optical solitons in a monomode fiber. J. Stat. Phys. 39, 597 (1985)
37. P.A. Clarkson, J.A. Tuszynski, Exact solutions of the multidimensional derivative nonlinear
Schrödinger equation for many-body systems of criticality. J. Phys. A 23, 4269 (1990)
38. E. Kengne, A. Lakhssassi, W.M. Liu, R. Vaillancourt, Phase engineering, modulational insta-
bility, and solitons of Gross-Pitaevskii-type equations in 1 + 1 dimensions. Phys. Rev. E 87,
022914 (2013)
39. M. Mirzazadeh, Y. Yıldırım, E. Yasar, H. Triki, Q. Zhoud, S.P. Moshokoa, M.Z. Ullahf, A.R.
Seadawy, A. Biswas, M. Belic, Optical solitons and conservation law of Kundu-Eckhaus equa-
tion. Optik 154, 551 (2018)
40. D. Peregrine, Water waves, nonlinear Schrödinger equations and their solutions. J. Austral.
Math. Soc. B 25, 16 (1983)
41. X.G. Geng, H.W. Tam, Darboux transformation and soliton solutions for generalized nonlinear
Schrödinger equations. J. Phys. Soc. Jpn. 68, 1508 (1999)
42. L.-C. Zhao, C. Liu, Z.-Y. Yang, Pair-tunneling induced localized waves in a vector nonlinear
Schrödinger equation. Commun. Nonlinear. Sci. Numer. Simulat. 20, 21–27 (2015)
43. J.K. Duan, Y.L. Bai, Rogue wave in coupled electric transmission line. Ind. J. Phys. 92(3), 369
(2018)
44. V.I. Shrira, V.V. Geogjaev, What makes the Peregrine soliton so special as a prototype of freak
waves? J. Eng. Math. 67, 11 (2010)
45. D.J. Kedziora, A. Ankiewicz, N. Akhmediev, The phase patterns of higher-order rogue waves.
J. Opt. 15, 064011 (2013)
46. P.A. Clarkson, Nonlinearity 5, 453 (1992)
47. H.N. Chan, K.W. Chow, D.J. Kedziora, R. H.J. Grimshaw, E. Ding, Rogue wave modes for a
derivative nonlinear Schrödinger model. Phys. Rev. E 89, 032914 (2014); S. Chen, Y. Zhou, L.
Bu, F. Baronio, J.M. Soto-Crespo, D. Mihalache, Opt. Express 27, 11370 (2019)
48. E. Kengne, W.M. Liu, Phase engineering chirped super rogue waves in a nonlinear transmission
network with dispersive elements. Adv. Theory Simul. 2100062 (2021)
49. T. Taniuti, N. Yajima, Perturbation method for a nonlinear wave modulation. J. Math. Phys.
10, 1369 (1969)
50. P.A. Clarkson, C.M. Cosgrove, Painleve analysis of the non-linear Schrodinger family of equa-
tions. J. Phys. A Math. Gen. 20, 2003 (1987)
51. K.S El-Labany, W.F. El-Taibany, N.A. El-Bedwehy, N.A. El-Shafeay, The role of superthermal
electrons on the escaping ions from The upper atmosphere of Titan and Venus. Alpharama J.
Basic Appl. Sci. 1, 99 (2020)
52. G. Huang, S. Zhang, B. Hu, Nonlinear excitations in ferromagnetic chains with nearest-and
next-nearest-neighbor exchange interactions. Phys. Rev. B 58, 9194 (1998)
53. V.H. Chu, C.C. Mei, The non-linear evolution of Stokes waves in deep water. J. Fluid Mech.
47, 337 (1971)
54. A.Al. Qarni, M.A. Banaja, H.O. Bakodah, A.A. Alshaery, Q. Zhou, A. Biswas, S.P. Moshokoa,
M.R. Belic, Bright optical solitons for Lakshmanan–Porsezian–Daniel model with spatio-
temporal dispersion by improved Adomian decomposition method. Optik 181, 891 (2019)
References 161

55. M.B. Hubert, S. Nestor, Douvagai, G. Betchewe, A. Biswas, S. Khan, S.Y. Doka, Q. Zhou,
M. Ekici, M. Belic, Dispersive solitons in optical metamaterials having parabolic form of
nonlinearity. Optik 179, 1009 (2019)
56. M. Matusovsky, B. Vaynberg, M. Rosenbluh, Phys. Rev. Lett. 77, 5198 (1996)
57. J. Moses, F.W. Wise, Controllable self-steepening of ultrashort pulses in quadratic nonlinear
media. Phys. Rev. Lett. 97, 073903 (2006)
58. I.I. Tendler, A. Hartford, M. Jermyn, E. LaRochelle, X. Cao, V. Borza, D. Alexander, P. Bruza,
J. Hoopes, K. Moodie, B.P. Marr, B.B. Williams, B.W. Pogue, D.J. Gladstone, L.A. Jarvis,
Experimentally observed Cherenkov light generation in the eye during radiation therapy. Int.
J. Radiat. Oncol. Biol. Phys. 106, 422 (2020)
59. A.-M. Wazwaz, Higher dimensional nonlinear Schrödinger equations in anomalous dispersion
and normal dispersive regimes: bright and dark optical solitons. Optik 222, 165327 (2020)
60. A. Kundu, J. Math. Phys. 25, 3433 (1984)
61. D.J. Kaup, A.C. Newell, An exact solution for a derivative nonlinear Schrödinger equation. J.
Math. Phys. 19, 798 (1978)
62. V.S. Gerdjikov, M.I. Ivanov, The quadratic bundle of general form and the nonlinear evolution
equations. Bulg. J. Phys. 10, 130 (1983)
63. H.H. Chen, Y.C. Lee, C.S. Liu, Integrability of nonlinear Hamiltonian systems by inverse
scattering method. Phys. Scr. 20, 490 (1979)
64. F. Baronio, S. Chen, P. Grelu, S. Wabnitz, M. Conforti, Baseband modulation instability as the
origin of rogue waves. Phys. Rev. A 91, 033804 (2015)
65. S. Chen, F. Baronio, J.M. Soto-Crespo, Y. Liu, and Ph. Grelu, Chirped Peregrine solitons in a
class of cubic-quintic nonlinear Schrödinger equations. Phy. Rev. E 93, 062202 (2016)
66. G.P. Agrawal, Nonlinear Fiber Optics, 4th ed. (Academic, 2007)
67. P. Marquié, J.M. Bilbault, M. Remoissenet, Generation of envelope and hole solitons in an
experimental transmission line. Phys. Rev. E 49, 828 (1994)
Chapter 5
Emission of Nonlinear Modulated Waves
in Multi-coupled Nonlinear Transmission
Networks

Abstract In the present chapter, we investigate the dynamics of nonlinear modu-


lated waves propagating along multi-coupled electric transmission networks. The
dynamics of slowly modulated waves propagating through different network sys-
tems are reduced to either two-dimensional higher order (HO) NLS equations or
generalized two-dimensional complex GL equations. The spatial wave solutions of
these amplitude equations are presented and their modulational instability is inves-
tigated. By means of perturbation approach, the transverse stability of spatial waves
propagating along the network systems under consideration are analyzed.

5.1 Transmission of Solitonlike Wave Signals in a


Two-Dimensional Lossless Dispersive Nonlinear
Transmission Network

This section deals with the dynamics of modulated waves in the two-dimensional
lossless network system of Fig. 5.2. The amplitude equation is reduced to a 2D NLS
equation without dissipative term. exact solitonlike solutions are presented and used
to study the transmission of solitonlike pulses through this network system.

5.1.1 Description of the and Network Equations

The physical system to be studied in this Section is a two-dimensional lossless electric


NLTN displayed in Fig. 5.2 [1]. It consists of M − 1 identical lines connected one
to another by linear capacitors C2 . Each single line is formed of N identical unit
cells such as that shown in Fig. 5.1a [1]. For investigating the characteristics of
modulated waves propagating through the multi-coupled network system of Fig. 5.2,
we introduce two discrete variables, one longitudinal variable, n, which specifies
the network nodes in the longitudinal direction which in this Section is considered
as the dominant direction of the wave propagation, and one transverse variable m,
which labels the position of lines in the transverse direction. Next, we denote by

© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2022 163
E. Kengne and W. Liu, Nonlinear Waves,
https://doi.org/10.1007/978-981-19-6744-3_5
164 5 Emission of Nonlinear Modulated Waves in Multi-coupled Nonlinear …

Fig. 5.1 a A schematic representation of the unit cell of a LC nonlinear transmission network with
dispersive element CS . The network is composed of N identical cells. b A schematic representation
of the unit cell of a nonlinear transmission electric network of the C type

Qn,m = Qn,m (Vn,m ) the electric charge at the (n, m)-th node, which is a function of
the voltage Vn,m across the capacitor of the n-th cell in line m. In our study, we
assume that wave propagate in both the longitudinal direction n and the transverse
direction m.
Applying Kirchhoff’s laws to the network system of Fig. 5.2 yields
 
∂ 2 Qn,m ∂2 1  
= CS 2 + Vn+1,m − 2Vn,m + Vn−1,m
∂t 2 ∂t L
∂2  
+ C2 2 Vn,m+1 + Vn,m−1 − 2Vn,m , (5.1)
∂t
where n = 1, 2, . . . , N , and m = 1, 2, . . . , M − 1. To derive the network equations
in terms of the propagating voltage Vn,m , we assume the following voltage-dependent
capacitance
  dQn,m C0
C Vn,m + V0 = = ,
dVn,m 1 + Vn,m V0

where V0 is the equilibrium


 voltage
 and C0 = C(V0 ). Under the condition |Vn,m | 
V0 , we expand C Vn,m + V0 into Taylor series and arrive to
   
C Vn,m + V0 = C0 1 − 2αV + 3βV 2 + · · · ; (5.2)

here, α and β are the nonlinearity coefficients which throughout this section are
assumed to be positive. Inserting now Eq. (5.2) into Eq. (5.1) yields the following
network equations
 
∂ 2 Vn,m ∂2  
+ 2
+ d0 2 2Vn,m − Vn−1,m − Vn+1,m
∂t 2 0
∂t
2  
∂ ∂2  2 
− γ0 2 Vn,m−1 − 2Vn,m + Vn,m+1 = 2 αVn,m − βVn,m
3
, (5.3)
∂t ∂t
5.1 Transmission of Solitonlike Wave Signals in a Two-Dimensional Lossless … 165

Fig. 5.2 A fragment of the system of a multi-coupled discrete nonlinear transmission electrical
network with dispersive elements coupled by linear capacitor C2 . The network is composed of
M − 1 coupled lines composed of N identical cells

where
1 C2 CS
0 = √ , γ0 = , d0 = (5.4)
C0 L C0 C0

are respectively the characteristic frequency of each line, the dimensionless coupling
parameter, and the dimensionless dispersive parameter.

5.1.2 Amplitude Equation for the Dynamics of Modulated


Waves in the Two-Dimensional Lossless Electric
Network

The main aim of this Section being the study of the dynamics of modulated waves in
the network system of Fig. 5.2 for long-wavelength transverse perturbations [3–7],
it is reasonable to start by finding the amplitude equation, that is, the continuum
equation that governs the dynamics of modulated waves along our network system.
Such continuum equation can be obtained with the use of the reductive perturbation
method in the semi-discrete limit. For this aim, we introduce the following three
slow continuous variables, x = (n − υg t), y = m, and τ =  2 t, where  is a
small parameter and υg is the longitudinal group velocity. Next, the solution of the
network equations (5.3) is sought in the general form
166 5 Emission of Nonlinear Modulated Waves in Multi-coupled Nonlinear …
 
Vn,m = A(x, y, τ ) exp (iθ ) +  2 φ(x, y, τ ) + B(x, y, τ ) exp (2iθ ) + c.c., (5.5)

with θ = kn + qm − ωt; here, k and q, the wavenumbers in respectively the n−


and m− directions and the angular frequency ω are assumed to satisfy the linear
dispersion relation

420 sin2 (k/2)


ω2 (k, q) = , (5.6)
1 + 4d0 sin2 (k/2) + 4γ0 sin2 (q/2)

from which the longitudinal group velocity υg is found to be


 2 
dω 0 − d0 ω2 sin k
υg = =  . (5.7)
dk ω 1 + 4d0 sin2 (k/2) + 4γ0 sin2 (q/2)

In Eq. (5.5), A, φ, and B are respectively the fundamental harmonic, the dc, and the
second harmonics.
It is clearly seen from Eq. (5.6) that the angular frequency ω decreases when
the dimensionless dispersive parameter γ0 increase. For a better understanding, we
depict in Fig. 5.3 the frequency curves versus the wavenumbers k and q for different
values of γ0 .
Under the above conditions and the following restriction


⎨ q ∈ 2π Z
(5.8)

⎩ f (k) = υg2 −20 cos k+d0 (4υg2 sin2 ( k2 )+4ωυg sin k+ω2 cos k )−γ0 ω2
20
= 0,

we arrive to the following two-dimensional NLS for the fundamental harmonic


A(x, y, τ )

Fig. 5.3 Angular frequency ω versus the longitudinal wavenumber k for q = π (a), and as a function
of wavenumber q for k = π (b), for γ0 = 0.1, 0.5, 1.5, 10. The coefficients 0 = 0.5 × 109 is used
here. Reprint from Ref. [2], Copyright 2022, with permission from American Physical Society
5.1 Transmission of Solitonlike Wave Signals in a Two-Dimensional Lossless … 167
 
∂A P ∂2 ∂2
i + + A + Q |A|2 A = 0, ; (5.9)
∂τ 2 ∂x2 ∂y2

where
4α 2 υ 2

ω
2 2

ωγ0 ω 3β + 2 −vg2 + 2 −d2α sin


( 0 0 ) k+ω2 2
P=− , Q=−  0 g  . (5.10)
1 + 4d0 sin2 (k/2) 2 1 + 4d0 sin2 (k/2)

During our calculations leading to the amplitude equation (5.9), the dc and the second
harmonics φ and B have been found to be

2αvg2 ω2 α
φ(x, y, τ ) = |A|2 , B(x, y, τ ) =   2 A2 .
υg2 − 20 d0 − 0 sin k + γ0 sin2 q − ω2
2

It is seen from Eq. (5.10) that dispersive coefficient P and the nonlinearity coeffi-
cient Q of the 2D NLS equation (5.9) are functions of the longitudinal wavenumber
k and take only negative values. It is important to note that the above amplitude
equation (5.9) is valid only for wavenumbers k and q that satisfy the condition (5.8).
Also, it is important to notice that equation f (k) = 0 may admit many real solutions
kl , as we can see from Fig. 5.4, so that the amplitude equation (5.9) should be valid
for all (k, q) = (kl , 2π Z).

5.1.3 Computational Simulations

As well as we know, the 2D NLS equation (5.9) under condition PQ > 0 admits the
following longitudinal bright soliton (that is, a soliton solution that depend on only
the time τ and the longitudinal variable x) [1]

Fig. 5.4 The plot of f (k) for


the transverse wavenumber
q = 2π and parameters
d0 = γ0 = 0.75,
0 = 0.5 × 109 . Reprint
from Ref. [2], Copyright
2022, with permission from
American Physical Society
168 5 Emission of Nonlinear Modulated Waves in Multi-coupled Nonlinear …
   
Amax exp iυe 2x − Pυp τ /4
A0 (x, τ ) = (5.11)
cosh [(2x − Pυe τ ) /2Se ]

in which υe and υp are respectively the envelope and phase velocities, Se =



P/Q/Amax is the spatial soliton’s extension, and Amax is the soliton amplitude.
The initial voltage for the numerical purpose can be obtained by perturbing the
bright soliton solution (5.11) as

Ap (x, y, τ ) = A0 (x, τ )
    
1 K K
+ δ  (x) exp i y + τ + ∗ (x) exp −i y + ∗ τ
2 Se Se
 
υe (2x − Pυe τ )
× exp i . (5.12)
4

Here, δ, K and  are respectively the perturbation amplitude, the perturbation


wavenumber and the perturbation complex angular frequency. From the linear sta-
bility analysis, we find the following special (, ):
⎛   ⎞
1 4QA2max 4 − K + υe P
2 2
(1 , ) = ⎝ , i ⎠, (5.13)
2 2x−Pυe τ 8
cosh 2Se
⎛ ⎞
1 2 υ 2
P − 4QA 2
K 2
(2 , ) = ⎝ − ,i e max ⎠, (5.14)
eτ 3 8
cosh2 2x−Pυ 2Se
⎛ ⎞
eτ  
sinh 2x−Pυ2Se 4QA2max 1 − K 2 + υe2 P
(3 , ) = ⎝ , i ⎠. (5.15)
cosh2 2x−Pυe τ
2Se
8

During the numerical/computational simulation, we use the following network


parameters:

C0 = 2 pF, CS = 1.5 pF, C2 = 1 pF, L = 1µH, α = 0.21 V−1 , β = 0.0197 V−2 ,

leading to the longitudinal wavenumber k = 1.42412.


To study analytically the evolution of bright soliton voltage through the network
system of Fig. 5.2, we use Eq. (5.12) with (, ) given by Eqs. (5.13)–(5.15) as
the fundamental harmonic, the input signal being obtained by setting n = 0 in Eq.
(5.12). Next, we go back to the original variables n, m, and t. Next, we specify values
of perturbation wavenumber K, complex frequency , and soliton parameters ve ,
vp , and Amax . The pulse then propagates in the dominant n-direction.
As an example of the computational simulations, we show in Fig. 5.5 the evolu-
tion of bright soliton along different network lines. Different plots of this Fig. 5.5,
generated with  = 0.005, Amax = 0.5, δ = 0.01, and K = 0.905, show the soliton
5.2 Coherent Structures for a Multi-coupled Nonlinear Transmission Network … 169

Fig. 5.5 Propagation of a bright soliton along different lines of the network. From the top to
bottom, the plots correspond to m = 1, m = 200, and m = 400, respectively. Reprint from Ref.
[2], Copyright 2022, with permission from American Physical Society

profile at time t = 100 µs. Plots of the top, middle, and bottom panels are obtained
with respectively (1 , ) , (2 , ) , and (3 , ) .
In conclusion, we have employed the reductive perturbation method in the semi-
discrete to reduce the dynamics of modulated waves propagating along the network
system of Fig. 5.2 to a two-dimensional NLS equation. Based on the longitudinal
bright soliton solution of this equation, we have carried out numerical simulations
to investigate the transmission of bright solitonlike waves in our network system.

5.2 Coherent Structures for a Multi-coupled Nonlinear


Transmission Network with Dissipative Elements

In the present Section, we treat the MI and the coherent structure for a multi-coupled
lossy mono-inductance transmission network. First, we show that the amplitude
equation for the dynamics of modulated waves propagating along this network system
is reduced to a system of generalized 2D CGL equation. Spatial wave solutions are
presented and their linear stability is investigated. Coherent structures for the found
2D CGL system are considered too, displaying numerical results for them.
170 5 Emission of Nonlinear Modulated Waves in Multi-coupled Nonlinear …

Fig. 5.6 a A schematic representation of an elementary cell of a one-dimensional RLC discrete


nonlinear transmission electric network, made of N identical cells. b A scheme of a unit cell of an
RC network, which is composed of M − 1 identical cells. Reprint from Ref. [14], Copyright 2022,
with permission from Springer

5.2.1 Introduction

Although single NLTNs have been studied intensively, just a few works are reported
for multi-coupled ones, either via the equations of the KdV type [7, 9] or the equa-
tions of complex Ginzburg-Landau type, including NLS equations [1, 2, 10]. These
works study theoretically and/or experimentally the dynamics of modulated non-
linear waves, including the cylindrical solitons, two-dimensional transmission net-
work. Employing various analytical methods, solitons and nonlinear resonance of
two-dimensional transmission lines including network with intraline resistances have
also been examined [11–13].
The physical system to be studied in this section is the multi-coupled lossy discrete
transmission electrical network depicted in Fig. 5.7 [14, 15]. This network system
consists of M − 1 identical lines coupled one to another by means of linear capacitors
C2 at each node. Each line of the system is formed of N identical cells such as that
shown in Fig. 5.6a. In this physical system, the nodes are labelled with the help of
two discrete variables, variables n and m that specify the nodes in respectively the
longitudinal direction and the transverse direction. We assume that waves propagate
in both the n- and the m-direction, the longitudinal direction being assumed to be the
dominant propagation direction. Next, we assume that the electric charge Qn,m at the
(n, m)-th node depends on the voltage Vn,m across the capacitor of the n-th cell in
line m. Moreover, we assume that the electric charge Qn,m is related to the nonlinear
voltage-dependent capacitance by the relationship [16]

C0
C(Vn,m ) =  p , (5.16)
Vn,m
1+ V0

where C0 and V0 are capacitance and voltage scales, the nonlinearity being repre-
sented by the real positive parameter p.
5.2 Coherent Structures for a Multi-coupled Nonlinear Transmission Network … 171

Fig. 5.7 A segment of a multi-coupled dissipative nonlinear transmission network, connected by


linear capacitor C2 . Reprint from Ref. [14], Copyright 2022, with permission from Springer

Applying the Kirchhoff’s laws to the network system of Fig. 5.7 yields

d 2 Qn,m 1   d2  
− Vn−1,m − 2Vn,m + Vn+1,m − C2 Vn,m−1 − 2Vn,m + Vn,m+1
dt 2 L1 dt 2
dVn,m R dQn RG RC2 d  
+G + + Vn,m − Vn,m−1 − 2Vn,m + Vn,m+1 = 0;
dt L1 dt L1 L1 dt
(5.17)

In Eq. (5.17), n = 1, 2, 3, . . . , N and m = 1, 2, 3, . . . , M − 1. Following the idea


of work [17], we apply the continuum limit to system (5.17) in both the n and m
directions, and treat the discrete variables n and m as continuous ones, setting n = x
and m = y so that Vn,m (t) = V (x, y, t). Under the condition that |V |  |V0 | , we
expand the voltage-dependent capacitance (5.16) as well as Vn±1,m and Vn,m±1 in the
Taylor series around V and use the network equations (5.17) to obtain the following
nonlinear partial differential equation:
 
∂2   1 ∂ 2V ∂ 4V RC0 ∂V RC0 ∂V 3
C0 V + bV 3
− − C 2 + G + +
∂t 2 L1 ∂x2 ∂t 2 ∂y2 L1 ∂t L1 ∂t
G RC2 ∂ 3 V
+R V − = 0, (5.18)
L1 L1 ∂t∂y2
 
where b ≡ −1/ 3V02 .
172 5 Emission of Nonlinear Modulated Waves in Multi-coupled Nonlinear …

5.2.2 Amplitude Equation for a Two-Dimensional Lossy


Network

To derive the amplitude continuum equation that governs the dynamics of modulated
waves in the network system of Fig. 5.7, we use the traditional perturbation technique.
For this purpose, we introduces two slow time scales T1 and T2 , and two slow
spatial variables, X1 and Y1 as follows: T1 = t and T2 =  2 t, X1 = x and Y1 = y.
The original variables x, y, and t are denoted respectively by X0 , Y0 , and T0 . Next,
we seek the solution of the network equations (5.18) in the following general form:
   
V =  1/2 u11 exp (iθ1 ) + v11 exp (iθ2 ) +  u22 exp (2iθ1 ) + v22 exp (2iθ2 )
   
+  3/2 u33 exp (2iθ1 ) + v33 exp (3iθ2 ) +  5/2 u42 exp (2iθ1 ) + v42 exp (2iθ2 )
+ c.c + · · · , (5.19)

where

θ1 = kX0 − ω1 T0 , θ2 = qY0 − ω2 T0 , ujk = ujk (X1 , Y1 , T1 , T2 ), vjk = vjk (X1 , Y1 , T1 , T2 ),

k and q are the wavenumbers in respectively the x− and the y− directions, with the
corresponding angular frequencies ω1 and ω2 . Assuming that G + RC0 /L1 = O ()
and RG/L1 ∼ 1 [18] and working with C2 = C0 /2, we obtain the following pair of
linear dispersion relations

k2 G   G
− C0 ω12 + + R = 0, C0 2 + q2 ω22 − 2R = 0, (5.20)
L1 L1 L1

and the following system of complex GL equations


 
∂u ∂ 2u ∂ 2u   ∂  2 
+ i P1 2 + P2 2 + γ1 u + D1 |u|2 + 2 |v|2 u + Q1
1 1 |u| + 2 |v|2 u = 0,
∂t ∂x ∂y ∂x
(5.21a)
 
∂v ∂ 2v ∂ 2v   ∂  2 
+ i P12 2 + P22 2 + γ2 v + D2 |v|2 + 2 |u|2 v + Q2 |v| + 2 |u|2 v = 0,
∂t ∂x ∂y ∂y
(5.21b)

if parameters k, q, ω1 , and ω2 are taken from the following conditions


 
8C2 L1 GR RG
k=− , ω1 = ,
1 − 4C0 L1 2C2 L1 (1 − 4C0 L1 )
C0 + 2q2  3
ω2 = , 2C0 L1 2 + q2 − RGq2 = 0. (5.22)
q

In the 2D complex GL system (5.21a)–(5.21b), we have set


5.2 Coherent Structures for a Multi-coupled Nonlinear Transmission Network … 173

√ −i k
√ qω
i 2 2 2 y
u11 e 2C0 L1 P1 ω1 , v = v11 e 2P2 (2+q ) ,
x
u=
1

k 2 − C0 L1 ω12 1 ω1 2 1
P11 = , P2 = , P1 = −   ,
2C02 L21 ω13 4 C0 L1 2 + q2 ω2
   
1 − q2 ω2 k R
P22 =  2 , Q 1 = − ω1 + i ,
2 + q2 4V02 C0 L1 ω12 L
 
5q R
Q2 =  2 ω 2 + i , (5.23)
2V02 2 + q2 L1
 
L1 G + RC0 k
γ1 = + iα − P11 α ,
2C0 L1 C0 L1 ω1
   
2L1 G + C0 R 2 − q2 qω2
γ2 =   − iβ P 2
β + ,
2C0 L1 2 + q2 2
2 + q2
   
1 R 1 R
D1 = iαQ1 − − iω1 , D2 = iβQ2 − 2   − iω2 .
2V02 L V0 2 + q2 L1


 2 
Because of the presence of the cubic derivative terms ∂x |u| + 2 |v|2 u and

 
∂y
|v|2 + 2 |u|2 v in Eqs. (5.21a)–(5.21b), this system is called generalized complex
GL system. In the following, any complex number f will be written in the canonical
form f ≡ f r + if i .
When u(x, y, t)v(x, y, t) = 0, let us say, for example v(x, y, t) = 0, the system
(5.21a)–(5.21b) turns into a single complex two-dimensional complex Ginzburg-
Landau equation with derivative terms:
   
∂u ∂ 2u ∂ 2u ∂u ∂u∗
+ i P11 2 + P21 2 + γ1 u + D1 |u|2 u + Q1 2 |u|2 + u2 = 0.
∂t ∂x ∂y ∂x ∂x
(5.24)

5.2.3 Modulational Instability in a Two-Dimensional Lossy


Transmission Network

We address here the modulational instability of spatial wave solutions of system


(5.21a)–(5.21b). The case of one-component complex GL equation (5.24) is also
investigated.

5.2.3.1 Modulational Instability for a Two-Component CGL System:


Zero Perturbation Wavenumber

First of all, we note that system (5.21a)–(5.21b) admits constant amplitude Stokes
wave solutions of the form
174 5 Emission of Nonlinear Modulated Waves in Multi-coupled Nonlinear …

u(x, y, t) = u0 exp i (k0 x + l0 y − ω0 t) , v(x, y, t) = v0 exp i (k0 x + l0 y − ω0 t) ,


(5.25)
where
 
  γ r Di + k0 Q1i
ω0 = − P11 k02 + P21 l02 − 1 r 1 ,
D1 − k0 Q1i
γr 2γ r
u02 =  r 2 −  r 1 , (5.26)
3 D2 − l0 Q2 i
3 D1 − k0 Q1i
γr 2γ2r
v02 =  r 1  −  ,
3 D1 − k0 Q1i 3 D2r − l0 Q2i

and k0 and l0 being the carrier wavenumbers to be taken from the condition
 2        
P1 − P11 k02 + P22 − P21 l02 D1r − k0 Q1i D2r − l0 Q2i − γ1r D1i + k0 Q1r
 
+ γ2r D2i + l0 Q2r = 0. (5.27)

Perturbing u and v as follow


 
u = u0 + u1 (x, y, t) ei(k0 x+l0 y−ω0 t) ,
 
v = v0 + v1 (x, y, t) ei(k0 x+l0 y−ω0 t) , (5.28)

and applying the linear stability analysis yield the following linear dispersion relation
for the perturbation
 r     √
= D1 − k0 Q1i u02 + D2r − l0 Q2i v02 i ± δ  , (5.29)

where
     2
 1 γ2r D1r − k0 Q1i γ1r D2r − l0 Q2i  r 
δ =− + − 2 γ1 + γ2
r
9 D2r − l0 Q2i D1r − k0 Q1i
      
4 γ2r D1r − k0 Q1i γ1r D2r − l0 Q2i
− − 2γ1
r
− 2γ2 .
r
(5.30)
3 D2r − l0 Q2i D1r − k0 Q1i

We can seen from Eq. (5.30) that for δ  to be non-positive, one must have
 r  
D1 − k0 Q1i D2r − l0 Q2i < 0. (5.31)

The case of δ  ≥ 0 If δ  ≥ 0, Eq. (5.29) leads to


    
Re (i) = Re(−i∗ ) = − D1r − k0 Q1i u02 + D2r − l0 Q2i v02 ,

and the above Stokes wave solutions will be stable under modulation if and only, if
5.2 Coherent Structures for a Multi-coupled Nonlinear Transmission Network … 175
 r   
D1 − k0 Q1i u02 + D2r − l0 Q2i v02 ≥ 0.

Thus, for the Stokes wave solutions (5.25)–(5.27) to be unstable under modulational
when δ  ≥ 0, it is necessary and sufficient that
 r   
D1 − k0 Q1i u02 + D2r − l0 Q2i v02 < 0.

The condition for the MI of Stokes wave solutions (5.25)–(5.27) in the case δ  ≥ 0
is then found to be
 2  2   
2γ2r D1r − k0 Q1i + 2γ1r D2r − l0 Q2i − (γ1 + γ2 ) D1r − k0 Q1i D2r − l0 Q2i < 0.

The case of negative δ  Considering now the case when δ  < 0, the angular frequency
of modulation  will be a pure imaginary number so that the Stokes wave solutions
(5.25)–(5.27) in the case of negative δ  will be unstable under modulation if the
following two conditions are satisfied
 r     
− D1 − k0 Q1i u02 + D2r − l0 Q2i v02 + −δ  > 0, (5.32)
 r    
D1 − k0 Q1i u02 + D2r − l0 Q2i v02 + −δ  < 0. (5.33)

The necessary condition for condition (5.33) to be satisfied is that


 r   
D1 − k0 Q1i u02 + D2r − l0 Q2i v02 < 0. (5.34)

Equation (5.32) under the condition (5.34) is always satisfied. As the conclusion, the
necessary and the sufficient conditions for the Stokes wave solutions (5.25)–(5.27)
to be unstable under modulation are respectively
 r   
D1 − k0 Q1i u02 + D2r − l0 Q2i v02 < 0, (5.35)
     
− D1r − k0 Q1i u02 + D2r − l0 Q2i v02 − −δ  > 0. (5.36)

Analyzing the above two cases leads to the following conclusion concerning the
MI.

Theorem I. If the carrier wavenumbers k0 and l0 and the carrier parameters u02 and
v02 satisfying respectively Eq. (5.27) and Eq. (5.26) verify the condition
     2
γ2r D1r − k0 Q1i γ1r D2r − l0 Q2i  r 
+ − 2 γ1 + γ2
r
D2r − l0 Q2i D1r − k0 Q1i
      
γ2r D1r − k0 Q1i γ1r D2r − l0 Q2i
+12 − 2γ1 r
− 2γ2r ≤ 0,
D2r − l0 Q2i D1r − k0 Q1i
176 5 Emission of Nonlinear Modulated Waves in Multi-coupled Nonlinear …

then the necessary and sufficient condition for the MI of Stokes wave solutions (5.25)–
(5.27) of system (5.21a)–(5.21b) is that
 2  2   
2γ2r D1r − k0 Q1i + 2γ1r D2r − l0 Q2i − (γ1 + γ2 ) D1r − k0 Q1i D2r − l0 Q2i < 0.
(5.37)

II. If the above the carrier wavenumbers k0 and l0 and the carrier parameters u02
and v02 satisfy the condition
     2
γ2r D1r − k0 Q1i γ1r D2r − l0 Q2i  r 
+ − 2 γ1 + γ2
r
D2r − l0 Q2i D1r − k0 Q1i
      
γ2r D1r − k0 Q1i γ1r D2r − l0 Q2i
+12 − 2γ1 r
− 2γ2r > 0,
D2r − l0 Q2i D1r − k0 Q1i

then the necessary and sufficient MI conditions for Stokes wave solutions (5.25)–
(5.27) of system (5.21a)–(5.21b) are respectively
 2  2   
2γ2r D1r − k0 Q1i + 2γ1r D2r − l0 Q2i − (γ1 + γ2 ) D1r − k0 Q1i D2r − l0 Q2i < 0,
(5.38)
 r  r 
D1 − k0 Q1 D2 − l0 Q2 < 0.
i i
(5.39)

5.2.3.2 Modulational Instability of One-Component Two-Dimensional


CGL Equation

The one-component GL equation (5.24) admits the continuous wave solution


 
u(x, y, t) = a0 exp i (k1 x + q1 y − ω1 t) ,

γ1r
a0 = , (5.40)
Q1i k1 − D1r
 
ω1 = − P11 k12 + P21 q12 + γ1i + D1i a12 + Q1r a12 k1 ;

with the carrier wavenumbers k1 and q1 , carrier amplitude a0 , and carrier frequency
ω1 . The MI of the CW solution (5.40) can be investigated with the help of the linear
stability analysis which consists of perturbing solution (5.40) by letting
 
u(x, y, t) = a0 + u1 (x, y, t) exp i (k1 x + q1 y − ω1 t) , (5.41)

u1 being the infinitesimal perturbation function. Before formulating the obtained


results, we introduce the following notations:
5.2 Coherent Structures for a Multi-coupled Nonlinear Transmission Network … 177
   
Z = D1r − k1 Q1i a02 + 2i a02 Q1r K − P11 k1 K − P21 q1 q , X = Re [c] , Y = Im [c] ,

with
  
c = −Z 2 + a04 Q1∗ K − k1 Q1∗ + iD1∗ (Q1 K + (k1 Q1 − iD1 ))
     
+ P11 K 2 + 2k1 K + P21 q2 + 2q1 q − (2Q1 K + k1 Q1 − iD1 ) a02
       
× P11 K 2 − 2k1 K + P21 q2 − 2q1 q + 2Q1∗ K − k1 Q1∗ − iD1∗ a02 .

Using these notations, the final result can be formulated in terms of the following
theorem.

Theorem For the CW solutions (5.40) of the one-component complex GL equation


(5.24) to be unstable under modulation, it is necessary that the carrier wavenumbers
k1 and q and the perturbation wavenumber K satisfy the inequality
  2 2
2 D1r − k1 Q1i a04 + 4 a02 Q1r K − P11 k1 K − P21 q1 q + a04 |Q1 |2 K 2 + P112 K 4 + P212 q4
 
+ 2k1 a04 Q1r D1i + a02 6k1 P11 Q1r K 2 + 8q1 P21 Q1r Kq − 2k1 P21 D1i q2 + k12 a04 Q1r2 + D1i2 a04
   
+2P11 P21 q2 K 2 − 4k1 q1 Kq − 2a02 K 2 P11 + P21 D1i > 0, (5.42)

and sufficient that


⎤  √ ⎡
 r  −X + X 2 + Y2
Re [Z] = D1 − k1 Q1i a02 ∈ ⎦−∞, − ⎣. (5.43)
2

5.2.4 Coherent Structures for a Two-Dimensional Lossy


Nonlinear Transmission Network

Following the idea developed in works [19, 20] on the coherent structures in one-
component complex GL equation, we seek traveling wave solutions of system
(5.21a)–(5.21b) in the following general form
 "   " 
u(x, y, t) = a(z) exp i F(z)dz − iω1coh t , v(x, y, t) = b(z) exp i G(z)dz − iω2coh t ,
(5.44)
where z = kx + qy − υcoh t is the propagating coordinate, F(z) and G(z) are the
local wavenumbers, υcoh is the propagating velocity, ω1coh and ω2coh are oscillating
frequencies, and k and q are the longitudinal and transverse wavenumbers. Injecting
ansatz (5.44) into system (5.21a)–(5.21b) yields
178 5 Emission of Nonlinear Modulated Waves in Multi-coupled Nonlinear …

da
= ξ,
dz
db
= η,
dz
dξ 1 #    
= aF 2 + 2 1 υcoh aF − D1i a3 + 2ab2 + ω1coh − γ1i a
dz k P + q P2
2 1
 1   $
−kQ1i 3a2 ξ + 2b2 ξ + 4abη − kQ1r a3 F + 2ab2 F , (5.45)
dη 1 #    
= bG 2 + 2 2 υcoh bG − D2i b3 + 2ba2 + ω2coh − γ2i b
dz k P + q P2
2 2
 1   $
−qQ2i 3b2 η + 2a2 η + 4abξ − qQ2r b3 G + 2ba2 G ,


dF X 1 ξ  
= −2F + 2 1 −υcoh + γ1r + D1r a2 + 2b2
dz a k P1 + q P2
2 1 a
  %
ξ  
+kQ1r 3aξ + 2b2 + 4bη − kQ1i a2 F + 2b2 F ,
a
dG η 1 & η  
= −2G + 2 2 −υcoh + γ2r + D2r b2 + 2a2 (5.46)
dz b k P1 + q P2
2 2 b
 η   
+qQ2r 3bη + 2a2 + 4aξ − qQ2i b2 F + 2a2 F .
b
Equations (5.46) feature formal singularities at a = 0 and b = 0 that can be
removed with the use of the blow-up (alias σ -process) transform [21] by setting

ξ η
= X; = Y. (5.47)
a b
The singularized (a, b, ξ, η, F, G) system (5.45)–(5.46) is then reduced the desin-
gularized (a, b, X , Y , F, G) system

da
= aX ,
dz
db
= bY ,
dz
dX 1 #  
= −X 2 + F 2 + 2 1 υcoh F − D1i a2 + 2b2 + ω1coh − γ1i
dz k P1 + q P2
2 1
 2   $
−kQ1 3a X + 2b2 X + 4b2 Y − kQ1r a2 F + 2b2 F ,
i

dY 1 #  
= −Y 2 + G 2 + 2 2 υcoh G − D2i b2 + 2a2 + ω2coh − γ2i
dz k P1 + q P2
2 2
 2   $
−qQ2 3b Y + 2aY + 4a2 X − qQ2r b2 G + 2a2 G ,
i
(5.48)
5.2 Coherent Structures for a Multi-coupled Nonlinear Transmission Network … 179

dF 1 #  
= −2FX + 2 1 −υcoh X + γ1r + D1r a2 + 2b2
dz k P + q P2
2 1
 2 1   $
+kQ1 3a X + 2b2 X + 4b2 Y − kQ1i a2 F + 2b2 F ,
r

dG 1 #  
= −2GY + 2 2 −υcoh Y + γ2r + D2r b2 + 2a2
dz k P +q P2 2
 2 1 2 2   $
+qQ2 3b Y + 2a Y + 4a2 X − qQ2i b2 F + 2a2 F .
r

The desingularized (a, b, X , Y , F, G) system (5.48) admits one invariant space,


namely a = 0, b = 0 on which system (5.48) reduce to the following two independent
(X , F) and (Y , G) systems,
'  
X  = −X 2 + k 2 P1 +q1
υcoh F + ω1coh − γ1i + F 2 ,
 ( 1 2)
2 P1
−1   (5.49)
F  = k 2 P11 + q2 P21 −υcoh X + γ1r − 2FX ,

and
'  −1  
Y  = −Y 2 + k 2 P12 + q2 P22 υcoh G + ω2coh − γ2i + G 2 ,
 −1   (5.50)
G  = k 2 P12 + q2 P22 −υcoh Y + γ2r − 2GY .
 
It is evident that 0, 0, X(, (
F, (
Y,G ( will be a fixed point (FP) of system (5.48) if and
   
only, if X (, (
F and ( ( are FPs of respectively of Eqs. (5.49) and (5.50), so that for
Y, G
 
the stability of the FP 0, 0, X (, (
F, (
Y,G ( of system (5.48) it is necessary and sufficient
   
that X (, (F and ( Y,G ( are stable FPs of systems (5.49) and (5.50), respectively.
Using
 the eigenvalues
  of the linearized systems associated with equilibrium points
X(, (
F and ( Y, G( , we find that if X ( > 0 and ( Y > 0 [X( < 0 or ( Y < 0], then the
 
equilibrium point 0, 0, X (, (
F, (
Y,G( of system (5.48) will be stable [unstable].
As an example of above analysis, we consider the stationary case where υcoh =
ω1coh = ω2coh = 0 and find
)  
( ( γ1i
X = X± = ± 21 |k 2 P1|γ+q
1|
2 P 1 − k 2 P 1 +q2 P 1 ,
2|
)  1 1 2

( γ
F =(F± = ± 21 |k 2 P1|γ+q
1|
i
2 P 1 + k 2 P 1 +q2 P 1 ,
1
1 | 2 1 2

)  (5.51)

( ( γ2i
Y = Y± = ± 21 |k 2 P2|γ+q2|
2 P 2 − k 2 P 2 +q2 P 2 ,
2|
)  1 1 2

(=G (± = ± 1 |γ2 | γ2 i
G 2 |k 2 P +q2 P |
2 2 + k P +q P
2 2 2 2 .
1 2 1 2

   
(+ , (
Therefore, 0, 0, X F± , ( (± is a stable FP, while 0, 0, X
Y+ , G (− , (
F± , ( (± is an
Y− , G
unstable FP of system (5.48).
180 5 Emission of Nonlinear Modulated Waves in Multi-coupled Nonlinear …

Fig. 5.8 Figures obtained for the stable FP (fixed point) (0, 0, 29462, 29462, 0.96825, 0.96825)
with initial conditions a(0) = b(0) = 0.0001, F(0) = G(0) = X (0) = Y (0) = 0.001. a, d Ampli-
tudes of u and v, respectively, as functions of z. b, e Plot of the local wavenumbers versus z. c, f F
and G curves in phase planes (a, F) and (b, G) , respectively. Reprint from Ref. [14], Copyright
2022, with permission from Springer

Fig. 5.9 Figures obtained for the unstable FP (0, 0, −29462, 29462, −0.96825, 0.96825) with
initial conditions a(0) = b(0) = 0.0001, F(0) = G(0) = X (0) = Y (0) = 0.001. a, d Amplitudes
of u and v versus z. b, e Plot of the local wavenumbers versus z. c, f F and G curves in the phase
plane (a, F) and (b, G) , respectively. Reprint from Ref. [14], Copyright 2022, with permission
from Springer
5.3 Soliton Signals in an Alternate Right-Handed … 181

In the stationary case υcoh = ω1coh = ω2coh = 0, we use the network parameters

C0 = 540pF, C2 = 270pF, L1 = 28μH, R = 105 , G = 10−4 −1 , b = 0.16V−1

to show in Figs. 5.8 and 5.9 the evolution of the wave amplitudes |u(z)| = a(z)
and |v(z)| = b(z), the local wavenumbers F = F(z) and G = G(z), and the local
wavenumbers F = F(a), and G = G(b) versus amplitudes a and b. In these figures,
a(z), b(z), F(z), and G(z) are numerical solutions of the initial-value problem for
the system (5.48). Different plots are generated with the numerical values of k 2 =
q2 = 10−7 .
In conclusion, we have presented the results for the modulated wavetrains in the
multi-coupled lossy network system of Fig. 5.7. To make this possible, we have
first reduced the dynamics of modulated waves propagating in this network system
to a system of generalized two-dimensional complex GL equations with two cubic
derivative terms [14]. The MI of Stokes wave for both the two-component and one-
component 2D complex GL system is investigated. The coherent structures generated
by the derived 2D complex GL system are considered too.

5.3 Soliton Signals in an Alternate Right-Handed and


Left-Handed Multi-coupled Lossy Nonlinear
Transmission Network

This Section deals with the analytical study of the dynamics of modulated waves in an
alternate right-handed and left-handed multi-coupled nonlinear discrete electrical lat-
tice with losses. The amplitude continuum equation is reduced to a two-dimensional
complex GL equation, which is used to study the MI phenomenon of the network
system under consideration. We show that the network system under consideration
can adopt purely right-handed, purely left-handed, or composite right-/left-handed.
Analytically, we establish that our network system may support purely backward
traveling waves, purely forward traveling waves, or composite backward/forward
traveling waves. Using both the exact and approximate solitonlike solutions of the
derived amplitude equation, we investigate analytically RH, LH, as well as RH/LH
behaviors of the system [22].

5.3.1 Description of the Model and Basic Equations

The physical model to be studied in this Section is that showed in Fig. 5.10c, formed
by M − 1 identical lines, transversely coupled to one another by linear capacitors C2 .
Each line of this system, as we can see from Fig. 5.10c, consists of N identical cells
of a Noguchi lossy electric network such as that depicted in Fig. 5.10a [23–25]. In
the physical system of Fig. 5.10c, we have used two discrete coordinates, n and m to
182 5 Emission of Nonlinear Modulated Waves in Multi-coupled Nonlinear …

label the nodes in respectively the longitudinal and transverse direction. Throughout
this Section, we assume that waves propagate in both the transverse and longitudinal
directions, the later direction being assumed as the dominant propagation direction.
To investigate the dynamics of modulated waves in the network system of
Fig. 5.10c, we assume that the electric charge at the (n, m)−th node is voltage-
varying and related to voltage Vn,m across the n−th capacitor of the m−th line by
the relationship [15, 26, 27]
  * *
Qn,m = C0 Vn,m − 2αVn,m
2
, *αVn,m *  C0 , (5.52)

C0 and α denoting respectively the characteristic capacitance and the nonlinear coef-
ficient of the electrical stored charge Qn,m . During the numerical computation, we
will mainly use the following network parameters [1, 28]

L = 200 µH,C0 = 370 pF,C2 = 100 pF,α = 0.21 V−1 . (5.53)

To derive the network equations, we apply Kirchhoff’s laws on the network of


Fig. 5.10c and eliminate the currents from the resulting equations [1, 2]. We then
arrive to the following system of ordinary difference-differential equations, known
as the network equations:

d 2 Vn,m d2  
2
+ λ S 2
2Vn,m − Vn−1,m − Vn+1,m
dt dt
dVn,m  
+ 2u0 σ + u02 2Vn,m − Vn−1,m − Vn+1,m
dt
d2   d 2 Vn,m
2
+ λ2 2 2Vn,m − Vn,m−1 − Vn,m+1 − α = 0. (5.54)
dt dt 2

Here, n = 1, 2, . . . , N , m = 1, 2, . . . , M − 1, u0 = √1
L1 C0
is the characteristic fre-
quency of the network, λS = CS
C0
is the dimensionless dispersive parameter, λ2 = CC20
is the dimensionless coupling parameter, and σ = 2uG0 C0 is the dimensionless con-
ductance.
To derive the amplitude continuum equation, we assume that voltage Vn,m is
small enough compared with its value V0 in the equilibrium state [28, 29]. Next, we
combine the continuum limit approximation with the standard perturbation theory
[29]. For this aim, we take into consideration the fact that the dominant propagation
direction is the longitudinal direction n and introduce the following slow variables,
1 1 3
ξ = ε 2 (n − υt) , η = ε 2 m, τ = ε 2 t, and Vn,m (t) = εu(ξ, η, τ ) where ε  1 and
υ is a real constant. Also, we assume that the modulated waves propagating along
our network system are sufficiently smooth. Under the above considerations, we find
that υ 2 = u02 , while u(ξ, η, τ ) satisfies the equation
5.3 Soliton Signals in an Alternate Right-Handed … 183

Fig. 5.10 a Schematic representation of one unit cell of the dissipative discrete nonlinear electrical
transmission network that consists of one linear inductance L shunted by one linear capacitance CS
(dispersive element) in the series branch and, one nonlinear capacitance C in the shunt branches
in parallel with one conductance G; b Schematic representation of the elementary cell of a 1D
series capacitor transmission network network; c Part of the system of the multi-coupled nonlinear
transmission network, coupled by a linear capacitor C2 . Reprint from Ref. [22], Copyright 2022,
with permission from Elsevier

∂u (1 + 12λS ) u02 ∂ 3 u λ2 u02 ∂ 3 u αu2 ∂u2


+ + + 0 + u0 σ1 u − +
λ0 = 0; (5.55)
∂τ 24υ ∂ξ 3 2υ ∂ξ ∂η 2 2υ ∂ξ

here, σ1 =  − 2 σ and +
3
λ0 is a constant of integration.

5.3.2 Complex Ginzburg-Landau Equation for the Dynamics


of Modulated Waves in a Multi-coupled Lossy Network

5.3.2.1 Two-Dimensional Complex GL Equation of Dynamics of


Damped Modulated Waves

In the present Section, we aim to investigate the dynamics of modulated waves with
angular frequency ω = ωp = 2π fp and wavenumbers k = kp and  = p . For this
purpose, two different slow times scales and two different slow spatial variables are
introduced as follows: T1 = ετ , T2 = ε2 τ, X1 = εξ, and Y1 = εη, where ε << 1 is
a small parameter, assumed to be the same as in the previous subsection. We denote
by T0 = τ , X0 = ξ, and Y0 = η the original scales. Next, we seek the solutions of
the network equations (5.55) as

+∞
,
u(X0 , X1 , Y0 , Y1 , T0 , T1 , T2 ) = εr ur (X0 , X1 , Y0 , Y1 , T0 , T1 , T2 ), (5.56)
r=1
184 5 Emission of Nonlinear Modulated Waves in Multi-coupled Nonlinear …

where ur (X0 , X1 , Y0 , Y1 , T0 , T1 , T2 ) are complex functions to be found. Asking that


Eq. (5.56) satisfies the network equations (5.55), we obtain for the determination of
u1 , u2 , and u3 , the following system of partial differential equations:

D0 u1 = 0, (5.57a)
∂u2
D0 u2 + D1 u1 + αu02 1 = 0, (5.57b)
∂X0
 2 
∂u1 ∂
D0 u3 + D1 u2 + D2 u1 + αu02 +2 (u1 u2 ) = 0, (5.57c)
∂X1 ∂X0

where D0 , D1 , and D2 are the differential operators defined as

∂ ∂3 (1 + 12λS ) u02 ∂ 3
D0 = 2υ + λ2 u02 + + 2u0 υσ1 , (5.58)
∂T0 ∂X0 ∂Y02 12 ∂X03
 
∂ (1 + 12λS ) u02 ∂ 3 ∂3 ∂3
D1 = 2υ + + λ 2 u 2
+ 2 ,
∂T1 4 ∂X02 ∂X1 0
∂Y02 ∂X1 ∂X0 ∂Y0 ∂Y1
(5.59)
∂ (1 + 12λS ) u02 ∂ 3 ∂3 ∂3
D2 = 2υ + + λ 2 u 2
+ 2λ 2 u 2
.
∂T2 4 ∂X0 ∂X12 0
∂X0 ∂Y12 0
∂X1 ∂Y0 ∂Y1
(5.60)

Next, we impose to u1 to be of the form


 
u1 = exp [−χ n] ψ exp [i (kX0 + Y0 − ωT0 )] + ψ ∗ exp [−i (kX0 + Y0 − ωT0 )] ,
(5.61)
where ψ = ψ(X , Y1 , T1 , T2 ) is a complex function with the complex conjugate ψ ∗ ,
and χ is a positive parameter that accounts for the dissipation. In the following, χ
will be referred to as the linear dissipation parameter (alias spatial linear gain).
Asking that Eq. (5.61) satisfies Eq. (5.57a), we obtain the complex linear dispersion
relation
     
u0 σ υχ 2 λ2 u02 χ u02 1 + 12λS χ 3
3 −i ω−i 3 −i k +i 1 −i k +i 1 = 0,
ε2 ε2 2υ ε2 2υ 12 ε2
(5.62)
that leads respectively to the following linear dispersion relation and the equation
for the linear dissipation parameter χ
 
u02 3 1 + 12λS 2 3σ υ 1
ω+ k − 2 λ2 − k − = 0, (5.63)
υ  3 u0 χ
42 λ2 − 8 + ε (12λS + 1) k 2 24σ υ
χ3 − 3 χ− =0 (5.64)
1 + 12λS u0 (1 + 12λS )
5.3 Soliton Signals in an Alternate Right-Handed … 185

Setting υ = u0 in Eq. (5.64) and solving the resulting equation for positive χ yield

⎡ ) ⎤ 13
( 4λ2 2 −8+(12λS +1)k 2 )
3

⎢ 12σ + 144σ 2 − (1+12λS ) ⎥


χ =⎢



1 + 12λS

⎡ ) ⎤ 13
( 4λ2 2 −8+(12λS +1)k 2 )
3

⎢ 12σ − 144σ 2 − (1+12λS ) ⎥


+⎢

⎥ ,
⎦ (5.65)
1 + 12λS

if
     
4+ 3
18σ 2 (1 + 12λS ) 4 2 − ελ2 2 + 3 144σ 2 (1 + 12λS )
0≤< ,0≤k<
2λ2  (1 + 12λS )
= kmax (), (5.66)


42 λ2 − 8 +  (12λS + 1) k 2
χ =2
1 + 12λS
⎡ ⎤

cos ⎣  ⎦
 / 42 ελ2 −8+ε(12λS +1)k 2 , (5.67)
4 λ2 − 8 +  (12λS + 1) k
2 2
1+12λS

  
4 2 − λ2 2 + 3 144σ 2 (1 + 12λS )
if < k 2, (5.68)
 (1 + 12λS )

and

6 3 18σ 2 (1 + 12λS ) 24σ
χ = χc > 0, with χc3 − χc ∓ = 0, (5.69)
1 + 12λS 1 + 12λS

if    3
144σ 2 (1 + 12λS ) + 4 2 − λ2 2 −  (12λS + 1) k 2 = 0. (5.70)

Here, k must satisfy the conditions


  

3 8 + 3 144σ 2 (1 + 12λS ) − (1 + 12λS ) χc2 8+ 3
144σ 2 (1 + 12λS )
<k ≤ 2
.
2 (1 + 12λS )  (12λS + 1)
(5.71)
186 5 Emission of Nonlinear Modulated Waves in Multi-coupled Nonlinear …
)  √ 
2 8+ 3 144σ 2 (1+12λS )
If 3(1+12λS )
< χc [see Eq. (5.69)], then χc will be called “critical linear
dissipation parameter”, and any wavenumbers k and  which simultaneously satisfy
Eqs. (5.70) and (5.71) will be denoted kc and c and will be referred to as the “critical
wavenumbers”. To show the effects of the dispersive parameter λS , the transverse
wavenumber , and the dimensionless conductance σ on the linear waves propagating
along the network system of Fig. 5.10c, we depict in Fig. 5.11 the variation of the
propagating frequency fp (top panels) and the linear dissipation parameter χ (bottom
panels) versus the longitudinal wavenumber k for different values λS , σ, and .
In our investigation, the following definitions will be used:
Definition: We say that the network system of Fig. 5.10c is right-handed [left-
handed] in a given region if ∂fp /∂k > 0 [∂fp /∂k < 0] in this region.
Definition: The network system of Fig. 5.10c is said to adopt purely right-handed
[left-handed] behavior or composite right-/left-handed behaviors if either ∂fp /∂k > 0
[∂fp /∂k < 0] in the whole region of k or ∂fp /∂k changes its sign in the region of
validity of k.
Definition: The terminology “modulated waves in RH [LH] region” will be used for
all modulated waves that propagate in right-handed [left-handed] networks.
It is important to note that for a given set of network parameters, the network
system of Fig. 5.10c may adopt an alternate right-handed and left-handed behavior.
This feature is well seen from plots of Fig. 5.11. The size of region (in term of
k) where the network may adopt either the right-handed, the left-handed, or the
composite right-/left-handed behaviors depends, as we can see from plots of Fig. 5.11,
on the value of the dispersive parameter λS , the dimensionless capacitance σ, and
the transverse wavenumber . Plots of Fig. 5.11 also reveal that the linear dissipation
parameter χ increases when either λS or  increases.
Taking into account the above found results and eliminating from Eqs. (5.57b) and
(5.57c) terms that may produce secular terms, we arrive, after letting κ = k + iε− 2 χ ,
1

 = ω − iε− 2 υχ , to
3

 
u2 = exp [−2χ n] N1 |ψ|2 + N2 ψ 2 exp [2i (kX0 + Y0 − ωT0 )] + c.c., (5.72)
∂ψ ∂ ψ ∂ ψ
2 2
∂ ψ 2
i + Pn 2 + Pm 2 + Pnm + Q |ψ|2 ψ = 0, (5.73)
∂τ ∂ξ ∂η ∂ξ ∂η

where

u02 1 + 12λS   1 ∂ 2
k + iε− 2 χ =
1
Pn = − , (5.74a)
2υ 4 2 ∂κ 2
u2   1 ∂ 2
Pm = − 0 λ2 k + iε− 2 χ =
1
, (5.74b)
2υ 2 ∂2
λ2 u02 ∂ 2
Pnm = − = , (5.74c)
υ ∂κ∂
5.3 Soliton Signals in an Alternate Right-Handed … 187

Fig. 5.11 Effects of the dispersive parameter λS on the propagating frequency fp and the linear
dissipation parameter χ. Top panels: Propagating frequency fp [in kHz] versus the the longitudinal
wavenumber k [in rad/cell] according to Eq. (5.63) for the network parameters ( 5.53), ε = 10−2 ,
and different values of: a λS , λS = 0 (1), λS = 0.1 (2), and λS = 0.2 (3) with σ = 0.461 × 10−2
and  = 0; b σ, σ = 0.461 × 10−2 (1), σ = 0.05 (2), and σ = 0.1 (3) with λS = 0.2 and  = 0; c
transverse wavenumber ,  = 0 (1),  = 1 (2), and  = 2 (3) with σ = 0.461 × 10−2 and λS = 0.2.
Bottom panels, d, e Variation of the linear dissipation parameter χ [in cell−1 ] given by Eq. (5.65)
versus the longitudinal wavenumber k [in rad/cell] for the network parameters (5.53) with ε = 10−2
and different values of (d): transverse wavenumber ,  = 0 (1),  = 1 (2), and  = 2 (3) with
σ = 0.461 × 10−2 and λS = 0.2; e parameter λS , λS = 0 (1), λS = 0.1 (2), and λS = 0.2 (3) with
σ = 0.461 × 10−2 and  = 0.1. f Behavior of the propagating frequency fp in the limit k → 0 for
very small transverse wavenumber  (that is, for  = 0 or  → 0). Reprint from Ref. [22], Copyright
2022, with permission from Elsevier

u02  
k + 3iε− 2 χ (2N1 + N2 ) exp [−2χ n] ,
1
Q = −αε2 (5.74d)
υ
with N1 and N2 given as

3εαu02 χ
N1 = , (5.75)
3υ (u0 σ − 2χ ) − (1 + 12λS ) u02 χ 3
 1 
4εαu0 iε 2 k − χ
N2 =  3 . (5.76)
1
4υσ + u0 (1 + 12λS ) χ − iε 2 k

As we can see from Eqs. (5.74a)–(5.74d), the imaginary parts Im(Pn ) = PnI and
Im(Pm ) = PmI of Pn and Pm are negative for υ = u0 ; this is clearly seen from plots
of Fig. 5.12. In the rest of this Section, we use υ = u0 and expression (5.65)–(5.66)
for χ . As we can see from plots of Fig. 5.13, the imaginary part Im(Q) = QI may
take any sign, positive, zero, or negative. When Im(Q) > 0, the imaginary parts PnI
188 5 Emission of Nonlinear Modulated Waves in Multi-coupled Nonlinear …

Fig. 5.12 Effects of parameters  and λS on the imaginary parts Im[Pn ] and Im[Pm ] of the dispersion
parameters Pn and Pm . Plots of dispersive damped coefficients Im[Pn ] (top panels) and Im[Pm ]
(bottom panels) [rad/s] of the two-dimensional complex GL equation ( 5.73) versus the longitudinal
wavenumber k for the network parameters ( 5.53) with ε = 10−2 and different value of a, c transverse
wavenumber ,  = 0 (1),  = 1 (2), and  = 2 (3) with σ = 0.461 × 10−2 and λS = 0.2; b, d
dispersive parameter λS , λS = 0 (1), λS = 0.1 (2), and λS = 0.2 (3) with σ = 0.461 × 10−2 and
 = 0.1. Reprint from Ref. [22], Copyright 2022, with permission from Elsevier

and PmI and QI will designate respectively the dispersion dissipation coefficients and
the nonlinear dissipation coefficient.
It should be noted that the mixed derivative term with coefficient Pnm can be
removed from Eq. (5.73) if the condition PPmR
nR
= PPmInI > 0 is satisfied, with the use of
the following characteristics of Eq. (5.73)
√ √ √ √
2a 2b 2a 2b
ξ= ξ− η and +
η= ξ+ η,
2 2 2 2
a2
where a and b are any two real parameters satisfying the condition b2
= PmR
PnR
.

5.3.2.2 Dissipative Effects on the Linear Waves

To study the dissipative effects on the linear waves propagating along the network
system of Fig. 5.10c, we consider linear waves with initial amplitude εVinit given as
1
Vn,m (t) = εVinit exp [−χ n] exp iε 2 {kn + m − (u0 k + ωε) t} + c.c. (5.77)
5.3 Soliton Signals in an Alternate Right-Handed … 189

 
Fig. 5.13 Effects of parameters , σ , and λS on the imaginary part Im Q of the nonlinearity
 
parameter Q of the two-dimensional complex GL equation (5.73). Plots of Im Q (rad/s/ V2 )
versus k for the network parameters (5.53) with ε = 10−2 and different values of a the transverse
wavenumber ,  = 0 (1),  = 1 (2), and  = 2 with σ = 0.1 and λS = 0.2; b the dimensionless
conductance σ , σ = 0.461 × 10−2 (1), σ = 0.01 (2), and σ = 0.1 with  = 1 and λS = 0.2; c
the dispersive parameter λS , λS = 0 (1), λS = 0.1 (2), and λS = 0.2 with σ = 0.461 × 10−2 and
 = 2. To generate different plots, we have used n = 2. Reprint from Ref. [22], Copyright 2022,
with permission from Elsevier

The propagating frequency of such a plane wave is found to be


1
f = fp = ε 2 (u0 k + ωε) /2π.

To investigate the dissipative effects on the linear waves (5.77), we have used
the network parameters (5.53) to display the evolution of the voltage Vn,m (t) in
Fig. 5.14 for given network line m as a function of the cell number n. Different plots
of Fig. 5.14 prove that the motion of linear waves in the network under consideration
can be observed only for small values of the dimensionless conductance σ such as
the experimental one [28]. Since the dissipation in the network enhances when either
CS of  increases, the lifetime of the linear waves decreases as at least one of these
two parameters increases.

5.3.3 Linear Stability Analysis: Modulational Instability of


Stokes Waves

In this subsection , we study the phenomenon of the modulational instability for the
network system of Fig. 5.10c. This phenomenon finds its origins in hydrodynamics
and can occurs in many other branches of nonlinear sciences such as biophysics
[30], nonlinear optics [31], plasma systems, electrical transmission networks, Bose–
Einstein condensates [21, 21, 32–37], and others.
First of all, we note that the amplitude equation (5.73) admits the CW solution of
the form
ψ(ξ, η, τ ) = ψ0 exp [i (λξ + μη − ντ )] , (5.78)

where λ, μ, ν, and ψ0 are all real parameters and satisfy the relationships
190 5 Emission of Nonlinear Modulated Waves in Multi-coupled Nonlinear …

Fig. 5.14 Effect of the dimensionless conductance σ on linear waves propagating through our
network system: Propagation at time t = 1s of a linear wave in n− direction of line m = 5 with
initial amplitude 1 V generating a for λS = 0.2 and different values of the transverse wavenumber
,  = 0,  = 1, and  = 2 leading to the propagating frequencies fp = 628.5 kHz, 629.4 kHz,
632.8 kHz, respectively; b for  = 5 and different values of the dispersive parameter λS , λS = 0,
λS = 0.07, and λS = 0.1, corresponding respectively to the propagating frequency fp = 875.4 kHz,
756.1 kHz, 705.8 kHz; c for λS = 0.05,  = 0.1, and three values of the propagating frequency fp ;
(d): for  = 5, λS = 0.1, fp = 705.8 kHz, and three values of the dimensionless conductance σ .
Plots (a), (b), and (c) are obtained with the dimensionless conductance σ = 0.461 × 10−2 . Different
plots are generated with the network parameter (5.53). Reprint from Ref. [22], Copyright 2022,
with permission from Elsevier

PmI 2 QI 2
ν = PnR λ2 + PmR μ2 + Pnm μλ − QR ψ02 and λ2 + μ − ψ = 0. (5.79)
PnI PnI 0

To investigate the linear stability of the plane wave solution (5.78), we can look
for a solution of Eq. (5.73) by perturbing the solution (5.78) as follows

ψ(ξ, η, τ ) = [ψ0 + δψ(ξ, η, τ )] exp [i (λξ + μη − ντ + ϑ(ξ, η, τ ))] , (5.80)

where δψ and ϑ are perturbation function and are assumed to be small in comparison
with the parameter of the carrier wave. If we insert Eq. (5.80) into Eq. (5.73) and
linearize the resulting equation with respect with both δψ and ϑ, we obtain a linear
differential system with constant coefficients (that is not presented here) in δψ and
ϑ. Applying the linear stability analysis to this system and denoting by K and  and
 the perturbation wavenumbers and frequency, we obtain, under the condition

PnI 2 PmI 2
K +  ≤ ψ02 , (5.81)
QI QI

the following MI gain


5.3 Soliton Signals in an Alternate Right-Handed … 191
 √
  −a + a2 + b2
(K, ) = ± (K, ) = 2 PnI K + PmI  − QI ψ0 ±
2 2 2
,
2
(5.82)
where parameters a and b are defined as
 2  
a = PnR K 2 + PmR 2 − 2QR ψ02 PnR K 2 + PmR 2
− QI2 ψ04 − 4 (λPnI K + μPmI )2 , (5.83)

b =4 λPnR PnI K + μPnR PmI K  + λPnI PmR K + μPmR PmI 
3 2 2 3

−QR ψ02 (λPnI K + μPmI ) . (5.84)

In conclusion, for the Stokes wave solution (5.78)–(5.79) to be unstable under mod-
ulation, the following conditions must simultaneously satisfy

 
PnI QI > 0, PmI QI > 0, PnR QR > 0, PmR QR > 0, ∀k ∈ kcr1 , kcr . (5.85)

To show the effects of the dispersive parameter λS on the MI, we use the network
parameters (5.53) with the experimental value of σ = 0.461 × 10−2 to show the
evolution of the MI growth rate (K, ) given by Eq. (5.82) in Fig. 5.15. Figure
5.15a shows that the MI growth rate decreases and the region of the validity of K
shrinks as λS increases. This result is also confirmed by Fig. 5.15b–d where the
growth rate of MI is plotted versus the wavenumbers K and  of the perturbation
for different values of the dispersive parameter λS .

5.3.4 Dissipative Effects on Electrical Modulated Wave


Propagation

Here, we turn to both the analytically and numerically studies of effects of the dissi-
pation on the motion of the electrical modulated waves in the multi-coupled electrical
network system of Fig. 5.10c. For this aim, we first find analytical and approximate
solitonlike solutions of the two-dimensional coupled GL equation (5.73).

5.3.4.1 Analytical Solitonlike Solutions of the Amplitude Equation

Due to a fine balance between nonlinearity, dispersion, gain, and loss of energy
[39–41], it is possible to build exact solitonlike solutions of the 2D dissipative GL
equation (5.73). For this aim, we look for exact traveling wave solutions of Eq. (5.73)
in the form,
QR
ψ(ξ, η, τ ) = ρ(ζ ) exp [i {α1 ξ + β1 η − μs τ }] , (5.86)
QI
192 5 Emission of Nonlinear Modulated Waves in Multi-coupled Nonlinear …

Fig. 5.15 a Modulational instability gain (K, ) versus the wavenumber K of the perturbation
for  = 0.1 and three values of the dispersive parameter λS . b–d Evolution of the MI growth rate
(K, ) versus K and  for respectively λS = 0.1, λS = 0.2, and λS = 0.3; the instability features
show the famous donut Benjamin–Feir instability region [38]. For generating different plots, we
have used the network parameters (5.53) with σ = 0.461 × 10−2 , ε = 10−2 ,  = 0, k = 2, and
the gain + (K, ) given in Eq. (5.82). The parameters of the carrier wave are taken as λ = 0.1,
μ = 0.2, and ψ0 = 1. Reprint from Ref. [22], Copyright 2022, with permission from Elsevier

where ζ = α0 ξ + β0 η − υ0 τ is the effective propagation distance, ρ(ζ ) is the wave


amplitude which varies as a function of the propagation distance, α0 , β0 , α1 , β1 , μs ,
and υ0 are six real constants. Asking that Eq. (5.86) satisfies the 2D GL equation
(5.73), we arrive to the following two ordinary differential equations for the wave
amplitude ρ(ζ ) :

  d 2ρ dρ
PnR α02 + PmR β02 + Pnm α0 β0 − 2 (PnI α0 α1 + PmI β0 β1 )
dζ 2 dζ
  Q 3
+ μs − PnR α12 − Pnm α1 β1 − PmR β12 ρ + R2 ρ 3 = 0, (5.87)
QI
QR   d 2
ρ
PnI α02 + PmI β02
QI dζ2
QR dρ
+ [2PnR α0 α1 + 2PmR β0 β1 + Pnm (α1 β0 + α0 β1 ) − υ0 ]
QI dζ
5.3 Soliton Signals in an Alternate Right-Handed … 193

QR   Q3
− PnI α12 + PmI β12 ρ + R2 ρ 3 = 0. (5.88)
QI QI

Bright solitonlike solutions To find bright solitonlike solution, we equate to zero


the coefficients of the first derivative d ρ/d ζ in system (5.87)–(5.88) and impose to
the resulting equations to coincide. Under the condition QI 1 < 0, we obtain
⎡ 0 ⎤
1 α12
0 1 λ + 1+12λ S
1   ⎢ β1 1 1 2 4 β12 ⎥
1 α12 5 + cosh ⎣ α0 2 2 β 2
2 ζ⎦
1 u0 λ2 + 1+12λ S 1+12λ
λ2 02 + 4 S
QI 1 1 4 β12 α0
(1
ρ(ζ ) = β exp [χn] 2− 1
χ  ,
Q1R Q1I ε 2 1+12λ
1 λ2 β1 + 4 α1
2 S 2
−1 + 7 cosh2 2 λ β 2 + 1+12λS α 2 ζ
2 0 4 0

(5.89)
(1 = ±β1 . The corresponding bright solitonlike solution of the equation (5.73)
where β
is
0  
1
1 4λ22 β02
1 u0 λ2 + 1+12λ
1 S α 2
ψ(ξ, η, τ ) = β (1 exp [χ n] 2− 0
χ
1
Q1I ε 2
⎡ 0 ⎤
1
1 4λ2 β0
2
⎢ 1 1+ 1+12λ   ⎥
⎢ 1 0 ξ + β0 η − υ0 τ ⎥
S α2
5 + cosh2 ⎢β1 1 21 2 α α ⎥
⎣ 2 β0 1+12λS 0 0 ⎦
+ 4λ   %
α02 2
4λ2 β0 μs
× ⎡ 0 ⎤ exp iβ1 − ξ +η− τ ,
1 1 + 12λS α0 β1
1 4λ2 β0
2
⎢ 1 1+ 1+12λ  ⎥
⎢ 1 1 S α2 β υ ⎥
−1 + 7 cosh2 ⎢β1 1 2 2 0 ξ + 0η− 0τ ⎥
α0 α0
⎣ 2 β0 1+12λS ⎦
+ 4λ
α02 2

(5.90)

where β1 = 0 is an arbitrary real parameter and υ0 /α0 , μs /β1 , and β0 /α0 are defined
as
    
υ0 4λ2 β02 4k β0 4λ22 β02 1 + 12λS
= −λ2 u0 β1  1 − + − ,
α0 1 + 12λS α 2 1 + 12λS α0 1 + 12λS α02 4
0
(5.91)
   
μs u0 β0 8λ22 4λ22 β02 χ QR1
= β1 + λ2 + √ −k , (5.92)
β1 2 1 + 12λS α0 1 + 12λS α 2 ε QI 1
0
α1 4λ2 β0
=− , (5.93)
β1 1 + 12λS α0
)  2
± 2 − 1+12λS √Q1R χ − k
β0 4λ 2 εQ1I
= , (5.94)
α0 χ √QR1 − k
εQI 1

with
194 5 Emission of Nonlinear Modulated Waves in Multi-coupled Nonlinear …
   
Q1R = Re Q exp [2χ n] , QI 1 = Im Q exp [2χ n] .

The necessary condition for equation (5.94) to have a physical sense is that  = 0.
Under conditions (5.91)–(5.94), equations (5.87) and (5.88) are the same, and takes
the common form

λ2 β12 + α1
1+12λS 2 1
d 2ρ 2ε 2 exp [−2χ n] 2
Q1R
= 4
ρ +   ρ3. (5.95)
dζ2 λ2 β02 + α0
1+12λS 2
4
u0 χ λ2 β02 + 4 S α02 Q1I
1+12λ

Under the condition Q1I < 0, Eq. (5.95) admits various types of solitonlike solutions
(5.89).
Kink solitonlike solutions Kink solitonlike solutions of the two-dimensional com-
plex GL equation (5.73) can be found by the direct method, letting


= A1 ρ + A2 ρ 2 . (5.96)

Asking now that Eq. (5.96) satisfies system (5.87)–(5.88), we find that
β β1 1+12λS
λ2 α0 + χ
A1 = 2α1
3α0 β02
0 α1
β
4
1+12λS

ε
,
λ2 k +2λ2  α0 + k
α02 0 4

QR Q2
A22 = 
β 2
β 1+12λ
 R2 ,
QI
(5.97)
u0 α02 λ2 k 02 +2λ2  α0 + 4 S k
α0 0

 
υ0 β0 β1 β0 β1
= 2PnR + Pnm + + 2PmR
α1 α0 α0 α1 α0 α1
  
β02 β0 β1
1+12λS
4
+ λ 2 α2
1+12λS
4
+ λ 2 α0 α1
+2  0
 χ 2, (5.98)
β02 β0
ε λ2 k α2 + 2λ2  α0 + 4 k 1+12λS

  2
0

μs QR β1 β1 QR
= PmR − PmI + Pnm + PnR − PnI , (5.99)
α1
2 QI α1 2 α1 QI

β0 β1
where α0
and α1
are solutions of the following algebraic system of equations
  2
QR β0 β0 QR
PmR − PmI + Pnm + PnR − PnI = 0, (5.100)
QI α0
2 α0 QI
⎛ ⎞
Q 2
⎝9PmI R + 8P 2
β0 ⎠ β1
2 2
16PmI PnI β0 β1
mI
+
PnI + PmI 02 α0 α1 PmI 02 + PnI α0 α1
QI2 β 2 2 2 β2
α0 α0
5.3 Soliton Signals in an Alternate Right-Handed … 195

QR2 8PnI2
+ 9PnI 2
+ β 2 = 0, (5.101)
QI PmI α02 + PnI
0

respectively. Now, integrating equation (5.96) yields

1 1
ρ (ζ ) = , (5.102)
A0 exp [−A1 ζ ] − 1 A2
A0 A1

where A0is a constant of integration to be taken from the condition A0 A1 A2 < 0. For
QR = Re Q > 0, a sufficient condition for A22 given by Eq. (5.97) to be positive is
that the wavenumbers k and  satisfy the additional condition

1 + 12λS 2
λ2 2 − k ≤ 0. (5.103)
4
Now, inserting Eq. (5.102) into Eq. (5.86), we obtain the following kink solitonlike
solution of the two-dimensional complex GL equation (5.73)

QR 1 1
ψ(ξ, η, τ ) = exp [i {α1 ξ + β1 η − μs τ }] ,
QI A0 exp [−A1 (α0 ξ + β0 η − υ0 τ )] − 1 A2
A0 A1
(5.104)
where A1 and A2 are real constants given by Eq. (5.97), A0 is an arbitrary real
constant to be taken from condition A0 A1 A2 < 0, υ0 and μs are real parameters given
respectively by Eqs. (5.98) and (5.99), and α0 , β0 , α1 , and β1 are determined through
β0 /α0 and β1 /α1 , solution of the algebraic system of equations given respectively
by Eqs. (5.100) and (5.101).

5.3.4.2 Approximate Solitonlike Solution of the Amplitude Equation

Here, we focus our attention to the approximate solitonlike solutions of the 2D


complex GL equation (5.73). For simplicity, we assume that  = 0 and build the
approximate bright soliton solutions with vanishing CW background. The first step
of building such approximate solution is to reduce the 2D Eq. (5.73) to the following
one-dimensional complex GL equation

∂U 1 ∂ 2U
i + + |U |2 U = i(U ). (5.105)
∂T 2 ∂X 2
Here,
 
X = X (ξ, η) = α2 ξ + β2 η, T = T (τ ) = 2τ PnR α22 + PmR β22 ,

QR 1 χ ∂ 2u QI
U (X , T ) =   ψ(ξ, η, τ ), (U ) = − √ − |u|2 u,
2 PnR α2 + PmR β2
2 2 2 k ε ∂X 2 QR
196 5 Emission of Nonlinear Modulated Waves in Multi-coupled Nonlinear …

and α2 and β2 are two real constants to be taken from the condition α2 β2 = 0. In
the following, we assume that |(U )|  1 (near-integrable condition). Under this
condition, we seek the approximate bright solitonlike solution of Eq. (5.105) in the
form [42, 43]

U (X , T ) = As (T )φ(Z) exp [i (ks (T )X − φ0 (T ))] ; (5.106)

here, Z = As (T ) [X − X0 (T )] , and As (T ), φ(Z), ks (T ), φ0 (T ), and X0 (T ) are real


functions of variable T to be determined. Asking that Eq. (5.106) satisfies Eq. (5.105),
we arrive to the following system of equations
      
A2s d 2φ QI 3 1 dAs 1 1 dAs dX0 dφ
d1 2 + 2 φ + − d1 ks2 φ + Z − As − ks = 0,
2 dZ QR As dT 2 As dT dT dZ
(5.107)
   
A2s d 2φ 1 dks dks d φ0 1 dφ
+ 2φ 3 − Z + X0 − + ks2 φ − d1 ks As = 0,
2 dZ 2 As dT dT dT 2 dZ
(5.108)
 √ 
where d1 = χ / k ε . For the simplicity, we choose φ(Z) as follow

1
φ(Z) = . (5.109)
cosh [Z]

Noticing that
 2
dφ dφ d 2φ
= −φ sinh [Z] ,
2
= φ 2 − φ 4 , 2 = φ − 2φ 3 ,
dZ dZ dZ

and inserting Eq. (5.109) into Eqs. (5.107) and (5.108), we get
      
QI 1 dAs 1  1 dAs dX0
A2s − d1 φ 3 + + d1 A2s − ks2 φ − Z − As − ks φ 2 sinh [Z] = 0,
QR As dT 2 As dT dT
(5.110)
 
1 dks dks d φ0
A2s 1
Z + X0 − − + ks2 φ − d1 ks As φ 2 sinh [Z] = 0,
As dT dT 2 dT 2
(5.111)

Next, we choose X0 from the condition

dX0
− ks = 0. (5.112)
dT
Two different physical situations can then occur according to the initial value of the
relative velocity ks0 , case ks = 0 and case ks = 0.
First case:ks = 0. By setting ks = 0, equation (5.111) leads to
5.3 Soliton Signals in an Alternate Right-Handed … 197

d φ0 A2
+ s = 0. (5.113)
dT 2
Multiplying then Eq. (5.107) by φ and integrating the resulting equation with respect
to Z from −∞ to +∞, we obtain

  "+∞   "+∞
QI dZ 1 1 dAs dZ
A2s − d1 4
+ + d1 As
2
= 0. (5.114)
QR cosh [Z] 2 As dT cosh2 [Z]
−∞ −∞

Because
"+∞ "+∞
dZ dZ 4
= 2 and = ,
cosh2 [Z] cosh4 [Z] 3
−∞ −∞

equation (5.114) leads to the following one

dAs 4QI − d1 QR 3
+ As = 0. (5.115)
dT 3QR

Integrating Eqs. (5.115), (5.113) and (5.112), we arrive to the following soliton
parameters

As0 3 ln [As (T )]
As (T ) = )   , φ0 (T ) =   − φ00 , X0 (T ) = X0 ,
χ
1+ 2
4 QQRI − χ
√ A2s0 T 2 4 QQRI − k √ ε
3 k ε
(5.116)
χ
with 4 QQRI − k √ ε
< 0; here, As0 , φ 00 , and X0 ar arbitrary real constants of integration.
The corresponding approximate bright solitonlike solution of Eq. (5.73) is
  
u0 1 + 12λS 2 As (τ ) exp [iϕ(ξ, η, τ )]
ψ(ξ, η, τ ) = exp [χ n] − k λ2 β22 + α2 ,
Q1R 4 cosh [As (τ ) [α2 ξ + β2 η − X0 ]]
(5.117)
where
As0
As (τ ) = )   , (5.118)
Q1I χ 
1 − 3 As0 4k Q1R − ε
2u0 2 √ α2 + λ2 β2 τ
1+12λS 2
4
2

 
3 1 Q1I χ
ϕ(ξ, η, τ ) =   ln + φ00 , Q1R < 0, 4 − √ < 0,
Q1I χ
2 4 Q1R − k √ε As (τ ) Q 1R k ε

(5.119)

with
198 5 Emission of Nonlinear Modulated Waves in Multi-coupled Nonlinear …

Q1R = QR exp [2χ n] , Q1I = QI exp [2χ n] , and As0 = As (0) > 0,

X0 and φ00 are arbitrary real constants, and α2 and β2 are real parameters to be taken
from condition α2 β2 = 0.
Second case:ks = 0.
To simplify Eqs. (5.110) and (5.111) in the case when ks = 0, we impose to φ0 ,
ks , and As to satisfy the following differential system

d φ0 dks 1 2 
− X0 + As − ks2 = 0, (5.120a)
dT dT 2
1 dAs 1  2 
+ d1 As − ks2 = 0. (5.120b)
As dT 2

Under conditions (5.120a) and (5.120b), we rewrite system (5.110)–(5.111) as fol-


lows:

 
QI 1 dAs
A2s − d1 φ 3 − Zφ sinh [Z] = 0, (5.121a)
QR As dT
1 dks
Zφ − d1 ks As φ 2 sinh [Z] = 0. (5.121b)
As dT

Now, if we multiply Eqs. (5.121a) and (5.121b) respectively by φ(Z) and ddZφ and
integrate the resulting equation with respect to Z from −∞ to +∞, we obtain that

 
dAs 4 QI
+ d1 − A3s = 0, (5.122a)
dT 3 QR
dks 2
− d1 ks A2s = 0. (5.122b)
dT 3
Imposing now to Eq. (5.120b) to coincide with Eq. (5.122a), we obtain the following
relationship between ks and As
 
8 QI 5
ks2 = − A2s .
3d1 QR 3

Integrating now equations (5.122a) and (5.122b) and going back to Eqs. (5.112) and
(5.120a), we arrive, under the condition

2 QI √
d1 = ⇔ 3QR χ = 2QI k ε, (5.123)
3 QR

to the soliton parameters


5.3 Soliton Signals in an Alternate Right-Handed … 199
)
As0 7
As (T ) = / , ks (T ) = As (T ), (5.124a)
1− 8 2 QI
A T 3
9 s0 QR

3 21 QR 1
X0 (T ) = − + X00 , (5.124b)
4 QI As (T )

15QR 21
φ0 (T ) = − ln [As (T )] − 4 X00 A5s (T ) − φ00 ; (5.124c)
4QI 4As0

here, As0 = As (0), X00 , and φ00 are arbitrary real constants. The corresponding
approximate solitonlike solution of the 2D complex GL equation (5.73) is found
to be

  
u0 1 + 12λS 2
ψ(ξ, η, τ ) = exp [χ n] − k λ2 β22 + α2
Q1R 4
As (τ ) exp [iϕ(ξ, η, τ )]
× & √  ; (5.125)
cosh As (τ ) α2 ξ + β2 η + 3 421 QQ1R 1
1I As (τ )
− X00

in Eq. (5.125), As and ϕ are given by

As0
As (τ ) = /   , Q1R < 0, Q1I < 0,
Q
1 + 89 u0 Q1R
1I
A2s0 k λ2 β22 + α2
1+12λS 2
4
τ
(5.126)
) √
7 15Q1R 21
ϕ(ξ, η, τ ) = As (τ ) (α2 ξ + β2 η) + ln [As (τ )] + 4 X00 A5s (τ ) + φ00 ,
3 4Q1I 4As0
(5.127)

where As0 = As (0) > 0, X00 , and φ00 are three arbitrary real parameters.
It is important here to note that the above found approximate bright solitonlike
solutions of the 2D complex GL equation (5.73) are not solutions in the classical
sense; they are rather solutions in the averaged sense.

5.3.4.3 Dissipative Modulated Waves Propagating Along


a Multi-coupled Lossy Electric Network
 
In the following, we employ, under the conditions Q1I = Im Q exp [2χ n] < 0 and
 2
Q
2 − 1+12λ
4λ2
S √ 1R χ − k
εQ1I
≥ 0, the above exact and approximate solitonlike solu-
tions to investigate the dynamics of soliton signals along the network system of
Fig. 5.10c. For this aim, we assume that |u1 |  1 and consider only the fundamental
harmonic in the decomposition (5.56) so that the voltage Vn,m (t) be approximated as
follows Vn,m (t) = εu(n, m, t). For the numerical purpose, we will use N = 1600 and
200 5 Emission of Nonlinear Modulated Waves in Multi-coupled Nonlinear …

M = 300, the network parameters L = 200 µH, C0 = 370 pF, α = 0.21 V−1 , the
dimensionless conductance σ = 0.461 × 10−2 , and the small parameter ε = 10−2 .
Under the above assumptions, the shape of the electrical envelope soliton associ-
ated with the exact bright solitonlike solution (5.90) is found to be approximated as
follows

⎡ ⎛ ⎞⎤
n m
⎢ ⎜ + ⎟⎥
⎢ ⎜ wdthn wdthm ⎟⎥
⎢ ⎜ 0 ⎟⎥
⎢ ⎜ ⎟⎥
⎢ ⎜ 1  2 ⎟⎥
2⎢ ⎜
5+cosh ⎢A(n)⎜ 1 υ0
⎟⎥
⎟⎥
⎢ ⎜ 1 Q1I u0 + α0 ε ⎟⎥



⎜ 1
− 2−  
⎟⎥
t⎟⎥



⎝ √ β0
2 ⎟⎥
⎠⎦
8ε ελ2 u0 χ α2 + 4λ2
2 1+12λS
0
Vn,m (t) = A(n) ⎡
⎛ ⎞

n m

⎢ ⎜ + ⎟⎥

⎢ ⎜ wdthn wdthm ⎟⎥
⎢ ⎜ 0 ⎟ ⎥
⎢ ⎜ ⎟⎥
⎢ ⎜ 1  2 ⎟⎥
⎢ ⎜
−1+7 cosh2 ⎢A(n)⎜ 1 ⎟ ⎥
⎢ ⎜ 1 Q1I u0 + αυ00 ε ⎟⎥
⎟⎥
−1
⎢ ⎜ ⎟⎥



⎜ 2− √  2  t⎟
⎟⎥

⎢ ⎝ β0 ⎠⎥
⎣ 8ε ελ2 u0 χ α2 + 4λ2
2 1+12λS

0

× cos [ϑ(n, m, t)] , (5.128)

where different parameters are given as follows:


1
ϑ = ε2
      % 
4λ2 β0 4λ2 β0 μs
k− β1 n + ( + β1 ) m u0 k − β1 + ε ω + β1 t ,
1 + 12λS α0 1 + 12λS α0 β1
(5.129)
0  
1
1 4λ2 β0
2
1 λ2 u0 1 + 1+12λ
1 S α 2
A(n) = 2ε 2 A0 exp [−χ n] 2− 1
0
χ,
Q1I ε 2
0  2 
1 √
1 β0 1+12λS
1 2 ελ2 u0 χ +
2 α02 4λ2
wdthn = 2ε − , (5.130)
Q1I
0  2 
1 √
1 β0 1+12λ
1 2 ελ2 u0 χ + 4λ S
2εα0 2 2
α0 2  
wdthm = − , Q1I = Im Q exp [2χ n] < 0. (5.131)
β0 Q1I

Here, β1 = A0 exp [−χ n] = 0, A0 = 0 being a free real parameter. Here, A(n) and
wdthn and wdthm are respectively the soliton amplitude and, the soliton width in lon-
gitudinal and transverse directions.
Evaluating the shape of the electrical envelop solitons associated with the approx-
imate solitonlike solution (5.125)–(5.127) under the above assumptions and the fol-
lowing restriction
5.3 Soliton Signals in an Alternate Right-Handed … 201

Fig. 5.16 a Profile at time t = 0.1 µs of a forward damped modulated bright soliton propagating at
frequency fp = −1180 kHz in the LH network system of Fig. 5.14c; b–d Spatiotemporal evolution
of forward damped modulated bright solitons propagating at frequency fp = −1180 kHz through
respectively lines m = 10, m = 70, and m = 130 of the network system of Fig. 4.14c. Different plots
are generated with the use of the approximate modulated bright soliton solution (5.133) with data (
5.134) of the network Eq. (5.54) for the solution parameters A0 = 10, X00 = 1, φ00 = 1, α2 = 0.2,
and β2 = 0.4. For generating different plots, we have used the network parameters (5.53) with
λS = 0.1. Reprint from Ref. [22], Copyright 2022, with permission from Elsevier

   
Q1R = Re Q exp [2χ n] < 0 and Q1I = Im Q exp [2χ n] < 0, (5.132)

we find
A(n, t)
Vn,m (t) = & √  cos [ϑ(n, m, t)] .
3 21Q1R 1 u0
n
cosh A(n, t) wndth + m
wmdth + 4Q1I A(n,t) − wndth t − X00 (k)
(5.133)
Here,

0
1  
1 1+12λS α2
2
1 k λ2 +
1 u0 4 β 2
A(n, t) = 2ε 2 A0 exp [−χ n] 1
1− Q 2 2
 ,
2 1R 3 Q 1+12λS α2
1 + 89 u0 ε 2 Q 1I A20 exp [−2χ n] k λ2 + 4 2 t
1R β2
0   0  
1 1
β2 1 u 1 + 12λ α 2 1 u0 α2
3 k λ + 1 + 12λS 2 ,
wndth = 2 2− 2−
0 S 2
ε 3 k λ2 + , w mdth = 2 ε 2
α2 Q1R 4 β22 Q1R 4 β22
202 5 Emission of Nonlinear Modulated Waves in Multi-coupled Nonlinear …

Fig. 5.17 Evolution of damped forward soliton in the network of Fig. 5.14c at frequency f = −1180
kHz for the network parameter (5.53) with the same data as in Fig. 5.20. Top panels [bottom
panels]: Temporal evolution of the bright solitary pulse voltages through cell n = 10 of different
lines [through different cell of line m = 33]: a m = 10; b m = 70; c m = 130; [d n = 10; e n = 50;
f n = 90]. Different plots are generated with the approximate modulated bright soliton solution
(5.133) with data (5.134) of the network Eq. (5.54). Reprint from Ref. [22], Copyright 2022, with
permission from Elsevier

X00
X00 (k) = √ , (5.134)
εwmdth
) 
1 7 n − u0 t m
ϑ(n, m, t) = ε 2 [kn − (u0 k + ωε) t] + ε + A(n, t) + φ00
3 wndth wmdth
)    
21 X00 exp [4χ n] A(n, t) 5 15Q1R A(n, t)
+ + ln .
ε 16ε 2 A40 wmdth 4Q1I wmdth

Here, A0 = 0, X0 , φ00 , α2 , and β2 are free constants, while wndth and wmdth are the
soliton width in the n− and m− directions, respectively.
It is well seen from Eqs. (5.133) and 5.134) that
(i) The wave amplitude A(n, t) is a function of both the cell number n and the time t.
(ii) The centre of the electrical envelope soliton when propagating in the longitudinal
direction n and in the transverse direction m is respectively

3 21Q1R wndth β2
ξndir (t) = u0 t − − m + wndth X00 (k)
4Q1I A(n, t) α2

and √
α2 3 21Q1R wmdth α2
ξmdir (t) = u0 t − − n + wmdth X00 (k).
β2 4Q1I A(n, t) β2

(iii) The soliton speed in the n− and m− directions is found to be respectively


5.3 Soliton Signals in an Alternate Right-Handed … 203

Fig. 5.18 Effect of the coupling parameter λ2 on the wave motion. Top panels: Spatiotemporal evo-
lution of the modulated damped bright solitons propagating at frequency fp = −1180 kHz along
line m = 10 of the network system of Fig. 5.14c for different values of the coupling parameter
λ2 , λ2 = 0.5C2 /C0 (a), λ2 = 3C2 /C0 (b), and λ2 = 6C2 /C0 (c). Middle panels [Bottom panels]:
Spatial profile at time t = 0.0625 µs [t = 0.125 µs] of the modulated damped bright solitons prop-
agating at frequency fp = −1180 kHz through line m = 10 of the network of Fig. 5.14c for different
values of the coupling parameter λ2 , λ2 = 0.5C2 /C0 (d), λ2 = 3C2 /C0 (e), and λ2 = 6C2 /C0 (f)
[λ2 = 0.5C2 /C0 (g), λ2 = 3C2 /C0 (h), and λ2 = 6C2 /C0 (i)]. Different plots are generated with
the use of the approximate modulated bright pulse signal solution (5.133) with data (5.134) of the
network Eq. (5.54) for the network parameters (5.53) and the same parameters as in Fig. 5.20,
except X00 , X00 = 5. Reprint from Ref. [22], Copyright 2022, with permission from Elsevier


3 21Q1R wndth dA(n, t)
υsoln = u0 +
4Q1I A2 (n, t) dt

and √
α2 3 21Q1R wmdth dA(n, t)
υsolm = u0 + .
β2 4Q1I A2 (n, t) dt

Dynamics of damped envelop soliton propagating along the multi-coupled lossy


electrical network Plots of Fig. 5.16 show a typical example of the dynamics of
damped envelop soliton propagating along a multi-coupled lossy electrical network
depicted on Fig. 5.10c. Generated with the exact electrical voltage (5.128), Fig.
Fig. 5.16a reveals that the wave amplitude decreases during its propagation along the
longitudinal direction in the network system. It is seen from Figs. 5.16b–d, obtained
with the approximate electrical voltage (5.133), that the soliton speed decreases when
the line number m increases. This behavior is clearly seen from plots of Fig. 5.17.
204 5 Emission of Nonlinear Modulated Waves in Multi-coupled Nonlinear …

Fig. 5.19 Effects of the linear dispersive parameter λS on the wave propagation. Top panels:
Spatiotemporal evolution of the forward damped modulated bright soliton propagating at frequency
f = −1180 through line m = 10 of our network system for the network parameters (5.53). Bottom
panels: Profile at time t = 0.1 µs of the bright soliton propagating at frequency f = −1180 kHz
through line m = 10 of the network system of Fig. 5.14c for the network parameters (5.53). Different
plots are generated with the same data as in Fig. 5.20 and different numerical values of the dispersive
parameter λS , λS = 0.05 (a, d), λS = 1 (b, e), and λS = 1.5 (c, f). Reprint from Ref. [22], Copyright
2022, with permission from Elsevier

Fig. 5.20 Plots of the frequency fp versus the longitudinal wavenumber k for different values
of the transverse wavenumber  according to the dispersion relation (5.63) for the experimental
dimensionless conductance σ = 0.461 × 10−2 and the network parameters L = 200 μH, C0 = 370
pF, CS = 0.01C0 , C2 = 100 pF, and α = 0.21 V−1 . Plots of panel (a) show that distinct modulated
waves can propagate at the same frequency. Reprint from Ref. [22], Copyright 2022, with permission
from Elsevier

Effects of the coupling parameter λ2 on the wave propagation To investigate the


effects of the coupling parameter λ2 = C2 /C0 on the wave propagation, we display
in Fig. 5.18 the evolution of bright solitonlike signals propagating at frequency fp =
−1180 kHz along line m = 10 of the network system of Fig. 5.10c. From plots of
Fig. 5.18, we can see that
(i) The wave amplitude and the soliton width increase with λ2 .
(ii) The soliton speed decreases as the coupling parameter λ2 increases.
5.3 Soliton Signals in an Alternate Right-Handed … 205

Fig. 5.21 Top panels: Spatial profile at time t = 0.2 μs of bright solitons propagating at frequency
fp = 2 Hz in the RH network (a) and in the LH network (b); Bottom panels: Spatial profile at time
t = 0.4 µs of bright solitons propagating at frequency fp = 2 Hz in the RH network (c) and in
the LH network (d). Different plots are obtained with the help of the bright solitonlike solution
(5.128) with data (5.129)–(5.131) for the transverse wavenumber  = 0.14 and the same network
parameters as in Fig. 5.24. Plots of left [right) panels correspond to k = 0.011 (k ≈ 0.078). Other
parameters used in depicting different plots are ε = 10−3 and A0 = 25. Reprint from Ref. [22],
Copyright 2022, with permission from Elsevier

Effects of the dispersive element CS on the wave propagation To study the effects
of the dispersive element CS on the wave propagation, we depict in Fig. 5.19 the
evolution of the approximate electrical voltage Vn,m (t) for different values of param-
eter λS = CS /C0 . Different plots of Fig. 5.19 show that the wave amplitude [speed]
increases [decreases] as parameter λS increases. Our study has revealed that the
dispersive element CS has not effect on the soliton width.
Right-handed versus left-handed networks, forward vs backward bright soli-
tons, and effects of the transverse wavenumber  on wave propagation For the
study of the alternately LH and RH behaviors of our network system, the behavior of
forward and backward solitons, as well as the effects of the transverse wavenumber 
on the envelope solitons propagating in the network system of Fig. 5.10c, we employ
here the exact electrical voltage Vn,m (t) given by Eq. (5.128). Before carrying out
this study, we first show in Fig. 5.20 the variation of the propagating frequency fp
versus k for different values of . Plots of Fig. 5.20 justify our choice on the exact
electrical voltage (5.128) for our study. It follows from Fig. 5.20 that for a given set
of the network parameters, there exists ]0, max ] so that the transverse wavenumber
 ∈]0, max ] with ]0, max ] =]0, 1 ] ∪ [1 , 2 ] ∪ [2 , max ] and the network system
depicted in Fig. 5.10c may support
(i) envelop solitons is purely RH behavior in ]0, 1 ],
(ii) bright solitons is purely LH behavior in [2 , max ], and
(iii) envelop solitons is composite RH/LH behavior in  ∈]1 , 2 [.
206 5 Emission of Nonlinear Modulated Waves in Multi-coupled Nonlinear …

Fig. 5.22 Top panels: Profile at time t = 0.2 μs of bright solitons propagating at frequency fp = 2
Hz along line m = 1 of our network system; Bottom panels: Profile at time t = 0.4 µs of bright
solitons propagating at frequency fp = 2 Hz along line m = 50 of our network system. In different
plots, the solid and the dashed lines are generated with respectively k = 0.011 and k = 0.078
corresponding to the right-handed regime and the left-handed regime, respectively. Different plots
are obtained with the use of the bright solitonlike solution (5.128) with data (5.129)–(5.131) for the
transverse wavenumber  = 0.14 and the same network parameters as in Fig. 5.24. Other parameters
used in generating different plots are ε = 10−3 and A0 = 25. Reprint from Ref. [22], Copyright
2022, with permission from Elsevier

The above enumerated behaviors can be well seen in Figs. 5.21 and 5.22.
To analyze the effects of the transverse wavenumber  on the wave propagation,
we display for different values of  the evolution of envelop solitons, propagating
at the same frequency fp = 6.048 Hz (plots of the left panels) and fp = −1180 kHz
(plots of the right panels) along line m = 50 in Fig. 5.23. From plots of Fig. 5.23,
we can see that for solitons propagating at the same frequency,
(i) the soliton amplitude in the RH (LH) region decreases (increases) as  increases;
(ii) the soliton width in the RH (LH) region decreases (increases) as  increases;
(iii) the speed of the backward (forward) envelop soliton in the RH (LH) region
increases (decreases) when  increases.
We end the present subsection with the study of the dynamics of composite for-
ward/backeward envelop solitons in the LH regime. For this purpose, we show in
Fig. 5.24 the spatial evolution at low frequencies fp = 4 Hz and fp = −4 Hz of respec-
tively the backward (left panels) and forward (right panels) envelop solitons in the
LH region. Plots of Fig. 5.24 reveal that the soliton amplitude and the soliton width
5.3 Soliton Signals in an Alternate Right-Handed … 207

Fig. 5.23 Left panels: Propagation at frequency fp = 6.048 Hz of backward bright solitons in
the RH region for a  = 0.13 and k = 0.0152, b  = 0.135 and k = 0.021, and c  = 0.14 and
k = 0.043. Right panels: Evolution at frequency fp = −1180 kHz of forward bright solitons in the
LH region for d  = 3 and k = 1.526, (e)  = 3.2 and k = 1.373, and (f)  = 3.4 and k ≈ 1.237.
Different plots are generated with the same network parameters as in Fig. 5.24, with parameters
ε = 10−3 and A0 = 25. Reprint from Ref. [22], Copyright 2022, with permission from Elsevier

of forward bright soliton are larger than those of backward solitons, while the speed
of the forward bright soliton is greater than that of the backward bright soliton.
Figure 5.25 shows the typical spatiotemporal evolution of composite back-
ward/forward bright solitons in the LH region. In this Fig. 5.25, envelop solitons
propagate along line m = 1 at low frequency fp = 4 Hz. Figure 5.25a, b shows the
propagation of backward [forward] bright solitons propagating at frequency fp = 4
Hz [fp = −4 Hz]. From these plots, we can see that the amplitude of the backward
bright soliton is greater than that of the forward soliton.
Kink soliton signals propagating in the multi-coupled lossy network An example
of the dynamics of kink soliton signals propagating through the network system of
Fig. 5.10c can be obtained with the use of the kink solitonlike solution (5.104). Using
this solution (5.104), we show in Fig. 5.26 the evolution of composite RH/LH (top
plots) and composite backward/forward in left-handed regime (bottom plots) kink
solitons propagating along line m = 1. It is seen from plots of Fig. 5.26 that the kink
208 5 Emission of Nonlinear Modulated Waves in Multi-coupled Nonlinear …

Fig. 5.24 Left panels: Profile at times t = 0.2 μs (a) and t = 0.4 µs (c) of the backward bright
soliton voltages propagating in the LH network at frequency fp = 4 Hz that corresponds to the
longitudinal wavenumber k = 0.04830; Right panels: Profile at times t = 0.2 µs (b) and t = 0.4
µs (d) of the forward bright soliton voltages propagating in the LH network at frequency fp = −4
Hz corresponding to the longitudinal wavenumber k = 0.08430. Reprint from Ref. [22], Copyright
2022, with permission from Elsevier

Fig. 5.25 Spatiotemporal evolution of a backward bright soliton and b forforward bright soliton
moving along line m = 1 at low frequency fp = 4 Hz. Different plots are generated with the trans-
verse wavenumber  = 0.142 and k = 0.04831 and k = 0.08430, respectively. Other parameters
are σ = 0.461 × 10−2 , λS = 0.01, A0 = 25, and ε = 10−3 . Reprint from Ref. [22], Copyright
2022, with permission from Elsevier
5.3 Soliton Signals in an Alternate Right-Handed … 209

Fig. 5.26 Spatiotemporal evolution of RH/LH (top panels) and backward/forward (bottom panels)
kink soliton moving along line m = 1 for  = 0.14, λS = 0.01, σ = 0.461 × 10−2 , ε = 10−3 , and
A0 = −250. a, b Evolution of kink solitons moving at frequency fp = 3 Hz in respectively the right-
and left-handed regimes for k = 0.01632 (a) and k = 0.07431 (b); c, d Evolution of the backward
(c) and forward (d) kink soliton propagating at frequency fp = 0.9 Hz generated with k ≈ 0.08593
(c) and k ≈ 0.08120 (d). Reprint from Ref. [22], Copyright 2022, with permission from Elsevier

soliton pulses propagating in the RH regime oscillate less in both time and space
than those in the LH regime.

