We investigate ion acoustic solitary waves (solitons) of varying amplitudes in a one-dimensional plasma using fully kinetic particle-in-cell simulations. The initial soliton conditions are based on the Korteweg–de Vries (KdV) equation, treating ions as a cold species and electrons with finite temperature. Our findings reveal that KdV solitons evolve nonlinearly to a saturated state at higher amplitude, deviating from KdV predictions for ion density and electric potential, and from the Boltzmann relation for electron density. At this saturated state, the KdV model cannot accurately describe the soliton behavior. For small amplitudes, Sagdeev's model describes the saturated state, but not the soliton width; for larger amplitudes, it models the width accurately, but not the amplitude. These discrepancies arise from assuming a Boltzmann relation for electron density, while electron trapping creates non-Boltzmann densities—a deviation that increases with soliton amplitude. Additionally, we observe that the soliton amplitude oscillates roughly at the electron bounce frequency. The soliton is better described by Schamel's electron density formulation and a modified KdV equation incorporating electron trapping. The soliton velocity matches best with predictions from Sagdeev's and Schamel's models. Moreover, the soliton speed–amplitude relationship differs from existing theoretical predictions. Finally, we find minimal ion and electron Landau damping effects.

Nonlinear waves in a plasma emerge from the collective interplay between charged particles and the self-consistent electromagnetic field within the plasma. Some examples of these waves include Alfvén waves, Langmuir waves, and ion acoustic waves.1–3 The study of nonlinear wave phenomena in plasma constitutes a pivotal aspect of plasma physics, given its broad relevance in fields such as astrophysics and space physics. Ion acoustic waves exhibiting nonlinear behavior, particularly in the form of solitons, have garnered extensive research interest due to their ubiquity and significance across a broad spectrum of plasma applications.4–28 The predominant approach in these studies has involved analytical and numerical studies grounded in fluid model descriptions of the plasma environment. The reductive perturbation method for weak nonlinearities6 and the fully nonlinear Sagdeev approach5,29 stand out as the classic methodologies. Using the reductive perturbation method, Washimi and Taniuti demonstrated that, when considering small but finite amplitude perturbations in such plasmas, the evolution of ion acoustic waves can be described by the integrable Korteweg–de Vries (KdV) equation, which admits a distinct class of soliton solutions as exact traveling wave solutions.6,30

These soliton structures arise from a precise balance between the opposing effects of nonlinear steepening and dispersive broadening of the wave. Specifically, in the KdV equation, the quadratic nonlinear term accounts for wave steepening, while the third-order dispersion term contributes to pulse spreading. Solitons possess several unique properties stemming from the mathematical characteristics and infinite number of conservation laws intrinsic to the KdV equation.19 Notable properties include the ability to propagate over long distances without degradation of their shape or amplitude, as well as the remarkable phenomenon of preserving individual identities upon colliding with other solitons. Ikezi et al.12 first experimentally detected the formation and propagation of solitons in an argon plasma and since then solitons have been commonly observed in both laboratory and space plasma environments.31–34 Recent investigations into the advancements in inertial confinement fusion,35 the complementary heating method in fusion devices,36 and techniques for accelerating charged particles37 have demonstrated close ties to the dynamics of solitons. A broad spectrum of research endeavors has been undertaken concerning solitons, encompassing their formation,38 propagation,39,40 reflection,41 collisions,42 decay,43,44 wave instabilities,45,46 etc.

One emerging area of interest is the idea of using precursor solitons to detect sub-centimeter orbital debris. Current ground-based methods and on-orbit detectors cannot detect such small debris, which can still prove fatal to spacecraft. Sen et al. numerically solved the forced Korteweg–de Vries (fKdV) equation and demonstrated the existence of precursor solitons created by orbiting charged debris.47,48 Tiwari and co-workers presented similar results from solving the full set of fluid equations.49,50 Jaiswal et al. experimentally validated the existence of these waves.51 Truitt and Hartzell used a similar approach to Sen by numerically solving the fKdV equation for varying space regions. They showed that precursor waves with detectable amplitudes could propagate indefinitely without changing shape or size.52 

However, these results were generated using the reductive perturbation method. It is essential to understand whether deviating from the fundamental assumptions underlying the fKdV equation, such as the small amplitude perturbation assumption or the idealized fluid model, can significantly influence the characteristics, formation, propagation, and stability of the nonlinear solution. To gain insights into this, a first-principles study initializing KdV solitons in a one-dimensional (1D) plasma and analyzing their properties seems appropriate. Therefore, a fully kinetic approach with minimal assumptions, such as a particle-in-cell (PIC) simulation, can be employed to investigate this phenomenon. This research motivation underpins the present study.

The study of solitons using various numerical methods has been previously conducted by a number of authors. Nopoush and Abbasi53 employed hybrid particle-in-cell (PIC) simulations to explore the generation of ion acoustic solitary waves (IASWs). Dharodi et al.54 demonstrated the formation of precursor solitons using fully kinetic PIC simulations and their dependence on the source shape. Kakad et al.20 examined the formation of ion acoustic solitary wave chains using fluid simulations. Qi et al.55 utilized hybrid PIC simulations to investigate soliton collisions. Sharma et al.56 also used hybrid PIC simulations to study the propagation of solitons with varying amplitudes. They showed that larger amplitude solitons initially grow as they propagate, while smaller amplitude solitons propagate without significant changes from their initial state.56 

Furthermore, Qi et al. conducted another study using hybrid PIC simulations to demonstrate that as the soliton amplitude increases, reductive perturbation models fail to accurately predict its behavior.57 They showed that when the amplitude surpasses a certain threshold, the wave becomes unstable and undergoes decay into a series of smaller amplitude waves.57 

In this work, we aim to study the propagation of Korteweg–de Vries (KdV) solitons launched in a one-dimensional (1D) plasma according to the solution of the KdV solution for the ion density, velocity, and potential using a fully kinetic plasma model. This study complements the works of Sharma et al.56 and Qi et al.57 who both modeled the electrons as a Boltzmann fluid. The validity of such hybrid models has not been verified for these types of studies. Their simulations ignored trapping effects and treated electrons as an isothermal background; therefore, they are unable to capture the nonlinear features and the deviation it causes. Furthermore, Kakad et al.58 showed that when the phase velocity of the soliton is greater than the ion acoustic velocity, it is necessary to include kinetic effects (including electrons) in the model. This is the case for the solitons in the present work, underscoring the reason for using a fully kinetic model as opposed to a hybrid one.

Ott and Sudan59 were one of the pioneers in looking at kinetic effects, namely, Landau damping of solitons. They derived a Korteweg–de Vries equation with a source term to model the lowest-order effects of resonant electrons, demonstrating the occurrence of linear Landau damping in solitons. However, they neglected trapped particle effects, which Meiss and Morrison60 later showed to be potentially as significant as linear Landau damping terms. They predicted that nonlinear effects become important when the Landau time, γL1=[k2πmeTe/(8mi2)]1/2, exceeds the electron bounce time, ωbe1=[me/(ek2ϕ)]1/2. This condition is met in this study; therefore, the role of these nonlinear trapping effects needs to be looked at.