5.3.5 Conclusion and Discussion

We have considered in this Section a multi-coupled Noguchi lossy NLTN with linear
dispersive elements. The dynamics of modulated waves propagating in this network
system are reduced to a two-dimensional complex GP equation whose coefficients
depend on the spatial linear gain χ . Using the amplitude continuum equation, we have
established the conditions under which Stokes waves will be unstable under modula-
tion. Next, exact and approximate solitonlike solutions of the amplitude continuum
equation are presented and used to investigate analytically the effects of various net-
work parameters as well as that of the transverse wavenumber  on the dynamics of
210 5 Emission of Nonlinear Modulated Waves in Multi-coupled Nonlinear …

modulated wave solitons of our network system. Using the sign of the longitudinal
group velocity, we have classified the network system under consideration into three
classes, the class of right-hand, left-hand, and composite left-/right-hand network.
We found that these classification depends on the transverse wavenumber . Our
results showed that the network system under consideration may support backward
waves, forward waves, as well as composite backward/forward waves.
It is important to notice that during the investigation of the characteristics of the
physical system of Fig. 5.10c, we have assumed that air and ceramic core inductors
do not exhibit any significant voltage noise. Also, this study has neglected the voltage
noise from a varactor in reverse bias. Since the NLTN is extremely susceptible to
bias voltage noise contributions [44], we believe that by taking into consideration
the noises in the physical system of Fig. 5.10c, the dynamics of the modulated waves
studied in this Section may be affected.

References

1. E. Kengne, S.T. Chui, W.M. Liu, Modulational instability criteria for coupled nonlinear trans-
mission lines with dispersive elements. Phys. Rev. E 74, 036614 (2006)
2. E. Kengne, V. Bozic, M. Viana, R. Vaillancourt, Transverse stability of solitary waves propa-
gating in coupled nonlinear dispersive transmission lines. Phys. Rev. E 78, 026603 (2008)
3. E.A. Kuznetsov, A.M. Rubenchik, V.E. Zakharov, Soliton stability in plasmas and hydrody-
namics. Phys. Rep. 142, 103–165 (1986)
4. W.S. Duan, Stability of dust acoustic waves in weakly two-dimensional dust plasma with
vortex-like ion distribution. Chin. Phys. Lett. 19, 452 (2002)
5. T. Kakutani, N. Yamazaki, Solitary waves on a two-layer fluid. J. Phys. Soc. Jpn. 45, 674 (1978)
6. T. Yoshinaga, T. Kakutani, Solitary and shock waves on a coupled transmission line. J. Phys.
Soc. Jpn. 49, 2072 (1980)
7. M.S. Ody, A.K. Common, M.I. Sobhy, Continuous symmetries of the discrete nonlinear tele-
graph equation. Eur. J. Appl. Math. 10, 265 (1999)
8. T.B. Benjamin, J.E. Feir, The disintegration of wavetrains on deep water. Part 1. J. Fluid Mech.
27, 417-430 (1967)
9. W.-S. Duan, Nonlinear waves propagating in the electrical transmission line. Europhys. Lett.
66, 192 (2004)
10. Yu.A. Stepanyants, Experimental investigation of cylindrically diverging solitons in an electric
lattice. Wave Motion 3, 335 (1981)
11. L.A. Ostrovski, V.V. Papko, Y.A. Stepanyants, Solitons and resonances in two dimensional
lattices. Sov. Phys. JETP 51, 417 (1980)
12. K. Narahara, Coupled nonlinear transmission lines for doubling repetition rate of incident pulse
streams. Prog. Electromagn. Res. Lett. 16, 69–78 (2010)
13. N.A. Akem, A.M. Dikandé, B.Z. Essimbi, Leapfrogging of electrical solitons in coupled non-
linear transmission lines: effect of an imperfect varactor. Appl. Sci. 2, 21 (2020)
14. E. Kengne, R. Vaillancourt, 2D Ginzburg-Landau system of complex modulation for coupled
nonlinear transmission lines. J. Infrared Milli. Terahz Waves 30, 679–699 (2009)
15. E. Kengne, W.M. Liu, Management of modulated wave solitons in a two-dimensional nonlinear
transmission network. Eur. Phys. J. B 92, 235 (2019)
16. A.C. Hicks, A.K. Common, M.I. Sobhy, Approximations to large amplitude solitary waves on
nonlinear electrical lattices. Physica D 95, 167–186 (1996)
17. A. Gammal, T. Frederico, L. Tomio, Ph. Chomaz, Atomic Bose-Einstein condensation with
three-body interactions and collective excitations. J. Phys. B 33, 4053 (2000)
References 211

18. E. Kengne, C.K. Kum, Modulational instability in nonlinear bi-inductance transmission line.
Int. J. Mod. Phys. B 19(26), 3961 (2005)
19. W. van Saarloos, P.C. Hohenberg, Fronts, pulses, sources and sinks in generalized complex
Ginzburg-Landau equations. Physica 56D, 303 (1992)
20. E. Kengne, C.N. Bame, Dynamics of modulated wave trains in a discrete nonlinear-dispersive
dissipative bi-inductance transmission line. Physica Scripta 71, 423–432 (2005)
21. Wim van Saarloos, Fronts, pulses, sources and sinks in generalized complex Ginzburg-Landau
equations. Physica D 56, 303 (1992)
22. E. Kengne, Modulational instability and soliton propagation in an alternate right-handed and
left-handed multi-coupled nonlinear dissipative transmission network. Chaos Solitons Fractals
146, 110866 (2021)
23. R. Hirota, K. Suzuki, The direct method in soliton theory. J. Phys. Soc. Jpn. 28, 1366 (1970)
24. R. Hirota, K. Suzuki, Theoretical and experimental studies of lattice solitons in nonlinear
lumped networks. Proc. IEEE 61, 1483 (1973)
25. A. Noguchi, Solitons in a nonlinear transmission line. Electron. Commun. Jpn. 57A, 9 (1974)
26. E. Kengne, R. Vaillancourt, 2D Ginzburg-Landau system of complex modulation for coupled
nonlinear transmission lines. J. Infrared Milli Terahz Waves 30, 679 (2009)
27. Y.A. Stepanyants, Experimental investigation of cylindrically diverging solitons in an electric
lattice. Wave Motion 3, 335 (1981)
28. P. Marquié, J.M. Bilbault, M. Remoissenet, Generation of envelope and hole solitons in an
experimental transmission line. Phys. Rev. E 49, 828 (1994)
29. D. Yemélé, P.K. Talla, T.C. Kofané, Dynamics of modulated waves in a nonlinear discrete LC
transmission line: dissipative effects. J. Phys. D: Appl. Phys. 36, 1429 (2003)
30. C.B. Tabi, R.Y. Ondoua, H.P. Ekobena, A. Mohamadou, T.C. Kofané, Energy patterns in cou-
pled α-helix protein chains with diagonal and off-diagonal couplings. Phys. Lett. A 380, 2374
(2016)
31. B.A. Malomed, Complex Ginzburg–Landau equation, in Encyclopedia of Nonlinear Science
(Routledge, New York, 2005), pp. 157–160
32. E. Kengne, W.M. Liu, Rogue waves of the dissipative Gross-Pitaevskii equation with distributed
coefficients. AIP Adv. 10, 045218 (2020)
33. C.S. Panguetna, C.B. Tabi, T.C. Kofané, Low relativistic effects on the modulational instability
of rogue waves in electronegative plasmas. J. Theor. Appl. Phys. 13, 237 (2019)
34. E. Kengne, R. Vaillancourt, B.A. Malomed, Coupled nonlinear Schrödinger equations for
solitary-wave and kink signals propagating in discrete nonlinear dispersive transmission lines.
Int. J. Mod. Phys. B 23, 133 (2009)
35. P.G. Kevrekidis, R. Carretero-Gonzalez, G. Theocharis, D.J. Frantzeskakis, B.A. Malomed,
Phys. Rev. A 68, 035602 (2003). Phys. Rev. A 68, 035602 (2003)
36. E. Kengne, W.M. Liu, Management of matter-wave solitons in Bose-Einstein condensates with
time-dependent atomic scattering length in a time-dependent parabolic complex potential. Phys.
Rev. E 98, 012204 (2018)
37. P. Grelu, N. Akhmediev, Dissipative solitons for mode-locked lasers. Nat. Photon. 6, 84 (2012)
38. B.F. Akers, Modulational instabilities of periodic traveling waves in deep water. Physica D
300, 26 (2015)
39. F. Bérard, C. Vandamme, S.C. Mancas, Two-dimensional structures in the quintic Ginzburg-
Landau equation. Nonlinear Dyn. 81, 1413 (2015)
40. O. Descalz, H.R. Brand, Dissipative soliton stabilization by several nonlinear gradient terms.
Chaos 30, 043119 (2020)
41. C. Cartes, O. Descalzi, H.R. Brand, Exploding dissipative solitons in the cubic-quintic complex
Ginzburg-Landau equation in one and two spatial dimensions. Eur. Phys. J. Spec. Top. 223,
2145 (2014)
42. A. Hasegawa, Y. Kodama, Solitons in Optical Communications (Oxford University Press,
Oxford, 1995)
43. Y.S. Kivshar, B.A. Malomed, Dynamics of solitons in nearly integrable systems. Rev. Mod.
Phys. 61, 763 (1989)
212 5 Emission of Nonlinear Modulated Waves in Multi-coupled Nonlinear …

44. J. Breitbarth, Z. Popović, in 2017 Joint Conference of the European Frequency and Time Forum
and IEEE International Frequency Control Symposium (EFTF/IFC) (2017), pp. 261–264
45. A.B. Aceves, G.G. Luther, C. De Angelis, A.M. Rubenchik, S.K. Turitsyn, Energy localization
in nonlinear fiber arrays: collapse-effect compressor. Phys. Rev. Lett. 75, 73 (1995)
46. H. Leblond, B.A. Malomed, D. Mihalache, Spatiotemporal vortex solitons in hexagonal arrays
of waveguides. Phys. Rev. A 83, 063825 (2011)
47. L. Hadžievski, A. Maluckov, A.M. Rubenchik, S. Turitsyn, Stable optical vortices in nonlinear
multicore fibers. Light Sci. Appl. 4, e314 (2015)
48. K. Krupa, A. Tonello, A. Barthélémy, T. Mansuryan, V. Couderc, G. Millot, P. Grelu, D.
Modotto, S.A. Babin, S. Wabnitz, Multimode nonlinear fiber optics, a spatiotemporal avenue.
APL Photon. 4, 110901 (2019)
Part II
Dynamics of Matter-Wave
and Magnetic-Wave Solitons
Chapter 6
Introduction

Abstract The present chapter introduces the main idea and content of the second
part of this book. Here, the scientific motivations of part II of the book are presented.
Main terminologies to be encountered in this part of the book are introduced. This
chapter 6 of the book also presents the main structure of the rest of this book as well
as main mathematical equations and techniques to be used in the last part of this
book.

Nonlinear Schrödinger equations are the most-studied nonlinear partial differential


equations in the physics community. They are basic models in a broad spectrum of
disciplines in applied mathematics, nonlinear physics, and engineering. Nonlinear
Schrödinger equations describe on a qualitative as well as on a quantitative level
a vast variety of phenomena such as nonlinear waves, second-order phase transi-
tions, superfluidity, superconductivity, strings in field theory, liquid crystals, and
Bose-Einstein condensates of ultracold atoms. In the context of Bose-Einstein con-
densation, the nonlinear Schrödinger equation contains an external potential; in this
context, they are often called the Gross-Pitaevskii (GP) equation. In general, the non-
linear Schrödinger equation with external potential in the context of BECs is used
to describe and understand some nonlinear phenomena, interesting and important
properties and characteristics of vortex states [1, 2] and localized waves including
rogue wave [3, 4], soliton [5, 6], breather [7] in BECs, which is the main nonlinear
system to be studied in this part of the book. Soliton solutions are commonly known
fundamental solutions of the GP equation [8, 9]. Various types of solitons have been
predicted and observed in optics and in BECs [10–14]. Particularly, many results
have been reported in the framework of the dynamical management of solitons in
many fields of science as solutions of NLS/GP equations with distributed coeffi-
cients [15]. A typical example is the dispersion management (DM) in fiber optics
[10, 15, 16]. The dynamical management of solitons helps to stabilize DM solitons
against various perturbations [17–20]. It is important to note that the DM scheme
was predicted to help to stabilize two-dimensional [21] and three-dimensional [22]
spatiotemporal solitons.
© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2022 215
E. Kengne and W. Liu, Nonlinear Waves,
https://doi.org/10.1007/978-981-19-6744-3_6
216 6 Introduction

The management of a time-varying coefficient in front of the cubic term of


NLS/GP equation, also called nonlinearity management (NM) also help in the dynam-
ical management of solitons in nonlinear systems [23–26].
In the context of optics, the concept of NM aims to create nontrivial spatiotemporal
light patterns, especially when the structure in light beams is induced by means of
linear techniques [27–37] . In the context of BECs, the cubic term of the GP equation
accounts for effects of two-body inter-atomic collisions treated in the mean-field
approximation, with the coefficient in front of it proportional to the s-wave collision
scattering length which can be experimentally controlled by means of the Feshbach-
resonance (FR) effect [38–50]. The nonlinearity management in the BECs context
may be realized using time-dependent and/or nonuniform control fields [11, 15, 46,
47, 51–55]. It should be noted that the FR technique as well as the quench method
can also be used for NM in the BECs context [56–59].
It has been established that in the case of the attractive interaction, matter-wave
solitons may be formed in an effectively one-dimensional condensate [12]; however,
the situation is not the same in the case of the two-dimensional geometry where the
attraction may results in the critical collapse of the condensate. For the formation of
matter-wave solitons in the case of 2D geometry, the FR technique was predicted to
be quite useful for the stabilization of 2D matter-wave solitons against the critical
collapse [24, 60–63]. It has been established that the NM scheme cannot stabilize
3D matter-wave solitons against the super critical collapse [64]; however, 3D matter-
wave solitons can be stabilized if the nonlinearity management is combined with a
quasi-1D spatially periodic potential applied to the condensate [11, 65–68]. For Bose-
Einstein condensations, a natural generalization of the scenarios of the nonlinearity
management is to combine the variable coefficient in front of the cubic nonlinear
term in the GP equation, imposed by the temporal and/or spatial modulation of the
scattering length, with a temporarily modulated potential trapping the condensate
[69–87].
In this part of the book, we also investigate the dynamics of magnetization in
ferromagnet with or without spin-transfer torque. For this purpose, the main physical
model to be used is the Heisenberg model of spin–spin interactions which can be
considered as the starting point for understanding the complex magnetic structures
[88]. The main mathematical equation to be used is the Landau–Lifshitz–Gilbert
(LLG) equation which describes the magnetization of a ferromagnet as a function of
space and time [89–92].
In this part of the book, we intend to develop basic theoretical results for the
dynamics matter-wave and magnetic-wave solitons of nonlinear systems in general
and of BECs in particular, as well as of optical fields trapped in external potentials,
combined with the time-modulated nonlinearity. The respective models are based on
one-component, two-component, and three-component non-autonomous GP/NLS
equations. The present part II of the book is composed of eight chapters which are
related by gradually increasing complexity of physical models presented in them.
Chapter 7 addresses the dynamics of one-dimensional condensates with time mod-
ulation of the scattering length and trapping potential. Here, the approach outlined
above leads to the construction and management of non-autonomous solitons of 1D
6 Introduction 217

cubic self-defocusing NLS equations with spatiotemporally modulated coefficients,


that may be transformed into the classical integrable NLS equation. In this setting,
matter-wave soliton solutions are constructed in an analytical form, and it is shown
that the instability of those solitons, if any, may be delayed or completely eliminated
by varying the nonlinearity’s strength in time. In this Chap. 7 , we also engineer non-
autonomous matter-wave solitons in BECs with spatially modulated local nonlin-
earity and a time-dependent harmonic-oscillator (HO) potential. The modulational
instability in that setting is considered too. Chapter 8 deals with the dynamics of
matter rogue waves in BECs trapped in time-dependent external potential. Chapter 9
deals with the soliton generation in multi-component Bose–Einstein condensates. In
Chap. 10 , we proceed to more general nonintegrable models, which are treated by
means of the semi-analytical variational approximation and direct numerical simu-
lations. In this case, we address the dynamics of two- and three-dimensional conden-
sates with the nonlinearity strength containing constant and harmonically varying
parts, which can be implemented with the help of ac magnetic field tuned to FR.
In particular, the spatially uniform temporal modulation of the nonlinearity may
readily play the role of an effective trap that confines the condensate, and some-
times enforces its collapse. Chapter 11 deals with the investigation of soliton states
in a model based on a set of three coupled GP equations modeling the dynamics
of the spinor BECs. Both nonintegrable and integrable versions of the system are
considered, and exact soliton solutions are demonstrated. The stability of the solu-
tions was checked, in most cases, by direct simulations and, in some cases, it was
investigated in a more rigorous form, based on linearized Bogoliubov–de Gennes
(BdG) equations for small perturbations. The motion of bright and dark matter-wave
solitons in 1D BECs in the presence of spin-orbit coupling (SOC) is also treated
in Chap. 11 . Here, we demonstrate that the spin dynamics of the SOC solitons is
governed by a nonlinear Bloch equation and affects the orbital motion of the solitons,
leading to SOC effects in the dynamics of macroscopic quantum objects. The models
addressed in Chap. 11 , as well as the applied methods and the obtained results, are
quite similar to those produced by the engineering. In particular, the macroscopic
SOC phenomenology is explained by the fact that an effective time-periodic force
produced by rotation of the soliton’s (pseudo-) spin plays the role of temporal man-
agement which affects motion of the same soliton. Chapters 12 and 13 deal with
the dynamics of magnetization in ferromagnet with or without spin-transfer torque.
Here, we mainly discuss the properties of dynamic magnetic solitons in uniaxial
anisotropic ferromagnetic nanowire driven by spin-transfer torque; we also analyze
the modulation instability and dark soliton on the spin wave (SW) background, which
shows the characteristic breather behavior of the soliton as it propagates along the
ferromagnetic nanowire. With stronger breather character, we engineer novel mag-
netic rogue waves (MRWs) and clarify its formation mechanism. The generation of
magnetic rogue waves, mainly arises from the accumulation of energy and magnons
toward to its central part [93], is also considered. Also, we theoretically investigate
in the last two of the book, the current-excited and frequency-adjusted ferromagnetic
resonance in magnetic trilayers.
218 6 Introduction

References

1. D.S. Wang, S.W. Song, B. Xiong, W.M. Liu, Quantized vortices in a rotating Bose-Einstein
condensate with spatiotemporally modulated interaction. Phys. Rev. A 84, 053607 (2011)
2. J.J. García-Ripoll, V.M. Pérez-García, Vortex bending and tightly packed vortex lattices in
Bose-Einstein condensates. Phys. Rev. A 64, 053611 (2001)
3. F. Yu, Matter rogue waves and management by external potentials for coupled Gross-Pitaevskii
equation. Nonlinear Dyn. 80, 685–699 (2015)
4. Y.V. Bludov, V.V. Konotop, N. Akhmediev, Matter rogue waves. Phys. Rev. A 80, 033610
(2009)
5. Y.Y. Yan, W.J. Liu, Soliton rectangular pulses and bound states in a dissipative system modeled
by the variable-coefficients complex cubic-quintic Ginzburg-Landau equation. Chin. Phys. Lett.
38, 094201 (2021)
6. A.R. Fritsch, M. Lu, G.H. Reid, A.M. Piñeiro, I.B. Spielman, Creating solitons with controllable
and near-zero velocity in Bose-Einstein condensates. Phys. Rev. A 101, 053629 (2020)
7. W.B. Cardoso, A.T. Avelar, D. Bazeia, Modulation of breathers in cigar-shaped Bose-Einstein
condensates. Phys. Lett. A 374, 2640–2645 (2010)
8. T. Dauxois, M. Peyrard, Physics of Solitons (Cambridge University Press, Cambridge, 2006)
9. J. Yang, Nonlinear Waves in Integrable and Nonintegrable Systems (SIAM, Philadelphia, 2010)
10. A. Hasegawa, Y. Kodama, Solitons in Optical Communications (Clarendon Press, Oxford,
1995)
11. Y.S. Kivshar, G.P. Agrawal, Optical Solitons: From Fibers to Photonic Crystals (Academic
Press, San Diego, 2003)
12. K.E. Strecker, G.B. Partridge, A.G. Truscott, R.G. Hulet, Bright matter wave solitons in Bose-
Einstein condensates. New J. Phys. 5, 73 (2003)
13. FKh. Abdullaev, A. Gammal, A.M. Kamchatnov, L. Tomio, Dynamics of bright matter wave
solitons in a Bose-Einstein condensate. Int. J. Mod. Phys. B 19, 3415–3473 (2005)
14. L. Salasnich, Bright solitons in ultracold atoms. Opt. Quant. Electron. 49, 409 (2017)
15. B.A. Malomed, Soliton Management in Periodic Systems (Springer, New York, 2006)
16. S.K. Turitsyn, B.G. Bale, M.P. Fedoruk, Dispersion-managed solitons in fibre systems and
lasers. Phys. Rep. 521, 135–203 (2012)
17. J.H.B. Nijhof, N.J. Doran, W. Forysiak, F.M. Knox, Stable soliton-like propagation in dispersion
managed systems with net anomalous, zero and normal dispersion. Electron. Lett. 33, 1726–
1727 (1997)
18. A.N. Niculae, W. Forysiak, A.J. Gloag, J.H.B. Nijhof, N.J. Doran, Soliton collisions with
wavelength-division multiplexed systems with strong dispersion management. Opt. Lett. 23,
1354–1356 (1998)
19. D.J. Kaup, B.A. Malomed, J. Yang, Collision-induced pulse timing jitter in a wavelength-
division-multiplexing system with strong dispersion management. J. Opt. Soc. Am. B 16,
1628–1635 (1999)
20. C.C. Mak, K.W. Chow, K. Nakkeeran, Soliton pulse propagation in averaged dispersion-
managed optical fiber system. J. Phys. Soc. Jpn. 74, 1449 (2005)
21. M. Matuszewski, M. Trippenbach, B.A. Malomed, E. Infeld, M. Skorupski, Two-dimensional
dispersion-managed light bullets in Kerr media. Phys. Rev. E 70, 016603 (2004)
22. L. Gao, K.H. Wagner, R.R. McLeod, All-optical Tb/S 3R wavelength conversion using
dispersion-managed light bullets. IEEE J. Sel. Topics Quant. Elect. 14, 625–633 (2008)
23. M. Desaix, D. Anderson, M. Lisak, Variational approach to collapse of optical pulses. J. Opt.
Soc. Am. B 8, 2082–2085 (1991)
24. I. Towers, B.A. Malomed, Stable (2+1)-dimensional solitons in a layered medium with sign-
alternating Kerr nonlinearity. J. Opt. Soc. Am. B 19, 537–543 (2002)
25. R.Y. Chiao, E. Garmire, C.H. Townes, Self-trapping of optical beams. Phys. Rev. Lett. 13,
479–482 (1964)
26. G. Fibich, The Nonlinear Schrödinger Equation: Singular Solutions and Optical Collapse
(Springer, Heidelberg, 2015)
References 219

27. R.J. Valkenburg, A.M. McIvor, Accurate 3D measurement using a structured light system.
Image Vision Comput. 16, 99–110 (1998)
28. J. Salvi, J. Pages, L. Batlle, Pattern codification strategies in structured light systems. Pattern
Recogn. 37, 827–849 (2004)
29. S. Zhang, H. Song and S. Peisen, Novel method for structured light system calibration. Opt.
Eng. 45, 083601 (2006)
30. J. Geng, Structured-light 3D surface imaging: a tutorial. Adv. Opt. Phot. 3, 128–160 (2011)
31. A. Forbes, A. Dudley, M. McLaren, Creation and detection of optical modes with spatial light
modulators. Adv. Opt. Phot. 8, 200–227 (2016)
32. ...H. Rubinsztein-Dunlop, A. Forbes, M.V. Berry, M.R. Dennis, D.L. Andrews, M. Mansuripur,
C. Denz, C. Alpmann, P. Banzer, T. Bauer, E. Karimi, L. Marrucci, M. Padgett, M. Ritsch-
Marte, N.M. Litchinitser, N.P. Bigelow, C. Rosales-Guzman, A. Belmonte, J.P. Torres, T.W.
Neely, M. Baker, R. Gordon, A.B. Stilgoe, J. Romero, A.G. White, R. Fickler, A. Willner, G.D.
Xie, B. McMorran, A.M. Weiner, Roadmap on structured light. J. Opt. 19, 013001 (2017)
33. D.L. Andrews, Structured Light and Its Applications: An Introduction to Phase-Structured
Beams and Nanoscale Optical Forces (Elsevier Academic Press, San Diego, 2008)
34. D.H. Froula, D. Turnbull, A.S. Davies, T.J. Kessler, D. Haberberger, J.P. Palastro, S.W. Bahk,
I.A. Begishev, R. Boni, S. Bucht, J. Katz, J.L. Shaw, Spatiotemporal control of laser intensity.
Nat. Phot. 12, 262–265 (2018)
35. D. Turnbull, S.W. Bahk, I.A. Begishev, R. Boni, J. Bromage, S. Bucht, A. Davies, P. Franke,
D. Haberberger, J. Katz, T.J. Kessler, A.L. Milder, J.P. Palastro, J.L. Shaw, D.H. Froula, Flying
focus and its application to plasma-based laser amplifiers. Plasma Phys. Contr. Fusion 61,
014022 (2019)
36. T.T. Simpson, D. Ramsey, P. Franke, N. Vafaei-Najafabadi, D. Turnbull, D.H. Froula, J.P.
Palastro, Nonlinear Spatiotemporal Control of Laser Intensity. Preprint arXiv:2009.11783
37. K. Henderson, C. Ryu, C. MacCormick, M.G. Boshier, Experimental demonstration of painting
arbitrary and dynamic potentials for Bose-Einstein condensates. New J. Phys. 11, 043030
(2009)
38. L.P. Pitaevskii, S. Stringari, Bose-Einstein Condensation (Oxford University Press, Oxford,
2003)
39. H. Feshbach, A.K. Kerman, R.H. Lemmer, Intermediate structure and doorway states in nuclear
reactions. Ann. Phys. 41, 230–286 (1967)
40. C. Chin, R. Grimm, P. Julienne, E. Tiesinga, Feshbach resonances in ultracold gases. Rev. Mod.
Phys. 82, 1225–1286 (2010)
41. G. Roati, M. Zaccanti, C. D’Errico, J. Catani, M. Modugno, A. Simoni, M. Inguscio, G.
Modugno, 39 K Bose-Einstein condensate with tunable interactions. Phys. Rev. Lett. 99, 010403
(2007)
42. E.R.I. Abraham, W.I. McAlexander, J.M. Gerton, R.G. Hulet, R. Côté, A. Dalgarno, Singlet
s-wave scattering lengths of 6 Li and 7 Li. Phys. Rev. A 53, R3713 (1996)
43. S.E. Pollack, D. Dries, M. Junker, Y.P. Chen, T.A. Corcovilos, R.G. Hulet, Extreme tunability
of interactions in a 7 Li Bose-Einstein condensate. Phys. Rev. Lett. 102, 090402 (2009)
44. F. Wang, X. Li, D. Xiong, D. Wang, A double species 23 Na and 87 Rb Bose-Einstein condensate
with tunable miscibility via an inter-species Feshbach resonance. J. Phys. B At. Mol. Opt. Phys.
49, 015302 (2016)
45. P.O. Fedichev, Yu. Kagan, G.V. Shlyapnikov, J.T.M. Walraven, Influence of nearly resonant
light on the scattering length in low-temperature atomic gases. Phys. Ref. Lett. 77, 2913–2916
(1996)
46. R. Ciuryło, E. Tiesinga, P.S. Julienne, Optical tuning of the scattering length of cold alkaline-
earth-metal atoms. Phys. Rev. A 71, 030701(R) (2005)
47. K. Enomoto, K. Kasa, M. Kitagawa, Y. Takahashi, Optical Feshbach resonance using the
intercombination transition. Phys. Rev. Lett. 101, 203201 (2008)
48. M. Yan, B.J. DeSalvo, B. Ramachandhran, H. Pu, T.C. Killian, Controlling condensate collapse
and expansion with an optical Feshbach resonance. Phys. Ref. Lett. 110, 123201 (2013)
220 6 Introduction

49. D.J. Papoular, G.V. Shlyapnikov, J. Dalibard, Microwave-induced Fano-Feshbach resonances.


Phys. Rev. A 81, 041603(R) (2010)
50. B. Marcelis, B. Verhaar, S. Kokkelmans, Total control over ultracold interactions via electric
and magnetic fields. Phys. Rev. Lett. 100, 153201 (2008)
51. S. Jochim, M. Bartenstein, A. Altmeyer, G. Hendl, S. Riedl, C. Chin, J.H. Denschlag, R. Grimm,
Bose-Einstein condensation of molecules. Science 302, 2101 (2003)
52. I. Vidanović, A. Balaž, H. Al-Jibbouri, A. Pelster, Nonlinear Bose-Einstein-condensate dynam-
ics induced by a harmonic modulation of the s-wave scattering length. Phys. Rev. A 84, 013618
(2011)
53. Y.V. Kartashov, B.A. Malomed, L. Torner, Solitons in nonlinear lattices. Rev. Mod. Phys. 83,
247–306 (2011)
54. S.-L. Zhang, Z.-W. Zhou, B. Wu, Phys. Rev. A 87, 013633 (2013)
55. E. Kengne, A. Lakhssassi, W.M. Liu, Non-autonomous solitons in inhomogeneous nonlinear
media with distributed dispersion. Nonlinear Dyn. 97, 449(R) (2019)
56. S.L. Cornish, S.T. Thompson, C.E. Wieman, Formation of bright matter-wave solitons during
the collapse of attractive Bose-Einstein condensates. Phys. Rev. Lett. 96, 170401 (2006)
57. L.W. Clark, A. Gaj, L. Feng, C. Chin, Collective emission of matter-wave jets from driven
Bose-Einstein condensates. Nature 551, 356–359 (2017)
58. H. Fu, L. Feng, B.M. Anderson, L.W. Clark, J. Hu, J.W. Andrade, C. Chin, K. Levin, Density
waves and jet emission asymmetry in Bose fireworks. Phys. Rev. Lett. 121, 243001 (2018)
59. J.H.V. Nguyen, M.C. Tsatsos, D. Luo, A.U.J. Lode, G.D. Telles, V.S. Bagnato, R.G. Hulet,
Parametric excitation of a Bose-Einstein condensate: from Faraday waves to granulation. Phys.
Rev. X 9, 011052 (2019)
60. FKh. Abdullaev, J.G. Caputo, R.A. Kraenkel, B.A. Malomed, Controlling collapse in Bose-
Einstein condensation by temporal modulation of the scattering length. Phys. Rev. A 67(2003),
013605 (2003)
61. H. Saito, M. Ueda, Dynamically stabilized bright solitons in a two-dimensional Bose-Einstein
condensate. Phys. Rev. Lett. 90, 040403 (2003)
62. G.D. Montesinos, V.M. Pérez-García, H. Michinel, Stabilized two-dimensional vector solitons.
Phys. Rev. Lett. 92(2004), 133901 (2004)
63. A. Itin, T. Morishita, S. Watanabe, Reexamination of dynamical stabilization of matter-wave
solitons. Phys. Rev. A 74, 033613 (2006)
64. B.A. Malomed, (INVITED) Vortex solitons: old results and new perspectives. Phys. D
399(2019), 108–137 (2019)
65. M. Matuszewski, E. Infeld, B.A. Malomed, M. Trippenbach, Fully three dimensional breather
solitons can be created using Feshbach resonances. Phys. Rev. Lett. 95, 050403 (2005)
66. M.A. Hoefer, J.J. Chang, C. Hamner, P. Engels, Dark-dark solitons and modulational instability
in miscible two-component Bose-Einstein condensates. Phys. Rev. A 84, 041605(R) (2011)
67. P.G. Kevrekidis, D.J. Frantzeskakis, Solitons in coupled nonlinear Schrödinger models: a survey
of recent developments. Rev. Phys. 1, 140–153 (2016)
68. C. D’Errico, A. Burchianti, M. Prevedelli, L. Salasnich, F. Ancilotto, M. Modugno, F. Minardi,
C. Fort, Observation of quantum droplets in a heteronuclear bosonic mixture. Phys. Rev. Res.
1, 033155 (2019)
69. Z.X. Liang, Z.D. Zhang, W.M. Liu, Dynamics of a bright soliton in Bose-Einstein condensates
with time-dependent atomic scattering length in an expulsive parabolic potential. Phys. Rev.
Lett. 94, 050402 (2005)
70. W.-P. Zhong, R.-H. Xie, M. Belić, N. Petrović, G. Chen, Exact spatial soliton solutions of
the two-dimensional generalized nonlinear Schrödinger equation with distributed coefficients.
Phys. Rev. A 78, 023821 (2008)
71. W.-P. Zhong, M.R. Belić, G. Assanto, B.A. Malomed, T.W. Huang, Self-trapping of scalar and
vector dipole solitary waves in Kerr media. Phys. Rev. A 83, 043833 (2011)
72. J. Belmonte-Beitia, V.M. Perez-Garcia, V. Vekslerchik, V.V. Konotop, Localized nonlinear
waves in systems with time- and space-modulated nonlinearities. Phys. Rev. Lett. 100, 164102
(2008)
References 221

73. S. Rajendrana, P. Muruganandam, M. Lakshmanan, Bright and dark solitons in a quasi-1D


Bose Einstein condensates modelled by 1D Gross Pitaevskii equation with time-dependent
parameters. Phys. D 239, 366–386 (2010)
74. Q.-Y. Li, Z.-D. Li, L. Li, G.-S. Fu, Nonautonomous bright and dark solitons of Bose-Einstein
condensates with Feshbach-managed time-dependent scattering length. Opt. Commun. 283,
3361–3366 (2010)
75. B. Li, X.-F. Zhang, Y.-Q. Li, W.M. Liu, Propagation and interaction of matter-wave solitons
in Bose-Einstein condensates with time-dependent scattering length and varying potentials. J.
Phys. B At. Mol. Opt. Phys. 44, 175301 (2011)
76. W.B. Cardoso, J. Zeng, A.T. Avelar, D. Bazeia, B.A. Malomed, Bright solitons from the non-
polynomial Schrödinger equation with inhomogeneous defocusing nonlinearities. Phys. Rev.
E 88, 025201 (2013)
77. R. Radha, P.S. Vinayagam, J.B. Sudharsan, W.-M. Liu, B.A. Malomed, Engineering bright
solitons to enhance the stability of two-component Bose-Einstein condensates. Phys. Lett. A
379, 2977–2983 (2015)
78. E. Kengne, W.M. Liu, B.A. Malomed, Spatiotemporal engineering of matter-wave solitons in
Bose-Einstein condensates. Phys. Rep. 899, 1–62 (2021)
79. M. Belić, N. Petrović, W.-P. Zhong, R.H. Xie, G. Chen, Analytical light bullet solutions to the
generalized (3+1)-dimensional nonlinear Schrödinger equation. Phys. Rev. Lett. 101, 123904
(2008)
80. N. Petrović, M. Belić, W.-P. Zhong, R.H. Xie, G. Chen, Exact spatiotemporal wave and soliton
solutions to the generalized (3+1)-dimensional Schrödinger equation for both normal and
anomalous dispersion. Opt. Lett. 34, 1609–1611 (2009)
81. W.-P. Zhong, M. Belić, Three-dimensional optical vortex and necklace solitons in highly non-
local nonlinear media. Phys. Rev. A 79, 023804 (2009)
82. W.-P. Zhong, M. Belić, G. Assanto, T. Huang, Three-dimensional spatiotemporal vector solitary
waves. J. Phys. B: At. Mol. Opt. Phys. 44, 095403 (2011)
83. S.-L. Xu, W.-P. Zhong, M. Belić, Three-dimensional spatiotemporal vector solitary waves in
coupled nonlinear Schrödinger equations with variable coefficients. J. Opt. Soc. Am. B 30,
113–122 (2013)
84. S.V. Manakov, On the theory of two-dimensional stationary self-focusing of electromagnetic
waves. Zh. Eksp. Teor. Fiz. 65, 505 (1973). Sov. Phys. JETP 38, 248 (1974)
85. S.K. Suslov, On integrability of nonautonomous nonlinear Schrödinger equations. Proc. Am.
Math. Soc. 140, 3067–3082 (2012)
86. A. Debnath, A. Khan, On solving cubic-quartic nonlinear Schrödinger equation in a cnoidal
trap. Eur. Phys. J. D 74, 184 (2020)
87. B.A. Malomed, Yu.A. Stepanyants, The inverse problem for the Gross-Pitaevskii equation.
Chaos 20, 013130 (2010)
88. Y.S. Kivshar, B.A. Malomed, Dynamics of solitons in nearly integrable systems. Rev. Mod.
Phys. 61, 763 (1989)
89. L.D. Landau, E.M. Lifshitz, The theory of the dispersion of magnetic. Phys. Z. Sowjetunion
8, 153 (1935)
90. A.M. Kosevich, B.A. Ivanov, A.S. Kovalev, Magnetic solitons. Phys. Rep. 194, 117 (1990)
91. H.J. Mikeska, M. Steiner, Solitary excitations in one-dimensional magnets. Adv. Phys. 40, 191
(1991)
92. T.L. Gilbert, A phenomenological theory of damping in ferromagnetic materials. IEEE Trans.
Magn. 40, 3443 (2004)
93. Li. Zai-Dong, He. Peng-Bin, Liu Wu-Ming, Dynamics of magnetization in ferromagnet with
spin-transfer torque. Chin. Phys. B 23(11), 117502 (2014)
Chapter 7
Dynamics of One-Dimensional
Condensates with Time Modulation
of the Scattering Length and Trapping
Potential

Abstract This chapter deals with the dynamics of one-dimensional Bose-Einstein


condensate models with varying scattering length trapped in time-dependent external
potentials. Through the Gross-Pitaevskii equations with either the cubic or cubic-
quintic nonlinearities, we apply diverse methods and techniques such as the direct
method (DM), the ansatz method (AM) or the phase-imprinting technique (PIT) to
investigate the generation of matter-wave solitons in one-component BECs. Also,
we study analytically the phenomenon of the modulational instability and the soli-
ton management for BECs with loss/gain of atoms. For BECs with both two- and
three-bodies inter-atomic interactions, we establish that the competing cubic-quintic
nonlinearity of the GP equation of the BEC models under consideration induces
propagating dark (bright) solitons and double-kink solitons in the NLS equation
with self-steepening and self-frequency shift.

7.1 Non-autonomous Solitons in Bose-Einstein Condensates


with a Spatially Modulated Scattering length

This Section deals with BEC models with a spatially modulated scattering length,
diverse versions of which have been elaborated theoretically by many scientists
[1–6]. We apply the phase-imprinting technique, a relatively new tool used for wave-
function engineering in BEC models, to reduce the GP equation that describes the
wave-function of such BEC models to a non-autonomous cubic derivative NLS equa-
tion with a time-dependent HO potential. It may be extended to manage the BEC
wave-function by means of absorption provided by proximity to a resonance with
frequencies of external laser illumination. The action of the phase-imprinting tech-
nique onto BEC amounts to modifying the phase pattern in the mean-field wave
function. As a result of this action, BEC atoms experience the action of a spatially
varying light-induced potential, and acquire thus the corresponding phase. One of
the main advantages of the application of PIT is that it conserves the total number of
BEC atoms.

© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2022 223
E. Kengne and W. Liu, Nonlinear Waves,
https://doi.org/10.1007/978-981-19-6744-3_7
224 7 Dynamics of One-Dimensional Condensates …

7.1.1 Introduction

In the NLS equation with distributed coefficients


 2   
∂φ(x, t) ∂ ∂ |φ|2 φ
i + + 2 
g (x, t) |φ(x, t)| + V (x, t) φ(x, t) + i β = 0,
∂t ∂x2 ∂x
(7.1)
the last term, generally referred to as the derivative cubic term, represents the delayed
nonlinear response of the system. Equation (7.1) is known to model a variety of phe-
nomena such as the propagation of finite-amplitude Alfvén waves in directions nearly
parallel to the external magnetic field in a plasma [7]. Other physical realizations
of the NLS equation of form (7.1) are provided by convection in binary fluids [8]
and wave propagation in electric nonlinear transmission networks [3]. Also, the NLS
equation of form (7.1) describes the behavior of large-amplitude magnetohydrody-
namic waves propagating in an arbitrary direction with respect to the magnetic field
in high-β plasmas [9, 10]. In the context of nonlinear optics, Eq. (7.1) can be sup-
plemented by the derivative self-steepening term to investigate short pulses with the
local Kerr nonlinearity [11–14]. In the special case of constant cubic nonlinearity,
Eq. (7.1) in the sense of the external potential reduces to the-well known integrable
derivative cubic NLS equation [15, 16]. In the context of the present section, two
basic questions arise, as concerns equations of type (7.1): (i) The question of the
introduction of a GP model that describes the impact of the cubic derivative nonlin-
earity on the condensates. (ii) The question of the impact of the derivative cubic term
of the GP equation on the modulational instability phenomenon in BEC. The present
section intends to give positive answer to the above two questions. To give adequate
answer to these two questions, we apply the PIT to reduce the BEC GP equation to
a NLS equation of type (7.1) [4, 5].

7.1.2 Cubic Inhomogeneous Nonlinear Schrödinger


Equation

For the nonlinearity management of matter-wave solitons of the BECs, the Feshbach-
resonance may be used by modulating the scattering length of inter-atomic collisions
temporarily, spatially, or spatiotemporally, leading to generation of various novel
nonlinear phenomena [17–19]. For example, time-dependent modulation of the scat-
tering length can be used to stabilize attractive two-dimensional BECs against the
critical collapse [20]. Also, time modulating of the scattering length can lead to the
creation of robust matter-wave breathers in one-dimensional BECs [21]. It has been
showed that for BECs with a spatially varying scattering length, atomic matter waves
exhibit novel features [1, 22–24].
7.1 Non-autonomous Solitons in Bose-Einstein Condensates … 225

The present Section starts with the one-dimensional GP equation


 2 
∂u(x, t) ∂
i + + g(x, t) |u(x, t)| + V (x, t) u(x, t) = 0,
2
(7.2)
∂t ∂x2

with the time-varying external harmonic potential

V (x, t) = −α(t)x 2 . (7.3)

As it is seen from Eq. (7.3), the strength of the harmonic trap α(t) may take any
sign, the negative sign corresponding to the confining potential, while positive one
is associated with the expulsive potential. For the experimental purpose, the strength
α(t) is typically fixed to a constant value.
In the case of the cigar-shaped Bose-Einstein condensates, the self-consistent
reduction of the three-dimensional GP equation to the one-dimensional one with the
external potential can be carried out by using a multiple-scale expansion [6]. For
this aim, one must introduce a small parameter δ defined as δ 2 = (N as /a0 )α  1
(where as is the s-wave scattering length), which can be used to evaluate the relative
strength of the two-body inter-atomic interactions as compared to the kinetic energy
of the atoms. This small parameter δ generally defines the ratio of the tight transverse

confinement to a characteristic scale along the longitudinal axis, as a⊥ /a0 ∼ δ α,
especially in the case of quasi-one-dimensional where the dynamics along the cigar’s
axis is of primary interest.
It is important to remember that the relation
   
δ r2 δx 1 2
(r, t) = √ exp(−iω⊥ t) exp − 2
u , δ ω⊥ t (7.4)
a⊥ as 2a⊥ a⊥ 2

can be used to connect the rescaled one-dimensional mean-field wave-function


u(x, t) of the condensate of the GP Eq. (7.2) to the underlying three-dimensional
order parameter (r, t). Here, r = (y, z), y and z being the coordinates in the trans-
verse plane, and ω⊥ is the confining HO frequency in this plane. In the following, we
assume that the external potential V (x, t) of the GP Eq. (7.2) is measured in units of
2 a⊥
2
/8m.
Under the above conditions, the functional parameter g(x, t) of the two-body
inter-atomic interaction is positive and negative for respectively the focusing or defo-
cusing nonlinearity. In this section, we consider a GP equation with a spatiotemporal
modulation of the interaction coefficient g(x, t), introduced as a linear function with
time-varying coefficients of the spatial variable x:

g(x, t) = x g0 (t),
g (t) +  (7.5)

where  g0 (t) are two real functional parameters.


g (t) and 
For the inhomogeneous NLS equation (7.2) with the time-dependent harmonic
potential (7.3) and the focusing sign of the nonlinearity, the phenomenon of the mod-
226 7 Dynamics of One-Dimensional Condensates …

ulational instability has been theoretically studied in work [25] and experimentally
demonstrated in work [26]. Following the results found in works [4] and [5], we
introduce in this Section modified version of the GP equation (7.2) with potential
(7.3) containing a cubic derivative term, and address the question of the MI of the
new GP equation.
Inhomogeneous NLS Equation with a Cubic Derivative Term Here, we employ
the PIT to reduce Eq. (7.2) with potential (7.3) to an inhomogeneous NLS equation
with a cubic derivative term. For this aim, we introduce a new wave-function, ψ(x, t),
related to the mean-field wave-function u(x, t) as follows [5]

u(x, t) = ψ(x, t) exp [−iθ (x, t)] , (7.6)

where θ (x, t) is the imprinted real phase to be constructed according to the following
two equations:

∂θ/∂ x = −3β(t) |ψ|2 , (7.7)


 
∂θ/∂t = iβ(t) ψ∂ψ ∗ /∂ x − ψ ∗ ∂ψ/∂ x + 9β 2 (t) |ψ|4 , (7.8)

β(t) being a time-varying real coefficient standing for the phase-imprint strength [4,
5].
Inserting ansatz (7.6) into the GP equation (7.2), we arrive at the derivative NLS
equation [5]
 2 
∂ψ ∂ ∂(|ψ|2 ψ)
i + + V (x, t) + g(x, t) |ψ|2
ψ + 4iβ(t) = 0. (7.9)
∂t ∂x2 ∂x

It follows from Eqs. (7.6)–(7.8) that |u(x, t)|2 = |ψ(x, t)|2 , meaning that the norm
under ansatz (7.6) is conserved, i.e., the total number N of atoms of the condensate
under transformation (7.6)–(7.8) is conserved.
Although the PIT scheme leading to Eq. (7.9) was not yet realized experimentally
in Bose-Einstein condensates, a similar experimental result, viz., creation of quasi-
one-dimensional dark solitons by means of sufficiently strong PIT was reported [27].
Modulational Instability in the Cubic Derivative NLS Equation
Before passing to the study of MI in the cubic derivative NLS (7.9), we first use
the results of works [3] and [10] to recapitulate the results obtained for the MI for a
similar equation with constant coefficients when the external potential is absent, that
is, an equation of the form
   
∂ψ ∂2 ∂ |ψ|2 ψ
i + + g0 |ψ| ψ + 4iβ
2
= 0, (7.10)
∂t ∂x2 ∂x

where g0 and β are two real constants. It is obvious that Eq. (7.10) admits the
continuous-wave solution
7.1 Non-autonomous Solitons in Bose-Einstein Condensates … 227

ψ(x, t) = φ0 exp iq x − i[q 2 + (4βq − g0 ) φ02 ]t , (7.11)

where φ0 and q are two arbitrary real constants. The MI of the CW solution (7.11)
can be analyzed by linear stability analysis [14] by perturbing the CW solution (7.11)
as follows
 
ψ(x, t) = [φ0 + ε(x, t)] exp iq x − i[q 2 + (4βq − g0 ) φ02 ]t , (7.12)

where ε(x, t) is a small complex perturbation satisfying the condition |ε(x, t)| 
|φ0 |. If we substitute ansatz (7.12) into Eq. (7.10) and neglect higher order in ε(x, t),
we derive to the Bogoliubov–de Gennes (BdG) equation:

∂ε ∂ 2ε   ∂ε ∂ε∗  
i + 2 + 2i q + 4βφ02 + 4iβφ02 + φ02 (g0 − 4βq) ε + ε∗ = 0.
∂t ∂x ∂x ∂x
(7.13)
Being a linear partial differential equation, we can seek eigenmodes of the perturba-
tion as
 
ε(x, t) = U1 exp [i (Qx − t)] + U2∗ exp −i Qx − ∗ t , (7.14)

to obtain the dispersion relation connecting the perturbation wavenumber Q and the
complex angular frequency of the perturbation:
  2  
− 2Q q + 4βφ02 = Q 2 φ02 16β 2 φ02 + 8qβ − 2g0 + Q 2 . (7.15)

It follows from Eq. (7.15) that the MI occurs when the perturbation wavenumbers Q
satisfies that condition
 
Q 2 < −2φ02 8β 2 φ02 + 4qβ − g0 , (7.16)

which may hold for independently on the sign of g0 and β, that is, for both focusing
and defocusing signs of the nonlinearities. Indeed, condition (7.16) is valid only
when 8β 2 φ02 + 4qβ − g0 < 0, that is, for values of β belonging to interval

q+ q 2 + 2g0 φ02 −q + q 2 + 2g0 φ02
Dβ ≡ − <β< . (7.17)
4φ02 4φ02

For such β to exist, it is necessary that

q 2 + 2g0 φ02 > 0. (7.18)

In the case of the focusing nonlinearity (g0 = +1), condition (7.18) is satisfied for
all q and φ0 satisfying the condition |q| + |φ0 | > 0. In the case of the defocusing
nonlinearity (g0 = −1), the necessary condition (7.18) holds only when q and φ0
228 7 Dynamics of One-Dimensional Condensates …

satisfy the constraint q 2 − 2φ02 > 0. Thus, Eq. (7.18) is the necessary condition for
MI of the CW solutions for the cubic derivative NLS equation (7.10). When β = 0,
Eq. (7.10) is just the standard NLS equation whose MI results from the Benjamin
and Feir [28] criterion g0 > 0.
Taking into account the condition (7.16), the MI growth rate (gain) for the deriva-
tive NLS equation in the free space reads
 
|Im | = |Q| 2g0 − 16β 2 φ02 − 8qβ φ02 − Q 2 . (7.19)

Introducing the function B(β) as

B(β) = 2g0 − 16β 2 φ02 − 8qβ,


 
it is obvious that B(β) reaches its maximum at the critical point βc = −q/ 4φ02 ,
which does not depend on g0 . This shows that the presence of the imprint parameter
β significantly modifies the instability domain and brings new effects. For a better
understanding, we consider, without loss of generality, the special case when g0 =
±1. For this special case, we depict in Fig. 7.1 the MI gain defined as per Eq. (7.19),
for different numerical values of the imprint parameter β with φ0 = 1 and q = 2. As
we can see from plots of Fig. 7.1, there are two scenarios, depending on whether the
imprint parameter β, belonging to interval (7.17), is above the critical value (β ≥ βc )
or below the critical value (β ≤ βc ). As we can see in the plots of the top panels of
Fig. 7.1 generated for β ≥ βc , the gain decreases with decreasing in the values of |β|.
Plots of the bottom panels, obtained with β ≤ βc , show that the MI gain increases
with the decreasing in the values of |β|. Thus, the imprint parameter β, when taken
above the critical value βc , softens MI; on the other hand, the MI enhances when β
falls below the critical value βc . When comparing plots of the left panels in 7.1 (for the
focusing nonlinearity with g0 = +1) with those of the right panels (the defocusing
nonlinearity with g0 = −1), we can see that the modulational instability is stronger
in the case of the focusing nonlinearity.
MI in the Inhomogeneous Cubic Derivative NLS Equation With the Harmonic
Potential Now, we turn to the MI for the NLS equation (7.9) with a cubic deriva-
tive term in the presence of an external HO potential. To examine the MI in this
general case, we follows the work [5] and introduce the following modified lens
transformation (LT)

1
ψ(x, t) = φ(X, T ) exp[i f (t)x 2 + η(t)]. (7.20)
(t)

Here, T = T (t), (t), η(t), and f (t) are real functions of time, and X (x, t) =
x −1 (t). The lens transformation was originally introduced by Talanov [30] as an
invariant transformation for the two-dimensional NLS equation, which changes the
scale of the coordinates and adds the radial chirp to the wave-function. Its significance
is stressed by its compatibility with adiabatic variation of the wave-function and with
7.1 Non-autonomous Solitons in Bose-Einstein Condensates … 229

Fig. 7.1 The MI gain as produced by Eq. (7.19) for three values of the imprint parameter β,
with φ0 = 1 and q = 2. The top panel includes three values of β > βc = −0.5: β = −0.4, 0.3,
and 0.2 (the blue, red, and green lines, respectively). The bottom panel includes three values of
β < βc = −0.5: β = −0.8, 0.7, and 0.6 (the blue, red, and green lines, respectively). The left
and right panels correspond to the focusing and defocusing nonlinearities, respectively. The figure
reproduces known results for MI produced by the derivative NLS equation with constant coefficients
[10]. Reprint from Ref. [29], Copyright 2022, with permission from American Physical Society

the dynamics driven by the critical collapse in two-dimensional geometry [31, 32].
It is important to note that the LT can be applied as well to the two-dimensional NLS
equation in the presence of the isotropic HO potential [32].
To preserve the scaling of Eq. (7.9) we impose that

dT 1
= 2 . (7.21)
dt (t)

Next, we demand that

df
+ 4 f 2 + α = 0, (7.22a)
dt
1 d
4 f (t) − = 0, (7.22b)
dt
dη 1 d
+ 2 f (t) − = 0, (7.22c)
dt dt
β(t) =  g (t) f −1 (t)/8. (7.22d)
230 7 Dynamics of One-Dimensional Condensates …

Inserting ansatz (7.20) in Eq. (7.9) and taking into account Eqs. (7.21)–(7.22d) yield

∂φ ∂ 2φ ∂(|φ|2 φ)
i + + λ 0 (t) |φ| 2
φ + iλ(t) = 0, (7.23)
∂T ∂ X2 ∂X
where real the time-varying functions λ0 (t) and λ1 (t) are given as

β(t)
λ0 (t) = 
g0 (t) exp[2η(t)], λ(t) = 4 exp[2η(t)]. (7.24)
(t)

This shows that the invariance of the inhomogeneous NLS equation with a cubic
derivative term with respect to LT is maintained.
Now, we can solve Eqs. (7.22b), (7.22c) and (7.21) in terms of f (t) to obtain

 t 
(t) = (0) exp 4 f (υ) dυ , (7.25a)
0
t
η(t) = 2 f (υ) dυ + η(0), (7.25b)
0
t  s 
−2
T (t) = (0) exp −8 f (υ) dυ ds + T (0). (7.25c)
0 0

We have thus obtained that the problem of finding time-varying parameters (t), η(t),
and T (t) is reduced to finding f (t), that is, to solving the Riccati equation (7.22a).
According to Ansatz (7.20), the time-varying parameter η(t) affects the total
number N of atoms as follows,

+∞ +∞
−1
N= |ψ| d x = (t) exp[η(t)]
2
|φ|2 d X,
−∞ −∞

if −1 (t) exp[η(t)] = constant. It follows from Eqs. (7.25a) and (7.25b) that

t
−1 −1
(t) exp[η(t)] = (0) exp[−2 f (υ) dυ + η(0)],
0

so that the total number N of the condensate atoms exponentially grows if f (t),
solution of the Riccati equation (7.22a), is negative, and exponentially decays if f (t)
is positive. In other words, negative f (t) represents the feeding of atoms into the
condensate, while positive f (t) implies loss of atoms. While f (t) depends on the
strength of the HO trap, α(t), its sign does not dependent on that of α(t).
7.1 Non-autonomous Solitons in Bose-Einstein Condensates … 231

To investigate the phenomenon of the MI for the NLS equation (7.23) with dis-
tributed coefficients, we introduce the perturbed solutions as [5]

 T 
φ = [φ0 + ε(X, T )] exp −i Q X − i (υ) dυ . (7.26)
0

Here, (T ) is a real time-dependent function representing the nonlinear frequency


shift, φ0 is a real constant, Q is the wavenumber of the carrier, and ε(X, T ) is a small
perturbation. Substituting Eq. (7.26) into Eq. (7.23) and linearizing the resulting
equation with respect to ε(X, T ), we obtain, for (T ) taken as

(T ) = Q 2 − λ0 (t)φ02 − Qφ02 λ(t), (7.27)

that the small perturbation ε(X, T ) satisfies the linear partial differential equation

∂ε ∂ 2ε  2  ∂ε 2 ∂ε
∗  
i + + 2i λφ − Q + iλφ + φ02 (λ0 + λQ) ε + ε∗ , (7.28)
∂T ∂X 2 0
∂X 0
∂X
∗ standing for the complex conjugation. We can seek the nontrivial solutions to Eq.
(7.28) as

 T   T 
ε = U1 exp i K X − i ω(υ) dυ + U2 exp −i K X + i ω∗ (υ) dυ . (7.29)

0 0

Here, K and ω are the wavenumber and the complex frequency of the perturbation,
T
K X − 0 ω(υ) dυ is the modulation phase, and U1 and U2 are two complex ampli-
tudes satisfying the condition that |U1 | + |U2 | > 0. To observe the phenomenon of
the MI, the complex angular frequency ω of the perturbation must have a nonzero
imaginary part. The following time-dependent dispersion relation is obtained by
inserting expression (7.29) into Eq. (7.28):
  2  
ω − 2K λφ02 − Q − K 2 K 2 + φ02 λ2 φ02 − 2Qλ − 2λ0 = 0. (7.30)

Asking that ω has a nonzero imaginary part leads to the following relationship on
φ0 , K , and Q :  
K 2 + φ02 λ2 φ02 − 2Qλ − 2λ0 < 0, (7.31)

which forms the MI criterion for the cubic derivative NLS equation (7.23). If the
criterion (7.31) holds, the local MI gain is given by

|Im ω(t)| = |K | φ02 2Qλ(t) + 2λ0 (t) − λ2 (t)φ02 − K 2 . (7.32)
232 7 Dynamics of One-Dimensional Condensates …

In the special case of constant λ0 and λ, it follows from Eqs. (7.21)–(7.22d) that
g0 (t) will satisfy the nonlinear second-order ODE
β must be constant, while α(t) and 
 
d 2
g0 dg0 2

g0 − 2 − 4α
g02 = 0. (7.33)
dt 2 dt

We thus obtain that in the special case of constant λ0 , λ and β, the problem of finding
(t), η(t), and T (t) amounts to solving Eq. (7.33). This Eq. (7.33) admits interesting
 
and important special solutions in  g0 (t) and α(t). For instance,  a0 exp 
g0 (t) =  λt

produces constant α = −λ /4.2

Following work [25], one of the most interesting cases in the setting with the
external harmonic potential is obtained by taking α(t) as follows:

α(t) = A(t + t ∗ )−2 , (7.34)

where A and t ∗ are two real constants; A and t ∗ determine the strength of the potential
and its width at t = 0, respectively. The corresponding  g0 (t) can be obtained by
inserting expression (7.34) in Eq. (7.33); this results to
 m √
4λ0 β t + t∗ −1 ± 1 − 16A 1

g0 (t) = ,m= , A< . (7.35)
λ (0) t∗ 2 16

Condition A < 16 1
allows us to investigate the MI for both the confining and expul-
sive potentials (A < 0 and A > 0, respectively). It is important to note that the
situation t ∗ < 0 describes BEC in a shrinking trap, while the situation of positive t ∗
corresponds to a broadening condensate. Combining expressions (7.35) with Eqs.
(7.22a)–(7.22d) determines all time-varying parameters; in particular,
 2m+1 
t∗ t
T (t) = +1 −1 .
(2m + 1) 2 (0) t∗

In order to secure
√ the variation of T from zero
√ to infinity, it is necessary to take
m = (−1 + 1 − 16A )/2 and m = −(1 + 1 − 16A)/2 in respectively the cases
of the broadening and shrinking trap. In the latter case, we focus on t varying from
0 to −t ∗ ; this provides the variation of T from 0 to +∞.
Considering the case of constant λ0 and λ, the MI growth rate does not depend
on time t , but depends on the phase-imprint parameter β as follows:

|Im ω(β)| = |K | φ02 −λ2 (β)φ02 + 2Qλ(β) + 2λ0 − K 2 . (7.36)

It is clearly seen from Eq. (7.36) that the variation of the gain, controlled by β, may
significantly modify the instability domain and therefore bring new effects. Indeed,
different numerical values of β lead to different instability diagrams, depending on
7.1 Non-autonomous Solitons in Bose-Einstein Condensates … 233

Fig. 7.2 The MI gain produced by Eq. (7.36) for three positive and three negative values of the
imprint parameter λ(β). a λ(β) = −0.5 (the solid line), λ(β) = −0.25 (the dashed line), and λ(β) =
−0.125 (the dotted-dashed line); b λ(β) = 0.125 (the solid line), λ(β) = 0.25 (the dashed line), and
λ(β) = 0.50 (the dotted-dashed line). Other parameters are Q = 1, φ0 = 1, and λ0 = 3/2. Reprint
from Ref. [29], Copyright 2022, with permission from American Physical Society

the sign of λ(β): negative λ(β) softens the instability, while positive one enhances the
instability. For a better understanding, we show in 7.2 the MI growth rate provided by
Eq. (7.36) versus the perturbation wavenumber K , for three values of negative λ(β)
and three values of positive λ(β). As we can see from 7.2a generated with negative
λ(β), the MI growth rate decreases with the increasing in the values of parameter λ;
from 7.2b, obtained with positive λ(β), we can see that the gain increases with the
increasing in the values of λ.
We can summarize the results of the MI in the case of constant λ0 and λ as follows.
For the occurrence of MI of Stokes wave solutions
 
φ = φ0 exp −i Q X − i Q 2 − λ0 φ02 − Qφ02 λ T ,

it is necessary and sufficient for the perturbation wavenumber K to satisfy the MI


criterion (7.31). Moreover, for given parameters φ0 , Q, and λ0 of the carrier, the
imprint parameter β should be chosen from the following condition:

λ2 (β)φ02 − 2Qλ(β) − 2λ0 < 0.

What happens when at least one of λ0 and λ is not constant. Considering for
example the strength α(t) of the HO potential to have the form

A 1
α(t) = ∗
, with A < ,
(t + t ) 2 16

we can integrate the Riccati equation (7.22a) and find its special solution as follows:

B 1± 1 − 16A
f (t) = , B= . (7.37)
t + t∗ 8
234 7 Dynamics of One-Dimensional Condensates …

With expression (7.37), the corresponding solution of Eq. (7.22a) reads

1 − 4B + C B (1 − 8B) (t + t ∗ )8B−1
f (t) = , (7.38)
C (1 − 8B) (t + t ∗ )8B + 4(t + t ∗ )

where
1 − 4B − 4 f (0)t ∗
C= .
f (0) (1 − 8B) (t ∗ )8B − B (1 − 8B) (t ∗ )8B−1

With the use of Eqs. (7.34) and (7.37), one obtains the following functional param-
eters from Eqs. (7.22a)–(7.22d):
 
t + t ∗ 4B
(t) = (0) , (7.39a)
t∗
 
 t + t∗ 
η(t) = 2B ln  ∗  + η(0), (7.39b)
t
g (t)(t + t ∗ )

β(t) = , (7.39c)
8B  
 ∗ 4B
1   t
T (t) = t + t∗ − t∗ . (7.39d)
(0) (1 − 4B) t + t∗

As we can see from Eqs. (7.39a)–(7.39d), β(t) is now a time-varying parameter and
depends on  g (t). In the case of positive t ∗ (broadening condensate), we take B < 1/4
and choose (0) so that T (t) will vary from 0 to +∞. For Bose-Einstein condensates
in the shrinking trap (t ∗ < 0), the appropriate choice of (0) and B > 1/4 demon-
strates that taking 0 ≤ t < t ∗ corresponds to the variation of T (t) from 0 to +∞.
It is obvious that when at least one of λ0 and λ varies with time, the MI growth rate
(gain) given by Eq. (7.32) will depend on time t. In such a situation, the variation of
the gain, related to the sign of t ∗ may significantly affect the MI domain and introduce
new effects: the instability is enhanced or attenuated by respectively t ∗ < 0 or t ∗ > 0,
as we can see from plots of Fig. 7.3. Figure 7.3 displays the MI gain produced by
Eq. (7.32), versus the perturbation wavenumber K , for three negative and three
positive values of t ∗ . Plots of Fig. 7.3a, realized with negative t ∗ (shrinking trap)
show that the MI gain increases with t ∗ . As we can see from plots of Fig. 7.3b,
generated with positive t ∗ (broadening trap), the MI gain decreases as t ∗ increases.
Different plots in this 7.3 are produced with the use of  g0 (t) = 2 exp (0.5t) and

g (t) = 2 exp (−t).
The above analysis informs us that the simplest and most interesting case in
the setting with the time-dependent external harmonic potential is the one with the
inverse-square time dependence of the trap strength given in Eq. (7.34) with A <
1/16. In this special case, the modified LT demonstrates the equivalence of the setting
to the cubic derivative NLS equation. The situation of derivative NLS equation with
distributed coefficients, suggests that frequencies of eigenmodes of the modulational
perturbations are either constant or effectively time-dependent.
7.1 Non-autonomous Solitons in Bose-Einstein Condensates … 235

Fig. 7.3 The MI gain produced by Eq. (7.32) for three positive and three negative values of t ∗ at
time t = 0. a The case of the shrinking trap with t ∗ = −0.6, −0.4, and −0.2 (the solid, dashed,
and dotted-dashed lines, respectively). b The case of the broadening trap with t ∗ = 0.1, 0.2, 0.4
(the solid, dashed, and dotted-dashed lines, respectively). Other parameters are√Q = 1, φ0 = 1,
A = −1, (0) = 1,√ η(0) = 0, 
g0 (t) = 2 exp (0.5t), 
g (t) = 2 exp (−t), B = (1 + 17 )/8 for plots
(a), and B = (1 − 17 )/8 for (b). Reprint from Ref. [29], Copyright 2022, with permission from
American Physical Society

7.1.3 Matter-Wave Solitons in an Inhomogeneous Nonlinear


Schrödinger Equation with the Spatiotemporal HO
Potential

In the present subsection, we follow the results of work [5] and present in the ana-
lytical form, matter-wave solitons of Eq. (7.23) when parameters λ0 and λ appearing
in Eq. (7.24) are all constant.
It is clearly seen that Eq. (7.23) admits the Stokes wave solutions of form
 
φ(X, T ) = φ0 exp −ik0 X + i λ0 φ02 + k0 λφ02 − k02 T , (7.40)

for constant φ0 and k0 , when λ0 and λ are all constant. Next, we seek a solution of
Eq. (7.23) in the Madelung form,

φ(X, T ) = R(X, T ) exp [i(X, T )] (7.41)

and arrive at the following system of equations for the real amplitude and phase:

− R∂/∂ T + ∂ 2 R/∂ X 2 − R (∂/∂ X )2 − λR 3 ∂/∂ X + λ0 R 3 = 0,


(7.42)
∂ R/∂ T + 2 (∂/∂ X ) (∂ R/∂ X ) + R∂ /∂ X + 3λR ∂ R/∂ X = 0.
2 2 2

It follows from the form of the Stokes wave solution (7.40) that solutions to system
(7.42) can be have the following traveling-wave form:
236 7 Dynamics of One-Dimensional Condensates …
 
R(X, T ) = φ0 + ρ(z), (X, T ) = −k0 z + ϕ(z) + λ0 φ02 + λk0 φ02 − k02 − k0 υ T,
(7.43)
where z = X − υT is the propagating coordinate in which υ is an arbitrary velocity.
If we insert ansatz (7.43) into the system of equations (7.42) and integrate the second
equation, we arrive to

dϕ C0 (υ + 2k0 ) 3λ
= + − (φ0 + ρ)2 , (7.44)
dz (φ0 + ρ)2 2 4

where C0 is a constant of integration. Inserting now Eq. (7.44) into the first equation
of system (7.42) yields the following elliptic ordinary differential equation
 2

= ζ 3 + 6
α ζ 4 + 4β γ ζ 2 + 4
δ ζ +
ε ≡ F(ζ ), (7.45)
dz

where

ζ (z) = [φ0 + ρ(z)]2 ,  =


α = −λ2 , β λυ−2λ0
,
ε= −4C02 ,
4
  (7.46)
4λ0 φ02 + 2λ 2k0 φ02 − C0 − 4υk0 − υ 2 − 4k02
γ =
 ;
6

here, 
δ and C0 are constants of integration. Solutions of Eq. (7.45) can be written in
terms of the Weierstrass elliptic function ℘ (z; g2 , g3 ) as [33, 34]

 
 d 1 1
f (ζ0 ) ℘ (z; g2 , g3 ) + F (ζ0 ) ℘ (z; g2 , g3 ) − F (ζ0 )
dz 2 24
1
+ F(ζ0 )F (ζ0 )
ζ (z) = ζ0 + 24 ,
 2
1 1
2 ℘ (z; g2 , g3 ) − F (ζ0 ) − F(ζ0 )F (I V ) (ζ0 )
24 48
(7.47)

where ζ0 is an arbitrary real constant, and the prime stands for d/dz. Invariants g2
and g3 of function ℘ (z; g2 , g3 ) are related to coefficients of F(ζ ) as [33]

g2 = 
α 
 − 4β δ + 3
γ 2 , g3 = 
α
γ 
 + 2β γ α
δ −δ2 − 
γ 3 −2 .
β (7.48)

To classify the behavior of solutions ζ (ξ ) and discriminate between periodic and soli-
tonlike solutions [34], we use the discriminant  of the Weierstrass elliptic function
℘ (z; g2 , g3 )
 = g23 − 27g32 . (7.49)
7.1 Non-autonomous Solitons in Bose-Einstein Condensates … 237

If  = 0, g2 ≥ 0, and g3 ≤ 0, ζ (ξ ) is a solitonlike solution and can be given by

F (ζ0 )
ζ (z) = ζ0 + √  , (7.50)
4 e1 − 1
24
F (ζ0 ) + 3e1 cosech2 3e1 z

if ζ0 is a simple zero of function F(ζ ) (here, e1 = (−g3 )1/3 ). In the following, the
constants of integration   are taken as 
δ and  δ =  = 0. In this special case, g2 =
3γ > 0, g3 = −
2
γ , and  = 0. Therefore, necessary condition for Eq. (7.50) to
3

define a nonnegative solitonlike solution is that  γ > 0 and 2β 2 − 3


αγ ≥ 0. Since
ζ (z) = [φ0 + ρ(z)] , the physical solution (7.50) must be nonnegative and bounded.
2

Considering properties of F(ζ ) [2], we obtain the conditions, expressed in terms


of coefficients of the basic equation, that determine the existence of the physical
solutions, (see Fig. 7.4 borrowed from Refs. [2, 29], which shows phase diagrams
associated to the physical solutions when  δ =  = 0).
Using the results of work [2], solitonlike solutions generated by Eq. (7.50) can be
cast in the following form:
    2   
γ 2β
  ± 4β 2 − 6
α
γ γ2 + 
3 γ − 1 cosh2 γ

3 3
z
2
ζ± (z) =       .
−3
α
γ 3 + 4β2 ± 2 4β2 − 6
α
γ −α
γ (5 + 
γ )2 cosh2 γ

3 3
z
2
(7.51)

Solutions (7.51) correspond to two simple roots of the polynomial F(ζ ) when
4β2 − 6αγ > 0, and are represented by phase diagrams (b), (c), (d), (f), (g), and (h)
in Fig. 7.4 (for further details, see Ref. [2]).
Now, we address the condition of the non-negativeness of solutions (7.51). For this
aim, we distinguish two cases, namely, case  γ = 1 associated with bright solitonlike
solution, and case 0 <  γ = 1 leading to both dark and bright solitonlike solutions.
(A): If 
γ = 1, function ζ (z) given in Eq. (7.51) will be nonnegative if and only if
the following two conditions are satisfied simultaneously:
 −1

α 2β 2 − 6
2 ± 4β α − 18
α < 2/3 (7.52)

and
  
 ± 4β
2β 2 − 6
α 2 ± 4β
2β 2 − 6
α − 18
α > 0. (7.53)

Particularizing parameters λ0 , φ0 , υ, k0 , and λ as follows λ0 = 1, φ0 = 1,


υ = 0.5, k0 = 1, and λ = 2.0625 leads to an example of the bright solitonlike
solution. With this set of parameters, conditions (7.52) and (7.53) are satisfied
for ζ+ (z) so that ζ+ (z) will be a nonnegative bounded bright soliton solution.
238 7 Dynamics of One-Dimensional Condensates …

Fig. 7.4 Phase diagrams associated with real bounded solutions, given by Eqs. (7.47) and (7.45),
with either 
δ =ε = 0 (see also Eq. (7.51) for this case) or 
δ = 0, 
ε < 0, and g2 = g3 = 0. Further
details can be found in Refs. [35] and [36] for the right and left plots, respectively. Reprint from
Ref. [29], Copyright 2022, with permission from American Physical Society

In Figs. 7.5, 7.6, and 7.7, we respectively show the effects of t ∗ , A, and β on
the density profile of the bright solitonlike solution, |u(x, t)|2 , at x = 3; here,
we use the numerical value (0) = 1. In each of Figs. 7.5, 7.6, and 7.7, the

Fig. 7.5 Density plots |u(x, t)|2 of solitary-wave solutions to Eq. (7.2) at x = 3 for three values
of parameter t ∗ in potential (7.34), given by expression (7.51) for ζ+ . The top and bottom panels
correspond to the confining potential (with A = −2) and expulsive one (with A = 1/17), respec-
tively, while the left and right panels correspond to broadening shrinking traps, respectively. Values
of other parameters are given in the text. Reprint from Ref. [29], Copyright 2022, with permission
from American Physical Society
7.1 Non-autonomous Solitons in Bose-Einstein Condensates … 239

left and the right panels are generated with respectively t ∗ > 0 (broadening)
and t ∗ < 0 (shrinking traps); plots of the top and bottom panels correspond
respectively to the confining (A < 0) and expulsive (A > 0) potentials. For
x = 3, Fig. 7.5 shows the time evolution of the wave density |u(x, t)|2 for three
different numerical values of t ∗ . It is clearly seen from different plots of Fig.
7.5 that the amplitude of the density profile decreases as t ∗ increases for the
broadening trap, and increases with the growth of t ∗ for the shrinking trap, when
A < 0 (in the case of the confining potential). In the case of expulsive poten-
tial (A > 0), the profile’s amplitude increases with t ∗ for both broadening and
shrinking traps. Figure 7.6 depicts the time evolution of the density |u(x, t)|2
for x = 3 and for three different values of A. Plots of this figure shows that, for
both the confining (top panels) and expulsive (bottom panels) potentials, the
profile’s amplitude decreases as parameter A increases; this situation happens
for both the broadening (plots of the left panels) and shrinking (plots of the right
panels) BEC traps. We can seen from plots of Fig. 7.7 where the wave density
|u(x, t)|2 is displayed for x = 3 and for different values of β that, irrespective
of the sign of A (confining/expulsive potential) and the sign of parameter t ∗
(broadening/shrinking trap), the density-profile’s amplitude decreases with the
increase of the values of the imprint parameter β.
(B): Let us now consider the situation when 0 <  γ = 1. In this situation, solutions
ζ± (z) given in Eq. (7.51) will be nonnegative if and only, if the following three
conditions are simultaneously satisfied:
γ2
3
(i) ≤ 1,
1−γ2
α
3 γ3
(ii)  < 1,
 2 
4β ± 2 4β − 62 α
γ −
αγ (5 + 
γ )2
     
(iii) 
γ  γ 2 − 1 2β  ± 4β2 − 6α
γ 2 − 
4β α
γ (5 +  2 − 6
γ )2 ± 2 4 β α
γ > 0.

For the following numerical values of various parameters, λ0 = 1, φ0 = 1, υ = 0.5, λ =


3.5625, and k0 = 1, conditions (i)–(iii) are simultaneously satisfied for ζ+ (z); therefore,
ζ+ (z) is an example of dark solitonlike solution to Eq. (7.45). With the help of this set of
parameters, we display in Figs. 7.5, 7.6, and 7.7, and in Figs. 7.8, 7.9, and 7.10 the temporal
[spatial] evolution of the wave intensity |u (x, t)|2 associated with the solitonlike solution
(7.51) for x = 3 [t = 4] and (0) = 1, and for different numerical values of parameters t ∗ ,
A, and β on the solitary-wave shape. Plots of the left and the right panels are generated for
respectively positive t ∗ (broadening trap) and negative t ∗ (shrinking trap). Plots of the top
panels are obtained for negative A (confining potential), while those of the bottom panels are
generated with positive A (expulsive potential). In particular, we show in Fig. 7.8 the spatial
profile at time t = 4 of the wave density |u(x, t)|2 for three different values of parameter t ∗ .
As we can see from plots of Fig. 7.8, the peak density in the case of confining potential (see
the plots of the top panels) decreases as t ∗ increases for the broadening trap (that is, when
t ∗ > 0), and increases with t ∗ for the shrinking trap (case when t ∗ < 0). In the case of the
expulsive potential (see the plots of the bottom panels), the peak density increases with t ∗
for both the broadening (t ∗ > 0) and shrinking (t ∗ < 0) traps. Figure 7.9 shows the spatial
profile at time t = 4 of the wave density |u(x, t)|2 for three different values of parameter A.
Plots of 7.9 demonstrate that independently of the sign of A (that is, for both the confining
and expulsive potentials), the peak density decreases as A increases, which happens for both
240 7 Dynamics of One-Dimensional Condensates …

Fig. 7.6 The same as in 7.9, but for three different values of parameter A in potential (7.34). The top
and bottom panels correspond to the confining and expulsive potential, respectively, while the left
and right panels are associated, severally, with the broadening trap (for t ∗ = 10) and the shrinking
one (for t ∗ = −10). Typical values of other parameters are given in the text. Reprint from Ref. [29],
Copyright 2022, with permission from American Physical Society

the broadening and shrinking traps. To show the effect of parameter β on the wave density,
we plot in Fig. 7.10, the spatial profile at time t = 4 the wave density |u(x, t)|2 for different
values of β. It is clearly seen from different plots of Fig. 7.10 that, irrespective of the sign of
A (the confining or expulsive potential) and the sign of t ∗ (the broadening or shrinking trap),
the density peak decreases with the increase in the numerical values of the imprint parameter
β.

7.1.4 Conclusion

We have investigated in this Section the MI of continuous wave states in the context
of the inhomogeneous cubic derivative NLS equations with external time-varying
harmonic potentials in the context of BECs. To make this investigation possible
for both attractive and repulsive nonlinearities, we have combined the phase-imprint
technique with a modified lens transformation. Such a combination has allowed us to
cast the problem in the form in which the cubic derivative NLS equation has constant
coefficients. The special case when the strength of the magnetic trap modulated in
time ∼ (t + t ∗ )−2 is investigated in details. The effects of both the PIT imprint
parameter and trap parameter t ∗ on the MI gain are investigated. In the case of
the constant MI gain, we have presented the analytical exact matter-wave soliton
solutions of the inhomogeneous NLS equation under the consideration. Also, the
effects of various parameters on the shape of matter-wave solitons are considered.
7.2 Soliton Management (SM) in One-Dimensional Bose-Einstein … 241

Fig. 7.7 The same as in Figs. 7.9 and 7.10, but for three different values of the imprint parameter
β in Eq. (7.7). The top and bottom panels correspond to the confining potential (with A = −2)
and expulsive one (with A = 1/17), respectively, while the left and right panels are associated,
severally, with the broadening trap (for t ∗ = 10) and the shrinking one (for t ∗ = −10). Typical
values of other parameters are given in the text. Reprint from Ref. [29], Copyright 2022, with
permission from American Physical Society

7.2 Soliton Management (SM) in One-Dimensional


Bose-Einstein Condensates with Two- and Three-Body
Inter-atomic Interactions

We consider in this Section the generalized cubic-quintic (GCQ) Gross-Pitaevskii


equation with external complex potential that describes the evolution of the wave-
function for Bose-Einstein condensates with two- and three-body inter-atomic inter-
actions in a spatiotemporal-varying potential consisting of parabolic, linear, and
complex terms. The complex term of the external potential relates to the feeding or
the loss of atoms by the condensates. The ansatz method is applied for investigating
the MI and for finding various types of solitonlike solutions including gray, kink, and
bright solitonlike solution of the GCQ-GP equation under certain parametric condi-
tions. We show that the complex potential seriously modifies the instability/stability
domain, while the linear part of the external potential has not effect on the stability
of the system. Using the found exact analytical solitonlike solutions of the GCQ-GP
equation, we investigate analytically the solitons control system; particularly, we
show that the soliton control system may relax the limitations to parametric condi-
tions. We then show how both the external harmonic and linear trapping potentials
can be used to manage the motion of matter-wave solitons in the systems. We show
that the amplitude of the matter-wave solitons keep no change in propagating in
242 7 Dynamics of One-Dimensional Condensates …

Fig. 7.8 Density |u(x, t)|2 at t = 4, as given by solution (7.51), for three different values of
parameter t ∗ appearing in potential (7.34). The top and bottom panels correspond, severally, to the
confining potential (with A = −2) and the expulsive one (with A = 1/17), while the left and right
panels are associated with the broadening and shrinking trap, respectively. Typical values of other
parameters are given in the text. Reprint from Ref. [29], Copyright 2022, with permission from
American Physical Society

the system; we establish that the total number of the condensate atoms decreases
(increases) when the condensate losses (gains) atoms. We investigate the effects of
the three-body inter-atomic interaction on the soliton motion and show that they are
responsible of the soliton compression.

7.2.1 Introduction

The soliton management/control plays an important role in the application of soli-


tons. During last decades, it has been extensively investigated theoretically [37–40]
and experimentally [41] for nonlinear phenomena described by a class of evolution
nonlinear partial differential equations (NPDEs) such as generalized NLS equations
with distributed coefficients. Such NPDEs of Schrödinger-type have been widely
investigated and find their applications in various branches of nonlinear science such
as nonlinear transmission networks, quantum physics, optical fiber, and fluid system
[42–47].
The present Section deals with the investigation the MI phenomenon and the soli-
ton management in BEC systems with spatiotemporal-dependent parabolic back-
ground when both the two- and three-body inter-atomic interactions are taken into
7.2 Soliton Management (SM) in One-Dimensional Bose-Einstein … 243

Fig. 7.9 The same as in 7.13, but for three different values of parameter A in potential (7.34).
The top and bottom panels correspond to the confining and expulsive potentials, respectively, while
the left and right panels are associated, severally, with the broadening trap (for t ∗ = 10) and the
shrinking one (for t ∗ = −10), respectively. Values of other parameters are given in the text. Reprint
from Ref. [29], Copyright 2022, with permission from American Physical Society

account. Such BEC systems can be described by the quasi-one-dimensional GP


equation [48]

∂ψ 1 ∂ 2ψ
i =− + g |ψ|2 ψ +  |ψ|4 ψ + V (x, t)ψ, (7.54)
∂t 2 ∂x2

where V (x, t) is the external potential that consists of an harmonic part, a linear part,
and a time-dependent complex part, defined as [43]

V (x, t) = k(t)x 2 + λ(t)x + iγ (t). (7.55)

In Eqs. (7.54) and (7.55), the t and x are respectively the temporal √ and spatial
coordinates,
√ measured in units of 1/ω ⊥ and a ⊥ , respectively. a ⊥ = /(mω⊥ ) and
a0 = /(mω0 ) denote the linear oscillator lengths in the transverse and cigar-
axis directions, respectively, where ω⊥ and ω0 are the corresponding harmonic-
oscillator frequencies and m is the atomic mass. Parameters g of the cubic nonlinearity
2
represents the two-body inter-atomic interactions, while parameter  = 3πmg2 1 N2 a 4 of

the quintic nonlinearity represents the three-body inter-atomic interactions; here, as
is the s-wave scattering length, a B is the Bohr radius, g1 is the effective three-body
contact interaction related to the GP equation, and N represents the total number of
atoms in the condensate. Throughout this Section, we consider g and  to be linked
244 7 Dynamics of One-Dimensional Condensates …

Fig. 7.10 The same as in Figs. 7.13 and 7.12, but for three different values of the imprint parameter
β in transformation (7.7). The upper and lower panels correspond to the confining potential (with
A = −2) and the expulsive one (with A = 1/17), respectively, while the left and right panels are
associated, severally, with yjr broadening trap (for t ∗ = 10) and the shrinking one (for t ∗ = −10).
Typical values of other parameters are given in the text. Reprint from Ref. [29], Copyright 2022,
with permission from American Physical Society

by the relationship || ≤ 01 % |g|, with 0 < 01 < 100 [49]. This means that the
parameter of the three-body inter-atomic interactions is small in comparison with
the parameter of the two-body inter-atomic interactions. Moreover, we assume that
both g and  are time-varying parameters [50, 51].
During our study, we impose√ the following restriction on the harmonic part of the
external potential (7.55) [52] |k(t)| = ω0 /ω⊥  1 so that k(t) can take any sign,
positive for confining potential, and negative for repulsive potential. It is important
to note that the dissipative parameter γ stands for the feeding of condensate atoms
if γ > 0 and loss of atoms in the condensate if γ < 0 [53, 54].

7.2.2 Modulational Instability in Bose-Einstein Condensates


Trapped in a Spatiotemporal-Dependent Dissipative
Potential

To investigate the dynamics of BECs whose wave-functions are governed by the


cubic-quintic GP equation (7.54)–(7.55), we employ the similarity transformation
technique [43, 55–58]. For this aim, we introduce the following suitable ansatz [59]
7.2 Soliton Management (SM) in One-Dimensional Bose-Einstein … 245
  
ψ(x, t) = (t)(X, T ) exp i (χ + λ0 T ) + γ (t)dt , (7.56)

1 dα 2 1 dβ
X (x, t) = α(t)x + β(t), χ (x, t) = − x − x + χ0 (t), (7.57)
2α dt α dt

where λ0 is a real constant, and T (t), (t), α(t), β(t), and χ0 (t) are functional
parameters of time t, solutions of the following ordinary differential system

dT
= α2 , (7.58)
dt
d 1 1 dα
− = 0, (7.59)
dt 2 α dt
   
d 1 dα 1 dα 2
− − 2k = 0, (7.60)
dt α dt α dt
 
d 1 dβ 1 dα dβ
− 2 −λ = 0, (7.61)
dt α dt α dt dt
 
dχ0 1 1 dβ 2
+ + λ0 α 2 = 0. (7.62)
dt 2 α 2 dt

Equation (7.59) admits the following special/particular solution that gives (t) in
terms of α(t) 
(t) = |α(t)|. (7.63)

Ansatz (7.56) under restrictions (7.57)–(7.62) maps the GP equation (7.54)–(7.55)


into the following cubic-quintic GP equation containing only two parameters

∂ 1 ∂ 2 
i + − g0 ||2  − 0 ||4  = 0, (7.64)
∂T 2 ∂
  X 2
   
g
g0 = exp 2 γ (t)dt , 0 =  exp 4 γ (t)dt . (7.65)
|α|

In our studies, we focus our attention to the situation when parameters g0 and 0
are real constants. It then follows from Eqs. (7.60) and (7.65) that the functional
parameters k, g, γ , and  must satisfy the following relationships
 2     
d2g dg 2 + 2g 2 dγ − 2 γ dg − 2γ 2 = 0, (t) =  exp −4
g − 2 − 2kg 0 γ dt .
dt 2 dt dt g dt
(7.66)

It follows from Eq. (7.65) that α(t) is defined as


  
g(t)
|α(t)| = exp 2 γ (t)dt , g0 g(t) > 0. (7.67)
g0
246 7 Dynamics of One-Dimensional Condensates …

Equation (7.64) admits the continuous wave solution


 
d 0 (t) 1 2
(X, T ) = φ0 exp [i (k0 X − 0 (t))] , = α2 k + g0 φ0 + 0 φ0 ,
2 4
dt 2 0
(7.68)
where φ0 and k0 stand for respectively the carrier amplitude and the carrier wavenum-
ber. Now, we perturb the CW solution (7.68) as follows

(X, T ) = [φ0 + φ] exp [i (k0 X − 0 (t))] , (7.69)

and impose that


⎡ ⎛ ⎞⎤ ⎡ ⎛ ⎞⎤
T T
⎢ ⎜ ⎟⎥ ⎢ ⎜ ⎟⎥
φ(X, T ) = b1 exp ⎣i ⎝ K X + (τ )dτ ⎠⎦ + b2∗ exp ⎣−i ⎝ K X + ∗ (τ )dτ ⎠⎦ , (7.70)
0 0

where b1 and b2 are two complex constants, and K and are the wavenumber and
frequency of the perturbation, respectively. Using the linear stability analysis, we
arrive to the following dispersion relation of the perturbation

K2 2  
( (T ) + k0 K )2 = K + 4φ02 20 φ02 + g0 . (7.71)
4
Therefore, for the cubic-quintic GP equation (7.64) to be unstable under modulation,
the wavenumber K of the perturbation must satisfy the condition
 
0 2
K < 2
−4g0 φ02 1 + 2 φ0 ,
g0

which is possible only for negative g0 (focusing cubic nonlinearity). The following
local growth rate (gain) of modulational instability is then obtained
$
K2  
|Im ( )| = |K | − − 20 φ02 + g0 φ02 , (7.72)
4
g0 and 0 being two real constants given by Eq. (7.65). It is important to note that the
local growth rate of MI (7.72) will be time-independent only when the relationships
(7.66) are satisfied under the condition g0 < 0.
Let us consider the situation when the relationships (7.66) are violated. In such a
situation, the local growth rate of MI (7.72) will depend on time. For simplicity, let
us consider the important case when k is constant, which is generally used in current
experiments [60]. Integrating Eqs. (7.60) and (7.58) when k is constant, we find that
α(t) and T (t) are given by

1 1 √ 
α(t) =  √  , T (t) = √ tan 2kt (7.73)
  2k
cos 2kt 
7.2 Soliton Management (SM) in One-Dimensional Bose-Einstein … 247

(2n+1)π
and have singularities at any t = tn = √ for nonnegative inter n. As the domain
2 2k  
of definition of both α(t) and T (t), we adopt t ∈ 0, t0 so that T (t) will vary from
0 to +∞. Therefore, the local growth rate of MI takes the following explicit form

|Im ( [T (t)])|
%       
|K |  √ 
 
= −K 2 − 4φ02 g(t) cos 2kt  + 2φ02 (t) exp 2 γ (t)dt exp 2 γ (t)dt ;
2
(7.74)

here, the constant wavenumber K of the modulation satisfies the restriction


       
 √  (t)
K 2 < −4φ02 g(t) cos 2kt  + 2φ02 exp 2 γ (t)dt exp 2 γ (t)dt .
g(t)
(7.75)
In what follows, we focus ourselves to the case where (t) = 01 % × g(t), with
0 < 01 % < 1 [49]. It is obvious that the strength of magnetic field k, the dissi-
pation coefficient γ , and the three-body inter-atomic interaction may affect the MI
phenomenon, as we can clearly see from Fig. 7.11 in which we have displayed the
instability growth rate given by Eq. (7.74) at time t = t0 for g(t) = exp [kt] and con-
stant loss/gain parameter γ . Fig. 7.11a, b show the MI gain for different numerical
values of the strength k. In Fig. 7.11c, d, we show the evolution of the MI growth
rate for different numerical values of parameter 01 of the three-body inter-atomic
interactions. The effects of the gain/loss parameter γ can be observed in Fig. 7.11e,
f where we depict the MI gain with three values of parameter γ . As we can see from
different plots of 7.11,
(i) when the condensate gains atoms, the harmonic trap softens the instability, as
we can well see from 7.11a generated for γ = 0.05;
(ii) when the condensate losses atoms, the harmonic trap relatively enhances the
instability, as we can see from 7.11b obtained for γ = −0.05. These two results
mean that for a BEC system with gain of atoms (γ > 0), the instability may be
suppressed by excess of the strength of the magnetic field.
(iii) the three-body interaction enhances the instability, independently on the sign of
parameter γ , although the three-body interaction strongly (weakly) enhances
the instability when the condensate gains (losses) atoms, as it is clearly seen
from Fig. 7.11c, d.
(iv) as it is well seen from Fig. 7.11e, f , the local MI gain increases when the
dissipation parameter γ increases.
(v) k and positive γ as well as k and 01 have opposite effects, while the positive
γ and 01 have the same effects on the MI.
248 7 Dynamics of One-Dimensional Condensates …

Fig. 7.11 Instability growth rate according to Eq. (7.74) at the limit t = tn=0 for φ0 = 1 with
g(t) = exp [kt] and constant dissipation parameter γ . a, b Effects of the strength of magnetic field
k on the MI for 01 = 10 and γ = 0.05 (a), γ = −0.05 (b), and three values of the strength of
magnetic field k = 0.001 (solid line), 0, 002 (dashed line, and 0.004 (dotted line); c, d effects of
parameter 10 of the three-body interaction on the MI for γ = 0.05 (c), γ = −0.05 (d), k = 0.001,
and three values of 10 = 10 (solid line), 10 = 15 (dashed line), and 20 (dotted line); e, f Effects
of the dissipation coefficient γ on the MI with k = 0.001, 10 = 10, and three values of γ : e case
of feeding of atoms, γ = 0 (solid line), γ = 0, 4 (dashed line) and γ = 0.8 (dotted line), f case of
feeding of atoms, γ = −0.09 (solid line), γ = −0.05 (dashed line), and γ = −0.01 (dotted line).
Reprint from Ref. [61], Copyright 2022, with permission from IOPScience
7.2 Soliton Management (SM) in One-Dimensional Bose-Einstein … 249

7.2.3 Soliton Management in Bose-Einstein Condensate


Systems with Two- and Three-Body Interactions
Trapped in a Spatiotemporal-Dependent Dissipative
Potential

In this subsection, we use the cubic-quintic GP equation (7.54)–(7.55) to investigate


analytically the formation of matter wave solitons in BEC systems when the loss/gain
of atoms is taken into account. We start by finding exact soliton solutions of wave
equation. We focus our attention to the situation when parameters g(t) and (t) of
respectively the two- and three-body interactions are related by |(t)| = 01 % ×
|g(t)| , (with 0 < 01 < 100 [49]). We also assume that the “integrable conditions”
(7.66) are satisfied. It then follows from Eqs. (7.65)–(7.67) that


+ 2γ 2 + k = 0, (7.76)
    dt      
75β0  α0  75  α0 
g(t) = exp −4 γ (t)dt , |α(t)| = exp −2 γ (t)dt ,
201  β0  201  β0 
(7.77)

where α0 = 83 0 and β0 = g0 . We can then rewrite Eq. (7.64) as

∂ 1 ∂ 2  3α0
i + − β0 ||2  − ||4  = 0. (7.78)
∂T 2 ∂X 2 8
Equation (7.78) contains two arbitrary constants, α0 and β0 .
Gray and Bright Soliton Solutions of the Cubic-Quintic GP Equation Assuming
conditions (7.66) to be satisfied, the traveling wave solutions of Eq. (7.78) can be
sought in the form

(X, T ) = ρ(ξ ) exp [i (υ X − ωT )] , (7.79)

where ξ = X − υT is the propagating coordinate, υ and ω being two constants and


nonnegative function ρ being any solution of the EODE
 2

= α0 ρ 4 + 4β0 ρ 3 + 6γ0 ρ 2 + 4δ0 ρ ≡ R(ρ). (7.80)

Here, δ0 ∈ R and γ0 = ( 3 ) . Special interesting solutions of Eq. (7.80) are given


2 υ 2 −2ω

by [62] 
6 d R 
ρ(ξ ) = ρ0 +  , (7.81)
2 
24℘ (ξ ; g2 , g3 ) − ddρR2  dρ ρ=ρ0
ρ=ρ0
250 7 Dynamics of One-Dimensional Condensates …

where ℘ (ξ ; g2 , g3 ) is the Weierstrass elliptic function [62], the primes denote dif-
ferentiation with respect to ρ, ρ0 is any simple root of polynomial R(ρ), and

g2 = 3γ02 − 4β0 δ0 , g3 = δ0 (2β0 γ0 − α0 δ0 ) − γ03 .

Equation (7.81) leads to solitonlike solutions of Eq. (7.80) if and only, if [63]

g2 ≥ 0, g3 ≤ 0, and = g23 − 27g32 = 0. (7.82)

We here notice that the general solution of the EODE (7.80) is given by Schürmann
[63]
√  
d℘ (ξ ;g2 ,g3 )
R(ρ0 ) dξ + 21 R (ρ0 ) ℘ (ξ ; g2 , g3 ) − 24
1 R (ρ ) + 1 R(ρ )R (ρ )
0 24 0 0
ρ(ξ ) = ρ0 +  2 ,
2 ℘ (ξ ; g2 , g3 ) − 24 1 R (ρ ) − 1 R(ρ )R (ρ )
0 48 0 0
(7.83)
where ρ0 is not necessary a zero of polynomial R(ρ), and the prime stands for
d/dρ. Under conditions (7.82), the following solitonlike solution is obtained from
Eq. (7.81) by taking
& '
3 1√
℘ (ξ ; g2 , g3 ) = e1 1+ √ , e1 = 3 −g3 : (7.84)
sinh 2
3e1 ξ 2
 
6R (ρ0 ) 72e1 1
ρ(ξ ) = ρ0 + 1− 48e1 +R (ρ0 )
√ ,
24e1 − R (ρ0 ) 24e1 − R (ρ0 ) 24e1 −R (ρ0 )
+ cosh2 3e1 ξ
(7.85)
where parameter ρ0 is any simple real zero of polynomial R(ρ).
Solving the equation  = (δ0 ) = 0 for real δ0 yield
     
2β0 27α0 γ0 − 16β02 ± 8β02 − 9α0 γ0 2 8β02 − 9α0 γ0
δ0 = , (7.86)
27α02

if 8β02 − 9α0 γ0 ≥ 0. For any δ0 = 0 given by Eq. (7.86) and satisfying the condition

γ03 + (α0 δ0 − 2β0 γ0 ) δ0 > 0,

ρ = ρ0 will be a simple zero of R(ρ) so that by letting ρ0 = 0 in Eq. (7.85) yields


the following solitonlike solution of Eq. (7.80)
 
2δ0 6e1 1
ρ(ξ ) = 1− 4e1 +γ0 √ , (7.87)
2e1 − γ0 2e1 − γ0 2e1 −γ0
+ cosh2 3e1 ξ

where
7.2 Soliton Management (SM) in One-Dimensional Bose-Einstein … 251

1 3 3
e1 = γ + (α0 δ0 − 2β0 γ0 ) δ0 .
2 0

Next, we choose υ and ω from the conditions g2 = 3γ02 − 4β0 δ0 > 0 and e1 > 0;
for such υ and ω, the above solution ρ(ξ ) is nonnegative and contains the gray
solitonlike solution
⎡ ⎤
8β0 ⎢
⎢1 − 3 ⎥
ρ(ξ ) = −  ⎥
⎦ , α0 > 0 and β0 < 0, (7.88)
3α0 ⎣ 2β 2
2 + cosh2 2 3α00 ξ

if we take γ0 = 0.
Inserting Eq. (7.87) into (7.79), we arrive to the following solitonlike solution of
the CQ-NLS equation (7.78)

(X, T )
( ⎡ ⎤
)
)
) 2δ0 ⎣ 6e1 1
=* 1− ⎦ exp [i (υ X − ωT )] . (7.89)
2e1 − γ0 2e1 − γ0 4e1 +γ0 + cosh2 √3e1 (X − υT )
2e −γ 1 0

Combining now ansatz (7.56) with Eq. (7.89) leads to the following solitonlike
solution of the GP equation (7.54)–(7.55):
(  
) 
) 3α0  δ 6e1 1
ψ(x, t) = 5*  0
 1− 4e1 +γ0 √
β  0 01 2e1 − γ0 2e1 − γ0 2e1 −γ0
+ cosh 2
3e1 (X − υT )
× exp [i (χ + υ X + (λ0 − ω) T )]| X =X (x,t), and χ=χ(x,t) ; (7.90)

here, λ0 , α0 and β0 < 0 are arbitrary real constants,


 2    
75α0
T = T (t) = T0 + exp −4 γ (t)dt dt, (7.91)
2β0 01

T0 being chosen from condition T (0) = 0, and γ (t) verifies the Eq. (7.76).
It is important to note that another interesting solitonlike solution of Eq. (7.80)
can be obtained from Eq. (7.85) by taking δ0 = 0. In this case, we obtain, for the
bright solitonlike solution

⎡ ⎤
6R (ρ0 ) ⎢
⎢1 − 36γ0 1 ⎥
ρ(ξ ) = ρ0 +  ⎥
⎦, (7.92)
12γ0 − R (ρ0 ) ⎣ 12γ0 − R (ρ0 ) 24γ0 +R (ρ0 ) 3γ ξ

12γ −R (ρ )
+ cosh 2
2 0
0 0

if
252 7 Dynamics of One-Dimensional Condensates …


2 2    −β0 ± β02 − 23 γ0 α0
γ0 = υ − 2ω > 0, β02 − υ 2 − 2ω α0 > 0, ρ0 = ,
3 2α0

where α0 , and β0 < 0 are real constants to be taken so that ρ(ξ ) should be nonneg-
ative.
Another interesting bright soliton solution of Eq. (7.92) can be found by the direct
method:
A
ρ(ξ ) = .
B + cosh2 [μξ ]

The result is
6γ0
ρ(ξ ) = ±     . (7.93)
− 4β02 − 6α0 γ0 ∓ 2β0 + 2 4β02 − 6α0 γ0 cosh2 3
2
γ 0 ξ

Various parameters from solution (7.93) must be choosing from its non-negativity.
For example, the positive solution

β0 6
ρ(ξ ) =   , α0 < 0, β0 < 0 (7.94)
α0 β02
−1 + 4 cosh2
−3 α0 ξ

β2
is obtained from Eq. (7.93) by letting γ0 = −2 α00 .
In the special situation when δ0 = 2β02 − 3γ0 α0 = 0, we find ρ0 = −2β0 /α0 to be
a double root of polynomial R(ρ). The corresponding nonnegative kink solitonlike
solution is obtained from Eq. (7.83) by setting ρ0 = −β0 /α0 :
⎛ ⎡% ⎤⎞
β0 ⎝ β 2
ρ(ξ ) = ρ± (ξ ) = − 1 ± tanh ⎣ 0 ⎦⎠
ξ , α0 > 0, β0 < 0. (7.95)
α0 α0

Dynamics of Matter Wave Solitons in BECs with Two- and Three-Body Inter-
actions With the use of the above found solitonlike solutions, we investigate here
the dynamics of matter-wave solitons in BECs with loss/gain of atoms when both
the two- and three-body inter-atomic interactions are taken into account. Param-
eters g and  are assumed to be linked by the relationship |(t)| = 01 % |g(t)|
(with 0 < 01 < 100), which means that the two-body inter-atomic interactions are
stronger than the three-body interactions [49]. Here, we focus our attention on the
situation when g(t) and γ (t) satisfy Eq. (7.77), while γ (t) and k(t) satisfy Eq. (7.76).
In the following, we focus our attention to the following three special cases:

(i) The case of constant k(t) [64]. Following the results of the experiment carried
out in work [64], we set k(t) = −2κ 2 (κ  0.05). Equations (7.76) and (7.77)
then give
7.2 Soliton Management (SM) in One-Dimensional Bose-Einstein … 253

  %  
75β0  α0   α0 
γ = ±κ, g(t) =   exp [∓4κt] , α(t) = ± 75   exp [∓2κt].
201 β  201 β 
0 0

(ii) Case of a finite blow-up time-varying k(t) [65] The case of a finite blow-up time-
varying k(t) corresponds to the situation when strength k(t) of the harmonic
(1−2k0 )
potential is of form k(t) = k0(t+t [60]. A similar case has been used when
0)
2

investigating the MI of a quasi-one-dimensional GP equations. For this form of


k(t), it follows from Eqs. (7.76) and (7.77) that
 
k0 75β0  α0  1
γ (t) = , g(t) =   , and
t + t0 201 β0 (t + t0 )4k0
%  
75  α0  1
α(t) = ±   .
2 β (t + t )2k0
01 0 0

Here, the blow-up time t = t0 is an arbitrary constant to be taken from the


condition t0 γ (t) > 0.
(iii) The case of the temporal periodic s-wave scattering length [66, 67] In this special
case, parameter g(t) of the two-body inter-atomic interactions is taken as

g = g01 (1 + m sin [ϑt]) , 0 < m < 1, g01 = 0.

Equations (7.76) and (7.77) then lead to


%  
ϑm cos [ϑt] 75  α0  
γ (t) = − , α(t) = ±   1 + m sin [ϑt], and
4 (1 + m sin [ϑt]) 201 β 
0

mϑ 2 m + 2m cos2 [ϑt] + sin [ϑt]


k=− .
4 (1 + m sin [ϑt])2

Before turning to the analysis of the soliton formation and its dynamics, we first
analyze the above found solitonlike solutions of Eq. (7.54). Under ansatz (7.56),
solitonlike solution ψ(x, t) of the GP equation under consideration reads

%  
3  α0 
ψ(x, t) = 5   ρ(ξ(x, t)) exp [i {χ + υ X + (λ0 − ω) T }] , (7.96)
201 β 
0
ξ(x, t) = α(t)x + β(t) − υT (t), T (t)
     
75α0 2
= T0 + exp −2 γ (t)dt dt (7.97)
201 β0
1 dα 2 1 dβ
X = α(t)x + β(t), χ = − x − x + χ0 (t), (7.98)
2α dt α dt
254 7 Dynamics of One-Dimensional Condensates …
   
λ(t)
β(t) = β01 + α 2 (t) C0 + dt dt, (7.99)
α(t)

where C0 and β01 are constants of integration, λ0 , α0 , β0 , υ, and ω are free real
constants to be chosen properly, T0 is a constant to be taken so that T (0) = 0, and
ρ(ξ(x, t) is any of found bright solitonlike solutions (7.87) and (7.92) of Eq. (7.80).
The soliton phase is found to be

ϕ(x, t) = χ (x, t) + υ X (x, t) + (λ0 − ω) T (t).

As we can see from Eq. (7.96),


(i) the soliton amplitude is inversely proportional to 01 so that any increasing in
the value of (t) will decrease the soliton amplitude;
(ii) the centre of the pulse is obtained from Eq. (7.97) as

υT (t) − β(t)
xc (t) = ,
α(t)

and satisfies the equation


··
x c + 2kxc (t) + λ = 0, (7.100)

meaning that the soliton centre behaves like a classical particle; therefore, param-
eter k and λ can be used to to manipulate the motion of matter-wave solitons in
BEC system under consideration.
(iii) the soliton width is proportional to
    
75  α0 
α(t) =   exp −2 γ (t)dt
201 β 
0

(this fact can be clearly seen from Eq. (7.97)) and therefore, we can employ
Eq. (7.96) to describe the compression of matter wave solitons in BECs under
consideration.
(iv) the total number of BEC atoms is given by

+∞    +∞
N = N (t) = |ψ(x, t)| d x = exp 2 γ (t)dt
2
ρ(ξ )dξ ;
−∞ −∞

hence, N (t) decreases (increases) for negative γ (t) (positive γ (t)).


In what follows, we intend to study the dynamics of BEC matter wave solitons for
each of the above enumerated cases. For this purpose we evaluate the exact BEC
density as
7.2 Soliton Management (SM) in One-Dimensional Bose-Einstein … 255
 
 α0 
|ψ(x, t)|2 =  β0  ρ(ξ(x, t)),
75
201   2  
 (7.101)
ξ(x, t) = α(t)x + β(t) − υ T0 + 201 β0
75α0
exp −2 γ (t)dt dt ,

where β(t) is any real solution of the differential equation (7.61). In our studies, we
assume parameter λ(t) of the linear potential to be

λ(t) = σ + η cos [θ t] . (7.102)

Expression (7.102) of parameter λ(t) means that the linear potential is induced by a
bias magnetic/laser field with strength σ /η and with the modulation frequency θ .
Application to BEC systems with a time-independent harmonic and linear
potentials As the example of BEC system with a time-independent harmonic and
linear potentials, we consider k(t) = −2κ 2 (κ = 0.05), γ (t) = ±κ [53, 60, 68],
leading to the following parameter g(t) of the two-body inter-atomic interactions
 
75β0  α0 
g(t) =   exp [∓4κt] .
201 β 
0

Therefore,
   
75  α0  75α0 2
α(t) =   exp [∓2κt] , T (t) = 1 (±1 ∓ exp [∓2κt]) ,
201 β  2κ 201 β0
0
(7.103)
   2  2  
75  α0  σ  α0  θ ± 4κ cos [θ t]
2
β(t) = − ± 2 + 2η    2 exp [∓2κt] .
401  β0  2κ β0 4κ 2 + θ 2
(7.104)

In this special case, the soliton width is found from Eq. (7.103) and reads
 
201  β0 
wsol = wsol (t) = w0   exp [±2κt] .
75 α 
0

For BEC with loss of atoms (γ (t) = −κ), wsol (t) → 0 as t → +∞, meaning that in
the case of BEC with loss of atoms, pulse width can be compressed quite effectively.

With the help of the solitonlike solutions (7.87) and (7.94), we show in Figs. 7.13
and 7.12 the evolution of respectively the gray and bright solitons (the evolution of
the bright soliton is obtained when δ0 is taken with sign “ +”). Plots of the left panels
correspond to BECs with gain of atoms, while those of the right panels are associated
with BECs with loss of atoms. It follows from Figs. 7.13 and 7.12 that both the gray
and the bright solitons move along a parabolic trajectory in the +x direction (the
−x direction) for BECs with gain (loss) of atoms. Plots (c) and (d) of 7.13 and Plots
256 7 Dynamics of One-Dimensional Condensates …

Fig. 7.12 a, b Evolution plot of the bright solitons obtained with the use of the soliton-like solution
(7.94) when γ0 = 2/3, α0 = −0.3, β0 = −0.25, υ = 0.25, 01 = 35, σ = 0.25, η = 3, and θ =
π/3. c–f Density plots of bright solitons associated with the soliton-like solution (7.94) with γ0 =
2/3, α0 = −0.3, β0 = −0.25, υ = 0.25, 01 = 35. Plots c, d are obtained when the effects of both
the bias magnetic and laser fields are neglected, that is, σ = η = 0, while plots e, f are obtained under
the combined effects of the bias magnetic and laser fields with σ = 0.25, η = 3, and θ = π/3. Plots
of the left panels correspond to BEC with feeding of atoms with the feeding parameter γ (t) = κ,
while the plots of the right panels are associated with BEC with loss of atoms with the loss parameter
γ (t) = −κ. Reprint from Ref. [61], Copyright 2022, with permission from IOPScience

(a), (b), (e), and (f) of Fig. 7.12 show respectively gray and bright solitons propa-
gating along oscillating trajectories; this behavior of the wave trajectory is due to
the temporal periodic modulation of the laser field. For the BECs with gain (loss)
of atoms, the soliton width increases (decreases) during its propagation, as it is well
seen from plots of the left and right panels, respectively. Different plots of Figs. 7.13
and 7.12 reveal that the soliton amplitude is not affected by the bias magnetic/laser
field; meanwhile the wave speed is affected by the bias magnetic/laser field.
Application to BECs with a Temporal Periodic Modulation of the s-Wave Scat-
tering Length
7.2 Soliton Management (SM) in One-Dimensional Bose-Einstein … 257

Fig. 7.13 Evolution plot of the gray solitons obtained with the use of the soliton-like solution
(7.87) when γ0 = −0.04, α0 = 0.5, β0 = −0.15, υ = 0.5, 01 = 15. Plots a, b are obtained when
the effects of both the bias magnetic and laser fields are neglected, that is, σ = η = 0, while plots
c, d are obtained under the combined effects of the bias magnetic and laser fields with σ = 0.25,
η = 0.9, and θ = π/3. Plots of the left panels are obtained with γ (t) = κ corresponding to BEC with
feeding of atoms, while the plots of the right panels are obtained with γ (t) = −κ corresponding to
BEC with loss of atoms. Reprint from Ref. [61], Copyright 2022, with permission from IOPScience
258 7 Dynamics of One-Dimensional Condensates …

For the second example, we follow Hao et al. [69] and consider a BEC system for
which parameter g(t) of the two-body inter-atomic interactions is of the form

g(t) = g01 (1 + m sin [ϑt]) exp [κt] ,

where 0 < m < 1, κ, ϑ, and g01 = 0 are free parameters. It then follows from Eqs.
(7.77) and (7.76) that
     
1 ϑm cos [ϑt] 75  α0   1
γ (t) = − κ+ , α(t) = 1 + m sin [ϑt] exp κt ,
4 1 + m sin [ϑt] 201  β0  2
 
κ 2ϑm cos [ϑt] ϑ 2 m 2 (m + sin [ϑt]) + m cos2 [ϑt]
k(t) = − κ+ − .
8 1 + m sin [ϑt] 8 (1 + m sin [ϑt])2

It is evident that for small values of parameter κ, the dissipative parameter γ (t) will
periodically change its sign; this corresponds to a BEC system which periodically
gains and losses atoms. Computing the soliton width in the present case gives
   
201  β0  1 1
wsol = wsol (t) = w0  √ exp − κt .
75  α  1 + m sin [ϑt] 2
0

Hence, wsol (t) → 0 as t → +∞ for positive κ. Next, it follows from Eqs. (7.97) and
(7.99) that

T (t)
⎡ ⎤
 2 mκϑ − κ − ϑ
2 2

1 75  α0  ⎢⎢ κ2 + ϑ2 ⎥
⎥,
=   ⎣ ⎦
κ 201 β0 κ + ϑ + mκ (κ sin [ϑt] − ϑ cos [ϑt])
2 2
+ exp [κt]
κ2 + ϑ2
  2 2
C0 75  α0  κ + ϑ + mκ (κ sin [ϑt] − ϑ cos [ϑt])
2
β(t) =β01 + exp [κt] ,
κ 201  β0  κ2 + ϑ2

where β01 and C0 are free constants.


Using the above data and the solitonlike solutions (7.88) and (7.94), we depict
respectively on the top panels and in the bottom panels of Fig. 7.14 the evolution
of the gray and the bright solitonlike waves, respectively. Plots of the left (right)
panels are generated with κ < 0 (κ > 0). As we can see from plots of Fig. 7.14, the
matter-wave solitons propagate along oscillating trajectories, while their profile keep
no change in propagation in the BEC system under consideration. Also, we can see
from plots of Fig. 7.14 that both the gray and the bright solitons have an increase (a
compression) in their width when κ < 0 (κ > 0).
In Fig. 7.15, we show the evolution of the gray and bright-matter wave solitons
generated with the above data and with respectively the solitonlike solution (7.88) and
(7.94) for different values of parameters α0 and β0 . From Fig. 7.15a–c showing the
7.2 Soliton Management (SM) in One-Dimensional Bose-Einstein … 259

Fig. 7.14 The dynamics of gray a, b and bright c, d matter wave solitons associated with respectively
the gray soliton-like solution (7.88) and the bright soliton-like solution (7.94) when 01 = 15,
m = 0.5, C0 = 0.1, ϑ = π/2. Plots of the left panels show the dynamics of matter wave solitons in
BECs with a decaying temporal periodic modulation of the s-wave scattering length with κ = −0.05,
while plots of the right panels depict the dynamics of matter wave solitons in BECs with a growing
temporal periodic modulation of the s-wave scattering length with κ = 0.05. Other parameters a,
b α0 = 0.5, β0 = −0.15, υ = 0.5, γ0 = 0, β01 = 150; c, d α0 = −0.3, β0 = −0.25, υ = 0.25,
β2
γ0 = −2 α00 , β01 = 10. Reprint from Ref. [61], Copyright 2022, with permission from IOPScience

density plots of the gray soliton for different positive values of parameters α0 , one can
see that the wave velocity increases with α0 , while the soliton width and the soliton
intensity decreases with the increasing in the values of α0 . Figure 7.15d–f which show
the density plots of the bright soliton for different negative values of parameter β0
reveal that the soliton intensity and soliton width increase with |β0 |, while the soliton
velocity decrease with the increasing in the values of |β0 |. Figure 7.15 also reveals
that parameters α0 and β0 do not affect the oscillations of the soliton trajectories.
Application to Bose–Einstein Condensate Trapped in Modulated Time-
Depending Linear Potentials
In order to apply our findings on BECs with loss/gain of atoms trapped in modulated
time-depending linear potentials when the two- and tree-body inter-atomic interac-
260 7 Dynamics of One-Dimensional Condensates …

Fig. 7.15 The density plots of (top panels) the gray matter wave solitons associated with solution
(7.88) for different values of parameter α0 and (bottom panels) the bright matter wave solitons
associated with solution (7.94) for three values of parameter β0 for 01 = 15, m = 0.5, C0 = 0.1,
and ϑ = π/2. Other parameters are (Top panels): γ0 = 0, β0 = −0.15, υ = 0.3, β01 = −100,
κ = −0.05, and α0 = 1.1 (a), 1.55 (b), 2 (c). (Bottom panels): γ0 = −2β02 /α0 , α0 = −0.3, υ =
0.25, β01 = −500, κ0.05, and β0 = −0.058 (d), −0.054 (e), −0.05 (f). Reprint from Ref. [61],
Copyright 2022, with permission from IOPScience

tions are taken into account, we consider an external potential V (x, t) without the
harmonic part, that is, when k(t) = 0. We then assume that the linear part of the
dissipative potential V (x, t) consists of a periodic linear magnetic/laser field with
strength σ/η and frequency θ , so that parameter λ(t) of the linear potential is given
by Eq. (7.102) [61, 70]. Then, we focus our attention to the situation when parame-
ters g(t) and (t) of respectively the two- and three-body inter-atomic interactions
are related by the condition |(t)| = 01 % |g(t)| . Next, we integrate Eqs. (7.76) to
arrive to either
 
1 75β0  α0  1
γ (t) = , g(t) = ,
2 (t + t0 ) 201  β0  (t + t0 )2
    
75  α0  1 75  α0  σ η cos [θ t]
α(t) = , β(t) = β + t − ,(7.105)
201  β0  |t + t0 | 201  β0  2
01
θ 2 (t + t0 )
    
75  α0  2 1 1
T (t) = − , when t0 = 0,
2  β 
01 0 t0 t +t 0

or
7.2 Soliton Management (SM) in One-Dimensional Bose-Einstein … 261
 
75β0  α0 
γ (t) = 0, g(t) = ,
201  β0 
    
75  α0  75  α0  2
|α(t)| = , T (t) = t, (7.106)
201  β0  201  β0 
  
75  α0  σ 2 η
β(t) = β01 + t − cos [θ t] , when t0 = 0,
2  β  2
01 0 θ2

where t0 and β01 are free constants (of integration). Then, When using parameters
given by Eq. (7.105), we follow Theocharis et al. [60] and take parameter t0 from
the condition t0 γ (t) > 0 for BECs with loss/gain of atoms. In the case when the set
of parameters is given by Eq. (7.105), our work time domain will be t ∈] − t0 , 0] if
t0 > 0 and t ∈ [0, −t0 [ if t0 < 0 so that T (t) will be a continuous function of time t
varying from zero to +∞. Computing the soliton width yields
 
201  β0 
wsol = wsol (t) = w0   |t + t0 | .
75 α 
0

With the set of parameters given in Eqs. (7.105) and (7.106), we display respec-
tively in Figs. 7.16 and 7.17 the spatiotemporal evolution of the matter-wave gray
soliton (plots of the top panels) and bright soliton (plots of the bottom panels). Pan-
els (a)–(c) [(d)–(f)]of these figures correspond to BECs with gain [loss] of atoms.
Plots (a) and (d) [(b) and (e)] are associated with BECs under the effects of the bias
magnetic field [laser field] only, while plots plots (c) and (f) are generated for BECs
with the combined bias magnetic and laser fields. It follows from plots of Figs. 7.16
and 7.17 that
(i) for a weak laser field, that is (η  0), the matter-wave solitons move with a con-
··
stant acceleration x = −σ along a parabolic trajectory whose direction depends
on the sign of the dissipative parameter γ (t), as one can observe from plots (a)
and (d);
(ii) when σ  0 (weak bias magnetic field), the matter-wave solitons move in +x
direction along an oscillatory trajectory, as it is well seen from plots (b) and (e);
(iii) for ησ = 0 (combination of the bias magnetic and laser fields effects), the mat-
ter wave solitons move along a parabolic oscillating trajectory whose direction
depends on the sign of the dissipative parameter γ (t), as we can see from plots
(c) and (f);
(iv) matter-wave amplitude keep no change in propagating in the BEC system, while
their speed changes according to the law

· d xc  η 
x c (t) = = − σ t + sin [θ t] .
dt θ

Now, we employ the kink solitonlike solution ρ(ξ ) = ρ+ (ξ ) given by Eq. (7.95)
to study the dynamics of matter-wave kink solitons. For this aim, we display in
Fig. 7.18 the spatiotemporal evolution of matter-wave kink solitons for the sets of
262 7 Dynamics of One-Dimensional Condensates …

Fig. 7.16 The dynamics of the matter wave gray soliton (top) and the matter wave bright soliton
(bottom) associated with respectively the soliton-like solutions (7.88) and (7.94) with υ = 0.5,
01 = 15, β01 = −10, and θ = 0.838. a–c The evolution plots of a matter wave gray soliton in a
linear trapping potential given by Eq. (7.102) either under a the effects of bias magnetic field, or
b the effects of laser field, or c the combined effects of both bias magnetic and laser fields. The
parameters are t0 = 100 (case of positive γ (t) corresponding to BEC system with gain of atoms),
γ0 = 0, α0 = 0.5, β0 = −0.15, and σ = 0.25 and η = 0 (a), σ = 0 and η = 1 (b), σ = 0.25 and
η = 1 (c). d–f The density plots of a matter wave bright soliton in a BEC system in a linear trapping
potential given by Eq. (7.102) either under (d) the effects of bias magnetic field, or (e) the effects
of laser field, or (f) the combined effects of both bias magnetic and laser fields. The parameters are
t0 = −50 (case of negative γ (t) corresponding to BEC system with loss of atoms), γ0 = −2β02 /α0 ,
α0 = −0.3, β0 = −0.05, and σ = 0.25 and η = 0 (d), σ = 0 and η = 1 (e), σ = 0.25 and η = 1
(f). Reprint from Ref. [61], Copyright 2022, with permission from IOPScience

parameters given by Eqs. (7.105) and (7.106) with different t0 , negative, zero, and
positive t0 corresponding respectively to γ (t) < 0, γ (t) = 0, and γ (t) > 0. Plots of
the top, middle, and bottom panels show the dynamics of matter-wave kink solitons
for respectively η = 0 and σ = 0, η = 0 and σ = 0, and σ η = 0. Various plots
reveal that (i) under the effects of the bias magnetic field only (η = 0 and σ > 0),
kink soliton moves along a parabolic trajectory in +x direction for BECs with loss
of atoms and in the −x direction for BECs with gain of atoms; (ii) when σ = 0 and
η = 0, the matter-wave kink soliton propagates along an oblique trajectory in +x
direction for nonnegative t0 and in −x direction for BECs with loss of atoms; (iii)
when γ (t) = 0, the matter-wave kink solitons propagate in −x direction, as we can
see from the plots of panels (t0 = 0) (iv) under the effects of the laser field (η = 0),
the matter-wave kink soliton propagates along an oscillating trajectory, as we can
see from the plots of the middle and bottom panels.
7.2 Soliton Management (SM) in One-Dimensional Bose-Einstein … 263

Fig. 7.17 Density plots of a–c gray solitons associated with the soliton-like solution (7.88) and
d–f bright solitons associated with the soliton-like solution (7.94) showing the effects of the bias
magnetic and laser fields on matter wave solitons propagating in a BEC system when both the gain
and loss of atoms are ignored, that is when t0 = 0. The parameters are the same as in Fig. 5.20.
Reprint from Ref. [61], Copyright 2022, with permission from IOPScience

7.2.4 Conclusion and Discussions

Using a generalized dissipative CQ GP equation with distributed coefficients, we


have investigated analytically the formation and the management of matter-wave
solitons in BECs with two- and three-body interactions when the gain/loss of the
condensate atoms is taken into account. For such BEC systems, we have studied
the phenomenon of the MI and found that the linear potential does not affect the
MI, while the feeding/loss of atoms seriously modifies the instability and stability
domain. The integrable condition for the GP equation under consideration is derived
and its exact solitonlike solutions including the gray, kink, and bright solutions are
presented. With the use of these exact solitonlike solutions, we have investigated both
the formation and the management of matter-wave solitons in our BEC systems. We
have found that (i) the external harmonic and linear trapping potentials can be use
to manage the matter-wave solitons in the BEC systems under consideration, (ii) the
found matter-wave gray and bright solitons execute periodic oscillations, (iii) the
wave amplitude keep no change, (iv) the soliton shape and speed change with time,
and (v) matter-wave solitons can be compressed by manipulating the three-body
inter-atomic interactions.
264 7 Dynamics of One-Dimensional Condensates …

Fig. 7.18 Dynamics of kink soliton associated with the kink soliton-like solution ρ(ξ ) = ρ+ (ξ )
given by Eq. (3.144) under the effects of (top) only the bias magnetic field, (middle) only the laser
field, (bottom) combined effects of the bias magnetic and laser fields. Plots of the first and the
third columns correspond respectively to BEC systems with loss and with gain of atoms, while
plots of the middle column corresponds to BEC system when the feeding and the loss of atoms are
β2
ignored. The parameters are α0 = 0.5, β0 = −0.5, γ0 = 23 α00 , υ = 0.15, 01 = 15, β01 = −10,
t0 = ±1000, σ = 0.25 (top, bottom), σ = 0 (middle), η = 0, η = 10 (middle, bottom), θ = 0.1
when t0 = 0, and either θ = π or θ = 0 when t0 = 0. Reprint from Ref. [61], Copyright 2022, with
permission from IOPScience

7.3 Chirped Solitons and Chirped Double-Kink Solitons


in Bose-Einstein Condensates with Time-Varying
Atomic Scattering Length in a Complicated External
Potential

This section deals with a one-dimensional cubic-quintic GP equation with distributed


coefficient containing a time-varying dissipative term. The integrable conditions of
such equation are derived. The concept of the “compensation” is introduced and
its condition is established. Exact solitonlike solutions with nonlinear chirping are
obtained. These solutions are used to describe the dynamics of chirped femtosecond
BEC solitons and chirped double-kink BEC solitons. We show that the chirping of
found chirped BEC solitons are localized and their feature depends on the sign of
the self-steepening parameter.
7.3 Chirped Solitons and Chirped Double-Kink Solitons … 265

7.3.1 Introduction

At very low temperatures, the matter-wave function (r, t) at position r and time t
of BECs trapped in an external potential V (r, t) can usually be described by the GP
equation [52, 71–76]
 
∂(r, t) 2 2
i = − ∇ + V (r, t) + g0 |(r, t)| + χ0 |(r, t)| (r, t),
2 4
∂t 2m
(7.107)
where g0 = 42 as (t)/m and χ0 are the strengths of respectively the two- and three-
body inter-atomic interactions,  and m are respectively the Planck constant and the
atomic mass, and as (t) is the s-wave scattering length. In this Section, we focus our
attention to BEC systems with external potential given by Anderson et al. [77]
m 2 2  2 
V (r, t) = ω x + ω⊥
2
y + z 2 + i
γ , with ω  ω⊥ ,
2
where  γ is the gain/loss parameter. As in the previous Section, we follow Gammal
et al. [75] and assume that χ0 = χ 0 g0 with 0 ≤ χ0 < 1. Next, we introduce the
transformation
   
1 y2 + z2 x
(r, t) = √ exp −iω⊥ t − 2
ψ , ω⊥ t
2πa B a⊥ 2a⊥ a⊥

to obtain the following GP equation with an external dissipative harmonic potential


[78–82]

∂ψ 1 ∂ 2ψ  
i + + g |ψ|2 ψ + χ
0 g |ψ|4 ψ − αx 2 + iγ ψ = 0. (7.108)
∂t 2 ∂x 2

Here, a B is the Bohr radius. Throughout this Section, we assume that

(i) t and x are√the temporal and spatial coordinates measured in respectively 1/ω⊥
and a⊥ = / (mω⊥ );
(ii) parameter g of the two-body inter-atomic interactions is a function of time t and
is given g = −2as /(3a B ) and can take any sign [83];
(iii) strength of the harmonic potential α expresses the trapping frequency in the x−
direction [84], is defined as α = ±ω2 /(2ω⊥2
), and can take any sign, positive for
confining potential and negative for repulsive potential;
(iv) parameter γ related to the feeding (γ > 0) or loss (γ < 0) of atoms in the
condensate [53, 68, 85].

Under the above conditions, we here intend to study the generation of chirped fem-
tosecond solitons and chirped double-kink solitons [86–92] in BECs with time-
dependent atomic scattering length in a time-varying complex potential when the
gain/loss of atoms is taken into account. For this purpose, we first reduce the GP
266 7 Dynamics of One-Dimensional Condensates …

equation (7.108) into a NLS equation with higher-order (HO) terms which may sup-
port the propagation of ultrashort solitons including dark/bright and double-kink
solitons [93–106]. Such a reduction can be made possible via the so-called modified
phase-imprint method. Once this reduction is done, the integrable and the compensa-
tion conditions for the derived higher-order NLS (HO-NLS) equation are presented.
Under the integrable conditions, we adopt a nonlinear chirping ansatz which differs
from that used by many authors [96, 100] to present various families of chirped
solitonlike solutions of the found HO-NLS equation. Using the found analytical
solutions of the HO-NLS equation, we study the dynamics of chirped femtosecond
solitons and chirped double-kink solitons in the BECs with time-varying scattering
length and loss/gain of atoms in a harmonic trapping potential.

7.3.2 Derivation of the Higher-Order Nonlinear Schrödinger


Equation

Integrable and Compensation Conditions for a GP Equation We start this Section


by first finding the integrable and compensation conditions for the GP equation
(7.108), conditions under which it can be reduced into a HO-NLS equation. For
this aim, we perform the following lens-type transformation with a phase-imprint
parameter [107]  
ψ(x, t) = φ(X, T ) exp i γ x 2 + θ , (7.109)

where T = T (t) is a real function t, X = X (x, t) = g(t) g0
x for any real constant g0
to be taken from the condition that g0 g(t) > 0 for all t, and θ = θ (X, T ) is a variable
phase-imprint parameter. Next, we select different functional parameters T, g, γ ,
and θ from the following conditions

dT g
= , (7.110a)
dt g0
1 dg
+ 4γ = 0, (7.110b)
g dt

+ 2γ 2 + α = 0, (7.110c)
dt
∂θ
= α0 |φ|2 , (7.110d)
∂X
    
∂θ 3 ∂φ ∗ ∂φ 1 3
=i α0 − β0 φ + φ∗ + g0 χ0 − β02 + α0 (β0 − α0 ) |φ|4 ,
∂T 2 ∂X ∂X 2 2
(7.110e)

where α0 = 0 and β0 two free parameters. Demanding under conditions (7.110a)–


(7.110e) that Eq. (7.109) satisfies the GP equation (7.108), we arrive to the following
7.3 Chirped Solitons and Chirped Double-Kink Solitons … 267

HO-NLS equation with self-steepening and self-frequency shift

∂φ 1 ∂ 2φ ∂  2  ∂ |φ|2
i + +g0 |φ| 2
φ + iα0 |φ| φ + i (β0 − 2α0 ) φ
∂T 2 ∂ X2 ∂X ∂X
1
+ (β0 − α0 ) (β0 − 2α0 ) |φ| φ = 0,
4
(7.111)
2
which in the context of nonlinear optics can be used to model the propagation of
femtosecond (ultrashort) pulses in a single-mode optical fiber [108, 108–112]. In the
below study, solitonlike solutions of the HO-NLS equation (7.111) will be referred
to as the femtosecond or double-kink solitonlike solutions, and any solitonlike solu-
tion of the GP equation (7.108) built with the use of these solutions will be called
femtosecond or double-kink solitonlike solutions of Eq. (7.108).
From the HO-NLS equation (7.111), one can derive a series of well-known inte-
grable equations of Schrödinger type. Indeed, when
(i) β0 − α0 = 0, Eq. (7.111) reduces to the Chen–Lee–Liu type NLS (CLL–
NLS) equation [113], the Kaup–Newell type NLS (KN–NLS) equation [114] can
be obtained from Eq. (7.111) by setting β0 − 2α0 = 0,
(ii) by setting β0 = 0, the HO-NLS equation (7.111) coincides with the Gerdjikov–
Ivanov (GI) equation [115].
Throughout this Section, conditions (7.110b) and (7.110c) under which the HO-
NLS equation with self-steepening and self-frequency shift (7.111) is derived will
be referred to as the integrable conditions of the GP equation (7.108). From Eq.
(7.110b), we can express parameter g(t) of the two-body inter-atomic interactions
in term of the gain/loss parameter γ (t) as follows
⎡ ⎤
t
g(t) = λ0 exp ⎣−4 γ (τ )dτ ⎦ ,
0

where λ0 is a free real constant. This means that under the above integrable conditions,
parameter γ (t) of the gain/loss of the condensate atoms can be used to manage the
matter-wave solitons (soliton peak and soliton compression) [53, 116]. Throughout
this Section, we understand by compensation process, any process which consists
of a balance of loss or gain effects with the effects of the s-wave scattering length
on the condensate. The compensation process will then ensures the stability of the
condensates over a longer interval of time. Therefore, Eq. (7.110b) which link g(t)
and γ (t) will be referred to as the “ compensation condition”.
Chirped Femtosecond and Chirped Double-Kink Solitonlike Solutions
of the HO-NLS Equation Now, we turn our attention to the exact solitonlike solu-
tions of the HO-NLS equation (7.111). For simplicity, we limit ourselves to traveling
wave solution whose form is

φ(X, T ) = ρ(ξ ) exp [i (χ (ξ ) − T )] . (7.112)


268 7 Dynamics of One-Dimensional Condensates …

Here, ξ = X − υT is the traveling coordinate, ρ and χ are reals functions of the


single variable ξ , and υ is a real parameter that determines the group velocity of
the wave packet. Next, we associate with the traveling wave solution (7.112) the
following frequency chirp

∂ d

δ ω(X, T ) = − (χ (ξ ) − T ) = − χ (ξ ) = −χ (ξ ).
∂X dξ

We call a solitonlike solution of form (7.112) a “ chirped solitonlike solution” if the


corresponding chirping is nonlinear.
The traveling wave solution (7.112) and ansatz (7.109) lead to the following
traveling wave solution of the GP equation by (7.108)
 
ψ(x, t) = ρ(ξ ) exp i γ x 2 + θ (X, T ) + χ (ξ ) − T . (7.113)

Using Eq. (7.110d), we find the corresponding chirping to be


 $ 
g 

δω(x, t) = − 2γ x + α0 ρ (ξ ) + χ (ξ )
2
t
. (7.114)
g0 ξ= g x− gυ g(τ )dτ
g0 0
0

In order to obtain exact chirped solitonlike solutions of Eq. (7.111), we ask that Eq.
(7.112) satisfies Eq. (7.111). We then obtain that ρ and χ must be solutions of the
following differential system:
 
dχ 1 dχ 2 1 d 2ρ dχ 3
ρ + υ ρ− ρ+ 2
− α0 ρ + g0 ρ 3
dξ 2 dξ 2 dξ dξ
1
+ (β0 − α0 ) (β0 − 2α0 ) ρ 5 = 0, (7.115a)
2
dρ 1 d 2χ dχ dρ dρ
−υ + ρ+ + (2β0 − α0 ) ρ 2 = 0. (7.115b)
dξ 2 dξ 2 dξ dξ dξ

A first integral of Eq. (7.115b) reads

dχ α0 − 2β0 2
= ρ + υ. (7.116)
dξ 2

Inserting now Eq. (7.116) into Eqs. (7.114) and (7.115a), we obtain respectively that
 $  
g 3α0 − 2β0 2
δω(x, t) = − 2γ x + υ+ ρ (ξ ) t
, (7.117)
g0 2 ξ = g x− υ g(τ )dτ
g0 g0
0

and

d 2ρ
+ b5 ρ 5 + b3 ρ 3 + b1 ρ =0, (7.118)
dξ 2
7.3 Chirped Solitons and Chirped Double-Kink Solitons … 269

3
b5 = α0 (α0 − β0 ) , b3 = 2 (g0 − α0 υ) ,b1 = 2 + υ 2 . (7.119)
2
It follows from Eq. (7.117) that the chirping amplitude can be controlled by varying
the free parameters g0 , α0 , and β0 . We can also see from Eq. (7.117) that the chirping
δω(x, t) consists of two parts, one part which is proportional to the wave density
|ψ(x, t)|2 = ρ 2 (x, t), and one linear time-dependent term.
To exploit some known classes of solutions of Eq. (7.118), especially its localized
solutions which can be expressed in terms of the Weierstrass elliptic function ℘ [35,
36, 117], we set r (ξ ) = ρ 2 (ξ ) reducing Eq. (7.118) to the following form
 2
dr
= α1r 4 + 4β1 r 3 + 6γ1r 2 + 4δ1 r ≡ R(r ), r = ρ2, (7.120)

8 b3 2
α1 = − b5 , β1 = − , γ1 = − b1 , (7.121)
3 2 3
where δ1 is a free constant of integration. As in the previous Sections, for Eq. (7.120)
to admit solitonlike solutions, it is sufficient

g2 = −4β1 δ1 + 3γ12 ≥ 0, g3 = 2β1 γ1 δ1 − α1 δ12 − γ13 ≤ 0, and  = g23 − 27g32 = 0. (7.122)

Under conditions (7.122), Eq. (7.120) admits the following solitonlike solution
[35, 36]

1 d R(r0 ) sinh2 3e1 ξ
r (ξ ) = r0 +   , (7.123)
4 dξ 3e + e − 1 d 2 R(r0 ) sinh2 √3e ξ
1 1 24 dξ 2 1


where e1 = 3 −g3 and r0 is any simple zero of R(r ). For solution (7.123) to be
nonnegative, additional restrictions must be imposed on b1 , b3 , b5 . In what follows,
we discuss, dependent on the behavior of the parameters b5 , b3 , b1 , and δ1 . We will
distinguish two special cases, the case b5 b3 b1 = 0 and the case b5 b3 b1 = 0.
Case b5 = 0 and b3 b1 = 0 Asking that  = 0 in the special case when b5 = 0 and
9γ 2
b3 b1 = 0, we obtain two special values for δ1 , namely δ1 = 0 and δ1 = 16β11 . Using
3γ1 2 +υ 2
Eq. (7.123) with the simple zero r0 = − 2β1
= α0 υ−g0
obtained for δ1 = 0 and Eq.
(7.123) yields
%
2 + υ 2 1
ρ(ξ ) = −   , 2 + υ 2 < 0, g0 − α0 υ > 0.
g0 − α0 υ cosh − 2 + υ 2 ξ

For this special solution ρ(ξ ) of Eqs. (7.118), (7.113) and (7.117) under the conditions
that 2 + υ 2 < 0 and g0 − α0 υ > 0 give the following solution of the GP equation
(7.108) with the corresponding chirping
270 7 Dynamics of One-Dimensional Condensates …

ψ(x, t)
%   
2 + υ 2 exp i γ x 2 + θ (X, T ) + χ (ξ ) − T
= −  & ' , (7.124)
g0 − α0 υ   g(t) υ  g(τ )dτ
t
cosh − 2 + υ 2 g x − g
0 0
0
δω(x, t)
⎡ ⎛ ⎞⎤
%  
⎢ ⎜ (3α0 − 2β0 ) 2 + υ 2 ⎟⎥
⎢ g(t) ⎜ ⎟⎥
= −⎢
⎢2γ (t)x +
⎜υ −
⎜  &

' ⎟⎥
⎟⎥ .
⎣ g0 ⎝   t ⎠⎦
2 (g0 − α0 υ) cosh − 2 + υ g(t) υ
g0 x − g0 g(τ )dτ
2 2
0
(7.125)

9γ 2
In the special situation when δ1 = 16β11 = 0, we apply the direct method to obtain
the following solitonlike solution of Eq. (7.118)
% $ 
2 + υ 2 2 + υ 2
ρ(ξ ) = tanh ξ , 2 + υ 2 > 0, g0 − α0 υ < 0.
2 (α0 υ − g0 ) 2
(7.126)
9γ 2
It is important to note that when δ1 = 16β11 = 0, r0 = 0 will be a simple zero
of R(r ). The corresponding solitonlike solution of Eq. (7.118) is obtained via Eq.
(7.120) and reads
%  √ 

1 3 2 + υ 2 sinh 2 + υ 2ξ
ρ(ξ ) =  , 2 + υ 2 > 0, g0 − α0 υ < 0.
2 α0 υ − g0 3 + 2 sinh [x]
2
(7.127)
The above two kink soliton solutions (7.126) and (7.127) of Eq. (7.118) yield the
following kink soliton solutions of the GP equation (7.108) with the corresponding
chirping
% ⎡$ ⎛% ⎞⎤
t
2 + υ 2 2 + υ 2 g(t) υ
ψ(x, t) = tanh ⎣ ⎝ x− g(τ )dτ ⎠⎦
2 (α0 υ − g0 ) 2 g0 g0
0
 
× exp i γ x 2 + θ + χ (ξ ) − T , (7.128)
 % &   '
g(t) (3α0 − 2β0 ) 2 + υ 2
δω(x, t) = − 2γ (t)x + υ+
g0 4 (α0 υ − g0 )
⎡$ ⎛ ⎞⎤⎤
$ t
2 + υ 2 g υ
× tanh2 ⎣ ⎝ x− g(τ )dτ ⎠⎦⎦ , (7.129)
2 g0 g0
0
7.3 Chirped Solitons and Chirped Double-Kink Solitons … 271

and
  
%  √ υ
t
 sinh 2 + υ 2 g(t)
x − g(τ )dτ
1 3 2 + υ 2 g0 g0
0
ψ(x, t) = %   
2 α0 υ − g0 √ t
υ
3 + 2 sinh2 2 + υ 2 g(t)
g0
x − g0
g(τ )dτ
0
 
× exp i γ x 2 + θ (X, T ) + χ (ξ ) − T , (7.130)
 % &  
g(t) 3 (3α0 − 2β0 ) 2 + υ 2
δω(x, t) = − 2γ (t)x + υ+
g0 4 (α0 υ − g0 )
   ⎞⎤
√ υ
t
sinh 2
2 + υ 2 g(t)
g0
x − g0 g(τ )dτ ⎟⎥
0 ⎟⎥
×    ⎟⎥ . (7.131)
√ υ
t ⎠⎦
3 + 2 sinh2 2 + υ 2 g(t)
g0
x − g0
g(τ )dτ
0

Here, α0 = 0, g0 = 0, υ, and are four parameters satisfying the conditions 2 +


υ 2 > 0 and g0 − α0 υ < 0.
Case b3 = 0 and b5 b1 = 0 In the special case when b3 = 0 and b5 b1 = 0, condition
 = 0 is satisfied only when either

$ %
γ1 2  2 + υ 2
δ1 = 0 or δ1 = ±γ1 −2 = ± 2 + υ 2
− .
α1 3 3α0 (α0 − β0 )
 
Assuming that α0 (α0 − β0 ) 2 + υ 2 < 0, we focus our attention to the situation
δ1 = 0 so that r0 = 0 will be a simple zero of R(r ). We then obtain, with the help
of Eq. (7.123), the following kink soliton solution of Eq. (7.118) if 2 + υ 2 > 0
and α0 (α0 − β0 ) < 0:
%    
2 2 + υ2 sinh2
2 2 + υ 2 ξ
ρ 2 (ξ ) = −    .
3 3α0 (α0 − β0 ) 2 + sinh2 2 2 + υ 2 ξ

The corresponding kink soliton solution of the GP equation defined (7.108) and the
corresponding chirping are given respectively as follows:

(   
)
) %   g(t) υ  t g(τ )dτ
) sinh2 2 2 + υ 2 x −
)2 2 + υ 2 g0 g0 0
ψ(x, t) = ) −   
*3 3α0 (α0 − β0 )   g(t) υ t
2 + sinh2 2 2 + υ 2 g x − g 0 g(τ )dτ
0 0
  
× exp i γ x 2 + θ (X, T ) + χ (ξ ) − T , (7.132)
272 7 Dynamics of One-Dimensional Condensates …
⎡ % ⎛ %
g(t) ⎝ 3α0 − 2β0 2 + υ 2
δω(x, t) = − ⎣2γ (t)x + υ+ −
g0 3 3α0 (α0 − β0 )
   ⎞⎤
  g(t) υ t
sinh2 2 2 + υ 2 g0 x − g0 0 g(τ )dτ ⎟⎥
×    ⎟ ⎥
⎠⎦ . (7.133)
  g(t) υ  t
2 + sinh2 2 2 + υ 2 g x − g 0 g(τ )dτ
0 0

Here, α0 = 0, β0 , , and υ are four constants satisfying the conditions 2 + υ 2 > 0


and α0 (α0 − β0 ) < 0.
Case b1 = 0 and b5 b3 = 0 In this special case,  = 0 is possible only when either

64 β13
δ1 = 0 or δ1 = − .
27 α12

β3
Choosing δ1 = − 64 1
27 α12
so that r0 = 0 should be a simple zero of R(r ) and using Eq.
(7.123) lead, under the conditions α0 (α0 − β0 ) < 0, g0 − α0 υ > 0, to
 
α0 υ − g0 sinh 2
2 (g0 − α0 υ) − 1
3α0 (α0 −β0 )
ξ
r (ξ ) =  .
3α0 (α0 − β0 ) 3 + sinh2 2 (g − α υ) − 1
ξ
0 0 3α0 (α0 −β0 )

The corresponding solitonlike solution of the GP equation given by (7.108) with the
corresponding chirping are given respectively as follows:

(  & '
)
) υ t
) 1
2 (g0 − α0 υ) − 3α (α −β ) g(t)
) sinh2 g0 x − g0 g(τ )dτ
) α0 υ − g0 0 0 0
0
ψ(x, t) = )  &

'
) 3α0 (α0 − β0 ) t
* 1 g(t) υ
3 + sinh2 2 (g0 − α0 υ) − 3α (α −β ) g x − g g(τ )dτ
0 0 0 0 0
0

  
× exp i γ x 2 + θ (X, T ) + χ (ξ ) − T , (7.134)

 % 
g(t) (3α0 − 2β0 ) (α0 υ − g0 )
δω(x, t) = − 2γ (t)x + υ+
g0 6α0 (α0 − β0 )

  t  ⎞⎤
g(t) υ
sinh2 2 (g0 − α0 υ) − 3α0 (α10 −β0 ) g0 x − g0 0 g(τ )dτ ⎟⎥
×   t  ⎠⎦ , (7.135)
g(t) υ
3 + sinh2 2 (g0 − α0 υ) − 3α0 (α10 −β0 ) g0 x − g0 0 g(τ )dτ

where α0 = 0, β0 , g0 = 0, and υ are real constants satisfying the conditions


α0 (α0 − β0 ) < 0 and g0 − α0 υ > 0.
7.3 Chirped Solitons and Chirped Double-Kink Solitons … 273

Fig. 7.19 Amplitude profile of the soliton solution in Eq. (7.136) for different values of the solution
2 β1 8 β1
2 3
parameters. a α1 = −1, β1 = 1, γ1 = 3 α1 , and δ10 = 27 α 2 ; b γ1 = δ10 = 1, α1 = 0.248872, and
1
β1 = 0.56137. Reprint from Ref. [106], Copyright 2022, with permission from Springer

Case b5 b3 b1 = 0 In this general case, one may solve equation  = 0 for δ1 . To any
positive solution δ10 of equation  = 0 that satisfies the conditions γ13 − 2β1 γ1 δ10 +
α1 δ10
2
> 0 and 3 γ13 − 2β1 γ1 δ10 + α1 δ10
2
− 21 γ1 > 0, r0 = 0 will be a simple zero of
R(r ) which together with Eq. (7.123) lead to the following double-kink-type soliton
solution of Eq. (7.118)
% √
δ10 sinh 3e1 ξ
ρ(ξ ) = ± √ , e1 = 3 γ13 − 2β1 γ1 δ10 + α1 δ10
2
.
e1 − 21 γ1 3e1
+ sinh2 3e1 ξ
e1 − 21 γ1
(7.136)
To observe the double-kink features of the solitonlike solution (7.136) of Eq. (7.118),
we plot in Fig. 7.19 the wave amplitude ρ(ξ ) = ρ+ (ξ ) given by Eq. (7.136) for
two sets of parameters α1 , β1 , γ1 , and δ10 leading to e −3e11 γ = 65 for Fig. 7.19a and
1 2 1
3e1
e1 − 21 γ1
= 1003 for Fig. 7.19b. As we can see from plots of Fig. 7.19, the interesting
double-kink feature of the solution (7.136) exists only for large enough values of
3e1
e1 − 21 γ1
.
With the use of the double-kink-type soliton solution (7.136), we obtain the fol-
lowing double-kink-type soliton solution of the GP equation (7.108) with the corre-
sponding frequency chirp:

 & '
√ t
% g(t) υ
sinh 3e1 g x− g 0
g(τ )dτ
0
3δ10 0
ψ(x, t) = (  & '
3e1 + 2 + υ 2 )
)
√ υ 
t
* 9e1
+ sinh2 g(t)
3e1 +2 +υ 2
3e1 g0 x − g0 g(τ )dτ
0

  
× exp i γ x 2 + θ (X, T ) + χ (ξ ) − T , (7.137)
274 7 Dynamics of One-Dimensional Condensates …

 % &
g(t) 3δ10 (3α0 − 2β0 )
δω(x, t) = − 2γ (t)x + υ+  
g0 2 3e1 + 2 + υ 2

 & ' ⎞⎤

√ υ 
t
g(t)
sinh2 3e1 g0 x − g0 g(τ )dτ ⎟⎥
0 ⎟⎥
×  & ' ⎟⎥
⎟⎥ ; (7.138)

2 √3e υ  g(τ )dτ ⎠⎦
t
9e1 g(t)
+ sinh 1 x −
3e1 +2 +υ 2 g0 g0
0

here,
$
2 3 3 9   27
e1 = − 2 + υ 2 + (g0 − α0 υ) 2 + υ 2 δ10 + α0 (α0 − β0 ) δ10
2
,
3 2 2

δ10 is any root of equation


   
27α12 δ12 + 4β1 16β12 − 27α1 γ1 δ1 + 18γ12 3α1 γ1 − 2β12 = 0,

and α0 = 0, g0 = 0, β0 , υ, and are five free constants to be chosen from conditions



1
γ13 − 2β1 γ1 δ10 + α1 δ10
2
> 0 and 3 γ13 − 2β1 γ1 δ10 + α1 δ10
2
− γ1 > 0,
2
α1 , β1 , and γ1 being given by Eq. (7.121).
In the special situation when δ1 = 0, interesting solitonlike solutions of Eq. (7.118)
can be sought by the direct method. As an example of kink solitonlike solution of
Eq. (7.118) under the condition δ1 = 0, we obtain
  
g0 − α0 υ g0 − α0 υ
ρ (ξ ) =
2
1 ± tanh √ ξ ,
4α0 (β0 − α0 ) 2 α0 (β0 − α0 )

under the conditions


 
α0 (β0 − α0 ) > 0, g0 − α0 υ > 0, and (g0 − α0 υ)2 − 4α0 2 + υ 2 (α0 − β0 ) = 0.

To the above kink soliton solution of Eq. (7.118) corresponds the following kink
solitonlike solution of the GP equation (7.108) and the corresponding chirping
( ⎛ ⎡ ⎛% ⎞ ⎤⎞
) t
)
1 ) g0 − α0 υ ⎝ g − α υ g(t) υ
1 ± tanh ⎣ √ ⎝ g(τ )dτ ⎠⎦⎠
0 0
ψ(x, t) = * x−
2 α0 (β0 − α0 ) 2 α0 (β0 − α0 ) g0 g0
0

  
× exp i γ x 2 + θ (X, T ) + χ (ξ ) − T , (7.139)
7.3 Chirped Solitons and Chirped Double-Kink Solitons … 275
 $ 
g 3α0 − 2β0 g0 − α0 υ
δω(x, t) = − 2γ x + υ+
g0 2 4α0 (β0 − α0 )

⎛ ⎡ ⎛% ⎞⎤⎞⎞⎤
t
g − α υ g(t) υ
× ⎝1 ± tanh ⎣ √ ⎝ g(τ )dτ ⎠⎦⎠⎠⎦ ,
0 0
x− (7.140)
2 α0 (β0 − α0 ) g0 g0
0

where α0 = 0, g0 = 0, β0 , υ, and are constants satisfying the conditions


 
α0 (β0 − α0 ) > 0, g0 − α0 υ > 0, 4α0 (α0 − β0 ) 2 + υ 2 − (g0 − α0 υ)2 = 0.

If various parameters satisfy the condition

γ1 > 0, β1 < 0, and 2β12 − 3γ1 α1 > 0,

then Eq. (7.118) will admit the bright solitonlike solution

3γ1 1
ρ 2 (ξ ) = − $
β1 6β12 −9γ1 α1 √
1+ 6β12
cosh 6γ1 ξ

which leads to the following bright solitonlike solution of the GP equation (7.108)
with the corresponding chirping

(
)
) 2 + υ 2
ψ(x, t) = 2) &  & ''
)
* 2 α υ − g −
  g(t) υ  g(τ )dτ
t
0 0 λ0 cosh 2 − 2 + υ 2
g0 x − g0
0
  
× exp i γ x 2 + θ (X, T ) + χ (ξ ) − T , (7.141)
 %
g(t)
δω(x, t) = − 2γ (t)x + (υ + (2β0 − 3α0 )
g0

⎞⎤
 
2 + υ 2 ⎟⎥
⎟⎥
×  & ' ⎟ ⎥,
⎟⎥ (7.142)
 
υ  g(τ )dτ ⎠⎦
t
g0 − α0 υ + 
λ0 cosh 2 − 2 + υ 2 g(t)
g0 x − g0
0

where  

λ0 = (g0 − α0 υ)2 − 4α0 (α0 − β0 ) 2 + υ 2 ,

and α0 = 0, g0 = 0, β0 , υ, and are real constants satisfying the conditions


 
2 + υ 2 < 0, g0 − α0 υ > 0, and (g0 − α0 υ)2 − 4α0 (α0 − β0 ) 2 + υ 2 > 0.
276 7 Dynamics of One-Dimensional Condensates …

Fig. 7.20 The dynamics of a double-kink soliton in a time-independent harmonic trapping potential
given by Eq. (7.137) and the corresponding frequency chirp given by Eq. (7.138). a, c Spatiotem-
poral evolution of a double-kink soliton and the corresponding chirp for g(t) = λ0 exp [4λt]. b, d
Evolution of a double-kink soliton and the corresponding chirp for g(t) = λ0 exp [−4λt]. Differ-
α 2 +0.62 218
ent plots are generated with υ = g0 +0.56137
2
α0 , = − 3+2υ
4 , δ1 = 1, β0 =
0
α0 , g0 = 1.8,
α0 = 0.5, λ0 = 0.1, and λ = 0.05. Reprint from Ref. [106], Copyright 2022, with permission from
Springer

7.3.3 Dynamics of Chirped Femtosecond Solitons and


Chirped Double-Kink Solitons in Bose-Einstein
Condensates with Time-Dependent Atomic Scattering
Length in a Complex Potential

Now, we use the found above exact solitonlike solutions of Eq. (7.108) to investigate
analytically the dynamics of chirped femtosecond solitons and chirped double-kink
solitons in BECs described by the GP equation (7.108). Analyzing the found above
exact solitonlike solutions ψ(x, t) of the GP equation (7.108) and the corresponding
chirping δω(x, t), we find that the motion of the centre ζ (t) of different matter-wave
solitons is given by
$ t
υ g0
ζ (t) = g(τ )dτ
g0 g(t)
0
7.3 Chirped Solitons and Chirped Double-Kink Solitons … 277

Fig. 7.21 Effects of the self-steepening coefficient α0 on the dynamics of a bright femtosecond
soliton in a time-independent harmonic trapping potential given by equation Eq. ( 7.141) and the
corresponding chirp defined by Eq. (7.142) for g(t) = λ0 exp [4λt] and g0 = 1.8 with two values
of α0 , α0 = 1 for plots of the top panels and α0 = −1 for plots (c) and (d). Other values of different
parameters are given in the text. Reprint from Ref. [106], Copyright 2022, with permission from
Springer

and satisfies the equation ddt ζ2 + 2αζ = 0. Therefore, the wave centre behaves like
2

a classical particle. We conclude that we can employ the external harmonic trap-
ping potential to manipulate the motion of chirped femtosecond solitons and chirped
double-kink solitons in BEC systems. Taking into account the above integrable con-
ditions, we find that
(i) the centre of chirped femtosecond solitons and chirped double-kink solitons
moves with the speed
⎛ ⎞
t $
dζ υ ⎝ g0
= 2γ g(τ )dτ + g ⎠ ;
dt g0 g(t)
0

(ii) the width of chirped


femtosecond solitons and chirped double-kink solitons is
g0
proportional to g(t) , meaning that the found above exact solution of the GP
equation may be used to describe the matter-wave compression;
(iii) the amplitude of matter-wave femtosecond pulses does not depend neither on
nonlinearity parameter g(t) nor on the gain/loss parameter γ (t).
An analysis of chirping δω(x, t) found in the previous subsection reveals that
278 7 Dynamics of One-Dimensional Condensates …

Fig. 7.22 Dynamics of the double-kink soliton with the corresponding chirp associated to solution
(7.137) with chirp (7.138) of the GP Equation (7.108) for the self-steepening coefficient having
different sign. Plots (a) and (b) are generated with α0 = 0.5, while (c) and (d) are obtained with
α0 = −0.5. Other parameters used in different plots are λ0 = −0.1, g0 = −1.8, m = 0.4, and
! = 2. Reprint from Ref. [106], Copyright 2022, with permission from Springer

(i) δω(x, t) depends both on the loss/gain parameter γ (t) and parameter g(t) of the
two-body inter-atomic interactions;
(ii) chirping δω(x, t) consists of two parts, one linear part which is proportional
to γ (t) = − 14 g1 dg , and one nonlinear part proportional to the wave intensity
dt
|ψ(x, t)|2 with coefficient g(t) g0
.
In the following, we illustrate the above study by some examples of 1D BEC
systems with different kinds of scattering length when the gain/loss of atoms is taken
into account. The following numerical values of different parameters in the case of
double-kink soliton solution (7.137) will be used

g0 + 0.56137 3 + 2υ 2 α 2 + 0.62 218


δ1 = 1, υ = , = − , andβ0 = 0 .
α0 4 α0

For the dynamics of bright femtosecond solitons, we will use either the set

1 1 3 + 2υ 2
α02 = 1, β0 = , υ = , g0 > 1, (g0 − 1)2 + 5 > 0, and = − ,
6α0 α0 4

or the set
7.3 Chirped Solitons and Chirped Double-Kink Solitons … 279

Fig. 7.23 Spatiotemporal evolution of the bright femtosecond soliton with the corresponding chirp
associated to solution (7.141) with chirp (7.142) of the GP Eq. (7.108) for the self-steepening
coefficient having different sign. Plots (a) and (b) are obtained with α0 = 1, while (c) and (d) are
generated with α0 = −1. Other parameters used in different plots are λ0 = 2.1, g0 = 1.8, m = 0.4,
and ! = 1. Reprint from Ref. [106], Copyright 2022, with permission from Springer

1 1 3α 2 + 2
α02 = 1, g0 > −1, β0 = , υ = − , and = − 0 2 .
α0 α0 4α0

BECs with Time-Independent Harmonic Potential Our first example the BECs
system with time-independent harmonic potential used in the creation of bright BEC
solitons [118]. For this example, we have α = −2λ2 with λ ≈ 0.05. It then follows
from the integrable conditions (7.110b) and (7.110c) that γ (t) = ±λ and g(t) =
λ0 exp [∓4λt] , where λ0 = 0 is a free constant to be taken from the condition λ0 g0 >
0. For the present example, we depict in Fig. 7.20 the dynamics of a double-kink
soliton (to panels) and the corresponding chirping (bottom panels). Plots of the
left panels [right panels] correspond to g(t) = λ0 exp [4λt] and γ (t) = −λ [g(t) =
λ0 exp [−4λt] and γ (t) = λ]. It is seen from different plots that for BEC with gain
of atoms, both the soliton and the corresponding chirping have double-kink feature
(plots of the left panels).
To show the effects of the self-steepening coefficient α0 on the dynamics of the
femtosecond soliton, we use the set of parameters
280 7 Dynamics of One-Dimensional Condensates …

Fig. 7.24 Spatiotemporal evolution of (left panels) the double-kink soliton given by Eq. (7.137)
and (right panels) the corresponding chirp defined by Eq. (7.138) for BEC systems with a time-
varying hyperbolic s-wave scattering length with g(t) = λ0 (1 + tanh [μt]). Plots (a) and (b) are
generated with a positive self-steepening coefficient α0 (α0 = 0.5), while (c) and (d) are obtained
with a negative self-steepening coefficient α0 (α0 = −0.5). Other parameter used in plots (a)–(d)
are λ0 = 0.3, g0 = 1.8, and μ = 2, and m = 0.99. Other parameters are given in the text. Reprint
from Ref. [106], Copyright 2022, with permission from Springer

1 1 3 + 2υ 2
α02 = 1, β0 = , υ = , g0 > 1, (g0 − 1)2 + 5 > 0, = −
6α0 α0 4

with g0 = 1.8 and α0 = ±1 to depict in Fig. 7.21 the spatiotemporal evolution of the
femtosecond soliton given by Eq. (7.141) and the corresponding chirping (7.142).
BECs with a Temporal Periodic Modulation of the S-Wave Scattering Length
Our second example is a BEC system with temporal periodic modulation of the s-
wave scattering length [119] with the parameter of the two-body interactions given
as
g(t) = λ0 (1 + m sin [! t]) ,

where 0 < m < 1, λ0 = 0 being any real constant. From the integrable conditions
(7.110b) and (7.110c), we obtain
7.3 Chirped Solitons and Chirped Double-Kink Solitons … 281

Fig. 7.25 The dynamics of (left panels)a bright femtosecond soliton defined by Eq. (7.141) and
(right panel) the corresponding frequency chirp given by Eq. (7.142) for BEC systems with a time-
varying hyperbolic s-wave scattering length with g(t) = λ0 (1 + tanh [μt]). Plots (a) and (b) of the
top panels are obtained with a positive self-steepening coefficient α0 (α0 = 1), while (c) and (d) are
generated with a negative self-steepening coefficient α0 (α0 = −1). Other parameter used in plots
(a)–(d) are λ0 = −1.1, g0 = −0.95, and μ = 2, and m = 0.99. Other parameters are given in the
text. Reprint from Ref. [106], Copyright 2022, with permission from Springer

 
m! cos [! t] m! 2 m 2 + cos2 [! t] + 2 sin [! t]
γ (t) = − and α(t) = − .
4 1 + m sin [! t] 8 (1 + m sin [! t])2

For the present example, we show respectively in Figs. 7.22 and 7.23 the dynamics
of a double-kink soliton and a bright femtosecond soliton with the corresponding
chirping. In these two constants, plots of the top [bottom] panels are generated with
positive [negative] self-steepening coefficient α0 . These two constants show chirped
solitons moving along an oscillatory trajectory in a direction that depends on the
sign of α0 . Independently of the sign of parameter α0 , both the double-kink soliton
and the corresponding chirping shown in Fig. 7.22 have double-kink feature. Plots of
Fig. 7.23 reveal that chirping associated with a bright soliton may have, depending
on the sign of parameter α0 , either the dark or the bright soliton feature.
BECs with a Time-Varying Hyperbolic S-Wave Scattering Length
To apply our results to BEC systems with a time-varying hyperbolic s-wave scattering
length, we follow Xue [120] and assume the parameter g(t) of the two-body inter-
atomic interactions to have the form
282 7 Dynamics of One-Dimensional Condensates …

Fig. 7.26 a, b Evolution plots of respectively the double-kink soliton given by Eq. (7.137) and the
corresponding chirp defined by Eq. (7.138) for δ1 = 1, υ = g0 +0.56137
2
α0 , = − 3+2υ
4 , and β0 =
α02 +0.62 218
α0 with α0 = 0.5, λ0 = −0.3, g0 = −1.8, m = 0.8, and μ = −0.1. c, d The evolution
plots of a bright femtosecond soliton defined by Eq. (7.141) and the corresponding frequency chirp
given by Eq. (7.142), respectively for α0 = 1, λ0 = 1.1, g0 = 0.95, m = 0.8, μ = −0.1, β0 = α10 ,
3α02 +2
υ = − α10 , and = − 4α02
. Other parameters are showed in the text. Reprint from Ref. [106],
Copyright 2022, with permission from Springer

g(t) = λ0 (1 + m tanh [μt]) ,

where λ0 μ = 0 and 0 < m < 1. It then follows from the integrable conditions
(7.110b) and (7.110c) that

μ 1
γ (t) = − and
4 (1 + tanh [μt]) cosh2 [μt]
μ2 3 + 4 cosh [μt] sinh [μt] (1 + tanh [μt])
α(t) = − .
8 cosh4 [μt] (1 + tanh [μt])2
References 283

With these functional parameters, we report respectively in Figs. 7.24 and 7.25 the
spatiotemporal evolution of the double-kink soliton and the femtosecond soliton. To
generate different plots of these two figures, we have used the following set of data

1 1 3α 2 + 2
α02 = 1, g0 > −1, β0 = , υ = − , and = − 0 2 .
α0 α0 4α0

Plots of the top (bottom) panels are obtained with positive (negative) self-steepening
coefficient α0 . The same results as in the previous example are obtained here: the
sign of the self-steepening coefficient α0 plays an important role in the direction of
the wave motion and the feature of the frequency chirp. It is sure that the sign of
parameter μ may also affect the wave width. For a better understanding, we display
in the top (bottom) panels of Fig. 7.26 the evolution plot of the double-kink soliton
(femtosecond soliton) and their corresponding chirp for a negative μ.

7.3.4 Conclusion and Discussions

In this last Section of this chapter, we have considered a cubic-quintic GP equation


with distributed coefficients to study the dynamics of matter-wave chirped solitons in
BECs with both two- and three-body interactions trapping in a time-varying complex
potential consisting of parabolic and complex terms. The integrable and the com-
pensation conditions are presented. Under these conditions, we have built chirped
solitonlike solutions of the GP equation under consideration. These solutions are used
to investigate the dynamics of chirped ultrashort pulses and chirped double-kink soli-
tons in various types of BECs. As the main results, we have obtained that chirping
associated to different solitonlike pulses consist of one linear part which depends
on both the parameter γ of the gain/loss of atoms and parameter g of the two-body
interactions, and one nonlinear part which is proportional to the wave density and
depend on the nonlinearity parameter g. The effects of the self-steepening coefficient
α0 on the BEC solitons and the corresponding chirping are also investigated.

References

1. Y.V. Kartashov, B.A. Malomed, L. Torner, Solitons in nonlinear lattices. Rev. Mod. Phys. 83,
247–306 (2011)
2. H.W. Schürmann, S.V. Serov, Traveling wave solutions of a generalized modified Kadomtsev-
Petviashvili equation. J. Math. Phys. 45, 2181 (2004)
3. E. Kengne, W.M. Liu, Exact solutions of the derivative nonlinear Schrödinger equation for a
nonlinear transmission line. Phys. Rev. E 73, 026603 (2006)
4. V.R. Kumar, R. Radha, M. Wadati, Phase engineering and solitons of Bose-Einstein conden-
sates with two- and three-body interactions. J. Phys. Soc. Jpn. 79, 074005 (2010)
284 7 Dynamics of One-Dimensional Condensates …

5. E. Kengne, A. Lakhssassi, W.M. Liu, R. Vaillancourt, Phase engineering, modulational insta-


bility, and solitons of Gross-Pitaevskii-type equations in 1+1 dimensions. Phys. Rev. E 87,
022914 (2013)
6. W.M. Liu, E. Kengne, Schrödinger Equations in Nonlinear Systems (Springer Nature Singa-
pore Pte Ltd., 2019)
7. K. Mio, T. Ogino, K. Minami, S. Takeda, A Perturbation method and its application to
obliquely propagating nonlinear Alfvén wave. J. Phys. Soc. Jpn. 41, 2114–2120 (1976)
8. W. Schopff, W. Zimmermann, Convection in binary fluids: amplitude equations, codimension-
2 bifurcation, and thermal fluctuations. Phys. Rev. E 47, 1739 (1993)
9. M.S. Ruderman, Propagation of solitons of the derivative nonlinear Schrödinger equation in
a plasma with fluctuating density. Phys. Plasmas 9, 2940 (2002)
10. A.M. Kamchatnov, New approach to periodic solutions of integrable equations and nonlinear
theory of modulational instability 286, 199–270 (1997)
11. N. Tzoar, M. Jain, Self-phase modulation in long-geometry optical waveguides. Phys. Rev.
A 23, 1266 (1981)
12. D. Anderson, M. Lisak, Nonlinear asymmetric self-phase modulation and self-steepening of
pulses in long optical waveguides. Phys. Rev. A 27, 1393 (1983)
13. G. Yang, Y.R. Shen, Spectral broadening of ultrashort pulses in a nonlinear medium. Opt.
Lett. 9, 510–512 (1984)
14. G.P. Agrawal, Nonlinear Fiber Optics (Academic Press, San Diego, 1995)
15. D.J. Kaup, A.C. Newell, Exact solution for a derivative non-linear Schrödinger equation. J.
Math. Phys. 19, 798–801 (1978)
16. R. Radhakrishnan, A. Kundu, M. Lakshmanan, Coupled nonlinear Schrödinger equations
with cubic-quintic nonlinearity: integrability and soliton interaction in non-Kerr media. Phys.
Rev. E 60, 3314 (1999)
17. K.E. Strecker, G.B. Partridge, A.G. Truscott, R.G. Hulet, Bright matter wave solitons in
Bose-Einstein condensates. New J. Phys. 5, 73 (2003)
18. S.-J. Wang, C.-L. Jia, D. Zhao, H.-G. Luo, J.-H. An, Dark and bright solitons in a quasi-one-
dimensional Bose-Einstein condensate. Phys. Rev. A 68, 015601 (2003)
19. S.L. Cornish, T.T. Sarah, C.E. Wieman, Formation of bright matter-wave solitons during the
collapse of attractive Bose-Einstein condensates. Phys. Rev. Lett. 96, 170401 (2006)
20. FKh. Abdullaev, J.G. Caputo, R.A. Kraenkel, B.A. Malomed, Controlling collapse in Bose-
Einstein condensation by temporal modulation of the scattering length. Phys. Rev. A 67(2003),
013605 (2003)
21. P.G. Kevrekidis, G. Theocharis, D.J. Frantzeskakis, B.A. Malomed, Feshbach resonance man-
agement for Bose-Einstein condensates. Phys. Rev. Lett. 90, 230401 (2003)
22. A.V. Carpentier, H. Michinel, M.I. Rodas-Verde, V.M. P érez-García, Analysis of an atom
laser based on the spatial control of the scattering length. Phys. Rev. A 74, 013619 (2006)
23. FKh. Abdullaev, Yu.V. Bludov, S.V. Dmitriev, P.G. Kevrekidis, V.V. Konotop, Generalized
neighbor-interaction models induced by nonlinear lattices. Phys. Rev. E 77, 016604 (2008)
24. H. Sakaguchi, B.A. Malomed, Solitons in combined linear and nonlinear lattice potentials.
Phys. Rev. A 81, 013624 (2010)
25. G. Theocharis, Z. Rapti, P.G. Kevrekidis, D.J. Frantzeskakis, V.V. Konotop, Modulational
instability of Gross-Pitaevskii-type equations in 1+1 dimensions. Phys. Rev. A 67, 063610
(2003)
26. J.H.V. Nguyen, D. Luo, R.G. Hulet, Formation of matter-wave soliton trains by modulational
instability. Science 356, 422–426 (2017)
27. N. Meyer, H. Proud, M. Perea-Ortiz, C. O’Neale, M. Baumert, M. Holynski, J. Kronjager,
G. Barontini, K. Bongs, Observation of two-dimensional localized Jones-Roberts solitons in
Bose-Einstein condensates. Phys. Rev. Lett. 119, 150403 (2017)
28. T.B. Benjamin, J.E. Feir, The disintegration of wave trains on deep water. J. Fluid Mech. 27,
417–430 (1967)
29. E. Kengne, A. Lakhssassi, R. Vaillancourt, W.M. Liu, Phase engineering, modulational insta-
bility, and solitons of Gross-Pitaevskii-type equations in 1+1 dimensions. J. Math. Phys. 54,
051501 (2013)
References 285

30. V. Talanov, Focusing of light in cubic media. J. Exp. Theor. Phys. Lett. 11, 199–201 (1970)
31. G. Fibich, The Nonlinear Schrödinger Equation: Singular Solutions and Optical Collapse
(Springer, Heidelberg, 2015)
32. G. Fibich, G. Papanicolaou, Self-focusing in the perturbed and unperturbed nonlinear
Schrödinger equation in critical dimension. SIAM J. Appl. Math. 60, 183–240 (1999)
33. E.T. Whittaker, G.N. Watson, A Course of Modern Analysis (Cambridge University Press,
Cambridge, 1927)
34. I.S. Gradshteyn, I.M. Ryzhik, Table of Integrals, Series, and Products, 7th edn. (Academic
Press, San Diego, 2007)
35. H.W. Schurmann, Traveling-wave solutions of the cubic-quintic nonlinear Schrödinger equa-
tion. Phys. Rev. E 54, 4312 (1996)
36. H.W. Schürmann, V.S. Serov, Traveling wave solutions of a generalized modified Kadomtsev-
Petviashvili equation. J. Math. Phys. 45, 2181 (2004)
37. V.I. Kruglov, A.C. Peacock, J.D. Harvey, Phys. Rev. Lett. 90, 113902 (2003)
38. V.N. Serkin, T.L. Belyaeva, JETP Lett. 74, 573 (2001)
39. A. Hasegawa, Phys. D (Amsterdam) 123, 267 (1998)
40. J. Li, K. Sun, X. Chen, Shortcut to adiabatic control of soliton matter waves by tunable
interaction. Sci. Rep. 6, 38258 (2016). https://doi.org/10.1038/srep38258
41. V.A. Bogatyrev, M.M. Bubnov, E.M. Dianov, A.S. Kurkov, P.V. Mamyshev, A.M. Prokhorov,
S.D. Rumyantsev, V.A. Semenov, S.L. Semenov, A.A. Sysoliatin, S.V. Chernikov, A.N.
Gur’yanov, G.G. Devyatykh, S.I. Miroshnichenko, J. Lightwave Technol. 9, 561 (1991)
42. B.-Q. Li, Y.-L. Ma, Extended generalized Darboux transformation to hybrid rogue wave and
breather solutions for a nonlinear Schrödinger equation. Appl. Math. Comput. 386, 125469
(2020)
43. E. Kengne, Rogue waves of the dissipative Gross-Pitaevskii equation with distributed coeffi-
cients. Eur. Phys. J. Plus 135, 622 (2020)
44. E. Kengne, W.M. Liu, Engineering rogue waves with quintic nonlinearity and nonlinear dis-
persion effects in a modified Nogochi nonlinear electric transmission network. Phys. Rev. E
102, 012203 (2020)
45. W. Guan, B. Li, Mixed structures of optical breather and rogue wave for a variable coefficient
inhomogeneous fiber system. Opt. Quant. Electron. 51, 352 (2019)
46. Y.-L. Ma, Interaction and energy transition between the breather and rogue wave for a gen-
eralized nonlinear Schrödinger system with two higher-order dispersion operators in optical
fibers. Nonlinear Dyn. 97, 95 (2019)
47. Y.-L. Ma, B.-Q. Li, Doubly periodic waves, bright and dark solitons for a coupled monomode
step-index optical fiber system. Opt. Quantum Electron. 50, 443 (2018)
48. FKh. Abdullaev, R.M. Galimzyanov, Kh.N. Ismatullaev, Collective excitations of a BEC under
an harmonic trap position jittering. J. Phys. B 41, 015301 (2008)
49. A. Gammal, T. Frederico, L. Tomio, Ph. Chomaz, Atomic Bose-Einstein condensation with
three-body interactions and collective excitations. J. Phys. B 33, 4053 (2000)
50. FKh. Abdullaev, R. Galimzyanov, The dynamics of bright matter wave solitons in a quasi
one-dimensional Bose-Einstein condensate with a rapidly varying trap. J. Phys. B 36, 1099
(2003)
51. G.S. Chong, W.H. Hai, Q.T. Xie, Exact wave solutions for Bose-Einstein condensates with
time-dependent scattering length and spatiotemporal complicated potential. Chin. Phys. Lett.
20, 2098 (2003)
52. F. Dalfovo, S. Giorgini, L.P. Pitaevskii, S. Stringari, Theory of Bose-Einstein condensation
in trapped gases. Rev. Mod. Phys. 71, 463 (1999)
53. E. Kengne, P.K. Talla, Dynamics of bright matter wave solitons in Bose-Einstein condensates
in an expulsive parabolic and complex potential. J. Phys. B 39, 3679 (2006)
54. L.-C. Zhao, Z.-Y. Yang, T. Zhang, K.-J. Shi, Dynamics of bright solitons in Bose-Einstein
condensates with complicated potential. Chin. Phys. Lett. 26, 120301 (2009)
55. Y.-L. Ma, Abundant excited optical breathers for a nonlinear Schrödinger equation with vari-
able dispersion and nonlinearity terms in inhomogenous fiber optics. Optik 201, 162821
(2020)
286 7 Dynamics of One-Dimensional Condensates …

56. B.-Q. Li, Y.-L. Ma, Excitation of optical rogue waves to a (2+ 1)-dimensional nonlinear
Schrödinger equation in nonlocal optical fibers. Optik 174, 178 (2018)
57. B.-Q. Li, J.-Z. Sun, Y.-L. Ma, Soliton excitation for a coherently coupled nonlinear
Schrödinger system in optical fibers with two orthogonally polarized components. Optik
175, 275 (2018)
58. N. Song, Y. Xue, Rogue waves of nonlinear Schrö dinger equation with time-dependent linear
potential function. Discrete Dyn. Nat. Soc. 2016, 7879517 (2016)
59. C. Sulem, P.L. Sulem, The Nonlinear Schrödinger Equation (Springer, Berlin, 1999)
60. G. Theocharis, Z. Rapti, P.G. Kevrekidis, D.J. Frantzeskakis, V.V. Konotop, Modulational
instability of Gross-Pitaevskii-type equations in 1 + 1 dimensions. Phys. Rev. A 67, 063610
(2003)
61. E. Kengne, W.M. Liu, Modulational instability and soliton control in a cubic-quintic dissipa-
tive Gross-Pitaevskii equation with distributed coefficients. J. Phys. B: At. Mol. Opt. Phys.
53, 215003 (2020)
62. K. Weierstrass, Mathematische Werke V (Johnson, New York, 1915)
63. H.W. Schürmann, Traveling-wave solutions of the cubic-quintic nonlinear Schrödinger equa-
tion. Phys. Rev. E 54, 4312 (1996)
64. L. Khaykovich, F. Schreck, G. Ferrari1, T. Bourdel, J. Cubizolles, L.D. Carr, Y. Castin, C.
Salomon, Formation of a matter-wave bright soliton. Science 296, 1290 (2002)
65. J.S. Baris, P.J. Baris, B. Ruchlewicz, On blow-up solutions of nonautonomous quadratic
differential systems. Differ. Equ. 42, 320 (2006)
66. G.S. Chong, W.H. Hai, T.Q. Xie, Stabilizations of two-dimensional trapped and untrapped
matter waves via a Feshbach resonance technique. Chin. Phys. Lett. 20, 2098 (2003)
67. H. Saito, M. Ueda, Dynamically stabilized bright solitons in a two-dimensional Bose-Einstein
condensate. Phys. Rev. Lett. 90, 040403 (2003)
68. D.S. Wang, X.-F. Zhang, P. Zhang, W.M. Liu, Matter-wave solitons of Bose-Einstein conden-
sates in a time-dependent complicated potential. J. Phys. B 42, 245303 (2009)
69. Lu. Ruiyu Hao, Zhonghao Li Li, Rongcao Yang, Guosheng Zhou, Optics Communications
245, 383 (2005)
70. Etienne Wamba, Timoléon C. Kofané, Alidou Mohamadou, Chin. Phys. B 21, 070504 (2012)
71. L.P. Pitaevskii, Vortex lines in an imperfect Bose gas. Sov. Phys. JETP 13, 451 (1961)
72. E.P. Gross, Structure of a quantized vortex in Boson systems. Nuovo Cimento 20, 454 (1961)
73. L.P. Pitaevskii, S. Stringari, Bose-Einstein Condensation (Oxford University Press, Oxford,
2003)
74. W.M. Liu, E. Kengne, Schrödinger Equations in Nonlinear System (Springer Nature: 1st ed.
2019 edition (29 Mar 2019))
75. A. Gammal, T. Frederico, L. Tomio, P. Chomaz, Atomic Bose-Einstein condensation with
three-body interactions and collective excitations. J. Phys. B At. Mol. Opt. Phys. 33, 4053
(2000)
76. A.L. Fetter, J.D. Walecka, Quantum Theory of Many-Particle Systems (McGraw-Hill, New
York, 1971)
77. B.P. Anderson, P.C. Haljan, C.A. Regal, D.L. Feder, L.A. Collins, C.W. Clark, E.A. Cornell,
Watching Dark solitons decay into vortex rings in a Bose-Einstein condensate. Phys. Rev.
Lett. 86, 2926 (2001)
78. J. He, E.G. Charalampidis, P.G. Kevrekidis, D.J. Frantzeskakis, Application to Bose-Einstein
condensates, rogue waves in nonlinear Schrödinger models with variable coefficients. Phys.
Lett. A 378, 577–583 (2014)
79. Chuanzhong Li, Jingsong He, Darboux transformation and positions of the inhomogeneous
Hirota and the Maxwell-Bloch equation. Sci. China Phys. Mech. Astron. 57(5), 898–907
(2014)
80. J.S. He, Y.S. Tao, K. Porsezian, A.S. Fokas, On the rogue wave management in an inhomo-
geneous nonlinear fibre with higher order effects. J. Nonlinear Math. Phys. 20(3), 407–419
(2013)
References 287

81. Jingsong He, Youying Wang, Linjing Li, Non-rational rogue waves induced by inhomogeneity.
Chin. Phys. Lett. 29, 060509 (2012)
82. J. He, Y. Song, C.G.L. Tiofack, M. Taki, Rogue wave light bullets of the three-dimensional
inhomogeneous nonlinear Schrödinger equation. Photon. Res. 9, 643–648 (2021)
83. V.M. Pérez-García, V.V. Konotop, V.A. Brazhnyi, feshbach resonance induced shock waves
in Bose-Einstein condensates. Phys. Rev. Lett. 92, 220403 (2004)
84. Xiong Bo, Liu Xun-Xu, Internal Josephson-like tunnelling in two-component Bose-Einstein
condensates affected by sign of the atomic interaction and external trapping potential. Chin.
Phys. 16, 2578 (2007)
85. E. Kengne, A. Lakhssassi, W.M. Liu, Non-autonomous solitons in inhomogeneous nonlinear
media with distributed dispersion. Nonlinear Dyn. 97, 449–469 (2019)
86. G.P. Agrawal, Applications of Nonlinear Fiber Optics (Academic, San Diego, 2001)
87. M. Wittmann, M.T. Wick, O. Steinkellner, P. Farmanara, V. Stert, W. Radloff, G. Korn, I.V.
Hertel, Generation of femtosecond VUV pulses and their application to time resolved spec-
troscopy in the gas phase. Opt. Commun. 173, 323–331 (2000)
88. H. Abramczyk, Introduction to Laser Spectroscopy 2005, 219 (2005)
89. Y. Liu, J.-G. Zhang, D. Tang, Optical waveform monitoring based on a free-running mode-
locked femtosecond fibre laser and four-wave mixing in a highly nonlinear fibre. Opto-
Electron. Rev. 24, 62–74 (2016)
90. Sorokina I.T., Sorokin E., Carrig T, Femtosecond pulse generation from a SESAM mode-
locked cr:znse laser, in Conference on Lasers and Electro-Optics/Quantum Electronics and
Laser Science Conference and Photonic Applications Systems Technologies, Technical Digest
(CD) (Optical Society of America, 2006), paper CMQ2
91. A. Starka, L. Correiaa, M. Teichmanna, S. Salewskia, C. Larsenb, V.M. Baev, P.E. Toscheka,
Intracavity absorption spectroscopy with thulium-doped fibre laser. Opt. Commun. 215, 113–
123 (2003)
92. V.L. Kalashnikov, A. Chernykh, Spectral anomalies and stability of chirped-pulse oscillators.
Phys. Rev. A 75, 033820 (2007)
93. Y. Kodama, Optical solitons in a monomode fiber. Stat. J. Phys. 39, 597–614 (1985)
94. Y. Kodama, A. Hasegawa, Nonlinear pulse propagation in a monomode dielectric guide. IEEE
J. Quantum Electron. 23, 510–524 (1987)
95. V.M. Vyas, P. Patel, P.K. Panigrahi, C.N. Kumar, W. Greiner, Chirped chiral solitons in the
nonlinear Schrödinger equation with self-steepening and self-frequency shift. Phys. Rev. A
78, 021803(R) (2008)
96. A. Goyal Amit, R. Gupta, C.N. Kumar, T.S. Raju, Chirped femtosecond solitons and double-
kink solitons in the cubic-quintic nonlinear Schrödinger equation with self-steepening and
self-frequency shift. Phys. Rev. A 84, 063830 (2011)
97. V.I. Kruglov, A.C. Peacock, J.D. Harvey, Exact self-similar solutions of the generalized non-
linear Schrödinger equation with distributed coefficients. Phys. Rev. Lett. 90, 113902 (2003)
98. A. Bouzida, H. Triki, M.Z. Ullah, Q. Zhouc, A. Biswas, M. Belic, Chirped optical solitons in
nano optical fibers with dual-power law nonlinearity. Optik 142, 77–81 (2017)
99. B. Younas, M. Younis, Chirped solitons in optical monomode fibres modelled with Chen-
Lee-Liu equation. Pramana J. Phys. 94, 3 (2020)
100. H. Triki, K. Porsezian, A. Choudhuri, T.P. Dinda, Chirped solitary pulses for a nonic nonlinear
Schrödinger equation on a continuous-wave background. Phys. Rev. A 93, 063810 (2016)
101. Kengne Emmanuel, Engineering chirped LambertW-kink signals in a nonlinear electrical
transmission network with dissipative elements. Eur. Phys. J. Plus 136, 266 (2021)
102. Keke Shi, Chuang Liu, James D. Biggs, Zhaowei Sun, Xiaokui Yue, Observer-based control
for spacecraft electromagnetic docking. Aerosp. Sci. Technol. 99, 105759 (2020)
103. Chuang Liu, George Vukovich, Zhaowei Sun, Keke Shi, Observer-based fault-tolerant attitude
control for spacecraft with input dela. J. Guid. Control. Dyn. 41, 2039 (2018)
104. Chuang Liu, Xiaokui Yue, Keke Shi, Zhaowei Sun, Inertia-free attitude stabilization for flex-
ible spacecraft with active vibration suppression. Int. J. Robust Nonlinear Control 29, 6311–
6336 (2019)
288 7 Dynamics of One-Dimensional Condensates …

105. Chuang Liu, Keke Shi, Xiaokui Yue, Zhaowei Sun, Inertia-free saturated output feedback
attitude stabilization for uncertain spacecraft. Int. J. Robust Nonlinear Control 30, 5101–5121
(2020)
106. E. Kengne, A. Lakhssassi, Compensation process and generation of chirped femtosecond
solitons and double-kink solitons in Bose-Einstein condensates with time-dependent atomic
scattering length in a time-varying complex potential. Nonlinear Dyn. 104, 4221 (2021)
107. V. Ramesh Kumar, R. Radha, M. Wadati, Phase engineering and solitons of Bose–Einstein
condensates with two- and three-Body interactions. J. Phys. Soc. Jpn. 79, 074005 (2010)
108. P.A. Clarkson, C.M. Cosgrove, Painleve analysis of the non-linear Schrödinger family of
equations. J. Phys. A: Math. Gen. 20, 2003 (1987)
109. S. Chen, Y. Zhou, L. Bu, F. Baronio, M. Soto-Crespo Jose, D. Mihalache, Super chirped rogue
waves in optical fibers. Opt. Express 27, 11370–11384 (2019)
110. H.M. Yin, B. Tian, C.R. Zhang, X.X. Du, X.C. Zhao, Optical breathers and rogue waves via
the modulation instability for a higher-order generalized nonlinear Schrödinger equation in
an optical fiber transmission system. Nonlinear Dyn. 97, 843–852 (2019)
111. C.Q. Song, H.Q. Zhao, Dynamics of various waves in nonlinear Schrödinger equation with
stimulated Raman scattering and quintic nonlinearity. Nonlinear Dyn. 99, 2971–2985 (2020)
112. G.P. Agrawal, Nonlinear Fiber Optics, 4th ed. (Academic, 2007)
113. H.H. Chen, Y.C. Lee, C.S. Liu, Integrability of nonlinear Hamiltonian systems by inverse
scattering method. Phys. Scr. 20, 490 (1979)
114. D.J. Kaup, A.C. Newell, An exact solution for a derivative nonlinear Schrödinger equation.
J. Math. Phys. 19, 798 (1978)
115. V.S. Gerdjikov, M.I. Ivanov, Hierarchies of Hamiltonian structures. Bul. J. Phys. 10, 130–143
(1983)
116. Z.X. Liang, Z.D. Zhang, W.M. Liu, Dynamics of a bright soliton in Bose-Einstein condensates
with time-dependent atomic scattering length in an expulsive parabolic potential. Phys. Rev.
Lett. 94, 050402 (2005)
117. E.T. Whittaker, G.N. Watson, A Course of Modern Analysis (Cambridge University Press,
Cambridge, 1927), pp.452–455
118. L. Khaykovich, F. Schreck, G. Ferrari, T. Bourdel, J. Cubizolles, L.D. Carr, Y. Castin, C.
Salomon, Formation of a matter-wave bright soliton. Science 296, 1290 (2002)
119. G.S. Chong, W.H. Hai, Q.T. Xie, Breathing bright solitons in a Bose-Einstein condensate.
Chin. Phys. Lett. 20, 2098 (2003)
120. J.-K. Xue, Controllable compression of bright soliton matter waves. J. Phys. B: At. Mol. Opt.
Phys. 38, 3841 (2005)
Chapter 8
Rogue Matter Waves in Bose-Einstein
Condensates Trapped in Time-Varying
External Potentials

Abstract Using the mapping techniques, we investigate analytically the dynamics


of rogue matter waves in Bose-Einstein condensates trapped in time-varying exter-
nal potentials. Rational solutions of the model equations are presented and used to
engineer first-order and second-order matter rogue waves in the BEC contexts when
the gain/loss of condensate atoms is taken into consideration.