Despite these theoretical advances, the time-dependent electron Landau damping problem remains largely unexplored using numerical methods such as PIC simulations. This gap in the literature also motivates our study, which aims to investigate the physics of nonlinear electron trapping and its effects on soliton properties, including the relationship between amplitude and speed. To our knowledge, this work represents the first comprehensive analysis of long-time soliton propagation using fully kinetic PIC simulations.

This paper is organized as follows: Sec. II provides a detailed description of the particle-in-cell (PIC) simulation employed in this study, including the simulation setup, initial conditions, and relevant parameters. Section III presents and discusses the results of our investigation, which is divided into several subsections. Subsection III A examines the full set of cold fluid equations and their ability to describe the soliton dynamics. Subsection III B focuses on the effects of electron trapping and is further divided into three sub-subsections. Sub-subsection III B 1 explores the impact of non-isothermal electrons on the soliton profile. Sub-subsection III B 2 introduces the modified KdV equation, which incorporates the nonlinearity arising from electron trapping. Sub-subsection III B 3 investigates the electron bounce frequency and its role in the soliton amplitude oscillations. Subsection III C analyzes the relationship between various soliton parameters, such as amplitude, width, and speed, and compares the findings with existing theoretical predictions. Subsection III D examines the effects of Landau damping on the soliton propagation. Finally, Sec. IV summarizes our conclusions, highlights the key findings of this study, and outlines potential directions for future research in the field of ion acoustic solitons.

The simulations were carried out using the open source PIC code called WarpX.61,62 We chose to use WarpX's electrostatic solver due to its highly parallel and highly optimized nature, which allows it to run efficiently on GPUs. This was an attractive choice considering the large number of particles, the high resolution of both time and spatial scales, and the extended duration of the total run time.

A finite difference scheme is used to solve Poisson's equation. The electric potential, ϕ, and the charge density, ρ, are calculated at the nodes, while the electric field, E, is calculated at the cell centers. Both electrons and ions are treated as particles in our simulations. To accelerate ion dynamics, we use a fictitious electron-to-ion mass ratio (me/mi) of 0.01. All simulations employ periodic boundary conditions with a domain length of L=10 m =1184λD, where λD is the Debye length defined as λD=ε0kBTe/(n0qe2). Here, kB is the Boltzmann constant, ε0 is the vacuum permittivity, qe is the elementary charge, Te is the electron temperature, and n0 is the background plasma density. A longer domain length (L=20 m) is used in some simulations where long-time effects are analyzed. We set Te=1500 K and n0=1011 m−3, approximating typical Low Earth Orbit (LEO) space plasma conditions. Initially, the ion temperature, Ti, is set to 0, representing a cold ion species consistent with the Korteweg–de Vries (KdV) formulation, but ion temperature is allowed to evolve with the simulation. A time step, dt=0.01ωpe is used, where ωpe=n0qe2/(ε0me)/(2π). The spatial discretization is dx = 0.0024 m = 0.29  λD. This dx was chosen to resolve for the narrow solitons.

The initial conditions are prescribed according to the solution of the Korteweg–de Vries (KdV) equation. The KdV equation is written as6 
(1)
where U represents the normalized ion number density, normalized electric potential, and normalized ion fluid velocity. Ion number density, electric potential, and ion fluid velocity are normalized by n0, kbTe/qe, and Cs, respectively. Cs=kBTe/mi is the ion acoustic speed. τ and ξ represent stretched nondimensional time-like and space-like variables, respectively. They are defined as τ=tωpi, where ωpi=n0qe2/(ε0mi)/(2π) and ξ=x/λDMτ, where M is the wave Mach number. The KdV equation admits a solution in the form of a soliton as
(2)
where N represents the amplitude of the soliton. In this paper, only the soliton's initial amplitude is represented by the nondimensional variable N. For the purpose of this study, we set N to 0.1, 0.25, 0.5, and 0.8 to investigate the effects of varying initial amplitudes. For all the simulations, x0 = 2 m.
The initial conditions for ion number density, electric potential, and ion fluid velocity are found by multiplying U(x,0) by their respective normalization quantities to find the following:
(3)
(4)
(5)
In order for Eq. (4) to be satisfied, the initial electron density (ne,0) must satisfy Poisson's equation. Therefore,
(6)

In similar research studies, it is common practice to set the initial ion and electron densities equal to each other (ni,0=ne,0).56–58 This approach is acceptable because the PIC code iteratively adjusts the particle positions until the desired potential, given by Eq. (7), is achieved. However, by utilizing Eq. (9) to calculate the initial electron densities, the simulation can be optimized by starting with the target potential from the outset, thereby reducing the computational time required. Figure 1 shows an example of initial conditions of the normalized ion and electron number density, and electric potential.

FIG. 1.

Sample normalized initial conditions (N = 0.5).

FIG. 1.

Sample normalized initial conditions (N = 0.5).

Close modal

The initial electron velocities are sampled from a Maxwellian distribution with an electron temperature (Te) of 1500 K. The random motion of the electrons gives rise to thermal fluctuations, which manifest as numerical noise in the simulation. These fluctuations effectively act as numerical collisions between particles, leading to the numerical damping of the soliton. We conducted initial studies to investigate the sensitivity of this numerical damping to the number of particles per cell (ppc) used in the simulation. The results of these preliminary studies prompted the use of a high ppc value of 2.75×105 to mitigate the effects of numerical damping and minimize disturbances to the soliton structure, an approach similar to that employed by Kakad et al.58 A shape factor of the 3rd order was used to minimize the effect of noise. The simulations were run on NERSC's Perlmutter, utilizing 32 nodes with 4 GPUs per node.

The evolution of the initial solitons for different values of N is depicted in Fig. 2. In all cases, the soliton initially grows until it reaches a saturated state, as exemplified in Fig. 3 for N = 0.5. The particles in the simulation self-adjust until the soliton satisfies the full set of nonlinear equations, which is consistent with the results presented in previous studies.56,57 Figure 4 demonstrates an approximately linear relationship between the starting and final amplitudes, similar to the findings of Sharma et al.56 However, the change in amplitude is found to be higher than what is reported in their work. Furthermore, the growth rate of the soliton from its initial state to its final state depends on the initial amplitude, with smaller amplitude solitons (e.g., N = 0.1) growing to their saturated state much slower than larger amplitude solitons.

FIG. 2.

The propagation of the ion density (normalized) for different values of N. The solitons grow until reaching saturated states with varying amplitudes and saturation times: (a) N = 0.1, amplitude = 0.125, ωpit170; (b) N = 0.25, amplitude = 0.345, ωpit20; (c) N = 0.5, amplitude = 0.82, ωpit15; (d) N = 0.8, amplitude = 1.52, ωpit10. The color gradient from cold to warm (blue to red) indicates time progression in each subplot.

FIG. 2.

The propagation of the ion density (normalized) for different values of N. The solitons grow until reaching saturated states with varying amplitudes and saturation times: (a) N = 0.1, amplitude = 0.125, ωpit170; (b) N = 0.25, amplitude = 0.345, ωpit20; (c) N = 0.5, amplitude = 0.82, ωpit15; (d) N = 0.8, amplitude = 1.52, ωpit10. The color gradient from cold to warm (blue to red) indicates time progression in each subplot.

Close modal
FIG. 3.

Soliton amplitude for N = 0.5 vs time. This corresponds to the data shown in Fig. 2.

FIG. 3.

Soliton amplitude for N = 0.5 vs time. This corresponds to the data shown in Fig. 2.

Close modal
FIG. 4.

Linear relationship between soliton starting amplitude and saturated amplitude.

FIG. 4.

Linear relationship between soliton starting amplitude and saturated amplitude.

Close modal

To further investigate the soliton propagation, a 2D FFT of the electric field data was performed, revealing the presence of plasma waves and ion acoustic waves in addition to the soliton propagation. This observation is in agreement with the findings of Kakad et al.58 and Sharma et al.56  Figure 5 presents an example of the power spectrum with N = 0.5.

FIG. 5.

2D FFT of the electric field data for N = 0.5. The bright yellow band represents soliton propagation, with the black dotted line indicating the average soliton velocity (1.2  Cs). Note that the soliton velocity rapidly attains a saturated state, exceeding the initial velocity, and maintains this constant saturated velocity throughout the simulation. The pink dotted line denotes the electron plasma frequency (ωpe). The green and cyan dotted lines illustrate the analytical dispersion relations for ion acoustic waves and Langmuir waves, respectively.

FIG. 5.

2D FFT of the electric field data for N = 0.5. The bright yellow band represents soliton propagation, with the black dotted line indicating the average soliton velocity (1.2  Cs). Note that the soliton velocity rapidly attains a saturated state, exceeding the initial velocity, and maintains this constant saturated velocity throughout the simulation. The pink dotted line denotes the electron plasma frequency (ωpe). The green and cyan dotted lines illustrate the analytical dispersion relations for ion acoustic waves and Langmuir waves, respectively.

Close modal

In every simulation, it is observed that a tail is produced behind the soliton as predicted by Keener and McLaughlin,63 Karpman and Maslove,64 and Watanabe.65 This tail grows in size until it eventually separates from the main soliton body and propagates at a lower amplitude and speed compared to the soliton. An example of this is shown in Fig. 6. The separation happens quicker for higher amplitude solitons.

FIG. 6.

Example of tail forming behind soliton. The left panel shows the soliton when the tail is still attached and the right panel shows the soliton after the tail has been detached.

FIG. 6.

Example of tail forming behind soliton. The left panel shows the soliton when the tail is still attached and the right panel shows the soliton after the tail has been detached.

Close modal
The KdV equation provides a valuable small-amplitude approximation for ion acoustic waves in plasma. However, to comprehensively understand the behavior of these waves across a broader range of amplitudes, it is necessary to examine the full set of nonlinear fluid equations. Similar studies were performed. In this section, we employ the “classic potential method” developed by Sagdeev and Galeev5 to numerically solve the complete nonlinear system. We then present a comparative analysis of three solutions: the KdV approximation, the full nonlinear solution (hereafter referred to as Sagdeev's solution), and our numerical results. The evolution of finite-amplitude ion acoustic waves in a plasma can be described by a basic fluid model. This model accounts for ion dynamics while treating the lighter electron species as having a Boltzmann distribution. For a one-dimensional scenario, the model equations can be expressed as
(7)
(8)
(9)
In these equations, ni represents the ion fluid density, u denotes the ion fluid velocity, and ϕ is the electrostatic potential. Equation (7) is the ion continuity equation, Eq. (8) represents the ion momentum equation, and Eq. (9) is Poisson's equation. The normalizations are as previously described in Sec. II. Transforming Eqs. (7)–(9) into the moving soliton frame by employing the coordinates ξ and τ, we can get the following:
(10)
(11)
(12)
Substituting (10) and (11) into (12), we get
(13)
This can be integrated to get
(14)

Sagdeev's solution for ϕ is obtained by solving Eq. (14) using the fourth-order Runge–Kutta (RK4) method. Figure 7 compares the results of solving (14) with the KdV solution and PIC simulation data after reaching saturation. The amplitude and position of the soliton were set to match the PIC data. For small amplitudes, Sagdeev's model and the KdV model show close agreement with each other; however, neither model accurately captures the steepening of the soliton potential (i.e., the width). As the amplitude increases, Sagdeev's model accurately describes the electric potential while the KdV model still cannot capture the steepening. The steepening of the soliton is a function of its nonlinearity. This suggests that at smaller amplitudes, there are nonlinearities not captured by either the Sagdeev or KdV model. As the amplitude grows, the standard higher order fluid nonlinearity begins to dominate, and the unmodeled nonlinearity has a diminished impact, allowing Sagdeev's model to accurately capture the electric potential. In Sec. III B 1, we propose that this unmodeled nonlinearity arises from electron trapping effects. To maintain consistency with previous studies,56,57 we initialized our simulations using KdV solitons. However, we also discovered that initializing solitons using Sagdeev's equations resulted in the same saturated soliton state as the one obtained from the KdV solitons. For the sake of conciseness, these additional results are not presented in this work.

FIG. 7.

Comparison of the soliton's electric potential produced by the PIC simulation against the KdV model and Sagdeev's model. Here the amplitude was set to match the PIC soliton amplitude.

FIG. 7.

Comparison of the soliton's electric potential produced by the PIC simulation against the KdV model and Sagdeev's model. Here the amplitude was set to match the PIC soliton amplitude.

Close modal

The resulting ϕ can then be substituted into Eqs. (11) and (10) to calculate ni. Additionally, M can be solved based on the amplitude of the electric potential by using (25). Figure 8 presents the ion density plots, revealing that the PIC soliton has evolved from the KdV profile to a new state. Here, the amplitude of the ion density is not fixed to the PIC result but is calculated self-consistently based on ϕ and M. The KdV profile fails to capture the steepening and amplitude of this ion density, with the discrepancy increasing as the amplitude grows. At smaller amplitudes, Sagdeev's model accurately captures the amplitude, but not the width, consistent with the results in Fig. 7. However, as the amplitude increases, the opposite effect is observed—Sagdeev's model captures the steepening well, but not the amplitude. As previously stated, this suggests the presence of an unmodeled nonlinearity that affects the steepening at smaller amplitudes but has a diminishing effect on the soliton width at higher amplitudes. In terms of the height, this unmodeled nonlinearity appears to have an increasing effect on the ion density amplitude at higher soliton amplitudes. The ion density given by the PIC simulation is lower than what is expected in Sagdeev's model. The lower ion density is attributed to a non-Boltzmann distribution of electrons at a lower electron density, leading to a reduction in ion density to compensate for this effect. Section III B 1 on electron trapping provides more details on this phenomenon.

FIG. 8.

Comparison of the soliton's ion density produced by the PIC simulation against the KdV model and Sagdeev's model. Here the amplitude was calculated based on the electric potential.

FIG. 8.

Comparison of the soliton's ion density produced by the PIC simulation against the KdV model and Sagdeev's model. Here the amplitude was calculated based on the electric potential.