8.1 Non-autonomous Rogue Matter Waves in Bose-Einstein


Condensates With Dissipative External Potential

A quasi-one-dimensional dissipative GP equation that describes the wave-function


of Bose-Einstein condensates trapped in time-varying external potentials under the
loss/gain of atoms is considered.The ansatz method is used to present the integrable
conditions. Using the similarity transformation technique, first-order and second-
order rational solutions are obtained and used to engineer first- and second-order
dissipative non-autonomous matter rogue waves in the BEC contexts. The control
of the scattering length, the external harmonic, and the linear parts of the potentials
are used to manage the motion and the background of dissipative non-autonomous
rogue matter waves. The effects of the two-body inter-atomic interactions as well as
those of the potential parameters on the motion of the BECs non-autonomous matter
waves are presented.

8.1.1 Introduction

Since the first experimental observation of Bose-Einstein condensates [1–4], a great


interest is paid to the subject of nonlinear excitations in cold atomic systems [5–
8]. To the best of our knowledge, only a few works have been reported on non-
autonomous/dissipative rogue matter waves in the context of BECs [9–11]. In this
Section, we develop mathematical models for generating non-autonomous dissipa-
tive rogue matter waves of BECs with complicated potentials. These mathematical

© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2022 289
E. Kengne and W. Liu, Nonlinear Waves,
https://doi.org/10.1007/978-981-19-6744-3_8
290 8 Rogue Matter Waves in Bose-Einstein Condensates …

models are then employed to build first- and second-order rational solutions of a dis-
sipative GP equation which governs the dynamics of BECs trapped in time-varying
external potentials when the loss/gain of condensate atoms is considered.

8.1.2 Model and Analytical Exact First- and Second-Order


Rational Solutions of the Gross-Pitaevskii Equation

Description of the Model The mathematical model to be studied in this Section is


the below dissipative quasi-one-dimensional distributed NLS equation with external
trap potential (alias Gross-Pitaevskii equation) which can be used to describe the
properties of weakly interacting bosonic gases trapped in a potential [12–14]

∂ψ 1 ∂ 2ψ  
i + + g(t) |ψ|2 ψ + k(t)x 2 + λ(t)x + iγ (t) ψ = 0, (8.1)
∂t 2 ∂x 2

where g(t), k(t), λ(t), and γ (t) are all real functions of time t. Function V (x, t)
defined as
V (x, t) = k(t)x 2 + λ(t)x + iγ (t) (8.2)

is the most physically relevant example of an external potential in the BEC context;
in this context, Eq. (8.1) describes the macroscopic wave function ψ(x, t) of the
condensate. In the BEC context, the nonlinearity parameter g(t) and parameters
k(t), λ(t), and γ (t) have the same physical meanings as in Chap. 7 [15–19]. Also,
the temporal and the spatial coordinates t and x are measured as described in Chap.
7. k(t)x 2 , λ(t)x, and iγ (t) are respectively the harmonic, linear, and dissipative
parts of potential (8.2). Throughout this Section, we assume that increasing the BEC
number of atoms will not lead to the collapse of the system [18–21]. Henceforth, Eq.
(8.1) will be referred to as the model equation.
First- and Second-Order Rational Solutions of the Model Equation For extracting
rational solutions of the model equation (8.1), we look for solutions of the form

ψ(x, t) = [A(t) + B(t)ψ1 (X, T ) + iC(t)ψ2 (X, T )] exp [iχ (x, t) + iλ0 T (t)] .
(8.3)
In Eq. (8.3), A, B, C, T , χ , ψ1 , ψ2 , and X = X (x, t) are real real functions and λ0
is a real control parameter. Next, we impose to X, χ , A, B, C to satisfy the following
conditions

X (x, t) = α(t)x + β(t), (8.4a)


1 dα 2 1 dβ
χ (x, t) = − x − x + χ0 (t), (8.4b)
2α dt α dt  
 
(A(t), B(t), C(t)) = (A0 , B0 , C0 ) |α| exp − γ (t)dt , (8.4c)

dT
− T0 α 2 (t) = 0, T0 = 0, (8.4d)
dt
8.1 Non-autonomous Rogue Matter Waves in Bose-Einstein… 291

where A0 , B0 , C0 , and T0 are real constants, and α, β, χ0 are solutions of the


differential system

d 2α 2 dα 2
− + 2kα = 0, (8.5)
dt 2 α dt
d 2β 2 dα dβ
− + λα = 0, (8.6)
dt 2 α dt dt

dχ0 1 1 dβ 2
+ = 0, (8.7)
dt 2 α dt

A0 , B0 , C0 , and T0 being real constants. Under the above conditions, we insert ansatz
(8.3) into the model equation (8.1) and obtain, under the restriction
  
g02 α 2 − g 2 exp −4 γ (t)dt = 0, g0 = 0 : g0 g(t) > 0, (8.8)
 2 
d2g dg dγ dg
g −2 + 2g 2 k − 2g g − 2γ + 2gγ 2 = 0, (8.9)
dt 2 dt dt dt

the following system of partial differential equations in ψ1 and ψ2 :


∂ψ2 B0 ∂ 2 ψ1

C0 T0 − + λ 0 T0 (A 0 + B0 ψ1 ) − g0 (A 0 + B0 ψ1 ) (A 0 + B0 ψ1 ) 2
+ C 0 ψ
2 2
2 = 0,
∂T 2 ∂ X2
(8.10a)
∂ψ1 C0 ∂ 2 ψ2

B0 T0 + − C0 λ0 T0 ψ2 + C0 g0 ψ2 (A0 + B0 ψ1 )2 + C02 ψ22 = 0, (8.10b)
∂T 2 ∂ X2
In Eq. (8.8), g0 is a free real constant to be taken from the condition g0 g(t) > 0
for all t. Throughout this Section, the condition (8.9) which links the parameter g of
the two-body inter-atomic integrations with the potential parameters k and γ will be
referred to as the “integrable condition” of the model equation (8.1).
Seeking the solution (ψ1 , ψ2 )T of system (8.10a)–(8.10b) in the form [22, 23],
T
4 8T
(ψ1 , ψ2 ) = −
T
, −  (8.11)
B0 1 + 2X 2 + 4T 2 C0 1 + 2X 2 + 4T 2

yields the following first-order rational solution of the model equation (8.1)
   
 4 + i8T
ψ(x, t) = 2g(t) exp −2 γ (t)dt 1 − exp [iχ + i T ] .
1 + 2 [α(t)x + β(t)]2 + 4T 2
(8.12)
It is important to notice that solution (8.11) of system (8.10a )–(8.10b) corresponds
to A0 = λ0 = 1, g0 = 21 , and T0 = 21 so that the first-order rational solution (8.12) of
the model equation (8.1) is valid for 0 < g0 g(t) = 21 g(t), for BECs with attractive
inter-atomic interactions.
To build the second-order rational solution of system (8.10a)–(8.10b), we consider,
for simplicity, the special case when A0 = λ0 = 1, g0 = 21 , and T0 = 21 . In this
292 8 Rogue Matter Waves in Bose-Einstein Condensates …

special case, we have g0 g(t) = 21 g(t) so that the corresponding second-order rational
solution of the model equation will be associated with BECs with attractive inter-
atomic interactions. Employing the direct method developed in Ref. [23], second-
order rational solution (ψ1 , ψ2 )T of system (8.10a)–(8.10b) has the form
 T
ψ11 (X, T ) ψ22 (X, T )
(ψ1 , ψ2 ) =
T
, , (8.13)
ψ12 (X, T ) ψ21 (X, T )

where ψ11 , ψi j , and ψ22 (X, T ) are given by


      2 
1
ψ11 (X, T ) = 3 1 − 4X 2 + 8T 2 8T 2 − 9 − 4 X 2 + 6T 2 ,
8B0
   2
3 3 22T 2 + 3X 2 6T 2 − X 2 8T 6 + 12T 6 X 2 + 6T 2 X 4 + X 6
ψi j (X, T ) = + + + , (8.14)
32 16 8 12

  2 
1
ψ22 (X, T ) = 15 − 8T 2 + 12X 2 − 4T 2 + 2X 2 .
4C0

The corresponding second-order rational solution of the model equation (8.1) reads

ψ(x, t)
⎡   2  2  ⎤
3 1 − 4X 2 + 8T 2 8T 2 − 9 − 4 X 2 + 6T 2 + 2i 15 − 8T 2 + 12X 2 − 4T 2 + 2X 2
= ⎣1 + ⎦
8ψ12 (X, T )
  

× 2g exp −2 γ (t)dt exp [iχ(x, t) + i T (t)] . (8.15)

8.1.3 Management of Dissipative Rogue Matter Waves


in Bose-Einstein Condensates with Complicated
Potential

Here, we show how to employ parameter g(t) of the two-body inter-atomic interac-
tions and the potential parameters k(t), λ(t) and γ (t) to control and manage the BECs
rogue matter waves [24, 25]. First, we note that, the amplitude of non-autonomous
rogue matter waves corresponding to the first- and the second-order solutions (8.12)
and (8.15) of the model equation (8.1) is proportional to
  

p(t) = 2g(t) exp −2 γ (t)dt ,
8.1 Non-autonomous Rogue Matter Waves in Bose-Einstein… 293

while their width is inversely proportional to α(t). Moreover, the wave centre is
found to be ξ(t) = −β(t)/α(t) and satisfies the equation

d 2ξ
− 2kξ − λ = 0, (8.16)
dt 2
which depends on both k and λ. This means that the we can manipulate the motion
of rogue matter waves by controlling both the external harmonic and linear trap-
ping potentials. Based on physical examples, we demonstrate in what follows the
engineering of rogue matter waves in one-component BECswith different kinds of
scattering length and complex trapping potential. Without loss of generality, we focus
our attention to only the case of positive α(t).
Dissipative Rogue Matter Waves in BEC System with Time-Independent Har-
monic Confining Potential Our first example is the BEC trapped in the time-
independent harmonic potential similar to the one used in Ref. [21]. We thus consider
a time-independent harmonic potential with strength k(t) = −2κ 2 , where κ  0.05.
Following Wu et al. [26], we employ for this example g(t) = exp [±2κt]. Asking
that the integrable condition (8.9) occurs yields

1√ 1√
γ (t) = γ− = κ − 2κ and γ (t) = γ+ = κ + 2κ. (8.17)
2 2
Solving now system (8.5)–(8.8), we find

√ 
(α, β, T )T = 2 exp 2κ − 4κ t , β01

√  ⎞T

√  1 − exp 2 2κ − 4κ t
+β02 exp 2 2κ − 4κ t , √ ⎠ , (8.18)
4κ − 2κ


(α, β, T )T = 2 exp − 2κt , β01

√ ⎞T

√ 1 − exp −2 2κt
+β02 exp −2 2κt , √ ⎠ , (8.19)

for respectively γ (t) = γ− and γ (t) = γ+ , where β01 and β02 are free constants to
be taken so that |β(t)| > 0.
The typical evolution of dynamics of non-autonomous dissipative first- and
second-order rogue matter waves for BECs with time-independent harmonic confin-
ing potential described above are showed in Fig. 8.1. Plots of this figure are generated
with γ (t) = γ+ , which correspond to g(t) = exp [2κt] , which is an increasing func-
tion of time. Plots of Fig. 8.1 show the first- and second-order rogue matter waves
embedded on an increasing CW background whose peaks are reached at time t  0.
It is also seen from plots of Fig. 8.1 that the first- and second-order rogue matter
294 8 Rogue Matter Waves in Bose-Einstein Condensates …

Fig. 8.1 Evolution of the atomic density left panels) and density plot (right panels) of the first-order
(top panels) and second-order (bottom panels) rogue matter waves associated respectively to the
rational solutions (8.12) and (8.15) of the GP equation (8.1). Different plots are generated for the
gain parameter γ (t) = γ+ with solution data (8.18) for β01 = 1.0 and β02 = 0.1. Reprint from Ref.
[9], Copyright 2022, with permission from Springer Nature

waves can concentrate the energy of the BEC system with gain of atoms in a small
region. We conclude from plots of Fig. 8.1 that with the increasing (decreasing) g(t),
the rogue matter wave is embedded on an increase (decrease) CW background.
Dissipative Rogue Matter Waves in BEC System with a Temporal Periodic Mod-
ulation of the S-Wave Scattering Length Following the idea developed in works
[27] and [28], we consider as our second example, a BEC system whose parameter
g of the two-body inter-atomic interaction has the form [29]

g(t) = 1 + m sin [ωt] , 0 < m < 1,

for the constant ω. For simplicity, we set λ(t) = 0 and γ (t) = γ = constant [17].
Using the integrable condition (8.9), we obtain
  2
1 dg 1 dg 1 d2g
k(t) = 2γ γ− + − . (8.20)
g dt g2 dt 2g dt 2

Going back to system (8.5)–(8.8) with the above k, g, and k, we find


8.1 Non-autonomous Rogue Matter Waves in Bose-Einstein… 295

 
α(t) = 2 (1 + m sin [ωt]) exp −2γ t ,


2 + m2 4m (ω cos [ωt] + 4γ sin [ωt])
β(t) = β01 − 2β02 +
4γ ω2 + 16γ 2


m 2 (ω sin [2ωt] − 2γ cos [2ωt])  
+  exp −4γ t , (8.21)
2 ω2 + 4γ 2


2 + m2 4m (ω cos [ωt] + 4γ sin [ωt]) m 2 (ω sin [2ωt] − 2γ cos [2ωt])
T (t) = T10 − + + 
4γ ω2 + 16γ 2 2 ω2 + 4γ 2
 
exp −4γ t ,

where, β01 and β02 are real constants to be taken from the condition |β(t)| > 0, and

2 + m2 4mω 2m 2 γ
T10 = + 2 − 2 ;
4γ ω + 16γ 2 2 ω + 4γ 2

this expression for T10 yields T (0) = 0.


With the above functional parameters with positive γ , we show in Fig. 8.2 the
evolution of the dissipative non-autonomous first-order (top panels) and second-order
(bottom panels) rogue matter waves form BECs with a temporal periodic modulation
of the s-wave scattering length. Plots of this figure show the emergence of two distinct
neighbor rogue waves with different amplitude. Also, we can see from Fig. 8.2a, c that
the waves are embedded on an oscillating decreasing nonzero background. Negative
γ will certainly lead to waves embedded on an oscillating increasing background.
Dissipative Rogue Matter Wave in BEC Trapped in a Temporal Periodic Modu-
lated Linear Complex Potential Our third example is a BEC trapped in a temporal
periodic modulated linear potential when the gain/loss of atoms is taken into consid-
eration. The magnitude of the linear potential is assumed to have the form [30–32]

λ(t) = σ + m cos [ωt] , (8.22)

where σ , m, and ω are real parameters.


Case of BECs with a time-independent s-wave scattering length In this case,
g(t) = g is a constant. Employing now Eqs. (8.5)–(8.9), we obtain

1 2g 2g 2 t
γ (t) = , α(t) = , T (t) = ,
2t + γ01 2t + γ01 γ01 (2t + γ01 )

σg g 8gβ02 + σ γ01 2
2gm cos [ωt]
β(t) = β01 − t− + 2 , (8.23)
2 4 (2t + γ01 ) ω 2t + γ01
296 8 Rogue Matter Waves in Bose-Einstein Condensates …

Fig. 8.2 Evolution of the atomic density (left panels) and density plots (right panels) of the first-
order rogue matter wave (top panels) and the second-order modulated rogue wave (bottom panels)
obtained with the use of the exact rational polynomial solutions (8.12) and (8.15), respectively, of
the GP equation (3.60) for β01 = β02 = −0.01, γ = 0.3, m = 1.3, and ω = 10. Reprint from Ref.
[9], Copyright 2022, with permission from Springer Nature

where γ01 , β01 , and β02 are three arbitrary real constants to be taken suitably.

Fig. 8.3 Evolution of the second-order rogue matter waves generated with the exact rational poly-
nomial solution (8.15) of the GP Eq. (8.1) for g = 1, β01 = 0.8, β02 = −0.2, and γ01 = 25. a
Second-order rogue matter wave under the effect of bias magnetic field, obtained with σ = 0.2 and
m = 0. b Second-order rogue matter wave under the effect of laser field generated with σ = 0,
m = 1, and ω = 1.5. c Second-order rogue matter wave under the combined effect of both the bias
magnetic and laser fields, generated with σ = 0.2, m = 1, and ω = 1.5. Reprint from Ref. [9],
Copyright 2022, with permission from Springer Nature
8.1 Non-autonomous Rogue Matter Waves in Bose-Einstein… 297

Fig. 8.4 Waveforms at time t0 = 0 of the first-order rogue matter wave associated with the exact
rational solution (8.12) for the solution parameters β01 = β02 = 0 and different values of g, γ01 ,
m, σ , and ω. a Waveforms of the first-rogue matter wave generated with m = 4.0, ω = 0.30,
σ = 0.20, γ01 = 25.0 and different numerical values of the nonlinearity parameter g, g = 0.25
(solid line), g = 0.30 (dashed line), and g = 0.35 (dotted line); b waveforms of the first-rogue
matter wave generated with m = 4.0, ω = 0.30, σ = 0.20, g = 0.25, and three numerical values
of the loss parameter γ01 , γ01 = 25.0 (solid line), γ01 = 26.0 (dashed line), and γ01 = 27.0 (dotted
line); c Waveforms of the first-rogue matter wave obtained with ω = 0.30, σ = 0.20, g = 0.25,
γ01 = 25.0, and different numerical values of the laser field parameter m, m = 3.90 (solid line),
m = 3.95 (dashed line), and m = 4.0 (dotted line); d waveforms of the first-rogue matter wave
generated with m = 4.0, ω = 0.30, g = 0.25, γ01 = 25.0, and three numerical values of parameter
σ of the magnetic field, σ = 0.15 (solid line), σ = 0.20 (dashed line), and σ = 0.25 (dotted line); e
waveforms of the first-rogue matter wave generated with m = 4.0, σ = 0.20, γ01 = 25.0, g = 0.25,
and for different values of the laser frequency ω, ω = 0.25 (solid line), ω = 0.30 (dashed line),
and ω = 0.35 (dotted line). Reprint from Ref. [9], Copyright 2022, with permission from Springer
Nature

A typical example of the evolution of the dynamics of dissipative non-autonomous


second-order rogue matter wave is depicted in Fig. 8.3 for various values of (σ, m) ,
(σ, m) = (0.2, 0) for Fig. 8.3a, (σ, m) = (0, 1) for Fig. 8.3b, and (σ, m) = (0.2, 1)
for Fig. 8.3c. Plots of Fig. 8.3 show how parameter σ and m of the linear potential
may affect to evolution of the rogue matter waves.
To show the effects of different potential parameters on the model equation as
well as those of the solution parameters on the waveform and the wave evolution
in BECs trapped in linear potentials, we have depicted in Fig. 8.4 the variation of
the first-order rogue wave. Plots of this figure show how much the first order rogue
matter waves are affected by parameters g, γ , m, σ, as well as ω.
Case of BECs with a time-dependent s-wave scattering length To study the case
of BECs with a time-dependent s-wave scattering length, we follow Kengne and
Talla [17] and consider the situation when g(t) = g01 exp [κt] with g01 |κ| > 0 and
γ (t) = κ/2. It follows from Eqs. (8.5)–(8.8) that
298 8 Rogue Matter Waves in Bose-Einstein Condensates …

Fig. 8.5 Generation of the first-order RW obtained with the use of the first-order rational polyno-
mial solution (8.12) of the GP equation (8.1) for g01 = 0.5 and γ (t) = −0.05 when the solution
parameters are taken as β10 = 0.1 and β02 = 0.2. a First-order dissipative forced RW under the
only effect of bias magnetic (m = 0), generated with σ = 0.25. b First-order dissipative forced RW
under the only effect of laser field (σ = 0), obtained with m = 4 and ω = π/2. c First-order dissipa-
tive forced RW under the combined effect of the bias magnetic and laser fields (mσ = 0), generated
with σ = 0.25, m = 4 and ω = π/2. Reprint from Ref. [9], Copyright 2022, with permission from
Springer Nature

2g01 m
α(t) = 2g01 , β(t) = β01 + 4g01
2
β02 t − g01 σ t 2 + cos [ωt] , T (t) = 2g01
2
t,
ω2
(8.24)
where β01 and β02 are two arbitrary real constants to be taken from the condition that
|β(t)| > 0.
It is evident that parameter g of the two-body interactions as well as different
parameters of the linear potential will affect the motion of rogue matter waves in
BEC systems under consideration. For a best understanding, we have displayed in
Figs. 8.5 and 8.6 the evolution of dissipative non-autonomous first- and second-order
rogue matter waves, respectively. Different plots are generated with a negative value
of γ (t).
Management of Dissipative Rogue Matter Waves in BEC System Trapped
in Combined Time-Independent Harmonic Repulsive and Modulated Linear
Potentials As the last example, we consider a BEC system trapped in combined time-
independent harmonic repulsive and modulated linear potentials. For this example,
we take the strength k(t) of the magnetic trap as k(t) = −2κ 2 (κ = 0) [21], and the
parameter g of the two-body interaction as [26]

g(t) = g01 exp [−2κt] (with g01 > 0).

The strength λ(t) of the linear potential is assumed to be of the form [30–32]

λ(t) = σ + m cos [ωt] ,

where σ, m, and ω are three real parameters. One can easily verify that by taking
γ (t) = −2κ [17], the integrable condition (8.9) will be satisfied. Integrating system
(8.5)–(8.8) yields
2
g01
α(t) = 2g01 exp [2κt] , T (t) = − (1 − exp [4κt]) , β(t) = β01

8.1 Non-autonomous Rogue Matter Waves in Bose-Einstein… 299

Fig. 8.6 Waveforms at time t0 = 0 of the second-order rogue matter wave generated with the help
of the rational polynomial solution (8.15) of the GP Eq. (8.1) for γ (t) = 0.05, κ = 0.1, β01 = 0.1,
and β02 = 0.2, and different numerical values of g01 , σ , m, and ω. a Waveforms of the second-order
rogue matter wave for σ = 0.25, m = 4, ω = π/2, and different values of the nonlinearity parameter
g01 , g01 = 0.25 (solid line), g01 = 0.30 (dashed line), and g01 = 0.35 (dotted line); b Waveforms
of the second-order rogue matter wave for g01 = 0.25, m = 4, ω = π/2, and different values of
the bias magnetic parameter σ, σ = 0.25 (solid line), σ = 0.30 (dashed line), and σ = 0.35 (dotted
line); c waveforms of the second-order rogue matter wave for g01 = 0.25, σ = 0.25, ω = π/2, and
different values of the laser parameter m, m = 3 (solid line), m = 4 (dashed line), and m = 5 (dotted
line); (d): Waveforms of the second-order rogue matter wave for g01 = 0.25, σ = 0.25, m = 4, and
different values the laser modulation frequency ω ω = 1.4 (solid line), ω = 1.5 (dashed line), and
ω = 1.6 (dotted line). Reprint from Ref. [9], Copyright 2022, with permission from Springer Nature


σ 2m g01
+g01 + cos [ωt] + β02 exp [2κt] exp [2κt] , (8.25)
2κ 2 ω2 + 4κ 2 κ

where β01 and β02 are two arbitrary real constants to be taken from the condition
|β(t)| > 0.
With the above data, we have displayed in Fig. 8.7 the dynamics of the second-
order dissipative forced rogue matter waves for γ (t) = −0.1 and different (σ, m, ω).
Plots of Fig. 8.7 show second-order dissipative forced rogue waves embedded on a
nonzero increasing backgrounds and localized in both the time t and the space x;
this means that the wave can concentrate the energy of the BEC system in a small
region, as it is clearly seen from Fig. 8.8. Different plots of Fig. 8.7 show how
much parameters σ, m, ω may affect the non-autonomous rogue matter waves in
BECs trapped in combined time-independent harmonic and linear potentials. These
300 8 Rogue Matter Waves in Bose-Einstein Condensates …

Fig. 8.7 Evolution of the dissipative forced second-order rogue matter waves generated with the
use of the rational polynomial solution (8.15) of the GP Eq. (8.1) for data (8.25) with g01 = 0.25
and κ = 0.05 [21] and the solution parameters β01 = 0.1 and β02 = 0.2. Different plots show
the evolution of the dissipative forced second-order RW for BECs trapped in a time-independent
harmonic repulsive potential a when both the bias magnetic and laser field are absent, σ = m = 0;
b under the effect of only the bias magnetic field, m = 0 and σ = 0.25; c under the effect of only
the laser field, σ = 0, m = 5 and ω = 3; d under the combined effects of the bias magnetic and
laser fields, σ = 0.25, m = 5 and ω = 3. Reprint from Ref. [9], Copyright 2022, with permission
from Springer Nature

parameters can also be used to manage the motion of rogue matter waves in such
systems.
Computing the amplitude of the first- and second-order rogue matter waves asso-
ciated with the above set of data, we find that the wave amplitude is proportional
to 
p(t) = 2g01 exp [3κt] .

Due to the presence of exp [nκt] in each of α(t), β(t), and T (t) [as √
we can see from
(8.25)], the fact that the wave amplitude is proportional to p(t) = 2g01 exp [3κt]
does not necessary lead to an increasing or decreasing in the wave amplitude.This
fact can be well seen from plots of Fig. 8.9. Plots of Fig. 8.9 reveal that the main
peak of the wave does not depend on the sign of parameter κ.
8.1 Non-autonomous Rogue Matter Waves in Bose-Einstein… 301

Fig. 8.8 Evolution of the second-order dissipative forced rogue matter wave generated with the
use of the exact rational polynomial solution (8.15) of the GP Eq. (8.1) with data ( 8.25) and for
the feeding parameter γ (t) = −0.1. The parameters are the same as those in Fig. 8.7. Reprint from
Ref. [9], Copyright 2022, with permission from Springer Nature

Fig. 8.9 Plots of the second-order dissipative forcing (top panels) and damped (bottom) rogue
matter waves generated with the help of the exact rational solution (8.15) for different values of
the feeding/loss parameter κ and with the same parameters as in Fig. 8.8d. a, d κ = ±0.01, b, e
κ = ±0.05; c, f κ = ±0.1. Positive [negative] κ is associated with BEC system with gain [loss] of
atoms (top panels) [bottom panels]. Reprint from Ref. [9], Copyright 2022, with permission from
Springer Nature
302 8 Rogue Matter Waves in Bose-Einstein Condensates …

8.1.4 Conclusion and Discussions

A quasi-one-dimensional GP equation which can be used to describe the evolution


of cigar shaped BECs with loss/gain of atoms is considered. The integrable condition
is presented and the first- and second-order rational solutions are built. These ratio-
nal solutions are used to investigate the dynamics of dissipative non-autonomous
first- and second-order rogue matter waves in the BEC context. Our results have
showed that
(i) the motion of dissipative rogue matter waves in the BEC systems can be manip-
ulated with the use of the both harmonic and linear parts of the potentials;
(ii) the motion of dissipative rogue wave of the BEC systems can be manipulated
with the control of the scattering length;
(iii) the CW backgrounds of the dissipative rogue waves as well as the wave width
can be seriously modified by the gain/loss of the condensate atoms.

8.2 Chirped Rogue Matter Waves in Bose-Einstein


Condensates with a Variable Scattering Length
in an Expulsive Potential

A cubic distributed NLS equation with parabolic external potential which governs the
macroscopic wave-function of one-component BECs trapped in a parabolic external
potential is considered. Using the mapping method, we find the integrable conditions
and reduce this equation to a cubic-quintic NLS equation with the Raman-like term
producing the self-frequency shift. The baseband modulational instability, respon-
sible of the generation of chirped rogue waves is investigated and its criterion is
presented. Under this criterion, we build first- and second-order rational solutions
with nonlinear chirping for the mathematical models under consideration. These
solutions are used to study analytically the dynamics of chirped rogue matter waves
in BECs with a variable scattering length in an expulsive harmonic potential.

8.2.1 Introduction

Extensive theoretical and experimental works have been devoted on the dynamics
of atomic matter waves in BECs including dark/bright solitons [33–37, 37–41], and
breather solitons and rogue waves [42–45], as well as gap solitons, four-wave mixing,
and many other coherent structure modes [46–51, 51–54] . The above enumerated
studies are based on the GP equation with variable coefficients which under some
restrictions can be reduced to the classical integrable NLS equation [55–60]. Like
many other nonlinear media, BECs may support the propagation of chirped rogue
matter waves embedded on a nonzero CW backgrounds [61–66]. To the best of
our knowledge, no much studies have been devoted to the dynamics of chirped
8.2 Chirped Rogue Matter Waves in Bose-Einstein Condensates… 303

rogue matter waves [22] in BECs with variable scattering lengths trapped in an
external parabolic potential. Based on the distributed cubic GP equation and using
the phase-engineering technique (PET) [60, 67, 68], we investigate in the present
Section the dynamics of chirped rogue matter waves in BECs with a time-varying
atomic scattering length when both the two- and three-body interactions are taken
into account [67, 69–72].
Thus, the mathematical models to be studied in this Section is the following
distributed NLS equation with an external parabolic potential [37, 73–76]:

∂ψ(x, t) ∂ 2 ψ(x, t)
i + + 2a(t) |ψ(x, t)|2 ψ(x, t) − k(t)x 2 ψ(x, t) = 0. (8.26)
∂t ∂x2
Throughout this Section, Eq. (8.26) will be considered in the BECs context for which
ψ(x, t) is the normalized mean-field wave function, and time t and coordinate x are
measured in the same units as in the previous two chapters, while the cubic term,
proportional to 2a(t) plays the role of the two-body inter-atomic interactions [60,
77], and throughout this Section is assumed to be positive (a(t) > 0). Therefore,
k(t)x 2 will be the harmonic external potential with magnitude k(t). Next, we will
assumed that the stability conditions are satisfied so that the increase of the number
of atoms will not lead to the any collapse of the system under consideration [38, 43,
67, 69, 70, 78]. Under the above assumptions, Eq. (8.26) is suitable to describe the
control and management of BECs [60, 67, 77].

8.2.2 Phase Engineering and Chirped Wave Solutions

Phase-engineering Technique To find the integrable conditions of Eq. (8.26) and


present its exact chirped wave solutions, we employ the PET [67] and seek the
solution of Eq. (8.26) in the form

1   
ψ(x, t) = √ (X, T ) exp i α(t)x 2 − 2θ (x, t) , (8.27)
(t)

where T = T (t),  = (t) and α(t) are real functions to be determined, X ≡ x/(t),
and θ (x, t) is the phase imprint, related to the new complex wave function  as
follows [67, 79]:

∂θ a3
= − ||2 , (8.28)
∂x 4
  
∂θ a3 1 ∂∗ ∗ ∂ 1 4a1 + a32
=i  −  − 4i xα(t) ||2
+ ||4 .
∂t 4 2 ∂X ∂X  82
(8.29)

Here, a1 and a3 the phase-imprint parameters.


304 8 Rogue Matter Waves in Bose-Einstein Condensates …

Next, we assume that α(t) = 0 and impose to the functional parameters T (t),
(t), and α(t) to satisfy the following differential system

dT 1
= 2, (8.30)
dt 
1 d
− 4α = 0, (8.31)
 dt

+ 4α 2 + k(t) = 0, (8.32)
dt
so that the model equation (8.26) is converted into the following cubic-quintic deriva-
tive NLS equation without external potential

∂ ∂ 2  ∂
i + + 2a(t) ||2  + a1 ||4  + ia3 ||2  = 0. (8.33)
∂T ∂ X2 ∂X

In Eq. (8.33), the cubic nonlinearity 2a(t) is positive, the quintic nonlinearity a1
and parameter a3 of the self-frequency shift are of any sign. In the BECs context,
Eq. (8.33) can be used to describe BEC systems with both two- and three-body
inter-atomic interactions.
In the following, we focus our attention to the special case when the phase-imprint
parameters a1 and a3 are related as

a1 = a32 /4. (8.34)

Next, we take the functional parameter (t) as follows


a2
(t) = , a2 > 0, (8.35)
2a(t)

where a2 is a free real parameter. We then find from Eqs. (8.30)–(8.32) that k(t)
and a(t) must satisfy the following constraint [integrable condition for the model
equation (8.26)]
dA 1 da
− A2 − 4k(t) = 0, A = ; (8.36)
dt a dt

under the restriction (8.36), we find T (t) and α(t) to be

t
4 1 da
T (t) = 2 a 2 (z)dz, α(t) = − . (8.37)
a2 4a dt
0


If parameter a3 of the self-frequency shift is chosen as a3 = −2 2a2 β, the dis-
√ cubic-quintic NLS equation (8.33) under the transformation τ = (a2 /2) T,
tributed
ζ = a2 /2X is reduced to the integrable Kundu-Eckhaus (KE) equation [67, 80–83]
8.2 Chirped Rogue Matter Waves in Bose-Einstein Condensates… 305

∂ ∂ 2  ∂
i + + 2 ||2  + 4β 2 ||4  − 4iβ ||2  = 0, (8.38)
∂τ ∂ζ 2 ∂ζ

which contains only one parameter, β. The cubic derivative term of Eq. (8.38) offsets
the modulation instability driven the three-body√interactions [67]. In the following,
we focus ourselves to the situation when |β| < 2/2 ≈ 0.70711, and set 2β 2 ≡ χ0
[69]. It follows from Eq. (8.27) that

+∞ +∞ +∞


1
|ψ(x, t)| d x =
2
|(X, T )| d x =
2
|(X, T )|2 d X,
(t)
−∞ −∞ −∞

since X ≡ x/(t), so that the total number of the condensate atoms under the trans-
formation (8.27) remains unchanged.
Baseband Modulational Instability Analysis Now, let us study the phenomenon
of the baseband modulational instability, a mechanism responsible of the generation
of rogue waves in nonlinear media [62–64, 84, 85]. This phenomenon implies that
the CW background may be unstable against perturbations with infinitesimally small
wavenumbers [62].
First of all, we note that Eq. (8.33) under condition (8.35) admits the CW solution
   
0 (X, T ) = ρ0 exp i 0 X + a2 ρ02 + a1 ρ04 − 20 T , (8.39)

where ρ0 and 0 are, respectively, the carrier amplitude and wavenumber and  ω=
a2 ρ02 + a1 ρ04 − 20 is the angular frequency. Perturbing the CW solution (8.39) as
  
(X, T ) = 0 (X, T ) 1 + b1 exp [i (X − K T )] + b2∗ exp −i X − K ∗ T
(8.40)
and using the linear stability analysis, we arrive to the following perturbation linear
dispersion relation

2  
2 K − 20 − a3 ρ02 − 2 + ρ02 2a2 + 4a1 − a32 ρ02 = 0, (8.41)

which defines complex angular frequency K in the limit  → +0 only under the
condition that 
2a2 + 4a1 − a32 ρ02 > 0. (8.42)

In the following, condition (8.42) will be referred to as the condition of the baseband
MI for the model equation. Under this condition, we derive the following growth
rate (gain) of the baseband MI

    
() =  2a2 + 4a1 − a32 ρ02 ρ02 − 2 , for 0 ≤  ≤ ρ0 2a2 + 4a1 − a32 ρ02 . (8.43)
306 8 Rogue Matter Waves in Bose-Einstein Condensates …

Fig. 8.10 Maps of the baseband modulational instability in the plane of  and ρ0 , as produced by
Eq. (8.43) for a2 = 2, a3 = −4β and four values of a1 : a a1 = −2β 2 , b a1 = 0, c a1 = 2β 2 , and d
a1 = 4β 2 , withβ = 1/3. In different plots, dashed lines correspond to the MI boundaries, defined

by either  = 2a2 ρ02 + 4a1 − a32 ρ04 or  = 0. Reprint from Ref. [10], Copyright 2022, with
permission from Elsevier

Using Eq. (8.43), we have displayed in Fig. 8.10 baseband MI map in the (, ρ0 )
plane. This figure shows that the baseband MI occurs not at all values of ρ0 , and the
instability region expands as a1 increases.
Chirped Rogue Waves
Under the integrable condition (8.36) and condition (8.42) of the baseband MI, we
intend here to build exact chirped RW solutions√of the model equation (8.26). In the
special case when a1 = a32 /4 = 0 and a3 = −2 2a2 β = 0, first-order RW solutions
of Eq. (8.33) are found to be [82, 83]
8.2 Chirped Rogue Matter Waves in Bose-Einstein Condensates… 307

4 − D1 + 8ia2 T
(X, T ) =
D1
     
 a2 2 + 4β 2   4 − D1 + 8ia2 T 2
exp i −β 2a2 X + T + β 2a2   , (8.44)
2 D1  dX
  2
D1 (X, T ) = 1 + 4a22 T 2 + 2a2 X + 2 2a2 βT , (8.45)

a2 being a free positive constant. Using this rogue wave solution, we find the following
first-order RW solution of the model equation (8.26):

! !
2a(t) (4 − D1 )2 + 64a22 T 2
ψ(x, t) =
a2 D12
   !
8a2 T 1 da 2 2
exp i arctan − x − 2β a(t)x
4 − D1 4a dt a2

 t !    
2 2 + 4β 2 2  4 − D1 + i8a2 T 2
+ a (z)dz − 2θ(x, t) + 2β
2
a(t)   dx ,
a2 a2  D1 
0

(8.46)

where D1 = D1 (X, T ) is given in Eq. (8.45) and X = X (x, t) and T = T (t) are
defined in Eq. (8.37). The corresponding chirping reads

!   
1 da 2 1 ∂ 8a2 T 
δω = x − 2a(t) β+√ × arctan . (8.47)
2a dt a2 2a2 ∂X 4 − D1  X =X (x,t), T =T (t)

Now, we apply the results from Ref. [82] on Eq. (8.38) to obtain the following
second-order rogue wave solutions of Eq. (8.33):
     
F2 + i G 2  a2 2 + 4β 2   F2 + i G 2 2
(X, T ) = exp i −β 2a2 X + T + β 2a2   dX ,
D2 2  D2 
(8.48)
where
"  
a2
F2 = 45 − 90a2 X 2 − 36a22 X 4 + 8a23 X 6 − 48a2 15β + 12βa2 X 2 − 4βa22 X 4 X T
2

   
−12a2 392 + 60β + 12a2 5 + 12β X − 4a22 1 + 20β 2 X 4 T 2
2 2 2 2

"
   
a2
−128βa23 9 5 + 4β 2 − 2a2 3 + 20β 2 X 2 X T 3 (8.49)
2

  
−48a24 11 + 120β 2 + 48β 4 − 2a2 1 + 20β 2 1 + 4β 2 X 2 T 4
"  2  3
a2
+768a25 β 1 + 4β 2 X T 5 + 64a26 β 1 + 4β 2 T 6 ,
2
308 8 Rogue Matter Waves in Bose-Einstein Condensates …
"
 a2 
G 2 = 24a2 15 + 12a2 X − 4a2 X T − 768a2 β
2 2 4 2
2a2 X 3 − 3X T 2 − 192a23
2
  
× 1 − 12β 2 + 2a2 12β 2 + 1 X 2 T 3
"
a2  2
−3072a2 β
4
1 + 4β 2 X T 4 − 384a25 1 + 4β 2 T 5 , (8.50)
2
"
a2 2 5 
D2 = 9 + 54a2 X + 2
+
12a22 X 4 + 48a2 β
8a23 X 6 4a2 X + 4a2 X 3 + 9X T
2
  
+12a22 33 + 36β 2 + 12a2 (2β − 1) (2β + 1) X 2 + 4a22 1 + 20β 2 X 4 T 2
"
3 a2
 2   
+128a2 3β 4β − 3 X + 2a2 β 3 + 20β 2 X 3 T 3 (8.51)
2

2  
+48a24 4β 2 − 3 + 2a2 1 + 20β 2 1 + 4β 2 X 2 T 4
 2 3
+384βa25 2a2 1 + 4β 2 X T 5 + 64a26 1 + 4β 2 T 6 .

Here, a2 is any positive constant, T = T (t) is given by Eq. (8.37), X = X (x, t) =


x/(t), where (t) is defined in Eq. (8.35). Going back to original variable x, t,
and ψ(x, t) , we obtain the following second-order RW solution of the model equa-
tion (8.26):

! !
2a(t) F22 + G 22
ψ(x, t) =
a2 D22
⎡ ⎛ !  t
 
G 1 da 2 2 2 + 4β 2
exp ⎣i ⎝arctan
2
− x − 2β
2
a(t)x + × a 2 (z)dz
F2 4a dt a2 a2
0

!    
2  F2 (X (x, t), T (t)) + i G 2 (X (x, t), T (t)) 2
−2θ (x, t) + 2β a(t)   dx .
a2  D2 (X (x, t), T (t)) 
(8.52)

The corresponding chirping reads


!   
1 da 2 1 ∂ G 2 
δω = x − 2a(t) β+√ × arctan .
2a dt a2 2a2 ∂X F2  X =X (x,t), T =T (t)
(8.53)

As we will see in the following, any point (x0 , 0) where F (x, t) and G(x, t)
vanish at point (x, t) = (x0 , 0) is a position of peak and hole of the chirping. At each
8.2 Chirped Rogue Matter Waves in Bose-Einstein Condensates… 309

of such a position (x0 , 0), the expression


 
∂ G 
lim arctan (8.54)
(x,t)→(x0 ,0) ∂X F  X =X (x,t), T =T (t)

is undefined.

8.2.3 Evolution of Chirped Rogue Waves Under the Action of


the Time-Dependent Atomic Scattering Length and
Parabolic Potential

Here, we employ the above exact chirped rogue wave solutions of the model equation
(8.26) to study analytically the evolution of chirped rogue matter waves in BEC with
time-varying atomic scattering length and external parabolic potential. Without the
loss of generality, we take the solution parameter a2 as a2 = 2. It is seen from the
found chirped rogue wave solutions
√ of the model equation Eq. (8.26) that the wave
amplitudes is proportional to a(t) and chirping δω(x, t) is free from the phase-
imprint parameter θ (x, t).
The following three special cases will be considered in our study:
(i) Case of BECs with time-independent expulsive potential : [39]

k(t) = −λ2 /4;

(ii) Case of BECs with a time-periodic modulation of the s-wave scattering length
[86]:
a(t) = r0 [1 + m sin(ωt)] with 0 < m < 1 and r0 > 0;

(iii) Case of BECs with time-varying external expulsive potential [87]:

μ20
k(t) = − [1 − tanh (μ0 t/2)] , with r0 > 0.
8
For the integrable condition (8.36) to be satisfied for each of cases (i), (ii), and
(iii) we choose a(t) and k(t) as [86, 87]:

a(t) = r0 exp(λt), r0 > 0,


 
k(t) = − mω2 /4 m + sin(ωt) + m cos2 (ωt) [1 + m sin(ωt)]−2 ,
a(t) = r0 [1 + tanh (μ0 t/2)] , r0 > 0,
310 8 Rogue Matter Waves in Bose-Einstein Condensates …

respectively. In the following, parameter r0 will be referred to as the amplitude


parameter. Using the above enumerated cases, we intend here to analyze in detail
how the first- and second-order chirped RWs get modified by a(t) and k(t).
Case (i): Evolution of Chirped Rogue Waves in BECs with Time-Independent
Expulsive Potential As the first example, we follow Liang et al. [38] and consider
BECs with a time-independent expulsive (anti-trapping) harmonic potential, mod-
eled by Eq. (8.26) with the potential magnitude k(t) = −λ2 /4, parameter a(t) of the
two-body interactions being
a(t) = r0 exp(λt). (8.55)
−1
It
2then follows
  (8.35) and (8.37) that (t) = r0 exp (−λt) and T (t) =
from Eqs.
r0 /2λ exp (2λt) − 1 .
For the present example, we use the first-order RW solution (8.46) and the corre-
sponding chirping (8.47) to show in Figs. 8.11 and 8.12 a typical evolution of respec-

Fig. 8.11 Left panels: spatiotemporal evolution of the first-order RW generated with the use of
the rational solution (8.46) for different values of parameter β. a β = −1/3; b β = 0; c β = 1/3.
Right panels: density distribution in the first-order obtained with the help of the rational polynomial
solution (8.46) for different values of β. d β = −0.7; e β = −1/3; f β = 0; g β = 1/3; h β = 0.7.
Other parameters used in generating different plots are r0 = 0.25 and λ = 0.02. Reprint from Ref.
[10], Copyright 2022, with permission from Elsevier
8.2 Chirped Rogue Matter Waves in Bose-Einstein Condensates… 311

Fig. 8.12 Spatiotemporal evolution (left panels) and top-view distributions (right panels) of chirp-
ing (8.47) associated with the first-order RW for with different values of parameter β. a, d β = 01/3;
b, e β = 0; c, f β = 1/3. Other parameters used in generating different plots are r0 = 0.25 and
λ = 0.02. Reprint from Ref. [10], Copyright 2022, with permission from Elsevier
312 8 Rogue Matter Waves in Bose-Einstein Condensates …

Fig. 8.13 Evolutive formation of the first-order RWs in BEC with the nonlinearity coefficient
exponentially increasing as per Eq. (8.55) and the time-independent trap’s frequency, k(t) = −λ2 /4.
Different plots are generated for β = 1/3, λ = 0.02, and different numerical values of the amplitude
parameter r0 in Eq. (8.55): r0 = 0.05 (a); r0 = 0.1 (b); r0 = 0.15 (c); r0 = 0.3 (d). Reprint from
Ref. [10], Copyright 2022, with permission from Elsevier

tively the first-order rogue matter waves and the distribution of the corresponding
chirping for different numerical values of parameter β, β = 0 corresponding to the
BEC Peregrine soliton. Different plots show of Fig. 8.11 feature one hump and two
valleys around the center. We can see from plots of Figs. 8.11 and 8.12 how much
parameter β affect the wave motion. For example, parameter β produces an essential
skew angle relative to the ridge of the RW in the anti-clockwise [clockwise] direction
for β > 0 [for β < 0], and the skew angle becomes larger with the increase of |β|
(this is clearly seen from plots of the bottom panels). We can also see from plots of
Figs. 8.11 and 8.12 that the shape of both the RW and the corresponding chirping
does not change drastically for β = 0. Left panels of Figs. 8.11 and 8.12 reveal that
the BECs RWs as well as the corresponding chirping in the present case are embed-
ded on an increasing nonzero CW background. As we can see from plots of Fig. 8.12,
the chirping is localized in both time and space, and exhibits two dark-bright doubly
localized structures, with the same location as in the two valleys of the corresponding
first-order RW.
To investigate the effects of the amplitude parameter r0 on the wave evolution,
we show in Figs. 8.13 and 8.14 the formation of respectively the first-order and
the second-order rogue matter waves for different numerical values of the amplitude
8.2 Chirped Rogue Matter Waves in Bose-Einstein Condensates… 313

Fig. 8.14 Evolutive formation of the second-order rogue matter waves in the BEC models with the
same a(t) and k(t) as in Fig. 8.12, the coefficients being β = 1/3 and λ = 0.02. Parameter r0 in Eq.
(8.55) takes values r0 = 0.01 (a); r0 = 0.05 (b); r0 = 0.1 (c); r0 = 0.2 (d); r0 = 0.6 (e); r0 = 1.0
(f). Reprint from Ref. [10], Copyright 2022, with permission from Elsevier

parameter r0 . Plots of Fig. 8.13 reveal that the peak value of the first-order RW grows
and its width compresses when parameter r0 increases. Plots of Fig. 8.14 show that
depending on the numerical value of the amplitude parameter r0 , the second-order
RW can either remain a single RW or split into two to three first-order RWs whose
width decreases with the increase of r0 . It is also seen from plots of Figs. 8.13 and
8.14 that the wave amplitude increases with the increase of r0 .
To show the effect of parameter β on the dynamics of the BEC second-order rogue
matter waves and the corresponding chirping in the present case, we display in Figs.
8.15 and 7.16 the dynamics of the second-order BEC RW and the corresponding
chirping associated with the exact rogue wave solution (8.52) and chirping (8.53)
for different numerical values of β. It is seen from different plots that β has almost
the same effects on the second-order BEC RWs as on the first-order ones and the
corresponding chirping: it does not impact dramatically the wave shape, but produces
a skew angle relative to the RW’s ridge or chirping’s ridge in the anti-clockwise
[clockwise] direction for β > 0 [β < 0]. Plots of Fig. 8.16 reveal that the chirping
314 8 Rogue Matter Waves in Bose-Einstein Condensates …

Fig. 8.15 Effects of strength β of the three-body inter-atomic interactions on the second-order
rogue matter waves moving in a BEC with the exponentially-varying atomic scattering length (see
Eq. (8.55) trapped in an expulsive time-independent parabolic potential. Plots of the left panels and
right panels are organized as in Fig. 8.11, but for the rational polynomial solution given by Eq.
(8.52), for β = −1/3 (a, d), β = 0 (b, e), and β = 1/3 (c, f). Parameters r0 and λ, as well as the
time-dependent ones, a(t) and k(t), are given in the text. Reprint from Ref. [10], Copyright 2022,
with permission from Elsevier
8.2 Chirped Rogue Matter Waves in Bose-Einstein Condensates… 315

Fig. 8.16 Spatiotemporal plot (left panels) and top-view distributions (right panels) of the frequency
chirp associated with the second-order RW solution (8.52), given by Eq. (8.53), for k = −λ2 /4 and
a(t) taken as per Eq. (8.55), cf. Fig. 8.12 for the first-order RWs. To generate different plots, we
have used the following numerical data: r0 = 0.25, λ = 0.02, and three values of β, viz., β = −0.7
(a, d), β = 0 (b, e), and β = 0.7.(c, f). Reprint from Ref. [10], Copyright 2022, with permission
from Elsevier

associated with the second-order BEC RWs is localized both in time and space.
Comparing plots of Figs. 7.15 and 8.16, we find that the second-order RW and the
corresponding chirping exhibit respectively four valleys and four dark-bright doubly
localized structures, located at the same position.
Case (ii): Evolution of Chirped Rogue Matter Waves in BECs with Time-
Periodic Modulation of the Scattering Length As the second example, we consider
a BEC system with a temporally periodic variation of the s-wave scattering length
[86]. For this special case, we assumed the parameter a(t) of the two-body interac-
tions to be given as
316 8 Rogue Matter Waves in Bose-Einstein Condensates …

Fig. 8.17 Left panels: top-view distribution for the first-order rogue matter wave generated with
the help of the exact rational solution (8.46) with r0 = 0.25, m = 0.4, and for different values of
strength β of the three-body inter-atomic interactions, viz., β = −0.7 (a); β = 0 (b); β = 0.7 (c).
Middle panels: Spatiotemporal evolution of rogue matter wave for the same rational solution with
r0 = 0.25, β = 1/3, and three values of parameter m, viz., m = 0.1 (d); m = 0.4 (e); m = 0.8 (f).
Right panels: Density plots for the same RW solution with β = 1/3, m = 0.4, and three values
of the amplitude parameter r0 , viz., r0 = 0.05 (g), r0 = 0.12 (h); r0 = 0.25 (i). Different plots are
generated with ω = 0.5. Reprint from Ref. [10], Copyright 2022, with permission from Elsevier

a(t) = r0 [1 + m sin (ωt)] , with 0 < m < 1, and r0 > 0. (8.56)

The corresponding potential strength k(t) is found from the integrable condition
(8.36) and reads
 
mω2 m + sin (ωt) + m cos2 (ωt)
k(t) = − . (8.57)
4 [1 + m sin (ωt)]2

To analyze how parameters a(t) and k(t) given by respectively Eq. (8.55) and
(8.56) modify the dynamics of the first- and second-order chirped BEC RWs, we
display respectively in Figs. 8.17 and 8.18 and 8.19 the evolution of first- and second-
order RWs and the distribution of the corresponding frequency chirps for different
numerical values of parameters β (top panels), m (middle panels), and r0 (bottom
panels), β = 0 corresponding to the BEC Peregrine solitons of the standard NLS
equation. Different plots show that the first- and the second-order RWs and the
corresponding chirping are embedded on a modulated increasing nonzero CW back-
ground. Plots of these three figures show that β produces a skew angle relative to the
8.2 Chirped Rogue Matter Waves in Bose-Einstein Condensates… 317

Fig. 8.18 Top panels: top-view distribution for the second-order rogue matter wave obtained with
the use of the exact rational solution (8.52) with r0 = 0.25, m = 0.4, and for different values of
strength β of the three-body interaction, viz., β = −1/3 (a); β = 0 (b); β = 1/3 (c). Middle panels:
spatiotemporal evolution of the second-order rogue matter wave generated with the same rational
solution with r0 = 0.25, β = 1/3, and three values of parameter m, viz., m = 0.1 (d); m = 0.5
(e); m = 0.9 (f). Bottom panels: Spatiotemporal evolution of the second-order rogue matter wave
obtained with the help of the same rational solution for β = 1/3, m = 0.2, and three values of the
amplitude parameter r0 , viz., r0 = 0.01 (g); r0 = 0.05 (h); r0 = 0.25 (i). Reprint from Ref. [10],
Copyright 2022, with permission from Elsevier

ridge of the RWs and their corresponding chirping, in the clockwise [anti-clockwise]
direction for β < 0 [β > 0]. It is clearly seen from different plots that the chirping
associated to first- and second-order RWs exhibit respectively two- and four-peak
dark-bright doubly localized structures, located at the same positions where valleys
of the corresponding RWs are located. Plots of the middle and bottom panels of Figs.
8.17 and 8.18 demonstrate how much the amplitude and the structure of the first-
and second-order RWs are affected by parameters m and r0 ; for example, the wave
amplitude increases with the increase in each of m and r0 , and the best structure of
the BEC RWs is obtained for small values of m and large values of r0 ; therefore, m
and r0 have the same effect on the waves’ amplitudes and opposite effect on their
structure.
Case (iii): Chirped Rogue Waves in Under the Action of the Scattering Length
and Expulsive Parabolic Potential Subjected to the Stepwise Temporal Modu-
lation As our last example, we consider a BEC system with time-varying scattering
length trapped in an expulsive parabolic potential subjected to the stepwise temporal
modulation. For this example, we choose a stepwise modulation profile defined as
[87, 88], we
318 8 Rogue Matter Waves in Bose-Einstein Condensates …

Fig. 8.19 Left panels: spatiotemporal distribution of the chirping (8.47) corresponding to the first-
order RW solution (8.46). Right panels: spatiotemporal distribution of the chirping (8.53) for the
second-order RW solution (8.52). Different plots are generated with r0 = 0.25, m = 0.4, ω = 0.5,
and three different values of strength β of the delayed nonlinear response, viz., β = −1/3 (a, d);
β = 0 (b, e); β = 1/3 (c, f). Reprint from Ref. [10], Copyright 2022, with permission from Elsevier

μ20
 μ 
0
k(t) = − 1 − tanh t , μ0 > 0. (8.58)
8 2
The integrable condition (8.36) in the present case then leads to

 μ 
0
a(t) = r0 1 + tanh t , (8.59)
2
where r0 > 0 is a free constant. As we will see in the following, the wave amplitude
will increase with the increase in the values of either μ0 or r0 .
To analyze the effects of parameters a(t) and k(t) as well as that of parameters β on
the BEC rogue matter waves and the corresponding chirping in the case of a stepwise
temporal modulation profile, we show for different numerical values of r0 , μ0 , and β
the evolution plots of the first- and second-order BEC RWs and the distribution of the
corresponding chirping in respectively Figs. 8.20 and 8.21 and 8.22. It is important
8.2 Chirped Rogue Matter Waves in Bose-Einstein Condensates… 319

Fig. 8.20 Evolution of the the first-order rogue matter wave generated the exact rational solution
(8.46) of Eq. ( 8.26) with the time-modulation format given by Eqs. ( 8.58) and (8.59). Top panels:
top-view distribution for r0 = 0.25, μ0 = 0.05, and different values of strength β of the three-
body interaction, viz., β = −1/3 (a); β = 0 (b); β = 1/3 (c). Middle panels: Spatiotemporal wave
evolution for r0 = 0.25, β = −1/3, and three values of μ0 , viz., μ0 = 0.1 (d); μ0 = 0.4 (e); μ0 = 5
(f). Bottom panels: spatiotemporal plots for β = −1/3, μ0 = 0.05, and three values of the r0 , viz.,
r0 = 0.05 (g), r0 = 0.1 (h); r0 = 0.25 (i). Reprint from Ref. [10], Copyright 2022, with permission
from Elsevier

to note that the situation when β = 0 corresponds to the Peregrine solitons of the
standard NLS equation. It is seen from plots of the middle and bottom panels of Figs.
8.20 and 8.21 that the first- and second-order rogue matter waves propagate on top
of a kink-shaped increasing nonzero CW background. Different plots of Figs. 8.20,
8.21, and 8.22 reveal that parameter β of the delayed nonlinear response does not
strongly affect the shape, but produces a skew angle relative to the ridge of the RWs
and that of the corresponding chirping in the clockwise [anti-clockwise] direction for
β < 0 [β > 0]. It is seen from plots of the middle and bottom panels of Figs. 8.20 and
8.21 that the wave amplitude increases with the increase in either r0 or μ0 , and that
depending on the numerical values of parameters μ0 and r0 , the first-order [second-
order] BEC rogue matter wave is composed of one hump and two valleys [features
three humps] located around the center; moreover, the hump width decreases with
the increase of either μ0 or r0 . The best structure of the first- and second-order BEC
rogue matter waves, as we can see from plots of the middle and bottom panels of Figs.
8.20 and 8.21, is obtained with small values of μ0 and large values of r0 ; moreover,
the wave width decreases with the increase of μ0 and r0 . Plots of Fig. 8.22 reveal that
320 8 Rogue Matter Waves in Bose-Einstein Condensates …

Fig. 8.21 Evolution of the second-order rogue matter wave as per the exact rational solution (8.52)
with the time-modulation format, defined by Eqs. ( 8.58) and (8.59). Top panels: top-view distri-
bution for r0 = 0.25, μ0 = 0.05, and different values of strength β of the three-body interaction,
viz., β = −1/3 (a); β = 0 (b); β = 1/3 (c). Middle panels: spatiotemporal plots for r0 = 0.25,
β = −1/3, and different values of μ0 , viz., μ0 = 0.1 (d); μ0 = 0.4 (e); μ0 = 5 (f). Bottom panels:
Spatiotemporal wave evolution for β = −1/3, μ0 = 0.05, and three values of r0 , viz., r0 = 0.03
(g), r0 = 0.1 (h); r0 = 0.25 (i). Reprint from Ref. [10], Copyright 2022, with permission from
Elsevier

chirping associated with the first- and second-order BEC RWs is localized in time
and space, and exhibits two and four dark-bright localized structures, respectively.

8.2.4 Conclusion and Discussions

In this Section, we have considered a one-dimensional cubic GP equation which


describes the evolution of the macroscopic wave-function for BEC systems with
time-varying atomic scattering lengths trapped in an expulsive parabolic potential.
The criterion of the baseband modulational instability is presented. The integrable
condition of this model equation is derived and first- and second-order rational solu-
tions with nonlinear chirping are built. With the use of these rational solutions, we
have investigated analytically the dynamics of first- and second-order chirped rogue
matter waves in various types of BEC systems. Effects of the potential magnitude
and parameter of the two-body inter-atomic interactions on the formation and motion
of BEC rogue matter waves are investigated. Also, the effects of the strength of the
delayed nonlinear response on the chirped BEC rogue matter waves are studied.
8.2 Chirped Rogue Matter Waves in Bose-Einstein Condensates… 321

Fig. 8.22 Top-view distribution of (left panels) the chirping (8.47) associated with the first-order
rational polynomial solution (8.46), and (right panels)the chirping (8.53) corresponding to the
second-order rogue wave solution (8.52). The step-like modulation format is taken as per Eqs.
(8.58) and (8.59). Different plots are obtained with r0 = 0.25, μ0 = 0.1, and three different values
of strength β of the delayed nonlinear response, viz., β = −0.7 (a, d); β = 0, (b, c) and β = 0.7
(e, f). Reprint from Ref. [10], Copyright 2022, with permission from Elsevier
322 8 Rogue Matter Waves in Bose-Einstein Condensates …

We have found that the chirping associated with the found rogue matter waves are
localized in time and space and exhibit two dark-bright localized structures for the
first-order rogue matter waves, and four dark-bright localized structures for second-
order rogue matter waves.

References

1. M.H. Anderson, J.R. Ensher, M.R. Matthews, C.E. Wieman, E.A. Cornell, Observation of
Bose-Einstein condensation in a dilute atomic vapor. Science 269, 198 (1995)
2. C.C. Bradley, C.A. Sackett, J.J. Tollett, R.G. Hulet, Evidence of Bose-Einstein condensation
in an atomic gas with attractive interactions. Phys. Rev. Lett. 75, 1687 (1995)
3. K.B. Davis, M.-O. Mewes, M.R. Andrews, N.J. van Druten, D.S. Durfee, D.M. Kurn, W.
Ketterle, Bose-Einstein condensation in a gas of sodium atoms. Phys. Rev. Lett. 75, 3969
(1995)
4. A. Mohamadou, E. Wamba, S.Y. Doka, T.B. Ekogo, T.C. Kofane, Generation of matter-wave
solitons of the Gross-Pitaevskii equation with a time-dependent complicated potential. Phys.
Rev. A 84, 023602 (2011)
5. E. Kengne, A. Lakhssassi, R. Vaillancourt, W.M. Liu, Phase engineering, modulational insta-
bility, and solitons of Gross-Pitaevskii-type equations in 1+1 dimensions. J. Math. Phys. 54,
051501 (2013)
6. E. Kengne, X.X. Liu, B.A. Malomed, S.T. Chui, W.M. Liu, Explicit solutions to an effec-
tive Gross-Pitaevskii equation: one-dimensional Bose-Einstein condensate in specific traps. J.
Maths. Phys. 49, 023503 (2008)
7. E. Kengne, W.M. Liu, Management of matter-wave solitons in Bose-Einstein condensates with
time-dependent atomic scattering length in a time-dependent parabolic complex potential. Phys.
Rev. E 98, 012204 (2018)
8. S. Serafini, L. Galantucci, E. Iseni, T. Bienaimé, R.N. Bisset, C.F. Barenghi, F. Dalfovo, G.
Lamporesi, G. Ferrari, Vortex reconnections and rebounds in trapped atomic Bose-Einstein
condensates. Phys. Rev. X 7, 021031 (2017)
9. E. Kengne, Rogue waves of the dissipative Gross-Pitaevskii equation with distributed coeffi-
cients. Eur. Phys. J. Plus 135, 622 (2020)
10. E. Kengne, Boris A. Malomed, W.M. Liu, Phase engineering of chirped rogue waves in Bose–
Einstein condensates with a variable scattering length in an expulsive potential. Commun.
Nonlinear. Sci. Numer. Simulat. 103, 105983 (2021)
11. Y.V. Bludov, V.V. Konotop, N. Akhmediev, Matter rogue waves. Phys. Rev. A. 80, 033610
(2009)
12. F. Dalfovo, S. Giorgini, L.P. Pitaevskii, S. Stringari, Theory of Bose-Einstein condensation in
trapped gases. Rev. Mod. Phys. 71, 463 (1999)
13. U. Roy, R. Atre, C. Sudheesh, C. Nagaraja Kumar, P.K. Panigrahi, Complex solitons in Bose–
Einstein condensates with two-and three-body interactions. J. Phys. B 43, 025003 (2010)
14. FKh. Abdullaev, M. Salerno, Gap-Townes solitons and localized excitations in low-dimensional
Bose-Einstein condensates in optical lattices. Phys. Rev. A 72, 033617 (2005)
15. FKh. Abdullaev, J.G. Caputo, R.A. Kraenkel, B.A. Malomed, Controlling collapse in Bose-
Einstein condensates by temporal modulation of the scattering length. Phys. Rev. A 67, 013605
(2003)
16. D.S. Wang, X.-F. Zhang, P. Zhang, W.M. Liu, Matter-wave solitons of Bose-Einstein conden-
sates in a time-dependent complicated potential. J. Phys. B 42, 245303 (2009)
17. E. Kengne, P.K. Talla, Dynamics of bright matter wave solitons in Bose-Einstein condensates
in an expulsive parabolic and complex potential. J. Phys. B 39, 3679 (2006)
References 323

18. Z.X. Liang, Z.D. Zhang, W.M. Liu, Dynamics of a Bright Soliton in Bose-Einstein condensates
with time-dependent atomic scattering length in an expulsive parabolic potential. Phys. Rev.
Lett. 94, 050402 (2005)
19. D. MollK, A.L. Gaeta, G. Fibich, Self-similar optical wave collapse: observation of the townes
profile. Phys. Rev. Lett. 90, 203902 (2003)
20. F.K. Abdullaev, A.M. Kamchatnov, V.V. Konotop, V.A. Brazhnyi, Adiabatic dynamics of peri-
odic waves in Bose-Einstein condensates with time dependent atomic scattering length. Phys.
Rev. Letts. 90, 230402 (2003)
21. L. Khaykovich, F. Schreck, G. Ferrari, T. Bourdel, J. Cubizolles, L.D. Carr, Y. Castin, C.
Salomon, Formation of a matter-wave bright soliton. Science 296, 1290 (2002)
22. N. Akhmediev, A. Ankiewicz, J.M. Soto-Crespo, Rogue waves and rational solutions of the
nonlinear Schrödinger equation. Phys. Rev. E 80, 026601 (2009); E. Kengne, W.M. Liu, Engi-
neering rogue waves with quintic nonlinearity and nonlinear dispersion effects in a modified
Nogochi nonlinear electric transmission network. Phys. Rev. E 102, 012203 (2020); D.H. Per-
grine, Water waves, nonlinear Schrödinger equations and their solutions. J. Austral. Math. Soc.
Ser. B 25, 16 (1983)
23. N. Akhmediev, A. Ankiewicz, J.M. Soto-Crespo, Rogue waves and rational solutions of the
nonlinear Schrödinger equation. Phys. Rev. E 80, 026601 (2009); E. Kengne, A. Lakhssassi,
W.-M. Liu, R. Vaillancourt, Phase engineering, modulational instability, and solitons of Gross–
Pitaevskii-type equations in 1+1 dimensions. Phys. Rev. E 87, 022914 (2013); J.M. Dudley,
G. Genty, F. Dias, B. Kibler, N. Akhmediev, Modulation instability, Akhmediev Breathers and
continuous wave supercontinuum generation. Opt. Express 17, 21497 (2009)
24. Z. Jie-Fang, C.-Q. Dai, Control of nonautonomous matter rogue waves. Acta Phys. Sin. 65,
050501 (2016)
25. S. Loomba, H. Kaur, R. Gupta, C.N. Kumar, T.S. Raju, Controlling rogue waves in inhomo-
geneous Bose-Einstein condensates. Phys. Rev. E 89, 052915 (2014)
26. L. Wu, J.-F. Zhang, L. Li, Modulational instability and bright solitary wave solution for Bose-
Einstein condensates with time-dependent scattering length and harmonic potential. New J.
Phys. 9, 69 (2007)
27. A. Ankiewicz, D.J. Kedziora, N. Akhmediev, Rogue wave triplets. Phy. Lett. A 375, 2782
(2011)
28. H. Saito, M. Ueda, Dynamically stabilized bright solitons in a two-dimensional Bose-Einstein
condensate. Phys. Rev. Lett. 90, 040403 (2003)
29. F.Kh. Abdullaev, J.C. Bronski, R.M. Galimzyanov, Dynamics of a trapped 2D Bose-
Einstein condensate with periodically and randomly varying atomic scattering length. cond-
mat/0205464
30. H.M. Li, F.M. Wu, Soliton solutions of Bose-Einstein condensate in linear magnetic field and
time-dependent laser field. Chin. Phys. Lett. 21, 1425 (2004)
31. Q. Yang, J.-F. Zhang, Bose-Einstein solitons in time-dependent linear potential. Opt. Commun.
258, 35 (2006)
32. E. Wamba, T.C. Kofané, A. Mohamadou, Matter-wave solutions of Bose-Einstein condensates
with three-body interaction in linear magnetic and time-dependent laser fields. Chin. Phys. B
21, 070504 (2012)
33. S. Burger, K. Bongs, S. Dettmer, W. Ertmer, K. Sengstock, A. Sanpera, G. V. Shlyapnikov,
M. Lewenstein, Dark solitons in Bose-Einstein condensates. Phys. Rev. Lett. 83, 5198 (1999);
J. Denschlag, J.E. Simsarian, D.L. Feder, Charles W. Clark, L.A. Collins, J. Cubizolles, L.
Deng, E.W. Hagley, K. Helmerson, W.P. Reinhardt, S.L. Rolston, B.I. Schneider, W.D. Phillips,
Generating solitons by phase engineering of a Bose-Einstein. Science 287, 97 (2000)
34. Th. Busch, J.R. Anglin, Motion of dark solitons in trapped Bose-Einstein condensates. Phys.
Rev. Lett. 84, 2298 (2000); C.K. Law, C.M. Chan, P.T. Leung, M.-C. Chu, Motional dressed
states in a Bose-Einstein condensate: superfluidity at supersonic speed. Phys. Rev. Lett. 85,
1598 (2000); S.-W. Song, D.-S. Wang, H. Wang, W.M. Liu, Generation of ring dark solitons
by phase engineering and their oscillations in spin-1 Bose-Einstein condensates. Phys. Rev. A
85, 063617 (2012)
324 8 Rogue Matter Waves in Bose-Einstein Condensates …

35. B.P. Anderson, P.C. Haljan, C.A. Regal, D.L. Feder, L.A. Collins, C.W. Clark, E.A. Cornell,
Watching dark solitons decay into vortex rings in a Bose-Einstein condensate. Phys. Rev. Lett.
86, 2926 (2001); B.P. Anderson, Dark Solitons in BECs: the first experiments. In: Emergent
Nonlinear Phenomena in Bose-Einstein Condensates. Atomic, Optical, and Plasma Physics,
ed. by P.G. Kevrekidis, D.J. Frantzeskakis, R. Carretero-González, vol. 45 (Springer, Berlin,
Heidelberg)
36. T. Bland, K. Pawł owski, M.J. Edmonds, K. Rzazewski, N.G. Parker, Interaction-sensitive
oscillations of dark solitons in trapped dipolar condensates. Phys. Rev. A 95, 063622 (2017);
Th. Busch, J.R. Anglin, Motion of dark solitons in trapped Bose-Einstein condensates. Phys.
Rev. Lett. 84, 2298 (2000)
37. W.-M. Liu, E. Kengne, Schrödinger Equations in Nonlinear Systems (Springer Nature Singa-
pore Pte Ltd., 2019); E. Kengne, A. Sheou, A. Lakhssassi, Dynamics of matter-wave solitons
in Bose-Einstein condensates with time-dependent scattering length and complex potentials.
Eur. Phys. J. B 89, 1 (2016)
38. E. Kengne, W.-M. Liu, Management of matter-wave solitons in Bose-Einstein condensates with
time-dependent atomic scattering length in a time-dependent parabolic complex potential. Phys.
Rev. E 98, 012204 (2018); Z.X. Liang, Z.D. Zhang, W.-M. Liu, Dynamics of a bright soliton
in Bose-Einstein condensates with time-dependent atomic scattering length in an expulsive
parabolic potential. Phys. Rev. Lett. 94, 050402 (2005)
39. K.E. Strecker, G.B. Partridge, A.G. Truscott, R.G. Hulet, Nature 417, 150 (2002); L.
Khaykovich, F. Schreck, G. Ferrari, T. Bourdel, J. Cubizolles, L.D. Carr, Y. Castin, C. Salomon,
Formation of a matter-wave bright soliton. Science 296, 1290 (2002)
40. P.G. Kevrekidis, G. Theocharis, D.J. Frantzeskakis, B.A. Malomed, Feshbach resonance man-
agement for Bose-Einstein condensates. Phys. Rev. Lett. 90, 230401 (2003); E. Kengne, A.
Lakhssassi, Matter-wave solitons of Bose–Einstein condensates with time-dependent complex
potentials. Inter. J. Mod. Phys. B 32, 1850184 (2018)
41. E. Kengne, W.-M. Liu, Modulational instability and soliton control in a cubic-quintic dis-
sipative Gross-Pitaevskii equation with distributed coefficients. J. Phys. B: At. Mol. Opt.
Phys. 53, 215003 (2020); E. Kengne, A. Lakhssassi, Compensation process and generation
of chirped femtosecond solitons and double-kink solitons in Bose–Einstein condensates with
time-dependent atomic scattering length in a time-varying complex potential. Nonlinear Dyn.
104, 4221 (2021)
42. A. Di Carli, C.D. Colquhoun, G. Henderson, S. Flannigan, G.L. Oppo, A.J. Daley, S. Kuhr, E.
Haller, Excitation modes of bright matter-wave solitons. Phys. Rev. Lett. 123, 123602 (2019)
43. O.V. Marchukov, B.A. Malomed, M. Olshanii, J. Ruhl, V. Dunjko, R.G. Hulet, V.A. Yurovsky,
Quantum fluctuations of the center of mass and relative parameters of nonlinear Schrödinger
breathers. Phys. Rev. Lett. 125, 050405 (2020); D. Luo, Y. Jin, J.H.V. Nguyen, B.A. Mal-
omed, O.V. Marchukov, V.A. Yurovsky, V. Dunjko, M. Olshanii, R.G. Hulet, Creation and
characterization of matter-wave breathers. Phys. Rev. Lett. 125, 183902 (2020)
44. E. Kengne, Rogue waves of the dissipative Gross–Pitaevskii equation with distributed coef-
ficients. Eur. Phys. J. Plus 135, 622 (2020); E. Kengne, A. Lakhssassi, W.-M. Liu, Non-
autonomous solitons in inhomogeneous nonlinear media with distributed dispersion. Nonlinear
Dyn. 97, 449 (2019); W.-R. Sun, L. Wang, Matter rogue waves for the three-component Gross–
Pitaevskii equations in the spinor Bose–Einstein condensates. Proc. R. Soc. A. 474, 20170276
(2018)
45. L. Wen, L. Li, Z.D. Li, S.W. Song, X.F. Zhang, W.-M. Liu, Matter rogue wave in Bose-Einstein
condensates with attractive atomic interaction. Eur. Phys. J. D 64, 473(2011); W.-R. Sun, B.
Tian, Y. Jiang, H.-L. Zhen, Rogue matter waves in a Bose-Einstein condensate with the external
potential. Eur. Phys. J. D 68, 282 (2014); L. Li, F. Yu, Non-autonomous multi-rogue waves for
spin-1 coupled nonlinear Gross-Pitaevskii equation and management by external potentials.
Sci. Rep. 7, 10638 (2017)
46. B. Eiermann, T. Anker, M. Albiez, M. Taglieber, P. Treutlein, K.P. Marzlin, M.K. Oberthaler,
Bright Bose-Einstein gap solitons of atoms with repulsive interaction. Phys. Rev. Lett. 92,
References 325

230401 (2004); X. Zhua, H. Li, Z. Shi, Defect matter-wave gap solitons in spin–orbit-coupled
Bose–Einstein condensates in Zeeman lattices. Phys. Lett. A 23, 3253 (2016)
47. L. Zeng, J. Zeng, Gap-type dark localized modes in a Bose–Einstein condensate with optical
lattices. Adv. Photon. 1, 046004 (2019); V. Delgado, A. Muñoz Mateo, Stable symmetry-
protected 3D embedded solitons in Bose–Einstein condensates. Sci. Rep. 8, 10940 (2018); R.
Ravisankar, T. Sriraman, L. Salasnich, P. Muruganandam, Quenching dynamics of the bright
solitons and other localized states in spin–orbit coupled Bose–Einstein condensates. J. Phys.
B: At. Mol. Opt. Phys. 19, 195301 (2020)
48. O. Morsch, M. Oberthaler, Dynamics of Bose-Einstein condensates in optical lattices. Rev.
Mod. Phys. 78, 179–215 (2006)
49. M.R. Matthews, B.P. Anderson, P.C. Haljan, D.S. Hall, C.E. Wieman, E.A. Cornell, Vortices
in a Bose-Einstein condensate. Phys. Rev. Lett. 83 , 2498 (1999); N.V. Hung, P. Szańkowski,
V.V. Konotop, M. Trippenbach, Four-wave mixing in spin–orbit coupled Bose–Einstein con-
densates. J. Phys. B: At. Mol. Opt. Phys. 22, 053019 (2020)
50. J.P. Burke Jr., P.S. Julienne, C.J. Williams, Y.B. Band, M. Trippenbach, Four-wave mix-
ing in Bose-Einstein condensate systems with multiple spin states. arXiv.org cond-mat,
arXiv:cond-mat/0404499; O. Danaci, C. Rios, R.T. Glasser, Generation of self-healing beams
via four-wave mixing optical mode conversion. Proceedings 9950, Laser Beam Shaping XVII,
99500D (2016)
51. D.S. Petrov, Quantum mechanical stabilization of a collapsing Bose-Bose mixture. Phys. Rev.
Lett. 115, 155302 (2015); D.S. Petrov, G.E. Astrakharchik, Ultradilute low-dimensional liq-
uids. Phys. Rev. Lett. 117, 100401 (2016); M. Tylutki, G.E. Astrakharchik, B.A. Malomed,
D.S. Petrov, Collective excitations of a one-dimensional quantum droplet. Phys. Rev. A 101,
051601(R) (2020)
52. C.R. Cabrera, L. Tanzi, J. Sanz, B. Naylor, P. Thomas, P. Cheiney, L. Tarruell, Quantum liquid
droplets in a mixture of Bose-Einstein condensates. Science 359, 301 (2018); P. Cheiney, C.R.
Cabrera, J. Sanz, B. Naylor, L. Tanzi, L. Tarruell, Bright soliton to quantum droplet transition
in a mixture of Bose-Einstein condensates, Bright soliton to quantum droplet transition in a
mixture of Bose-Einstein condensates. Phys. Rev. Lett. 120, 135301 (2018); G. Semeghini,
G. Ferioli, L. Masi, C. Mazzinghi, L. Wolswijk, F. Minardi, M. Modugno, G. Modugno, M.
Inguscio, M. Fattori, Self-bound quantum droplets in atomic mixtures. Phys. Rev. Lett. 120,
235301 (2018)
53. G. Ferioli, G. Semeghini, L. Masi, G. Giusti, G. Modugno, M. Inguscio, A. Gallemí, A. Recati,
and M. Fattori, Collisions of self-bound quantum droplets, Phys. Rev. Lett. 122, 090401 (2019);
C. D’ Errico, A. Burchianti, M. Prevedelli, L. Salasnich, F. Ancilotto, M. Modugno, F. Minardi,
C. Fort, Observation of quantum droplets in a heteronuclear bosonic mixture. Phys. Rev. Res.
1, 033155 (2019)
54. I. Ferrier-Barbut, H. Kadau, M. Schmitt, M. Wenzel, T. Pfau, Observation of quantum droplets
in a strongly dipolar Bose gas. Phys. Rev. Lett. 116, 215301 (2016); L. Chomaz, S. Baier, D.
Petter, M.J. Mark, F. Wächtler, L. Santos, F. Ferlaino, Quantum-fluctuation-driven crossover
from a dilute Bose-Einstein condensate to a macrodroplet in a dipolar quantum flui. Phys. Rev.
X 6, 041039 (2016)
55. V.E. Zakharov, S.V. Manakov, S.P. Novikov, L.P. Pitaevskii, Solitons: Inverse Scattering Method
(Nauka Publishers, Moscow, 1980 (Consultants Bureau, New York, English translation, 1984)
56. M.J. Ablowitz, P.A. Clarkson, Solitons, Nonlinear Evolution Equations and Inverse Scattering
(Cambridge University Press, New York, 1991)
57. J. P. Ying, Fission and fusion of solitons for the (1+1)-dimensional Kupershmidt Equation.
Commun. Theor. Phys. 35, 405 (2001); C. Bai, H. Zhao, The study of soliton fission and fusion
in (2+ 1)-dimensional nonlinear system. Eur. Phys. J. D 39, 93 (2006)
58. S. Wang, X.Y. Tang, S.Y. Lou, Soliton fission and fusion: Burgers equation and Sharma–Tasso–
Olver equation. Chaos, Solitons & Fractals, 21, 231 (2004); Y.-L. Ma, B.-Q. Li, Mixed lump
and soliton solutions for a generalized (3+ 1)-dimensional Kadomtsev-Petviashvili equation.
AIMS Math. 5, 1162 (2020)
326 8 Rogue Matter Waves in Bose-Einstein Condensates …

59. A. Ankiewicz, A. Chowdury, Superposition of solitons with arbitrary parameters for higher-
order equations. Zeitschrift für Naturforschung A 71, 647 (2016); J. Lin, Y.S. Xu, F.M. Wu,
Evolution property of soliton solutions for the Whitham-Broer-Kaup equation and variant
Boussinesq equation. Chin. Phys. 12, 1049 (2003); W. Rui and Y. Zhang, Adv. Differ. Equ.
2020, 195 (2020)
60. E. Kengne, W.-M. Liu, B.A. Malomed, Spatiotemporal engineering of matter-wave solitons in
Bose-Einstein condensates. Phys. Rep. 899, 1–62 (2021)
61. Yu.V. Bludov, V.V. Konotop, N. Akhmediev, Matter rogue waves. Phys, Rev. A 80, 033610
(2009)
62. S. Chen, F. Baronio, J.M. Soto-Crespo, P. Grelu, D. Mihalache, Versatile rogue waves in scalar,
vector, and multidimensional nonlinear systems. J. Phys. A: Math. Theor. 50, 463001 (2017)
63. B. Kibler, A. Chabchoub, A. Gelash, N. Akhmediev, V.E. Zakharov, Superregular breathers
in optics and hydrodynamics: omnipresent modulation instability beyond simple periodicity.
Phys. Rev. X 5, 041026 (2015)
64. J.M. Dudley, G. Genty, F. Dias, B. Kibler, N. Akhmediev, Modulation instability, Akhmediev
breathers and continuous wave supercontinuum generation. Opt. Express 17, 21497 (2009);
E. Kengne, Modulational instability and soliton propagation in an alternate right-handed and
left-handed multi-coupled nonlinear dissipative transmission network. Chaos Solitons Fractals
146, 110866 (2021)
65. K. Manikandan, P. Muruganandam, M. Senthilvelan, M. Lakshmanan, Manipulating matter
rogue waves and breathers in Bose-Einstein condensates. Phys. Rev. E 90, 062905 (2014);
A.R. Osborne, Nonlinear Ocean Waves (Academic Press, New York, 2009)
66. Yu. V. Bludov, V. V. Konotop, and N. Akhmediev, Matter rogue waves. Phys, Rev. A 80, 033610
(2009)
67. V.R. Kumar, R. Radha, M. Wadati, Phase engineering and solitons of Bose-Einstein condensates
with two and three-body interactions. J. Phys. Soc. Jap. 79, 074005 (2010); E. Kengne, W.-M.
Liu, Engineering rogue waves with quintic nonlinearity and nonlinear dispersion effects in a
modified Nogochi nonlinear electric transmission network. Phys. Rev. E 102, 012203 (2020)
68. E. Kengne, W.M. Liu, Phase engineering chirped super rogue waves in a nonlinear transmission
network with dispersive elements. Adv. Theory Simul. 2021, 2100062 (2021); S.-W. Song, D.-
S. Wang, H. Wang, W.M. Liu, Generation of ring dark solitons by phase engineering and their
oscillations in spin-1 Bose-Einstein condensates. Phys. R. A 85, 063617 (2012)
69. A. Gammal, T. Frederico, L. Tomio, Ph. Chomaz, Atomic Bose-Einstein condensation with
three-body interactions and collective excitations. J. Phys. B 33, 4053 (2000)
70. F.K. Abdullaev, A. Gammal, L. Tomio, T. Frederico, Stability of trapped Bose-Einstein con-
densates. Phys. Rev. A 63, 043604 (2001)
71. G. P. Agrawal, Nonlinear Fiber Optics, 4th ed. (Academic Press, 2007); Yu.S. Kivshar, G.P.
Agrawal, Optical Solitons: From Fibers to Photonic Crystals (Academic, 2003)
72. A.S. Reyna, C.B. de Araújo, High-order optical nonlinearities in plasmonic nanocomposites–a
review. Adv. Opt. Phot. 9, 720 (2017)
73. V.M. Perez-Garcia, H. Michinel, H. Herrero, Bose-Einstein solitons in highly asymmetric traps.
Phys. Rev. A 57, 3837 (1998); U. Al. Khawaja, Lax pairs of time-dependent Gross–Pitaevskii
equation. J. Phys. A 39, 9679 (2006)
74. E. Kengne, P.K. Talla, Dynamics of bright matter wave solitons in Bose–Einstein condensates
in an expulsive parabolic and complex potential. J. Phys. B 39, 3679 (2006); A. Mohamadou,. E.
Wamba, S.Y. Doka, T.B. Ekogo, T.C. Kofane, Generation of matter-wave solitons of the gross-
Pitaevskii equation with a time-dependent complicated potentia. Phys. Rev. A 84, 023602
(2011)
75. V.I. Kruglov, A.C. Peacock, J.D. Harvey, Exact self-similar solutions of the generalized nonlin-
ear Schrödinger equation with distributed coefficients. Phys. Rev. Lett. 90, 113902 (2003); M.
Desaix, L. Helczynski, D. Anderson, M. Lisak, Propagation properties of chirped soliton pulses
in optical nonlinear Kerr media. Phys. Rev. E 65, 056602 (2002); H. Triki, Anjan Biswas, D.
Milović, M. Beliće, H. Triki, Anjan Biswas, D. Milović, M. Beliće, Opt. Commun. 366, 362
(2016). Opt. Commun. 366, 362 (2016)
References 327

76. A.A. Goyal, R. Gupta, C.N. Kumar, Chirped femtosecond solitons and double-kink solitons in
the cubic-quintic nonlinear Schrö dinger equation with self-steepening and self-frequency shift.
Phys. Rev. A 84, 063830 (2011); H. Kumar, F. Chand, Dark and bright solitary wave solutions
of the higher order nonlinear Schrödinger equation with self-steepening and self-frequency
shift effects. J. Nonlinear Opt. Phys. Mater. 22, 1350001 (2013)
77. B.A. Malomed, Soliton Management in Periodic Systems (Springer, New York, 2006)
78. J. Cuevas, P.G. Kevrekidis, B.A. Malomed, P. Dyke, R.G. Hulet, Interactions of solitons with
a Gaussian barrier. New J. Phys. 15, 063006 (2013)
79. S. Burger, K. Bongs, S. Dettmer, W. Ertmer, K. Sengstock, A. Sanpera, G.V. Shlyapnikov, M.
Lewenstein, Dark solitons in Bose-Einstein Condensates. Phys. Rev. Lett. 83, 5198 (1999)
80. A. Kundu, Landau–Lifshitz and higher-order nonlinear systems gauge generated from nonlinear
Schrödinger-type equations. J. Math. Phys. 25, 3433 (1984); F. Calogero, W. Eckhaus, Inverse
problems nonlinear evolution equations, rescalings, model PDES and their integrability: I.
Inverse Probl. 3, 229 (1987)
81. R.S. Johnson, On the modulation of water waves in the neighbourhood of kh 1.363. Proc. R.
Soc. Lond. A 357, 131 (1977); P.A. Clarkson, J.A. Tuszynski, Exact solutions of the multidi-
mensional derivative nonlinear Schrodinger equation for many-body systems of criticality. J.
Phys. A 23, 4269 (1990); D. Qiu, J. He, Y. Zhang, K. Porsezian, The Darboux transformation
of the Kundu–Eckhaus equation. Proc. R. Soc. A. 471, 0236 (2015); A. Bekir, E.H.M. Zahran,
Optik 223, 165233 (2020)
82. X. Wang, B. Yang, Y. Chen, Y. Yang, Higher-order rogue wave solutions of the Kundu-Eckhaus
equation. Phys. Scr. 89, 095210 (2014)
83. E. Kengne, W.-M. Liu, Transmission of rogue wave signals through a modified Noguchi elec-
trical transmission network. Phys. Rev. E 99, 062222 (2019); E. Kengne, W.-M. Liu, Exact
solutions of the derivative nonlinear Schrödinger equation for a nonlinear transmission line.
Phys. Rev. E 73, 026603 (2006)
84. A. Mohamadou, E. Wamba, S.Y. Doka, T.B. Ekogo, T.C. Kofané, Generation of matter-wave
solitons of the Gross-Pitaevskii equation with a time-dependent complicated potential. Phys.
Rev. A 84, 023602 (2011); E. Kengne, C. Tadmon, T. Nguyen-Ba, R. Vaillancourt, Higher
order bright solitons and shock signals in nonlinear transmission lines. Chin. J. Phys. 47 (5),
698703 (2009)
85. F. Baronio, S. Chen, Ph. Grelu, S. Wabnitz, M. Conforti, Baseband modulation instability as the
origin of rogue waves. Phys. Rev. A 91, 033804 (2015); F. Baronio, M. Conforti, A. Degasperis,
S. Lombardo, M. Onorato, S. Wabnitz, Vector rogue waves and baseband modulation instability
in the defocusing regime. Phys. Rev. Lett. 113, 034101 (2014)
86. H. Saito, M. Ueda, Dynamically stabilized bright solitons in a two-dimensional Bose-Einstein
condensate. Phys. Rev. Lett. 90, 040403 (2003); G.S Chong, H. HaiW, Q.T Xie, Breathing
bright solitons in a Bose-Einstein condensate. Chin. Phys. Lett. 20, 2098 (2003)
87. S. Rajendran, P. Muruganandam, M. Lakshmanan, Bright and dark solitons in a quasi 1D
Bose-Einstein condensates modelled by 1D Gross-Pitaevskii equation with time-dependent
parameters. Phys. D 239, 366 (2010)
88. D.H. Peregrine, J. Austral, Water waves, nonlinear Schrödinger equations and their solutions.
Math. Soc. B: Appl. Math. 25, 16 (1983)
Chapter 9
Dynamics of Matter-Wave Solitons
in Multi-component Bose-Einstein
Condensates

Abstract Multi-coupled equations of Gross-Pitaevskii type that model binary (two-


component) BECs are considered. Employing various analytical techniques such as
the technique of Darboux transform (DT), exact analytical soliton solutions are pre-
sented. With the help of these soliton solutions, we investigate analytically the soliton
management and stability of binary BECs with time-varying intrinsic interactions
trapped in an external time-varying HO potential. The effects of time-modulated HO
potential and the modulated nonlinearity on the BECs stability are studied too.