Close modal

1. Non-isothermal electrons

In deriving Eq. (14), an isothermal electron equation of state (i.e., the Boltzmann relation) was assumed. This assumption aligns with the hybrid PIC simulations carried out by Sharma et al.56 and Qi et al.,57 where electrons are treated as a Boltzmann fluid. However, Schamel66 demonstrated that electron trapping within the soliton potential well leads to deviations from the Boltzmann relation for the electron density. This deviation introduces a new nonlinear term that is not accounted for in Sagdeev's method. Trapped particles are those that resonate with and move along with the potential well of the soliton. In phase space, these particles manifest as either a hump or a hole. Our simulations clearly exhibit electron trapping. Figure 9 illustrates the propagation of electron vortices in the x–v (phase) space, with each snapshot corresponding to a different time stage in the simulation for N = 0.5. As the soliton amplitude increases, the potential well deepens, enabling the trapping of more electrons. This trapping phenomenon creates a nonlinearity that causes the electron density to deviate from the Boltzmann relation. Figure 10 shows this deviation where the observed electron density is lower than the value predicted by the Boltzmann relation using the computed ϕ.

FIG. 9.

Distribution of electrons in xv (phase) space centered around the soliton. Here we see electrons exhibiting particle trapping. The y axis is the electron velocity normalized by the ion acoustic velocity. The colorbar is kept constant and represents the fraction of electrons in the entire simulation.

FIG. 9.

Distribution of electrons in xv (phase) space centered around the soliton. Here we see electrons exhibiting particle trapping. The y axis is the electron velocity normalized by the ion acoustic velocity. The colorbar is kept constant and represents the fraction of electrons in the entire simulation.

Close modal
FIG. 10.

Comparison of the soliton's electron density from Schamel's density model against the PIC results. Here β was iteratively picked until the error between the models was at a minimum.

FIG. 10.

Comparison of the soliton's electron density from Schamel's density model against the PIC results. Here β was iteratively picked until the error between the models was at a minimum.

Close modal

Electron trapping effects manifest through the electron density. By forgoing the assumption of isothermal electrons and instead utilizing the numerical electron density, we can capture the nonlinearity that is not represented in Fig. 7. Figure 11 illustrates this result, where a modified version of Eq. (14), incorporating a general electron density, is solved using the electron density provided by the PIC simulation. This approach captures the width of the electric potential more accurately than in Fig. 7. The improved accuracy is due to the inclusion of the nonlinearity arising from electron trapping in the numerical electron density, which effectively models the steepening of the soliton. Solving Poisson's equation with this electron density and potential should also yield the correct ion density with the appropriate amplitude and width. However, it is important to note that this approach relies on the numerical result for electron density from the PIC simulation and does not employ an analytical model.

FIG. 11.

Comparison of the soliton's electric potential produced by the PIC simulation against the model with numerical electron density. Here the amplitude was set to match the PIC soliton amplitude.

FIG. 11.

Comparison of the soliton's electric potential produced by the PIC simulation against the model with numerical electron density. Here the amplitude was set to match the PIC soliton amplitude.

Close modal
Schamel66 developed an analytical approach by constructing a distribution function for trapped and free electrons. The distribution function is designed to be continuous at the transition point between free and trapped electrons, and the slope of the trapped electron distribution function in the velocity space is required to be finite. Using this distribution function, Schamel derived a general electron density function given by
(15)
where W(x) represents the Dawson integral. The parameter β is a trapping parameter that describes the distribution function around the soliton speed67 and represents the nonlinearity created by the trapped electrons. Setting β=1 results in the standard isothermal formulation. In the regime where β1, the trapping nonlinearity is negligible. As β decreases toward negative values, the nonlinearity increases. Figure 10 presents the results of using Eq. (15). We found that there exists a value of β that can closely model the electron density inside the soliton. The β values shown in Fig. 10 reveal that β decreases as the soliton amplitude increases, but not proportionally. This observation is consistent with the results presented in Figs. 7 and 8. For smaller amplitudes, the nonlinearity arising from electron trapping, although small, is nontrivial compared to the standard nonlinearity. As a result, there is a mismatch in the width of the soliton. In contrast, for larger amplitudes, the nonlinearity increases but remains small compared to the standard nonlinearity, thus having a negligible effect on the soliton width. However, as previously observed, the deviation from a Boltzmann relation increases with increasing amplitude, causing a mismatch in the soliton density amplitudes. It is important to note that the value of β was determined iteratively to minimize the error between the analytical model and the numerical result. This iterative approach allows for a more accurate representation of the electron density profile within the soliton. The parameter β introduces a modification to the Maxwellian distribution for trapped electrons, allowing for different temperatures. A flat-topped electron distribution is obtained when β=0, while negative values of β indicate a depleted population of trapped electrons, represented by a hole in the trapped region.66 Deviation from the Boltzmann relation, which assumes an isothermal electron equation and represents a profile of “maximum electron trapping,” is a consequence of the underpopulation of trapped electrons. As β decreases from 1 toward negative values, the discrepancy between the electron density profile and the Boltzmann relation increases, signifying the growing dominance of nonlinearity due to electron trapping. By adjusting the value of β, the model can better capture the deviation from the standard KdV solution and more accurately describe the electron density distribution in the presence of electron trapping. However, it is crucial to acknowledge that the determination of the optimal β value through an iterative process relies on the availability of numerical results for comparison. This highlights the importance of combining analytical models with numerical simulations to gain a more comprehensive understanding of the soliton dynamics and the role of electron trapping in shaping the electron density profile. Further research could explore the development of more sophisticated methods for determining the trapping parameter β, such as deriving a functional relationship between β and the soliton parameters or other relevant plasma parameters. This would enable a more predictive approach to modeling the electron density in the presence of trapped electrons, without the need for extensive numerical simulations.

2. The modified KdV equation

Similar to how the KdV formulation is a small amplitude approximation of Sagdeev's model, a modified KdV equation that takes the small amplitude limit of the full set of cold fluid equations but uses the electron density given in Eq. (15) instead of the Boltzmann relation can be derived. Schamel66 performed this derivation and obtained the following modified KdV equation:
(16)
Equation (16) is called the Schamel-KdV equation and its general solution is
(17)
where
(18)
(19)
(20)

In Eq. (17), when |b|O(N), the strength of the trapping nonlinearity is on the same order as the usual nonlinearity. Furthermore, the nonlinearity due to trapping becomes dominant when |b|O(N).67 Note that setting β=1 would simply yield the standard KdV solution. Figures 12 and 13 present the comparisons of using Eq. (17). These figures demonstrate that the sech4 solution given by the modified KdV equation provides a better fit for the data compared to the sech2 formulation of the KdV solution shown in Figs. 7 and 8. In Figs. 12 and 13, the soliton amplitudes were set to match the PIC result to directly compare the shape of the profile even though, strictly speaking, for the small amplitude limit, the amplitudes for the electric potential and ion densities are given by the same value. For smaller amplitudes, this model is able to accurately capture the soliton profile, while the error in the width increases slightly for larger amplitudes. This behavior is expected since the model is based on a small amplitude approximation. It is important to note that to achieve the best fit, the β values used (shown in Table I) were different from those presented in Fig. 10. This implies that the “true” β value does not produce the “best” solution. This discrepancy suggests that while the sech4 function is the best approximation to the soliton, the underlying partial differential equation (PDE) that describes it may be imprecise or incomplete. This could also be a limitation of using a KdV-type approximation, but it is interesting to note that the model still captures the shape of the soliton quite well. The determination of consistent β values to describe all the soliton parameters is left for future work.

FIG. 12.

Comparison of the soliton's electric potential produced by the PIC simulation against the solution of the modified KdV equation (Schamel's solution). Here the amplitude was set to match the PIC soliton amplitude.

FIG. 12.

Comparison of the soliton's electric potential produced by the PIC simulation against the solution of the modified KdV equation (Schamel's solution). Here the amplitude was set to match the PIC soliton amplitude.

Close modal
FIG. 13.

Comparison of the soliton's ion density produced by the PIC simulation against the solution of the modified KdV equation (Schamel's solution). Here the amplitude was set to match the PIC soliton amplitude.

FIG. 13.

Comparison of the soliton's ion density produced by the PIC simulation against the solution of the modified KdV equation (Schamel's solution). Here the amplitude was set to match the PIC soliton amplitude.

Close modal
TABLE I.

Relationship between starting amplitude, final ion density, final electric potential, and β values.

Starting Final ion Final electric β values
Amplitude Density Potential Ion density Electric
(N) (n0/n0) (eϕ/kBTe) Profile Potential profile
0.1  0.125  0.11  0.1  0.3 
0.25  0.345  0.27  −0.3  0.0 
0.5  0.82  0.54  −2.0  −0.5 
0.8  1.52  0.82  −4.0  −1.2 
Starting Final ion Final electric β values
Amplitude Density Potential Ion density Electric
(N) (n0/n0) (eϕ/kBTe) Profile Potential profile
0.1  0.125  0.11  0.1  0.3 
0.25  0.345  0.27  −0.3  0.0 
0.5  0.82  0.54  −2.0  −0.5 
0.8  1.52  0.82  −4.0  −1.2 

The KdV formulation increasingly deviates from the saturated soliton as the amplitude rises. While it is known that the KdV formulation begins to break down at higher amplitudes, the deviation presented in this paper is more significant than previously reported.56–58 From an engineering perspective focused on designing sensors for soliton detection,47,52,68 this deviation might seem relatively minor. However, in our effort to understand the fundamental physics of soliton propagation, stability, and damping, we propose that the nonlinearity due to electron trapping must be accounted for in the model. In summary, the modified KdV equation, which incorporates the effects of electron trapping through the sech4 solution, provides a better approximation to the soliton profile compared to the standard KdV equation. However, the discrepancy in the β values used to obtain the best fit suggests that further refinements to the underlying PDE may be necessary to fully capture the soliton dynamics in the presence of electron trapping.

3. The electron bounce frequency

We saw electron trapping in the phase space shown in Fig. 9. As the soliton well reaches a saturated state, the regions of lower electron density are filled, and the phase space becomes mixed, remaining invariant for the remainder of the simulation. This phase space smearing effect at saturation was predicted by Meiss and Morrison,60 who also anticipated that the soliton amplitude would oscillate at the electron bounce frequency (also known as the trapping frequency) ωbe=[(ek2ϕ)/me]1/2. Here, we are using k=2π/solitonwidth. To verify this prediction, we performed an FFT analysis on the soliton amplitude data presented in Fig. 3. The resulting frequency spectrum, shown in Fig. 14, reveals that the soliton amplitude indeed oscillates at a frequency close to the electron bounce frequency, ωbe. This finding corroborates the theoretical predictions made by Meiss and Morrison, providing further insights into the dynamics of electron trapping and soliton propagation in the phase space.

FIG. 14.

FFT of soliton amplitude for N = 0.5. The dominant frequencies seen here are for ω roughly equal to ωbe.

FIG. 14.

FFT of soliton amplitude for N = 0.5. The dominant frequencies seen here are for ω roughly equal to ωbe.

Close modal

The KdV equation predicts that normalized potential, ion density, and velocity should be equivalent in soliton solutions. However, our study shows results that deviates from this expectation for larger amplitudes. Table I presents the quantitative relationships between soliton starting amplitude, final ion density, final electric potential, and corresponding β values. Here, we see discrepancies in the growth patterns of ion density and electric potential. While ion density exhibited substantial increases (but not as much as predicted by Sagdeev), the electric potential remained relatively constant, approximating its initial amplitude. This suggests a weaker coupling between these parameters than previously theorized and provides further evidence for the breakdown of the conventional KdV relationship under the conditions examined. Furthermore, the electric potential profile, like the ion density profile, demonstrates a better fit to a Schamel soliton model compared to the KdV soliton. One interesting thing to note here is the divergence in β values that optimally fit the electric potential and ion density data for a given amplitude N. Table I delineates the distinct β values for ion density and electric potential profiles.

The soliton speed remains relatively constant throughout the simulation, as evident in Fig. 5, where the slope of the bright yellow band (or the black dotted line) represents the soliton's velocity. Interestingly, the soliton speed reaches a saturated state much faster than the ion density and electric potential, almost immediately after the simulation commences. To compare the observed soliton velocity against theoretical predictions, we can use the quantities presented in Table I and perform three comparisons with (1) KdV, (2) Schamel, and (3) Sagdeev.

According to the KdV model, the Mach number of the soliton (Vsoliton/Cs) is given by
(21)
where ψ represents the soliton amplitude. Theoretically, ψ should be the same for both ion density and electric potential, but as previously stated, this is not the case. Consequently, we calculate two different MKdV values for comparison — one using the final ion density and another using the final electric potential.
Schamel66 derived a Mach number equation (22) for the full set of nonlinear equations with the non-Boltzmann electron density and from the modified KdV equation (16), which takes into account the modified soliton inertia due to electron trapping:
(22)
where
(23)
and ne is as defined in (15),
(24)
The parameter b in Eq. (24) encapsulates the magnitude of electron trapping. However, as seen in Table I, the β values differ for the ion density and electric potential profiles. To calculate MSchamel,mKdv, we have assumed that the β corresponding to the electric potential is the “true” value. For MSchamel,full, we take the values presented in Fig. 10.
Finally, Sagdeev and Galeev5 derived an expression for the propagation of high-amplitude waves as a function of the peak electric potential, given by
(25)
This equation relates the Mach number (MSagdeev) to the ion acoustic speed (Cs), electron temperature (Te), ion mass (mi), and the peak electric potential (ψ). By employing these three theoretical models and utilizing the relevant quantities from Table I, we can compare the observed soliton velocity from the simulation with the predicted values, providing valuable insights into the accuracy and applicability of each model in describing the soliton's propagation characteristics.

The comparison between the observed soliton speed from the PIC simulation and the theoretical predictions is presented in Fig. 15. For small-amplitude solitons, the KdV, the full Schamel, and Sagdeev models show relatively good agreement with the PIC soliton speed. However, as the soliton amplitude increases, this agreement diminishes. Among the three models, the Sagdeev model and the full Schamel model appears to provide the closest match to the PIC soliton speed and the PIC soliton always remains to lay in between these two models. As the amplitude increases, the trend seems to suggest that the soliton speed is better captured by the full Schamel model but further work is needed to verify this. Interestingly, when using the ion density to calculate the soliton speed, the theoretical models tend to overshoot the observed speed, while using the electric potential leads to an underestimation. Consequently, the PIC soliton speed is sandwiched between these two estimates in all cases. It is noteworthy that the mKdV Schamel formulation predicts a significantly higher speed than what is observed in the PIC simulation. This is particularly intriguing because the Schamel soliton provides a good fit to the saturated soliton shape, yet it does not accurately model the soliton's speed. This discrepancy is probably due to using inconsistent β values and provides further evidence that the “best” fit β is not the “true” β. This mismatch could also suggest that the inertia created due to electron trapping may be greater than what has been theorized by Schamel,66 resulting in a slower soliton speed. The theoretical models predict that the soliton speed increases with increasing soliton amplitude, and this trend is indeed observed in the simulation results. However, the dynamic evolution of the soliton speed as a function of the soliton amplitude is not captured in these findings. The results presented here focus on the final, saturated state of the soliton and do not provide insights into the transient behavior of the soliton speed during the growth phase. The soliton speed reaches saturation almost immediately after the simulation starts and therefore, to fully understand the relationship between soliton amplitude and speed, further investigation into the temporal dynamics of the soliton propagation is necessary. This would require a detailed analysis of the soliton speed at different stages of its evolution at much smaller timescales, from the initial formation to the final saturated state.