9.1 Soliton Management in a Binary Bose-Einstein


Condensate

By means of an integrable system of coupled GP equations of the Manakov’s type


[1], we carry out in this Section the analytical investigation of dynamics of matter-
wave solitons in a binary BEC. Using the DT, we build exact analytical solutions
with/without the cross-phase modulation (XPM) interaction between the two com-
ponents for soliton set on top of a plane-wave background. The effects of XPM inter-
action between the two components are investigated. We show that in the presence
of XPM, the built solutions exhibit properties different from those in the single-
component GP.

9.1.1 The Model and Analysis

The mathematical model to be used in this Section is the following coupled one-
dimensional GP equations which governs the dynamics of two-component self-
attractive BECs with negative scattering lengths:

© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2022 329
E. Kengne and W. Liu, Nonlinear Waves,
https://doi.org/10.1007/978-981-19-6744-3_9
330 9 Dynamics of Matter-Wave Solitons in Multi-component …
 
∂φ1 1 ∂ 2 φ1 a21
i =− −g |φ |
1 +
2
|φ 2 |2
φ1 ,
∂t 2 ∂x2 a12
(9.1)
 
∂φ2 1 ∂ 2 φ2 a22
i =− −g |φ2 | + |φ1 | φ2 ,
2 2
∂t 2 ∂x2 a12

where a11 and a22 and a12 are respectively the negative scattering lengths of intra-
and inter-component atomic collisions, the spatial coordinate x is measured in units
of x0 ∼ 1 µm, time t is measured in units of mx02 /, with atomic mass m, φ j are

measured in units of n 0 with n 0 being the maximum density of the initial distribution
of the condensate, and the nonlinearity parameter (interaction constant) g is defined
as g = 4π n 0 x02 |a12 | .
When a11 = a22 = a12 , the coupled GP equations (9.1) are integrable [1]. Exact
two-component solution of system (9.1) accounting for a soliton interacting with a
plane wave in the special case a11 = a22 = a12 can be generated by means of DT
method:    
4iηG 1 2ηC j G 2 i
φ j = A j exp (iϕ) 1 + + √ exp ϕ ; (9.2)
F g 2

here, j = 1, 2,
 
F = D exp θ1 + A2 D exp (−θ1 ) − 2i A2 L − L sin ϕ1 + C exp (−θ2 ) , (9.3a)
G 1 = L exp θ1 − A L exp (−θ1 ) − |L| exp (−iϕ1 ) + A exp (iϕ1 ) ,
2 2 2
(9.3b)
   
i 1 i 1
G 2 = L exp − (ϕ1 + ϕ2 ) + (θ1 − θ2 ) + A2 exp (ϕ1 − ϕ2 ) − (θ1 + θ2 ) , (9.3c)
2 2 2 2
  1 
ϕ ≡ kx + g A21 + A22 − k 2 t, (9.3d)
2

with θ1 , θ2 , ϕ1 , ϕ2 , L , and D given as

1
θ1 ≡ M I x + [(ξ − k) M I + ηM R ] t − θ10 , (9.4)
2
θ2 ≡ ηx + ηξ t + θ10 , (9.5)
1
ϕ1 ≡ M R x + [(ξ − k) M R − ηM I ] t + ϕ10 , (9.6)
2
1 2 
ϕ2 ≡ ξx + ξ − η2 t − ϕ10 , (9.7)
2

L ≡ M R + ξ + k + i (M I + η) , M ≡ (k + ξ + iη)2 + A2 = M R + i M I , (9.8)

 
D ≡ |L|2 + A2 , A ≡ 4g A21 + A22 , C = |C1 |2 + |C2 |2 . (9.9)

All parameters k, θ10 , ϕ10 , A1 , and A2 are free real constants, while C1 and C2 are
two free complex constants to be chosen from the condition
9.1 Soliton Management in a Binary Bose-Einstein Condensate 331

A1 C1 + A2 C2 = 0. (9.10)

In the case of a vanishing CW background, that is, when the background amplitude
A1 = A2 = 0, solution (9.2) leads to
√ 
φ j = ηε j / g cosh θ exp (−iϕ2 ) , with θ = η (x + ξ ) − θ0 , j = 1, 2, (9.11)

where θ0 is a free real constant, and ε1 and ε2 are arbitrary complex constants sat-
isfying the relationship |ε1 |2 + |ε2 |2 = 1. Solution (9.11) is a stable two-component
soliton; the soliton speed, width, and amplitudes associated with the √exact solution
(9.11) are respectively Vsol = −ξ , wsol = η−1 , and As1,2 = ηε1,2 / g and satisfy
the relationship A2s1 + A2s2 = η2 /g. For the exact soliton solution (9.11), the pro-
portional to the number Q of atoms in the binary condensate (total norm of the
two-component soliton) is given as

+∞
  2
Q ≡ Q1 + Q2 = |φ1 |2 + |φ2 |2 d x = 2 |η| /g, with Q j = 2 |η| ε j /g.
−∞

Computing the momentum M and Hamiltonian H of the system, we obtain respec-


tively

+∞
i    
M ≡ M1 + M2 = − φ 1 φ1,x − φ1 φ 1,x + φ 2 φ2,x − φ1 φ 2,x d x = Vsol Q
2
−∞
+∞
1    2 2
H ≡ φ1,x 2 + φ2,x 2 − g |φ1 |2 + |φ2 |2 2 d x = M − g Q 3 .
2 2Q 24
−∞

Now, let us consider the situation of vanishing soliton amplitudes, A1s,2s = 0 (for
example, η = 0). In this special case, solution (9.2) reduces to the Stokes wave (alias
CW),
(φ1 , φ2 )T = (A1 , A2 )T exp [iϕ] ,

with amplitudes A1 and A2 , wavenumber k, and angular frequency


 

= g A21 + A22 − k 2 /2.

Therefore, the exact solution (9.2) describes a two-component soliton propagating


on a non-vanishing CW background. The above condition (9.10) is the condition
of the XPM interaction between the two components of the CW background. When
C1 = C2 = 0, the second term in solution (9.2) vanishes, and each soliton component
will be embedded in its own CW background (a more detailed consideration of this
special case is given below). In the general case when C1 C2 = 0, the second term in
332 9 Dynamics of Matter-Wave Solitons in Multi-component …

solution (9.2) does not vanish. We then conclude that condition (9.10) of the XPM
interaction implies that coefficients C1 and C2 depend on amplitudes A1 and A2 of
the CW.
For a better understanding of the behavior of solution (9.2), it is relevant to start
with the special case when each soliton component will be embedded in its own CW
background, that is, C1 = C2 = 0. In this special case, solution (9.2) reduces to

φ j = A j exp (iϕ) (1 − 2ηW ) , j = 1, 2, (9.12a)


 
1 a cosh θ1 + sin ϕ1 b sinh θ1 + c cos ϕ1
W ≡ −i , (9.12b)
A cosh θ1 + a sin ϕ1 cosh θ1 + a sin ϕ1

with
     
a = −i A L − L /D, b = A L + L /D, and c = A2 − |L|2 /D.

It is well seen that θ2 and ϕ2 are absent from expressions (9.12a) and (9.12b), and the
two components share a common shape. Solution (9.2) in the present special case
(C1 = C2 = 0) is then similar to that for the one-component BEC.
If M I = 0, equation (9.12a) leads to the total norm of the soliton component of
the solution, which coincides with the norm above the CW level:

+∞  
 2   η b2 + c2
|φ1 |2 − A1 + |φ2 |2 − A22 d x = I, (9.13)
g |M I |
−∞
+∞
η + A (cosh x) sin (Bx + )
I ≡ d x, (9.14)
cosh x + a sin (Bx + )
−∞

where  
= −(1/2)ηM I 1 + B 2 t + Bθ10 + ϕ10 and B = M R /M I .

In obtaining the above expression, we have used the following relations

|φ1 (±∞, t)|2 = A21 and |φ2 (±∞, t)|2 = A22

for asymptotic values of the fields. One can easily verify that integral (9.14) is free
of , meaning that the total norm of the soliton is a conserved quantity.
If now M I = 0, it is obvious that the soliton speed vanishes and the CW back-
ground completely traps the soliton [2]. In particular, setting ξ = −k and A2 > η2
yields M I = 0, and W in solution (9.12a) turns to

1 η cosh θ1 + A sin ϕ1 − i M R sinh θ1


W = , (9.15)
A A cosh θ1 + η sin ϕ1
9.1 Soliton Management in a Binary Bose-Einstein Condensate 333

with
θ1 ≡ (1/2)ηM R t − θ10 and ϕ1 ≡ M R (x − kt) + ϕ10 ,

where M R = A2 − η2 , and θ10 and ϕ10 are free real constants. It is evident that
expression (9.15) is periodic in variable x with period = 2π/M R , and aperiodic
in time t. For a better understanding of the modulational instability development
provided by the above solution, we assume that

ε = exp (−θ10 )  1 (9.16)

is a small amplitude of the initial perturbation that triggers the onset of the MI
and linearize solution (9.12a) at the initial time t = 0 with respect to ε, taking Eq.
(9.15) into account. We then obtain the following CW with a small modulational
perturbation added to it

φ j (x, 0) ≈ A j exp (iϕ) [ρ − εχ sin (M R x + ϕ10 )] , (9.17)

where
 
ρ ≡ A2 − 2η2 − 2iηM R /A2 with |ρ|2 = 1,
χ ≡ 4ηM R (M R − iη) /A3 .

Carrying out direct numerical simulations of the underlying equations (9.1) demon-
strates that the evolution of initial configuration (9.17) closely follows the exact
solution provided by Eqs. (9.12a) and (9.15) [2].

9.1.2 Results

It is important to remember that the above results have been obtained under the
Manakov integrable conditions a11 = a22 = a12 which, in reality, may not be exactly
satisfied. Since the difference of the scattering lengths in a BEC mixture of two
different hyperfine states of the same atomic species is very small, we can still use
the Manakov’s system as a good approximation. As an illustration,we have set

a11 = −1.03, a12 = −1, a22 = −0.97, (9.18)

and solve numerical the coupled GP system (9.1). The obtained results yield the
picture of the modulational instability development displayed in Fig. 9.1. Comparing
the exact solution when a11 = a22 = a12 demonstrates that the two solutions are
virtually indistinguishable.
Now, we consider solution (9.2) in a more general situation when (|A1 | + |A2 |)
(|C1 | + |C2 |) > 0. Under condition (9.10) of the XPM interaction, it is possible
334 9 Dynamics of Matter-Wave Solitons in Multi-component …

Fig. 9.1 Plots of the spatial profile of the exact solution ( 9.12a) [dotted lines] and plots of the
numerical results obtained under nearly integrable condition (9.18) [solid lines] (exact and numerical
plots completely overlap). Different plots are generated for the parameters taken as η = 0.8, k =
−0.1, g = 1, A1 = 1, A2 = 1.2, θ10 = 6, the latter one representing amplitude (9.16) of the initial
perturbation which triggers the onset of modulational instability, with ε = 2.4788 × 10−3 and
ϕ10 = 0. Reprint from Ref. [2], Copyright 2022, with permission from American Physical Society

to explicitly study the effects of the XPM interaction between the plane waves in
the two-component BEC. In this general case, for condition (9.10) to be satisfied,
√ √
we set C1 = 4 g A2 C and C2 = −4 g A1 C, where C = 0 is an arbitrary complex
constant. To analyze the following two representative situations in detail, we further
fixe ξ = −k in solution (9.2).
(i) Case A2 = η2  
In the present case, we have 4 A21 + A22 = A2s1 + A2s2 and solution (9.2) can be
written as follows
 √ 
φ j = − exp (iϕ) A j tanh (θ2 /2) + (−1) j 2 A3− j cosh−1 (θ2 /2) exp (iϕ3 ) ,
(9.19)

where
 
θ2 ≡ η (x − kt) + θ10 , ϕ ≡ kx + η2 /4 − k 2 /2 t, ϕ3 ≡ η2 t/8 + ϕ10 ,

and θ10 and ϕ10 are free real constants. For solution (9.19), the density distribution
is found to be  
|φ1 |2 + |φ2 |2 = A22 + A21 1 + sech2 (θ2 /2) .

As it is clearly seen from Fig. 9.2. solution (9.19) appears as a superposition of bright
and dark solitons, produced by the action of XPM. In this figure, we can see that
the solution with A1 A2 = 0 is a complex consisting of bright and dark solitons in
9.1 Soliton Management in a Binary Bose-Einstein Condensate 335

Fig. 9.2 Spatiotemporal distribution of densities of the two components in solution (9.19), with
phase difference π between the components. Different plots are generated for parameters taken
as k = −0.1, g = 1, A1 = 1, A2 = 0.8, θ10 = −2, and ϕ10 = 0. Reprint from Ref. [2], Copyright
2022, with permission from American Physical Society

the first

and second species, or vice versa. It follows from condition A2 = η2 that

|η| = 4g A21 + A22 ; this means that the soliton width can be controlled by the
amplitude of the CW background. Also, we can observe from plots of Fig. 9.2 a shift
of the soliton’s peak due to the action of XPM.
(ii) Case A2 > η2
In the present case, we have
 
4 A21 + A22 > A2s1 + A2s2 , (9.20)

and solution (9.2) can be written as

φ j = A j exp (iϕ) (1 − 2ηW1 ) − (−1) j 2η A3− j W2 , (9.21)

where
1 η cosh θ1 + A sin ϕ1 − i M R sinh θ1
W1 = , (9.22a)
A A cosh θ1 + η sin ϕ1 + A exp (−θ2 )
1 (M R + iη) exp [(−iϕ1 + θ1 ) /2] + A exp [(iϕ1 − θ1 )]
W2 =
A A cosh θ1 + η sin ϕ1 + A exp (−θ2 )

 
−θ2 + i (ϕ − ϕ2 )
× exp , (9.22b)
2

with

θ1 ≡ (η/2) M R t − θ10 , θ2 ≡ x − kt + θ20 , ϕ1 ≡ M R (x − kt) + ϕ10 ,


1 2  
ϕ2 ≡ −kx + k − η2 t − ϕ20 , and M R = A2 − η2 ;
2
336 9 Dynamics of Matter-Wave Solitons in Multi-component …

Fig. 9.3 Soliton profile at different times generated with the use of the exact solution (9.21) for η =
−1.5, k = −0.1, g = 1, A1 = 1, A2 = 1.2, θ10 = −6, ϕ10 = 0, and θ20 = −5, ϕ20 = 0. Reprint
from Ref. [2], Copyright 2022, with permission from American Physical Society

Fig. 9.4 Soliton profile at different times generated with the numerical solution of Eqs. (9.1) with
the nonlinearity constants taken as per Eq. (9.18). Other parameters are identical to those in Fig. 9.3.
Reprint from Ref. [2], Copyright 2022, with permission from American Physical Society

θ10 , θ20 , ϕ10 , and ϕ20 being free real constants. It follows from Eqs (9.22a) and (9.22b)
that W2 → 0 in the limit θ2 → ±∞, and W1 → W in the limit θ2 → +∞, and W1 →
0 in the limit θ2 → −∞. From the form of solution (9.12a) and expression (9.15),
we can then conclude that solution (9.21) describes partial modulational instability,
since the MI growth rate is restrained in the limit of θ2 → −∞, as illustrated by
Fig. 9.3. However, when the integrable condition a11 = a12 = a22 is not satisfied, the
numerical solution of the model equation (9.1) will conspicuously differ from the
exact solution (9.21), as we can see from Fig. 9.4.
9.1 Soliton Management in a Binary Bose-Einstein Condensate 337

Fig. 9.5 Soliton profile at different times generated with the general-form analytical solution ( 9.2).
Parameters are η = 1.5, ξ = −0.4, k = −0.5, g = 1, A1 = 1, A2 = 1.2, 10 = 5, 10 = 0, C1 = 1,
and C2 is determined by constraint (9.10). Reprint from Ref. [2], Copyright 2022, with permission
from American Physical Society

Finally, let us consider solution (9.2) in the general case. It then follows from Eqs.
(9.4) and (9.5) that the solution contains terms with different velocities,

V1 = −(1/2) [(ξ − k) + ηM R /M I ] and V2 = −ξ,

leading to splitting of the soliton part of the solution on top of the continuous wave
background in two wave packets, high-frequency one and low-frequency one, as we
can see from plots of Fig. 9.5 where exact analytical solution is obtained under the
integrable condition with ai j = 1. Also, we can observe from plots of Fig. 9.6 that
plots generated with the use of exact solution are compared to those obtained with
the numerical solution of Eqs. (9.1), with the normalized scattering lengths chosen
as in Eq. (9.18) when the integrable condition is not satisfied. It is clearly seen from
plots of Fig. 9.6 that the soliton part of the numerical solution, built on top of the
continuous wave background, also splits in two wave packets, stable high-frequency
and unstable low-frequency ones.

9.1.3 Conclusion and Discussions

We have considered a coupled GP equation which describes binary BECs. The condi-
tion of the XPM interaction is presented in an analytical form. Under the integrable
condition that leads to a Manakov system, exact analytical soliton solutions are
presented, in the form of a soliton embedded on a non-vanishing continuous wave
background. We found that under the integrable condition, the exact soliton solution
can be used to describe the MI development which in physical applications can be
338 9 Dynamics of Matter-Wave Solitons in Multi-component …

Fig. 9.6 Soliton profile at different times generated with the numerically found solution of Eqs.
(9.1) with a11 = −1.03, a12 = −1, a22 = −0.97. Other parameters are identical to those in Fig. 9.5.
Reprint from Ref. [2], Copyright 2022, with permission from American Physical Society

used to generate a soliton train. We also found that even if the integrable condition
is not satisfied, the wave complex can be generated as a robust one. We have shown
that interaction between two continuous wave amplitudes gives rise to a phase dif-
ference between the components and can cause splitting of the soliton and formation
of a complex pattern. Although our study was carried out for positive g, it is impor-
tant to note that the model system (9.1) is also integrable in the case of negative g
when either a11 = a22 = a12 , corresponding to a binary self-repulsive condensate,
or a11 = a22 = −a12 , corresponding to a mixture of two self-repulsive condensates
that attract each other [2–5].

9.2 Soliton Stability in Binary Bose-Einstein Condensate


Under Temporal Modulation

In this Section, the mathematical model to be considered is a coupled one-dimensional


GP equations whose nonlinearities exponentially decay with time. Such a coupled
GP system may be used to describe the dynamics of two-component BECs with
time-varying intrinsic attractive interactions trapped in an expulsive time-depending
HO potential. The integrable condition of the model is presented in an explicit form.
Although solitons associated with solutions of this mathematical model are subject
to decay, we show that the robustness of bright solitons can be enhanced, making
their lifetime longer. We show that a combination of the expulsive time-modulated
HO potential with the modulated nonlinearity may sustain stable BECs, while it
quickly decays in the time-independent potential. To confirm the analytical results,
numerical simulations on the model equation are considered. We also show that the
stability of BEC soliton is not destroyed by an addition of noise [6].
9.2 Soliton Stability in Binary Bose-Einstein Condensate Under Temporal Modulation 339

9.2.1 The Physical Model and the Lax Pair

The model to be studied here pertains to a two-component BEC, having equal atomic
masses and attractive interactions in both components [7, 8], trapped in an external
HO potential. The evolution of the mean-field wave functions of the setting is thus
described by a coupled 1D GP equation which, in the scaled form, can be written as
 
∂ψ1 1 ∂2 1 2
i = − + b11 |ψ1 | + b12 |ψ2 | +
1 (t)x ψ1 = 0,
2 2 2
∂t 2 ∂x2 2
(9.23)
 
∂ψ2 1 ∂2 1 2
i = − + b22 |ψ2 | + b21 |ψ1 | +
2 (t)x ψ2 = 0,
2 2 2
∂t 2 ∂x2 2

where ψ j denotes the mean-field wave function of the j-th component normalized
as follows
+∞ +∞
|ψ1 | d x = 1 and
2
|ψ2 |2 d x = N2 /N1 .
−∞ −∞

In Eq. (9.23), b j j = 4a j j Ni /r⊥ and b jk = 4a jk Ni /r⊥ are respectively the self-phase


modulation (SPM) and the cross-phase modulation (XPM) coefficients in which a j j
and a jk are the respective scattering lengths, while r⊥ is the transverse-component
radius. For the integrability purpose, the symmetric system is considered, with

b11 = b22 = b21 = b12 ≡ −g, and


2j (t) = ω2j (t)/ω⊥
2
,

where ω j and ω⊥ stand for respectively the frequencies of the trapping potential
in the longitudinal and transverse directions. Throughout this Section,√time t and
coordinate x are assumed to be measured in units of 2/ω⊥ and r⊥ = / (mω⊥ ),
respectively.
Next, we assume that the confining and expulsive signs of the potential correspond
to λ2 < 0 and λ2 > 0, respectively, and set


21 (t) =
22 (t) ≡ −λ2 (t);

moreover, we assume time-varying parameter g(t) of interaction and strength λ2 (t)


of the external potential. Under all the above assumptions, replacing t by 2t, the
model equation (9.23) takes the following form:

∂ψ j ∂ 2ψ j 
i + + 2g(t) ψ j 2 + ψ3− j 2 ψ j + λ2 (t)x 2 ψ j = 0, j = 1, 2. (9.24)
∂t ∂x2
It is clearly seen that the nonlinearity in the coupled GP equations (9.24) takes
the Manakov’s form [1], and therefore is an integrable system. In the following, we
340 9 Dynamics of Matter-Wave Solitons in Multi-component …

focus our attention to the case of positive time-varying parameter g(t) of interaction;
this corresponds to the attractive sign of the SPM and XPM interactions. Under the
above special integrable condition imposed on g(t) and λ(t), system (9.24) admits
a representation in the form of the Lax pair,

x + U  = 0, t + V  = 0, (9.25)

where  is the three-component Jost function with components φ1 , φ2 , φ3 , that is,


 = (φ1 , φ2 , φ3 )T , and U and V are two operators defined as
⎛ ⎞
iζ (t) Q 1 Q2
U = ⎝ −Q ∗1 −iζ (t) 0 ⎠ , (9.26)
−Q ∗2 0 −iζ (t)
⎛ ⎞
v11 v12 v13
V = ⎝ v21 v22 v23 ⎠ , (9.27)
v31 v32 v33 )

with
i i i ∂ Q1
v11 = −iζ 2 (t) + i(t)xζ (t) + Q 1 Q ∗1 + Q 2 Q ∗2 , v12 = (t)x Q 1 − ζ (t)Q 1 + ,
2 2 2 ∂x

i ∂ Q2 i ∂ Q1
v13 = (t)x Q 2 − ζ (t)Q 2 + , v21 = (t)x Q ∗1 + ζ (t)Q ∗1 + ,
2 ∂x 2 ∂x
i i
v22 = iζ 2 (t) − i(t)xζ (t) − Q 1 Q ∗1 , v23 = − Q 2 Q ∗1 ,
2 2
i ∂ Q ∗2 i
v31 = −(t)x Q ∗2 + ζ (t)Q ∗2 + , v32 = − Q 1 Q ∗2 ,
2 ∂x 2
 
i 1 i
v33 = iζ 2 (t) − i(t)xζ (t) − Q 2 Q ∗2 , Q 1 = √ ψ1 (x, t) exp (t)x 2 ,
2 g(t) 2
 
1 i
Q2 = √ ψ2 (x, t) exp (t)x . 2
g(t) 2

Using the compatibility condition ∂∂ ∂  2

x∂t
= ∂t∂ x
, we arrive to the following zero-
curvature equation
Ut − Vx + [U, V ] = 0,

which shows that the spectral parameter ζ (t) must be taken from the following
nonisospectral condition
  
ζ (t) = μ exp − (t)dt ; (9.28)

here, μ is a hidden complex constant and (t) is a free function of time, which can
be used to define the trap strength λ(t) as follows
9.2 Soliton Stability in Binary Bose-Einstein Condensate Under Temporal Modulation 341

d
λ2 (t) =  2 (t) − , (9.29)
dt

under the condition that the time-varying strength λ(t) of the trap and the time-
dependent interaction strength g(t) must satisfy the following “integrable condition”:
 2
dg d2g
λ (t)g (t) = 2
2 2
− g(t) . (9.30)
dt dt 2

Thus, the coupled 1D GP system (9.24) under the integrable condition (9.30) will
be integrable. In the special case of a time-independent trap, λ(t) = const ≡ c1 , Eq.
(9.30) admits the special solution g(t) = exp [c1 t]. In what follows, we assume the
integrable condition (9.30) to be satisfied. It is important to note that if the integrable
condition (9.30) is slightly violated, the result will depend on the accumulation of the
deviation from the integrability over a characteristic time scale, T , of the dynamical
regime. In other words, if the deviation from the integrability is characterized by
difference λ(t) from the value given in Eq. (9.30), the system will remain close to
the integrability under the following “nearly integrable condition”
T


λ(t)dt  1. (9.31)


0

9.2.2 Analytical and Numerical Results for Two-Component


Bright Solitons in the Integrable System

Under the integrable condition (9.30), we use the gauge-transformation approach to


look for bright soliton solutions of the coupled GP equations (9.24) as

2 β1 (t)  
ψ1(1) = √ ε(1) exp i −ξ1 + (t)x 2 /2 , (9.32a)
g(t) 1 cosh θ1
2 β2 (t)  
ψ2(1) = √ ε2(1) exp i −ξ1 + (t)x 2 /2 , (9.32b)
g(t) cosh θ1

where
 
θ1 = 2β1 x + 4 α1 β1 dt − 2δ1 , ξ1 = 2α1 x + 2 (α12 − β12 )dt − 2χ1 ,

with    
α1 = α10 exp (t)dt , β1 = β10 exp (t)dt ,
342 9 Dynamics of Matter-Wave Solitons in Multi-component …

Fig. 9.7 a, b Spatiotemporal evolution of bright solitons obtained with the use of the exact analytical
soliton solution of Eqs. (9.24) under the integrability condition (9.30), with g(t) taken as per Eq.
(9.33) (and λ2 = 1/16). c, d Spatiotemporal evolution of bright solitons generated with the use of
numerical solution for the same parameters. Reprint from Ref. [6], Copyright 2022, with permission
from Elsevier

while δ1 and χ1 are free constants, and ε1(1) and ε1(1) are coupling coefficients satisfying
the restriction 2 2
(1) (1)
ε1 + ε2 = 1.

As our first example, we consider the case of BEC systems with time-decaying
strength g(t) of the interaction, and adopt the dependence of the nonlinearity coef-
ficient having the form
g(t) = 0.5 exp [−0.25t] . (9.33)

For the nonlinearity parameter given in Eq. (9.33), Eq. (9.30) renders the external HO
potential time-independent and expulsive, with λ2 = 1/16 > 0. The corresponding
density profile obtained with the use of the exact analytical solution, given by Eqs.
(9.29) and (9.32a), (9.32b), is depicted in Figs. 9.7a, b. Its counterpart produced with
the use of the numerical solution of Eq. (9.24) is displayed in Figs. 9.7c, d. From
plots of these figures, we observe perfect agreement between the analytical and
numerical solutions. This agreement implies the stability of the analytical solution.
Replacing the strength g(t) of the time-modulated interaction given in Eq. (9.33) by
the following one
g(t) = 0.5 exp [−0.5t] (9.34)
9.2 Soliton Stability in Binary Bose-Einstein Condensate Under Temporal Modulation 343

Fig. 9.8 The same as in Fig. 9.7, but obtained with g(t) given by Eq. (9.34), with λ2 = 1/4. Reprint
from Ref. [6], Copyright 2022, with permission from Elsevier

Fig. 9.9 The same as in Fig. 9.7, but generated with g(t) given by Eq. (9.35), with λ2 = 0.81.
Reprint from Ref. [6], Copyright 2022, with permission from Elsevier
344 9 Dynamics of Matter-Wave Solitons in Multi-component …

Fig. 9.10 a, b The analytical soliton solution of Eqs. (9.24), given by Eqs. (9.32a) and (9.32b),
with g(t) and λ2 (t) taken as per Eqs. (9.36) and (9.37). c, d The numerically generated counterpart
of the same solution. Reprint from Ref. [6], Copyright 2022, with permission from Elsevier

in Eq. (9.30), we obtain a stronger expulsive potential with λ2 = 1/4, and the con-
densate will quickly spread out, as we can see from Fig. 9.8a, b generated with the
exact analytical solution, and its numerical counterpart in Fig. 9.8c, d. As seen in
Fig. 9.9, this trend continues if the time dependence (9.34 is replaced by

g(t) = 0.5 exp [−0.9t] , (9.35)

for which Eq. (9.30) yields with λ2 = 0.81.


In order to enhance the stability of the condensates, we can switch on the time
dependence of the HO strength, taking
 
g(t) = 0.5 exp −0.125t 2 ; (9.36)

solving Eq. (9.30) for this g(t) yields the following time-dependent strength of the
expulsive potential
1 t2
λ2 = + . (9.37)
4 16
Using the corresponding exact analytical solution generated by Eqs. (9.32a) and
(9.32b), we show in Fig. 9.10a, b the spatiotemporal evolution of bright solitons. Its
numerical counterpart is displayed in Fig. 9.10c, d. From these figures, we conclude
that the correctness and stability of the exact analytical solution is corroborated by
its numerical counterpart. Next, for the much steeper modulation of parameter g(t)
of interaction taken as
9.2 Soliton Stability in Binary Bose-Einstein Condensate Under Temporal Modulation 345

Fig. 9.11 The same as in Fig. 9.9, but for g(t) and λ2 (t) taken as per Eqs. (9.38) and (9.39). Note the
essential difference of the spatiotemporal shape of the solution in comparison with that displayed
in Fig. 9.9. Reprint from Ref. [6], Copyright 2022, with permission from Elsevier

 
g(t) = 0.5 exp −2.5t 2 , (9.38)

it follows from the integrable condition (9.30) that

λ2 = 5 + 25t 2 . (9.39)

Using the corresponding exact analytical solution given by Eqs. (9.32a) and (9.32b)
and its numerical counterpart, we show in Fig. 9.11 the spatiotemporal evolution of
bright solitons which demonstrates non-monotonous evolution in time. Indeed, for
both the analytical and numerical cases, the soliton’s wave fields shrink and then
expand.
Considering now the modulation of the interaction coefficient with the form
 
g(t) = 0.5 exp −12.5t 2 , (9.40)

the integrable condition (9.30) gives

λ2 = 5 + 25t 2 . (9.41)

The corresponding exact analytical solution is depiceted in Fig. 9.12. Comparing


plots of Fig. 9.11 with the corresponding plots of Fig. 9.12, we can see that the shape
of bright soliton remains qualitatively similar to that in Fig. 9.11.
Comparing the results obtained for BEC with attractive interactions trapped in a
time-varying expulsive trap with those resulting from BEC with attractive interactions
346 9 Dynamics of Matter-Wave Solitons in Multi-component …

Fig. 9.12 The same as in Fig. 9.10, but for g(t) and λ2 (t) given by Eqs. (9.40) and ( 9.41). Reprint
from Ref. [6], Copyright 2022, with permission from Elsevier

Fig. 9.13 The same as in Fig. 9.9c, d, but in the case when strong white random noise, with a
standard spectral width, is added to the simulations. Reprint from Ref. [6], Copyright 2022, with
permission from Elsevier

Fig. 9.14 The same as in Fig. 9.10c, d, but in the case when strong random noise is added to the
simulations. Reprint from Ref. [6], Copyright 2022, with permission from Elsevier
9.2 Soliton Stability in Binary Bose-Einstein Condensate Under Temporal Modulation 347

Fig. 9.15 The same as in Fig. 9.11c, d, but in the case when strong random noise is added to the
simulations. Reprint from Ref. [6], Copyright 2022, with permission from Elsevier

trapped in a time-independent expulsive trap, we conclude that the first one, adjusted
to the integrable system based on Eqs. (9.24) and (9.30), is more long-lived.
The stability of the exact soliton solution produced by Eqs. (9.24) and (9.30) can
also be confirmed by adding random noise to the simulations. Example of obtained
results are showed in Figs. 9.13, 9.14 and 9.15. As we can see from these figures,
the stability of the evolving condensates is not broken by the white noise. These
figures also reveal that the stability of the evolving condensates depend neither on
the particular correlation structure of the noise nor on its spectral width. This conclu-
sion confirms the robustness of a two-component condensate in the time-dependent
expulsive HO potential in comparison with its counterpart in the time-independent
trap. These last results indicate that the life span of the two-component BEC with the
time-modulated attractive interactions trapped in the time-dependent HO potential
can be increased. The findings of this Section may be realized experimentally in
condensates composed of 39 K [9], 85 Rb [8], and 7 Li [10] atoms.

9.2.3 Conclusion

In this Section, we have considered a coupled one-dimensional GP equation for


studying the dynamics of a two-component BEC with time-varying intrinsic attrac-
tive interactions trapped in an expulsive time-depending HO potential. The inte-
grable condition is presented explicitly in terms of the nonlinearity parameter and
the strength of the external potential. Under the found integrable condition, exact
analytical bright soliton solutions are built and used for studying the dynamics of
various binary BECs with time decaying interaction. We have showed that under the
integrable condition and for BEC in time-dependent expulsive HO potential, bright
solitons stay stable for a reasonably large interval of time, compared to the con-
densate in the time-independent expulsive HO potential. Analytical results and their
stability are confirmed by numerical simulations carried out on the model equation
in both the case when the noise is absent and the case when strong random noise is
added to the simulations.
348 9 Dynamics of Matter-Wave Solitons in Multi-component …

References

1. S.V. Manakov, On the theory of two-dimensional stationary self-focusing of electromagnetic


waves. Zh. Eksp. Teor. Fiz. 65, 505 (1973); Sov. Phys. JETP 38, 248 (1974)
2. L. Li, B.A. Malomed, D. Mihalache, W.M. Liu, Exact soliton-on-plane-wave solutions for
two-component Bose-Einstein condensates. Phys. Rev. E 73, 066610 (2006)
3. V.G. Makhankov, N.V. Makhaldiani, O.K. Pashaev, On the integrability and isotopic structure
of the one-dimensional Hubbard model in the long wave approximation. Phys. Lett. A 81, 161
(1981)
4. V.M. Pérez-García, J.B. Beitia, Symbiotic solitons in heteronuclear multicomponent Bose-
Einstein condensates. Phys. Rev. A 72, 033620 (2005)
5. S.K. Adhikari, Bright solitons in coupled defocusing NLS equation supported by coupling:
application to Bose-Einstein condensation. Phys. Lett. A 346, 179–185 (2005)
6. R. Radha, P.S. Vinayagam, J.B. Sudharsan, W.-M. Liu, B.A. Malomed, Engineering bright
solitons to enhance the stability of two-component Bose-Einstein condensates. Phys. Lett. A
379, 2977–2983 (2015)
7. E.R.I. Abraham, W.I. McAlexander, J.M. Gerton, R.G. Hulet, R. Côté, A. Dalgarno, Singlet
s-wave scattering lengths of 6 Li and 7 Li. Phys. Rev. A 53, R3713 (1996)
8. S.L. Cornish, N.R. Claussen, J.L. Roberts, E.A. Cornell, C.E. Wieman, Stable 85 Rb Bose-
Einstein condensates with widely tunable interactions. Phys. Rev. Lett. 85, 1795 (2000)
9. G. Roati, M. Zaccanti, C. D’Errico, J. Catani, M. Modugno, A. Simoni, M. Inguscio, G.
Modugno, 39 K Bose-Einstein condensate with tunable interactions. Phys. Rev. Lett. 99, 010403
(2007)
10. K.E. Strecker, G.B. Partridge, A.G. Truscott, R.G. Hulet, Bright matter wave solitons in Bose-
Einstein condensates. New J. Phys. 5, 73 (2003)
Chapter 10
Dynamics of Higher-Dimensional
Condensates with Time Modulated
Nonlinearity

Abstract The present Chapter deals generally with more general nonintegrable
models, and particularly, with higher-order GP equations, that may model the dynam-
ics of higher-dimensional Bose-Einstein condensates. These models are treated by
means of both the variational approximation (VA) and direct computational simula-
tions. We mainly focus our attention on two- and three-dimensional condensates with
time modulated nonlinearity. Effects of various parameters of the model equations
on the dynamics of matter wave solitons are investigated.

10.1 Dynamics of Two- and Three-Dimensional


Bose-Einstein Condensates with Time Modulated
Nonlinearities

In this Section, we study the dynamics of two- and three-dimensional condensates


with time modulated inter-atomic interactions that consist of one constant and one
time-periodic oscillating terms. Analytical investigations are carried out by means of
either the semi-analytical variational approximation or an averaging method; these
theoretical results are confirmed with systematic direct simulations of the GP equa-
tion [1–4, 6]. All these methods used in the two-dimensional case reveal the existence
of stable self-confined states when the external trap is absent; this results is in agree-
ment with similar results obtained early for (2+1)D spatial solitons in nonlinear optics
[7]. When the external trap potential is absent, the variational approximation in the
case of 3D condensates also predicts the existence of self-confined state [8]. With
the use of direct computatiuonal simulations, we demonstrate that the stability in the
3D free space is limited in time, eventually switching into collapse.

© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2022 349
E. Kengne and W. Liu, Nonlinear Waves,
https://doi.org/10.1007/978-981-19-6744-3_10
350 10 Dynamics of Higher-Dimensional Condensates with Time Modulated …

10.1.1 The Model and Variational Approximation (VA)

The mathematical model used in this Section is the following mean-field GP equation
without an external trapping potential, that describes single-particle wave function
in its usual form [9]:

 
∂ψ(r,t) 2 2 4π 2 as
i = − ∇ + g |ψ(r,t)|2 ψ(r,t) with g = , (10.1)
∂t 2m m

where as and m are respectively the atomic scattering length and atomic mass, and
g is the strength of the two-body inter-atomic interactions which is assumed to be
time modulated has the form

g = g0 + g1 sin (χ t) . (10.2)

In Eq. (10.2), g0 and g1 are respectively the amplitude of the dc and ac parts, while χ
is the ac-modulation frequency. In the following, we demonstrate in some detail that
the trapping potential may be replaced, in some special situation, by a combination
of the temporal modulation of the nonlinearity coefficient (10.2) with the dc and ac
parts as in Eq. (10.4).
Next, we introduce the following transformation

 = 2gn 0 /, t  = t, and r = r 2m/, (10.3)

where n 0 is the largest value of the condensate density. It is evident that employing
the transformation (10.3) with the model equation (10.1) and omitting the primes
lead to the following scaled equation for isotropic states in which only the radial
coordinate is kept:

 2 
∂ψ(r,t) ∂ D−1 ∂
i =− + ψ(r, t) − [λ0 + λ1 sin (ωt)] |ψ(r, t)|2 ψ(r, t),
∂t ∂r 2 r ∂r
(10.4)
where D is the spatial dimension (D = 2 or 3), λ0 ≡ −g0 / (), λ1 ≡ −g1 / (),
and ω ≡ χ / . It is important to notice that positive [negative] λ0 in Eq. (10.4)
correspond to the self-focusing [self-defocusing] nonlinearity. In what follows, we
set, without loss of generality, |λ0 | = 1 so that λ0 remains a sign-defining parameter.
We start our investigation by applying the VA to Eq. (10.4) [1, 7, 10–14]. For this
purpose, we should use the following Lagrangian density generating Eq. (10.4):
   
i ∂ψ ∗ ∂ψ ∗  ∂ψ 2 1
L (ψ) = ψ − ψ −   + λ(t) |ψ|4 , (10.5)
2 ∂t ∂t ∂r  2

where
10.1 Dynamics of Two- and Three-Dimensional Bose-Einstein … 351

λ(t) = λ0 + λ1 sin (ωt) ,

and the (∗ ) stands for the complex conjugation. Next, we choose the variational ansatz
for the wave function as the Gaussian [10]
 
r2 1
ψG (r, t) = A(t) exp − 2 + ib(t)r 2 + iδ(t), (10.6)
2a (t) 2

where A, a, b, and δ are the time-varying amplitude, width, chirp, and overall phase,
respectively.
Following the approach used in Ref. [12], we insert the ansatz (10.6) in the
Lagrangian density (10.5) and calculate the respective effective Lagrangian. We
then arrive to
∞

L eff = C D L ψg r D−1 dr, (10.7)


0

where C D = 2π or 4π in respectively the two-dimensional or three-dimensional


cases. Finally, we derive the evolution equations for various functional parameters
A(t), a(t), b(t), and δ(t) of ansatz (10.6) from the expression of L eff given by Eq.
(10.7) when using the corresponding Euler-Lagrange equations. In what follows, we
present the analytical and numerical results separately for the two-dimensional and
three-dimensional cases.

10.1.2 Two-Dimensional Case

In the two-dimensional case, theoretical studies are done by means of both the VA and
the averaging method, applied to the GP equation; theoretical results are confirmed by
direct numerical simulations. First of all, we note that the application of the averaging
method to the 2D equation (10.4) without using VA in the case of the high-frequency
modulation is possible [1, 6, 15, 16].

10.1.2.1 The Variational Approximation

For the two-dimensional GP equation, we calculate from Eq. (10.7) the effective
Lagrangian and arrive to
 
1 4 2 db 2 2 dδ 1
L (2D)
eff = π − a A − a A − A 2
− a 4 2 2
A b + λ(t)a 2 4
A . (10.8)
2 dt dt 4

Using the Euler-Lagrange equations obtained from Eq. (10.8), we find that the total
number of atoms N in the condensate is conserved:
352 10 Dynamics of Higher-Dimensional Condensates with Time Modulated …

πa 2 A2 ≡ N = const; (10.9)

the corresponding width a and chirp b are found to satisfy the differential system

da db 2 λ(t)N
= 2ab, = 4 − 2b2 − ,
dt dt a 2πa 4
and the following closed-form evolution equation for the width:

d 2a 2 (2 − λ(t)N /2π )
= , (10.10)
dt 2 a3
which by setting

≡ 2 [λ0 N /(2π ) − 2] , = −λ1 N /π, (10.11)

can be rewritten as
d 2a −
+ sin (ωt)
= . (10.12)
dt 2 a3
In the absence of the ac component, Eq. (10.12) conserves the energy,

E 2D = (1/2) (da/dt)2 −
a −2 .

Evidently, E 2D → −∞ [+∞] at the limit a → 0, if


> 0 [
< 0], meaning that in
the absence of the ac component, the two-dimensional pulse is expected to collapse
[spread out] at
> 0 [
< 0]. The situation
= 0 corresponds to the critical norm
[17] whose numerically exact value is N = 1.862 [18]; note that the variational
equation (10.11) yields N = 2 (if λ0 = +1) [11].
Let us consider the situation when the ac component of g(t) oscillates at a high
frequency; in this situation, we can set

a(t) = a + δa, with |δa|  |a| , (10.13)

where, a is a function of a slow time scale and δa is a rapidly varying function with
zero mean value. Under all these assumptions, we can integrate Eq. (10.12) using
the Kapitsa averaging method [1, 6, 15]. Inserting Eq. (10.13) into Eq. (10.12), we
obtain, after straightforward manipulations, the following system of ODEs for a and
δa:

d 2a

2
= −
a −3 + 6a −5 δa 2
− 3 δa sin (ωt)
a −4 , (10.14a)
dt
d2
δa = 3δa
a −4 + sin (ωt) a −3 , (10.14b)
dt 2
where · · ·
stands for averaging over period 2π/ω. Equation (10.14b) admits the
special solution
10.1 Dynamics of Two- and Three-Dimensional Bose-Einstein … 353

sin (ωt)
δa(t) = −
. (10.15)
a ω2 + 3a −4

Inserting now Eq. (10.15) into Eq. (10.14a), we obtain the following evolution equa-
tion for the slow variable a:

d 2a 1 3
2 3 2
= 3 −

2 + . (10.16)
dt 2 a ω2 a 4 + 3
2 ω2 a 4 + 3

In the limit a → 0, Eq. (10.16) reduces to the following one


 
d 2a 1 2 2 − 6
2
= 3 −
+ = . (10.17)
dt 2 a 6
6
a 3

It follows from Eq. (10.17) that if 2 > 6


2 (meaning that the amplitude of the high-
frequency ac component is large enough), then the behavior of the condensate in the
limit a → 0 is exactly opposite to that which would be expected in the presence of
the dc component only; this means that in the case of positive
, rebound occurs
rather than the collapse, and vice versa in the case of negative
.
Now, in the limit a → ∞ , Eq. (10.16) takes the asymptotic form

d 2a
= −
/a 3 ,
dt 2
showing that the condensate remains self-confined if
> 0 (in the case when
> 0,
the norm exceeds the critical value). The above asymptotic results guarantee the
stability of the behavior of the condensate for any a, solution of Eq. (10.16), the
collapse and spreading out being ruled out if

> 6
> 0. (10.18)

In the following, we illustrate the above results in terms of the experimentally


relevant setting for the condensate of 7 Li with the critical number ∼ 1500 atoms.
We then conclude that, for 1800 atoms, we must add a periodic modulation with
amplitude = 0.98 for the stabilization, as one can see from Eq. (10.11) where
λ0 = 1 has been used. Indeed, it follows from conditions (10.18) that for small
[large] a, the right-hand side (rhs) of Eq. (10.16) is positive [negative], that is, da dt
increases [decreases] with the time t increases, meaning that Eq. (10.16) must give
rise to a stable fixed point (FP). In fact, the rhs of Eq. (10.16) under condition (10.18)
vanishes at exactly one fixed point,
  
3 2 3 4
ω2 a 4 = + 3 − − 3
.
2 (10.19)
4
16
2
354 10 Dynamics of Higher-Dimensional Condensates with Time Modulated …

The stability of the FP (10.19) can be easily checked from the calculation of an
eigenfrequency of small oscillations around it.
Now, we turn to the direct numerical simulations of Eq. (10.12). Carrying out
direct numerical/computational simulations on Eq. (10.12), we obtain results that
are in exact agreement with those provided by the averaging method. In other words,
we obtain that stable states with α(t) performing small oscillations around point
(10.19) [1, 15]. We ask ourselves if the same results can be obtained in the 3D
situation. The answer to this question, as we will see in what follows, is no!
In 3D situation, we also needs, as in the above analysis, an approximate form of
Eq. (10.16) valid when the norm is closed to the critical value (that is, in the limit

→ 0), and for large enough ω. The obtained approximate form of Eq. (10.16)
reads
d 2a
3 2
= − + . (10.20)
dt 2 a3 2 ω2 a 7

Next, we estimate the value of the amplitude of the high-frequency ac component


necessary to stop the collapse. First, we note that in physical units, a characteristic
trap frequency is  ∼ 100 Hz. Then, we consider a high modulation frequency ∼ 3
kHz, and the scaled modulation frequency taken as ω 30. Assuming the norm
to be N /2π = 2.2, we find from Eq. (10.11) that
= 0.4; these numerical values
correspond to the condensate of 7 Li with 1800 atoms, the critical number being
1500, with modulation parameters taken as λ0 = 1, λ1 = 2.3, and = 10. From
Eq. √(10.19), we find the stationary value of the condensate width to be ast = 0.8l,
l = m/ being the healing length.
Based on the VA and assuming that the norm slightly exceeds the critical value,
the following important result is obtained:
Result: The combining ac and dc components of the two-body inter-atomic inter-
actions corresponding to the attraction in the two-dimensional Gross-Pitaevskii equa-
tion without the trapping term may replace the collapse by a stable solitonlike oscilla-
tory state that confines itself. A similar result was reported in Ref. [7] when studying
cylindrical solitons in a bulk nonlinear-optical medium.

10.1.2.2 Averaging of the Two-Dimensional Gross-Pitaevskii Equation


and Hamiltonian

Here, we assume the ac frequency ω to be large, and rewrite the two-dimensional


GP equation (10.4) in the simplified form

∂ψ
i + ∇ 2 ψ + λ(ωt) |ψ|2 ψ = 0. (10.21)
∂t
Next, we employ the multiscale approach for deriving the amplitude equation for
slow variations of the field. The approach consists of expanding the solution in a
power series of 1/ω. For this purpose, we introduce slow temporal variables Tk as
10.1 Dynamics of Two- and Three-Dimensional Bose-Einstein … 355

Tk ≡ ω−k t, k = 0, 1, 2, . . ., and the fast time is ζ = ωt. Focusing our attention to


the case when the dc part of the two-body inter-atomic interactions corresponds to
attraction between the atoms (λ0 = 1), we seek the solution of Eq. (10.21) in the
general form as
 2
1 1
ψ(r, t) = A(r, Tk ) + u 1 (ζ, A) + u 2 (ζ, A) + · · · , (10.22)
ω ω

with u k
= 0, where · · ·
stands for the average over the period of the rapid mod-
ulation.
Using the approach developed in Ref. [19], the first and second corrections are
found to be


u 1 = −i (μ1 − μ1
) |A| A, μ1 ≡
2
[λ(τ ) − λ1
] dτ,
0

u 2 = (μ2 − μ2
) 2i |A|2 At + i A2 A∗t + ∇ 2 (|A|2 A) (10.23)
 
1
− |A| A4
(μ1 − μ1
) − 2M + λ
(μ2 − μ2
) ,
2
2

where,



μ2 ≡ (μ1 − μ1
) ds, and M = (1/2) μ21
− μ1
2 = (1/2) λ2
− 1 .
0

With the use of these results, we derive at order (1/ω)2 to the following evolution
equation for A(x, T0 ), which is valid for both the 2D and 3D cases:

∂A  2

i + ∇ 2 A + |A|2 A + 2M |A|6 A − 3 |A|4 ∇ 2 A + 2 |A|2 ∇ 2 |A|2 A


∂t
ω
+ A2 ∇ 2 |A|2 A∗ = 0, (10.24)

being the same amplitude of the ac component as in Eq. (10.11). In both 2D and
3D cases, the quasi-Hamiltonian can be represented in the following form
  2 
4 ∂A δ Hq
1 + 6M |A| = −i ∗ , (10.25a)
ω ∂t δA
   2  2  
1

Hq = d V |∇ A|2 −2M |A|8 − |A|4 + 4M ∇ |A|2 A  ,
ω 2 ω
(10.25b)
356 10 Dynamics of Higher-Dimensional Condensates with Time Modulated …

where d V is the infinitesimal volume in the 2D or 3D space. It follows from Eqs.


(10.25a) and (10.25b) that d Hq /dt = 0.
For investigating the 2D case, we apply the modulation theory developed in Ref.
[20], and seek the solution in the form of a modulated TS:

ψ(r, t) = exp [it] RT (r ),

RT (r ) being any physical solution of the boundary-value problem,



d 2 RT 1 d RT d RT 
+ − RT + RT = 0,
3
= 0, RT (r )|r =∞ = 0. (10.26)
dr 2 r dr dr r =0

The corresponding norm N and the Hamiltonian H read

∞ ∞  2 
d RT 1 4
NT ≡ RT2 (r |)r dr = Nc ≈ 1.862, HT = − RT (r ) r dr = 0,
dr 2
0 0
(10.27)
respectively. Equation (10.24) shows an increase of the critical norm for the collapse,
as opposed to the classical value in Eq. (10.27). With the use of Eq. (10.22), we
compute the critical norm and find [21]

∞  2
Ncrit = |ψ|2 r dr = N T + 2IM , I = 11.178.
ω
0

Now, let us compute the phase chirp in the soliton. The mean value of the phase
chirp b in the soliton is found to be
∞
Im [(∂ψ/∂r ) ψ ∗ ] r dr
b= 0
∞ .
2
0 r dr
|ψ|2

By using expression (10.23) for the first correction, we arrive to


 ∞ 
∞ 2  2 4
3 0 r dr R (R ) − (1/4) 0 dr R
b = − ( /ω) B M (μ1 − μ1
) , B ≡ ∞ ≈ 0.596.
2 2
0 r dr R

For developing a general analysis, we assume that the solution with the norm
close to the critical value can be approximated as a modulated TS. In other words,
we assume that

r 2 da dσ
A(r, t) ≈ [a(t)]−1 RT [r/a(t)] exp (i S) , S = σ (t) + , = a −2 (10.28)
4a dt dt
10.1 Dynamics of Two- and Three-Dimensional Bose-Einstein … 357

with some function a(t). Assuming the initial norm to be close to the critical value
(|N − Nc |  Nc ) and using the same strategies used in Ref. [20], we derive the
following evolution equation for a(t):

d 2a 2
a3 = −β0 + f 1 (t), (10.29)
dt 2 4M0 ω2

where
∞
2 N − Nc 1
β0 = β(0) − f 1 (0), β(0) = , M0 ≡ r 3 dr RT2 ≈ 0.55,
4M0 ω 2 M0 4
0

f 1 (t) being an auxiliary function given as


   
1
f 1 (t) = 2a(t)Re d xd y F(A T ) exp [−i S] {RT + ρ∇ R R (ρ)} . (10.30)

For the harmonic modulation, equations of the lowest-order of approximation lead


to
d 2a
1 C 2
= − + , (10.31)
dt 2 a3 ω2 a 7
where

1 = (N − Nc ) /M0 − C 2 /(ω2 a04 )

and C is defined as [1]

∞  
3 
2 
3 1 8
C≡ dρ 2ρ RT RT − ρ RT RT − ρ RT ≈ 39.
4 2 3
(10.32)
M0 8
0

It then follows from the averaged equation that the collapse can be stopped by the
rapid modulations of the nonlinear term in the two-dimensional GP equation. Com-
paring Eq. (10.31) with its counterpart (10.20) derived by means of averaging equa-
tion (10.12), we find that when numerical coefficients in the second terms are different
due to the different profiles of the Gaussian and TS, then both approaches lead to the
same behavior near the collapse threshold.
Further, we estimate the FP as per numerical simulations performed in work
[22] where the following parameters were used, λ = 1 + at 0 < t < T , and λ =
1 − at T < t < 2T with the following numerical values: T = = 0.1, N /(2π ) =
11.726/(2π ), with the critical number Nc = 11.68/(2π ). Using these numerical
values, we obtain ac = 0.49, which agrees with the value ac ≈ 0.56 produced by
computational numerical simulations.
358 10 Dynamics of Higher-Dimensional Condensates with Time Modulated …

It is important to notice that based on representation (10.22) for the wave function,
one can also apply the averaging procedure directly to the Hamiltonian of Eq. (10.4).
The averaged Hamiltonian, which can be use to explain the possibility to arrest the
collapse in the presence of the rapid modulation of the nonlinearity strength [23], is
then found to be
    2   2 

2 1
H= d xd y |∇ A| + 2M
2  2 
∇ |A| A − |A| − 6M 4
|A| .
8
ω 2 ω
(10.33)
We note that, if a given field configuration has compressed itself to a spot with size
ρ, where A is ∼ ℵ, the conservation of norm N yields the relationship

ℵ2 ρ D ∼ N , (10.34)

where D is the space dimension. For the strongest collapse-driving and collapse-
arresting terms, the same estimate H− and H+ in the Hamiltonian yields
 2  2
H− ∼ − ℵ8 ρ D , H+ ∼ ℵ6 ρ D−2 . (10.35)
ω ω
If we employ relation (10.34) to eliminate the amplitude from Eq. (10.35), we con-
clude that H± asymptotically scale as ρ −5 in the limit ρ → 0 (case of the catastrophic
self-compression); depending on details of the configuration, the collapse may there-
fore be arrested. It is important to note that in the 3d situation, the collapse cannot
be prevented. Indeed, in the limit ρ → 0, the collapse-driving term in the 3D case
diverges as ρ −9 , while the collapse-arresting one scales ∼ ρ −8 .

10.1.2.3 Numerical Simulations

Here, we intend to check the existence of stable self-confined solitonlike oscillating


states by direct numerical simulations of the 2D equation (10.4).
The above analytical study predicts the existence of stable self-confined solitonlike
oscillating states in region (10.18) when
(i) the dc part of the nonlinearity corresponds to self-attraction and
(ii) the amplitude of the ac component is small enough.
Now, we intend to check this result by carrying out direct numerical simulations on
the 2D equation (10.4). A typical example of the obtained results is displayed in
Fig. 10.1 showing the formation of a self-confined condensate under a combination
of the self-focusing dc and sufficiently strong ac components of the nonlinearity
when the external trap is absent. Comparing Fig. 10.1a with Fig. 10.1b showing
respectively the collapsing state in the absence of the ac modulation at t ≈ 0.3 and
the radial profile of the stable state in the presence of the ac modulation formed by the
same input at t ≈ 0.6, it is clearly seen that when the ac term is taken into account,
the pulse is stabilized for about 40 ac-modulation periods, after which it decays [4].
10.1 Dynamics of Two- and Three-Dimensional Bose-Einstein … 359

Fig. 10.1 A typical example of the formation of a self-confined condensate in simulations of the
2D equation (10.4). a Collapsing state at t ≈ 0.3 when the ac modulation is absence; b radial
profile of the stable state formed by the same input at t ≈ 0.6, in the presence of the ac term in
the nonlinearity coefficient. Different plots are generated with the following numerical values of
parameters λ0 = 2.4, λ1 = 0.85, ω = 100π , and N = 5. Reprint from Ref. [1], Copyright 2022,
with permission from American Physical Society

10.1.3 The Three-Dimensional Gross-Pitaevskii Model

Similar to the 2D situation, we separately consider results produced by the analytical


approximations (VA and averaging method) and by direct numerical simulations in
the 3D case.

10.1.3.1 VA and Averaging

In the case of the 3D GP equation, we compute the effective Lagrangian (10.7) and
obtain [1, 5]
 
1 3 2 3 3 db dδ 1 3
L (3D)
eff = π 2 A a − a2 −2 + √ λ(t)A2 − 2 − 3b2 a 2 . (10.36)
2 2 dt dt 2 2 a

The corresponding Euler-Lagrange equations yields the norm conservation


3
π 2 A2 a 3 ≡ N = const,

the following evolution equation for the width a of the condensate,

d 2a 4 λ(t) N
= 3 − √ 3 4, (10.37)
dt 2 a 2 2π 2 a

and the following relations including the width a and the chirp b,
360 10 Dynamics of Higher-Dimensional Condensates with Time Modulated …

da db 2 λ(t)N
= 2ab, = 4 − 2b2 − √ 3 .
dt dt a 2 2π 2 a 5

Equation (10.37) for the wave width in the 3D situation is different from its counter-
part (10.10) corresponding to the 2D situation.
Following the same procedures as in the 2D case, the amplitudes of the dc and ac
components of the nonlinearity are normalized as follows
√ √

≡ λ0 N / 2π 3 and ≡ −λ1 N / 2π 3 .

Under such a normalization, equation (10.37) takes the following scaled form

d 2a 4 −
+ sin (ωt)
= 3+ . (10.38)
dt 2 a a4
Setting = 0 in Eq. (10.38) (case of the absence of the ac term), we arrive the
following equation for the energy conservation
 2
1 da 1
E 3D = + 2a −2 −
a −3 . (10.39)
2 dt 3

It follows from Eq. (10.39) that in the limit a → 0, E 3D → −∞ [+∞], if


> 0
[
< 0]; this corresponds respectively to the collapse or decay of the pulse.
Solving numerically Eq. (10.38) without averaging, we have found, as one can see
from Fig. 10.2, the existence of a region in the parameter space where the condensate,
which would decay for negative
, may be stabilized by the ac component, provided
that its amplitude is sufficiently large. Figure 10.2 is generated for different initial
conditions, a(t = 0) = 0.3, 0.2, or 0.13 and da/dt (t = 0) = 0. It is clearly seen from

Fig. 10.2 The Poincaré


section in the plane of
(a, da/dt ) for
= −1,
= 100, ω = 104 π ,
generated with the help of
the numerical solution of the
variational equation (10.38)
with different initial
conditions given in the text.
Reprint from Ref. [1],
Copyright 2022, with
permission from American
Physical Society
10.1 Dynamics of Two- and Three-Dimensional Bose-Einstein … 361

Fig. 10.3 The region in the ( , ω/π ) parameter plane where the numerical solution of Eq. (10.38)
with
= −1 predicts stable quasiperiodic solutions in the 3D case. Here, crosses mark points
show different positions where stable solutions were actually obtained, while stars correspond to
minimum values of the amplitude of the ac-component eventually leading to the collapse of the
solution of equation (10.4) with
= −1. Reprint from Ref. [1], Copyright 2022, with permission
from American Physical Society

Fig. 10.2 that independently on the initial conditions, the solution remains bounded
with quasiperiodic oscillations, avoiding the collapse or decay.
The corresponding stability region in the parameter plane (ω/π, ), generated with
the numerical solution is small, as we can see Fig. 10.3. Here, we can also see that for
maintaining the stability, the numerical values of the frequency and amplitude of the
ac component must be very large. It is important to note that for frequencies above
106 π , the condensate width a(t) becomes very small in the course of the evolution,
meaning that the collapse may happen in the solution of the full 3D equation (10.4).
Using the VA, we found that the stability is predicted only when
< 9. If
≥ 9,
the VA predicts solely the collapse.
Because the frequency ω is large enough in the stability region as we can see in
Fig. 10.3, we also apply the averaging method to this. For this aim, we follow the
above procedures done for the 2D case and find the rapidly oscillating correction
δa(t) to the solution as
sin [ωt] a
δa = − , (10.40)
ω2 a 5 − 12a + 4

while the resulting evolution equation for the slow variable a(t) is found to be

d 2a −4 2 2 6a − 5

=a 4a −
+
4
+
2

2 .
dt 2 ω2 a 5 − 12a + 4
ω2 a 5 − 12a + 4

(10.41)
362 10 Dynamics of Higher-Dimensional Condensates with Time Modulated …

In the limit a → 0, Eq. (10.41) turns to


 
d 2a 1 3 2
= 4 −
+ . (10.42)
dt 2 a 16

Equation (10.42) plays the same role as Eq. (10.17) in the 2D situation, but leads to
√ case of negative
and with large enough amplitude
only one property (result). In the
of the ac component ( > (4/ 3) |
|), we conclude from Eq. (10.42) that collapse
takes place instead of spreading out. Other results obtained from Eq. (10.17) in the
2D with the use of an equation similar to the averaged equation (10.41) are wrong
(see Figs. 10.2 and 10.3). In particular, analyzing in details the rhs of Eq. (10.41),
we see that it predict neither a stable FP for negative
nor for positive
, and
this contradicts what is revealed by the numerical simulations of Eq. (10.38). This
failure follows from the existence of singular points in Eqs. (10.40) and (10.41),
independently of the sign of
.

10.1.3.2 Direct Simulations of the GP Equation in the 3D Case

It is necessary to compare the analytical results to the numerical ones.


Ignoring the ac component ( = 0), the numerical simulations under the condition

< 0 show straightforward decay. In the presence of an ac component with a large


enough amplitude, transient stabilization of the condensate takes place, confirming
thus the analytical results obtained by means of the VA. Meanwhile, the stabilization
is not permanent since the collapse occurs after some periods.
A typical example of the obtained results for the 3D case with
< 0 is displayed
in Fig. 10.4 for N = 1,
= −1, and ω = 104 π. Plots of Fig. 10.4 show the evolution
of the density radial profiles |u(r )|2 at different times. It follows from different plots
of Fig. 10.4 that a basic characteristic of the system is a dependence of the minimum
, which gives rise to the collapse at fixed
= −1, versus ω. This dependence

Fig. 10.4 The evolution plot of the density radial profile, |u(r )|2 , in the presence of the strong
and fast ac modulation (ω = 104 π , = 90). The profiles of |u(r )|2 are shown at different times:
t = 0.007 (a), 0.01 (b), and 0.015 (c). Reprint from Ref. [1], Copyright 2022, with permission from
American Physical Society
10.2 Stable Vortex Modes in Two-Dimensional Bose-Einstein Condensates 363

is showed in Fig. 10.3 by many star points. It is evident that as a function of the
frequency ω, the minimum value of necessary for the collapse increases with the
increase in ω. The numerical simulations have also showed that the collapse in the
case
> 0 cannot be predicted.

10.1.4 Conclusion and Discussions

In this Section, we have considered as the model equation, a higher-dimensional GP


equation with a time-varying two-body inter-atomic interactions without an exter-
nal potential. Assuming the strength of the two-body inter-atomic interactions to be
formed of two parts, one constant (dc) and one time-varying (ac) parts, the model
equation is used to investigate analytically and numerically the dynamics of both
2D and 3D BECs. The effects of both the dc and ac on the condensates are studied.
We found that in the 2D case, with the ac component of the two-body inter-atomic
interactions, it is possible to maintain the condensate without an external trap in a
stable self-confined state. For the condensates without a trap to maintain a stable
self-confined state in the 3D situation, we found via the VA that the dc part of the
nonlinearity must correspond to repulsion between atoms; direct numerical simu-
lations reveals that the stability of the self-confined condensate in this case is just
partial, that is, limited in time. From our investigations, we conclude that the spa-
tially uniform ac magnetic field may play the role of an effective trap that confines
the condensate, and sometimes can enforce its collapse.

10.2 Stable Vortex Modes in Two-Dimensional


Bose-Einstein Condensates

In this Section, we consider a physical model governed by a two-dimensional GP


equation with spatial-varying parameters written in the scaled form. Exact analytical
vortex-soliton (VS) solutions and approximate fundamental soliton solutions for two-
dimensional BECs with a spatially modulated attractive and/or repulsive nonlinearity
when the external trapping potential is either absent or present. By means of the
discrete energy spectrum of a related linear Schrödinger equation, the number of
vortex-soliton modes is presented and the stability of built VS with vorticity higher
than or equal to two are reported [24, 25]. The number of VS solutions in the case of
BECs with attractive nonlinearity is found to be infinite, while for BEC systems with
repulsive nonlinearity, there exists no more than a finite number of VS solutions.
364 10 Dynamics of Higher-Dimensional Condensates with Time Modulated …

10.2.1 Model Description and Main Transformation

The mathematical model used for investigating stable vortex modes for BECs with
a spatially modulated attractive and repulsive nonlinearities trapped in a harmonic
potential is the two-dimensional GP equation whose scaled form reads

∂ψ
i = −∇ 2 + g(r ) |ψ|2 + V (r ) ψ. (10.43)
∂t
Physically, the GP equation (10.43) describes the BECs macroscopic wave func-
tion ψ = ψ(r, θ, t), ∇ 2 is the two-dimensional Laplacian, and g(r ) and V (r ) are
respectively the nonlinearity coefficient (alias strength of the two-body inter-atomic
interactions) and the strength of the external harmonic potential, functions of the
spatial radial coordinate r .
To present exact analytical solutions of Eq. (10.43), we use the following ansatz

ψ(r, θ, t) = φ(r ) exp [i Sθ − iμt] (10.44)

in which φ(r ), θ, S, and μ are respectively the stationary wave function (spatial-
varying amplitude), the azimuthal angle, an integer vorticity, and the chemical poten-
tial. Imposing to Eq. (10.44) to satisfy the model equation (10.43) yields the following
ODE for φ(r )
 
d 2φ 1 dφ S2
μφ = − − + g(r )φ 3 + + V (r ) φ. (10.45)
dr 2 r dr r2

Equation (10.45) is subject to the following restriction:


(i) φ(r ) must vary as r [S] at the limit r → 0 when the integer vorticity S = 0;
(ii) in the case when the integer vorticity S =0, φ(r ) must satisfy the following

homogeneous boundary condition at r = 0, dφ dr r =0
= 0;
(iii) φ(r ) must be localized: φ(r )|r =∞ = 0.
In the special situation when the strength V (r ) of the external potential V (r ) has
the form g0
g(r ) = 2 6 , |g0 | > 0, g(0)g (+∞) ∈ R, (10.46)
r ρ (r )

(where g0 = is a free real constant accounting for the nonlinearity strength), we seek
special solution of equation (10.45) in the following form

r
ds
φ(r ) = ρ(r )U [R(r )] , with R(r ) = . (10.47)
sρ 2 (s)
0

It follows from Eqs. (10.46) and (10.47) that the nonlinearity functional parameter
g(r ) as well as R(r ) do not change sign; therefore, ρ(r ) has a constant sign. For
10.2 Stable Vortex Modes in Two-Dimensional Bose-Einstein Condensates 365

g(0)g (+∞) to be a finite quantity, we must have ρ(r ) ∼ r −a , with a ≥ 1/3, and
ρ(+∞) = 0. Under these conditions on ρ(r ), g(r ) will be bounded and the integral
(10.47) that defines R(r ) will converge. Asking now that equation (10.47) satisfies
the ODE (10.45), we find that ρ(r ) must be a nonzero solution of the ODE
 
ρ S2 E
ρ + + μ − V (r ) − 2 ρ = 2 3 , (10.48)
r r r ρ

and U (R) must be any solution of the elliptic ordinary differential equation (EODE)

d 2U
− + g0 U 3 = EU, (10.49)
d R2
where E is an arbitrary real constant.

10.2.2 Exact Vortex-Soliton Solutions for the Attractive


Nonlinearity (g0 < 0) when E = 0

In the special case when E = 0 and g0 < 0, Eq. (10.48) is solvable. In the case of
the harmonic potential V = kr 2 , the solution ρ(r ) of Eq. (10.48) can be written in
terms of the Whittaker’s M and W functions [26, 27] as
    
−1 μ √ |S| √ 2 μ √ |S| √ 2
ρ(r ) = r c1 M k, , kr + c2 W k, , kr .
4 2 4 2

The above restrictions on ρ(r ) require μ < μ0 = 2 (1 + |S|) k and c1 c2 > 0. In
the absence of the trap (k = 0), ρ degenerates to

ρ(r ) = c3 I S −μr + c4 K S −μr ,

with μ < μ0 = 0, where I S and K S are respectively the modified Bessel and Hankel
functions, and constants C3 and C4 satisfy the condition c3 c4 > 0. It is evident that at
the limit r → 0, we have ρ(r ) ∼ r −[S] , leading to g(r ) ∼ r 6|S|−2 and R(r ) ∼ r 2|S| at
the limit r → 0; moreover ρ(r ) → ∞ as r → ∞. Hence, the respective nonlinearity
is localized and bounded, and R(r ) is also bounded. For the boundary conditions
φ(0) = φ(∞) = 0 to be satisfied, we choose as an exact solution to Eq. (10.49), the
following function
η   √  √ 
U (R) = n √ cn nη R − K 1/ 2 , 1/ 2 (10.50)
−g0
366 10 Dynamics of Higher-Dimensional Condensates with Time Modulated …

Fig. 10.5 a Evolution plot of the exact vortex-solitons when the external potential is absent. (Inset)
The corresponding profiles of the attractive nonlinearity coefficient. c Exact vortex-solitons of
different radial quantum number n with S = 1, the respective nonlinearity-coefficient profile being
depicted by the solid line in the inset of (a). Parameters are c3,4 = −μ = −g0 = 1. (b, d) The same
as (a) and (c) when the trapping potential is present, V = 10−2 r 2 , with c1,2 = 3. Reprint from Ref.
[24], Copyright 2022, with permission from American Physical Society


in which n = 2, 4, 6, . . . , n/2 is the radial quantum number, η ≡ K (2/ 2)/R(r √=
∞), cn(ξ, m) is the Jacobi elliptic cosine function with modulus m, and K (1/ 2)
is the complete elliptic integral of the first kind.
As we can see from Eq. (10.50), U (R) ∼ R at the limit R = 0; this implies that
at the limit r → 0, the amplitude of the exact VS is ρU ∼ r |S| , as it should be. We
have thus showed that for given parameters μ and S and nonlinearity strength g0 (and
k, in the case when the trap is taken into account), one can build an infinite number
of exact vortex-soliton solutions with n/2 bright rings surrounding the vortex core,
as it is clearly seen in Fig. 10.5. It is important here to note that these VS solutions
share the same chemical potential, but their energies increase with the increase of
even number n.
10.2 Stable Vortex Modes in Two-Dimensional Bose-Einstein Condensates 367

10.2.3 Exact Analytical Vortex-Soliton Solutions


for the Repulsive Nonlinearity (g0 > 0)

Now, let us we consider the case of the repulsive nonlinearity, g0 > 0, and assume that
the system is confined by a harmonic trap. For the EODE (10.49) to admit elliptic-
function solutions, we must take the free constant E from the condition E > 0; under
this restriction, Eq. (10.48) becomes a nonlinear equation, which in general can be
solved only in a numerical form. To build the VS solutions in the case of repulsive
nonlinearity, we require ρ ∼ r −|S| (for S = 0) at r → 0 so that at the limit r → 0,
the nonlinear term of Eq. (10.48) may be neglected. Under these assumptions, ρ(r )

is similar to the Neumann function, Y S μr , at the limit r → 0 for μ > 0 (that


VS solutions do not exist when μ < 0). Due to the presence of the harmonic trap,
we necessary have ρ → ∞ as r → ∞. Further, under the restriction E > 0, term
Er −2 ρ −3 in equation (10.48) guarantees the sign definiteness of the spatial-varying
function ρ(r ). Therefore, we can integrate numerically Eq. (10.48) to obtain R(r )
and g(r ) if we take small r0 as an initial point and then use as initial conditions,
the Neumann function and its derivative at r = r0 . Vortex-soliton solutions can then
be built in the numerical form, using the following exact analytical solution of Eq.
(10.49) 
 
2 E − B2 
U (R) = sn B R, E/B 2 − 1 , (10.51)
g0

with    √
E/2 < B ≡ n K E/B 2 − 1/R(∞) < E; (10.52)

solution (10.51)–(10.52) is subject to a constraint with even numbers n.


The maximal value of n is computed from Eq. (10.52) as

n < n max = 2R(∞) E/π.

This means that in the case of repulsive nonlinearity, there exists no more than a finite
number of the VS modes (if for example n max < 2, there will not be VS mode). The
number of the numerical VS solutions versus μ is depicted in Fig. 10.6a. As we can
see from plots of Fig. 10.6a, the cutoff value μ0 of the chemical potential in the case
of repulsive nonlinearity is the same as that for the exact VS solutions built when the
attractive nonlinearity was considered. It is also seen from plots of Fig. 10.6a that the
number of VS solutions jumps at points

μ = μ(S)
j ≡ 2 (2 j + |S| − 1) k, j = 1, 2, 3, . . . ;

moreover, point μ(S)


j coincides with the j-th energy eigenvalue of the vortex state
in the corresponding linear Schrödinger equation. We can then conclude that the
number of numerically VS solutions in the interval μ(S) (S)
j < μ ≤ μ j+1 is exactly j,
368 10 Dynamics of Higher-Dimensional Condensates with Time Modulated …

Fig. 10.6 a The number of numerically found VS modes versus the chemical potential in the
case of the repulsive nonlinearity, with the harmonic trap. b The largest instability growth rate for
numerically found vortices with S = 2, E = 1, and k = 10−2 . Reprint from Ref. [24], Copyright
2022, with permission from American Physical Society

Fig. 10.7 a Modulation profiles of the repulsive nonlinearity for numerically found vortices, which
are displayed in b for S = 1 and in c for S = 2, in the presence of the harmonic trap. Open circles,
squares, solid circles, and triangles denote solutions with n = 2, 4, 6, and 8, respectively. Different
plots are obtained with parameters E = 1, μ = 2, k = g0 = 0.01. Reprint from Ref. [24], Copyright
2022, with permission from American Physical Society

associated with expression (10.51) for U [R] in which n are taken as n = 2, 4, 6, . . . ,


2 j. A similar result has been reported for the 1D GP equation with the optical-lattice
potential [28]. A characteristic example of the numerically found VS solutions is
depicted in Fig. 10.7 where we can see four VS solutions.
All the above solutions are built when the integer vorticity S = 0. It is important
to note that it is still possible to build VS solution when S = 0. This is demonstrated
in the special case of a two-tier nonlinearity, with constant gr :

gr for 0 ≤ r < r0 ,
g(r ) = (10.53)
g0 r −2 ρ −6 for r ≥ r0 .
10.2 Stable Vortex Modes in Two-Dimensional Bose-Einstein Condensates 369

Exact solutions for r ≥ r0 can be built in the same way as above, except that now
R(r ) must be defined as
r
ds
R(r ) ≡ .
sρ 2 (s)
r0

Assuming r0 to be small enough√in comparison with the spatial scale of the external
potential, as for example r0  1/k in the presence of a harmonic trap V (r ) = kr 2 ,
we can approximate φ in the case of r < r0 by a constant as follows

φ(r ) = [μ − V (0)] /gr .

Because φ(r ) and φ  (r ) must be continuous at r = r0 , one then requires φ  (r0 ) = 0


and dU (0)/d R = 0; these two conditions yield

μ − V (0)
gr = .
[ρ(r0 )U (0)]2

The solution to Eq. (10.49) in this special case is then given by Eqs. (10.51) and
(10.52), where B R must be replaced by B R + K (E/B 2 − 1), and n = 1, 3, 5, . . .
for the case of repulsive nonlinearity, so that make φ(∞) = 0. In the same way,
we can build VS solutions in the case of attractive nonlinearity. Examples of found
fundamental solutions in the case of repulsive two-tier nonlinearity are displayed in
Fig. 10.8.
Next,we carry out the linear stability analysis and direct simulations to verify the
stability of the built VS solutions. By means of the azimuthal modulational instability
(AMI) in the case of the attractive nonlinearity in the absence of the trap potential,
we lead to the breaking up and the collapse of higher-order VS solutions. The lowest-
order exact VS solutions with n = 2 are found to be stable when the harmonic trap
is present, if μ → μ0 , the wave amplitude remaining large as we can see from plots
of Fig. 10.9. Similar result has not yet been reported when either S = 1 or S < 2
[29–31].
Under the action of the both repulsive nonlinearity and external harmonic trap,
our numerical simulations showed that VS solutions can be stable for every n at
which they exist, within some region of values of μ , as we can clearly see from
Fig. 10.6b. Also, as we can see from plots of Figs. 10.6b and 10.10, VS solutions
with fewer rings may be less stable than their counterparts with the largest number of
rings. The numerical results plotted in Fig. 10.10a reveal that unstable VS solutions
with n = 2 either exhibit a quasistability or split into vortices with lower topological
charges, which periodically break up and recover the axial symmetry. This finding is
similar to what was already reported in the case of the attractive nonlinearity [29–32].
Nevertheless, unstable VS solutions with n = 4, 6, . . . , as one can clearly see from
plots of Fig. 10.10b, c, ultimately evolve into vortices located close to zero-amplitude
points.
370 10 Dynamics of Higher-Dimensional Condensates with Time Modulated …

Fig. 10.8 Evolution plots of fundamental solitons supported by the repulsive two-tier nonlinearity
[see Eq. (10.53)]. The constant values of gr for n = 1, 3, 5, 7, and 9 are, respectively, 0.0200, 0.0201,
0.0224, 0.0325, and 0.1093. Different plots are generated with the parameters ρ(r0 ) = 1, ρ(r0 ) = 0,
r0 = E = 1, k = g0 = 0.01, μ = 2. Reprint from Ref. [24], Copyright 2022, with permission from
American Physical Society

Fig. 10.9 a Plot of the largest instability growth rate for exact vortex-solitons with S = 3 and
the attractive nonlinearity; b plot os a stable vortex (solid lines, with circles showing its profile
at t = 200, when it was initially perturbed by random noise), and c plot of the corresponding
nonlinearity coefficient when μ = 0.7. Here g0 = −1000, k = 0.01, and c1,2 = 3. Reprint from
Ref. [24], Copyright 2022, with permission from American Physical Society

10.2.4 Conclusion

In this Section, we have studied a physical model governed by a 2D GP with spatially-


varying parameters. The cases of both the attractive and repulsive nonlinearities as
well as the case of the presence/absence of an external trapping potential are con-
sidered. The strategy for constructing exact analytical VS solutions are presented.
This strategy consists of reducing the two-dimensional GP equation to the solvable
uncoupled system formed by Eqs. (10.48) and (10.49). We also showed how to build
approximate fundamental soliton solutions in the case of two-dimensional axisym-
metric profiles of the nonlinearity coefficient, and harmonic trapping potential. Our
results showed that BEC systems with attractive/repulsive two-body inter-atomic
interactions (nonlinearity) support a finite number of exact VS solutions. We have
References 371

Fig. 10.10 a Quasistable evolution of a numerically found vortex with n = 2, in the case of the
self repulsion, at t = 1700. b, c Evolution of unstable vortices with n = 4 and 6 at t = 60 and 70,
respectively. Different plots are generated with the following numerical parameters S = 2, E = 1,
g0 = k = 0.01, μ = 1.82. The vortex with n = 8 are stable. Reprint from Ref. [24], Copyright
2022, with permission from American Physical Society

presented stable VS solutions with vorticity S ≥ 2; vortex-solitons corresponding


to higher-order radial states are also reported. The strategy and the methodology
elaborated for finding VS solutions in this Section can be applied to other models
[33, 34].

References

1. F.Kh Abdullaev, J.G. Caputo, R.A. Kraenkel, B.A. Malomed, Controlling collapse in Bose-
Einstein condensation by temporal modulation of the scattering length. Phys. Rev. A 67, 013605
(2003)
2. H. Saito, M. Ueda, Dynamically stabilized bright solitons in a two-dimensional Bose-Einstein
condensate. Phys. Rev. Lett. 90, 040403 (2003)
3. G.D. Montesinos, V.M. Pérez-García, H. Michinel, Stabilized two-dimensional vector solitons.
Phys. Rev. Lett. 92, 133901 (2004)
4. A. Itin, T. Morishita, S. Watanabe, Reexamination of dynamical stabilization of matter-wave
solitons. Phys. Rev. A 74, 033613 (2006)
5. E. Kengne, W.M. Liu, B.A. Malomed, Spatiotemporal engineering of matter-wave solitons in
Bose-Einstein condensates. Phys. Rep. 899, 1–62 (2021)
6. F.K. Abdullaev, B.B. Baizakov, M. Salerno, Stable two-dimensional dispersion-managed soli-
ton. Phys. Rev. E 68, 066605 (2003)
7. I. Towers, B.A. Malomed, Stable (2+1)-dimensional solitons in a layered medium with sign-
alternating Kerr nonlinearity. J. Opt. Soc. Am. B 19, 537–543 (2002)
8. S.K. Adhikari, Stabilization of bright solitons and vortex solitons in a trapless three-dimensional
Bose-Einstein condensate by temporal modulation of the scattering length. Phys. Rev. A 69,
063613 (2004)
9. R.A. Battye, N.R. Cooper, P.M. Sutcliffe, Stable Skyrmions in two-component Bose-Einstein
condensates. Phys. Rev. Lett. 88, 080401 (2002)
10. D. Anderson, M. Lisak, Nonlinear asymmetric self-phase modulation and self-steepening of
pulses in long optical waveguides. Phys. Rev. A 27, 1393 (1983)
11. D. Anderson, Variational approach to nonlinear pulse propagation in optical fibers. Phys. Rev.
A 27, 3135–3144 (1983)
12. B.A. Malomed, Variational methods in nonlinear fiber optics and related fields. Prog. Opt. 43,
71–193 (2002)
372 10 Dynamics of Higher-Dimensional Condensates with Time Modulated …

13. V.M. Pérez-García, H. Michinel, J.I. Cirac, M. Lewenstein, P. Zoller, Dynamics of Bose-
Einstein condensates: variational solutions of the Gross-Pitaevskii equations. Phys. Rev. A 56,
1424 (1997)
14. M. Desaix, D. Anderson, M. Lisak, Variational approach to collapse of optical pulses. J. Opt.
Soc. Am. B 8, 2082–2085 (1991)
15. FKh. Abdullaev, A. Gammal, A.M. Kamchatnov, L. Tomio, Dynamics of bright matter wave
solitons in a Bose-Einstein condensate. Int. J. Mod. Phys. B 19, 3415–3473 (2005)
16. Yu. Kivshar, S. Turitsyn, Spatiotemporal pulse collapse on periodic potentials. Phys. Rev. E
49, 2536 (1994)
17. R.Y. Chiao, E. Garmire, C.H. Townes, Self-trapping of optical beams. Phys. Rev. Lett. 13,
479–482 (1964)
18. G. Fibich, The Nonlinear Schrödinger Equation: Singular Solutions and Optical Collapse
(Springer, Heidelberg, 2015)
19. T.S. Yang, W.L. Kath, Analysis of enhanced-power solitons in dispersion-managed optical
fibers. Opt. Lett. 22, 985 (1997)
20. G. Fibich, G.C. Papanicolaou, A modulation method for self-focusing in the perturbed critical
nonlinear Schrödinger equation. Phys. Lett. A 239, 167 (1998)
21. F.Kh, Abdullaev, J.G. Caputo, Validation of the variational approach for chirped pulses in fibers
with periodic dispersion. Phys. Rev. E 58, 6637 (1998)
22. L. Bergé, V.K. Mezentsev, J.J. Rasmussen, P.L. Christiansen, Yu.B. Gaididei, Self-guiding light
in layered nonlinear media. Opt. Lett. 25, 1037 (2000)
23. L. Bergé, Wave collapse in physics: principles and applications to light and plasma waves.
Phys. Rep. 303, 259 (1998)
24. L. Wu, L. Li, J.-F. Zhang, D. Mihalache, B.A. Malomed, W.M. Liu, Exact solutions of the Gross-
Pitaevskii equation for stable vortex modes in two-dimensional Bose-Einstein condensates.
Phys. Rev. A 81, 061805(R) (2010)
25. D.-S. Wang, S.-W. Song, B. Xiong, W.M. Liu, Quantized vortices in a rotating Bose-Einstein
condensate with spatiotemporally modulated interaction. Phys. R. A 84, 053607 (2011)
26. D.-S. Wang, S.-W. Song, B. Xiong, W.M. Liu, Quantized vortices in a rotating Bose-Einstein
condensate with spatiotemporally modulated interaction, Phys. Rev. A 84, 053607 (2011)
27. E.T. Whittaker, G.N. Watson, A Course in Modern Analysis, 4th edn. (Cambridge University
Press, Cambridge, 1990)
28. Y. Zhang, B. Wu, Composition relation between gap solitons and Bloch waves in nonlinear
periodic systems. Phys. Rev. Lett. 102, 093905 (2009)
29. T.J. Alexander, L. Bergé, Ground states and vortices of matter-wave condensates and optical
guided waves. Phys. Rev. E 65, 026611 (2002)
30. D. Mihalache, D. Mazilu, B.A. Malomed, F. Lederer, Vortex stability in nearly-two-dimensional
Bose-Einstein condensates with attraction. Phys. Rev. A 73, 043615 (2006)
31. L.D. Carr, C.W. Clark, Vortices in attractive Bose-Einstein condensates in two dimensions.
Phys. Rev. Lett. 97, 010403 (2006)
32. B. LeMesurier, P. Christiansen, Regularization and control of self-focusing in the 2D cubic
Schrodinger equation by attractive linear potentials. Physica D 184, 226 (2003)
33. M. Vengalattore, R.S. Conroy, W. Rooijakkers, M. Prentiss, Ferromagnets for integrated atom
optics. J. Appl. Phys. 95, 4404 (2004)
34. M. Vengalattore, M. Prentiss, A reciprocal magnetic trap for neutral atoms. Eur. Phys. J. D 35,
69 (2005)
Chapter 11
Engineering Matter-Wave Solitons
in Spinor Bose-Einstein Condensates

Abstract We consider in this chapter a (non)integrable system of three nonlinearly


coupled GP equation, which describes the dynamics of matter-wave solitons in a
three-component spinor Bose-Einstein condensate. One-, two-, and three-component
soliton solutions of the polar and ferromagnetic (FM) types are presented. Applying
the Darboux transform (DT) in the case of an integrable system, exact analytical
soliton solutions that display full nonlinear evolution of MI of CW states are also
obtained. Using the Bogoliubov-de Gennes (BdG) equations for small perturba-
tions, the multistability of the solitons is investigated analytically and numerically.
The global stability of the multi-component solitons leading to the ground-state and
metastable soliton states of the FM and polar types is considered too. Also, the effects
of the nonlinearity parameter on the structural stability of the solitons are analyzed
(Li et al. in Phys Rev A 72:033611 [1]).

11.1 Formulation of the Model

The physical system to be investigated in this Chapter is an effectively one-


dimensional BEC loaded in a cigar-shaped trap, elongated in x and tightly confined
in the transverse plane (y, z) [2, 3]. In this physical system, we denote by

→  T
 (x, t) = +1 (x, t), 0 (x, t), −1 (x, t) ,

the one-dimensional three-component wave function describing the atoms in the


hyperfine state with atomic spin F = 1. Here, components +1 , 0 , and −1 cor-
respond to the three values of the vertical spin projection, m F = +1, 0, −1 [4]. The
mathematical model associated with the physical system under consideration is the
following three-component system of GP equations [4–7]

© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2022 373
E. Kengne and W. Liu, Nonlinear Waves,
https://doi.org/10.1007/978-981-19-6744-3_11
374 11 Engineering Matter-Wave Solitons in Spinor …

Fig. 11.1 The evolution plot of the single-component polar soliton generated with the use of the
exact analytical single-component soliton solution (11.13) with a small random perturbation added,
at t = 0, to the φ+1 and φ−1 components. Parameters used in different plots are ν = 1, a = −0.5,
and μ = −1. Reprint from Ref. [1], Copyright 2022, with permission from American Physical
Society

∂±1 2 ∂ 2 ±1  
i =− + (c0 + c2 ) |±1 |2 + |0 |2 ±1 + (c0 − c2 )
∂t 2m ∂ x 2

× |∓1 |2 ±1 + c2 ∗∓1 20 ,


∂0  2 ∂ 2 0  
i =− + (c0 + c2 ) |+1 |2 + |−1 |2 0 (11.1)
∂t 2m ∂ x 2
+c0 |0 |2 0 + 2c2 +1 −1 ∗0 ,

in which
c0 = (g0 + 2g2 ) /3 and c2 = (g2 − g0 ) /3 (11.2)

denote effective constants of respectively the spin-independent and spin-exchange


interactions [1, 8]. In the context of fiber optics, c2 stands for the coefficient of
the four-wave mixing, while c0 + c2 and c0 − c2 account for the SPM and XPM
interactions. Denoting respectively by m, a f , and a⊥ (with f = 0, 2) the atomic
mass, the s-wave scattering length in the channel with total hyperfine spin f, and the
size of the transverse ground state, parameters g0 and g2 in Eq. (11.2) are defined as
follows
42 a f
gf =   , f = 0, 2,
ma⊥ 2
1 − ca f /a⊥
11.1 Formulation of the Model 375

where c = −ζ (1/2) ≈1.46. Assuming time and length to be measured respectively


in units of / |c0 | and 2 /2m |c0 | and using the following transformation for the
wave function  T

→ √
 → φ+1 , 2φ0 , φ−1 ,

the model system (11.1) takes the following scaled form:

∂φ±1 ∂ 2 φ±1  
i =− − (ν + a) |φ±1 |2 + 2 |φ0 |2 φ±1 − (ν − a) |φ∓1 |2±1 φ±1
∂t ∂x 2

−2aφ∓1 φ02 , (11.3)
∂φ0 ∂ φ0
2  
i = − 2 − 2ν |φ0 |2 φ0 − (ν + a) |φ+1 |2 + |φ−1 |2 φ0 − 2aφ+1 φ−1 φ0∗ ,
∂t ∂x
where a ≡ −c2 /c0 and ν ≡ −sgn(c0 ).
The Hamiltonian that gives birth to system (11.3) is a dynamical invariant of the
model (d H/dt = 0) and reads

+∞

∂φ+1 2 ∂φ−1 2 1  
H = d x +

− (ν + a) |φ+1 |4 + |φ−1 |4

∂x ∂x 2
−∞

∂φ0 2
− (ν − a) |φ+1 | |φ−1 | + 2
2 2 − ν |φ0 |4 − (ν + a) (11.4)
∂x
    2  
∗ ∗
× |φ+1 |2 + |φ−1 |2 |φ0 |2 − a φ+1 φ−1 φ02 + φ+1 φ−1 φ0∗ .

One can easily verify that system (11.3) conserves the momentum, the solution’s
norm, and the total magnetization. In other words, the following three equations
hold: (i) the conservation of momentum:

+∞ 
∗ ∂φ+1 ∗ ∂φ−1 2 ∂φ0
P =i φ+1 + φ−1 + 2φ0 d x,
∂x ∂x ∂x
−∞

(ii) the conservation of the solution’s norm, proportional to total number of atoms:

+∞
 
N= |φ+1 (x, t)|2 + |φ−1 (x, t)|2 + 2 |φ0 (x, t)|2 d x, (11.5)
−∞

and
(iii) the conservation of the total magnetization:
376 11 Engineering Matter-Wave Solitons in Spinor …

Fig. 11.2 The evolution plots of norms N0 (solid line), N± (dashed line) of respectively the φ0
and φ±1 components, and total norm N (dotted line), as per Eqs. ( 11.17) and (11.5), in the two-
component polar soliton solution given by Eq. (11.18) perturbed by a small random perturbation
introduced in the φ0 component. Different plots are generated with the use of parameters ν = 1 and
∗ < 0; b a = 1.5 for φ φ ∗ > 0; c a = −1.5 for φ φ ∗ < 0; d
μ = −1 and a a = 1.5 for φ+1 φ−1 +1 −1 +1 −1

a = −1.5 for φ+1 φ−1 > 0. Reprint from Ref. [1], Copyright 2022, with permission from American
Physical Society

+∞
 
M= |φ+1 (x, t)|2 − |φ−1 (x, t)|2 d x. (11.6)
−∞

Setting φ0 = 0 in system (11.3) yields the following system of two nonlinear


partial differential equations, having the same behavior (form) as a system which
describes light transmission in bimodal nonlinear optical fibers [1]:

∂φ±1 ∂ 2 φ±1
i =− − (ν + a) |φ±1 |2 φ±1 − (ν − a) |φ∓1 |2 φ±1 . (11.7)
∂t ∂x2
In the context of nonlinear optical fibers, the two modes φ−1 and φ+1 represent either
different wavelengths or two orthogonal polarizations. a = ν/5 and a = −ν/3 are
associated with the particular cases of respectively two linear polarizations and two
circular polarizations. In system (11.7), the two nonlinear terms, proportional to
11.2 Exact Analytical One-, Two-, and Three-Component Soliton Solutions 377

(ν + a) and (ν − a), account for respectively the SPM and XPM interactions of the
two waves. The modulation instability of continuous wave states in system (11.7)
under the condition ν + a = 0 (XPM-coupled system) has been studied in detail [9].
In the special case when ν = −1 and |a| < 1, the nonlinearity in system (11.7) will be
self-defocusing, and the single NLS equation would show no MI. The XPM-coupled
system (11.7) in the special case when ν = −1 and |a| < 1 of course will give rise
to MI; this means that the XPM interaction is stronger than SPM interaction, that is,
0 < a < 1 [10].

11.2 Exact Analytical One-, Two-, and Three-Component


Soliton Solutions

Following Ieda et al. [11], we here focus our attention, for simplicity, to the special
case of a physically possible integrable model (with a = ν = 1), which corresponds
to c2 = c0 leading to 2g0 = −g2 > 0. For condition 2g0 = −g2 > 0 to be satisfied,
we impose the following conditions on the scattering lengths

3ca0 a2
a⊥ = , a0 a2 (a2 − a0 ) > 0. (11.8)
2a0 + a2

It is important to note that it is also possible to derive and study the stability of exact
analytical soliton solutions of both polar and FM types in the non-integrable case
a
= ν. Following the approach elaborated by Zaidong et al. [1], we present in the
following several soliton species and results for their stability.

11.2.1 Single-Component FM Solitons

A one-component FM soliton can be found by means of the direct method by seeking,


under the conditions a + ν > 0 and μ < 0, the single-component solution in the form
  T
2μ exp [−iμt]
(φ−1 , φ0 , φ+1 ) =
T
− 0, 0, √  . (11.9)
1+ν cosh −μx

In solution (11.9), the chemical potential μ is a parameter of the soliton family.


Solution (11.9) corresponds to the zero-velocity soliton. Solutions leading to moving
solitons can be generated from Eq. (11.9) by means of the Galilean transformation.
In the special case when ν = ±1, condition a + ν > 0 means that a > 1 [a > −1] in
the cases of repulsive (ν = −1) [attractive (ν = +1)] spin-independent interaction.
Using Eqs. (11.5) and (11.4), the norm N and energy H of the above soliton are
found to be
378 11 Engineering Matter-Wave Solitons in Spinor …

4 −μ (ν + a)2 3
N= ,H =− N . (11.10)
ν+a 48

Linearizing system (11.3) about the solution (11.9) leads to the following BdG equa-
tions for small perturbations φ−1 and φ0 :

∂φ−1 ∂ 2 φ−1
=− − (ν − a) |φ+1 |2 φ−1 , (11.11)
∂t ∂x2
∂φ0 ∂ 2 φ0
= − 2 − (ν + a) |φ+1 |2 φ0 . (11.12)
∂t ∂x
For the soliton solution (11.9), a = 1 so that system (11.11)–(11.12) is a system
of decoupled BdG equations. Therefore, solution (11.9) is stable against small per-
turbations. As we well know, envelop soliton solution of the single standard NLS
equation is always stable, so that solution (11.9) cannot be unstable against small
perturbations of φ+1 .

11.2.2 Single-Component Polar Solitons

By setting φ±1 (x, t) = 0, we obtain the following simplest polar soliton for ν = +1
and μ < 0
 T
√ exp [−iμt]
(φ−1 , φ0 , φ+1 ) = −μ 0,
T
√ , 0 . (11.13)
cosh −μx

Linearizing system (11.3) about the soliton solution (11.13) gives rise to a coupled
system of BdG equations for small perturbations φ±1 of the other fields:
 
∂ ∂ 2 χ±1 2μ  ∗

i + μ χ±1 = − + √  (1 + a) χ±1 + aχ∓1 , (11.14)
∂t ∂x 2 cosh 2
−μx

where (∗ ) stands for the complex conjugation and

χ±1 (x, t) ≡ φ±1 (x, t) exp [iμt] . (11.15)

The parametric gain that comes from the x-dependent term (last term) of Eq. (11.14)
may be a source of instability. To understand this instability, we can investigate
system (11.14) at x = 0; we then arrive to the following system:

dχ±1  ∗

i = μ (1 + 2a) χ±1 + 2aχ∓1 (11.16)
dt
11.2 Exact Analytical One-, Two-, and Three-Component Soliton Solutions 379

Fig. 11.3 The evolution of the three-component polar soliton, generated with the use of the exact
analytical soliton solution (11.23) under the action of a small random perturbation initially added

to the φ0 component. Parameters use to obtain different plots are ν = 1, μ = −0.8, and ε = 2/ 5
and a a = 0.5; b a = 1.5; c a = −0.5; d a = −1.5. The meaning of the solid, dashed, and dotted
curves is the same as in Fig. 11.2. Reprint from Ref. [1], Copyright 2022, with permission from
American Physical Society

which illustrates the qualitative mechanism of the parametric instability. One can
easily show that the zero equilibrium point (χ−1 , χ+1 ) = (0, 0) of the linear equations
(11.16) is unstable through a double eigenvalue in the region of a < −1/4.
By means of direct numerical simulations of Eqs. (11.3), we have checked the sta-
bility of the single-component polar soliton (11.13). In these numerical simulations,
we have added at time t = 0 a small random perturbation with values distributed uni-
formly between 0 and 0.03 in components φ+1 and φ−1 . The result of this numerical
investigation showed that the polar soliton (11.13) is always unstable, as shown in
Fig. 11.1. Figure 11.1d depicts the time evolution of the following total number N0
and N±1 of atoms of different components of the condensate under consideration:

+∞ +∞
N0 = |φ0 (x, t)| d x, N±1 =
2
|φ±1 (x, t)|2 d x. (11.17)
−∞ −∞
380 11 Engineering Matter-Wave Solitons in Spinor …

Fig. 11.4 The same as in Fig. 11.2, but generated with the three-component polar soliton solution
given by Eq. (11.24). Different
√ plots are generated with the following numerical values of parameters
ν = 1, μ = −0.8, ε = 2/ 5, and a = 0.5 in (a), a = 1.5 in (b), a = −0.5 in (c), and a = −1.5 in
(d). The meaning of the solid, dashed, and dotted curves is the same as in Figs. 9.2 and 9.3. Reprint
from Ref. [1], Copyright 2022, with permission from American Physical Society

11.2.3 Two-Component Polar Solitons

We now turn our attention to two-component polar solitons. By setting ν = +1 and


considering a zero φ0 (x, t), we arrive to the following two-component polar solitons:

√ exp [−iμt]
φ0 (x, t) = 0, φ+1 (x, t) = ±φ−1 (x, t) = −μ √ . (11.18)
cosh −μx

In the case of the two-component polar solitons (11.18), the corresponding BdG
equations for small perturbations in the φ0 component reads
 
∂ ∂ 2 χ0 2μ  
i + μ χ0 = − 2 + √  (1 + a) χ0 + aχ0∗ . (11.19)
∂t ∂x cosh 2
−μx

As in the case of one-component polar solitons, the instability of solution (11.18) is


possible due to the parametric gain produced by the last term of Eq. (11.18). Similar
11.2 Exact Analytical One-, Two-, and Three-Component Soliton Solutions 381

to the case of Eq. (11.14), this instability can be well understood by studying Eq.
(11.19) at x = 0; setting x = 0, Eq. (11.19) turns to the following equation which
illustrates the qualitative mechanism of the parametric instability:

dχ0  
i = μ (1 + 2a) χ0 ± 2aχ0∗ . (11.20)
dt
It is obvious that the zero equilibrium point χ0 = 0 of Eq. (11.20) for either sign ± is
unstable in the region a < −1/4. Following the same way as for the one-component
polar soliton (11.13), we can test the stability of the soliton (11.18) by means of
direct numerical simulations of Eqs. (11.3) with a small uniformly distributed random
perturbation added to the φ0 component. This numerical study shows that the two-
component polar soliton (11.18) is unstable in the region of |a| ≥ 1, as shown in
Fig. 11.2, and it is stable at |a| < 1 (the corresponding figure is not displayed here).
It is important to notice here that the above results found numerically on the
soliton stability can be explained analytically in the special case when a = +1.
Indeed, writing the perturbation in its canonical form

χ0 (x, t) ≡ χ1 (x, t) + iχ2 (x, t) ,

and looking for a χ1,2 (x, t) as

χ1,2 (x, t) = U1,2 (x) exp [σ t]

where σ stands for the instability growth rate, we arrive at the following differential
system for the eigenfunctions U1 and U2
 
d2 2μ (1 + 2a)
− σ U2 = − 2 − μ + √  U1 ,
dx cosh2 −μx
(11.21)
 
2
d 2μ
−σ U1 = − −μ+ √  U2 .
dx2 cosh2 −μx

At the limit σ = 0, system (11.21) decouples, and each equation of the system
becomes an explicitly solvable and produces a series of zero eigenvalue at a =
an ≡ n (n + 3) /4, n = 0, 1, 2, . . . . Note that system (11.21) becomes symmetric at
point a0 = 0 when σ = 0. The zero crossings at other points an = n (n + 3) /4 with
vanishing σ implies stability changes. In particular, at the critical point a1 = 1, the
two-component polar soliton (11.18) is destabilized as observed in the simulations;
this destabilization corresponds to eigenfunctions
√ 
−μx
tanh 1
U1 (x) = √  , U2 (x) = √ . (11.22)
cosh −μx cosh −μx
382 11 Engineering Matter-Wave Solitons in Spinor …

Fig. 11.5 The evolution of the three-component polar soliton generated with the use of the exact
analytical soliton solution given by Eq. (11.25) with initially added random perturbations. Parame-
ters used here are ν = 1, μ+1 = −1.21, μ−1 = −0.25, and a = 0.5 in (a, c) or a = −0.5 in (b, d).
In cases (a) and (b), the random perturbation was added to the φ0 component, and in cases (c) and
(d) it was added to components φ±1 . The curves labelled by 1, 2, 3, and 4 represent, respectively,
the evolution of norms N+1 , N−1 , N0 , and N defined by Eqs. (11.17) and (11.5). Reprint from Ref.
[1], Copyright 2022, with permission from American Physical Society

Each of critical points an with n > 1 implies additional destabilizations through the
emergence of new unstable eigenmodes of the already unstable soliton.
It is important to notice here that the destabilization of the two-component polar
soliton at a = −1, observed in the simulations, can be explained by means of the
bifurcation theory.

11.2.4 Three-Component Polar Solitons

Now, we turn to the search of three-component polar solitons when φ0 φ±1


= 0.
Setting ν = +1, one family of the three-component solitons of the polar type is
found to be (11.23)
11.2 Exact Analytical One-, Two-, and Three-Component Soliton Solutions 383

√ exp [−iμt]   T,


(φ−1 , φ0 , φ+1 )T = −μ √  ∓ε, 1 −
2 , ±ε (11.23)
cosh −μx

where ε is a free real parameter to be taken as −1 < ε < +1. Exactly as the above
found one- and two-component polar solitons, the three-component polar solitons
(11.23) contains one parameter, the chemistry parameter μ < 0 and does not explic-
itly depend on parameter a.
Another three-component polar soliton solution, independent of a and containing
one parameter, μ is found to be

√ exp [−iμt]   T
(φ−1 , φ0 , φ+1 )T = −μ √  ±ε, i 1 − ε2 , ±ε . (11.24)
cosh −μx

Comparing the three-component polar soliton solutions (11.23) and (11.24), it is seen
that the two solutions differ by the sign of the φ±1 components and a phase shift of
π/2 in the φ0 component.
It is possible to build another family of one parameter three-component polar
soliton solutions which explicitly depend on parameter a. An example of such species
of three-component polar solitons is found to be
⎛  ⎞T

√ exp [−iμt] μ+1 μ−1
(φ−1 , φ0 , φ+1 )T = −μ±1 √  ⎝1, i − , 1⎠ , (11.25)
cosh −μx μ±1

where μ±1 are two free negative parameters to be taken from the condition

√ √ 2 2μ
−μ+1 + −μ−1 + = 0. (11.26)
ν+a

It follows from Eq. (11.26) and the negativity of the chemical potential μ that ν + a >
0, for example, ν = −1, corresponding to repulsive spin-independent interaction.
Solving Eq. (11.26) in the chemical potential μ, and inserting the result in Eq.
solution (11.25), we find that the three-component polar soliton solution (11.25)
contains exactly two free parameters, μ−1 and μ+1 . On the other hand, each of the
two three-component polar soliton solutions (11.23) and (11.24) also contains exactly
two free parameters, μ and ε. Thus, we have built three families of three-component
polar soliton solutions each of which contains exactly two free parameters; two
of these three families of three-component polar soliton solutions do not contain
explicitly parameter a, while one of them contains a explicitly.
The stability of all the above found three-component polar soliton solutions was
tested in direct simulations and reported in Ref. [1]. For the direct numerical sim-
ulations for testing the stability of solutions (11.23) and (11.24), a small random
perturbation was performed only in the φ0 component, and it was found that both
these types are unstable, as shown in Figs. 11.3 and 11.4.
384 11 Engineering Matter-Wave Solitons in Spinor …

For testing the stability of the three-component polar soliton solution given by Eq.
(11.25), small random perturbations were performed in all the three components; it
has been found that this three-component polar soliton (11.25) is completely stable.
An example of the obtained numerical simulations is reported in Fig. 11.5. It is seen
from plots of Fig. 11.5 that the added small perturbation on each soliton compo-
nent induces only small oscillations of the amplitudes of the corresponding soliton
component.

11.2.5 Multistability of Solitons

As we have seen in the above analysis, the one-component ferromagnetic soliton


(11.9), the two-component polar solitons (11.18) in the regions of −1 < a < +1 ,
as well as the three-component polar soliton (11.25) may all be stable in the same
parameter region. In the situation when these three solitons are stable in the same
parameter region, one may identify which solitons are “more stable” and “less stable.”
For this aim, one may fix the soliton’s norm (11.5) and compare respective values of
the Hamiltonian (11.4) for these three solutions. In this comparison, the ground-state
solution is assumed to correspond to a minimum of H at given N .
Substituting each derived solutions in Eq. (11.4) reveals that all the one-, two-,
and two first three-component polar solitons (11.13), (11.18), (11.23), and (11.24)
which do not explicitly contain a produce identical relations between N , μ, and H :

√ 1
N (μ) = 4 −μ, H (N ) = − N 3 . (11.27)
48
The situation is not the same for the one-component ferromagnetic soliton (11.9) for
which N (μ) and H (N ) are given in Eq. (11.10) and differ from those given in Eq.
(11.27). Computing N (μ) and H (N ) in the case of the stable three-component polar
soliton (11.25) which depends explicitly on a, we found that μ, N , and H are related
by exactly the same form as Eq. (11.10) for the FM soliton.
Comparing expression (11.27) to expression (11.10), we find that the one-
component FM soliton (11.9) and the stable three-component polar soliton given
by (11.25) simultaneously provide the minimum of energy when ν = +1 and a > 0,
which corresponds to attractive spin-independent and attractive spin-exchange inter-
actions between atoms. Each of the two species (11.9) and (11.25) in the cases
v = +1 and a > 0 plays the role of the ground state in its own class of the solitons.
In the above studies, we found that the two-component soliton (11.18) is also stable
in the region of 0 < a < 1, but it corresponds to higher energy; this means that the
two-component soliton (11.18) in region 0 < a < 1 represents a metastable state.
Also, the FM soliton (11.9) and the three-component polar soliton (11.25) exist
simultaneously when ν = −1 and a > 1 under the condition that ν + a is positive.
Because other soliton solutions (11.13) , (11.18) , (11.23) , and (11.24) do not exist
11.2 Exact Analytical One-, Two-, and Three-Component Soliton Solutions 385

Fig. 11.6 The evolution of the two-component polar soliton on the CW background, obtained with
the help of the exact analytical soliton solution given by Eq. (11.29), under the action of initial
random perturbations added to the φ0 and φ±1 components. Different parameters used in various
plots are ν = +1, q = −1/2, and μ = −1. Reprint from Ref. [1], Copyright 2022, with permission
from American Physical Society

when ν = −1, the two soliton species (11.9) and (11.25) provide for the energy
minimum.
The situation ν = +1 and −1 < a < 0 correspond to attractive spin-independent
and repulsive spin-exchange interactions in Eq. (11.7); in this situation, the Hamil-
tonian (11.27) is smaller than the competing one (11.10) for the single-component
FM soliton. In this situation, we then conclude that the two-component polar soli-
ton (11.18) plays the role of the ground state. It has been showed that also the
single-component soliton (11.9) and the three-component polar soliton (11.25) are
also stable in the region ν = +1 and −1 < a < 0, but correspond to greater energy;
these two solitons (11.9) and (11.25) then represent metastable states in this region.
The above studies have showed that there are no stable solitons when ν = +1 and
a < −1.
Using the dependence N (μ) for each solution family of found solitons and
employing the well-known known Vakhitov-Kolokolov criterion which gives the
necessary stability condition for a soliton family supported by a self-attractive non-
linearity (this criterion reads d M/dμ < 0) [12, 13], one can state an additional char-
acteristic of the soliton stability. Although this criterion guarantees that the soliton
is always stable against perturbations with real eigenvalues, it does not say anything
on the oscillatory perturbations modes that appear due to the complex eigenval-
ues. Computing d M/dμ from Eqs. (11.10) and (11.27), we find that d M/dμ < 0.
This means that the above solitons may be unstable only against temporal oscillat-
ing perturbations that grow in time. This features/behaviors of unstable solitons are
illustrated Figs. 11.1, 11.2, 11.3 and 11.4 generated with direct simulations.

11.2.6 Finite-Background Solitons

Let us now study the evolution of solitons embedded on a non-vanishing continuous


wave (CW) background. For simplicity, we focus our attention to the special case
when ν = 1 and a = −1/2. In this special case, we can build exact analytical solu-
tions for solitons sitting on a non-vanishing CW background. First of all, we note
386 11 Engineering Matter-Wave Solitons in Spinor …

Fig. 11.7 The same as in Fig. 11.6, but for the finite-background soliton (11.30). Parameters are
ν = a = 1 and μ = −0.36. Reprint from Ref. [1], Copyright 2022, with permission from American
Physical Society

that in this special case, Eq. (11.3) admits the CW solution



μ
(φ−1 , φ0 , φ+1 ) = − (1, 0, 1)T exp [−iμt]
T
(11.28)
2

Using the CW solution (11.28) as the seed solution and letting φ0 = 0 yield the fol-
lowing two-component polar soliton solutions with a non-vanishing CW background
attached on it:

(φ−1 , φ0 , φ+1 )T
 T
√ 1 1 1 1
= −μ √ ∓ √  , 0, √ ± √  exp [−iμt] .
2 cosh −μx 2 cosh −μx
(11.29)

When ν = a = 1, it is also possible to use the seed solution



1 μ
(φ−1 , φ0 , φ+1 ) =
T
− (1, 1, 1)T exp [−iμt]
2 2

to build the following family of three-component polar soliton solution with a CW


background attached on it:
 T
φ−1 , φ0 , φ+1
√  T
−μ 1 1 1 1 1 1
= √ ± √ , √ ∓ √ , √ ± √ 
2 2 cosh −μx 2 cosh −μx 2 cosh −μx
exp [−iμt] . (11.30)

Performing direct simulations to Eq. (11.3), we have studied the stability of the
two two- and three-component polar solitons (11.29) and (11.30) by simultaneously
perturbing the φ0 and φ±1 components by adding at the initial time t = 0 small
11.3 The Darboux Transform and Nonlinear Development of Modulational Instability 387

Fig. 11.8 The nonlinear development of the modulation instability, generated with the use of exact
analytical soliton solution (11.41), under conditions k = 2η and αc2 + βc2 > ξ 2 . Parameters are
k = 0.6, η = 0.3, ξ = 1, α = β = γ = exp (−8), αc = 1.2, and βc = 0. Reprint from Ref. [1],
Copyright 2022, with permission from American Physical Society

uniformly distributed random perturbations. Examples of obtained results are illus-


trated in Figs. 11.6 and 11.7 showing the evolution of respectively two- and three-
component polar solitons embedded on a non-vanishing CW background. These
two figures show that both the two- and three-component polar soliton solutions are
unstable, although the character of the instability is different. It is seen from plots
of Fig. 11.6 [11.7] that the soliton’s core seems stable [features oscillatory instabil-
ity], but the CW background appears to be modulationally unstable [stable under
the modulation]. Because the soliton’s core is attached to its CW background, we
conclude that both the two- and three-component solitons are unstable.

11.3 The Darboux Transform and Nonlinear Development


of Modulational Instability

In this Section, we focus our attention on the integrable case of Eq. (11.1) with
c2 + c0 = 0, when c0 < 0, which in terms of Eq. (11.3) corresponds to ν = a = 1.
This special case which corresponds to the attractive interactions, makes it possible to
develop deeper analysis of MI. We will also assumed that the physically constraint
(11.8) imposed on the scattering lengths of collisions between atoms is satisfied.
Under all these assumptions, Eqs. (11.3) can be reduced to the following completely
integrable 2 × 2 matrix NLS equation [11]
 
∂Q ∂ 2 Q φ+1 φ0
i + + 2QQ Q = 0, Q ≡

. (11.31)
∂t ∂x2 φ0 φ−1

Darboux transform (DT) for the NLS equation (11.31) can be derived from the
respective Lax pair [11, 14]

→ −

∂ → ∂
− −

= U , = V ; (11.32)
∂x ∂t
in Eq. (11.32), U and V are defined as
388 11 Engineering Matter-Wave Solitons in Spinor …

U = λJ + P, V = 2iλ2 J + 2iλP + iW (11.33)


   
I 0 0 Q
J= ,P = , (11.34)
0 −I −Q† 0
 
QQ† ∂ Q
W = ∂ † ∂x † , (11.35)
∂x
Q −Q Q



where  = (1 , 2 )T is the matrix eigenfunction corresponding to λ, 1 and 2
are 2 × 2 matrices, λ is the spectral parameter, and I and 0 are respectively the unit
and zero matrices. It is evident that the Lax pair (11.32) is an over-determined linear
system and Eq. (11.31) is tantamount to the compatibility condition system (11.32),

∂U ∂V
− + UV − VU = 0.
∂r ∂x
By virtue of the Lax pair (11.32), we introduce a transformation in the form of
 


 −
→ −1 λ1 I 0
 = (λ − S)  , S = DD , = , (11.36)
0 λ1 I

D being a nonsingular matrix satisfying the first order partial differential equation

∂D
= JD + PD.
∂x
Next, letting


  
∂ −
→  0 Q1
= U  , U = λJ + P1 , P1 ≡ , (11.37)
∂x −Q†1 0

yields
P1 = P + JS − SJ. (11.38)

Moreover, the following involution property of the above linear equations is satisfied.


Involution property: If  = (1 , 2 )T is an eigenfunction with the eigenvalue λ,
 T
then −2∗ , 1∗ will be an eigenfunction with the eigenvalue −λ∗ . Taking then D
as follows    
1 −2∗ λI 0
D= , = ,
2 1∗ 0 −λ∗ I

we obtain    
I0  ∗
 −S11 S12
S=λ + λ+λ ,
0I S21 −S22

the matrix elements of S being given by


11.3 The Darboux Transform and Nonlinear Development of Modulational Instability 389
 
−1 −1
 
−1 −1
S11 = I + 1 2−1 1∗ 2∗ , S12 = 2 1−1 + 1∗ 2∗ ,
 −1
−1  
−1 −1
S21 = 2∗ 1∗ + 1 2−1 , S22 = I + 2 1−1 2∗ 1∗ .

The Darboux transform for Eq. (11.31) finally follows from Eq. (11.38), taking the
form of  
Q1 = Q+2 λ + λ∗ S12 . (11.39)

Combining Eqs. (11.39) and (11.32), we deduce that Q1 defined by Eq. (11.39) is
a new solution for Eq. (11.42), as soon as the seed solution Q of the NLS equation
(11.31) is known. It is important to note that if we take the trivial zero state as the
seed solution, solution (11.39) will be just a one-soliton solution of the NLS equation
(11.31). Further, using the solution Q1 as the new seed solution of Eq. (11.31), we
derive from Eq. (11.39) another new solution of Eq. (11.31), which now will be a
two-soliton solution. Following the same procedure, we can use the (n − 1)-soliton
solution to build the next solution, the n-soliton solution of the NLS equation (11.31).
Thus, we have shown how to apply the DT to construct the multisoliton solution of
the NLS equation (11.31).
In the following, we associate to the NLS equation (11.31) nonzero boundary
conditions (BCs) and consider the corresponding solution. First of all, we note that
Eq. (11.31) admits the following CW solution with constant densities, appearing as
its simplest solution with nonzero BCs:
 
βc αc    
Qc = −Ac exp [iϕc ] , Ac ≡ , ϕc ≡ kx + 2 αc2 + βc2 − k 2 t,
αc −βc
(11.40)
where Ac with real constant elements αc and βc (with |αc | + |βc | > 0) is the matrix
amplitude (carrier amplitude) and k is the wavenumber (carrier wavenumber). In
solution (11.40), we can easily verify that the constant densities of components φ±1
are equal, while their signs are opposite.
Taking the CW solution with constant densities Qc as the seed solution and apply-
ing the above DT, we solve the linear system (11.32) and employ Eq. (11.39) to obtain
the following new family of solutions of the NLS equation (11.31):
  −1
Q1 = Ac + 4ξ I + AA∗ A exp (iϕc ) , (11.41)

where
  −1
1 1
A =  exp [θ − iϕ] + Ac Ac  exp [θ − iϕ] + I , (11.42)
κ κ
θ = M I x + [2ξ M R − (k + 2η) M I ] t, (11.43)
ϕ = M R x − [2ξ M I + (k + 2η) M R ] , (11.44)
  
M = (k + 2iλ)2 + 4 αc2 + βc2 = M R + i M I , (11.45)
390 11 Engineering Matter-Wave Solitons in Spinor …

κ being the spectral parameter given as

1
κ≡ (ik − 2λ + i M) with λ = ξ + iη,
2
and  being an arbitrary 2 × 2 symmetric matrix defined as
 
βα
= .
αγ

When ξ = 0, solution (11.41) reduces to the CW (11.40). On the other hand, if


Ac = 0, solution (11.41) will be, as we have described in the previous Section, one-
soliton solution of the NLS equation (11.31):
   
4ξ 1 exp (−θ1 ) + σ y ∗1 σ y exp (θs ) det 1
Q1 = exp (iϕs ) , (11.46)
exp (−2θs ) + 1 + exp (2θs ) |det 1 |2

where  
θs = 2ξ (x − 4ηt) − θ0 , ϕs = 2ηx + 4 ξ 2 − η2 t,

θ0 being a free real constant which determines the initial position of the soliton, σ y
is the Pauli matrix, and 1 is the polarization matrix [15] defined as
 
 −1/2 β1 α1
1 = 2 |α|2 + |β|2 + |γ |2 ≡ .
α1 γ1

We can then conclude that solution (11.41) describes one-soliton embedded on a


non-vanishing CW background.
It is important to note that the three-component polar soliton solution (11.30)
found in the previous Section is not a special case of solution (11.41); indeed, the
background fields in components φ±1 in solution (11.30) have the same sign, which
is not the case for solution (11.41).
Let us note that the exact solutions (11.41) can be used to describe the onset and
nonlinear development of MI of the continuous wave states. A necessary condition
for this situation to happen is that different parameters are taken from the condition
M I = 0, which is possible only when Im(k + 2iλ) = 0. This leads to k = 2η. To
observe the instability of the CW background, we must have αc2 + βc2 − ξ 2 > 0. It is
evident that when k = 2η and αc2 + βc2 − ξ 2 > 0, solution (11.41) will be periodic in
the spatial coordinate and aperiodic in the temporal one. Since under these condition
θ will not depend on x, we conclude that the solution features no localization. It is
important to note that under the above conditions, we will have M I = 0 and
 
M R2 = 4 αc2 + βc2 − 4ξ 2 (11.47)
11.3 The Darboux Transform and Nonlinear Development of Modulational Instability 391

Fig. 11.9 Evolution plots of the density distribution in the soliton, generated with the use of
exact analytical solution (11.41) by adding a random time-varying perturbation of the nonlinear
coupling constant a, with the perturbation amplitude ±5%. Different plots are generated with the
following numerical values of parameters: ν = a = 1 (as concerns the unperturbed value of a),
ξ = 1, η = −0.03, and θ0 = 0. a The ferromagnetic soliton, obtained with α1 = 0.48, β1 , and γ1
determined by conditions α12 = β1 γ1 and 2 |α1 |2 + |β1 |2 + |γ1 |2 = 1; b the polar soliton, generated
with α1 = 0.53 and β1 = 0.43, γ1 being determined by condition 2 |α1 |2 + |β1 |2 + |γ1 |2 = 1. In
different panels, solid and dotted curves show respectively the exact solutions and the perturbed
ones, generated by numerical simulations. Reprint from Ref. [1], Copyright 2022, with permission
from American Physical Society

will not vanish. To show in details that solution (11.41) under the above conditions
will develop a particular mode of the nonlinear development of MI, we can take a
special case with  = εE, where ε is a small amplitude of the initial perturbation
added to the CW background and E is a matrix with all elements equal to 1. This
choice of  will help us to initiate the onset of MI. Linearizing the initial value of
solution (11.41) with respect to the small parameter
, we arrive to
392 11 Engineering Matter-Wave Solitons in Spinor …
 
Q1 (x, 0) ≈ Ac ρ −
χ1 E cos [M R x] − εχ2 σ z σ x Ac exp [i M R x] exp [ikx] ,
(11.48)
where
 
ρ ≡ 1 − ξ (2ξ + i M R ) / αc2 + βc2 with |ρ| = 1,
 
χ1 ≡ ξ M R (2iξ − M R ) / αc2 + βc2 ,
  
χ2 ≡ βc χ1 / αc2 + βc2 , M R = 2 αc2 + βc2 − ξ 2 ,

and σ x,z are the Pauli matrices. It is obvious that the approximation (11.48) consists
of a non-vanishing CW background, namely, its first term, two spatially modulated
terms (perturbations). Equation (11.48) therefore means that a small periodic per-
turbation in the CW background may lead to modulational instability. A compari-
son of the exact solution given by Eq. (11.41) in the special case when k = 2η and
αc2 + βc2 − ξ 2 > 0 with the results obtained from direct numerical simulations of Eq.
(11.38) under a given initial condition, we obtain that the two solutions are very close
each to another, both showing the development of the MI. For a better understanding,
we have used the exact solution (11.41) when k = 2η and αc2 + βc2 − ξ 2 > 0 to depict
in Fig. 11.8 the nonlinear development of the modulational instability. The nonlinear
MI shown in Fig. 11.8 means that the condensate periodically transfers atoms from
the spin state m F = 0 into the spin states m F = ±1 and vice versa.
In the general case when a
= 1 (that is, 2g0
= −g2 ), one can directly analyze the
onset of the MI from Eqs. (11.3) using the linear stability analysis [16]. In this way,
one should consider the CW solution (11.40) with k = 0 and perturb it as follows
 
    b+1 (x, t) b0 (x, t)
 = (Ac + B) exp 2iν αc2 + βc2 t , B =
Q ,
b0 (x, t) b−1 (x, t)

where matrix B with elements b±1 (x, t) and b0 (x, t) is a weak matrix of perturba-
tions. Using the linear stability method and denoting by K and ω the perturbation
wavenumber and frequency, respectively, yield the following two branches of the
dispersion relation:
 
ω2 = K 2 K 2 − 4aαc2 − 4aβc2 , (11.49)
 
ω2 = K 2 K 2 − 4ναc2 − 4νβc2 . (11.50)

For the MI to occur, at least one of the two ω2 must be a negative quantity, which
is possible as soon as at least one of the following two condition
   
K 2 < 4a αc2 + βc2 , K 2 < 4ν αc2 + βc2 (11.51)

holds. Note that if aν < 0, then one and only one of conditions (11.51) is satis-
fied. This means that, for example, the MI may occur in the case of the repulsive
spin-independent interactions in Eqs. (11.3) when a is positive. Also the MI may
References 393

occur if a > 0 and ν > 0, as for example, when a = v = 1 (case of attractive spin-
independent interactions in Eqs. (11.3)) with positive a.
For exploring the stability of the exact analytical soliton solutions (11.46) of
both the ferromagnetic and polar types rise up the question of the soliton to survive
random deviations of the coupling constant a from the value a = 1 corresponding
to the integrability of the system. Figure 11.9 shows the evolution of the density
distribution in the soliton generated from the exact soliton solution (11.46) by adding
a random time-dependent perturbation of the coupling constant a. As we can see from
plots of Fig. 11.9, solitons of both types are robust against random changes of the
nonlinear coupling constant a.

11.4 Conclusion

Based on a system of (non)integrable three nonlinearly coupled GP equations, we


have presented in this chapter soliton states of spinor BEC with atomic spin F = 1.
In the case when the model equations are nonintegrable, various types of exact
analytical soliton solutions are reported, including single-component ferromagnetic
soliton solutions, one-, two-, and three-component polar soliton solutions. Their
linear stability are checked by direct numerical simulations and, in some special
cases, in an analytical form by means of BdG equations. The global stability of the
solitons is analyzed too. In the special case when the model system is integrable, we
employed the DT to present methodically the construction of multisoliton solutions
that describe multisoliton embedded on a non-vanishing CW background. Their
stability against a random time-dependent perturbation of the nonlinear coupling
parameter a is investigated numerically and analytically. We found that the one-
soliton solutions of both the FM and polar types are robust against random changes
parameter a.

References

1. L. Li, Z. Li, B.A. Malomed, D. Mihalache, W.M. Liu, Exact soliton solutions and nonlinear
modulation instability in spinor Bose-Einstein condensates. Phys. Rev. A 72, 033611 (2005)
2. W.M. Liu, E. Kengne, Schrödinger Equations in Nonlinear Systems (Springer Nature Singapore
Pte Ltd., 2019)
3. S.W. Song, L. Wen, C.F. Liu, S.-C. Guo, W.M. Liu, Ground states, solitons and spin textures
in spin-1 Bose-Einstein condensates. Front. Phys. 8(3), 302–318 (2013)
4. T.-L. Ho, Spinor Bose condensates in optical traps. Phys. Rev. Lett. 81, 742–745 (1998)
5. Y. Kawaguchi, M. Ueda, Spinor Bose-Einstein condensates. Phys. Rep. 520, 253–381 (2012)
6. D.M. Stamper-Kurn, M. Ueda, Spinor Bose gases: symmetries, magnetism, and quantum
dynamics. Rev. Mod. Phys. 85, 1191 (2013)
7. S.W. Song, L. Wen, C.F. Liu, S.C. Guo, W.M. Liu, Ground states, solitons and spin textures in
spin-1 Bose-Einstein condensates. Front. Phys. 8(3), 302–318 (2013)
394 11 Engineering Matter-Wave Solitons in Spinor …

8. H. Pu, C.K. Law, S. Raghavan, J.H. Eberly, N.P. Bigelow, Spin-mixing dynamics of a spinor
Bose-Einstein condensate. Phys. Rev. A 60, 1463 (1999)
9. G.P. Agrawal, Nonlinear Fiber Optics (Academic Press, San Diego, 1995)
10. G.P. Agrawal, Modulation instability induced by cross-phase modulation. Phys. Rev. Lett. 59,
880–883 (1987)
11. J. Ieda, T. Miyakawa, M. Wadati, Exact analysis of soliton dynamics in spinor Bose-Einstein
condensates. Phys. Rev. Lett. 93, 194102 (2004)
12. N.G. Vakhitov, A.A. Kolokolov, Stationary solutions of the wave equation in a medium with
nonlinearity saturation. Radiophys. Quantum Electron. 16, 783 (1973)
13. L. Bergé, Wave collapse in physics: principles and applications to light and plasma waves.
Phys. Rep. 303, 259 (1998)
14. T. Tsuchida, M. Wadati, The coupled modified Korteweg-de Vries equations. J. Phys. Soc. Jpn.
67, 1175 (1998)
15. J. Ieda, T. Miyakawa, M. Wadati, Exact analysis of soliton dynamics in Spinor Bose-Einstein
condensates. Phys. Rev. Lett. 93, 194102 (2004)
16. N.P. Robins, W. Zhang, E.A. Ostrovskaya, Y.S. Kivshar, Modulational instability of spinor
condensates. Phys. Rev. A 64, 021601(R) (2001)
17. L. Wen, Q. Sun, Y. Chen, D.-S. Wang, J. Hu, H. Chen, W.M. Liu, G. Juzeliūnas, B.A. Malomed,
A.-C. Ji, Motion of solitons in one-dimensional spin-orbit-coupled Bose-Einstein condensates.
Phys. Rev. A 94, 061602(R) (2016)
18. Y.-J. Lin, K. Jiménez-García, I.B. Spielman, Spin-orbit-coupled Bose-Einstein condensates.
Nature (London) 471, 83 (2011)
19. V. Galitski, I.B. Spielman, Spin-orbit coupling in quantum gases. Nature 494, 49–54 (2013)
20. H. Zhai, Degenerate quantum gases with spin-orbit coupling: a review. Rep. Progr. Phys. 78,
026001 (2015)
21. L.W. Cheuk, A.T. Sommer, Z. Hadzibabic, T. Yefsah, W.S. Bakr, M.W. Zwierlein, Spin-
injection spectroscopy of a spin-orbit-coupled Fermi gas. Phys. Rev. Lett. 109, 095302 (2012)
22. T. Busch, J.R. Anglin, Motion of dark solitons in trapped Bose-Einstein condensates. Phys.
Rev. Lett. 84, 2298–2301 (2000)
23. S. V. Manakov, On the theory of two-dimensional stationary self-focusing of electromagnetic
waves. Zh. Eksp. Teor. Fiz. 65, 505 (1973); Sov. Phys. JETP 38, 248 (1974)
24. S. Burger, K. Bongs, S. Dettmer, W. Ertmer, K. Sengstock, A. Sanpera, G.V. Shlyapnikov,
M. Lewenstein, Dark solitons in Bose-Einstein condensates. Phys. Rev. Lett. 83, 5198–5201
(1999)
Chapter 12
Engineering Magnetic Solitons
in Nonlinear Systems

Abstract In this chapter, we consider a number of mathematical models which


describe various nonlinear phenomena, including the dynamics of dissipative mag-
netic matter-wave solitons in a spinor polariton Bose-Einstein condensate, nonlinear
magnetization dynamics (NMDs) of the classical ferromagnet with two single-ion
anisotropies in an external magnetic field, and linear and nonlinear magnetization
dynamics in the presence of spin-polarized current. With the use of these mathemat-
ical models, we investigate (i) the dynamics of dissipative magnetic polariton soliton
in multi-component BECs, (ii) the effects of spin transport on nonlinear excitations
in a ferromagnetic nanowire with non uniform magnetization, and (iii) the effects
of a spin-polarized current on the nonlinear excitation of the magnetization in the
ferromagnetic metal nanowire.

12.1 Dynamics of Dissipative Magnetic Matter-Wave


Solitons in a Spinor Polariton Bose-Einstein
Condensate

In the present Section, we consider a driven-dissipative GP equation coupled to the


rate equation to investigate the dynamics of dissipative magnetic matter-wave soli-
tons in a spinor polariton Bose-Einstein condensate. We mainly discuss a scenario,
different from the conventional single-channel one for dissipative solitons; in such
a scenario, double balances rely on the presence of multiple collective excitation
channels in open-dissipative quantum systems. With the help of this scenario, we
prove, using exact analytical soliton solutions of the mathematical model under con-
sideration, the existence of dissipative magnetic solitons (MSs) in a spinor polariton
BEC [1].

© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2022 395
E. Kengne and W. Liu, Nonlinear Waves,
https://doi.org/10.1007/978-981-19-6744-3_12
396 12 Engineering Magnetic Solitons in Nonlinear Systems

12.1.1 Introduction and Model Equations

In most non-dissipative nonlinear systems such as Hamiltonian systems, a single


balance between dispersion and nonlinearity generally gives rise to solitons [2].
Due to the presence of loss and/or gain of matter or energy in dissipative nonlinear
systems, solitons usually exist within a finite lifetime [3, 4]. For dissipative nonlinear
systems, dissipative solitons (solitons that do not decay with time) may be realize by
a best balances of gain and dissipation, along with dispersion and nonlinearity [5].
In the realization of dissipative solitons, exact solutions of the model equation play
important roles; indeed, exact analytical solitonlike solutions of the model equation
can be employed for clarifying the bi-channel double balance of a dissipative MS,
as shown in Fig. 12.1. Figure 12.1 is generated under the conditions that

(i) the condensed polaritons with loss rate γC continuously replenish from reservoir
polaritons at rate R;
(ii) the reservoir polaritons decay at rate γ R while subjected to a uniform pumping
P;
(iii) the polaritons in the reservoir and condensate interact with each other modeled
by the spin-independent interaction parameter g R ;
(iv) there exist two excitation channels on top of the steady state, namely, the density
and the spin polarization which are usually coupled;
(v) nonlinear spin-polarization dynamics occurs in an isolated channel, leading to
a soliton profile, when the gain balances loss and fixes the background density;
the appearance of the soliton profile manifests the interplay of nonlinearity and
dispersion.

The aim of this Section is to employ the exact analytical solution of the model equa-
tion for investigating analytically the dynamics of dissipative MSs on top of a spin-
balanced density background for spinor polariton Bose-Einstein condensates [6–9].
The dissipative magnetic solitons found in the present Section are differ from con-
ventional dissipative solitons generated with the use of solutions of Gross-Pitaevski
equation [10], as well as from dissipative solitons in open-dissipative polaritonic
systems [11, 12].
The physical system to be studied in this Section is a spinor polariton BEC
formed under uniform non-resonant pumping in a wire-shaped microcavity [13].
For the exciton-polariton BEC, we assume the order parameter ψ(x, t) to be a two-
component complex vector with spin-up and spin-down wave functions ψ↑↓ = ψ1,2 :
ψ(x, t) ≡ [ψ1 (x, t), ψ2 (x, t)]T [14, 15, 0-15]. Denoting by n R the density of the
reservoir polaritons, the mathematical model that governs the evolution of the wave
functions ψ1 and ψ2 is found to be the driven-dissipative GP equations coupled to
the rate equation for n R [16, 17]:
 
∂ψ1 2 ∂ 2
i = − + g |ψ1 |2
+ g12 |ψ2 | 2
+ V R + D s ψ1 , (12.1)
∂t 2m ∂ x 2
12.1 Dynamics of Dissipative Magnetic Matter-Wave Solitons in a Spinor … 397

Fig. 12.1 Schematics of


bi-channel double-balance
scenario for producing
dissipative magnetic soliton
in a spinor polariton BEC
when the nonresonant
pumping is taken into
consideration. Reprint from
Ref. [1], Copyright 2022,
with permission from
American Physical Society

 
∂ψ2 2 ∂ 2
i = − + g |ψ2 |2
+ g12 |ψ 1 | 2
+ V R + D s ψ2 , (12.2)
∂t 2m ∂ x 2
∂n R   
= P − γ R + R |ψ1 |2 + |ψ2 |2 n R , (12.3)
∂t
with

VR = g R n R , Ds = i (Rn R − γC ). (12.4)
2

In system (12.1)–(12.3), m = 10−4 m e is the effective mass of polaritons, where m e is


the free electron mass, g12 and g are respectively the constants of the two-body inter-
actions for opposite-spin and same-spin polaritons which in this Section are assumed
to be respectively negative and positive and satisfy the constraint |g12 |  g, g R is
the spin-independent parameter of the interaction between condensates and reservoir
polaritons. γC is the loss rate of the condensed polaritons, R is the rate of reservoir
polaritons, whose gain and loss process is captured by Ds. P is the continuous-
wave pumping and γ R is the decay rate of the reservoir. In the following, the density
 2
of spin-up (spin-down) components will be denote by n 1(2) = ψ1(2)  . Under the
above restriction on g12 and g, the steady state condensate of system (12.1)–(12.3)
is linearly polarized with a stochastic polarization direction [18]. For the realization
of dissipative MSs on top of a spin-balanced density background, we assume two
excitation channels corresponding to the density n = n 1 + n 2 and spin polarization
n 1 − n 2 . Under all the above assumptions, we intend to build exact analytical vector
 T
soliton solutions ψ S ≡ ψ1S , ψ2S to Eqs. (12.1)–(12.3) that satisfies the condition

D S ψ S = 0. (12.5)
398 12 Engineering Magnetic Solitons in Nonlinear Systems

12.1.2 Exact Analytical Solutions and Dissipative Magnetic


Polariton Solitons in a Spinor Polariton Bose-Einstein
Condensate

To find exact analytical solutions of system (12.1)–(12.3) that generate dissipative


magnetic polariton solitons, we propose the ansatz

ϕ
n0   r ϕg μ R
ψ1 = 1 + δn 1 exp i + − t ,
2 2 2 
(12.6)

ϕ
n0   r ϕg μ R
ψ2 = 1 + δn 2 exp i − + − t ,
2 2 2 

where n 0 is a constant density, μ R = g R γC /R, δn 1,2 (η = x − υt) is defined from


the system n 1 = n 0 (1 + δn 1 )/2 and n 2 = n 0 (1 − δn 2 )/2 and satisfy the boundary
conditions
lim δn 1(2) (η) = 0,
η→±∞

ϕr and ϕg label respectively the relative and global phases of components ψ1 and ψ2
and satisfy the BC

lim ϕr (g) (η) = 0.
η→±∞ ∂η

Under the above BCs, Eq. (12.6) gives an exact analytical soliton solution of system
(12.1)–(12.3) as soon as n 0 = γPC − γRR and n R = γC /R, and

1 − U2
δn 1 = δn 2 =
√ , (12.7a)
cosh ξηs 1 − U 2

√ ⎤
sinh ξηs 1 − U 2
ϕr = arctan ⎣ ⎦+ π, (12.7b)
U 2
⎡√
√ ⎤ √ 
1 − U 2 tanh ξηs 1 − U 2 1 − U 2
ϕg = − arctan ⎣ ⎦ − arctan , (12.7c)
U U

with respectively the dimensionless velocity U and the spin healing length ξs defined
respectively as

2m 
U =υ , ξs = / 2mn 0 (g − g12 ).
n 0 (g − g12 )
12.1 Dynamics of Dissipative Magnetic Matter-Wave Solitons in a Spinor … 399

Fig. 12.2 Properties of a moving dissipative magnetic soliton. a Spatiotemporal density distribution
generated with the use of Eq. (12.7a). b Stokes parameters obtained with the help of Eq. (12.8).
c Distribution of density of n 1 = |ψ1 |2 and n 2 = |ψ2 |2 , relative phase ϕr and global phase phase
ϕg at dimensionless time of t/τ = 15. Panel (c) compares exact analytical solutions (solid lines)
to numerical solutions (dashed lines) of Eqs. (12.1)–(12.3) (see main text). To generate all these
plots, different parameters are chosen as γC = 0.01 ps−1 , R = 0.01 ps−1 mm2 , g = 0.01m e Vµm 2 ,
P = 0.41 ps−1 mm2 , γ R /γC = 40, g12 /g = −0.1, and U = 0.6. Reprint from Ref. [1], Copyright
2022, with permission from American Physical Society

The spatiotemporal density distribution of a moving soliton with U = 0.6 is illus-


trated in Fig. 12.2a. It is seen from this figure that the wave shape is preserved
during the wave motion. Figure 12.2c shows the soliton profile at t/τ = 15 when
τ = /(g − g12 )n 0 . We can see from curves of Fig. 12.2c that the relative phase
exhibits an exact π -jump, which analytically can be obtained as

lim ϕr − lim ϕr = π.
η→+∞ η→−∞

Dashed lines in Fig. 12.2c correspond to the numerical solutions of system (12.1)–
(12.3) generated by taking the initial order parameter from Eqs. (12.6)–(12.7c) for
t = 0 and by letting n R (0) = γC /R. Plots of Fig. 12.2c show perfect agreement
between the analytical and numerical solution at time t/τ = 15 . Also, we have
numerically checked the stability of the MS by keeping n 0 = n 1 (0) + n 2 (0) fixed
and perturbing n 1 (0) − n 2 (0) from Eq. (12.7a).
Solution (12.6)–(12.7c) is a MS; in other words, this solution describes a localized
spin polarization sitting on top of a linearly polarized background state. We can
characterize the polarization property of the polariton soliton associated with the
soliton solution (12.6)–(12.7c) by means of the Stokes parameters [19]. Computing
the distribution of respectively the linear and the circular polarization degrees of
from the soliton solution (12.6)–(12.7c) , we arrive to
   
2 ψ1∗ ψ2 |ψ1 |2 − |ψ2 |2
Sx (η) = , Sz (η) = . (12.8)
n0 n0

As we can see from different curves in Fig. √12.2b, the moving soliton is strongly
elliptically polarized for η < lω (here, lω = 1 − U 2 is the soliton width). It is also
seen from Fig. 12.2b that, the maximum value of the circular polarization degree
400 12 Engineering Magnetic Solitons in Nonlinear Systems

Sz (0) = 1 − U 2 , is reached at the soliton center. Away from the center, the soliton
becomes linearly polarized. Computing the total spin polarization degrees of the
soliton, we arrive to 
dηSz (η) = π ξs ,

which does not depend on the velocity U .


Following the strategy elaborated in works [9, 17, 20], we have computed the
energy E of the soliton and found
 2 2 
 ∂
E= d xψ † − 2m ∂x2
ψ + g−g 4
12
d x [n 1 − n 2 ]2

+ g+g4
12
d x [n − n 0 ]2 . (12.9)

To show that the above magnetic polariton soliton (12.6)–(12.7c) preserves its energy,
we employ Eqs. (12.1)–(12.3) along with soliton solutions (12.6)–(12.7c) and obtain
    
dE ∂ψ1∗ ∂ψ2∗
= −2Re Ds + dx = 0. (12.10)
dt ∂t ∂t

Since the moving magnetic polariton soliton (12.6)–(12.7c) preserves its energy, we
conclude that (12.6)–(12.7c) is a dissipative soliton.
In the case studied in this Section, the constant density n 0 = γPC − γRR of the
dissipative magnetic soliton associated with solution (12.6)–(12.7c) is enabled by
the stable balance between the gain and loss. On the other hand, we can employ Eq.
(12.5) to obtain the following EODE for spin-polarization Sz :
 2
d Sz (y)  
+ Sz4 (y) − 1 − U 2 Sz2 (y) = 0, y = η/ξs . (12.11)
dy

Real solution of Eq. (12.11) results directly from the competition between the non-
linear interaction and the kinetic energy. It is important to note that the dissipative
MS relies on respective balances in two decoupled channels [12, 21, 22]. We also
notice that exact solution of form (12.6)–(12.7c) has been reported for equilibrium
atomic two-component BECs, when g − g12  g [6, 18, 23–25].
In the following, we analyze the properties of (i) linear spin polarization excitation,
(ii) linear response function, and (iii) the excitation spectrum for a spinor polariton
BEC. For this purpose, we consider, for the description of the spinor polariton BEC,
Eq. (12.6) in which we assume that δn  1 and ϕ A(B)  1. Inserting Eq. (12.6) into
the GP system (12.1)–(12.3) and following the standard BdG approach, we obtain,
for the excitation spectra ωq , the following equation:


 2  3  2
ωq − (ω S )2 ωq + i (Rn 0 + γ R ) ωq
  
+ Rn 0 γC + (ω B )2 ωq + ic(q) = 0, (12.12)
12.1 Dynamics of Dissipative Magnetic Matter-Wave Solitons in a Spinor … 401

where
     
ω S = εq0 εq0 + (g − g12 ) n 0 , ω B = εq0 εq0 + (g + g12 ) n 0 , (12.13)
c(q) = − (Rn 0 + γ R ) (ω B )2 + 2gn 0 γC εq0 , (12.14)

with the free-particle energy εq0 = 2mq . The quadratic equation from Eq. (12.12)
2 2

yields ωq = ±ω S for the energy of the spin-polarization excitation. Evidently, the
cubic equation from Eq. (12.12) reflects the coupled linear excitations in the reservoir
and density channel of polariton BEC.
ω S given by Eq. (12.13) is a real quantity, independently on the fact that the
reservoir is fast or slow compared to the polariton BEC, as one can see from Fig. 12.3a.

Fig. 12.3 Visualization of the decoupling of spin-polarization channel from the density channel
through linear excitations. a, b real part (a) and imaginary part (b) of the energy of density excitation
ω D and spectrum ωs of spin polarization excitation. Solid curves depict numerical solutions of
the Bogoliubov’s equations, and the curve with circles indicate analytical solutions generated with
the used equations (12.15). c, d Density static structure factor S D (q) given by Eq. (12.16) and
spin-density static structure factor SS (q) defined by Eq. (12.17) when the spinor polariton BEC
is subjected to a perturbation in the form λ exp [i (q x − ωt)] +H.c (c) and λσz exp [i (q x − ωt)]
+H.c (d) (see the main text). Different plots are generated with the following numerical values of
parameters: γC = 0.01ps−1 , ps−1 µm2 ps−1 µm2 , g = 0.01 m e Vµm2 , P = 0.41ps−1 µm2 , γ R /γC
= 40, and g12 /g = −0.1. Reprint from Ref. [1], Copyright 2022, with permission from American
Physical Society
402 12 Engineering Magnetic Solitons in Nonlinear Systems

Generally, the linear density excitation exhibits a complex energy ω D . In the case
of the linear density excitation, adiabatic elimination of the reservoir enables the
complex expression

ω D = −i ± ω0 , (12.15)
2
where
  
ω0 = εq0 εq0 + (g + g12 ) n 0 − 2g R /R −  2 /4,
 = n 0 n 0R R 2 / (γ R + n 0 R) .

Because ω0 is a real quantity for |q| ≤ qc , ω D will be purely imaginary due to the
polariton loss, with

qc = m α 2 +  4 /4 − α 2 /2 .

For a best understanding, we have displayed respectively in Fig. 12.3a, b the real and
the imaginary parts of the linear density excitation ω D in the fast reservoir limit
along with the numerical solutions (solid lines) of the BdG equations. It is clearly
seen from these figures that the analytical and the numerical solutions agree well
with each other.
Concerning the linear response, we subject the spin polariton BEC to an exter-
nal density perturbation described by the expression λ exp [i (q x − ωt)] + H.c with
λ  1; next we assumed that the linear response in this BEC system is characterized
by the spin density static structure factor SS (q) and the density static structure factor
S D (q) and [26]. For simplicity, we set SS (q) = 0 and focus our attention on the case
of a fast reservoir limit, leading to the following analytical expression for S D (q):
⎧ ε0

+2|ω0 |


q
log , if q < qc ,

⎪ π|ω 0 | −2|ω 0 |


⎨ 0
S D (q) = 4εq
, if q = qc , (12.16)

⎪ π




ε0
⎩ 1 + 1 tan−1 4ω02 −2 q
, if q > qc .
2 π 4ω0 ω0

The result is shown in Fig. 12.3c where the density static structure factor S D (q)
given by Eq. (12.16) is depicted. We can see from Fig. 12.3c that we have limq→∞ S D
(q) = 1. Now, let us consider the situation when a perturbation λσz exp
[i (q x − ωt)] + H.c is present. In this case, wee set S D (q) = 0 and find the spin
density static structure factor SS (q) to be

q2
SS (q) =  . (12.17)
2mω S
12.2 Nonlinear Magnetization Dynamics of a Classical Ferromagnet 403

The result is depicted in Fig. 12.3d. As we can see from Fig. 12.3d, SS (q) → 1 as
q → ∞. These two results (limq→∞ S D (q) = limq→∞ SS (q) = 1) corroborate that
a spin polarization perturbation only induces spin excitations.

12.1.3 Conclusion and Discussions

Using exact analytical soliton solution of a driven dissipative two-component GP


equations, we have introduced dissipative magnetic polariton soliton which manifests
a bi-channel double balance. The properties of linear spin polarization excitation and
of linear response function, as well as that of the excitation spectrum for a spinor
polariton BEC are analyzed. The found here bi-channel double mechanism and its
variants may lead to new dissipative solitons in the system under consideration. It is
important to note that in a broader context, multi-component dissipative nonlinear
system are widely seen, including mode-locked lasers and optical micro-resonators
[27, 28].

12.2 Nonlinear Magnetization Dynamics of a Classical


Ferromagnet

We investigate the effect of an external magnetic field on nonlinear excitations in a


classical ferromagnet with two single-ion anisotropies under the action of an external
magnetic field. By means of the stereographic projection, the effect of an external
magnetic field on the integrability of the system is analyzed and the integrable con-
ditions of the equation of motion are presented. The exact analytical Jost solutions
are presented and their properties as well as the scattering data are investigated in
detail. Using the Binet-Cauchy formula, multisoliton solutions are investigated. Our
results show that the external magnetic field can affect the motion of the center, the
amplitude, and the width of the MSs, but has no effect on the soliton shape [29].

12.2.1 Introduction

Introduced in the last century by Landau and Lifshitz, the Landau–Lifshitz (LL) equa-
tion is a fundamental tool in the magnetic recording industry [30]. It can be used
to describe the nonlinear magnetization dynamics (NMDs) of the classical ferro-
magnet [30]. Its special solutions, including single-solitonlike and multi-solitonlike
solutions, have been reported in a number of research works [31–45]. Also, special
solutions of LL equation for the special initial condition have been considered by
several researchers [46–60] .
404 12 Engineering Magnetic Solitons in Nonlinear Systems

In all the enumerated works on the solutions of LL equation for NMDs, the
anisotropy in an external magnetic field was neglected. When the anisotropy is added
in the external magnetic field, analytical investigation of NMDs of the ferromagnet
becomes difficult. In this situation the equation of motion for the NMDs of the
ferromagnet differs from those of an isotropic ferromagnet and cannot be solved
neither by separating variables in moving coordinate, by Darboux transform (DT),
nor by an usual form of inverse scattering transformation (IST) [32]. Considering
exact analytical solutions of the LL equation under various external actions, a general
theory with terms of the continuous spectrum as a starting point, as we will see in
the following, is necessary. It is evident that the system, under an external magnetic
field, the initial condition of the Landau-Lifschitz equation of a ferromagnet with
an anisotropy will be changed, and the system will become non-integrable. In this
Section, the effect of the magnetic field on the integrability of the system will be
discussed and single- and multisoliton solutions of the equation of motion will be
built [34, 60–64].

12.2.2 Macroscopic Description and Equation of Motion

The physical model to be investigated in this Section is the classical ferromagnet


with two single-ion anisotropies in an external magnetic field. For the macroscopic
description, the dynamics of the classical ferromagnet is assumed to be determined by
giving at each point of the magnetization vector M = (Mx , M y , Mz ). For a classical
ferromagnet with two single-ion anisotropies in an external magnetic field, the micro
magnetism (alias energy of the ferromagnet) E, assumed to be composed of three
terms, namely, an external magnetic field E ex , an anisotropic energy E an , and a
Zeeman energy E Z , takes the following form [29]

 ! 
α ∂M ∂M 3 1
E = E ex + E an + EZ = d x − βx Mx2 d 3 x
2 k
∂ x k ∂ x k 2
 
1
− βz Mz2 d 3 x − μ B M · Bd 3 x, (12.18)
2

where μ B is the Bohn magneton. The integral of motion of Eq. (12.18) reads

M2 ≡ M02 = constant.

The quantity M0 in the ground state coincides with the spontaneous magnetization

2μ B S
M0 =
a3
12.2 Nonlinear Magnetization Dynamics of a Classical Ferromagnet 405

with the atomic spin S and inter-atomic spacing a. In the limit bβx = 0, the biaxial
anisotropic ferromagnet will reduce into an uniaxial anisotropic ferromagnet with
an anisotropy axis coinciding with the z axis: when βz > 0 [βz < 0], an anisotropy
is of an easy-axis type and its magnetization vector in the ground state is directed
along the z axis [lies in the easy plane in the absence of an external magnetic field
and can be directed arbitrarily in this plane]. In the absence of the anisotropic energy
(that is, E an = 0), a crystal is called an isotropic ferromagnet.
As a function of space coordinates and time, the magnetization vector of the
classical ferromagnet M(x, t), function of both the spatial coordinates x and time t,
is governed by the Landau-Lifschitz equation

∂M 2μ B δE
= M× . (12.19)
∂t  δM
Assuming the spatial coordinate x and time t to be measured in unit of respectively
l0 = (a/βz )1/2 and ω0 = (2μ B βz M0 ) /, we deduce from Eqs. (12.18) and (12.19)
the following equation of motion
 2 
∂M ∂ M
=M× + J M + μ B B , with J = diag(Jx , Jy , Jz ), (12.20)
∂t ∂x2

in which the matrix J is related to the anisotropic constants. The terms proportional
to the Bohn magneton μ B of Eq. (12.20) describes various external actions such as
the magnetic field (the case to be treated in this Section), the magnetic impurities,
and the dissipative loses. It has been established that the motion equation (12.20)
in the special case B = 0 is exactly integrable [48]; in this case, Eq. (12.20) can be
reduced into a sine-Gordon equation in the limit Jx  Jy  Jz , when oscillations
M are localized near an easy plane yz. In the limit B = 0 the motion equation (12.20
turns to a NLS equation in the limit Jx ≈ Jy  Jz , when oscillations of M are
localized in the vicinity of the vacuum state M(x, t) = (0, 0, M0 ). Also, the motion
equation for an isotropic ferromagnet in an external magnetic field is completely
integrable, if βz = 0 [54]. The motion equation (12.20) under a zero magnetic field
is equivalent to a NLS equation [36]. Therefore, the motion equation (12.20) is the
most general motion equation describing the nonlinear magnetization dynamics of
classical ferromagnet with two single-ion anisotropies in an external magnetic field
[61].
We start by considering the effect of an external magnetic field on the integrabil-
ity of the motion equation (12.20). For magnetic fields with either an easy-plane or
an easy-axis (that is, with rotational symmetry), we can, by going over to a rotat-
ing coordinate frame, generalize the analysis of results of work [48]. Introducing a
stereographic projection P(x, y) of the unit sphere of magnetization vector onto a
complex plane as follows [46, 65]

Mx + i M y
P (x, y) = , (12.21)
1 + Mz
406 12 Engineering Magnetic Solitons in Nonlinear Systems

it follows from the motion equation Eq. (12.20) that


     
1 − P ∗ x P, P ∗ − 1 − P 2 ∗x P, P ∗ = 0,
2

     
−i 1 + P ∗  y P, P ∗ − i 1 + P 2 ∗y P, P ∗ = 0,
2
(12.22)
   
P ∗ z P, P ∗ − P∗z P, P ∗ = 0,

where i with i = x, y, z  y and z can be written as


" #2
  ∂P ∂2 P ∂2 P
i P, P ∗ = i 1 + |P|2 + 1 + |P|2 − 2P ∗ + 2Ji P 1 − |P|2
∂t ∂x2 ∂x2
1 1 
+μ B 1 + |P|2 Bx 1 − P2 + i B y 1 + P2 − Bz P , (12.23)
2 2

where
Jx = Jz − Jy , Jy = Jy − Jx , Jz = Jy − Jx .

For system (12.23) to be consistent, we must have


   
i P, P ∗ = 0 and i∗ P, P ∗ = 0. (12.24)

Under the restrictions (12.24), the evolution equation for the stereographic projection
P(x, t) given by Eq. (12.21) in the presence of the general direction of an external
magnetic field is found to be
 2 2
 
2 ∂P
  2
2 ∂ P ∗ ∂ P
 
i 1 + |P| + 1 + |P| − 2P + 2Ji P 1 − |P|2
∂t ∂x 2 ∂x 2
 
  1   1  
+μ B 1 + |P| 2
B 1 − P + i B 1 + P − B P = 0.
x 2 y 2 z
(12.25)
2 2

For analyzing the effect of an external magnetic field on the integrability of the motion
equation, we will use the evolution equation (12.25) of the stereographic projection.
When B x = B y = 0, that is, B = [0, 0, B z (t)] corresponding to an external field
directed along an anisotropic axi, we can remove the magnetic field term from Eq.
(12.25) by using the following gauge transformation
  
$ = P exp iμ B dt B z (t) ;
P→P

the system then becomes integrable. It is evident that in the situation when only
B y (t) = 0, that is, B = [0, B y (t), 0] (the case when the magnetic field is transverse),
the above gauge transformation cannot free Eq. (12.25) of the magnetic field term.
In this case of B = [0, B y (t), 0], the combined Galilean and gauge invariance of
12.2 Nonlinear Magnetization Dynamics of a Classical Ferromagnet 407

the Landau Lifschitz equation will be broken, there will not exist Lax pairs, and the
system will be non-integrable.
Although the introduction of an angular variable $ϕ = ϕ − ω B t in the polar coor-
dinates (θ, ϕ) free Eq. (12.20) in terms (θ , $
ϕ ) from B, it is important to note that
NMDs of the classical ferromagnet with an easy-axis plane is very sensitive to an
external magnetic field, independently on the nature of the magnetic field (weak or
strong). A perpendicular to an easy-axis plane external magnetic field does not alter
the axial symmetry corresponding to the z-axis; in such a situation, the form of the
ground state will depend on the strength of an external field. Let us denote by Bc the
critical of the external magnetic field B:

[(Jx − 2Jz )M]


Bc = .
μB

The magnetization vector M in the ground state under the condition that Bz < Bc
will deviate from an easy plane; in such a situation, M will be characterized by an
inclination θ = θ0 = arccos(B z /Bc ) to the z axis (the ground state is referred to as
“easy cone” when Bz < Bc ). The angle ϕ remains arbitrary. With the increase in the
external magnetic field, the angular opening of the easy cone becomes smaller, and
vector M in a non-excited ferromagnet with an easy plane lies along the z-axis.
Let us consider, in the context of the experiments [63, 64], the situation when
either B = [B x (t), 0, 0] or B = [0, B y (t), 0], corresponding to the situation when
an external magnetic field lies in an easy plane. In this context, we consider as samples
of a ferromagnet with an easy plane, Cs Ni F3 and (C6 H11 NH3 )CuBr3 , and assumed
an external field to be applied as a rule in an easy plane. With these two samples,
we obtain that the presence of an external field, which lies in an easy plane, it is
not easy to find exact soliton solutions of the corresponding LL equation. It is not
possible to free equation (12.25) from the magnetic-field term by using the above
gauge transformation. This means that in general, a ferromagnet with a uniaxial
anisotropy in a transverse magnetic field is non-integrable; it may become integrable
only in the absence of either an external field or an anisotropic interaction.
If we introduce the following two equations for 2 × 2 matrices (x, t; μ, λ)

∂(x, t; μ, λ)
= L (μ, λ) (x, t; μ, λ),
∂x
∂(x, t; μ, λ)
= A (μ, λ) (x, t; μ, λ), (12.26)
∂t
where coefficients L (μ, λ) and A (μ, λ) in the Lax pairs, functions of the spectral
parameters λ and μ, are defined as

L (μ, λ) = −iρns (μ, λ) Mx σx − iρds (μ, λ) M y σ y − iρcs (μ, λ) Mz σz ,


A (μ, λ) = 2iρ 2 ds (μ, λ) cs (μ, λ) Mx σx + 2iρ 2 ns (μ, λ) cs (μ, λ) M y σ y
+2iρ 2 ns (μ, λ) ds (μ, λ) Mz σz
408 12 Engineering Magnetic Solitons in Nonlinear Systems
 
∂ Mz ∂ My
−iρns (μ, λ) M y − Mz σx
∂x ∂x
 
∂ Mx ∂ Mz
−iρds (μ, λ) Mz − Mx σy (12.27)
∂x ∂x
 
∂ My ∂ Mx
−iρcs (μ, λ) Mx − My σz ,
∂x ∂x

with σi (i = x, y, z) being the Pauli metrics, ns(μ, λ), ds(μ, λ), and cs(μ, λl)
being elliptical functions, while μ and ρ being defined as

Jy M − Jx M 1
μ= , ρ= √ ,
2ρ 2 Jz M − Jx M

it is obvious that the motion equation (12.20) may be represented as a compatibility


condition
∂L ∂A
− + [L , A] = 0
∂t ∂x
of Eq. (12.26) with data (12.27). In the following, we restrict ourselves to the situation

when λ satisfies the conditions |Reλ| ≤ 2K and |Imλ|  ≤ 2K , K (μ) being a complete
elliptic integral of the first kind and K  (μ) = K 1 − μ2 . This restriction on λ
follows from the double-periodicity of coefficients L (μ, λ) and A (μ, λ) as function
of λ.
When the spectral parameters λ and μ satisfy the following relation
%
μ2 + 4ρ 2 , for βz < 0 (an easy plane)
λ =
2
(12.28)
μ2 − 4ρ 2 , for βz > 0 (an easy axis),

with ρ given as
% 1√
4 √(Jx
− Jz ) M, for βz < 0 (an easy plane),
ρ= (12.29)
1
4 (Jz − Jx ) M, for βz > 0 (an easy plane),

the Lax pairs for an uniaxial anisotropic ferromagnet in an external magnetic field
can be rewritten as follows:

L (μ, λ) = −iμMx σx − iμM y σ y − iλMz σz ,


 
∂ Mz ∂ My
A (μ, λ) = 2iμλMx σx + 2iμλM y σ y + 2iμ Mz σz − iμ M y
2
− Mz σx
∂x ∂x
   
∂ Mx ∂ Mz ∂ My ∂ Mx
−iμ Mz − Mz σ y − iλ Mz − Mz σz ,(12.30)
∂x ∂x ∂x ∂x

Next, we introduce a Riemann surface. For this aim, we assumed the spectral param-
eters λ and μ to be functions of a new introduced variable k, called the affine variable:
12.2 Nonlinear Magnetization Dynamics of a Classical Ferromagnet 409
& &
(2ρ (k 2 +1)) , 4ρk
, for an easy plane,
λ = λ(k) = k 2 −1 μ = μ(k) = k 2 −1 (12.31)
k −ρ k 2 +ρ 2
, for an easy axis.
2 2

k
, k

Then, we associate with the equation of motion (12.20) two types of BCs, namely,
the boundary condition of the first type, applies at x → ±∞ and the BC of the second
type, at y → ±∞.
The BC of the first type

M|x→±∞ = M0 = (0, 0, M0 ) (12.32)

corresponds to a breatherlike solution (alias magnetic soliton). Integrating Eq. (12.26)


under the BC of the first type (12.32) yields the Jost solutions 0± (x, k), chosen as
0± (x, k) → E(x, k)as x → ±∞, where

E(x, k) = exp[−iρcs(k)M0x σz ], (12.33)

while %   '
2ns(k)ds(k)
0± (x, k) exp −iρcs(k)M0 x − t σz
cs(k)
( 
for Im k ∈ 0, 2K  . E(x, t) given by Eq. (12.33) contains two independent solu-
tions E 1 (x, k) and E 2 (x, k):

E 1 (x, k) = (E 11 (x, k), E 21 (x, k))T , E 2 (x, k) = (E 12 (x, k), E 22 (x, k))T .

Also, each of 0+ (x, k), 0 (x, k), and 0− (x, k) has two independent solutions,
namely, 0+1 (x, k) and 0+2 (x, k), 01 (x, k) and 02 (x, k), and 0−1 (x, k) and
0−2 (x, k), respectively.
Under an external magnetic field, vector M in the ground state of a ferromagnet
with an easy plane will deviate from an easy plane, and will be characterized by
an inclination θ0 [φ0 ] to the z− axis [x-axis], where the asymptotic magnetization
vector M lies on the surface of an easy cone. In this situation, the simplest solution
of the equation of motion (12.20) is found to satisfy the BC of the first type

M|x→±∞ = M0 = (M0 sin θ0 cos φ0 , M0 sin θ0 sin φ0 , M0 cos [θ0 ]). (12.34)
p
To the boundary condition (12.34) correspond Jost solutions 0± (x, k) of Eq. (12.26)
that may be chosen as
p
0± (x, k) → E p (x, k) as x → ±∞,

where

2ρk 2ρk  
E p (x, k) = exp −i 2 M0 sin [θ0 x] cos [φ0 σx ] − i 2 M0 sin [θ0 x] sin φ0 σ y
k −1 k −1
410 12 Engineering Magnetic Solitons in Nonlinear Systems

ρ k2 + 1  
−i M0 cos θ0 xσz ⎦ , (12.35)
k2 − 1

while
&    
p 2ρk 2ρ k 2 + 1
0 (x, k) = exp −i 2 M0 sin [θ0 ] cos [φ0 ] x − t σx
k −1 k2 − 1
   
2ρk 2ρ k 2 + 1
−i 2 M0 sin [θ0 ] sin [φ0 ] x − t σy
k −1 k2 − 1
    )
ρ k2 + 1 8ρk 2
−i M0 cos [θ0 ] x − 4 t σz .
k2 − 1 k −1

As it has been stated above, the magnetization will be far from an easy plane when an
external magnetic field increases; in the special case of B z  Bc , the magnetization
will lie along the z axis. When the magnetic fields vanish, magnetization will lie on
an easy plane and takes the form

M0 = (M0 cos [φ0 ] , M0 sin [φ0 ] , 0).

As in the above analysis, E p (x, k) given by Eq. (12.35) contains two independent
p p
solutions E 1 (x, k) and E 2 (x, k) each of which has two components:

p  p p T p  p p T
E 1 (x, k) = E 11 (x, k), E 21 (x, k) , E 2 (x, k) = E 12 (x, k), E 22 (x, k) .
p p p
The same, each of solutions 0+ (x, k), 0 (x, k), 0− (x, k) has two independent
p p p p p
solutions 0+1 (x, k) and 0+2 (x, k), 01 (x, k) and 02 (x, k), and 0−1 (x, k) and
p
0−2 (x, k), respectively.
Because the z axis is an easy axis in a ferromagnet, we choose the boundary
condition of the first type as

M|x→±∞ = M0 = (0, 0, M0 ). (12.36)

To the BC of the first type (12.36) correspond Jost solutions 0±


a
(x, k) of Eq. (12.26)
which may be chosen as

0±
a
(x, k) → E a (x, k) as x → ±∞,

where  
k2 − ρ2
E a (x, k) = exp −i M0 xσz , (12.37)
2k

while
12.2 Nonlinear Magnetization Dynamics of a Classical Ferromagnet 411
&   2 2  )
k2 − ρ2 k + ρ2
0a (x, k) = exp −i M0 x −  2  t σz .
2k k k − ρ2

Similarly, E a (x, k) given by Eq. (12.37) has two independent solutions E 1a (x, k) and
E 2a (x, k), each of which has two components
 a T  a T
E 1a (x, k) = E 11 (x, k), E 21
a
(x, k) , E 2a (x, k) = E 12 (x, k), E 22
a
(x, k) ,

while each of 0+ a


(x, k), 0a (x, k), and 0−
a
(x, k) has two independent solutions
0+1 (x, k) and 0+2 (x, k), 01 (x, k) and 02 (x, k), and 0−1
a a a a a
(x, k) and 0−2
a
(x, k),
respectively.
Using the standard procedures of characteristic theory yields the following integral
representation

∞
+ (x, k) = E(x, k) + λ dy K + (x, y)E(y, k),
x
(12.38)
x
− (x, k) = E(x, k) + λ dy K − (x, y)E(y, k),
−∞

where the kernels K + (x, y) and K − (x, y) implicitly depend on magnetization M(x),
but do not depend on eigenvalue λ; moreover kernels K + (x, y) and K − (x, y) satisfy
the following BC of the second type

K ± (x, y)| y=±∞ = 0.

p
For a ferromagnet with an easy plane in an external magnetic field, ± (x, k) can
be also taken from the following integral representation
  ∞
p ρ k2 + 1 p,d
+ (x, k) = E (x, k) +
p
dy K + (x, y)E p (y, k)
k2 − 1
x
∞
2ρk p,nd
+ dy K + (x, y)E p (y, k),
k2 −1
x
(12.39)
  x
p ρ k2 + 1 p,d
− (x, k) = E (x, k) +
p
dy K − (x, y)E p (y, k)
k2 − 1
−∞
x
2ρk p,nd
+ dy K − (x, y)E p (y, k),
k2 −1
−∞
412 12 Engineering Magnetic Solitons in Nonlinear Systems

p
where kernels K ± (x, y) satisfy the BC of the second type

K ± (x, y) y=±∞ = 0.
p

In the case of a ferromagnet with an easy axis in an external magnetic field, ±a (x, k)
can be taken as
∞
k2 − ρ2
+a (x, k) = E (x, k) +
a
dy K +a,d (x, y)E a (y, k)
2k
x
∞
k2 + ρ2
+ dy K +a,nd (x, y)E a (y, k),
2k
x
(12.40)
x
k2 − ρ2
−a (x, k) = E a (x, k) + dy K −a,d (x, y)E a (y, k)
2k
−∞
x
k2 + ρ2
+ dy K −a,nd (x, y)E a (y, k),
2k
−∞

where kernels K ±a,d (x, y) satisfy the BC of the second type



K ±a (x, 0) y=±∞ = 0.

In the above Eqs. (12.39) and (12.40), the superscripts d and nd denote respectively
the diagonal and nondiagonal parts of the matrix.

12.2.3 Soliton Solutions

Using the results of the previous subsection, we investigate here soliton solutions.
These soliton solutions can be built using the inverse scattering problem, which
consists of constructing the magnetization vector M(x, t) from the time-dependent
scattering data. Such an inverse scattering problem can be carried out either by means
of the Gel’fand-Levitan-Marchenko (GLM) equation or a linear integral equation.
The GLM equation in the reflectionless case (case where the reflectional coeffi-
cient r (k, t) = 0) reads

K 11 (x, t) + K 12 (x, t) N  (x, t) = 0,


(12.41)

K 12 (x, t) − G (x, t) − K 11 (x, t) N (x, t) = 0,
12.2 Nonlinear Magnetization Dynamics of a Classical Ferromagnet 413

where K 11 (x, t) and K 12 (x, t) are defined as


&   )
det I + N  (x, t) M  (x, t)
K 11 (x, t) = i −1 ,
det [I + N  (x, t) N  (x, t)]
(12.42)

T

 
det I + N (x, t) N (x, t) + H (x) G (x, t)
K 12 (x, t) = − 1,
det [I + N  (x, t) N  (x, t)]

with
M  (x, t) = N  (x, t) + i H (x)T G (x, t),

N  (x, t) and N  (x, t) being two N × N matrices.


For obtaining  K i j (x,

t), we will use the

 Binet
Cauchy formula

 to
compute

det I + N (x, t) N (x, t) , det I + N (x, t) M (x, t) , and
T
det I + N  (x, t) N  (x, t) + H (x) G (x, t) .
First, let us set  
0 = det I + N  N  . (12.43)

Using the Binet-Cauchy formula yields


! ! !
0 = 1 + γ0 (n 1 , n 2 , ..., nr ; m 1 , m 2 , ..., m r ) .
r =1 1≤n 1 <n 2 <...<nr ≤N 1≤m 1 <m 2 <...<m r ≤N
(12.44)
Computing γ0 (n 1 , n 2 , ..., n r ; m 1 , m 2 , ..., m r ) for ferromagnet with two single-ion
anisotropies in an external magnetic field, we have

** * *
γ0 (n 1 , n 2 , ..., nr ; m 1 , m 2 , ..., m r ) = (−1)r f n f m αn αm
n m n<n  m<m 

 2
ρ 2 cs (kn ) − cs (kn  ) [cs (km ) − cs (km  )]2
×  2 , (12.45)
cs (km ) − cs (km )

where
* cs (kn ) − cs (km )
αn =   ,
m=n
ρ cs (km ) − cs (km ) cs (kn ) − cs (kn )

*
N
cs (kl )
fn = bn Hn2 .
l=1
cs (kl )
414 12 Engineering Magnetic Solitons in Nonlinear Systems

Computing now γ0 (n 1 , n 2 , ..., n r ; m 1 , m 2 , ..., m r ) for an uniaxial anisotropic ferro-


magnet in an external magnetic field, we arrive to

γ0 (n 1 , n 2 , ..., nr ; m 1 , m 2 , ..., m r )

2 2 2 2

r ++ + +
4 2 2 2

⎪ 4ρ 2 km kn −1 km −1 kn  −kn km  −km

⎪ (−1) f f α α ,


n m n m
km (km2 +1) k 2 −1
2 2
kn  −1
2
2 −1
2
2 −k 2


n m n<n  m<m  n km  km n

for an easy plane,
=  2 
2 2
 2  2

⎪ ++ + + 4ρ 2 kn km +1 kn kn  −1 −kn  kn −1
2 2
km km 2 −1 −k
m  km −1

⎪ (−1)r f n f m αn αm

,



2

⎪ n m n<n  m<m  km kn +1 kn (km
2 2 −1)−k 2
m kn −1

for an easy axis.
(12.46)

where


⎪ + (kn2 −1) kn 2 −1 km 2 −1 (km2 −kn2 )

⎨ 2 2 , for an easy plane,
m=n 2ρ (km −1) kn −kn km −kn
2 2 2
αn =

⎪ + (kn2 −1) kn −1 km −1 [kn (km2 −1)−km (kn2 −1)]
2 2



2
2 , for an easy axis,
m=n 4ρ (km −1) kn km −1 −km (kn −1) kn kn −1 −kn (kn −1)
2 2 2 2

(12.47)
⎧ N 2 2
⎪ + (kl +1) kl +1

⎨ 2 bn H 2 , for an easy plane,
l=1 (kl −1) kl −1
2 n
fn =

⎪ +N
kl (kl2 −1)
⎩ b H 2 , for an easy axis.
2 n n
l=1 kl kl −1

Secondly, let us introduce 1 by


 
1 = det I + N  M  ; (12.48)

computing the rhs of Eq. (12.48), we find that 1 can be written as


! ! !
1 = 1 + γ1 (n 1 , n 2 , ..., nr ; m 1 , m 2 , ..., m r ) .
r =1 1≤n 1 <n 2 <...<nr ≤N 1≤m 1 <m 2 <...<m r ≤N
(12.49)

Evaluating γ1 (n 1 , n 2 , ..., n r ; m 1 , m 2 , ..., m r ) for the classical ferromagnet with two


single-ion anisotropies in an external magnetic field, we obtain
** * *
γ1 (n 1 , n 2 , ..., nr ; m 1 , m 2 , ..., m r ) = (−1)r f n f m αn αm
n m n<n  m<m 

 2
ρ 2 cs (km ) cs (kn ) − cs (kn  ) [cs (km ) − cs (km  )]2
×  2 .
cs (kn ) cs (km ) − cs (km )
(12.50)
12.2 Nonlinear Magnetization Dynamics of a Classical Ferromagnet 415

Computing now γ1 (n 1 , n 2 , ..., n r ; m 1 , m 2 , ..., m r ) for an uniaxial anisotropic ferro-


magnet in an external magnetic field, we arrive to

γ1 (n 1 , n 2 , ..., nr ; m 1 , m 2 , ..., m r )


2 2 2
2
++ + +
2 2

⎪ ρ 2 km kn −1 kn  −kn km  −km
2

⎪ (−1)r f n f m αn αm ,

⎪ n m n<n  m<m  kn (km −1) kn  −1 km  −1 km −kn
2 2 2 2

2 2


for an easy plane,
=
⎪  2  2
2 2
 2  2 (12.51)
⎪ (−1)r + + + + f f α α
2
⎪ 16ρ 2 km +1 kn −1 kn kn  −1 −kn  kn −1 2 −1 −k
km km  m  km −1


2 ,

⎪ n m n<n  m<m 
n m n m
(km2 −1) kn +1 kn (km2 −1)−km kn −1
2 2


for an easy axis.

Third, let us turn to the evaluation of expression



T

det I + N  (x, t) N  (x, t) + H (x) G (x, t)

in Eq. (12.42). In order to calculate this expression, we will introduce a N × (N + 1)


matrix Q  and (N + 1) × N matrix Q  , as follows

Q nm = Nnm

, Q n0 = −i Hn , Q nm = Nnm

, Q n0 = i G n , n, m = 1, 2, ..., N .
 
Then one can write det I + Q  Q  as
  ! ! !
det I + Q  Q  = 1 + Q  (n 1 , n 2 , ..., nr ; m 1 , m 2 , ..., m r )
r=1 1≤n 1 <n 2 <...<nr ≤N 1≤m 1 <m 2 <...<m r ≤N

×Q  (m 1 , m 2 , ..., m r ; n 1 , n 2 , ..., nr ) , (12.52)

where the sum is decomposed into two parts: one part for m 1 = 0 and the second
one for to m 1 ≥ 1. Except for the same extended to m 1 = 0, Eq. (12.52) coincides
with (12.43). Therefore,

   
det I + Q  Q  − det I + N  N 

! !
=
r=1 1≤n 1 <n 2 <...<nr ≤N
!
Q  (n 1 , n 2 , ..., nr ; 0, m 2 , ..., m r ) Q  (0, m 2 , ..., m r ; n 1 , n 2 , ..., nr ) . (12.53)
1≤m 1 <m 2 <...<m r ≤N

Setting now    
2 = det I + Q  Q  − det I + N  N  , (12.54)
416 12 Engineering Magnetic Solitons in Nonlinear Systems

we find that
! ! !
2 = γ2 (n 1 , n 2 , ..., n r ; 0, m 2 , ..., m r ) .
r =1 1≤n 1 <n 2 <...<nr ≤N 1≤m 1 <m 2 <...<m r ≤N
(12.55)
Evaluating γ2 for the classical ferromagnet with two single-ion anisotropies in an
external magnetic field, we obtain

** * *
γ2 (n 1 , n 2 , ..., n r ; 0, m 2 , ..., m r ) = (−1)r +1 f n f m αn αm
n m n<n  m<m 

 2
ρ 2 cs (km ) cs (kn ) − cs (kn  ) [cs (km ) − cs (km  )]2
×  2 . (12.56)
cs (kn ) cs (km ) − cs (km )
Calculating nor γ2 for an uniaxial anisotropic ferromagnet in an external magnetic
field, we arrive to

γ2 (n 1 , n 2 , ..., nr ; 0, m 2 , ..., m r )

⎧ 2
2 2 2
2
++ + +
2 2

⎪ ρ 2 km kn −1 kn  −kn km  −km
2

⎪ (−1)r +1 f n f m αn αm 2 2 2 ,

⎪ n m n<n  m<m  kn (km2 −1) k 2 −1 2 −1 2 −k 2

⎪ n
km  km n

for an easy plane,
=
⎪  2  2
2 2
 2  2 (12.57)
⎪ (−1)r +1 + + + + f f α α
2
⎪ 16ρ 2 km +1 kn −1 kn kn  −1 −kn  kn −1 2 −1 −k
km km  m  km −1


2 ,

⎪ n m n<n  m<m 
n m n m
(km −1) kn +1 kn (km −1)−km kn −1
2 2

2 2

for an easy axis.

Here, f n = exp [−1n + i2n ].


For obtaining K 11 and K 12 , it is sufficient to insert Eqs. (12.43), (12.48), and
(12.54) into Eq. (12.42). To obtain multisoliton solutions in the classical ferromagnet
with two single-ion anisotropies in an external magnetic field, we use the following
relations
   −1
M (x, t) = i K (x, x, t) − σz σz i K (x, x, t) − σz , (12.58)

resulting to
 
21 2
(Mn )x = Re ,
|1 |2 + |2 |2
 
21 2
(Mn ) y = Im , (12.59)
|1 |2 + |2 |2
12.2 Nonlinear Magnetization Dynamics of a Classical Ferromagnet 417

|1 |2 − |2 |2
(Mn )z = M0 − .
|1 |2 + |2 |2

In the case of an uniaxial anisotropic ferromagnet in an external magnetic field,


the multisoliton solutions can be written as follows:


⎨ M0 sin θ0 cos φ0 − Re 21 2
, for an easy plane,

|1 | +|2 |
2 2
(Mn )x =
⎩ Re 2 
1 2
, for an easy axis;
| |2 +| |21 2



⎨ M0 sin θ0 cos φ0 − Im 21 2
, for an easy plane,

|1 | +|2 |
2 2
(Mn ) y = (12.60)
⎩ Im 21 2
, for an easy axis;
| |2 +| |21 2

& |1 |2 −|2 |2


M0 cos θ0 − |1 |2 +|2 |2
, for an easy plane,
(Mn )z = |1 |2 −|2 |2
M0 − |1 |2 +|2 |2
, for an easy axis.

If we consider the z axis as the polar axis in the polar coordinates, we arrive to
the following multisoliton solutions of the classical ferromagnet with two single-ion
anisotropies in an external magnetic field:

2 |2 |2
cos [θ ] = , ϕ = − arg [2 ] − arg [1 ] . (12.61)
|1 |2 + |2 |2

For an uniaxial anisotropic ferromagnet, the multisoliton solutions are found to be




⎨ cos [θ0 ] − 2|2 |2
, %
|1 |2 +|2 |2 − arg [2 ] − arg [1 ] , for an easy plane,
cos [θ ] = ϕ= (12.62)

⎩ 1 − 2| 2|
2
− arg [2 ] − arg [1 ] for an easy axis
|1 |2 +|2 |2

1 and 2 being given respectively by Eqs. (12.49) and (12.55).


n = 1 corresponds to the single-soliton solutions of the classical ferromagnet with
two single-ion anisotropies in an external magnetic field which can take the following
form:

ns (k1 )2 sinh [1 ] sin [2 ] + ns (k1 )2 cosh [1 ] cos []
(M1 )x = ,
4ns (k1 )2 cosh2 [1 ] + 4ds (k1 )2 sinh2 [1 ] + cs (k1 )2

ds (k1 )2 sinh [1 ] cos [2 ] − ds (k1 )2 cosh [1 ] sin [2 ]
(M1 ) y = , (12.63)
4ns (k1 )2 cosh2 [1 ] + 4ds (k1 )2 sinh2 [1 ] + cs (k1 )2
418 12 Engineering Magnetic Solitons in Nonlinear Systems

2cs (k1 )2


(M1 )z = M0 − ,
4ns (k1 )2 cosh2 [1 ] + 4ds (k1 )2 sinh2 [1 ] + cs (k1 )2

where

1 = 2ρcs (k1 ) (x − V1 t − x10 ) , 2 = 2ρcs (k1 ) (x − V2 t − x20 ) (12.64)

and  
 2 cs (k1 )2 − cs (k1 )2 + 4ρ 2
V1 = 4cs (k1 ) , V2 = . (12.65)
ρcs (k1 )

Computing now the single-soliton (one-soliton) solutions of a uniaxial anisotropic


ferromagnet in an external magnetic field for an easy plane and easy axis, we arrive
respectively to


 2
2k12 4k12 + k12 − 1 sin2 [2 ]
(M1 )x = M0 sin [θ0 ] cos [φ0 ] −    ,
k 2 − 12 k 2 cosh2 [1 ] + k 2 sin2 [2 ]
1 1 1

(M1 ) y = M0 sin [θ0 ] sin [φ0 ]


   
2k1 k1 4k1 k1 sinh [1 ] cos [2 ] + |k1 |4 − 1 cosh [1 ] sin [2 ]
−  2    , (12.66)
k − 12 k 2 cosh2 [1 ] + k 2 sin2 []2
1 1 1
(M1 )z = M0 cos [θ0 ]
     
4k1 k1 k1 |k1 |2 + 1 sinh [1 ] sin [2 ] − k1 k12 − 1 cosh [1 ] cos [2 ]
−  2    ,
k − 12 k 2 cosh2 [1 ] + k 2 sin2 [2 ]
1 1 1

and
2
16k12 k12 sinh [1 ] sin [2 ] + |k1 |4 − 1 cosh [1 ] cos [2 ]
(M1 )x =  2  2 ,
|k1 |4 − 1 cosh2 [1 ] + 16k12 k12 sinh2 [1 ] + 4k12 |k1 |2 + 1

2
16k12 k12 sinh [1 ] cos [2 ] − |k1 |4 − 1 cosh [1 ] sin [2 ]
(M1 ) y =  2  2 , (12.67)
|k1 |4 − 1 cosh2 [1 ] + 16k12 k12 sinh2 [1 ] + 4k12 |k1 |2 + 1

2
2k12 |k1 |2 + 1
(M1 )z =  2  2 .
|k1 |4 − 1 cosh2 [1 ] + 16k12 k12 sinh2 [1 ] + 4k12 |k1 |2 + 1
12.2 Nonlinear Magnetization Dynamics of a Classical Ferromagnet 419

Here,
⎧  2
⎨ 8ρk1 (2|k1 | 2+1) (x − V1 t − x10 ) , for an easy plane,

1 = 8ρk| (1|k1 ||2 +1)
k −1

⎩ 1
(x − V1 t − x10 ) , for an easy axis,
|k12 −1|2
(12.68)

⎪ 8ρk1 (|k1 |2 −1)
⎨ (x − V2 t − x20 ) , for an easy axis,
2 = |k12 −1| 2

⎪ 8ρk1 (|k1 |2 −1)


⎩ (x − V2 t − x20 ) , for an easy plane,
|k12 −1|2


⎨ 2ρ [(|k1 | −1)2+4k12 (|k1 | −1)] , for an easy plane,
⎪ 4 2 2

|k1 −1|
V1 = 2ρk  (|k1 |2 −1
⎪ )
⎩ 1 2 2 , for an easy axis,
|k1 −1|
(12.69)


⎪ 2ρ 4k12 (|k1 |2 −1) −4k12 (|k1 |2 +1) +|k1 −1|4
2 2

⎨ , for an easy axis,
k1 (|k1 |2 −1)|k12 −1 |
2
V2 =

⎪ 2ρ (|k1 | −1) −4k1 (|k1 | +1)


2 2 2 2

⎩ , for an easy plane.
(|k1 |2 −1)|k12 −1|2

The found above results reveal that the nonlinear magnetization of the classical
ferromagnet with an anisotropy under the action of an external magnetic field depends
essentially on the two velocities V1 and V2 given in Eqs. (12.65) and (12.69), the
center of nonlinear magnetization moving with a constant velocity V1 , while its shape
changes with any change in the velocity V2 . For a better understanding, we display,
with the use of the exact soliton solutions given in Eqs. (12.66) and (12.67), in
Figs. 12.4, 12.5, 12.6 and 12.7 typical graphical illustrations showing the motion of
the center and the change of shape of the z− component of nonlinear magnetization
field (M1 )z . Figures 12.4 and 12.6 [12.5 and 12.7] are associated with a ferromagnet
with an easy plane [axis].
By taking now the z− axis as the polar axis in the polar coordinates, we can write
the one-soliton solutions as follows

2cs (k1 )2


cos [θ ] = 1 − ,
4ns (k1 )2 cosh2 [1 ] + 4ds (k1 )2 sinh2 [1 ] + cs (k1 )2
(12.70)
ds (k1 )2 sinh [1 ] cos [2 ] − ds (k1 )2 cosh [1 ] sin [2 ]
tanh [ϕ] = .
ns (k1 )2 sinh [1 ] sin [2 ] + ns (k1 )2 cosh [1 ] cos [2 ]

In the case of an uniaxial anisotropic ferromagnet, the one-soliton solutions for an


easy plane and for an easy axis can be written respectively as
420 12 Engineering Magnetic Solitons in Nonlinear Systems

Fig. 12.4 Some graphical illustrations of the motion of the center and the change of shape of the
z component of the nonlinear magnetization (M1 )z defined by Eq. (12.66) in a ferromagnet with
an easy plane. The plot is generated with the following numerical values of parameters: θ0 = 30◦ ,
ρ = 0.1, k1 = 0.1, k1 = 0.2, x10 = 0, and x20 = 0. Reprint from Ref. [29], Copyright 2022, with
permission from American Physical Society

Fig. 12.5 Some graphical illustrations of the motion of the center and the change of shape of the
z component of the nonlinear magnetization (M1 )z given by Eq. (12.67) in a ferromagnet with an
easy axis. For generating this figure, we have used the following values of parameters: ρ = 0.1,
k1 = 0.1, k1 = 0.2, x10 = 0, and x20 = 0. Reprint from Ref. [29], Copyright 2022, with permission
from American Physical Society
12.2 Nonlinear Magnetization Dynamics of a Classical Ferromagnet 421

Fig. 12.6 Some graphical illustrations of the motion of the center and the change of shape of the z
component of the nonlinear magnetization (M1 )z expressed by Eq. (12.66) in a ferromagnet with
an easy plane. To generate this plot, we have used the following numerical values of parameters:
θ0 = 30◦ , ρ = 0.3, k1 = 0.1, k1 = 0.2, x10 = 0, and x20 = 0. Reprint from Ref. [29], Copyright
2022, with permission from American Physical Society

Fig. 12.7 Some graphical illustrations of the motion of the center and the change of shape of the
z component of the nonlinear magnetization (M1 )z given by Eq. (12.67) in a ferromagnet with an
easy axis. The plot is generated for the numerical values of parameters taken as: ρ = 0.3, k1 = 0.1,
k1 = 0.2, x10 = 0, and x20 = 0. Reprint from Ref. [29], Copyright 2022, with permission from
American Physical Society
422 12 Engineering Magnetic Solitons in Nonlinear Systems

cos [θ ] = cos [θ0 ]


     
2 k1 |k1 |2 + 1 sinh [1 ] sin [2 ] + k1 |k‘ |2 − 1 cosh [1 ] cos [2 ]
−  2    ,
k − 12 k 2 cosh2 [1 ] + k 2 sin2 [2 ]
1 1 1
(12.71)
 
sin [θ0 ] sin [φ0 ] − 8k1 k1 sinh [1 ] cos [2 ] − 2 |k1 − 1 cosh [1 ] sin [2 ]
|4
tan [ϕ] =
 2 ,
sin [θ0 ] cos [φ0 ] − 2k12 4k12 + k12 − 1 sin2 [2 ]

and
 
2k12 |k1 |2 + 1
cos [θ ] = 1 −  2  2 ,
|k1 |4 − 1 cosh2 [1 ] + 16k12 k12 sinh2 [1 ] + 4k12 |k1 |2 + 1
(12.72)
  2
16k12 k12 sinh [1 ] cos [2 ] − |k1 |4 − 1 cosh [1 ] sin [2 ]
tan [ϕ] =  2 .
16k12 k12 sinh [1 ] sin [2 ] + |k1 |4 − 1 cosh [1 ] cos [2 ]

The following property if found:

cos (−x, −t) = cos (x, t) , (12.73)

meaning that the z component Mz of the nonlinear magnetization M in the presence


of an external magnetic field is a symmetric function of both x and t; on the other
hand, property (12.73) in terms of the orientation means that the orientation of the
nonlinear magnetization M in the plane orthogonal to the anisotropic axis changes
with an external field, and will turn to a constant in the absence of an external field.
Now, let us analyze the feature of the above found soliton solutions. If we set all
values as zero (x10 = 0, x20 = 0) in the moving coordinates of the soliton, we obtain

2cs (k1 )2


cos [θ ] = 1 −   ,
4ns (k1 ) cosh 2ρcs (k1 ) x
2 2
 
+ 4ds (k1 )2 sinh2 2ρcs (k1 ) x + cs (k1 )2
(12.74)
   
ds (k1 )2 sinh 2ρcs (k1 ) x cos 2ρcs (k1 ) (x − V2 t)
   
− ds (k1 )2 cosh 2ρcs (k1 ) x sin 2ρcs (k1 ) (x − V2 t)
tan [ϕ] =     ,
ns (k1 )2 sinh 2ρcs (k1 ) x sin 2ρcs (k1 ) (x − V2 t)
   
+ ns (k1 )2 cosh 2ρcs (k1 ) x cos 2ρcs (k1 ) (x − V2 t)

and, for respectively an easy plane and an easy axis,


12.2 Nonlinear Magnetization Dynamics of a Classical Ferromagnet 423
        
2k1 |k1 |2 + 1 8ρk1 |k1 |2 + 1 8ρk1 |k1 |2 − 1
sinh  x tan   (x − V2 t)
|k1 |4 − 1 k 2 − 12 k 2 − 12
  1    1
2k1 |k1 |2 − 1 8ρk1 |k1 |2 + 1
+ cosh  2  x
|k1 |4 − 1 k − 12
1
cos [θ] = cos [θ0 ] − ⎧
  

⎪ ,

 2 2 ⎨ 2 8ρk1 |k1 |2 +1 8ρk1 |k1 |2 −1 ⎬
k −1 k cosh2 x +k 2 sin2 (x−V2 t)
1 ⎪

1
|1 |
k 2 −1
2 1
|1 |
k 2 −1
2 ⎪


  
4k1 k1 8ρk1 |k1 |2 + 1
tan [ϕ] = sin [θ0 ] sin [φ0 ] − sinh  2  x
|k1 |4 − 1 k − 12
1
       
8ρk1 |k1 |2 + 1 8ρk1 |k1 |2 − 1
− cosh  2  x tan  2  (x − V2 t) ,
k − 12 k − 12
1 1
(12.75)

and
2
4k12 |k1 |2 + 1
2  2
|k1 |4 − 1
cos [θ ] = 1 − ⎡ ⎤ ⎡ ⎤ 2 ,
8ρk1 |k1 |2 +1 16k 2 k 2 8ρk1 |k1 |2 +1 4k12 |k1 |2 +1
cosh2 ⎣   x ⎦ + 1 1 2 sinh2 ⎣   x⎦ + 2
 2 2  2 2
k1 −1 |k1 |4 −1 k1 −1 |k1 |4 −1
⎡ ⎤ ⎡ ⎤ ⎡ ⎤
8ρk1 |k1 |2 +1 2 8ρk1 |k1 |2 +1 8ρk1 |k1 |2 −1
16k12 k12 sinh⎣  2 x ⎦− |k1 |4 −1 cosh⎣  2 x ⎦ tan⎣  2 (x−V2 t)⎦
 2   2   2 
k1 −1 k1 −1 k1 −1
tan [ϕ] = ⎡ ⎤ ⎡ ⎤ ⎡ ⎤.