FIG. 15.

Comparison of the soliton Mach number (M=Vsoliton/Cs) obtained from the PIC simulation with theoretical predictions from the KdV, Schamel, and Sagdeev models for different values of the initial soliton amplitude (N). The KdV model is evaluated using both the maximum ion density (ni,max) and the maximum electric potential (ϕmax). The Schamel model is evaluated using both the full set of nonlinear equations and using the modified KdV equation with β values from the electric potential.

FIG. 15.

Comparison of the soliton Mach number (M=Vsoliton/Cs) obtained from the PIC simulation with theoretical predictions from the KdV, Schamel, and Sagdeev models for different values of the initial soliton amplitude (N). The KdV model is evaluated using both the maximum ion density (ni,max) and the maximum electric potential (ϕmax). The Schamel model is evaluated using both the full set of nonlinear equations and using the modified KdV equation with β values from the electric potential.

Close modal

The investigation of soliton damping is of particular significance to the small debris research community. The concept of utilizing solitons for debris detection47,52 relies heavily on whether the soliton persists long enough to be detected or if it damps away prematurely. Moreover, the damping rate can serve as an additional data point to help uniquely identify the characteristics of the debris, such as its size, material composition, charge, and velocity. Consequently, this section will concisely explore the effects of soliton damping.

Ion Landau damping can occur when a wave has a slow phase velocity that matches the thermal velocity of ions. The ion acoustic wave, in particular, is significantly influenced by Landau damping. If the electron temperature (Te) is less than or equal to the ion temperature (Ti), the phase velocity lies in the region where the velocity distribution function has a negative slope. Consequently, ion Landau damping occurs when TeTi. However, this condition is not met in our simulations. Table II presents the final ion temperature at the end of each simulation for various starting amplitudes. The ion temperature exhibits a slight increase from 0, with larger amplitudes resulting in a more pronounced increase. Nevertheless, the ion temperature remains substantially lower than the electron temperature. Thus, under the conditions investigated in our study, we conclude that ion Landau damping does not play a significant role.

TABLE II.

Relationship between starting amplitude, final ion temperature, and electron temperature. The temperatures are in units of Kelvin (K).

Starting amplitude (N) Final ion temperature (K) Electron temperature (K)
0.1  0.064  1500 
0.25  0.55  1500 
0.5  1.74  1500 
0.8  3.9  1500 
Starting amplitude (N) Final ion temperature (K) Electron temperature (K)
0.1  0.064  1500 
0.25  0.55  1500 
0.5  1.74  1500 
0.8  3.9  1500 

Ions are not the only possible resonant particles. Ott and Sudan59 were the first to study linear electron Landau damping of ion-acoustic solitons. They derived a modified Korteweg–de Vries (KdV) equation with a source term that models the lowest-order effects of resonant electrons. Their equation includes the lowest-order nonlinear terms and terms corresponding to the linear dispersion relation for ion-acoustic waves. The derivation is justified by a formal procedure in which these terms in the equation are of the same order. Ott and Sudan obtained a solution to this modified KdV equation, which predicts an amplitude that decays to zero over time. However, they neglected trapped-particle effects, which are of the same order as the linear Landau-damping terms included in their modified KdV equation.

To address this gap, Meiss and Morrion60 derived a new time-dependent solution that incorporates effects associated with electron orbits from electron trapping. Their work predicts that nonlinear effects from trapped particles eventually balance out the damping from untrapped particles, leading to saturation at a finite amplitude. This is consistent with our observed results. They predicted that N ∼  (me/mi)2 is an effective threshold for the effect of electron linear Landau damping, which is met for all cases in this study. Above this threshold, the effects from electron Landau damping are negligible. They also showed that the soliton amplitude oscillates and decreases from its initial value before reaching a saturated amplitude, with this effect becoming more pronounced as γL/ωbe increases. In our simulations thus far, this ratio has been approximately zero, and therefore, the dip and oscillation of the amplitude are not observed.

To isolate this effect, we performed another simulation with me/mi=N=0.1, resulting in γL/ωbe=0.18, where the nonlinear electron Landau damping should be more visible. The result of this simulation is shown in Fig. 16 and is compared against the analytical solution given by Meiss and Morrison.60 In this figure, ψ(t) represents the soliton ion density amplitude as a function of time. Since the ion mass is only ten times heavier than the electron mass, the noise from random electron fluctuations is amplified by the random ion motion. Therefore, the data were averaged to reduce noise. The final saturation amplitude differs from the theoretical prediction, but the overall behavior of the soliton amplitude, including the dip and oscillation, seems to be captured. It is also interesting to note that the amplitude experiences a sudden dip right after the simulation starts and then climbs back up before slowly dipping again. We believe this is a result of the particles in the simulation quickly readjusting themselves after the start of the simulation to fit the prescribed fields. An explanation of the differing final amplitudes will be left for future work.

FIG. 16.

Comparison of soliton amplitude vs time for γL/ωbe=0.18 with the analytical solution. Here, ψ(t) represents the ion density amplitude as a function of time.

FIG. 16.

Comparison of soliton amplitude vs time for γL/ωbe=0.18 with the analytical solution. Here, ψ(t) represents the ion density amplitude as a function of time.

Close modal

As previously mentioned, for the data presented in Fig. 2, electron Landau damping effects are not observed. However, we wanted to investigate whether any effects could be detected if the simulation was run for a significantly longer time. As a test case, we ran the simulation for N = 0.5 for γLt>5 (ωpit>325). Figure 17 shows a small damping of the amplitude on this time scale. We believe that this damping is due to numerical noise acting as collisions, which can damp the wave.

FIG. 17.

Soliton amplitude of the ion density for N = 0.5 vs time. For this longer time simulation, a slight damping of the soliton amplitude is observed.

FIG. 17.

Soliton amplitude of the ion density for N = 0.5 vs time. For this longer time simulation, a slight damping of the soliton amplitude is observed.

Close modal

In this study, we found that the soliton's behavior is quite sensitive to simulation parameters such as the number of particles, cell size, time step, domain length, etc. Therefore, these parameters were carefully chosen through multiple iterations until convergence was achieved. However, there is always the possibility that some of these results may be influenced by numerical effects, and the authors acknowledge this possibility.