8ρk1 |k1 | +1
2 
8ρk1 |k1 | −1
2 2 8ρk1 |k1 |2 +1
16k12 k12 sinh⎣   x ⎦ tan ⎣   (x−V 2 t)⎦ + |k 1 | 2 −1 cosh ⎣   x⎦
 2 2  2 2  2 2
k1 −1 k1 −1 k1 −1
(12.76)

We have also found that the amplitudes and widths of the MSs under the action of
an external magnetic field are periodic function of time t. With the use of Eqs. (12.75)
and (12.76), we have showed in Fig. 12.8 some graphical illustrations of the change
of amplitude and width of the z component (M1 )z of the nonlinear magnetization M
in the case of an easy plane. It is seen from Fig. 12.8 that the amplitude and shape of
the (M1 )z also changes with a velocity V2 and, the shape of soliton is not symmetrical
with respect to the soliton center. In the case of an easy axis, the shape of the soliton
is symmetrical with respect to the soliton center, as we can clearly see from Fig. 12.9.
It is evident that at the limit ρ → 0, all soliton solutions in an uniaxial anisotropic
ferromagnet reduce to those in an isotropic ferromagnet. If for example, setting
ρ → 0 in Eqs. (12.66) and (12.67) yields
424 12 Engineering Magnetic Solitons in Nonlinear Systems

Fig. 12.8 Some graphical illustrations of the change of amplitude and width of the z component of
the nonlinear magnetization (M1 )z generated with the use of the expression given in Eq. (12.75) in
a ferromagnet with an easy plane, for the following numerical values of the parameters: θ0 = 30◦ ,
ρ = 0.2, k1 = 0.1, k1 = 0.2, x10 = 0, and x20 = 0. Reprint from Ref. [29], Copyright 2022, with
permission from American Physical Society

Fig. 12.9 Some graphical illustrations of the change of amplitude and width of the z component
of the nonlinear magnetization (M1 )z obtained with the help of the expression given in Eq. (12.76)
in a ferromagnet with an easy axis, for the following parameters: ρ = 0.2, k1 = 0.1, k1 = 0.2,
x10 = 0, and x20 = 0. Reprint from Ref. [29], Copyright 2022, with permission from American
Physical Society
12.2 Nonlinear Magnetization Dynamics of a Classical Ferromagnet 425

Fig. 12.10 Some graphical illustrations of the motion of the center and the change of shape of
the z component of the nonlinear magnetization (M1 )z generated with the use of Eq. (12.77) in
an isotropic ferromagnet. Different parameters used in this plot are taken as: ρ = 0, k1 = 0.1,
k1 = 0.2, x10 = 0, and x20 = 0. Reprint from Ref. [29], Copyright 2022, with permission from
American Physical Society

Fig. 12.11 Some graphical illustrations of the amplitude and width of the z component of the
nonlinear magnetization (M1 )z given in Eq. (12.78) in an isotropic ferromagnet, which do not
change periodically with time, for the parameters ρ = 0, k1 = 0.1, k1 = 0.2, x10 = 0, and x20 = 0.
Reprint from Ref. [29], Copyright 2022, with permission from American Physical Society
426 12 Engineering Magnetic Solitons in Nonlinear Systems

Fig. 12.12 Some graphical illustrations of the change of the z component of the total magnetic
momentum Pz defined in Eq. (12.80) in a ferromagnet with an easy plane. This plot is generated with
the numerical values of the parameters taken as: θ0 = 30◦ , ρ = 0.1, k1 = 0.1, k1 = 0.2, x10 = 0,
and x20 = 0. Reprint from Ref. [29], Copyright 2022, with permission from American Physical
Society

2k1   
(M1 )x = sech2 k1 x − 4k1 t − x10
|k1%|2      
  k 2
× k1 sinh k1 x − 4k1 t − x10 sin k1 x − 2 k1 − 1 t − x20
k1
     '
  k 2
+k1 cosh k1 x − 4k1 t − x10 cos k1 x − 2 k1 − 1 t − x20 ,
k1
2k1   
(M1 ) y = sech2 k1 x − 4k1 t − x10 (12.77)
|k1%|2      
  k 2
× k1 sinh k1 x − 4k1 t − x10 cos k1 x − 2 k1 − 1 t − x20

k1
     '
  k 2
−k1 cosh k1 x − 4k1 t − x10 sin k1 x − 2 k1 − 1 t − x20 ,
k1
2k    
(M1 )z = M0 − 12 sech2 k1 x − 4k1 t − x10 .
|k1 |

These results coincide with Eq. (27a) in Ref. [55], obtained by means of the
method of an inverse scattering transformation (IST). Our study also shows that the
center and shape of the z component Mz of nonlinear magnetization under the action
of an external magnetic field do not move with the two velocities V1 and V2 , as it is
clearly seen from Fig. 12.10. If we consider the z axis as the polar axis in the polar
coordinates, we arrive to

2k12   
cos [θ ] = 1 − sech2 k1 x − 4k1 t − x10 ,
|k1 |2
 " # 
k 2
tan [ϕ] = ϕ0 + k1 x − 2k1 1 − 12 t − x20
k1
1
+  . (12.78)
k1   
tan tanh k1 x − 4k1 t − x10
k1
12.2 Nonlinear Magnetization Dynamics of a Classical Ferromagnet 427

Fig. 12.13 Some graphical illustrations of the change of the z component of the total magnetic
momentum Pz defined by Eq. (12.80) in a ferromagnet with an easy axis, for the following numerical
values of parameters: ρ = 0.1, k1 = 0.1, k1 = 0.2, x10 = 0, and x20 = 0. Reprint from Ref. [29],
Copyright 2022, with permission from American Physical Society

Equation (12.78) means that under an external magnetic field, the amplitudes and
widths of the z component Mz of the nonlinear magnetization M are not periodic
functions of time t. For a better understanding, we depict in Fig. 12.11 some graph-
ical illustrations of the amplitudes and width of component (M1 )z of the nonlinear
magnetization M defined by Eq. (12.78) in the case of an isotropic ferromagnet. In
the limit t → 0, the obtained here results are equivalent to Eq. (22) obtained in Ref.
[33] with the method of separating variables in moving coordinates.
Evaluating the total magnetic momentum, we find

P = M0 d x (1 − cos θ ) ∇ϕ, (12.79)

which is not a constant under the action of an external magnetic field, but depends
on time. We also find that

Pz = M0 d x (1 − cos θ ) (12.80)

is not a constant. In Figs. 12.12 and 12.13, we show some graphical illustrations of
the Pz (z component of the total magnetic momentum P) defined by Eq. (12.80)
for respectively an easy plane and an easy axis. Pz varies periodically with time in
an anisotropic ferromagnet with an easy plane (Fig. 12.12), but is aperiodic in time
for an anisotropic ferromagnet with an easy axis (Fig. 12.13). In the latter case, Pz
decreases as a function of time t and has the sense of the mean number of spins
428 12 Engineering Magnetic Solitons in Nonlinear Systems

deviated from the ground state in localized magnetic excitations. This feature has
not yet been observed in the study of all other nonlinear problems in magnetism [33].

12.2.4 Asymptotic Behavior of Multisoliton Solutions

Assuming kn to be all positive and k1 > k2 > ... > k N , we denote by n the vicinity
of x = xin + Vin t (i = 1, 2). For large enough time t, we separate these vicinities
from left to right as  N ,  N −1 , ..., 1 . In  j , there are the following limits:

(x − Vin t − xin0 ) → −∞ and | f n | → ∞, if n < j;


(x − Vim t − xim0 ) → ∞ and | f m | → ∞, if m > j,

while

0 ∼ γ0 (1, 2, ..., j − 1; 1, 2, ..., j − 1) + γ0 (1, 2, ..., j; 1, 2, ..., j) ,


1 ∼ γ1 (1, 2, ..., j − 1; 1, 2, ..., j − 1) + γ1 (1, 2, ..., j; 1, 2, ..., j) ,
2 ∼ γ2 (1, 2, ..., j; 0, 1, 2, ..., j − 1) .

To obtain the asymptotic behavior of multisoliton solutions, we substitute these


asymptotic expressions into Eqs. (12.43), (12.48), and (12.54) and obtain, for the
classical ferromagnet with two single-ion anisotropies in an external magnetic field,
the following relations
       
cs k j ns k j  2 ns k j  2 2cs k j
 
0 ∼ 1 +     F j ,  1 ∼ 1 +  
  F j , 2 ∼   Fj ,
cs k j ns k j ns k j ns k j

where       
*
j−1
*
N
cs k j − cs (kn ) cs k j − cs (km )
Fj =       .
n=1 m= j+1
cs k j − cs (kn ) cs k j − cs (km )

In the same way, we find, for an uniaxial anisotropic ferromagnet in an external


magnetic field, for 0 , 1 , 2 , and F j that
⎧ ⎧
2



⎪   k j k j 2 +1 ⎪


 2 k 2 −1 k j +1
⎪ 1 +  F j 2
⎪ , for an easy axis, ⎪
⎪ 1 + F j  j , for an easy axis,
⎨ k j k 2j +1 ⎨ k 2 +1 k −1
2
0 ∼ 1 ∼ j j
⎪ 2 ⎪ 2 2
⎪ 1 +  F 2 k j k j +1 , for an easy axis,





 2
 
k +1 k j −1
j , for an easy axis,

⎩ j
k k +1
2 ⎩ 1 + Fj
⎪ 2
k 2j −1 k j +1
j j
12.2 Nonlinear Magnetization Dynamics of a Classical Ferromagnet 429

4k j k j k j −1
2



⎪ F   , for an easy axis,

⎪ j  2
⎨ k j k 2j −1
2 ∼  2 
(12.81)
 
⎪ 4k j k j −1 k j  +1
2



⎪ 2 , for an easy axis,


Fj
2

 
k j +1 k 2j −1

⎪ 2 −1 k 2 −1 k 2 −k 2 k 2 −k 2
⎪ f +
⎪ j−1 + N km n n j m j

⎪ , for an easy axis,
⎨ j n=1 m= j+1 k 2 −1 km 2 −1 kn 2 −k 2 k 2 −k 2
Fj = n
j m j


⎪ +
j−1 +
N 2 −1 k 2 −1 k k 2 −1 −k k 2 −1 k k 2 −1 −k k 2 −1
km


n j n


n j

j m


m j
, for an easy axis.

⎩ fj 2 2 2 2 2 2
n=1 m= j+1 kn −1 km −1 k j kn −1 −kn k j −1 k j km −1 −km k j −1

From the above results, we affirm the existence of multisoliton solutions in a strict
sense for the classical ferromagnet with two single-ion anisotropies in an external
magnetic field. At the limit t → ±∞, M appears to be the trains of N separating
one-solitons. We find that the trains

(i) at t → −∞ turn out to be trains


(ii) at t → +∞, the number and the shape of MSs remain unchanged after the col-
lision, while the position of center after the collision is displaced in the traveling
coordinates.

In the course of time from t → −∞ to t → ∞, the total displacement of the center


of the jth peak is calculated and is given by

⎧ ⎡    ⎤ ⎡    ⎤ ⎫
⎨ *  cs k j − cs (kn ) 
j−1 *N  cs k − cs (k )  ⎬
1  j m 
Xj =   ln ⎣    ⎦ − ln ⎣    ⎦ . (12.82)
cs k j ⎩ n=1
 cs k j − cs (kn ) 
m= j+1
 cs k j − cs (km )  ⎭

 
It is obvious that the angle tan [ϕ] = arctan M y /Mx contains a linear term in
time t, meaning that the x− and the y− components Mx and M y of the nonlinear
magnetization manifest themselves as solitons. Evaluating the total phase shift of the
jth peak, we arrive to
⎧ ⎡   ⎤ ⎡   ⎤⎫
⎨ j−1
* cs k j − cs (kn ) *N
cs k j − cs (km ) ⎬
 j = 2 arg ⎣   ⎦ − arg ⎣   ⎦ . (12.83)
⎩ cs k j − cs (kn ) cs k j − cs (km ) ⎭
n=1 m= j+1

The total displacement of the center X j and the total phase shift  j of the jth peak in
the course of time from t → −∞ to t → +∞ for a uniaxial anisotropic ferromagnet
in an external magnetic field are found to be
430 12 Engineering Magnetic Solitons in Nonlinear Systems
⎧   ⎧    ⎫
 2 2 ⎨ j−1  2 2 2   2 2 −1 ⎬

⎪ k j −1 +  kn −k j kn −1  +
N  km −k 2j km

⎪ −   , for an easy plane,


ln
2ρk j k j ⎩ n=1  k 2 −1 kn 2 −k 2 
ln  2 2 
m= j+1  km −k j km −1 ⎭
⎪ 2

⎪ n j

⎪  2 ⎡ 


⎨  2   2 2 2 
k j −1 +  kn −1 k j kn −1 −kn k j −1 
j−1
   ⎣ln
Xj =
⎪   2  2
2
2 −1 

⎪ 2ρk j k j  +1 n=1  k n −1 k j k n −1 −k n k

⎪ 
⎤
j

⎪ 


2
 km −1 k j km 2 −1 −k k 2 −1 

⎪ +
N

m j
⎦ , for an easy axis,

⎩ − ln  2
2
2 
m= j+1  km −1 k j km −1 −km k j −1 
(12.84)
⎧ ⎧ ⎡ ⎤ ⎡ ⎤⎫
⎪ ⎨ 2 2
+ kn −k j kn −1
j−1 2
+N 2
km −k 2j km 2 −1 ⎬

⎪ ⎣ ⎦
− arg ⎣ ⎦ , for an easy plane,
⎪ 2 arg


⎪ ⎩ k 2 −1 k 2 −k 2 k 2 −k k
2
−1 ⎭
⎪ ⎡ ⎛n=1 n

n

j
m= j+1 m

j m


⎨ +
j−1 k n
2
−1 k k
j n
2 −1 −k k 2 −1
n j
 j = 2 ⎣arg ⎝


⎪ 2 2
n=1 kn −1 k j kn −1 −kn k j −1
2

⎪ ⎛
⎞⎤



⎪ +N 2
km −1 k j km 2 −1 −k k 2 −1

⎪ ⎝

m j
⎠⎦ , for an easy axis.

⎩ − arg 2
2 2
m= j+1 km −1 k j km −1 −km k j −1

At the limit ρ → 0, the displacement X j of the center and the phase shift  j of
the jth peak of an isotropic ferromagnet in an external magnetic field are found to be
⎛    ⎞
1 ⎝ *  kn − k j  * k − k 
j−1 N
 m j ⎠
X j =  ln   − ln   ,
kj  kn − k j   km − k j 
n=1 m= j+1

(12.85)
⎡ " j−1  # ⎛  ⎞⎤

* kn kn − k j *N
km km − k j
 j = 2 ⎣arg − arg ⎝ ⎠⎦ ,
n=1
kn − k j m= j+1
k m − k j

respectively. These results coincide with Eqs. (28a) and (28b) of work [55], obtained
by means of the method of an IST.

12.2.5 Conclusion

By means of the stereographic projection we have presented the integrable conditions


of the equation of motion for the NMDs of a classical ferromagnet with two single-
ion anisotropies in an external magnetic field. Exact analytical soliton solutions,
including Jost soliton and multisoliton solutions are built. The effect of a magnetic
field on the magnetic single-solitons as well as magnetic multisoliton is investigated
in detail. We found that the presence of an external magnetic field affects essentially
two parameters of the nonlinear magnetization:
(i) one of the two parameters affects the motion of the wave center moving with a
constant velocity, while another parameter affects the wave shape;
12.3 Nonlinear Magnetization Dynamics in the Presence of Spin-Polarized Current 431

(ii) the amplitude and width of the nonlinear magnetization are periodic function of
time, while the wave shape, also a function of time, is unsymmetric with respect
to its center.
The total magnetic momentum P is derived and its z component Pz are studied in
detail. Studying the asymptotic behavior of multisoliton solutions in the long-time
limit, we have calculated the total displacement X j of the center as well as the phase
shift  j of the jth peak.

12.3 Nonlinear Magnetization Dynamics in the Presence


of Spin-Polarized Current

In this Section, we discuss both a linear and nonlinear magnetization dynamics in a


uniaxial ferromagnetic nanowire injected with current. The nonlinear magnetization
dynamics is assumed to be described by a modified Landau-Lifshitz-Gilbert (LLG)
equation with additive terms. By means of a stereographic projection, exact analytical
soliton solutions of the LLG equations are obtained in both the cases of high and of
low quality factors Q. We investigate analytically the effect of spin transport on the
MSs and show that the spin current can change the soliton velocities and affect the
solitons collision with a phase shift. The effect of ac current on the wave center is
investigated too.

12.3.1 Introduction

In ferromagnetic multilayers, the spin-polarized current causes many unique phe-


nomena, such as spin-wave excitation, magnetization switching and reversal, and
enhanced Gilbert damping[66–71] [67–72]. In the mechanism for magnetization
reversal in multilayers, an electric current flows perpendicular to the layers [73, 74].
In the presence of spin-polarized current, the NMDs is mostly governed by a gen-
eralized LLG equation with additive terms related to either spin-polarized current,
spin accumulation, or spin transfer torque [72, 75–80].
This Section deals with the nonlinear excitations in a uniaxial ferromagnetic
nanowire injected with current [81, 82]. The NMDs of this model is assumed to
be described by the LLG equation. By means of a stereographic projection [83],
the model equation (LLG equation) is reduced into a nonlinear equation of com-
plex function whose soliton solutions for both a high-Q model and a low-Q one are
presented (here, Q is the so-called “quality factor”). With the use of these soliton
solutions, we investigate analytically the effect of a spin-polarized current on the
excitation of propagating solitons in the ferromagnetic metal nanowire.
432 12 Engineering Magnetic Solitons in Nonlinear Systems

12.3.2 Long Ferromagnetic Nanowire with a Uniform Cross


Section (High-Q Model, Q > 1)

We start our study with the situation of a very long ferromagnetic nanowire with
a uniform cross Section, as depicted in Fig. 12.14 [84]. The nanowire of Fig. 12.14
can be viewed as infinite in length. In this nanowire of Fig. 12.14, the x− and the
z-axes are assumed to be respectively the flow direction (long length of the wire)
and the directions of both the uniaxial anisotropy field and the external field. Next,
it is assumed that the electronic current flows in the x-direction. Throughout this
nanowire, the magnetization is assumed to be nonuniform only in the direction of
current, that is, in the x-direction. Therefore, the current injected in this physical
system can be polarized and produces a spin torque

∂M μB
τb = b J , with b J = P je (12.86)
∂x eMs

that acts on the local magnetization; here, P is the spin polarization of the current,
je stands for the density of the electric current, e is the magnitude of electron charge,
Ms is the saturation magnetization, μ B is the Bohr magneton, and M is the nonlinear
magnetization vector [46]. Under all these considerations, the modified LLG equation
with the spin torque 12.86 reads

∂M α ∂M
= −γ M × He f f + M× + τb . (12.87)
∂t Ms ∂t

Here, He f f is the effective magnetic field including the external field, the anisotropy
field, the demagnetization field, and the exchange field, α is the Gilbert damping
parameter, and γ is the gyromagnetic ratio. In the context of this subsection, the
effective field He f f will be assumed to be
 2   
2A ∂ 2M HK
He f f = + − 4π Mz + Hext ez , (12.88)
Ms ∂x2 Ms

where HK , Hext , A, and ez are respectively the anisotropy field, the applied field,
the exchange constant, and the unit vector along the z-direction. Throughout this
Section, we denote by Q the “quality factor”, defined as

Hk
Q= . (12.89)
4π Ms

Dependent on the value of the quality factor Q, we will focus our attention in two
special cases, the case Q > 1 corresponding to “Long ferromagnetic nanowire”, and
the case when Q < 1 which corresponds to low or “short ferromagnetic nanowire”.
It follows from Eq. (12.89) that the situation Q > 1 corresponds to the case when
the uniaxial anisotropy field is larger than the magnetization field (this situation
12.3 Nonlinear Magnetization Dynamics in the Presence of Spin-Polarized Current 433

Fig. 12.14 Geometry of the


nanowire. Reprint from Ref.
[84], Copyright 2022, with
permission from American
Physical Society

occurs for some metallic magnetic films [85] and multilayers [86]). Under the con-
dition Q > 1, dark magnetic solitons for the Mz component can be obtained. Before
turn to the search of exact soliton solutions of the LLG equation (12.87), let us rescale
the time coordinate and space coordinate by characteristic time as follows:

1 2A
t0 = , l0 = .
γ (HK − 4π Ms ) (HK − 4π Ms ) Ms

Letting M = Ms m, the LLG equation (12.87) takes the following simplified form
[84]
" #  
∂m ∂2m Hext ∂m b J t0 ∂m
=− m× − m z + (m × ez ) + αm × + . (12.90)
∂t ∂x2 H K − 4π Ms ∂t l0 ∂ x

Next, we assume that [83]  2


M
m =2
= 1. (12.91)
MS

Equation (12.91) means that the magnetization magnitude is a constant at temper-


atures well below the Curie temperature.Taking into consideration the condition
(12.91), we introduce the stereographic transformation
mx my
ψ= +i , (12.92)
1 + mz 1 + mz

where m x , m y , and m z are respectively the x−, y−, and z− component of the
magnetization vector m. We then find that the complex function ψ is described by
the following equation:
 2
∂ψ ∂ 2ψ ψ∗ ∂ψ Hext
i (1 − iα) = − 2 − ψ
∂t ∂x2 1 + ψψ ∗ ∂x HK − 4π Ms
434 12 Engineering Magnetic Solitons in Nonlinear Systems

1 − ψψ ∗ b J t0 ∂ψ
− ∗
ψ +i . (12.93)
1 + ψψ l0 ∂ x

In the following, we will get linear magnetization dynamics (LMDs) and NMDs
from the complex function ψ.
Following Li and Zhang [76], we take the spin wave (SW) ansatz
 
m = δm x ex + δm y e y exp [i (k · r − ωt)] + ez , (12.94)

where δm x and δm y stand for the rescaled amplitude of the SWs in respectively the
x and y directions, and ω is a complex angular frequency. Inserting ansatz (12.94)
into Eq. (12.92) yields
 
δm x ex + δm y e y
ψ = ψ0 exp [i (k · r − ωt)] with ψ0 = . (12.95)
2
Inserting now ψ from Eq. (12.95) into Eq. (12.93) and linearizing the resulting equa-
tion with respect with ψ0 yields the dispersion relation
 
ω = (1 + iα) ω0 1 + l02 (k x − b J Ms /γ A)2 , (12.96)

where
γ (HK − 4π Ms + Hext ) − b2J Ms /2γ A
ω0 = ,
1 + α2
(12.97)
16γ 2 A2
l02 =  2 ,
8γ Ms A (HK − 4π Ms + Hext ) − Ms2 b2J

b J being given in Eq. (12.86). Using Eqs. (12.86) and (12.97), one can easily verify
that the spin-polarized current changes the energy gap. The gap vanishes for large
current density
e 
je ≥ 2 AL s (HK − 4π Ms + Hext ).
P
This means that without other stimulation, such as thermal effects, the spin-polarized
current excites SWs when its density increases beyond a critical value. The Gilbert
damping gives the energy gap a rescaling and damps the SW.
Now, we turn to the investigation of the nonlinear excitation of the LLG equations.
For investigating the nonlinear excitation of LLG equations, we assume that |ψ|2 
1, or, its equivalent form, m 2x + m 2y  m 2z . This condition corresponds to a deviations
of magnetization from the equilibrium direction along the anisotropy axis. Let us
assume that the wavelength l0 and the wavenumber k of the SW satisfy the condition
l0 k  1 [87]. By keeping only the nonlinear terms of the order of the magnitude
of |ψ|2 ψ, we can obtain equation (12.93) in a more simple form. Neglecting the
damping, we arrive to the following NLS equation with one linear derivative term
12.3 Nonlinear Magnetization Dynamics in the Presence of Spin-Polarized Current 435
 
∂ψ ∂ 2ψ t0 ∂ψ Hext
i = +2 |ψ| 2
ψ + ib J − 1+ ψ. (12.98)
∂t ∂x2 l0 ∂ x HK − 4π Ms

Applying the Hirota bilinear method (HBM) [88], we can easily obtain one- and
two-soliton solutions of Eq. (12.98). The one-soliton solution is
 %   '
kR Hext
ψ(x, t) =   exp i η I + 1 + t , (12.99)
cosh η R + C2 HK − 4π Ms

where
     
t0 2 |λ|
η = k x + b J − ik t , exp [C] = exp , (12.100)
l0 4k 2R

subscripts R and I denoting respectively the real and the imaginary parts, and k is a
free complex parameter. Generally, the free complex parameter k can be determined
with the help of the initial soliton location and amplitude. Using now the expression of
ψ given in Eq. (12.99), we obtain the solution of magnetization though its components
as
     
2k R Hext C
mx = cos η I + 1 + t cosh η R + ,
 HK − 4π Ms 2
     
2k R Hext C
my = sin η I + 1 + t cosh η R + , (12.101)
 HK − 4π Ms 2
  
1 C
mz = −k 2R + cosh2 η R + ,
 2
 
where  = k 2R + cosh2 η R + C2 .
The one-soliton solution (12.101) represents a 2π domain wall (DW) [89]. The
height of this one-soliton coincides with the rescaled magnetization, and its velocity
V is defined as
l0
V = −b J − 2k I , with k I = Im [k] .
t0

From the expression for the velocity V , it is clearly seen that the spin-polarized current
alters the soliton velocity as −b J , proportional to the electron current density. From
the solution (12.101), we can see that the x− and the y− component of the nonlinear
magnetization vector m, m x and m y are oscillate periodically with respect with time
t. This behavior constitutes one of the effects of the external fields on the nonlinear
magnetization vector m. If b J , instead of being a constant, is an alternating currents
(ac), that is, has the form
P je μ B
bJ = cos [ωt] ,
eMs

the only change to the solutions will be


436 12 Engineering Magnetic Solitons in Nonlinear Systems
 
bJ
η=k x+ sin [ωt0 t] − ikt .
ωl0

In this situation, the center of the mass of solitons vibrate with the same frequency
as the alternating currents. Considering the effect of nonlinear excitations on the
spin be polarized current, the spin-current density in ferromagnetic material with
nonuniform magnetization and adiabatic approximation will then be [76]
μB
jx = P je m,
e
so that the single-soliton excitation in nanowire gives spin current a magnetic pulse
like the optical pulse in nonlinear media.
Following the same procedure, the two-soliton solutions are obtained to be
   
g2 + g3 Hext
ψ= exp i 1 + t ,
1 + f2 + f4 HK − 4π Ms

and NMDs
 T
 T ψ + ψ∗ ψ − ψ ∗ 1 − ψψ ∗
mx , m y, mz = , −i , , (12.102)
1 + ψψ ∗ 1 + ψψ ∗ 1 + ψψ ∗

where

g1 = λ1 exp [η1 ] + λ2 exp [η2 ] ,


    
g3 = exp [η1 + η2 ] exp η1∗ + θ1 + exp η2∗ + θ2 , (12.103)
 ∗
  ∗
  ∗

f 2 = exp η1 + η1 + R1 + exp η2 + η2 + R2 + exp η1 + η2 + δ
 ∗ 
+ exp η1 + η2 + δ ∗ ,
 
f 4 = exp η1 + η1∗ + η2∗ + η2 + R3 ,

with

|λ1 |2 λ2 (k1 − k2 )2 |λ2 |2 λ1 (k1 − k2 )2


exp [θ1 ] =     2
, exp [θ1 ] =   2 ,
k1 + k1∗ k1∗ + k2 k2 + k2∗ k2∗ + k1
|λ1 |2 |λ2 |2
exp [R1 ] =  2
, exp [R2 ] =  2 ,
k1∗ + k1 k2∗ + k2
λ1 λ∗2
exp [δ] =  2 , (12.104)
k1 + k2∗
 4
|λ1 |2 |λ2 |2 k1∗ − k2∗ 
exp [R3 ] =  2  2  4 ,
k 1 + k ∗ k 2 + k ∗ k 1 + k ∗ 
1 2 2
ηi = ki [x + (b J t0 /l0 − iki ) t] ,
12.3 Nonlinear Magnetization Dynamics in the Presence of Spin-Polarized Current 437

Fig. 12.15 Evolution of dark two-solitons showing the elastic collision of two dark solitons of
m z component for k1 = 1 + 0.5i, k2 = 2 − 0.5i, λ1 = 0.4 + 2i, λ2 = 2i. Numerical values of
the materials parameters of CoPt 3 are taken as: HK = 33778 Oe, A = 1.0 × 10−6 erg/cm, Ms =
1125 G, γ = 1.75 × 107 Oe−1 s−1 , P = 0.35. The numerical values of other parameters used
for generating this figure are: Hext = 2000 Oe, je = 106 A/cm2 , leading to t0 = 2.91 × 10−12
s, l0 = 3.01 × 10−7 cm, b J = 18 cm/s. Reprint from Ref. [84], Copyright 2022, with permission
from American Physical Society

k1 , k2 , λ1 , and λ2 being four complex parameters which can be determined by the


initial conditions. The two-soliton solution in the limits of time t → ±∞ is reduced
to two one-solitons having the same form as those given in Eq. (12.101). In other
words, the two-soliton solution is asymptotically combined with two single-solitons.
Carrying out the asymptotical investigation on the two-solitons, we found that mag-
netic two-solitons collide without amplitude switching, but with a displacement δφ
of the solitons center, where

R3 − R2 − R1
δφ = .
2
For a better understanding, we have showed in Fig. 12.15 the elastic collision of two
dark two-solitons of the z− component m z of the nonlinear magnetization vector m.
As in the case of the one-solitons associated with the exact solution (12.101), the
width and the amplitudes of the x− and y components m x and m y of the nonlinear
magnetization vector m vary periodically with time; this behavior of m x and m y
is due of the oscillating factor induced by the external and anisotropic fields. It is
seen from the analytical expression of the two-solitons solutions that the two-soliton
excitations bring the spin current two magnetic single-solitons propagating with
different velocities.
438 12 Engineering Magnetic Solitons in Nonlinear Systems

12.3.3 Short Ferromagnetic Nanowire with a Uniform Cross


Section (Low-Q Model, Q < 1)

Let us now consider the situation when the uniaxial anisotropy field is smaller than the
magnetization field such as in permalloy, that is, the situation when Q < 1 and discuss
the nonlinear solutions of Eq. (12.87). In this case, we introduce the characteristic
time t0 and length l0 as follows:

1 2A
t0 = , and l0 = .
γ (4π Ms − HK ) 4π Ms − HK

Following the same procedure as in the previous subsection, we transform the LLG
Eq. (12.87) into the following nonlinear Schrödinger equation with one linear deriva-
tive term and an external linear potential
 
∂ψ ∂ 2ψ t0 ∂ψ Hext
i = − 2 |ψ| 2
ψ + ib J + 1 − ψ. (12.105)
∂t ∂x2 l0 ∂ x 4π Ms − HK

The single-soliton solution of Eq. (12.105) reads


  
ζ
ψ = −λ exp [i] 1 + exp [i] + (1 − exp [i]) tanh ,
2

where
 
t0 Hext
 = kx + 2λ2 + k 2 + kb J + 1 + t + 0 ,
l0 4π Ms − HK
 √ 
P 4λ2 − P 2
 = arctan , (12.106)
P 2 − 2λ2
   
t0
ζ =P x− 4λ − P − 2k + b J
2 2 t + ζ0 ,
l0

k, λ, P, 0 , and ζ0 being arbitrary real constants. The corresponding exact analytical


solutions of the NMDs are then found to be
%  '
λ ζ
mx = cos [ + ] − cos [] + (cos [ + ] + cos []) tanh ,
 2
%  '
λ ζ
my = sin [ + ] − sin [] + (sin [ + ] + sin []) tanh ,
 2
(12.107)
%   '
1   ζ
mz = 4 1 − λ2 + P 2 1 − tanh2 ,
4 2
12.3 Nonlinear Magnetization Dynamics in the Presence of Spin-Polarized Current 439

where %   '
1   ζ
= 4 1 + λ2 − P 2 1 − tanh2 .
4 2

As we can see from Eq. (12.106) and solutions (12.107), the spin-polarized current
alters the soliton velocity and induces the oscillation of the two components m x and
m y of the nonlinear magnetization vector m.
The two-soliton solutions of Eq. (12.105) are found to be

λ (1 + g1 + g2 )
ψ =− exp [i] ,
1 + f1 + f2

f 1 , f 2 , g1 , and g2 being defined as

g1 = exp [i1 ] exp [ζ1 ] + exp [i2 ] exp [ζ2 ] ,


g2 = C exp [i (1 + 2 )] exp [ζ1 + ζ2 ] ,
f 1 = exp [ζ1 ] + exp [ζ2 ] , f 2 = A exp [ζ1 + ζ2 ] ,

where
   
t0
ζ j = Pj x − − 4λ2 − 2k + kb J
P j2 t + ζ j0 ,
l0

  
 j = arctan P j 4λ2 − P j2 / P j2 − 2λ2 , j = 1, 2,
  2
(P1 − P2 )2 + 4λ2 − P12 − 4λ2 − P22
C=   2 ,
(P1 + P2 )2 + 4λ2 − P12 − 4λ2 − P22

P1 , P2 , ζ10 , and ζ20 being four arbitrary real constants. The corresponding nonlinear
evolution of magnetization is defined as
 T
 T ψ + ψ∗ ψ − ψ ∗ 1 − ψψ ∗
mx , m y, mz = , −i , .
1 + ψψ ∗ 1 + ψψ ∗ 1 + ψψ ∗

By means of the asymptotical analysis, we lead to the conclusion that the magnetic
two-solitons collide without amplitude switching, but with phase shift δφ = ln [C].
After the collision of magnetic two-solitons, the solitons centers displace with ln [A].
We have depicted the evolution of two bright two-solitons in Fig. 12.16, showing the
elastic collision of two bright two-solitons of an m z component of the nonlinear
magnetization vector. The numerical values of the material parameters of permalloy
used in Fig. 12.16 are: A = 1.3 × 10−6 erg/cm, HK = 10 Oe, Ms = 800 G, je = 106
440 12 Engineering Magnetic Solitons in Nonlinear Systems

Fig. 12.16 The elastic


collision of two bright
solitons of an m z component
for P1 = 1.2, P2 = 3.9,
λ = 2, k = 1,
0 = 2, ζ10 = 1, ζ20 = 3.
Other parameters used for
generating this plot are given
in the text. Reprint from Ref.
[84], Copyright 2022, with
permission from American
Physical Society

A/cm2 , P = 0.5 and Hext = 2000 Oe. The numerical values of the characteristic time
t0 and length l0 are t0 = 4.5510 − 11 s, l0 = 1.1 × 410−6 cm, while b J = 51 cm/s.
As in the case Q > 1, let us take a look on the situation when the external magnetic
field deviates from the z axis; in such a situation, the LLG equations generally do not
have exact analytical solutions [90], except in some special cases. Let us consider
for example, the case when Q < 1 and Hx ex + Hz ez = Hext and Q < 1. Taking into
account the spin-polarized current, we find the following dynamic soliton solution

 T
mx , m y , mz
" #T
1 − D 2 cosh2 [ξ ]   1 − D 2 cosh2 [ξ ]   2D cosh [ξ ]
= cos γ H t
z 0 t , sin γ H t
z 0 t , ,
1 + D 2 cosh2 [ξ ] 1 + D 2 cosh2 [ξ ] 1 + D 2 cosh2 [ξ ]

where
  
 t0 γ Hx t0
ξ= 1 − γ t0 Hx x + bJ t , D = .
l0 1 − γ Hx t0

Analyzing this dynamic soliton, we can well see that (i) the depth and width of
the trough decrease with the increase in Hx , and (ii) the amplitudes of m x and m y
components of the nonlinear magnetization vector m periodically oscillate with time
with frequency γ Hz . According to the above found dynamic solution, when the
external field perpendicular to wire’s axis vanishes (Hext = 0), the magnetization
rotates purely in a plane parallel to the wire’s axis. This found dynamic solution is
displayed in Fig. 12.17 with the same material parameters as in Fig. 12.16: A = 1.3 ×
10−6 erg/cm, HK = 10 Oe, Ms = 800 G, je = 106 A/cm2 , P = 0.5 and Hext =
2000 Oe.
References 441

Fig. 12.17 Evolution of


Two neighbor bright
two-solitons of m z
component with Hy = 1000
Oe. Different parameters
used for generating this plot
are given in the text. Reprint
from Ref. [84], Copyright
2022, with permission from
American Physical Society

12.3.4 Conclusion

We have considered a uniaxial ferromagnetic nanowire injected with current whose


NMDs are described by a modified LLG equation with a spin torque. By means
of a stereographic projection, we have reduced this modified LLG equation with a
spin torque into a nonlinear partial differential equation of the complex function ψ.
Applying the HBM, we have built two families (one of high-Q and another of low-Q)
of single- and two-soliton solutions that describe NMDs in ferromagnetic nanowire
with spin-polarized current. Our results showed that the found MSs propagate along
the wire’s axis with time-varying velocities (due to the presence of spin torque),
the change in the wave velocity being proportional to the current density, while the
waves amplitudes are modulated by spin-polarized current. The centers of the MSs
are found to vibrate periodically in time with a frequency that coincides with that of
the alternating currents. Interestingly, we found that magnetic multisolitons solitons
collide elastically with phase shift.

References

1. C. Jia, R. Wu, Y. Hu, W.-M. Liu, Z. Liang, Dissipative Magnetic Polariton Soliton,
arXiv:1907.13300v1 (cond-mat.quant-gas) 31 Jul. 2019
2. B.A. Malomed, D. Mihalache, F. Wise, L. Torner, Spatiotemporal optical solitons. J. Opt. B
Quantum Semiclassical Opt. 7, R53 (2005)
3. L.M. Sieberer, M. Buchhold, S. Diehl, Keldysh field theory for driven open quantum systems.
Rep. Prog. Phys. 79, 096001 (2016)
4. E. Kengne, W.M. Liu, B.A. Malomed, Spatiotemporal engineering of matter-wave solitons in
Bose-Einstein condensates. Phys. Rep. 899, 1–62 (2021)
5. T.J. Kippenberg, A.L. Gaeta, M. Lipson, M.L. Gorodetsky, Dissipative kerr solitons in optical
mi croresonators. Science 361, 567 (2018)
6. T. Byrnes, N.Y. Kim, Y. Yamamoto, Exciton polariton condensates. Nat. Phys. 10, 803 (2014)
442 12 Engineering Magnetic Solitons in Nonlinear Systems

7. P. Cilibrizzi, H. Ohadi, T. Ostatnicky, A. Askitopoulos, W. Langbein, P. Lagoudakis, Linear


wave dynam ics explains observations attributed to dark solitons in a polariton quantum fluid.
Phys. Rev. Lett. 113, 103901 (2014)
8. F. Pinsker, Approximate solutions for half-dark solitons in spinor non-equilibrium polariton
conden sates. Ann. Phys. 362, 726 (2015)
9. L.A. Smirnov, D.A. Smirnova, E.A. Ostro vskaya, Y.S. Kivshar, Dynamics and stability of dark
solitons in exciton-polariton condensates. Phys. Rev. B 89, 235310 (2014)
10. N. Akhmediev, V.V. Afanasjev, Novel arbitrary amplitude soliton solutions of the cubic-quintic
complex ginzburg-landau equation. Phys. Rev. Lett. 75, 2320 (1995)
11. X. Ma, S. Schumacher, Vortex multistabil ity and bessel vortices in polariton condensates.
Phys. Rev. Lett. 121, 227404 (2018)
12. M. Sich, D.N. Krizhanovskii, M.S. Skolnick, A.V. Gor bach, R. Hartley, D.V. Skryabin, E.A.
Cerda-Méndez, K. Biermann, R. Hey, P.V. Santos, Observation of bright polariton solitons in
a semiconductor microcavity. Nat. Photon. 6, 50 (2011)
13. E. Wertz, L. Ferrier, D.D. Solnyshkov, R. Johne, D. San vitto, A. Lemaître, I. Sagnes, R.
Grousson, A.V. Ka vokin, P. Senellart, G. Malpuech, J. Bloch, Spontaneous formation and
optical manipulation of extended polariton condensates. Nat. Phys. 6, 860 (2010)
14. T.C.H. Liew, O.A. Egorov, M. Matuszewski, O. Kyri ienko, X. Ma, E.A. Ostrovskaya, Instabil-
ity induced formation and nonequilibrium dynamics of phase defects in polariton condensates.
Phys. Rev. B 91, 085413 (2015)
15. A. Askitopoulos, K. Kalinin, T.C.H. Liew, P. Cilibrizzi, Z. Hatzopoulos, P.G. Savvidis, N.G.
Berloff, P.G. Lagoudakis, Nonresonant optical control of a spinor po lariton condensate. Phys.
Rev. B 93, 205307 (2016)
16. X. Xingran, H. Ying, Z. Zhang, Z. Liang, Spinor polariton condensates under nonresonant
pumping: steady states and elementary excitations. Phys. Rev. B 96, 144511 (2017)
17. X. Xingran, L. Chen, Z. Zhang, Z. Liang, Dark-bright solitons in spinor polariton conden sates
under nonresonant pumping. J. Phys. B: At. Mol. Opt. Phys. 52, 025303 (2019)
18. I.A. Shelykh, A.V. Kavokin, Y.G. Rubo, T.C.H. Liew, G. Malpuech, Polariton polarization sen-
sitive phenomena in planar semiconductor microcavi ties. Semicond. Sci. Technol 25, 013001
(2010)
19. M. Sich, L.E. Tapia-Rodriguez, H. Sigurds son, P.M. Walker, E. Clarke, I.A. Shelykh, B. Royall,
E.S. Sedov, A.V. Kavokin, D.V. Skryabin, M.S. Skolnick, D.N. Krizhanovskii, Spin domains
in one-dimensional conser vative polariton solitons. ACS Photon. 5, 5095–5102 (2018)
20. S. Yuri, Kivshar, X. Yang, Perturbation induced dynamics of dark solitons. Phys. Rev. E 49,
1657 (1994)
21. O.A. Egorov, A.V. Gorbach, F. Lederer, D.V. Skryabin, Two-dimensional localization of exciton
po laritons in microcavities. Phys. Rev. Lett. 105, 073903 (2010)
22. F. Pinsker, H. Flayac, On-demand dark soliton train manipulation in a spinor polariton con-
densate. Phys. Rev. Lett. 112, 140405 (2014)
23. I. Danaila, M.A. Khamehchi, V. Gokhroo, P. Engels, P.G. Kevrekidis, Vector dark-antidark
solitary waves in multicomponent bose-einstein condensates. Phys. Rev. A 94, 053617 (2016)
24. T. Congy, A.M. Kamchatnov, N. Pavloff, Dispersive hydrodynamics of nonlinear polarization
waves in two-component Bose-Einstein condensates. SciPost Phys. 1, 006 (2016)
25. I. Carusotto, C. Ciuti, Quantum fluids of light. Rev. Mod. Phys. 85, 299 (2013)
26. Ph. Nozieres, D. Pines, The Theory of Quantum Liquids, vol. II (Addison Wesley, New York,
1990)
27. P. Grelu, N. Akhmediev, Dissipative soli tons for mode-locked lasers. Nat. Photon. 6, 84 (2012)
28. H.-G. Purwins, H.U. Bödeker, Sh. Amiranashvili, Dissipative solitons. Adv. Phys. 59, 485
(2010)
29. W-M. Liu, W.-S. Zhang, F.-C. Pu, X. Zhou, Nonlinear magnetization dynamics of the classical
ferromagnet with two single-ion anisotropies in an external magnetic field. Phys. Rev. B 60,
12893 (19999)
30. L.D. Landau, E.M. Lifschitz, On the theory of the dispersion of magnetic permeability in
ferromagnetic bodies. Phys. Z. Sowjetunion 8, 153 (1935)
References 443

31. K. Nakumura, T. Sasada, Solitons and wave trains in ferromagnets. Phys. Lett. 48A, 321 (1974)
32. M. Lakshmanan, T.W. Ruijgrok, C.J. Thompson, On the dynamics of a continuum spin system.
Phys. A 84, 577 (1976)
33. J. Tjon, J. Wright, Solitons in the continuous Heisenberg spin chain. Phys. Rev. B 15, 3470
(1977)
34. A.M. Kosevich, B.A. Ivanov, A.S. Kovalev, Nonlinear localized magnetization wave of a fer-
romagnet as a bound state of a large number of magnons. Pis’ma Zh. E ksp. Teor. Fiz. 25, 516
(1977) [JETP Lett. 25, 486 (1977)]
35. H.J. Mikeska, Solitons in a one-dimensional magnet with an easy plane. J. Phys. C 11, L29
(1978)
36. K.A. Long, A.R. Bishop, Nonlinear excitations in classical ferromagnetic chains. J. Phys. A
12, 1325 (1979)
37. V.E. Zakharov, L.A. Takhtajan, Equivalence of the nonlinear Schrödinger equation and the
equation of a Heisenberg ferromagnet. Theor. Math. Phys. 38, 17 (1979)
38. B.A. Ivanov, A.M. Kosevich, I.M. Babich, Localized nonlinear oscillations in ferromagnetic.
Pis’ma Zh. Eksp. Teor. Fiz. 29, 777 (1979) [JETP Lett. 29, 714 (1979)]
39. M.M. Bogdan, A.S. Kovalev, Exact multisoliton solution of one-dimensional Landau-Lifshitz
equations for an anisotropic ferromagnet. Pis’ma Zh. É ksp. Teor. Fiz. 31, 453 (1980) [JETP
Lett. 31, 424 (1980)]
40. M. Svendsen, H.C. Fogedby, Phase shift analysis of the Landau-Lifshitz equation. J. Phys. A
26, 1717 (1993)
41. K. Nakumura, T. Sasada, Gauge equivalence between one-dimensional Heisenberg ferromag-
nets with single-site anisotropy and nonlinear Schrodinger equations. J. Phys. C 15, L915
(1982)
42. A. Kundu, O. Pashaev, Comments on the gauge equivalence between Heisenberg spin chains
with single-site anisotropy and nonlinear Schrodinger equations. J. Phys. C 16, L585 (1983)
43. G.R.W. Quispel, H.W. Capel, The Anisotropic Heisenberg spin chain and the nonlinear
Schrödinger equation. Phys. A 117, 76 (1983)
44. L.G. Potemina, Excitation of spin-waves by a varying magnetic-field in a biaxial ferromagnetic
with a moving domain-wall. Zh. Eksp. Teor. Fiz. 90, 964 (1986) [Sov. Phys. JETP 63, 562
(1986)]
45. Y.S. Kivshar, Perturbation theory based on the Riemann problem for the Landau-Lifshitz equa-
tion. Phys. D 40, 20 (1989)
46. M. Lakshmanan, Continuum spin system as an exactly solvable dynamical system. Phys. Lett.
61A, 53 (1977)
47. L.A. Takhtajan, Integration of the continuous Heisenberg spin chain through the inverse scat-
tering method. Phys. Lett. 64A, 235 (1977)
48. H.C. Fogedby, Solitons and magnons in the classical Heisenberg chain. J. Phys. A 13, 1467
(1980)
49. E. K. Sklyanin, Algebras I the defining equations (unpublished)
50. A.B. Borisov, Direct and inverse scattering problem for the Landau-Lifshitz equation. Dokl.
Akad. Nauk. SSSR 288, 1339 (1986) [Sov. Phys. Dokl. 31, 482 (1986)]
51. A.V. Mikhailov, Reduction in integrable systems. The reduction group. Pis’ma Zh. Eksp. Teor.
Fiz. 32, 187 (1980) [JETP Lett. 32, 174 (1980)]; Phys. Lett. 92A, 51 (1982)
52. Y.L. Rodin, The Riemann boundary problem on Riemann surfaces and the inverse scattering
problem for the Landau-Lifschitz equation. Phys. D 11, 90 (1984)
53. A.E. Borovik, N-soliton solutions of the nonlinear Landau-Lifshitz equation. Pis’ma Zh. Éksp.
Teor. Fiz. 28, 629 (1978) [JETP Lett. 28, 581 (1978)]
54. A.E. Borovik, S.I. Kulinich, Integration of the nonlinear dynamics of a uniaxial ferromagnet
by the method of the inverse scattering problem. Pis’ma Zh. Éksp. Teor. Fiz. 39, 320 (1984)
[JETP Lett. 39, 384 (1984)]
55. F.C. Pu, X. Zhou, B.Z. Li, Multi-soliton solutions for a classical ferromagnetic chain. Commun.
Theor. Phys. 2, 797 (1983)
444 12 Engineering Magnetic Solitons in Nonlinear Systems

56. Z.Y. Chen, N.N. Huang, Z.Z. Liu, An inverse scattering transform for the Landau-Lifschitz
equation for a spin chain with an easy axis. J. Phys.: Condens. Matter 7, 4533 (1995)
57. H. Yue, X.J. Chen, N.N. Huang, An inverse scattering transform for the Landau-Lifshitz equa-
tion for a spin chain with an easy plane. J. Phys. A 31, 2491 (1998)
58. N.N. Huang, Z.Y. Chen, Z.Z. Liu, Exact soliton solutions for a spin chain with an easy plane.
Phys. Rev. Lett. 75, 1395 (1995); N.N. Huang, Z.Y. Chen, Z.Z. Liu, The method of Darboux
transformation matrix for solving the Landau-Lifschitz equation for a spin chain with an easy
plane. J. Phys. A 28, 4063 (1995)
59. W.M. Liu, X.B. Wang, F.C. Pu, N.N. Huang, Nonlinear dynamics of the magnetization in an
anisotropic ferromagnet with a magnetic field. Phys. Rev. E 54, 4612 (1996); W.M. Liu, X.B.
Wang, F.C. Pu, and N.N. Huang, Solitons in a uniaxial anisotropic Heisenberg spin chain with
Gilbert damping in an external magnetic field. Phys. Rev. E 55, 1375 (1997)
60. H. Yue, N.N. Huang, An inverse scattering transform for the Landau-Lifshitz equation for a
spin chain with an easy plane. J. Phys. A 30, 331 (1997)
61. W.M. Liu, B.L. Zhou, Solitons in an order-parameter-preserving antiferromagnet. J. Phys.:
Condens. Matter 5, L149 (1993)
62. W.M. Liu, B.L. Zhou, Nonlinear excitations in the fcc antiferromagnet CeAs. Phys. Lett. A
184, 487 (1994)
63. J.K. Kjems, M. Steiner, Evidence for soliton modes in the one-dimensional ferromagnet
CsNiF3 . Phys. Rev. Lett. 41, 1137 (1978)
64. L.J. de Jongh, C.A.M. Milder, R.M. Cornelisse, A.J. van Duyn eveldt, J.P. Renard, Energy
absorption from an oscillating magnetic driving field by soliton motions in the quasi one-
dimensional ferromagnet [(CH3 )4 NNiCl3 (TMNC). Phys. Rev. Lett. 47, 1672 (1981)
65. E. Magyari, H. Thomas, R. Weber, Comment on “Landau-Lifshitz equation of ferromagnetism:
exact treatment of the Gilbert damping.” Phys. Rev. Lett. 56, 1756 (1986)
66. J.C. Slonczewski, Excitation of spin waves by an electric current. J. Magn. Magn. Mater. 195,
L261 (1999)
67. J.Z. Sun, Current-driven magnetic switching in manganite trilayer junctions. J. Magn. Magn.
Mater. 202, 157 (1999)
68. C. Heide, P.E. Zilberman, R.J. Elliott, Current-driven switching of magnetic layers. Phys. Rev.
B 63, 064424 (2001)
69. S. Zhang, P.M. Levy, A. Fert, Mechanisms of spin-polarized current-driven magnetization
switching. Phys. Rev. Lett. 88, 236601 (2002)
70. M. Tsoi, V. Tsoi, J. Bass, A.G.M. Jansen, P. Wyder, Current-driven resonances in magnetic
multilayers. Phys. Rev. Lett. 89, 246803 (2002)
71. T.Y. Chen, Y. Ji, C.L. Chien, M.D. Stiles, Current-driven switching in a single exchange-biased
ferromagnetic layer. Phys. Rev. Lett. 93, 026601 (2004)
72. Y. Tserkovnyak, A. Brataas, G.E.W. Bauer, Enhanced Gilbert damping in thin ferromagnetic
films. Phys. Rev. Lett. 88, 117601 (2002)
73. J.C. Slonczewski, Current-driven excitation of magnetic multilayers. J. Magn. Magn. Mater.
159, L1 (1996)
74. L. Berger, Emission of spin waves by a magnetic multilayer traversed by a current. Phys. Rev.
B 54, 9353 (1996)
75. Ya.. B. Bazaliy, B.A. Jones, S.C. Zhang, Modification of the Landau-Lifshitz equation in the
presence of a spin-polarized current in colossal- and giant-magnetoresistive materials. Phys.
Rev. B 57, R3213 (1998)
76. Z. Li, S. Zhang, Domain-wall dynamics and spin-wave excitations with spin-transfer Torques.
Phys. Rev. Lett. 92, 207203 (2004)
77. E.B. Myers, D.C. Ralph, J.A. Katine, R.N. Louie, R.A. Buhrman, Current-induced switching
of domains in magnetic multilayer devices. Science 285, 867 (1999)
78. J.Z. Sun, D.J. Monsma, M.J. Rooks, R.H. Koch, Batch-fabricated spin-injection magnetic
switches. Appl. Phys. Lett. 81, 2202 (2002)
79. J.E. Wegrowe, X. Hoffer, Ph. Guittienne, A. Fabian, L. Gravier, T. Wade, Current driven
resistance changes in low resistance x area magnetic tunnel junctions with ultra-thin (Math
processing error) barriers, and. J. Ph. Ansermet. J. Appl. Phys. 91, 6806 (2002)
References 445

80. M. Tsoi, A.G.M. Jansen, J. Bass, W.C. Chiang, M. Seck, V. Tsoi, P. Wyder, Excitation of a
magnetic multilayer by an electric current. Phys. Rev. Lett. 80, 4281 (1998)
81. Linear and Nonlinear Spin Waves in Magnetic Films and Super Lattices, ed. by M.G. Cottam
(World Scientific, Singapore, 1994), p. 421
82. Z.D. Li, J.Q. Liang, L. Li, W.M. Liu, Soliton solution of continuum magnetization equation in
a conducting ferromagnet with a spin-polarized current. Phys. Rev. E 69, 066611 (2004)
83. M. Lakshmanan, K. Nakamura, Landau-Lifshitz equation of ferromagnetism: exact treatment
of the Gilbert damping. Phys. Rev. Lett. 53, 2497 (1984)
84. P.-B. He, W.M. Liu, Nonlinear magnetization dynamics in a ferromagnetic nanowire with spin
current. Phys. Rev. B 72, 064410 (2005)
85. Y. Yamada, W.P. Van Drent, E.N. Abarra, T. Suzuki, High perpendicular anisotropy and
magneto-optical activities in ordered Co3 Pt alloy films. J. Appl. Phys. 83, 6527 (1998)
86. L. Belliard, J. Miltat, V. Kottler, V. Mathet, C. Chappert, T. Valet, Stripe domains morphology
versus layers thickness in CoPt multilayers. J. Appl. Phys. 81, 5315 (1997)
87. A.M. Kosevich, B.A. Ivanov, A.S. Kovalev, Magnetic solitons. Phys. Rep. 194, 117 (1990)
88. R. Hirota, Exact envelope-soliton solutions of a nonlinear wave equation. J. Math. Phys. 14,
805 (1973)
89. W.M. Liu, B. Wu, X. Zhou, D.K. Campbell, S.T. Chui, Q. Niu, interacting domain walls in an
easy-plane ferromagnet. Phys. Rev. B 65, 172416 (2002)
90. Dynamical Problems in Soliton Systems, ed. by S. Takeno (Springer, Berlin, 1984), p. 210
Chapter 13
Current Driven Dynamics
of Magnetization in Ferromagnet
with Spin Transfer Torque

Abstract This Chapter focuses on the dynamics of magnetization in ferromagnet


described by the generalized Landau–Lifshitz–Gilbert equation. By means of Hirota
bilinear method (HBM), exact analytical single- and two-soliton solutions of the
model equation are presented. Using these solutions, we investigate analytically the
nonlinear excitation of magnetization in ferromagnetic nanowire with spin polarized
current. The effect of adiabatic and non-adiabatic spin torque on dynamics magnetic
solitons (MSs) is studied too. We also, through the exact spin wave (SW) solutions of
the LLG equation, investigate the properties of the dynamic spin waves of magneti-
zation in ferromagnet. We show that the wave amplitude has the spatial and temporal
period on the spin wave background.

13.1 Introduction

The state of a magnetic crystal is described by a nonlinear magnetization vector


M which, as a function of both space and time can be governed by equations of
Landau–Lifshitz–Gilbert type, the LLG equations having the form [1–5]

∂M α ∂M
= −γ M × Heff + M× , (13.1)
∂t Ms ∂t

where, as in the previous chapter, M ≡ M(x, t) is the localized nonlinear magnetiza-


tion vector, γ , α, Ms , and Heff are respectively the gyromagnetic ratio, the damping
parameter, the saturation magnetization, and the effective magnetic field including
the exchange field, the external field, the demagnetization field, and the anisotropy
field [6–14].
Landau–Lifshitz–Gilbert equations (13.1) are nonlinear partial differential equa-
tion (NPDE) in which time t only plays the role of an independent variable, since its
coefficients do not contain explicitly the time. Their soliton solutions are therefore
called “ autonomous magnetic domain wall (DW) or soliton solutions” or “dynamic
soliton solutions” [15, 16]. During their propagation, these autonomous solitons may
completely preserve their velocities and localized form [17–24]. In magnetic mate-
rials, the velocities of the traveling autonomous solitons conventionally depends on
© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2022 447
E. Kengne and W. Liu, Nonlinear Waves,
https://doi.org/10.1007/978-981-19-6744-3_13
448 13 Current Driven Dynamics of Magnetization in Ferromagnet …

the strength of a constant magnetic field below a certain value (alias the Walker
break down value or critical value) above which the autonomous soliton develops a
time-dependent complex internal structure resulting in DW velocity oscillation [6,
7, 18, 19, 25–27]. In general, when the coefficients of a NPDE depend explicitly on
time t (such a NPDE can be obtained during the mathematical modeling of a physical
phenomenon when some external time-varying forces are applied on it), its soliton
solutions are called “ non-autonomous solitons” [16, 36, 37]. In the context of the
ferromagnetic nanowire, non-autonomous solitons can be realized by applying the
time-varying magnetic field/current to the local nonlinear magnetization [38–40].
Of course, applying time-varying magnetic field/current to the local magnetization
will result in the fast DW propagation with an optimal field or current pulse. Based
on the LLG equations (13.1) or its generalization/modification, we investigate in
this Chapter the NMDs of a magnetic domain in the presence of adiabatic and non-
adiabatic spin-transfer torque resulting in domain wall and non-autonomous solitons
in ferrimagnets. We also discuss some of the fundamental properties unique to mag-
netic multilayers.

13.2 Current Driven Dynamics of Domain Wall


in Ferrimagnets

In this Section, we consider a ferromagnetic metallic film in the presence of adiabatic


and non-adiabatic spin-transfer torque. The nonlinear magnetization dynamics in
this film is described by the generalized Landau–Lifshitz–Gilbert equations (model
equations). Assuming the magnetization magnitude to be a constant at temperatures
well below the Curie temperature, we reduce the model equations into a coupled first-
order PDEs with respect with a polar angle θ and azimuth angle φ, functions of spatial
variable x and time t. In the special case of an anisotropic ferromagnetic nanowire
driven only by spin-transfer torque, the critical current condition is reported both
theoretically and numerically. Exact DW solutions as well as exact non-autonomous
soliton solutions are presented and used to study the dynamics of nonlinear excitation
in the film. Analytical results are confirmed by direct numerical simulations.

13.2.1 Domain-Wall Resonance Induced by Spin-Polarized


Current

The physical model to be studied in this Section is the ferromagnetic metallic film in
the xy plane displayed in Fig. 13.1, with the easy axis z perpendicular to the film. We
assume, for this film, that going through the film along the x direction, the in-plane
current exerts spin-transfer torque on the local nonlinear magnetization. Taking the
13.2 Current Driven Dynamics of Domain Wall in Ferrimagnets 449

Fig. 13.1 Geometry of the film with stripe-domain structures. The easy axis is perpendicular to the
film plane and defined as the z direction. The current is applied along the film and perpendicular
to the domain wall, and  is the thickness of domain wall. Reprint from Ref. [5], Copyright 2022,
with permission from American Physical Society

adiabatic and non-adiabatic spin-transfer torque into consideration, the NMDs in this
film will be described by the following generalized LLG equations

∂M δH α ∂M
= γM × + M× + Ta + Tn , (13.2)
∂t δM Ms ∂t

H being the energy density. The adiabatic spin torque, Ta , caused by the non-
equilibrium conduction electrons is given as
   
bJ ∂M τex 
Ta = − M 2
M × M × , with ξ = , τex = ,
1 + ξ2 s ∂x τsf Jex

where Jex , je , and τsf are respectively the exchange coupling strength, the electric
current density, and the spin-flip relaxation time. The adiabatic and non-adiabatic
spin torque, Tn , describes the effect of the non-equilibrium conductive electrons and
is given by  
ξ bJ ∂M
Tn = − Ms M× .
1+ξ 2 ∂x

Here, M is, of course, the nonlinear magnetization vector.


Equation (13.2) under the condition M2 = Ms2 = const below the Curie temper-
ature can be reduced into the following two coupled differential equations for θ and
φ:
 
γ
∂θ
∂t + α ∂φ δH
∂t sin θ = − Ms sin θ δφ +
bJ
1+ξ 2
∂θ
∂x + ξ ∂φ
∂x sin θ ,
  (13.3)
∂φ γ δH ∂φ
∂t sin θ − α ∂θ
∂t = Ms δθ + bJ
1+ξ 2 ∂x sin θ − ξ ∂θ
∂x ,
450 13 Current Driven Dynamics of Magnetization in Ferromagnet …

where θ and φ are the polar angle and azimuth angle, defined relative
to the z axis, respectively. Next, we seek the dynamic wall structure as
[6, 41]  
x − L(t)
θ(x, t) = 2 arctan exp , φ = φ(t), (13.4)
(t)

where L = L(t) and  = (t) represent the normal displacement coor-


dinate of the wall and the wall width, respectively. Next, we assume
that the distortion of wall width is small during the wall motion, that is,
|d /dt| / |dL/dt|  1; this means that (t) varies less than L(t). If
we insert Eq. (13.4) into Eq. (13.3), we obtain the following nonlinear
differential system for L and φ:
dL dφ bJ
= 2π γ Ms  sin [2φ] + α − , (13.5)
dt dt 1 + ξ2
dφ 8π Ms α dL bJ ξ 1
= −γ L− − . (13.6)
dt lW  dt 1 + ξ2 
In system (13.5)–(13.6), lW stands for the length of one domain. Due to
the reigning momentum transfer for the non-adiabatic electrons [12],
the non-adiabatic torques will generate global pressure on the walls.
Also, the applied field and magneto-static restoring potential will gen-
erate an instantaneous pressure on the DWs. Assuming that the initial
current and perpendicular magnetic field are applied on the walls and
that the wall lies in the pinning center, the initial velocity υ0 of wall
can be computed using system (13.5)–(13.6):

(1 + αξ ) bJ
υ0 = −   .
1 + ξ 2 1 + α2

Also, we assume that the azimuth angle φ starts precessing at the


Larmor frequency
(α − ξ ) bJ
fφ =  .
 1 + ξ2

It is evident that the velocity decreases when φ increases.


Next, we focus on the case of small deviations of φ from equilibrium,
and assume that the wall velocity varies linearly with φ. Linearizing
Eq. (13.5) with respect with φ and using Eq. (13.6) to eliminate φ from
13.2 Current Driven Dynamics of Domain Wall in Ferrimagnets 451

the resulting equation yield


 
d 2L dL 2Ms ξ 1 + ξ α ∂bJ
m 2 + b + kL = −  bJ + ,
dt dt γ 1 + ξ2  4π γ Ms ∂t
(13.7)
where
 

2Ms k 1 + α2 16π Ms2 A


b=α + ,m = ,k = , and  = .
γ 4π γ Ms 2π γ 
2 lW K
In the case of a direct current, non-adiabatic torque produces a global
pressure on the wall, while the adiabatic torque either transfers mag-
netic momentum to the wall or only twists the magnetization of wall
[12, 14, 35]. In the ac (alternating current) case, it follows from Eq.
(13.7) that both the non-adiabatic spin torque and adiabatic one con-
tribute to the global motion of wall.
When a direct current is applied, Eq. (13.7) describes damped oscil-
lation of the wall. Deemed as a particle in the dc case, only non-
adiabatic spin torque produces a pressure on the domain wall. This
means that in the case of a direct current, only non-adiabatic spin torque
affects wall motion. The direct current changes the equilibrium loca-
tion of the oscillation with the evaluated as follows, when the driving
field is absent:
je ξ PμB lW
L0 =  .
8π eγ Ms2 1 + ξ 2

If L0 ≥ lW , the wall will depinned, leading to the following current


cri for depinning
density jdc

8π 1 + ξ 2 eγ Ms2
jdc =
cri
. (13.8)
ξ PμB
Let us now consider the case of alternating current. Taking je =
jac cos ωt with oscillating frequency ω in the ac case, equation (13.7)
will describe forced damped oscillations. Assuming that 4mk − b2 >
0, we arrive to
452 13 Current Driven Dynamics of Magnetization in Ferromagnet …

  √
b 4mk − b2
L(t) = L0 exp − t cos t
2m 2m

2PμB /γ e 1 + ξ 2 ξ 2 /2 + (1 + αξ/4π γ Ms )2 ω2 jac
−  2
k − mω2 + b2 ω2
 
× cos ωt − δ − δ  , (13.9)

where
 
bω (1 + αξ ) ω
δ = arctan 2 , δ  = arctan  ,
k − mω2 1 + ξ 2 4π γ ξ Ms

and L0 is a real constant which can be determined from the initial


conditions. For large b/2m (this is the case for alternating current), the
first term of L(t) will rapidly decrease to zero. From Eq. (13.9), we
evaluate the resonance frequency ω and obtain


 42
ω = 4π γ Ms  2
+ ξ 2 + ξ 4 − ξ 2,
lW

and smaller current leads to a larger displacement. Solving the equa-


tion Lmax = lW in jac leads to the following critical alternating current
cri that can depin DW in the presence of resonance:
density jac
 ⎡ 

    ⎤
 ⎢    l 2
l
 ⎢ 2 π π ξ 4 − α 2 ξ 4 W + 4ξ 2 W + 4 ⎥ ⎥ ψri
ξ W l  2 
cri
jac =  ⎢ ⎢
⎥j .
⎥ dc
2  ⎣ ⎦
 2
2 lW
+ α − 2π ξ −4

(13.10)

Evaluating the resonant frequency ω and the critical alternating cur-


cri for the material parameters of Co Pt with l = 1 µm,
rent density jac 3 W
we find ω = 12.1 GHz and jac cri = 0.18j cri . In the absence of currents,
dc
to counter the restoring force, the driving field is not less than 8π Ms .
For example, for the material parameters of Co3 Pt, the value of the
13.2 Current Driven Dynamics of Domain Wall in Ferrimagnets 453

driving field is found to be 2.1 T. If we combine the magnetic field and


spin current, we can also evaluate the numerical value of the critical
current density. When we apply a magnetic field perpendicular to the
film and add the term 2Ms H⊥ to the right hand side of Eq. (13.7), we
arrive to critical direct current density due to driving field and current
cri :
jdc
cri
jdc = [1 − H⊥ /(8π Ms )]jdc
cri
.

13.2.2 Screw-Pitch Effect and Velocity Oscillation of Domain


Wall

Here, we study analytically the screw-pitch and velocity oscillation


of DW. For anisotropic ferromagnetic nanowire driven only by spin-
transfer torque, we report analytically critical current condition and
study the behavior of the ferromagnetic nanowire relatively to the
critical current. When the spin polarized current is below the criti-
cal current, the ferromagnetic nanowire admits only the final static
domain wall solution below the critical current; this implies that the
spin-polarized current cannot drive DW move continuously. Above the
critical value of the spin polarized current, the dynamics of DW exhibits
the novel screw-pitch effect: the DW velocity and width with period-
ically oscillate in time. Detail theoretical and computational studies
show that the above novel phenomenon arises from the natural con-
junction action of Gilbert-damping and spin-transfer torque . By means
of the Runge–Kutta–Munthe–Kaas (RKMK) method [42, 43], our the-
oretical prediction may be numerically confirmed.
Now, let us consider an infinitely long uniaxial anisotropic ferro-
magnetic nanowire. We assume that in such a uniaxial anisotropic fer-
romagnetic nanowire, the electronic current flows along the long length
of the wire defined as the x-direction; this x-direction is also assumed to
be the easy axis of anisotropy ferromagnet. Next, we assume the mag-
netization to be nonuniform only in the x-direction (direction of the
current). Taking into account the slow spatial variation of the magneti-
zation, we take the adiabatic limit so that the dynamics of the localized
magnetization can be described by the modified LLG equation with
spin-transfer torque, that is, the following first-order partial differential
equation in nonlinear magnetization vector M [11–13]:
454 13 Current Driven Dynamics of Magnetization in Ferromagnet …

∂M α ∂M
= −γ M × Heff + M× + τb . (13.11)
∂t Ms ∂t
In Eq. (13.11), Heff stands for the effective magnetic field which, in
the case of an uniaxial ferromagnetic nanowire can take the form

2A ∂ 2 M Hx Mx
Heff = + ex − 4π Mz ez ,
Ms2 ∂x2 Ms

where A, Hx , and ei (i = x, y, z) are the exchange constant, the


anisotropy field, and the unit vector, respectively. Denoting by m the
normalized magnetization, that is, m = M/Ms , Eq. (13.11) will take
the following simplified dimensionless form:
∂m  ∂m ∂m
α1 = −m × heff − αm× m × heff + αb1 m× + b1 ,
∂t ∂x ∂x
 (13.12)
where α1 = 1 + α and b1 = bJ t0 /l0 , bJ are given in advance, and
2

heff is the dimensionless effective field. Throughout this subsection, the


time t and space coordinate x have been rescaled by the characteristic
time 
1 A
t0 = and l0 = ,
16π γ Ms 8π Ms2

respectively. The dimensionless effective field heff under the above


assumptions on the temporal and spatial variables t and x becomes

heff = ∂ 2 m/∂x2 + Cx mx ex − Cz mz ez ,

with Cz = 0.25 and Cx = Hx / (16π Ms ).


In the following, the exact DW solutions for Eq. (13.12) are pre-
sented, and are used to study analytically the dynamics of magnetiza-
tion driven by spin-transfer torque. For this aim, we use the ansatz
sin [φD ] cos [φD ]
mx = tanh [D ] , my = , mz = , (13.13)
cosh [D ] cosh [D ]
where φD = k1 x − ω1 t, k1 and ω1 are some real parameters to be deter-
mined. Asking that Eq. (13.13) satisfies Eq. (13.12) leads to
13.2 Current Driven Dynamics of Domain Wall in Ferrimagnets 455

k12 = Cx + Cz cos2 [φD ] , (13.14)



−ω1 1 + α 2 = b1 k1 + Cz sin [φD ] cos [φD ] , (13.15)

αb1 k1 cos [φD ] = α Cx − k12 sin [φD ] , (13.16)
αb1 k1 sin [φD ] = −αCz sin [φD ] cos [φD ] .
2
(13.17)

Using Eqs. (13.14)–(13.17), we can derive three cases of DW solu-


tions of Eq. (13.12).
(i) Case of the absence of damping (α = 0): Setting α = 0 in system
(13.14)–(13.17) yields the exact solution
 Cz
k1 = ± Cx + Cz cos2 [φD ], ω1 = −b1 k1 − sin [2φD ] , (13.18)
2
where φD is a free angle. It follows from this solution (13.18) that in
the absence of damping, the spin-transfer torque contributes a dimen-
sionless velocity −b1 only. Computing the velocity of DW, we find
Cz
ν=− sin [2φD ] − b1 .
2k1
It is clearly seen that the DW velocity is formed of two parts, the first
part depending on angle φD , while the second one coming from the
spin-transfer torque; therefore, the velocity v of the domain wall can
be affected by adjusting the angle φD and the spin-transfer torque.
(ii) Case of the absence of spin torque (b1 = 0): When b1 = 0, sys-
tem (13.14)–(13.17) will lead to
π 
ω1 = 0, φD = ± , and k1 = ± Cx .
2
This is just a static domain wall solution. By means of the Runge–
Kutta–Munthe–Kaas method method [42, 43], the direct simulations
for Eq. (13.12) with various initial conditions, we show that the damp-
ing drives the change of φD , which in turn affects the velocity and width
of the DW defined by 1/ |k1 | . conditions φD = ± π2 and ω1 = 0 means
that the
√ DW stops completely, and its width achieves the maximum
value Cx . This confirms Walker’s analysis [16] leading to the con-
clusion that the damping prevents DW from moving without the spin-
transfer torque or external magnetic field. It has been however shown
456 13 Current Driven Dynamics of Magnetization in Ferromagnet …

that the presence of damping is prerequisite for the novel screw-pitch


property of DW driven by spin-transfer torque.
(iii) Case of the presence of damping and spin-transfer torque (αb1  =
0). When αb1  = 0, we solve system (13.14)–(13.17) and obtain
1   2b1 k1
k1 = ± B1 − B2 , ω1 = 0, sin [2φD ] = − , (13.19)
2 Cz
where  2
B1 = 2Cx + Cz − b21 , B2 = Cz − b21 − 4Cx b21 ,

and the signs “ −” and “ +” denote anti-kink and kink solution, respec-
tively. Inserting Eq. (13.19) into Eq. (13.13) with these signs “ −” and
“ +” leads to the static head-to-head and tail-to-tail Neel domain wall,
respectively.
Analyzing Eq. (13.19), the critical spin polarized current condition
is found to be    l0
bJ ≤ Cx + Cz − Cx ; (13.20)
t0
condition (13.20) means that the critical current condition is deter-
mined by the anisotropic parameter Cx , the demagnetization parameter
Cz , and the character velocity l0 /t0 . Under the condition
  2
b21 ≤ Cx + Cz − Cx

(that is, below the critical current), the width of the DW falls into the
range of
1 1 1
√ ≤ ≤ .
Cx + Cx Cz |k1 | Cx
Solving the last equation in Eq. (13.19) for φD , we obtain four solu-
tions:
 
π 1 2b1 k1
φD = ± + arcsin for k1 > 0, and
2 2 Cz
 
π 1 2b1 |k1 |
φD = ± − arcsin for k1 < 0.
2 2 Cz
13.2 Current Driven Dynamics of Domain Wall in Ferrimagnets 457

The above result shows that below the critical spin polarized current,
condition
∂m
m × heff = b1 (13.21)
∂x
guarantees the final equilibrium domain wall solution. This condition
(13.21) clearly shows that the spin-transfer torque has the following
two interesting effects: (i) the term b1 ∂m/∂x in Eq. (13.12) plays the
anti-precession role counteracting the precession driven by the effec-
tive field heff , while term αb1 m × (∂m/∂x) in Eq. (13.13) has the anti-
damping effect counteracting the damping term −αm × (∂m/∂x); (ii)
the spin-polarized current below its critical value cannot drive DW
move continuously without the applied external magnetic field.
As we will see in the next Section, the dynamics of DW exhibits
two novel properties when the spin-polarized current exceeds the crit-
ical value. In the situation when the spin-polarized current exceeds the
critical value, the precession term −m×heff cannot be counteracted
by spin-transfer torque, and equation (13.12) does not admit static
DW solutions. Since the magnitude of the magnetization is constant,
we have m · (∂m/∂x) = 0; this shows that m⊥∂m/∂x, that is, either
∂m/∂x = 0 or the direction of m is always perpendicular to the direc-
tion of ∂m/∂x. As we well know, a magnetic domain wall separates
two opposite domains by minimizing the energy, while the direction of
magnetic moments gradually changes, so that ∂m/∂x  = 0. It is clear
that the normalized magnetization out of region of DW will site at the
easy axis, that is, mx = 1 (mx = −1), in which we have ∂m/∂x = 0.
Analyzing with details Eq. (13.12) under the above consideration,
we find that the motion of magnetic moment in the DW center will
not stop, except it falls into the easy axis (x-direction), that is, out
of the range of DW. Going back to Eq. (13.12), we can see that the
above found novel screw-pitch effect on the domail velocity and width
occurs even at the conjunct action of the damping and spin-transfer
torque. This theoretical prediction is confirmed by performing direct
simulations by means of the RKMK method on Eq. (13.12) with an
arbitrary initial condition [42, 43]; in these numerical simulations, we
consider only current that exceeds the critical value. Figures 13.2a–c
and Fig. 13.2d show the time evolution of the normalized magnetiza-
tion m and the displacement of domain wall center, respectively. All
the results in Fig. 13.2 are in entire agreement with our above theo-
458 13 Current Driven Dynamics of Magnetization in Ferromagnet …

Fig. 13.2 Dynamics of DW above the critical current. a–c Evolution of the normalized magneti-
zation m. d Displacement of DW driven only by spin-transfer torque. The parameters are α = 0.2,
Cx = 0.05, Cz = 0.25, bJ = 0.6, and the initial angle φ = 0.01π . Reprint from Ref. [5], Copyright
2022, with permission from American Physical Society

retical analysis. Figure 13.3 displays the evolution of cos [φD ] and the
evolution of the velocity and width of the DW velocity. As we can see
from Fig. 13.3, the periodic change of cos [φD ] leads to the temporal
periodically oscillation of DW width and velocity. From the third term
of Eqs. (13.12) and (13.18), cos [φD ] undergoes uneven change as one
can clearly see in Fig. 13.3a. It is clearly seen from Fig. 13.2d that the
DW displacement increases rapidly at first period and then slowly.

13.2.3 Non-autonomous Helical Motion of Magnetization in


Ferromagnetic Nanowire

Here, we focus on an infinite biaxial anisotropic ferromagnetic nanowire.


The easy axis is assumed to be along the wire defined as the z-direction,
while the hard axis is supposed to be perpendicular to the wire axis,
as depicted in Fig. 13.4. We denote by M(z, t) the localized nonlinear
magnetization and by Ms the saturation magnetization . We assume M
13.2 Current Driven Dynamics of Domain Wall in Ferrimagnets 459

Fig. 13.3 a Evolution of cos [φ] and periodic oscillation of DW velocity. b Temporal periodic
oscillation of DW width. The numerical values of different parameters are the same as those in
Fig. 11.2. Reprint from Ref. [5], Copyright 2022, with permission from American Physical Society

to be nonuniform only in the z-direction. Next, we adopt Slonczewski


torque [8] TSTT which generally takes the form of

TSTT = γ m × (m × S) ,

where γ and
M(z, t)
m(z, t) = ,
Ms
are the gyromagnetic ratio and the normalized magnetization, respec-
tively, and

S = SL ez + ST (z, t)m × ez , with SL,T = jL,T P ,
2d |e|
the subscript T representing the temporal and spatial dependent trans-
verse density, and L representing the constant longitudinal density
along the z-axis. The parameters d and 2 are the thickness of the free
magnetic layer and e the magnitude of electron charge, respectively.
Under all these considerations, the contribution of the torque effect is
separated from the reference layer rotation.
Considering a time-depending spin-polarized current, the dynamics
of non-autonomous magnetic DW can be governed by the following
modified LLG equation with spin-transfer torque [4, 8, 10]
460 13 Current Driven Dynamics of Magnetization in Ferromagnet …

Fig. 13.4 Schematic representation of the ferromagnetic nanowire and the coordinate system.
The easy axis is assumed to be perpendicular to the nanowire and defined as the z-direction. SL
and ST stand for the constant longitudinal density along the z axis and the temporal and spatial
dependent transverse density, respectively. Reprint from Ref. [5], Copyright 2022, with permission
from American Physical Society

∂M α ∂M
= −γ M × Heff + M× + TSTT . (13.22)
∂t Ms ∂t
Here, Heff is the effective field having the form

Kx M x Kz A ∂ 2M
Heff = e x − M e
z z − + Hext . (13.23)
μ0 Ms2 μ0 Ms2 μ0 Ms2 ∂z 2

In Eq. (13.23), Kx and Kz denote the energetic anisotropy constant along


the hard and easy axis, respectively. A, Kx , Kz and ei (i = x, y, z) are
the exchange interaction coefficient, the energetic anisotropy constant
along the hard axis, the energetic anisotropy constant along the easy
axis and the unit vector, respectively. Considering in this subsection
an external magnetic field Hext with temporal and spatial dependence,
we can reduce Eq. (13.22) to the following dimensionless form
1 ∂m   
= −m × heff − αs + m× αheff + s , (13.24)
 ∂t
∂ 2m Kx,z
heff = − kx mx ex + kz mz ez + hext with kx,z ≡ ,
(13.25)
∂z 2 μ0 Ms
Hext
hext = ≡ −hL ez − hT m × ez , (13.26)
Ms
13.2 Current Driven Dynamics of Domain Wall in Ferrimagnets 461

where heff and hext are the dimensionless effective


 field, s ≡ S/Ms is
the dimensionless spin torque, and  = 1/ 1 + α . The spatial coor-
2

dinate x and time t in Eq. (13.24) have been rescaled by the character-

istic length l0 = A/ μ0 Ms2 and time t0 = 1/(γ Ms ), respectively.
Writing m in the system of the spherical coordinates, that is,

m = (sin [θ] cos [φ] , sin [θ] sin [φ] , cos θ)T ,

equation (13.24) in the coordinates θ (z, t) and φ (z, t) becomes

∂φ ∂θ 1 ∂ 2θ
sin [θ] − α = kx cos [φ] + kz sin [2θ] − 2 − (hL − ST ) sin [θ] ,
2
∂t ∂t 2 ∂z
(13.27)
∂φ ∂θ 1 ∂ 2θ
α sin [θ] + = (kx sin [2φ] + hT + SL ) sin [θ] − 2 .
∂t ∂t 2 ∂z
(13.28)

To integrate Eqs. (13.27) and (13.28), we assume that


z − L(t)
ln [tan [θ/2]] = ,
W0
where L(t) is the DW center, the constant width W0 , and the angle
φ(t) of DW are to be determined. Imposing to the width of the DW to
remain constant, we obtain that
1
W0 = √
kx + kz
and the spatiotemporal transverse spin current ST should have the form

ST (z, t) = −kx sin2 [φ] tanh [θ] .

We then obtain that, the DW profile satisfies the equation

∂ 2θ
(kx + kz ) sin [2θ] − 2 = 0,
∂z 2
with the BC of θ = 0 and π at long distance. Under all these consid-
erations, we find from Eqs. (13.27) and (13.28) that the DW velocity
462 13 Current Driven Dynamics of Magnetization in Ferromagnet …

and DW’s rotational motion obey the following differential system for
L, φ, and
 
∂L W0 ∂φ
V≡ =− hL + , (13.29)
∂t α ∂t
∂φ
= C + Cm sin [2φ] , (13.30)
∂t
where Cm and C are two parameters defined as
1
Cm = αkx and C = α(SL + hT ) − hL . (13.31)
2
Equations (13.29) and (13.30) constitute the novel motion of non-
autonomous DW. We can clearly see from Eq. (13.29) that the velocity
of the DW consists of two parts: the coherent DW rotation velocity
∂φ/∂t and the longitudinal external field hL along the easy axis. Also,
we can see from these two equations that the damping parameter α also
affects the DW velocity, while the DW width W0 is determined by the
energetic anisotropy constants. As it is clearly seen from Eq. (13.30),
the velocity of the DW has two characteristics, neamely, the constant
velocity C and the temporal periodic velocity with amplitude Cm ;
the constant velocity arises from the external driving force, while the
periodic temporal one is induced by the internal transverse anisotropy
parameter kx . From Eqs. (13.29) and (13.30), we derive the following
DW displacement:
⎡ ⎤
t
W0 ⎣
L(t) = − hL t +  (C + Cm sin [2φ]) dt ⎦ . (13.32)
α
0

The rotating and moving magnetization in DW described by Eqs.


(13.29) and (13.30) will be referred to as the non-autonomous heli-
cal motion or screw-pitch effect [44].
It is obvious that by tuning SL and hext , the coherent DW rotation
will exhibit different feature, which in turn affects the velocity and dis-
placement of the DW. Depending on C and Cm given by Eq. (13.31), we
can distinguish three cases of solutions of Eq. (13.30). For simplicity,
we assume in the following discussion that hT ,L and SL are constant.
13.2 Current Driven Dynamics of Domain Wall in Ferrimagnets 463

(i) Case of small applied current and external field (C < Cm ). Inte-
grating Eq. (13.30) yields

CB2 − 2Cm B + C  √ 
sin [2φ] = − , with B = exp −2 κ1 κ2 t ,
Cm B − 2CB + Cm
2
(13.33)
where κ1 = Cm − C and κ2 = Cm + C. At the limit t → ∞, B → 0
and sin [2φ] → −C/Cm ; this means that the DW stops the rotation
and the cant angle attains its extremum. This fact is confirmed in Eq.
(13.30). Computing the DW velocity in the limit t → ∞, we find V =
−W0 hL /α; this means that in the limit t → ∞, the DW velocity can be
controlled by adjusting the longitudinal external field hL . Integrating
Eq. (13.32) in this case yields the following DW displacement L(t):
 
  
W0 κ1 ζ (t)
L(t) = − hL t + arctan − arctan √ ,
α κ2 κ1 κ2
with ζ (t) = Cm B − C.

(ii) The parameter kx of the internal transverse anisotropy is balanced


by the applied current and external field (C = Cm ). When C = Cm , we
find the following special solution of Eq. (13.30)

1 − (2Ct)2
sin [2φ] = . (13.34)
1 + (2Ct)2
The DW solution (13.34) implies that the DW stops the rotation in the
limit t → ∞, and at the same time, the cant angle is sin [2φ] → −1.
Computing the DW velocity in the limit t → ∞, we obtain the same
result as in the above case (i). Meanwhile, the DW displacement is
found to be
W0
L(t) = − [hL t + arctan [2Ct]] .
α
It follows from the results obtained for C ≤ Cm that a critical value
for current and/or field exists. When the external filed is neglected,
this critical condition reads SLc = kx /2. When the spin-polarized cur-
rent is below its critical value, the DW finally stops the coherent rota-
tion, resulting in the static DW solution with an arbitrary initial DW
state. The DW velocity in both cases is found to be controlled finally
464 13 Current Driven Dynamics of Magnetization in Ferromagnet …

by the longitudinal external filed hL for which the damping factor α


makes it enhance. For the cases when either C < Cm or C = Cm , we
have displayed in Fig. 13.5 the graphic illustration of the DW rotation
·
velocity ∂φ/∂t = φ, the DW velocity V , and the displacement L. One
can easily see from Fig. 13.5a that the DW rotation velocity increases
with increasing C initially. Finally, the DW stops rotation and the DW
velocity and displacement are proportional to the longitudinal filed
value hL . This graphical feature of the DW rotation and DW velocity
and displacement confirms the theoretical findings as we can see from
Eqs. (13.29)–(13.32).
Let us consider the case C > Cm (beyond the critical value). When
C > Cm , the DW always rotates with the easy axis with velocity
∂φ/∂t and possesses the non-autonomous helical motion when both
the applied current and external field are increased. If we integrate Eq.
(13.30), we obtain the solution

κ2 + κ1 tan2 [ωt] √
sin [2φ] = − , with ω =  −κ1 κ2 . (13.35)
κ2 − κ1 tan2 [ωt]
The extrema of sin [2φ] are reached when sin [2φ] = ±1; solving this
equation yields t = tn = nπ/ (2ω) (n = 0, 1, ...). Inserting Eq. (13.30)
into Eq. (13.35), we find that the DW rotation velocity ∂φ/∂t consists
of two parts, the temporal periodic velocity with the amplitude Cm
and frequency ω, controlled by the longitudinal spin-polarized cur-

·
Fig. 13.5 Graphic illustration of a domain wall rotation velocity φ = ∂φ/∂t, b domain wall velocity
V , and c displacement L in the cases C ≤ Cm . Different parameters used in generating different
plots are: α = 0.1, kx = 0.86, and kz = 5 × 10−4 . Reprint from Ref. [5], Copyright 2022, with
permission from American Physical Society
13.2 Current Driven Dynamics of Domain Wall in Ferrimagnets 465

rent SL , and the constant velocity C, related to the internal trans-
verse anisotropy parameter kx . This interesting phenomenon occurs
independently on the presence or absence of external filed. When the
external field is neglected, only the longitudinal spin-polarized current
SL obtained from Eq. (13.29) controls the DW velocity and rotation.
When increasing parameter C, the DW rotation becomes more rapid;
this can be realized by adjusting the external driving force.
Using the expressions of C and Cm from Eq. (13.31), the DW velocity
in Eq. (13.29) reads
 
1
V = −W0  αhL + SL + hT + kx sin [2φ] . (13.36)
2
Equation (13.36) shows the contribution of the longitudinal spin-
polarized current SL , the longitudinal external field hL , and the transver-
sal external field hT to the DW velocity: SL and hT contribute to the fast
DW velocity, while hL is weakened by the damping factor α. Compared
the above results obtained in the case C > Cm with those obtained when
C ≤ Cm , we find that the DW velocity becomes slow with the same
the longitudinal external field hL . For the rotating motion of magne-
tization obtained in the case C > Cm , the DW displacement L(t) in
Eq. (13.32) consists two parts, a linear and a nonlinear parts. We can
also see from Eq. (13.32) that for each time period T0 = π/ω, the DW
moves through the same displacement.
The above properties of the non-autonomous DW motion obtained
beyond the critical value (C > Cm ) are illustrated in Figs. 13.6 and
13.7. As we can see from Fig. 13.6 obtained when C is slightly above
the critical value that the DW rotates periodically and exhibits a nonuni-
form velocity in each period of time. It firstly rotates slowly, and finally,
rotates rapidly in the middle of each period. The DW displacement
and velocity in turn take the same vision. Figure 13.7 generated with
C Cm reveals that the DW possesses the sinusoidal oscillation nearly
for DW rotation and velocity. At the same time, the DW displacement
takes the linear response to time. This last feature is theoretically con-
firmed by the following special form of L(t) obtained from Eq. (13.32)
with κ1∼ κ2 :
tW0 (ω + hL )
L(t) = − .
α
466 13 Current Driven Dynamics of Magnetization in Ferromagnet …

Fig. 13.6 Non-autonomous domain wall motion in the case of C > Cm , namely, C is slightly
·
greater than the critical value. a Domain wall rotation velocity φ = ∂φ/∂t, b domain wall velocity
V , and c domain wall displacement L. Parameters used for generating different plots are the same as
those in Fig. 13.5. Reprint from Ref. [5], Copyright 2022, with permission from American Physical
Society

·
Fig. 13.7 Non-autonomous motion of domain wall rotation velocity φ = ∂φ/∂t, domain wall
velocity V , and displacement in the case of C Cm . Different parameters used in generating these
plots are the same as those in Fig. 13.5. Reprint from Ref. [5], Copyright 2022, with permission
from American Physical Society

Let us take a look on the motion of the DW in the limit t → ∞ (this


may help us in a deeply analysis of the relationship between C and Cm ).
It is evident that in the limit t → ∞, the critical point for both the DW
rotation velocity and the DW velocity occurs by C = Cm . As we can
see from Fig. 13.8a, the DW will stop when C ≤ Cm and, as we can see
from Eq. (13.30), the rotation velocity is zero in the limit t → ∞, that
13.2 Current Driven Dynamics of Domain Wall in Ferrimagnets 467

·
Fig. 13.8 Graphic illustration of a DW rotation velocity φ = ∂φ/∂t and b DW velocity V versus
C. The parameter SL ranges form 0 to 2. Other parameters used for generating these plots are:
α = 0.1, k1 = 0.86, k2 = 5 × 10−4 , hL = 0.01, and hT = 0.1. Reprint from Ref. [5], Copyright
2022, with permission from American Physical Society

is, ∂φ/∂t|t→∞ = 0. In this limit t → ∞, it follows from Eq. (13.29)


that the DW velocity takes the form V (t)|t→∞ = −W0 hL /α and can
be well controlled by adjusting hL . Since the DW motion is periodic in
time above the critical value (C > Cm ), the average rotation velocity
in the limit t → ∞ can be defined as follows:

∂φ 
≡ ∂φ/∂t and V (t)|t→∞ ≡ V
∂t t→∞

for each period of time π/ω.


According to the analysis done for C > Cm , the above found non-
autonomous DW solution under the action of different external field
and spin-polarized current with the spatial dependence will bring forth
the complicated interesting motion. For a better clarifying of this
scheme, the expressions of hext and s can be projected into the follow-
ing special normal coordinate system, (m, ∂ m/∂z, m × ∂ m/∂z), where
∂ m/∂z ≡ (∂m/∂z)/ |∂m/∂z|, as shown in Fig. 13.9a. Figure 13.9b
shows a plan view in which we can clearly see that both the spin
polarization of current and the external magnetic field fall into the
plane of (∂ m/∂z, m × ∂ m/∂z). It follows from Eq. (13.24) that the
spin polarization of current and only the applied field that fall into the
plane of (∂ m/∂z, m × ∂ m/∂z) can contribute to the dynamics of local
magnetization. It is also seen from Eq. (13.24) that the longitudinal
part SL of current results in the adiabatic spin-transfer torque, while its
468 13 Current Driven Dynamics of Magnetization in Ferromagnet …


Fig. 13.9 Schematic of the special normal coordinate system m, ∂∂zm , m × ∂∂zm , where ∂∂zm ≡
 ∂m  ∂m −1
  . The polarization of current S and the external magnetic field Hext fall into the plane
∂z
 ∂z
of ∂∂zm , m × ∂∂zm . a Three-dimensional diagram and b a plan view. Reprint from Ref. [5], Copyright
2022, with permission from American Physical Society

Fig. 13.10 Geometry of a very long ferromagnetic nanowire. The magnetic anisotropy field HK
and the applied magnetic field Hext are along the vertical direction of film defined as the the z-axis.
The applied current je flows along the nanowire, considered as the x- direction. Reprint from Ref.
[29], Copyright 2022, with permission from American Physical Society

spatial dependence part ST (z, t) plays the same role of non-adiabatic


spin-transfer torque.

13.3 Current Driven Dynamics of Soliton

In this Section, we consider a very long ferromagnetic nanowire (which


can be viewed as infinite in length) with uniform cross section, as
shown in Fig. 13.10. The electronic current flows along the long length
of the wire which is defined as the x-direction. We consider the z-axis
to be taken as the directions of uniaxial anisotropy field and exter-
nal field.The magnetization is assumed to be nonuniform only in the
direction of current. Under these considerations, the nonlinear mag-
13.3 Current Driven Dynamics of Soliton 469

netization dynamics will be governed by the following modified LLG


equation with the spin torque produced by the injected in the nanowire
current after its polarization (model equation)
∂M α ∂M
= −γ M × Heff + M× + τb . (13.37)
∂t Ms ∂t
In Eq. (13.37), the effective magnetic field Heff can be written as
  
2A ∂ 2 M HK
Heff = 2 2 + − 4π Mz + Hext ez ,
Ms ∂x Ms

where A, Hext , HK , and ez are the exchange constant, the applied field,
the anisotropy field, and the unit vector along the z-direction.
In the following, we discuss both the LMDs and the NMDs in a
uniaxial ferromagnetic nanowire injected with current. By means of a
stereographic projection [45], we transform the model equation into
a nonlinear equation of complex function; using this last equation,
we obtain high-Q model and the low-Q model of dark/bright soliton
solutions with the help of which we investigate the effect of spin-
polarized current on the MSs of the magnetization in the ferromagnetic
metal nanowire; here,
HK
Q= (13.38)
4π Ms
is the quality factor.

13.3.1 Dark Solitons

In the case when Q > 1 which corresponds to the situation when the
uniaxial anisotropy field is larger than the demagnetization field, such
as the case for some metallic magnetic films [46] and multilayers [47],
dark MSs for mz component of the nonlinear magnetization vector
m = M/MS can be obtained. For this aim, the time coordinate and
space coordinate are rescaled by characteristic time and length as
  
t0 = 1/ γ (HK − 4π Ms ) and l0 = 2A/ [(HK − 4π Ms ) Ms ],
470 13 Current Driven Dynamics of Magnetization in Ferromagnet …

and set M = Ms m. The LLG equation (13.37) can then be simplified


and take the following form:
 
∂m ∂ 2m Hext
= −m × − mz + (m × ez )
∂t ∂x2 HK − 4π Ms
∂m bJ t0 ∂m
+αm × +  . (13.39)
∂t l0 ∂x

Taking into account the fact that m2 = (M/Ms)2 = 1, that is, the mag-
nitude of the magnetization is a constant at temperature well below the
Curie temperature, we introduce the following stereographic transfor-
mation for the complex function ψ [45],
mx my
ψ= +i . (13.40)
1 + mz 1 + mz
LMDs and NMDs will be obtained from ψ.
As we well know, the exact spin wave solutions of the LLG equation
can be obtained in the presence of spin polarized current [28]. Because
the deviation of magnetization from the direction of the field is small
when Q > 1 (high magnetic field), we only consider low-energy exci-
tation. We then seek for the nonlinear magnetization vector m using
the following ansatz

m = δmx ex + δmy ey exp [i (kx − ωt)] + ez , (13.41)

where δmx and δmy stand for the rescaled amplitudes of the spin wave
in the x- and y-directions, respectively, and k and ω are the wavenumber
and angular frequency of the spin wave, respectively. Equation (13.41)
corresponds to the following stereographic transformation

ψ = ψ0 exp [i(kx − ωt)] with ψ0 = (δmx + iδmy )/2.

Asking the Eq. (13.41) to satisfy the linear part of Eq. (13.39) yields
the following dispersion relation of the SW:
 2
M
ω = (1 + iα) ω0 1 + l02 ks − bJ A
s
, (13.42)
γ
13.3 Current Driven Dynamics of Soliton 471

where
HK − 4π Ms + Hext Ms
ω0 = γ − b2
J  A, and
1 + α2 2γ 1 + α 2
  2 16γ 2 A2
l0 = .
8γ 2 Ms A (HK − 4π Ms + Hext ) − Ms2 b2J

We can easily see that the spin-polarized current changes the energy
gap, which vanishes for large current density
e 
je ≥ 2AMs (HK − 4π Ms + Hext ).
P
This means that the spin-polarized current without other stimulation
such as thermal effects will excite SWs when its density increases
beyond a critical value.
Assuming that m2x + m2y  m2z , or |ψ|2  1 (case of small devia-
tions of magnetization from the equilibrium direction), and consider-
ing the long wavelength approximation [2] yield the following NLS
equation:
 
∂ψ ∂ 2ψ Hext t0 ∂ψ
i = − 1+ ψ + 2 |ψ|2 ψ + ibJ .
∂t ∂x 2 HK − 4π Ms l0 ∂x
(13.43)
Using the HBM [48], we can easily derive the one- and two-soliton
solutions of Eq. (13.43). The one-soliton solution reads
 #  $ 
kR Hext
ψ= ! " exp i ηI + 1 + t ,
cosh ηR + C2D HK − 4π Ms

where    
t0 exp [2 |λD |]
η = k x + bJ − ik t , exp [CD ] = ,
l0 4kR2

with R and I standing for the real and imaginary parts, respectively, and k being a free
complex parameter (wavenumber). Using the initial soliton location and amplitude,
the complex parameter k can be determined. Using the expression for ψ given by Eq.
(13.40) yields the following one-soliton solution of magnetization:
472 13 Current Driven Dynamics of Magnetization in Ferromagnet …
     
2kR Hext CD
mx = cos ηI + 1 + t cosh ηR + ,
 HK − 4π Ms 2
     
2kR Hext CD
my = sin ηI + 1 + t cosh ηR + , (13.44)
 HK − 4π Ms 2
2kR2
mz = 1 − , with  = cosh2 [ηR + CD /2] + kR2 .


The one-soliton solution (13.44) represents a 2π DW with rescaling


magnetization height which constitutes the bound state of double π
wall [49]. The DW velocity is
l0
V = −bJ − 2kI , (13.45)
t0
and consists of two parts. It is clearly seen from Eq. (13.45) that the
spin-polarized current alters the soliton velocity V as −bJ , proportional
to the electron current density. Due to the external field, the mx and my
components of the MS oscillate periodically with time.
In the situation when alternating currents are applied, that is,
μB
bJ = Pje cos [ωt] ,
eMs
we obtain the same solution as above with the only change in the center
of mass  
bJ
η=k x+ sin [ωt0 t] − ikt .
ωl0
Thus, in the case of alternating currents, the center of mass of the
MS vibrates with the same frequency as alternating currents. As well
as we know, the spin current density in ferromagnetic material with
nonuniform magnetization and adiabatic approximation is given as
[28]
μB
jx = Pje m,
e
so that the corresponding one-soliton excitation in nanowire will pro-
vide spin current with a magnetic pulse like the optical pulse in non-
linear media.
13.3 Current Driven Dynamics of Soliton 473

13.3.2 Bright Soliton

In the case of the low-Q mode (Q < 1) corresponding to the situation


when the uniaxial anisotropy field is less than the demagnetization field
such as permalloy, the SW solutions of Eq. (13.37) are different. In the
present situation, the characteristic time and length to be used are

1 2A
t0 = and l0 = .
γ (4π Ms − HK ) (4π Ms − HK )Ms

Using the same approach as in the high-Q mode, the stereographic


transformation (13.40) is introduced and used to transform the LLG
equation (13.37) into the NLS equation
 
∂ψ ∂ 2ψ Hext t0 ∂ψ
i = + 1− ψ − 2 |ψ|2 ψ + ibJ .
∂t ∂x 2 4π Ms − HK l0 ∂x
(13.46)
As the one-soliton solution of Eq. (13.46), we obtain
  
ζ
ψ = −λB exp [iB ] (1 + exp [iB ]) + (1 − exp [iB ]) tanh ,
2
where
 
t0 Hext
B = kx + 2λ + k + kbJ + 1 +
2 2
t + 0 ,
l0 4π Ms − HK
⎡ ⎤
PB 4λ2B − PB2
B = arctan ⎣ ⎦,
PB2 − 2λ2B
   
t0
ζB = PB x − 4λB − PB − 2k + bJ
2 2 t + ζ0 ,
l0

k, λB , PB , 0 , and ζ0 being all real constants. We then obtain the


following solutions of magnetization:
474 13 Current Driven Dynamics of Magnetization in Ferromagnet …
  
λB ζ
mx = cos [B + B ] − cos [B ] + (cos [B + B ] + cos [B ]) tanh ,
 2

  
λ ζ
my = sin [B + B ] − sin [B ] + (sin [B + B ] + sin [B ]) tanh , (13.47)
 2

 
1    ζ
mz = 4 1 − λ2 + PB2 − PB2 tanh2 ,
4 2

where   
1  2 ζB
= 4 1 + λB − PB + PB tanh
2 2 2
.
4 2
In this solution, not only the velocity changes, but also the spin-
polarized current induces the oscillation of mx and my components
of the magnetization.

13.3.3 Current Driven Interaction of Spin Wave and Soliton

Let us now investigate the properties of obtained MS on a nonlinear


SW background. It is evident that m ≡ (mx , my , mz ) = (0, 0, 1) forms
the ground state of system, and Eq. (13.39) admits two types of excited
states, namely, the spin wave solution and MS solution. In the case of
high enough magnetic field, the deviation of magnetization from the
ground state is small for these two types of solutions. To obtain these
excited states, we make the transformation

ψ = mx + imy , mz = 1 − |ψ|2 . (13.48)

If we insert Eq. (13.48) into Eq. (13.39), we obtain the following NLS
equation in ψ
 
∂ψ ∂ 2ψ ∂ 2 mz ∂ψ ∂mz t0 ∂ψ
i = mz 2 − ψ − α mz − ψ + ibJ
∂t ∂x ∂x 2 ∂t ∂t l0 ∂x
 
Hext
− mz + ψ. (13.49)
HK − 4π Ms
13.3 Current Driven Dynamics of Soliton 475

Two special solutions of Eq. (13.49) can be obtained easily: one is the
trivial solution, ψ = 0, which corresponds to the ground state m =
(0, 0, 1), and the other is bulk spin wave excitations,

ψ = Ac exp [i(−kc x + ωc t)] ,

which corresponds to the temporal periodic spin current

mx = Ac cos [−kc x + ωc t] ,
my = Ac sin [−kc x + ωc t] , (13.50)

mz = 1 − A2c ,

where ωc , kc , and Ac are the frequency of spin wave , the dimensionless


wavenumber, and the transverse wave amplitude satisfying the condi-
tion Ac  1, respectively. In the case of the attractive interaction, the
nonlinear spin waves in ferromagnet with anisotropy lead to MS.
Our main purpose in the present subsection is to seek the exact
analytical soliton solution of magnetization on a nonlinear spin wave
(NSW) background that can be used to describe soliton propagation
in a uniaxial ferromagnetic nanowire with spin torque. It is evident
that the NLS equation (13.49) is not an exactly integrable equation.
Imposing some restrictions of this equations, special exact analytical
solutions can be obtained. In the following, we consider the case of
the long-wavelength approximation [2] when damping is neglected,
and assume the dimensionless wavenumber kc to satisfy the condition
kc  1. Inserting mz from Eq. (13.48) into Eq. (13.49) and keeping
only nonlinear terms of the order of the magnitude of |ψ|2 ψ yield the
following integrable equation:

∂ψ ∂2 1 ∂
i = 2 ψ + ψ |ψ|2 + iAJ ψ − ω0 ψ, (13.51)
∂t ∂x 2 ∂x
where
Hext t0
ω0 = 1 + and AJ = bJ .
HK − 4π Ms l0
By means of the Hirota methods [29, 48], we can easily obtain the
soliton solutions on the background of the ground state, ψ = 0. Next,
the straightforward Darboux transformation [51–54] can be used to
476 13 Current Driven Dynamics of Magnetization in Ferromagnet …

build general expressions of soliton solution of Eq. (13.51), with the


help of which we can discuss the properties of soliton solution on the
NSW background.
The DT mainly consists of firstly transforms the nonlinear equation
into the Lax representation, and then in terms of a series of transforma-
tions; finally, with the use of an obvious seed solution, one can build
algebraically the soliton solution of the nonlinear equation. Consider-
ing the spectral parameter
μ1 ν1
λ = λ1 ≡ +i
2 2
with real parameters μ1 and ν1 , we obtain the following one-soliton
solution
2μ1
mx = Ac cos [ϕ] + (Q1 cos [ϕ] − Q2 sin [ϕ]) ,
1
2μ1
my = Ac sin [ϕ] + (Q1 sin [ϕ] + Q2 cos [ϕ]) , (13.52)
1

   
2μ1 Q1 2 2μ1 Q2 2
mz = 1 − Ac + − ,
1 1

where

ϕ = −kc x + ωc t,
 
1 2
Q1 = Ac L1R cosh [θ1 ] + |L1 | + Ac cos [1 ] ,
2
4
  
1 2
Q2 = Ac L1I sinh θ1 + |L1 | − Ac sin [1 ] ,
2
4
 
1
1 = |L1 |2 + A2c cosh [θ1 ] + Ac L1R cos [1 ] ,
4
with

θ1 = 2D1R x + 2 (D1 δ1 )R t + 2x0 , 1 = 2D1I x + 2 (D1 δ1 )I t − 2ϕ0 .


(13.53)
In Eq. (13.53), subscript R and I stand for the real part and imaginary
part, respectively. The other parameters are taken as
13.3 Current Driven Dynamics of Soliton 477

 2
kc A2
D1 = i + λ1 − c , δ1 = −2iλ1 − kc + AJ , (13.54a)
2 4
   
kc 1  C2  1  C2 
L1 = −i − D1 − λ1 , x0 = − ln   , ϕ0 = arg   , (13.54b)
2 2 C1 2 C1

where C1 and C2 are two free complex constants.


The exact analytical solution (13.52) has the following two impor-
tant properties: When μ1 = 0, the solution (13.52) reduces to SW
background
ψ = Ac exp [i(ωc t − kc x)] ,

where the magnon density is also constant. On the other hand, when the
SW amplitude and wavenumber are zero, that is, Ac = kc = 0, solution
(13.52) turns to the soliton solution
2μ1
mx = cos [1 + η] ,
cosh [θ1 ]
2μ1
my = sin [1 + η] , (13.55)
cosh [θ1 ]

4μ21
mz = 1 − ,
cosh2 [θ1 ]

where
 
2
θ1 = μ1 x + (2ν1 + AJ ) t + x0 ,
μ1
#   $
1  2 2
 1 = ν1 x − μ − ν1 − A J t − ϕ 0 ,
2
(13.56)
ν1 1 ν1
η = ω0 t.

We then conclude that the exact soliton solution (13.52) describes a


one-soliton solution for the spin-polarized current in ferromagnetic
nanowire propagating on the periodic spin current background (13.50)
[29].
It is seen from the soliton solution (13.55) that the spatially localized
excitation [49] denoted by the transverse amplitude 2μ1 deviates from
the ground state m = (0, 0, 1). The mx and my components precess
around the component mz with the frequency
478 13 Current Driven Dynamics of Magnetization in Ferromagnet …

1 = ν12 − μ21 + AJ + ω0 .

Moreover, this soliton solution is characterized by the velocity of soli-


ton center v1 = −(2ν1 + bJ t0 /l0 ) and the soliton width 1/μ1 . In addi-
tion, the soliton solution (13.55) corresponds to a static magnetic soli-
ton with three integrals of the motion and the uniform magnon density
|ψ1 |2 = 2u1 along the direction of soliton propagation. For this static
magnon soliton, the carrier wavenumber ks,1 = −ν1 and the carrier
frequency 1 satisfy the following dispersion law
Hext
1 = ks,1
2
− ks,1 AJ − μ21 + 1 + ;
HK − 4π Ms
this dispersion law shows that the magnetic field contributes to the
precession frequency only. Computing the magnetic soliton energy,
we find
2
2 t0 Hext 1
E1 = −bJ 2 − μ21 + 1 + + m∗ ν12 ,
4l0 HK − 4π Ms 2

where m∗ is the dimensionless effective mass of soliton and takes the


value m∗ = 1/2.
It follows from Eqs. (13.55) and (13.56) that the velocity and the
precessional frequency of soliton on a background of the ground state
m = (0, 0, 1) can be changed by adjusting the bJ term [13, 29]. The
bright soliton obtained with the use of the exact analytical solution
(13.52) is characterized by the wavenumber ks = −2D1I , the width
1/(2D1R ), the envelope velocity v1 = −(D1 δ1 )R /D1R , and the initial
position −x0 /D1R of the center. The initial position of the soliton cen-
ter moves x0 (2/μ1 − 1/D1R ) by the SW and offers a new way for
controlling the soliton in space.
It follows from the expressions of D1 and δ1 given by Eq. (13.54a)
that the soliton velocity and width are modulated by the amplitude Ac
and wavenumber kc of SW, as shown in Fig. 13.11. As we can clearly
see from Figs. 13.11a and c, the absolute values of the soliton velocity
and width increase with increasing Ac . It is seen from Fig. 13.11b that
the value of kc near −ν1 has evident effect on the soliton velocity. In
the limit kc = −ν1 , the soliton width reaches its maximal value, as one
can see from Fig. 13.11d.
13.3 Current Driven Dynamics of Soliton 479

Fig. 13.11 a Velocity versus amplitude Ac for kc = 0.1 and ν1 = −0.12 (red line), and kc = −0.1
and ν1 = 0.12 (blue dotted line). b Velocity versus spin wavenumber kc for Ac = 0.06 and ν1 =
−0.15. c Width versus amplitude Ac for kc = −0.1 and ν1 = 0.12. d Width versus spin wavenumber
kc for Ac = 0.06 and ν1 = −0.15. The numerical values of different parameters are: μ1 = 0.1,
l0 = 2 × 10−8 cm, t0 = 5.7392 × 10−12 s, and bJ = 52 cm/s. Reprint from Ref. [57], Copyright
2022, with permission from American Physical Society

Analyzing Eq. (13.53) one can see that in the case when D1I δ1I =
δ1R D1R , that is, (D1 δ1 )R = 0, the wave parameter θ1 does not depend
on time t; this means that the wave velocity −(D1 δ1 )R/D1R becomes
zero, which implies that the soliton is trapped in space by the nonlinear
SW. Mathematically, this condition reads
 
eMs l0 D1I
je = −μ1 − ν1 + k c ; (13.57)
PμB t0 D1R

Equation (13.57) is determined by the wave amplitude, the character-


istic velocity l0 /t0 , the nonlinear SW and the soliton wavenumber, and
the parameters (eMs /PμB ). From a bulky computation, we find that
the above condition in Eq. (13.57) for μ1  ν1 and Ac  kc reduces
to
l0 eMs
je ≈ −2ν1 .
t0 PμB
When Ac = 0, that is, the SW amplitude vanishes, the trapping condi-
tion in Eq. (13.57) becomes
480 13 Current Driven Dynamics of Magnetization in Ferromagnet …

l0 eMs
je = −2ν1 ,
t0 PμB
and is determined by the soliton wavenumber ν1 , the parameter
(eMs /PμB ), and the characteristic velocity l0 /t0 . This means that in the
special case when Ac  kc and μ1  ν1 , the background has almost no
effect on the trapping condition. For the materials of Co3 Pt alloy films
[46] with high perpendicular anisotropy, we use the following numer-
ical values of parameters: HK = 1 × 104 Oe, A = 1.0 × 10−6 erg/cm,
4π Ms = 1 × 102 Oe, γ = 1.76 × 107 Oe−1 s−1 , P = 0.35, as well as
the dimensionless parameters kc = 0.05, Ac = 0.02, ν1 = −0.12, and
μ1 = 0.1. Computing the critical electric current trapping soliton with
these parameters, we find je = 1.867 × 104 A/cm2 . Here, it should be
noted, as we can see from Eqs. (13.52) and (13.19) and the expres-
sions of δ1 given by Eq. (13.54a), that the term bJ changes the soliton
velocity and the frequency affecting the soliton energy. This last soli-
ton property is characterized by the temporal and spatial period along
the propagating direction,
(D1 δ1 )R x0
x=− t− ,
D1R D1R
denoted by
D1R (D1 δ1 )R
π  2 and π  ,
δ1I D1R + D1I
2 δ1I
2 D2 + D2
1R 1I

respectively.
Some novel properties of solution (13.52) can be explained by dis-
cussing the special case kc = −ν1 and analyzing in detail two repre-
sentative situations with Ac , μ1 > 0: (i) the SW amplitude Ac exceeds
the half of the transverse amplitude 2μ1 of soliton (μ21 < A2c ); (ii) the
SW amplitude Ac is smaller than the half of transverse amplitude 2μ1
of soliton (μ21 > A2c ).
(i) In the special case when μ21 < A2c , the exact analytical solution
(13.52) reduces to
13.3 Current Driven Dynamics of Soliton 481

mx = R1 cos [ϕ] − R2 sin [ϕ] ,


my = R1 sin [ϕ] + R2 cos [ϕ] , (13.58)

ζ1 (Ac cosh [θ1 ] cos [1 ] − μ1 )
mz = 1 − A2c ,
(Ac cosh [θ1 ] − μ1 cos [1 ])2

where ϕ is given by Eq. (13.53), and the other parameters are given as

cosh [θ1 ]
ζ1 = 4μ1 κ1 , κ = A2c − μ21 , R1 = −Ac + 2κ12
2
,
Ac cosh [θ1 ] − μ1 cos [1 ]
sinh [θ1 ]
R2 = −2μ1 κ1 ,
Ac cosh [θ1 ] − μ1 cos [1 ]
with
 
2ϕ0
θ1 = μ1 κ1 t + 2x0 , 1 = κ1 x − ν1 t − . (13.59)
κ1

Analyzing Eq. (13.59) reveals that the exact analytical solution (13.58)
is periodic in the space coordinate with period 2π/κ1 , and aperiodic
in the time, as we can see from Fig. 13.12. We can clearly see from
Fig. 13.12 that the soliton background becomes unstable under mod-
ulation. This means that the solution (13.58) can be use for describ-
ing the MI phenomenon [54]. Our calculations show that along the
soliton propagation direction, the component
mz reaches (a) the max-
imum mz = 1 at cos [1 ] = 2μ21 − A2c /μ1 Ac when A2c /4 < μ21 <
A2c ; these points are 
not excited even on the SW background, and (b)
the minimum mz = 1 − (2μ1 + Ac )2 at sin [1 ]= 0. When μ21 <
A2c /4, the maximum of the component mz is mz = 1 − (2μ1 − Ac )2
and is reached at 1 = 0, while its minimum is mz = 1 − (2μ1 + Ac )2
and is reached at 1 = π. From these results, we conclude that the lin-
ear combination of transverse amplitude of SW and magnetic soliton
can be obtained in these special cases A2c < 4μ21 < 4A2c and 4μ21 < A2c .
482 13 Current Driven Dynamics of Magnetization in Ferromagnet …

Fig. 13.12 Evolution plot of the component mz defined in Eq. (13.58) with condition of kc = −ν1
and μ21 < A2c . The plot is generated with the following numerical values of parameters: μ1 = 0.12,
ν1 = −0.1, Ac = 0.16, bJ = 34.8 cm/s, l0 = 2 × 10−10 m, t0 = 5.7392 × 10−12 s, x0 = −1.27,
and ϕ0 = 0. Reprint from Ref. [57], Copyright 2022, with permission from American Physical
Society

Fig. 13.13 Evolution plot of the component mz given in Eq. (13.60) with the condition of kc = −ν1
and μ21 > A2c . The parameters used for generating this figure are: μ1 = 0.1, ν1 = −0.08, Ac = 0.06,
bJ = 41.8 cm/s, l0 = 2 × 10−10 m, t0 = 5.7392 × 10−12 s, x0 = 3, and ϕ0 = 0. Reprint from Ref.
[57], Copyright 2022, with permission from American Physical Society

(ii) In the special case when μ21 > A2c , the solution (13.52) becomes

mx = R1 cos [ϕ] − R2 sin [ϕ] ,


my = R1 sin [ϕ] + R2 cos [ϕ] , (13.60)

ζ2 (μ1 − Ac cosh [θ1 ] cos [1 ])
mz = 1 − A2c ,
(μ1 cosh [θ1 ] − Ac cos [1 ])2

where
13.3 Current Driven Dynamics of Soliton 483

ζ2 = 4μ1 κ22 , κ2 = μ21 − A2c ,
2κ22 cos [1 ]
R1 = −Ac + , 1 = −μ1 κ2 t − 2ϕ(13.61)
0,
μ1 cosh [θ1 ] − Ac cos [1 ]
2μ1 κ2 sin [1 ] x0
R2 = , θ1 = κ2 (x − ν1 t) + 2 .
μ1 cosh [θ1 ] − Ac sin [1 ] κ2
As we can see from expressions (13.60) and (13.61), the main charac-
teristics of soliton solution can be summarized as follows:
(α) the envelope velocity ν1 = −(2ν1 + bJ t0 /l0 ) of the soliton is the
same on both the ground state background m = (0, 0, 1) and the back-
ground of a periodic spin current in Eq. (13.50);
(β) the amplitude of component mz given in Eq. (13.60) periodically
oscillates in time as we can see from Fig. 13.13. A detail computation
shows that the amplitude of component mz given in Eq. (13.60) reaches
its minimum at θ1 = 0 and is given by
  2
4μ1 μ1 − A 2
c
mz = 1 − A2c − , (13.62)
μ1 − Ac cos [1 ]

and its maximum



μ21 A2c sin2 [1 ]
mz = 1− 2 (13.63)
μ1 − A2c cos2 [1 ]

at
2μ1 Ac cos [1 ]
cosh [θ1 ] = − .
Ac cos [1 ] μ1
In Fig. 13.14a, we display the evolution along the propagation direc-
tion of the minimum amplitude of mz and maximum amplitude of com-
ponent mz by Eqs. (13.62) and (13.63), respectively; in this Fig. 13.14a,
the solid line shows the evolution of the SW intensity, that is, the
amplitude of the SW background. Figure 13.14b shows the locations
of minimum and maximum amplitudes in the (t, x) plane. Compar-
ing the narrowness, shape, and depth of the soliton, we can see from
Fig. 13.14 that the narrower the soliton, the sharper the peak, and the
deeper the two dips at the wings of the bright soliton. This obtained
484 13 Current Driven Dynamics of Magnetization in Ferromagnet …

Fig. 13.14 a Evolution plot of the minimum amplitude of the component mz given in Eq. ( 13.62)
(dashed line), the maximum amplitude of the component mz defined in Eq. (13.63) (dotted line), and
the background amplitude of the component mz (solid line) given by Eq. (13.50). b Location of the
minimum amplitude (solid line) and the maximum amplitude (dotted line) in the time-propagation
distance plane. Different plots are generated with the same as those in Fig. 13.14. Reprint from Ref.
[57], Copyright 2022, with permission from American Physical Society

feature illustrates the characteristic breather behavior of the soliton


embedded on a non-vanishing CW background of a periodic solution
for ferromagnetic nanowire under consideration.

13.3.4 Dynamics of Magnetic Rogue Wave

As in the previous subsection, we can easily find two basic solutions


of Eq. (13.51) that may help in investigating the dynamics of magnetic
rogue wave (MRW). The first one is ψ = 0, and corresponds to the
ground state m = (0, 0, 1), while the second one is the SW defined as

ψ = Ac exp [i(ωc t − kc x)] ,

where kc and ωc are the dimensionless wavenumber and frequency,


respectively. As we well know, the interaction between SW and a
dynamic soliton in perpendicular anisotropic ferromagnetic nanowire
can result in the novel MRW [55]. Such a MRW can be realized by
adjusting the relation of amplitude and wavenumber for SW and a
dynamic soliton.
We can observe from Eq. (13.52) that the exact solution com-
monly exhibits a breather character and a soliton with time periodic
modulation amplitude, which plays the role of SW background and
denotes the periodization process of magnetic soliton. The soliton
13.3 Current Driven Dynamics of Soliton 485

Fig. 13.15 Breather velocity Vq and width W versus the amplitude Ac and wavenumber kc of spin
wave. a, d Soliton velocity Vq and width W versus Ac and kc . The red star signs the location of
the rogue wave. b, c Dependence of velocity on Ac and kc . e, f Dependence of velocity on Ac and
kc . Different plots are generated with parameters u1 = 0.2, ν1 = −0.3, and AJ = 0.2. Reprint from
Ref. [57], Copyright 2022, with permission from American Physical Society

solutions are characterized by the soliton width 1/D1R , the breather


propagation velocity vq ≡ −(D1 δ1 )R /D1R , and the slope direction
Vθ ≡ D1R x + (D1 δ1 )R t. Using the expressions of D1 and δ1 given in
Eq. (13.54a), we compute the soliton velocity and width and find that
they are modulated by the amplitude Ac and wavenumber kc of SW as
one can see in Fig. 13.15a and d. As it is clearly seen from Fig. 13.15c
and f, the absolute values of the soliton velocity and width increase
with increasing modulation parameter Ac and, as Ac increases con-
tinuously, values of parameter kc near −ν1 will seriously affect the
soliton velocity. In the limit kc = −ν1 , the soliton velocity and width
reach their maximal values, as we can see in Fig. 13.15b and e. The
phenomenon shown in Fig. 13.15 describes to magnonic spin-transfer
torque [56] denoting the transfer of spin angular momentum from SW
background to a dynamic soliton.
From the above discussion on the exact analytical solution of Eq.
(13.52), we already know that under the condition of ν1 = −kc , the crit-
ical point u1 = Ac forms a dividing line between the MI phenomenon
(u1 > Ac ) and the periodization phenomenon (u1 < Ac ) [57, 58]. When
the breather character depends on parameter u1 of modulation to a dif-
ferent extent, one may observe different physical behaviors, as shown
486 13 Current Driven Dynamics of Magnetization in Ferromagnet …

in Fig. 13.16 where we have depicted the evolution of a magnetic soli-


ton in the limit u1 → A± c . As we can see from plots of Fig. 13.16, two
different asymptotic behaviors appear in the limit processes u1 → A− c
[u1 → A+ c ] for ν1 = −kc . Figure 13.16(a) [11.16(b)] obtained in the
limit u1 → A− +
c [u1 → Ac ] demonstrates a spatial periodic process of
a soliton [a localized process of the NSW background along the slope
direction Kβ = −(D1 δ1 )I /D1I for t0 = 0], where we can observe that
the spatial separation of adjacent peak and each peak value increases
rapidly as the modulation parameter u1 approaches Ac . In the case
when u1 → A+ c , each peak value decreases rapidly. More interestingly,
the following MRW corresponding to the pentagram sign indicated in
Figs. 13.15a and d can be obtained in the limit case of u1 → Ac [55]:

4 1 − itA2c
ψ = Ac exp [iϕ] , with ζ = AJ − 2kc .
1 + A2c (x + ζ t)2 + t 2 A4c
(13.64)
Equation (13.64) shows the typical rogue wave (RW). It follows from
this RW that the magnons accumulated from SW background con-
verge a single hump whose the critical amplitude is AQ = 3Ac . It is
clearly seen from Eq. (13.64) that the RW is localized in both space
and time, and the wave localization is captured completely at x = 0 and
t = 0 by SW background. As shown in Fig. 13.17 where the magnon
density distribution against the background for the breather solution
in Eq. (13.52) is displayed, the realization of MRW showed in Eq.
(13.64) attributes to the fact that the aggregation of magnons gradu-
ally increases when u1 approaches A− c and gradually decreases as u1
+
approaches Ac . As u1 → Ac , the temporal localization magnetic rogue
wave is excited and the magnon density peak is found to be |ψ|2 = 9A2c .
With the use of the magnon density distribution against the back-
ground
 2
ρq (x, t) = |ψ(x, t)|2 −  ψ(x, t)|x=±∞  ,

we can deeply study the properties of the MRW in the ferromagnetic


nanowire under consideration. The formative mechanism of MRW in
ferromagnetic nanowire is demonstrated in Fig. 13.17. We can see from
Fig. 13.17a that the critical peak of the MRW will appear under the
condition of u1 → Ac . Also, as we can see from plots of Fig. 13.17, the
magnons in the background gradually gather toward each individual
13.3 Current Driven Dynamics of Soliton 487

Fig. 13.16 Evolution plot of magnetic soliton in the limit processes u1 → A− +


c (a) and u1 → Ac
(b). Other parameters use in generating these plats are: Ac = 0.2, AJ = 0.2, kc = −ν1 = 0.003,
and x0 = t0 = 0. Reprint from Ref. [55], Copyright 2022, with permission from American Physical
Society

Fig. 13.17 Magnon density distribution against the background for different parameter u1 , which
ranges from 0.09 to 0.29 with a step of 0.05. Inset: Magnon density distribution against the back-
ground for the excited formation of magnetic rogue wave. Other parameters used in generating
different plots are: Ac = 0.2, AJ = kc = 0.1 and x0 = t0 = 0. Reprint from Ref. [55], Copyright
2022, with permission from American Physical Society

central part and the envelope becomes sharper when parameter u1 of


modulation approaches the amplitude parameter Ac .
Combining equation (13.64) and the quantity
 2
ρ = |ψ(x, t)|2 −  ψ(x, t)|x=±∞ 

yields the magnon density distribution ρ in a MRW defined as


488 13 Current Driven Dynamics of Magnetization in Ferromagnet …

1 − 2
ρ = 8A2c , with 1 = 1 + t 2 A4c and 2 = A2c (x + ζ t)2 .
(1 + 2 )2
(13.65)
It follows from Eq. (13.65) that the integral for arbitrary time is

+∞
ρ(x, t)dx = 0.
−∞

Using now the condition of ρQ (±1/Ac , 0) = 0, we compute the spatial


width of the hump part in Rw and find 2/Ac . Carrying out a detail
computation, we find the relation S1 + S2 = S3 at a fixed time; this
means that the loss of magnons in background completely transfers to
hump.
It is interesting to know how RW gathers magnons and energy
toward its central part from the background. Mathematically, this can
be explained by the following quantity
 2
 
δ(x, t) ≡ lim ψ(x, t) − ψ(x, t)|x=lψ  .
lQ →±∞

It then follows from Eq. (13.64) that


1
δ(x, t) = 16A2c . (13.66)
(1 + 2 )2
Equation (13.66) denotes the nonuniform exchange of magnons between
RW and its background for different spin currents as shown in Fig. 13.18.
We can find from Eq. (13.66) that
(i) the spin current can control the accumulation and dissipation rate
of magnons, and
(ii) the critical current condition reads AJc = 2kc .
Below (above) the critical value, the magnon exchange decreases (is
accelerated) with increasing current term AJ . In other word, the cases
where the current is below and above the critical current attribute oppo-
site role to spin-transfer torque, as we can see in Fig. 13.18f. We can
well see from Figs. 13.18a–e that the MRW can be created in the differ-
ent direction for (x, t) plane. In the limit AJ = 2kc , the time of magnon
accumulation/dissipation reaches its maximum. As one can see from
13.3 Current Driven Dynamics of Soliton 489

Fig. 13.18 a–e Formation region in (x, t) plane for MRW with different current. The parameter AJ
ranges form 0 to 0.8 with a step of 0.2. f Nonuniform exchange of magnons between RW and the
SW background for different spin current. The inset in panel (f) denotes the maximal accumulation
(or dissipation) process for the critical current value AJ = 2kc . Other parameters used for generating
these plots are: Ac = 0.2 and kc = 0.2. Reprint from Ref. [55], Copyright 2022, with permission
from American Physical Society

Fig. 13.18f, propagating on its background, the magnons accumulate


to the central part when t < 0, which mathematically means that

+∞
8π Ac
ξ(x, t) = δ(x, t)dx =  .
1 + t 2 A4c
−∞

When t > 0, the magnons in the hump start to dissipate into the back-
ground so that the hump gradually decays. The magnetic rogue wave
disappears ultimately before t → +∞.
By means of numerical simulations and with the use of some specific
initial conditions, the excitation of RWs can be recovered. For a better
understanding of such a process of MRW, we consider the initial value
of solution, approximated by

ψ(x, 0) = (ρ + εχ cos [ϕ1 ]) exp [iϕ] , (13.67)

where
2κ1 (κ1 − iu1 ) − A2c 4κ1 u1 (κ1 − iu1 )
ρ= , ε = exp [−x0 ] , χ = , ϕ1 = κ1 x
Ac A2c
490 13 Current Driven Dynamics of Magnetization in Ferromagnet …

Fig. 13.19 a Numerical evolution of the magnetic rogue wave (MRW). b Comparison of magnon
density between the numerical solution (in solid curve) and ideal solution (in ‘◦’). To generate these
plots, we have used the following numerical values of parameters: Ac = 0.2, u1 = 0.19, x0 = 0.72,
AJ = 0.01, kc = 0.2, and ω = 0.2. Reprint from Ref. [55], Copyright 2022, with permission from
American Physical Society


with κ1 = A2c − u12 . We find by direct numerical simulations that
the solution of Eq. (13.51) that satisfies the initial condition (13.67)
is in perfect agreement with the exact analytical solution (13.64), as
we can see from Fig. 13.19. Here, Fig. 13.19a shows the numerical
evolution of the magnetic rogue wave, while Fig. 13.19b compares the
magnon density generated with numerical solution and that generated
with exact analytical solution (13.64). From plots of Fig. 13.19, we
conclude that a small initial perturbation with a small modulation can
induce the generation and breakup of a near-ideal RW.

13.4 Ferromagnetic Resonance in Magnetic Trilayers

Based on the Landau–Lifshitz–Gilbert–Slonczewski (LLGS) equation


[8], we perform in this Section a theoretical study on the current-
driven ferromagnetic resonance [62, 63, 65–71, 84–86] in a magnetic
trilayer [72] with a tilted polarizer [73–86] and a perpendicular analyzer
[63, 83, 87, 88]. Linearizing the model equation, the expression of
output voltage for arbitrary anisotropy in the free and pinned layers is
presented and applied to achieve the ferromagnetic resonance spectra.
13.4 Ferromagnetic Resonance in Magnetic Trilayers 491

13.4.1 Ferromagnetic Resonance in a


Perpendicular-Analyzer Magnetic Trilayer

The physical model to be studied here is a magnetic trilayer with two


ferromagnetic metallic layers such as that depicted in Fig. 13.20. In
this model, the two ferromagnetic metallic layers are separated by a
non magnetic metallic or insulating spacer.We assume that the film
lies in the (x, y) plane and denote by n = nx , ny , nz the free-layer
easy axis. Next, we assume to be fixed the direction of the pinned-layer
magnetization so that various orientations can be chosen by fabrication
[82, 83].
Under the effects of the spin torque, the NMDs in the free layer is
governed by the LLGS equation [8]
∂m γ ∂E ∂m 
= m× + αm× + γ aJ m× m × mp , (13.68)
∂t Ms ∂m ∂t
where γ , Ms , α are the gyromagnetic ratio, the saturation magneti-
zation of the free layer, and the Gilbert damping parameter, respec-
tively, m = mx , my , m
z is the unit vector of the free magnetization,
mp = mpx , mpy , mpz is the unit vector of the pinned magnetization,
and E is the magnetic energy which includes anisotropic and demag-
netization energy and is defined as

E = K 1 − (m · n)2 + 2π Ms2 cos2 [θ] ,

Fig. 13.20 Schematic


drawing of a magnetic
trilayer and coordinate
system. Applied current
flows along the vertical
direction of film (z axis).
Reprint from Ref. [5],
Copyright 2022, with
permission from American
Physical Society
492 13 Current Driven Dynamics of Magnetization in Ferromagnet …

K, θ, and φ being the anisotropy constant, the polar angle of m, and the
azimuthal angle of m, respectively. Throughout this Section, the spin-
torque factor aJ appearing in the LLGS equation (13.68) is assumed
to have the general form
g 1
aJ = J, with g =   ,
2edMs −4 + 3 + m · mp (1 + P)3 / 4P 3/2

where d , P, and J are the thickness of the free layer, the spin polar-
ization, and the current density, respectively. When the current flows
from the fixed layer to the free one, the current density J will be taken
positive. Also, we assume the current to consist of two parts, the dc
part Jdc and the alternating one with the angular frequency ω:

J = Jdc + Jac exp [iωt] .

Also, aJ is assumed to dependent on the materials and the angle


between the pinned and free magnetization.
Introducing the spherical coordinates θ, and φ, the LLGS equation
(13.68) is transformed into the following two coupled equations for θ,
and φ:
 
∂θ ∂φ 
+ α sin θ = γ HK nr ny cos [φ] − nx sin [φ]
∂t ∂t

+γ aJ −mp2 cos [θ] + mpz sin [θ] ,
(13.69)
 
∂φ ∂θ 
sin [θ] −α = −γ HK nr nr + 2π Ms sin [2θ] + γ aJ mp1 ,
∂t ∂t
where

nr = nx sin [θ] cos [φ] + ny sin [θ] sin [φ] + nz cos [θ] ,
nr = nx cos [θ] cos [φ] + ny cos [θ] sin [φ] − nz sin [θ] ,
mp1 = mpx sin [φ] − mpy cos [φ] , mp2 = mpx cos [φ] + mpy sin [φ] ,
2K
HK = (anisotropy field).
Ms
13.4 Ferromagnetic Resonance in Magnetic Trilayers 493

In the absence of the current, the following two coupled equations


determine the initial direction of free magnetization:

nr ny cos [φ] − nx sin [φ] = 0 and Hk nr nr + 2π Ms sin [2θ] = 0.

In the situation when we have parallel analyzer, that is, when either
nx = 1 or ny = 1, the nonlinear magnetization vector m will be
along the x− direction, which is possible only for θ = π/2 and
φ ∈ {0, π/2, π, 3π/2}. In the case of perpendicular analyzer, that is,
when nz = 1„ the nonlinear magnetization vector m will along the z
axis, if HK > 4π Ms , which is possible only when θ ∈ {0, π }. On the
contrary, the magnetization vector m lies in the film plane. In the pres-
ence of the current, the nonlinear magnetization vector m will deviate
from this equilibrium direction; let us denote by θ0 and φ0 the new
equilibrium direction. If we then excite the free magnetization by a
small ac, it will oscillate around θ0 and φ0 . Hence, we introduce the
ansatz

θ = θ0 + θ exp [iωt] and φ = φ0 + φ exp [iωt] . (13.70)

Now, if we insert ansatz (13.70) into equation (13.69) and keep only
the linear terms in θ , φ, and Jac , we arrive to the equilibrium
equations which yield the linear solution of m. We then find that the
new equilibrium direction of the free magnetization is determined by

HK nr0 q2 − adc p3 = 0, − HK nr0 nr0 + 2π Ms sin [2θ0 ] + adc p1 = 0,
(13.71)
where
g0 Jdc
adc = , nr0 = nx sin [θ0 ] cos [φ0 ] + ny sin [θ0 ] sin [φ0 ] + nz cos [θ0 ] ,
2edMs
nr0 = nx cos [θ0 ] cos [φ0 ] + ny cos [θ0 ] sin [φ0 ] − nz sin [θ0 ] ,
q1 = nx cos [φ0 ] + ny sin [φ0 ] , q2 = ny cos [φ0 ] − nx sin [φ0 ] ,
p1 = mpx sin [φ0 ] − mpy cos [φ0 ] , p2 = mpx cos [φ0 ] + mpy sin [φ0 ] ,
p3 = p2 cos [θ0 ] − mpz sin [θ0 ] .

The complex quantities θ and φ appearing in Eq. (13.70) are then


found to be
494 13 Current Driven Dynamics of Magnetization in Ferromagnet …

p3 A22 + p1 A12 p3 A21 + p1 A11


θ = − γ aac , φ = , (13.72)
A22 A11 − A12 A21 A22 A11 − A12 A21
where different parameters are set as follows:

A11 = iω − γ HK nr0 q2 − γ aac p4 + βg0 p32 ,

A12 = iαω sin [θ0 ] − γ HK q22 sin [θ0 ] − nr0 q1
−γ adc (p1 cos [θ0 ] − βg0 p1 p3 sin [θ0 ]) ,
%  &
A21 = −iαω + γ HK n2 r0 − nr0 + 4π Ms cos [2θ] + γ adc βg0 p1 p3 ,
2

A22 = iω sin [θ0 ] + γ HK q2 nr0 cos [θ0 ] + nr0 sin [θ0 ]

−γ adc p2 + βg0 p12 sin [θ0 ] ,

with
1
g0 = , p4 = p2 sin [θ0 ] + mpz cos [θ0 ] ,
−4 + β (3 + p4 )
(1 + P)3 g0 Jac
β= , aac = .
4P 3/2 2edMs
Since θ and φ are complex quantities, the response of the free
magnetization to the alternating current includes two parts, namely,
the in-phase part and the out-of-phase part.
The ferromagnetic resonance signals in the experimental context
can be detected by a direct voltage generated through the rectification
of the ac by the time-dependent magneto-resistance [62, 63]. In this
context, the resistance of the device is assumed, when an ac is applied,
to vary with the precession of free magnetization [62] as follows:
R 
R = R0 + 1 − m · mp
2
R
= R0 + (1 − p4 − 1 cos [ωt] − 2 sin [ωt]) , (13.73)
2
where
   
1 = Re p3 θ − p1 φ sin [θ0 ] , 2 = Im p3 θ − p1 φ sin [θ0 ] ,
13.4 Ferromagnetic Resonance in Magnetic Trilayers 495

and R0 and R are the resistance of the trilayer when the free and
pinned magnetization are parallel and the increase in resistance when
they become anti-parallel, respectively. If we denote by S the cross
section area of the device, the output direct voltage Vdc can then be
given in terms of the time-averaged value of RJac S cos [ωt] as

T
1 RJac S
Vdc = RJac S cos [ωt] = γ aac τ, (13.74)
T 4
0

where τ is free from ω, Jdc , n, and mp and is given as


 
p3 (p3 (p3 A22 + p1 A12 ) + p1 (p3 A21 + p1 A11 )) sin [θ0 ]
τ = Re .
A22 A11 − A12 A21
(13.75)
Therefore, the output signal only reflects the in-phase response of the
free magnetization to the alternating current. With the use of Eqs.
(13.74 and (13.75), it is possible to study with detail the current
driven ferromagnetic resonance in magnetic trilayers with arbitrary
anisotropy. Without loss of generality, we focus on the situation when
nx = 1, which corresponds to the trilayers with a tilted polarizer and a
parallel analyzer.
In our discussion on the current driven ferromagnetic resonance in
magnetic trilayers with arbitrary anisotropy, we focus on the case when
nz = 1 corresponding to the trilayer with a perpendicular analyzer. In
this special case, the anisotropy field favors the free magnetization in
the vertical direction; in the same time, the demagnetizing field favors
in-plane. In our computation, we limit ourselves to the special case of
a single-axis anisotropy (HK > 4π Ms ). As we well know, the critical
current for switching is proportional to HK − 4π Ms for perpendicular
magnetized devices and can be adjusted to a low value [87].
Due to the wide range of the alternating current frequency, the ferro-
magnetic resonance can be realized more effectively just by adjusting
the ac frequency [62–64]. In the following, the frequency-adjusted
ferromagnetic resonance for different directions of the pinned mag-
netization without magnetic field is studied. We take [Co/Ni] × 4
multilayers as an illustrative example [87] with the following numer-
ical values of parameters: γ = 1.75 × 107 Oe−1 ·s−1 , HK = 10154
Oe, P = 0.35, Ms = 650 Gs, d = 3 nm, and α = 0.1. Computing
496 13 Current Driven Dynamics of Magnetization in Ferromagnet …

the effective anisotropy field for this material along the z axis yields
Hk − 4π Ms = 1986 Oe. For the above material, parameter aJ of the
spin-torque has the dimension of magnetic field. For example, comput-
ing aJ for the current density J = 0.2 A/μm2 , we find that aJ varies
from 538 Oe to 2186 Oe by changing the included angle between the
pinned and free magnetization. Setting the included angle to zero, aJ
takes the minimal value, while its maximum value is obtained for 180◦ .
In our discussion, we will take Jac = 0.01 A/μm2 . The correspond-
ing aac then ranges from 27 Oe to 110 Oe and satisfies that condition
aac  (HK − 4π Ms ).
Rotating the coordinate system and making the y axis perpendicular
to the film plane avoid the singularity at θ = 0◦ . It is clearly seen from
Fig. 13.21 where Vdc / (RJac S) is depicted versus the ac frequency
that an increasing in the dc values from negative to positive values turns
the resonant peak sharper. When the current is above 0.089 A/mm2 , the
peak begins to become blunt, as we can clearly see in the inset (c) where
the line-width decreases with the increase in current from negative to
positive values and increases when Jdc > 0.089 A/mm2 . Depending
on the direction of the pinned magnetization and the current, the spin
torque acts as a damping torque or a precessional one, meaning that
it plays dual roles in driving the free magnetization. Also, Fig. 13.21
reveals that the dc can be used to control the effective damping. One
can see from inset (b) of Fig. 13.21 that the resonant frequency can be
adjusted by the dc. The maximum value of the frequency is reached at
Jdc = 0.003 A/mm2 and is found to be about 34.58 GHz.
Now, let us compare the ferromagnetic resonance modes of the per-
pendicular and parallel analyzers. For this aim, we have taken CoFeB
[64] as an example and depicted the spectra of trilayer with a parallel
analyzer in the inset (a) of Fig. 13.21 for θp = 60◦ and φp = 30◦ . As
we can see from different curves in inset (a), the resonance peak is
asymmetric [symmetric] for a parallel [perpendicular] analyzer. Here,
we can also see that independently on the direction of mp , the peak for
the perpendicular analyzer is always symmetric, which is not the case
for parallel analyzer.
Figure 13.22 shows the effects of the pinned magnetization direction
on the ferromagnetic resonance curves. The ferromagnetic resonance
peak in the (y, z) plane becomes blunter and blunter with the angle
θp increasing from 0◦ to 90◦ . As we can see from Fig. 13.22, one can
13.4 Ferromagnetic Resonance in Magnetic Trilayers 497

Fig. 13.21 Ferromagnetic resonance spectra for different dc currents in the case when θp = 60◦
and φp = 30◦ . Here, Jdc ranges from −0.04 A/mm2 to 0.04 A/mm2 in 0.02 A/mm2 steps. Plot (a)
shows the spectra of trilayer with a parallel analyzer. Plots (b) and (c) show the dependence of the
resonant frequency and the resonant line-width on the dc current density, respectively. Reprint from
Ref. [103], Copyright 2022, with permission from American Physical Society

Fig. 13.22 Ferromagnetic resonance spectra for different directions of the pinned magnetization
in the case when Jdc = 0.2 A/μm2 . φp = 90◦ and θp takes values from 0◦ to 90◦ in 15◦ steps.
Plots (a) and (b) show the dependence of the equilibrium location on θp ; plots (c) and (d) show
the dependence of the resonant location and line-width on θp . Reprint from Ref. [103], Copyright
2022, with permission from American Physical Society
498 13 Current Driven Dynamics of Magnetization in Ferromagnet …

use both the magnitude of spin torque and changing the direction of
spin torque to control the ferromagnetic resonance. One can clearly
observe in Figs. 13.22a and b the variation of the equilibrium direction
of the free magnetization when the direction of spin torque changes.
Figure 13.22c and d showing the effects of θp on the resonant location
ωr and line-width ω reveal that the resonant location ωr and line-
width ω have the same behavior: they are the largest with mp in the
film and decrease when the pinned magnetization deviates from this
direction. Moreover, the line-width, as we can see from Fig. 13.22d, is
most in the parallel configuration.

13.4.2 Stability Analysis of Perpendicular Magnetic Trilayers

We now focus on the linear stability analysis of the magnetic trilayers


with perpendicular anisotropy [100–107]. The magnetic device to be
considered here is that shown in Fig. 13.23. This device is formed of
two ferromagnetic metallic layers, the free layer and the pinned layer,
separated by a thin normal metallic/insulating spacer layer. For this
device, the easy axis is assumed to be perpendicular to the film plane
and defined as the y− direction, so that the film lies in the (x, z) plane.
We assume for this device that each of the magnetic anisotropy in
pinned layer and the free layer is perpendicular to the film. When the
electrical current flows perpendicularly through the trilayer, the mag-

Fig. 13.23 Schematic


drawing of a magnetic
trilayer and coordinate
system. For this device, the
easy axis is perpendicular to
the film plane and is defined
as the y direction. The
current is applied
perpendicularly to the film
plane. Reprint from Ref.
[104], Copyright 2022, with
permission from American
Physical Society
13.4 Ferromagnetic Resonance in Magnetic Trilayers 499

netization in the free layer undergoes a spin-transfer torque generated


by current, while spin polarization occurs in the pinned layer. Nonlin-
ear magnetization dynamics of free layer is assumed to be governed
by the below dimensionless LLG equation for which we have included
the Slonczewski and field-like spin torques [14, 91],
   
dm dm J m× m × ey βJ m × ey
− αm× = −m × heff + + .
dτ dτ Jp 1 + λp my jp 1 + λp my
(13.76)
In the dimensionless LLG equation (13.76), α, β, and τ = 4π γ Ms t
are the Gilbert damping constant, the relative strength of field-like
torque to Slonczewski one, and a rescaled variable representing time,
respectively; m denotes the unit vector of free-layer magnetization, the
effective field heff is the sum of anisotropy field, demagnetizing field,
and external magnetic field and takes the form
HK H0
heff = (κmy + h0 )ey with κ = hk − 1, hk = , and h0 = ,
4π Ms 4π Ms
H0 , HK , and ey being the external magnetic field, the anisotropy field,
and the unit vector along the y direction, respectively. On the rhs of
the LLG equation (13.76), the second term stands for the Slonczewski
torque, while the third one is the field-like spin torque. J denotes the
dc current density, which will be taken positive when the current flows
from the pinned layer to the free one, λp and Jp are two parameters
given as

(1 + P)3 3 (1 + P)3 Ms2
λp =  , and Jp = − 8 4π de ,
3 (1 + P)3 − 16P 3/2 2P 3/2 

P, , e, d , and Ms , and d being the spin polarization ratio, the reduced


Planck constant, the electron charge, the film thickness of the free layer,
and the saturation magnetization of the free layer.
In terms of the polar angle θ and the azimuthal angle φ, we reduce
the LLG equation (13.76) into the following pair of coupled differential
equations
500 13 Current Driven Dynamics of Magnetization in Ferromagnet …

∂θ ∂φ
+α sin [θ] = (κ sin [θ] sin [φ] + h0 ) cos [φ]
∂t ∂t
J (cos [θ] sin [φ] + β cos [φ])
− ,
Jp 1 + λp sin [θ] sin [φ]
(13.77)
∂θ ∂φ
−α + sin [θ] = − (κ sin [θ] sin [φ] + h0 ) cos [θ] sin [φ]
∂t ∂t
J (cos [φ] − β cos [θ] sin [φ])
− .
Jp 1 + λp sin [θ] sin [φ]

It is evident that (θ0 , φ0 ) = (π/2, π/2) and (θ0 , φ0 ) = (π/2, 3π/2)


are two equilibrium points (fixed point) of system (13.77); therefore,
m = ey and −ey are equilibrium directions of magnetization dynamics
of the LLG equation (13.76). It is also evident that the magnetization
under a small perturbation will oscillate slightly around these equilib-
rium directions which can be stable or unstable. In the following, we
investigate the linear stability of these equilibrium directions ey and
−ey .
Denoting by δθ and δφ the small perturbations for θ and φ, respec-
tively, we perturb the fixed point (FP) (θ0 , φ0 ) = (π/2, π/2) as follows

θ = π/2 + δθ and φ = π/2 + δφ. (13.78)

Inserting Eq. (13.78) into Eq. (13.77) and linearizing the resulting
system with respect with both δθ and δφ yield the following linear
differential equations for the perturbation δθ and δφ

d δθ d δφ J βJ
+α =  δθ − κ + h0 −  δφ,
dτ dτ Jp 1 + λ p Jp 1 + λ p
(13.79)

d δθ d δφ βJ J
−α + = κ + h0 −  δθ +  δφ.
dτ dτ Jp 1 + λ p Jp 1 + λ p

Next, we seek nontrivial solutions of linear differential system (13.79)


in the form (δθ, δφ)T = (θ, φ)T exp [iωt] where θ and φ are
the perturbation amplitudes and ω is the perturbation complex angular
frequency. Inserting these δθ and δφ into system (13.79) and asking
13.4 Ferromagnetic Resonance in Magnetic Trilayers 501

that |θ| + |φ| > 0 yield the following dispersive equation for per-
turbation
⎛ ! "⎞
βJ
μ − J 1+λJ
αμ + κ + h0 − J 1+λ
p( ! p) p( p) ⎠
det ⎝ βJ
"
−αμ − κ + h0 − J 1+λ μ − J 1+λ
J
p( p) p( p)
= 0, μ = iω. (13.80)

Solving equation (13.80) in μ = iω leads to



b± 
iω = μ =  , (13.81)
2 1 + α2

where
2
J (1 + αβ) J (α − β)
b=2  − α (κ + h0 ) , and  = −4  + κ + h0 .
Jp 1 + λ p Jp 1 + λ p

Because  < 0, the FP (θ0 , φ0 ) = (π/2, π/2) is an unstable focus


when b > 0 and a stable one when b < 0. The condition for the insta-
bility is obtained as

α (h0 + κ) 1 + λp
J > Jc =
1
Jp . (13.82)
1 + αβ
In the same manner, we study the linear stability of the FP (θ0 , φ0 ) =
(π/2, 3π/2) and obtain the following condition for its instability

 α (h0 − κ) 1 − λp
J < Jc =
2
Jp . (13.83)
1 + αβ
For spin-transport devices such as that showed in Fig. 13.23, the
incorporation of free layers within the magnetic layers affects the spin
polarization. The incorporation of the free layer may yield high spin-
orbit scattering and therefore reduce spin-torque efficiencies [87]. The
situation is different for many perpendicularly magnetized devices.
For a better understanding of the process that generally occurs for
spin-transport devices, we consider as an example the [Co/Ni]×4 mul-
tilayers [87] and use conditions (13.82) and (13.83) to discuss the
nonlinear magnetization dynamics for this device. For the [Co/Ni]×4
502 13 Current Driven Dynamics of Magnetization in Ferromagnet …

Fig. 13.24 Phase diagram


defined in parameter space
spanned by β , h0 , and J.
The blue and red surfaces
represent the instability
boundaries of ‘quasi P’ and
‘quasi-AP’ states,
respectively. Reprint from
Ref. [104], Copyright 2022,
with permission from
American Physical Society

multilayers, we will, during our computational calculations, use the


following relative parameters: the gyromagnetic ratio will be taken as
γ = 1.75 × 107 Oe−1 s−1 , the anisotropic field Hk = 10154 Oe, the
spin polarization P = 0.35, the Gilbert damping parameter α = 0.1,
the thickness of free layer d = 3 nm, and the saturate magnetization
will be Ms = 650 G. For this material, the numerical values of parame-
ters λp and Jp are λp = 0.6048, Jp = 1.89 A/mm2 , while the effective
anisotropy field κ = 0.2431 and along the y axis.
Inserting the above relative parameters and other parameters for
the [Co/Ni]×4 multilayers into Eqs. (13.82) and (13.83), we find that
the instability current is proportional to the external magnetic field.
In Fig. 13.24, the phase diagram is displayed in (J , β, h0 ) parameter
space. In this Fig. 13.24, the instability current of equilibrium direc-
tions y and −y, are represented in the blue and red surfaces, respec-
tively. When we adjust the both current and external magnetic field,
we can observe several magnetic states. It is seen from Fig. 13.24 that
if J < Jc1 [J > Jc2 ], the free-layer magnetization is near the y [−y]
direction. These two regions are denoted by ‘quasi-P’ and ‘quasi-AP’,
respectively. In the situation when the external field and current satisfy
the condition Jc1 < J < Jc2 , the equilibrium directions ey and −ey are
all unstable, the out-of-plane precessional states can emerge (this new
state is denoted by ‘OPP’) in Fig. 13.24. We can clearly see that in
other region, the two stable equilibrium directions ey and −ey coexist
and are sensitive to the initial state; here, we use the same notation,
‘quasi-P’ or ‘quasi-AP’. The system is in ‘quasi-P’ [quasi-AP’] state
with the initial free-layer magnetization along the y [−y] direction.
13.4 Ferromagnetic Resonance in Magnetic Trilayers 503

Fig. 13.25 Dynamic evolutions of the free-layer magnetization for points ‘a’, ‘b’, ‘c’ in Fig. 13.25.
a, a Magnetization precession with the initial direction of free-layer magnetization along the y and
−y axis, respectively. b, c Magnetization reversal from y to −y and from −y to y, respectively.
Reprint from Ref. [104], Copyright 2022, with permission from American Physical Society

We have thus obtain that depending on the spin polarized current and
external magnetic field, the switching from stable states to precessional
ones and the reversal between y and −y orientations can be realized.
As we can see from Fig. 13.24, the system enters the out-of-plane
precessional state without reference to the initial direction when β, h0 ,
and J are located in the region ‘OPP’. For β, h0 , and J are located
in ‘quasi-AP’ region, we find that the magnetization reversal from
‘quasi-P’ to ‘quasi-AP’ [‘quasi-AP’ to ‘quasi-P’] happens if the initial
direction is along the y [−y] axis.
As one can see from Fig. 13.24, the instability current becomes
weakly dependent on the field-like spin torque. The instability current
of each of quasi-antiparallel state and quasi-parallel state is inversely
proportional to the relative strength β of field-like torque to Slon-
czewski one; we found that this instability current can be decreased by
10% when we increase β from 0 to 1.
The validity of the phase diagram showed in Fig. 13.24 can be con-
firmed as follows. We randomly pick out three points a, b, and c
from Fig. 13.24 and plot the corresponding dynamic evolutions of the
free-layer magnetization in Fig. 13.25. Figures 13.25a and a show the
dynamic oscillating evolutions of the free-layer magnetization gener-
ated with the initial direction of free-layer magnetization along the y
504 13 Current Driven Dynamics of Magnetization in Ferromagnet …

and −y axis, respectively. In Figs. 13.25b and c, we show the magneti-


zation reversal trajectory from y to −y and from −y to y, respectively.
It is important to underline that we cannot apply the linearizing
method for analyzing the stability of out-of-plane precession [86,
107]. In the case of out-of-plane precession, one employs the Poincare
Bendixson theorem to specify the precessional region [108, 109] . We
find that the static states (fixed points) or limit cycle (self-oscillation)
are the only possible final states of the system and chaos is precluded
for free-layer magnetization and therefore, the steady precession must
emerge when none of the static solutions, that is, FPs and limit cycles,
are stable.
Taking into account the rotational symmetry of the problem [110],
limit cycle solutions m0 (τ ) can be assumed to be in the form of magne-
tization precessing around the y axis with a constant angular velocity.
In other words, we suppose that the limit cycle solution m0 (τ ) satisfies
the following first order differential equation
d m0 (τ )  
= ω m0 (τ ) × ey . (13.84)

Inserting this Eq. (13.84) into Eq. (13.68), we find the following two
equations for ω:
βJ 1
ω = −κmy − h0 + , (13.85)
Jp 1 + λp my
J
ω=−  . (13.86)
αJp 1 + λp my

Imposing to the two expression for ω to coincide, we can exclude my


from Eqs. (13.85) and (13.86); focusing on −1 ≤ my ≤ 1, we obtain
the precession frequency of the ‘OPP’ state as follows

α 2 J 2 κ − h λ 2 + 4καJ λ J (1 + βα)
κ − h0 λp p 0 p p p
ω= + .
2λp (1 + βα) 2αJp λp (1 + βα)
(13.87)
In Fig. 13.26, we display the frequency ω versus the external mag-
netic field and the current. We can see from this Fig. 13.26 that ω
increases with increasing external magnetic field. One can clearly see
13.5 Conclusion 505

Fig. 13.26 Precession


frequency ω versus J and h0
for β = 0.1. The inset shows
the dependence of precession
frequency ω on β for point
‘a’ shown in Fig. 13.24.
Reprint from Ref. [104],
Copyright 2022, with
permission from American
Physical Society

from the inset of Fig. 13.26 that the velocity of magnetization pre-
cession diminishes with increasing field-like spin torque. Increasing
parameter β from 0 to 1 can decrease the precession frequency for
more than 10%. The dependence of precession frequency on β can
be analyzed qualitatively. From Eq. (13.68), the influence of the field-
like spin torque on precession is related to the directions of effec-
tive field and current. Here, the precession exists in the region that
heff = κmy + h0 < 0 and J < 0. Thus, the precession produced by the
field-like spin torque is opposite to the effective field-induced preces-
sion. Because the effective field-induced torque is the leading cause of
precession, the frequency decreases with increasing strength of field-
like spin torque.

13.5 Conclusion

In this last Chapter of the Book, we have mainly investigated the


dynamics of magnetization in ferromagnet. Firstly, we have investi-
gated the effects of adiabatic and non-adiabatic spin torque on motion
and distortion of domain walls. The equilibrium motion of magnetiza-
tion versus the critical current is analyzed. Secondly, we have used the
Hirota bilinear method to discuss the properties of dynamic magnetic
one- and two-solitons in uniaxial anisotropic ferromagnetic nanowire
driven by spin-transfer torque. The interaction of a spin wave and these
solitons in a uniaxial ferromagnetic nanowire is studied with details.
506 13 Current Driven Dynamics of Magnetization in Ferromagnet …

Carrying out linear stability analysis, the MI of dark soliton propagat-


ing on the nonlinear spin wave background is analyzed too. Lastly, the-
oretical investigations are done for the current-excited and frequency-
adjusted ferromagnetic resonance in magnetic trilayers, especially, for
devices formed of one free layer and one pinned layer, separated by a
spacer layer. A particular attention has been paid on the stability analy-
sis of perpendicular magnetic trilayers. Also, the effects of the critical
current and precession frequency on the relative strength of field-like
torque to Slonczewski one have been investigated.

References

1. L.D. Landau, E.M. Lifshitz, Theory of the dispersion of magnetic. Phys. Z. Sowjetunion 8,
153 (1935)
2. A.M. Kosevich, B.A. Ivanov, A.S. Kovalev, Magnetic solitons. Phys. Rep. 194, 117 (1990)
3. H.J. Mikeska, M. Steiner, Solitary excitations in one-dimensional magnets. Adv. Phys. 40,
191 (1991)
4. T.L. Gilbert, A phenomenological theory of damping in ferromagnetic materials. IEEE Trans.
Magn. 40, 3443 (2004)
5. Z.-D. Li, P.-B. He, W.-M. Liu, Dynamics of magnetization in ferromagnet with spin-transfer
torque. Chin. Phys. B 23(11), 117502 (2014)
6. N.L. Schryer, L.R. Walker, The motion of 180◦ domain walls in uniform dc magnetic fields.
J. Appl. Phys. 45, 5406 (1974)
7. A.P. Malozemoff, J.C. Slonczewski, Magnetic Domain Walls in Bubble Material (Academic,
New York, 1979)
8. J.C. Slonczewski, Current-driven excitation of magnetic multilayers. J. Magn. Magn. Mater.
159, L1 (1996)
9. L. Berger, Emission of spin waves by a magnetic multilayer traversed by a current. Phys. Rev.
B 54, 9353 (1996)
10. G. Tatara, H. Kohno, J. Shibata, Microscopic approach to current-driven domain wall dynam-
ics. Phys. Rep. 468, 213 (2008)
11. Y.B. Bazaliy, B.A. Jones, S.C. Zhang, Modification of the Landau-Lifshitz equation in the
presence of a spin-polarized current in colossal- and giant-magnetoresistive materials. Phys.
Rev. B 57, R3213 (1998)
12. G. Tatara, H. Kohno, Theory of current-driven domain wall motion: spin transfer versus
momentum transfer. Phys. Rev. Lett. 92, 086601 (2004)
13. Z. Li, S. Zhang, Domain-wall dynamics and spin-wave excitations with spin-transfer torques.
Phys. Rev. Lett. 92, 207203 (2004)
14. S. Zhang, P.M. Levy, A. Fert, Mechanisms of spin-polarized current-driven magnetization
switching. Phys. Rev. Lett. 88, 236601 (2002)
15. N.J. Zabusky, M.D. Kruskal, Interaction of “ Solitons” in a collisionless plasma and the
recurrence of initial states. Phys. Rev. Lett. 15, 240 (1965)
16. V.N. Serkin, A. Hasegawa, T.L. Belyaeva, Nonautonomous solitons in external potentials.
Phys. Rev. Lett. 98, 074102 (2007)
17. A. Hubert, R. Schaferd, Magnetic Domains: The Analysis of Magnetic Microstructures
(Springer, New York, 2000)
18. T. Ono, H. Miyajima, K. Shigeto, K. Mibu, N. Hosoito, T. Shinjo, Propagation of a magnetic
domain wall in a submicrometer magnetic wire. Science 284, 468 (1999)
References 507

19. G.S.D. Beach, C. Nistor, C. Knutson, M. Tsoi, J.L. Erskine, Dynamics of field-driven domain-
wall propagation in ferromagnetic nanowires. Nat. Mater. 4, 741 (2005)
20. L. Thomas, M. Hayashi, X. Jiang, R. Moriya, C. Rettner, S.S.P. Parkin, Oscillatory dependence
of current-driven magnetic domain wall motion on current pulse length. Nature 443, 197
(2006)
21. A. Thiaville, Y. Nakatani, J. Miltat, Y. Suzuki, Micromagnetic understanding of current-driven
domain wall motion in patterned nanowires. Europhys. Lett. 69, 990 (2005)
22. J. Grollier, P. Boulenc, V. Cros, A. Hamzic, A. Vaurès, A. Fert, G. Faini, Switching a spin
valve back and forth by current-induced domain wall motion. Appl. Phys. Lett. 83, 509 (2003)
23. C.H. Marrows, Spin-polarised currents and magnetic domain walls. Adv. Phys. 54, 585 (2005)
24. S.S.P. Parkin, M. Hayashi, L. Thomas, Magnetic domain-wall racetrack memory. Science
320, 190 (2008)
25. J.S. Yang, C. Nistor, G.S.D. Beach, J.L. Erskine, Magnetic domain-wall velocity oscillations
in permalloy nanowires. Phys. Rev. B 77, 014413 (2008)
26. X.R. Wang, P. Yan, J. Lu, High-field domain wall propagation velocity in magnetic nanowires.
Europhys. Lett. 86, 67001 (2009)
27. J.A. Katine, F.J. Albert, R.A. Buhrman, Current-driven magnetization reversal and spin-wave
excitations in Co Cu Co pillars. Phys. Rev. Lett. 84, 3149 (2000)
28. Z. Li, S. Zhang, Domain-wall dynamics driven by adiabatic spin-transfer torques. Phys. Rev.
B 70, 024417 (2004)
29. P.B. He, W.M. Liu, Nonlinear magnetization dynamics in a ferromagnetic nanowire with spin
current. Phys. Rev. B 72, 064410 (2005)
30. H. Koo, C. Krafft, R.D. Gomez, Current-controlled bi-stable domain configurations in ele-
ments: an approach to magnetic memory devices. Appl. Phys. Lett. 81, 862 (2002)
31. A. Yamaguchi, T. Ono, S. Nasu, K. Miyake, K. Mibu, T. Shinjo, Real-space observation of
current-driven domain wall motion in submicron magnetic wires. Phys. Rev. Lett. 92, 077205
(2004)
32. E. Saitoh, H. Miyajima, T. Yamaoka, G. Tatara, Current-induced resonance and mass deter-
mination of a single magnetic domain wall. Nature 432, 203 (2004)
33. C.K. Lim, T. Devolder, C. Chappert, J. Grollier, V. Cros, A. Vaures, A. Fert, G. Faini, Domain
wall displacement induced by subnanosecond pulsed current. Appl. Phys. Lett. 84, 2820
(2004)
34. J. Ohe, B. Kramer, Dynamics of a domain wall and spin-wave excitations driven by a meso-
scopic current. Phys. Rev. Lett. 96, 027204 (2006)
35. X. Waintal, M. Viret, Current-induced distortion of a magnetic domain wall. Europhys. Lett.
65, 427 (2004)
36. V.N. Serkin, A. Hasegawa, Exactly integrable nonlinear Schrodinger equation models with
varying dispersion, nonlinearity and gain: application for soliton dispersion. IEEE J. Quantum
Electron. 8, 418 (2002)
37. Q.Y. Li, Z.D. Li, S.X. Wang, W.W. Song, G.S. Fu, Nonautonomous solitons of Bose-Einstein
condensation in a linear potential with an arbitrary time-dependence. Opt. Commun. 282,
1676 (2009)
38. F. Zhao, Z.D. Li, P.B. He, Q.Y. Li, W.M. Liu, Nonautonomous helical motion of magnetization
in ferromagnetic nanowire driven by spin-polarized current and magnetic field. Eur. Phys. J.
B 84, 197 (2011)
39. Z.Z. Sun, J. Schliemann, Fast domain wall propagation under an optimal field pulse in mag-
netic nanowires. Phys. Rev. Lett. 104, 037206 (2010)
40. A. Goussev, J.M. Robbins, V. Slastikov, Domain-wall motion in ferromagnetic nanowires
driven by arbitrary time-dependent fields: an exact result. Phys. Rev. Lett. 104, 147202 (2010)
41. J.C. Slonczewski, Modes of natural vibration for magnetic domains. J. Magn. Magn. Mater.
23, 305 (1981)
42. H. Munthe-Kaas, Lie-butcher theory for Runge-Kutta methods. BIT 35, 572 (1995)
43. K. Engo, On the construction of geometric integrators in the RKMK class. BIT 40, 41 (2000)
508 13 Current Driven Dynamics of Magnetization in Ferromagnet …

44. Z.D. Li, Q.Y. Li, X.R. Wang, W.M. Liu, J.Q. Liang, G.S. Fu, Screw-pitch effect and velocity
oscillation of a domain wall in a ferromagnetic nanowire driven by spin-polarized current. J.
Phys.: Condens. Matter 22, 216001 (2010)
45. M. Lakshmanan, K. Nakamura, Landau-Lifshitz equation of ferromagnetism: exact treatment
of the Gilbert damping. Phys. Rev. Lett. 53, 2497 (1984)
46. Y. Yamada, W.P. Van Drent, E.N. Abarra, T. Suzuki, High perpendicular anisotropy and
magneto-optical activities in ordered alloy films. J. Appl. Phys. 83, 6527 (1998)
47. L. Belliard, J. Miltat, V. Kottler, V. Mathet, C. Chappert, T. Valet, Stripe domains morphology
versus layers thickness in CoPt multilayers. J. Appl. Phys. 81, 5315 (1997)
48. R. Hirota, J. Math. Phys. 14, 805 (1973)
49. W.M. Liu, B. Wu, X. Zhou, D.K. Campbell, S.T. Chui, Q. Niu, Interacting domain walls in
an easy-plane ferromagnet. Phys. Rev. B 65, 172416 (2002)
50. S. Takeno, Dynamical Problems in Soliton Systems (Springer, Berlin, 1984)
51. V.B. Matveev, M.A. Salli, Springer Series in Nonlinear Dynamics (Springer, Berlin, 1991)
52. C.H. Gu, H.S. Hu, Z.X. Zhou, Darboux Transformation in Soliton Theory and its Geometric
Applications (Shanghai Scientific and Technical Publishers, Shanghai, 1999)
53. L. Li, Z.H. Li, S.Q. Li, G.S. Zhou, Modulation instability and solitons on a CW background
in inhomogeneous optical fiber media. Opt. Commun. 234, 169 (2004)
54. S.Q. Li, L. Li, Z.H. Li, G.S. Zhou, Properties of soliton solutions on a CW background in
optical fibers with higher-order effects. J. Opt. Soc. Am. B 21, 2089 (2004)
55. F. Zhao, Z.D. Li, Q.Y. Li, L. Wen, G.S. Fu, W.M. Liu, Magnetic rogue wave in a perpendicular
anisotropic ferromagnetic nanowire with spin-transfer torque. Ann. Phys. 327, 2085 (2012)
56. P. Yan, X.S. Wang, X.R. Wang, All-magnonic spin-transfer torque and domain wall propaga-
tion. Phys. Rev. Lett. 107, 177207 (2011)
57. Z.D. Li, Q.Y. Li, L. Li, W.M. Liu, Soliton solution for the spin current in a ferromagnetic
nanowire. Phys. Rev. E 76, 026605 (2007)
58. Z.D. Li, Q.Y. Li, P.B. He, Z.G. Bai, Y.B. Sun, Interaction of a nonlinear spin-wave and
magnetic soliton in a uniaxial anisotropic ferromagnet. Ann. Phys. 322, 2945 (2007)
59. E.B. Myers, D.C. Ralph, J.A. Katine, R.N. Louie, R.A. Buhrman, Current-induced switching
of domains in magnetic multilayer devices. Science 285, 867 (1999)
60. S.I. Kiselev, J.C. Sankey, I.N. Krivorotov, N.C. Emley, R.J. Schoelkopf, R.A. Buhrman, D.C.
Ralph, Microwave oscillations of a nanomagnet driven by a spin-polarized current. Nature
425, 380 (2003)
61. K.J. Lee, A. Deac, O. Redon, J.P. Nozieres, B. Dieny, Excitations of incoherent spin-waves
due to spin-transfer torque. Nat. Mater. 3, 877 (2004)
62. A.A. Tulapurkar, Y. Suzuki, A. Fukushima, H. Kubota, H. Maehara, K. Tsunekawa, D.D.
Djayaprawira, N. Watanabe, S. Yuasa, Spin-torque diode effect in magnetic tunnel junctions.
Nature 438, 339 (2005)
63. J.C. Sankey, P.M. Braganca, A.G.F. Garcia, I.N. Krivorotov, R.A. Buhrman, D.C. Ralph,
Spin-transfer-driven ferromagnetic resonance of individual nanomagnets. Phys. Rev. Lett.
96, 227601 (2006)
64. G.D. Fuchs, J.C. Sankey, V.S. Pribiag, L. Qian, P.M. Braganca, A.G.F. Garcia, E.M. Ryan,
Z.P. Li, O. Ozatay, D.C. Ralph, R.A. Buhrman, Spin-torque ferromagnetic resonance mea-
surements of damping in nanomagnets. Appl. Phys. Lett. 91, 062507 (2007)
65. W. Chen, J.M.L. Beaujour, G. De Loubens, A.D. Kent, J.Z. Sun, Spin-torque driven ferro-
magnetic resonance of synthetic layers in spin valves. Appl. Phys. Lett. 92, 012507 (2008)
66. H. Kubota, A. Fukushima, K. Yakushiji, T. Nagahama, S. Yuasa, K. Anda, H. Maehara, Y.
Nagamine, K. Tsunekawa, D.D. Djayaprawira, N. Watanabe, Y. Suzuki, Quantitative measure-
ment of voltage dependence of spin-transfer torque in MgO-based magnetic tunnel junctions.
Nat. Phys. 4, 37 (2008)
67. J.C. Sankey, Y.T. Cui, J.Z. Sun, J.C. Slonczewski, R.A. Buhrman, D.C. Ralph, Measurement
of the spin-transfer-torque vector in magnetic tunnel junctions. Nat. Phys. 4, 67 (2008)
68. H.W. Xi, Y.M. Shi, K.Z. Gao, Spin-current effect on ferromagnetic resonance in patterned
magnetic thin film structures. J. Appl. Phys. 97, 033904 (2005)
References 509

69. L. Torresa, G. Finocchio, L. Lopez-Diaz, E. Martinez, M. Carpentieri, G. Consolo, B. Azzer-


boni, Micromagnetic modal analysis of spin-transfer-driven ferromagnetic resonance of indi-
vidual nanomagnets. J. Appl. Phys. 101, 09A502 (2007)
70. J.N. Kupferschmidt, S. Adam, P.W. Brouwer, Theory of the spin-torque-driven ferromagnetic
resonance in a ferromagnet/normal-metal/ferromagnet structure. Phys. Rev. B 74, 134416
(2006)
71. A.A. Kovalev, G.E.W. Bauer, A. Brataas, Current-driven ferromagnetic resonance, mechanical
torques, and rotary motion in magnetic nanostructures. Phys. Rev. B 75, 014430 (2007)
72. M. Farle, Ferromagnetic resonance of ultrathin metallic layers. Rep. Prog. Phys. 61, 755
(1998)
73. S.C. Shin, A.K. Agarwala, Magnetic anisotropy in TbFe thin films prepared at oblique inci-
dence. J. Appl. Phys. 63, 3645 (1988)
74. A. Layadi, A theoretical investigation of the effect of the oblique anisotropy axis on the
ferromagnetic resonance linewidth. J. Appl. Phys. 86, 1625 (1999)
75. A. Layadi, Ferromagnetic resonance modes in single and coupled layers with oblique
anisotropy axis. Phys. Rev. B 63, 174410 (2001)
76. K.Z. Gao, H.N. Bertram, Magnetic recording configuration for densities beyond 1 Tb/in2 and
data rates beyond 1 Gb/s. IEEE Trans. Magn. 38, 3675 (2002)
77. J.P. Wang, Y.Y. Zou, C.H. Hee, T.C. Chong, Y.F. Zheng, Approaches to tilted magnetic record-
ing for extremely high areal density. IEEE Trans. Magn. 39, 1930 (2003)
78. Y.Y. Zou, J.P. Wang, C.H. Hee, T.C. Chong, Tilted media in a perpendicular recording system
for high areal density recording. Appl. Phys. Lett. 82, 2473 (2003)
79. J.P. Wang, Tilting for the top. Nat. Mater. 4, 191 (2005)
80. M. Albrecht, G. Hu, I.L. Guhr, T.C. Ulbrich, J. Boneberg, Magnetic multilayers on
nanospheres. Nat. Mater. 4, 203 (2005)
81. A.K. Singh, J. Yin, H.Y.Y. Ko, T. Suzuki, Fabrication and characterization of granular-type
(FePt/Fe3 Pt) and ( FePt/fcc FePt)-tilted magnetic recording media. J. Appl. Phys. 99, 08E704
(2006)
82. Y. Zhou, C.L. Zha, S. Bonetti, J. Persson, J. Aermanb, Spin-torque oscillator with tilted fixed
layer magnetization. Appl. Phys. Lett. 92, 262508 (2008)
83. R. Sbiaa, R. Law, L. Tan Ei, T. Liew, Spin transfer switching enhancement in perpendicular
anisotropy magnetic tunnel junctions with a canted in-plane spin polarizer. J. Appl. Phys. 105,
013910 (2009)
84. P.B. He, Z.D. Li, A.L. Pan, Q. Wan, Q.L. Zhang, R.X. Wang, Y.G. Wang, W.M. Liu, B.S. Zou,
Theory of ferromagnetic resonance in magnetic trilayers with a tilted spin polarizer. Phys.
Rev. B 78, 054420 (2008)
85. P.B. He, Z.D. Li, A.L. Pan, Q. Wan, Q.L. Zhang, R.X. Wang, Y.G. Wang, W.M. Liu, B.S. Zou,
Current-driven ferromagnetic resonance in magnetic trilayers with a tilted spin polarizer. J.
Appl. Phys. 105, 043908 (2009)
86. R.X. Wang, P.B. He, Z.D. Li, A.L. Pan, Q.H. Liu, Phase diagram of magnetic multilayers
with tilted dual spin torques. J. Appl. Phys. 109, 039905 (2011)
87. S. Mangin, D. Ravelosona, J.A. Katine, M.J. Carey, B.D. Terris, E.E. Fullerton, Current-
induced magnetization reversal in nanopillars with perpendicular anisotropy. Nat. Mater. 5,
210 (2006)
88. T. Seki, S. Mitani, K. Yakushiji, K. Takanashi, Magnetization reversal by spin-transfer torque
in 90 configuration with a perpendicular spin polarizer. Appl. Phys. Lett. 89, 172504 (2006)
89. S. Urazhdin, O. Birge Norman, W.P. Pratt, J. Bass, Jr., Current-driven magnetic excitations in
permalloy-based multilayer nanopillars. Phys. Rev. Lett. 91, 146803 (2003)
90. S. Petit, C. Baraduc, C. Thirion, U. Ebels, Y. Liu, M. Li, P. Wang, B. Dieny, Spin-torque
influence on the high-frequency magnetization fluctuations in magnetic tunnel junctions.
Phys. Rev. Lett. 98, 077203 (2007)
91. Z. Li, S. Zhang, Z. Diao, Y. Ding, X. Tang, D.M. Apalkov, Z. Yang, K. Kawabata, Y. Huai,
Perpendicular spin torques in magnetic tunnel junctions. Phys. Rev. Lett. 100, 246602 (2008)
510 13 Current Driven Dynamics of Magnetization in Ferromagnet …

92. A.M. Deac, A. Fukushima, H. Kubota, H. Maehara, Y. Suzuki, S. Yuasa, Y. Nagamine, K.


Tsunekawa, D.D. Djayaprawira, N. Watanabe, Bias-driven high-power microwave emission
from MgO-based tunnel magnetoresistance devices. Nat. Phys. 4, 803 (2008)
93. Z.M. Zeng, P. Upadhyaya, P. Khalili Amiri, K.H. Cheung, J.A. Katine, J. Langer, K.L. Wang,
H.W. Jiang, Enhancement of microwave emission in magnetic tunnel junction oscillators
through in-plane field orientation. Appl. Phys. Lett. 99, 032503 (2011)
94. Y. Zhou, J. Akerman, Perpendicular spin torque promotes synchronization of magnetic tunnel
junction based spin torque oscillators. Appl. Phys. Lett. 94, 112503 (2009)
95. Z.Z. Zhang, Y.C. Qiu, Q.Y. Jin, Y.W. Liu, Micromagnetic study of fieldlike spin torque effect
on the magnetization switching in tunnel junctions with perpendicular anisotropy. Appl. Phys.
Lett. 97, 172501 (2010)
96. Y. Zhou, Effect of the field-like spin torque on the switching current and switching speed of
magnetic tunnel junction with perpendicularly magnetized free layers. J. Appl. Phys. 109,
023916 (2011)
97. H. Meng, J.P. Wang, Spin transfer in nanomagnetic devices with perpendicular anisotropy.
Appl. Phys. Lett. 88, 172506 (2006)
98. A.D. Kent, B. Oyilmaz, B.E. Hel, Spin-transfer-induced precessional magnetization reversal.
Appl. Phys. Lett. 84, 3897 (2004)
99. K.J. Lee, O. Redon, B. Dieny, Analytical investigation of spin-transfer dynamics using a
perpendicular-to-plane polarizer. Appl. Phys. Lett. 86, 022505 (2005)
100. Ya.B. Bazaliy, B.A. Jones, S.C. Zhang, Towards metallic magnetic memory: how to interpret
experimental results on magnetic switching induced by spin-polarized currents. J. Appl. Phys.
89, 6793 (2001)
101. J. Grollier, V. Cros, H. Jaffres, A. Hamzic, J.M. George, G. Faini, J. Ben Youssef , H. Le Gall,
A. Fert, Field dependence of magnetization reversal by spin transfer. Phys. Rev. B 67, 174402
(2003)
102. H. Morise, S. Nakamura, Stable magnetization states under a spin-polarized current and a
magnetic field. Phys. Rev. B 71, 0144439 (2005)
103. R.X. Wang, P.B. He, Q.H. Liu, Z.D. Li, A.L. Pan, B.S. Zou, Y.G. Wang, Tilted spin torque-
driven ferromagnetic resonance in a perpendicular-analyzer magnetic trilayer. J. Magn. Magn.
Mater. 322 , 2264 (2010)
104. R.X. Wang, J.L. Zhao, P.B. He, G.N. Gu, Z.D. Li, A.L. Pan, Q.H. Liu, Stability analysis of
perpendicular magnetic trilayers with a field-like spin torque. J. Magn. Magn. Mater. 327,
132 (2013)
105. M.D. Stiles, J. Miltat, Spin Dynamics in Confined Magntic Structures III (Hillebrands, B.,
Thiaville, A., ed.) (Springer, New York, 2006)
106. U. Ebels, D. Houssameddine, I. Firastrau, D. Gusakova, C. Thirion, B. Dieny, L.D. Buda-
Prejbeanu, Macrospin description of the perpendicular polarizer-planar free-layer spin-torque
oscillator. Phys. Rev. B 78, 024436 (2008)
107. Y. Zhou, S. Bonetti, C.L. Zha, J. Akerman, New J. Phys. 11, 103028 (2009)
108. G. Bertotti, C. Serpico, I.D. Mayergoyz, A. Magni, M. d’Aquino, R. Bonin, Magnetization
switching and microwave oscillations in nanomagnets driven by spin-polarized currents. Phys.
Rev. Lett. 94, 127206 (2006)
109. L. Perko, Differential Equations and Dynamical Systems (Springer, New York, 1996)
110. G. Bertotti, R. Bonin, C. Serpico, M. d’Aquino, I.D. Mayergoyz, Spin-wave analysis of
uniaxial nanopillar devices. J. Appl. Phys. 105, 07D104 (2009)
Index

A Dissipative network, 65, 68, 72, 209


Azimuthal modulational instability, 369 Dissipative solitons, 5, 47, 191, 395, 398,
400
Domain-wall resonance, 448
B
Baseband Modulational Instability (Base-
band MI), 54, 65, 110, 146, 153, 156, E
302, 305, 320 Electrical network, 20
BdG equation, 378, 402 Electric nonlinear transmission network,
Bi-inductance transmission network, 93 124, 136
Bogoliubov–de Gennes, 217, 227 Elliptic ordinary differential equation, 365
Bose-Einstein condensates, 215, 223, 241, Envelope soliton, 11, 15–17, 24, 80, 84, 88,
244, 249, 264, 276, 289, 292, 302, 90, 113, 141, 478
329, 338, 349, 363, 373, 395, 396, Envelope soliton-like, 68, 82, 83, 85, 87, 91,
398 93
Bright soliton, 27

C F
Chen-Lee-Liu equation, 5 Femtosecond pulse, 21
Chen-Liu equation, 5 Femtosecond soliton, 23, 265, 276, 283
Chirped bright soliton, 28 Femtosecond soliton, 276
Chirped double-kink solitons, 264 Feshbach-resonance, 216
Chirped femtosecond soliton, 22 Finite-Background, 385
Chirped solitons, 4, 264, 266 Frequency chirp, 23, 44
Continuous-wave, 226
CW background, 8, 13, 17, 23, 26, 27, 29,
44, 137, 152, 302, 331, 335, 391 G
Gilbert-damping, 453
Ginzburg-Landau equation, 5, 74, 93, 94,
D 172, 183, 396
Darboux transform, 8, 329, 373, 387 GP equations, 217, 329, 339, 341, 349, 357,
Darboux transformation, 14, 404, 476 359, 364, 368, 373, 403
Dark soliton, 27 Gross-Pitaevskii equation, 241, 263, 265,
Dissipative elements, 47, 66, 68, 72, 80, 92, 290, 354, 370
169 Group velocity dispersion, 21, 22

© The Editor(s) (if applicable) and The Author(s), under exclusive license 511
to Springer Nature Singapore Pte Ltd. 2022
E. Kengne and W. Liu, Nonlinear Waves,
https://doi.org/10.1007/978-981-19-6744-3
512 Index

H Magnetic trilayers, 495, 498


Harmonic-oscillator frequency, 243 Magnetic-wave, 210, 216
Harmonic-oscillator potential, 217 Magnetization dynamics, 431
Harmonic potential, 225, 228, 234, 240, 265, Mathematical modeling, 448
279, 293, 302, 364, 365 Matter-waves, 215–217, 223, 224, 235, 240,
Heisenberg model, 216 302, 329, 373, 395
Hirota bilinear method, 435, 447, 505 Metastable soliton, 373
Hirota method, 475 MI criterion, 68, 79, 99, 231
Hole soliton, 15, 24, 25, 80, 113 MI gain, 232, 234, 240
MI growth rate, 57, 68, 80, 113, 127, 191
MI regions, 11, 15, 17
I Modulated waves, 3, 5, 49, 52, 75, 79, 92,
Integrable condition, 109, 249, 263, 266, 94, 110, 126, 156, 165, 172, 181, 186,
267, 279, 282, 283, 289, 291, 294, 189, 191
299, 304, 307, 310, 316, 322, 333, Modulated wavetrains, 93, 104, 181
336, 338, 341, 347, 403, 430 Modulational Instability (MI), 4, 7, 8, 11,
Inter-atomic interaction, 223, 225, 241, 243, 15, 17, 23, 44, 75, 77, 79, 80, 85, 92,
244, 247, 252, 255, 258, 261, 263, 96, 104, 110, 112, 113, 163, 169, 173,
265, 278, 289, 292, 303, 322, 349, 175 189, 217, 226, 227 244, 253, 263,
355, 363, 371 306, 335, 377 387
Intrapulse Raman scattering, 22 Multi-component BEC, 217, 329
Multi-component NLTN, 5
Multisoliton solution, 403
J Multistability, 384
Jost solutions, 403

N
K Near-integrable condition, 196
Kerr nonlinearities, 22 NLS equation, 8, 11, 12, 15, 19, 23, 24, 31,
Korteweg-de Vries (KdV), 7, 170 34, 39, 42, 45, 47, 112, 124, 126, 138,
Kundu–Eckhaus, 125 145, 163, 169, 215, 217, 223, 225,
228, 230, 234, 240, 251, 264, 267,
377, 387, 405
L Noguchi, 7, 124, 209
Lambert W-kink pulses, 58, 63, 65 Noguchi electric network, 23
Lambert W-kink signals, 47, 65 Noguchi NLTN, 8
Lambert W-kink solitary-wave, 47 Non-autonomous, 125
Lambert W-kink solutions, 54, 59, 61, 62 Non-autonomous domain wall, 462, 467
Landau–Lifshitz–Gilbert equations, 216, Non-autonomous domain wall motion, 465,
431, 447, 448 466
Linear magnetization dynamics, 434 Non-autonomous dynamics, 459
Linear Schrödinger equation, 367 Non-autonomous equations, 216
Linear waves, 67, 71, 72, 188 Non-autonomous Gross-Pitaevskii, viii
Long ferromagnetic nanowire, 431 Non-autonomous helical motion, 458, 462,
Lossless network, 13, 156 464
Lossless NLTN, 7, 163 Non-autonomous matter-wave solitons, 217
Lossy network, 49, 54, 58, 63, 90, 94, 172, Non-autonomous motion, 466
181, 183, 209 Non-autonomous rogue wave, 124, 127, 289,
292, 295, 298, 302
Non-autonomous solitons, 216, 223
M Nonlinear magnetization, 395, 403, 417–
Magnetic solitons, 217, 395, 399, 400, 409, 419, 424, 426, 431
431, 433, 441, 469, 474, 478, 481, Nonlinear modulated wave, 3, 4, 9, 23, 30,
484 43, 44, 47, 67, 72, 124, 163
Index 513

Nonlinear partial differential equation, 3, 22, S


73, 75, 96, 109, 215, 441, 448 Schrödinger equation, 8, 44, 80, 215, 224,
Nonlinear Schrödinger equation, 22 266, 434, 438
Nonlinear Transmission Networks Self-frequency, 26, 45
(NLTNs), 3–5, 7, 9, 47, 48, 52, 67, Self-frequency shift, 22, 31, 65
78, 92, 104, 109, 124, 137, 145, 146, Self-phase modulation, 22
156, 163, 169, 177, 181, 189, 224, Self-steepening, 26, 31, 65
242 Self-steepening effect, 138
Nonlinear wave, 3 Self-steepening term, 21
Self-steeping, 45, 65
Single-Component soliton, 377
O Soliton management, 241, 243, 249, 329
One-component BEC, 332 Solitons collision, 431
One-dimensional BEC, 278, 293 Soliton solutions, 431
One-dimensional condensate, 216, 223, 241 Spin-orbit coupling, 217
One-dimensional network, 47, 156 Spin-orbit scattering, 501
One-dimensional NLTN (1D NLTN), 5 Spin-polarized current, 395, 431, 434, 435,
Ordinary differential equations, 99, 104, 196 439, 448, 453, 457, 465, 467, 469,
471, 472, 474, 477
Spin-transfer torque, 216, 217, 447, 453–
P 455, 457, 459, 467, 485, 488, 499
Partial differential equations, 242, 291 Stereographic projection, 403, 431
Poincare Bendixson theorem, 504 Stokes wave, 56, 68, 79, 99
Polar solitons, 378, 379, 382, 384, 390 Super rogue wave, 137, 145, 146, 149, 156,
157

Q
Quasi-one-dimensional, 243, 290, 302 T
Transverse stability, 163
Two-dimensional network, 165, 172
R
Raman scattering, 22
Rogue waves, 4, 109, 113, 116, 118, 124, V
127, 145, 151, 152, 155, 157, 217, Vakhitov-Kolokolov criterion, 385
289, 292, 297, 302, 306, 309, 317, Vortex-soliton, 363, 365, 367
484, 486, 488
Runge–Kutta–Munthe–Kaas, 453
Runge–Kutta–Munthe–Kaas (RKMK) W
method, 453, 455 Wave propagation, 205

You might also like