This comprehensive study, employing fully kinetic particle-in-cell (PIC) simulations, provides valuable insights into the behavior of ion acoustic solitary waves (solitons). The key findings reveal that solitons initialized according to the Korteweg–de Vries (KdV) solution undergo nonlinear evolution, reaching a saturated state at a higher amplitude that deviates from the KdV soliton for ion density and electric potential, as well as from the Boltzmann relation for electron density. This deviation intensifies with increasing soliton amplitude, accompanied by stronger nonlinear effects arising from electron trapping. The KdV model fails to capture the width of the soliton for all amplitudes and cannot model the different valued amplitudes for the ion density and electric potential. Solving the full set of nonlinear fluid equations (Sagdeev's model) allows for the approximation of the electric potential and ion density for small amplitudes, albeit with a slightly inaccurate width. As the amplitude grows, Sagdeev's model accurately captures the electric potential and the width of the ion density, but not its amplitude. This suggests that both the KdV model and Sagdeev's model fail to capture a nonlinearity that causes these discrepancies. We demonstrated that these discrepancies primarily arise from incorrectly assuming isothermal electrons with a Boltzmann description for the electron density. The deviation from the Boltzmann relation increases with increasing soliton amplitude and is caused by electron trapping effects. Electron trapping inside the soliton was observed through the phase space, and it was shown that this trapping induces an oscillation of the soliton amplitude at the electron bounce frequency. In correcting the electron density by using the numerical density in solving the full set of nonlinear fluid equations, we were able to model the soliton more accurately. Schamel derived an equation for the electron density under the influence of electron trapping, which includes a parameter β to describe the additional nonlinearity from trapping. We found that there exists a β value such that the electron density closely resembles the electron density from the PIC simulation. This β decreases away from 1 as the soliton amplitude increases, suggesting increasing nonlinearity. Furthermore, Schamel derived a modified KdV (mKdV) equation, which is a small amplitude approximation of Sagdeev's model with his electron density model instead of a Boltzmann description. We discovered that there exists a β value that enables the mKdV model to capture the soliton shape much better than the KdV model. This approximation works well for all amplitudes considered in this study, although it performs best for smaller amplitudes, as expected from a small amplitude approximation. However, like the KdV model, the mKdV model also assumes the same amplitude for electric potential and ion density, thus, failing to capture the saturated state. Interestingly, the higher amplitude ion density can still be approximated by the mKdV model by using a different β value, suggesting that the final state of the soliton is best approximated by an analytical form of sech4.

The soliton velocity was compared against predictions from the KdV, Schamel, and Sagdeev models. For small amplitudes, all models agree relatively well with the PIC soliton speed, but the agreement diminishes as amplitude increases. The Sagdeev and the full Schamel models provide the closest match, with the PIC speed lying between them. The mKdV Schamel formulation predicts significantly higher speeds than observed, despite the Schamel soliton fitting the saturated shape well. This discrepancy likely stems from using inconsistent β values. Although the models predict increasing speed with amplitude, which is observed in the simulations, the dynamic evolution of the speed is not captured in these findings, as they focus on the saturated state. Further investigation into the temporal dynamics at smaller timescales is necessary to fully understand the amplitude-speed relationship.

While the differences between the models mentioned above might seem relatively minor from an engineering perspective, we believe that these discrepancies highlight the limitations of widely used models, such as the KdV model and Sagdeev's model, in capturing the fundamental physics of soliton formation, stability, and propagation. The primary reason for these limitations is that these models do not account for the nonlinearity arising from electron trapping. We conclude that Schamel's model, which incorporates the full set of nonlinear fluid equations, is a more suitable alternative to Sagdeev's model, and the modified-KdV (mKdV) equation is a better alternative to the standard KdV model. However, even Schamel's model has its limitations. These include the inability to accurately predict the β parameter and the inconsistencies associated with requiring different β values for modeling the electric potential and ion density profiles. Furthermore, the mKdV model fails to accurately capture the differing amplitudes between the electric potential and ion density, as well as the soliton speed. Additionally, Schamel's model does not include other potential nonlinearities, such as ion trapping, ion reflection, and radiation of the soliton. In light of these findings, we recommend that any future model development for soliton propagation should take electron trapping effects into account. By incorporating these nonlinear effects, more accurate and comprehensive models can be developed to better understand the complex dynamics of solitons in plasma environments.

Our study also examines Landau damping effects and concludes that both ion and electron Landau damping appear to be minimal under the simulated conditions. By using a smaller ion mass, we were able to better observe the effects of electron Landau damping as predicted by Meiss and Morrison. While we observed saturation at a smaller amplitude than the initial value and an oscillation of the amplitude, the final predicted amplitude did not match our simulation. Nonetheless, the possibility of numerical effects influencing the results is acknowledged, emphasizing the importance of careful parameter selection to mitigate numerical effects. The conditions used in these simulations were not conducive to Landau damping. Although we used realistic electron temperatures that mimic LEO (1500 K) conditions, using a cold ion species is not representative of a general plasma. Similarly, using a fictitious ion mass could also create artificial results. We anticipate that the use of realistic mass ratios in our simulations will lead to several effects: an increase in electron trapping, a decrease in electron Landau damping, and the bounce frequency remaining the same. Exploring the effects of Landau damping in more depth for a general plasma with realistic temperature ratios will be the focus of our next study.

In conclusion, this work provides a comprehensive, first-principles investigation into the propagation of KdV solitons using fully kinetic PIC simulations. The findings highlight the importance of nonlinear electron trapping effects in soliton dynamics and the limitations of conventional models such as KdV, Sagdeev's model and hybrid PIC simulations in capturing these phenomena. The insights gained from this study contribute to a deeper understanding of soliton behavior in plasma environments and lay the groundwork for further research into the fundamental physics of soliton propagation, shape, growth, stability, and damping.

The authors would like to express their sincere gratitude to Dr. Remi Lehe for his invaluable support in running simulations on WarpX and for providing insightful guidance during the data analysis process. His expertise and assistance have been instrumental in the successful completion of this study.

We would also like to extend our heartfelt appreciation to Dr. Sarvershar Sharma, Dr. Sudip Sengupta, Dr. Gurudas Ganguli, and Dr. Abhijit Sen for their thought-provoking discussions and invaluable insights. Their contributions have significantly enriched this work, and their knowledge and expertise have been essential in shaping the direction and quality of our research.

This work was supported by MIT Lincoln Laboratory in collaboration with IARPA's SINTRA program.

This research was supported in part by the Knight-Hennessy Fellowship program at Stanford University. This research was supported in part by the Exascale Computing Project (No. 17-SC-20-SC), a collaborative effort of the U.S. Department of Energy (DOE) Office of Science and the National Nuclear Security Administration; the U.S. Department of Energy, Office of Science, Office of Advanced Scientific Computing Research, Exascale Computing Project under Contract No. DE-AC02-05CH11231.

This research used resources of the National Energy Research Scientific Computing Center, a DOE Office of Science User Facility using NERSC Award No. NERSC DDR-ERCAP0030179.

The authors have no conflicts to disclose.

Ashwyn Sam: Conceptualization (lead); Data curation (lead); Formal analysis (lead); Investigation (lead); Methodology (lead); Software (lead); Validation (lead); Visualization (lead); Writing – original draft (lead). Prabhat Kumar: Data curation (supporting); Methodology (supporting); Software (supporting); Visualization (supporting). Alex C. Fletcher: Conceptualization (equal); Formal analysis (supporting); Investigation (supporting); Methodology (supporting); Supervision (supporting); Validation (supporting); Visualization (supporting); Writing – review & editing (equal). Chris Crabtree: Conceptualization (equal); Formal analysis (supporting); Investigation (supporting); Methodology (supporting); Supervision (equal); Validation (supporting); Writing – review & editing (equal). Nicolas Lee: Funding acquisition (equal); Project administration (supporting); Writing – review & editing (supporting). Sigrid Elschot: Conceptualization (supporting); Funding acquisition (lead); Project administration (lead); Supervision (lead); Writing – review & editing (equal).

The data that support the findings of this study are available from the corresponding author upon reasonable request.

1.
B.
Buti
, “
Ion acoustic waves in a collisional plasma
,”
Phys. Rev.
165
,
195
201
(
1968
).
2.
S.
Bardwell
and
M. V.
Goldman
, “
Three-dimensional Langmuir wave instabilities in type III solar radio bursts
,”
Astrophys. J.
209
,
912
(
1976
).
3.
L.
Chen
and
F.
Zonca
, “
Physics of Alfvén waves and energetic particles in burning plasmas
,”
Rev. Mod. Phys.
88
,
015008
(
2016
).
4.
C. F.
Kennel
and
R. Z.
Sagdeev
, “
Collisionless shock waves in high plasmas: 1
,”
J. Geophys. Res.
72
,
3303
3326
, https://doi.org/10.1029/JZ072i013p03303 (
1967
).
5.
R. Z.
Sagdeev
and
A. A.
Galeev
,
Nonlinear Plasma Theory
(
Benjamin
,
1969
).
6.
H.
Washimi
and
T.
Taniuti
, “
Propagation of ion-acoustic solitary waves of small amplitude
,”
Phys. Rev. Lett.
17
,
996
998
(
1966
).
7.
T.
Taniuti
and
C.-C.
Wei
, “
Reductive perturbation method in nonlinear wave propagation. I
,”
J. Phys. Soc. Jpn.
24
,
941
946
(
1968
).
8.
N. J.
Zabusky
and
M. D.
Kruskal
, “
Interaction of ‘solitons’ in a collisionless plasma and the recurrence of initial states
,”
Phys. Rev. Lett.
15
,
240
243
(
1965
).
9.
N. J.
Zabusky
, “
Solitons and bound states of the time-independent Schrödinger equation
,”
Phys. Rev.
168
,
124
128
(
1968
).
10.
V. I.
Karpman
, “
An asymptotic solution of the Korteweg-De Vries equation
,”
Phys. Lett. A
25
,
708
709
(
1967
).
11.
C. S.
Gardner
,
J. M.
Greene
,
M. D.
Kruskal
, and
R. M.
Miura
, “
Method for solving the Korteweg-de Vries equation
,”
Phys. Rev. Lett.
19
,
1095
1097
(
1967
).
12.
H.
Ikezi
,
R. J.
Taylor
, and
D. R.
Baker
, “
Formation and interaction of ion-acoustic solitons
,”
Phys. Rev. Lett.
25
,
11
14
(
1970
).
13.
T.
Nagasawa
and
Y.
Nishida
, “
Experiments on the ion-acoustic cylindrical solitons
,”
Plasma Phys.
23
,
575
595
(
1981
).
14.
T. E.
Sheridan
,
S.
Yi
, and
K. E.
Lonngren
, “
On the origin of the ion acoustic soliton
,”
Phys. Plasmas
5
,
3165
3170
(
1998
).
15.
M.
Widner
,
I.
Alexeff
,
W. D.
Jones
, and
K. E.
Lonngren
, “
Ion acoustic wave excitation and ion sheath evolution
,”
Phys. Fluids
13
,
2532
2540
(
1970
).
16.
D. A.
Kurtze
and
D. C.
Hong
, “
Traffic jams, granular flow, and soliton selection
,”
Phys. Rev. E
52
,
218
221
(
1995
).
17.
Y.
El-Zein
,
T. E.
Sheridan
,
K. E.
Lonngren
, and
W.
Horton
, “
Excitation of ion acoustic solitons from grids
,”
J. Plasma Phys.
61
,
161
168
(
1999
).
18.
K. E.
Lonngren
, “
Ion acoustic soliton experiments in a plasma
,”
Optical Quantum Electron.
30
,
615
630
(
1998
).
19.
R.
Davidson
,
Methods in Nonlinear Plasma Theory
(
Elsevier
,
2012
).
20.
A.
Kakad
,
Y.
Omura
, and
B.
Kakad
, “
Experimental evidence of ion acoustic soliton chain formation and validation of nonlinear fluid theory
,”
Phys. Plasmas
20
,
062103
(
2013
).
21.
K. E.
Strecker
,
G. B.
Partridge
,
A. G.
Truscott
, and
R. G.
Hulet
, “
Formation and propagation of matter-wave soliton trains
,”
Nature
417
,
150
153
(
2002
).
22.
B.
Lefebvre
,
L.-J.
Chen
,
W.
Gekelman
,
P.
Kintner
,
J.
Pickett
,
P.
Pribyl
, and
S.
Vincena
, “
Debye-scale solitary structures measured in a beam-plasma laboratory experiment
,”
Nonlinear Processes Geophys.
18
,
41
47
(
2011
).
23.
G. S.
Lakhina
,
A. P.
Kakad
,
S. V.
Singh
, and
F.
Verheest
, “
Ion- and electron-acoustic solitons in two-electron temperature space plasmas
,”
Phys. Plasmas
15
,
062903
(
2008
).
24.
G. S.
Lakhina
,
S. V.
Singh
,
A. P.
Kakad
,
F.
Verheest
, and
R.
Bharuthram
, “
Study of nonlinear ion- and electron-acoustic waves in multi-component space plasmas
,”
Nonlinear Processes Geophys.
15
,
903
913
(
2008
).
25.
T. K.
Baluku
,
M. A.
Hellberg
, and
F.
Verheest
, “
New light on ion acoustic solitary waves in a plasma with two-temperature electrons
,”
Europhys. Lett.
91
,
15001
(
2010
).
26.
F.
Verheest
,
M. A.
Hellberg
,
N. S.
Saini
, and
I.
Kourakis
, “
Large acoustic solitons and double layers in plasmas with two positive ion species
,”
Phys. Plasmas
18
,
042309
(
2011
).
27.
F.
Verheest
,
M. A.
Hellberg
, and
I.
Kourakis
, “
Electrostatic supersolitons in three-species plasmas
,”
Phys. Plasmas
20
,
012302
(
2013
).
28.
V. V.
Prudskikh
, “
Large-amplitude ion acoustic solitons in a bi-ion plasma
,”
Plasma Phys. Rep.
35
,
1051
1057
(
2009
).
29.
R. Z.
Sagdeev
, “
Cooperative phenomena and shock waves in collisionless plasmas
,”
Rev. Plasma Phys.
4
,
23
(
1966
).
30.
D. J.
Korteweg
and
G.
De Vries
, “
XLI. On the change of form of long waves advancing in a rectangular canal, and on a new type of long stationary waves
,”
London, Edinburgh, Dublin Philos. Mag. J. Sci.
39
,
422
443
(
1895
).
31.
K. E.
Lonngren
, “
Soliton experiments in plasmas
,”
Plasma Phys.
25
,
943
982
(
1983
).
32.
Y.
Nakamura
,
T.
Ito
, and
K.
Koga
, “
Excitation and reflection of ion-acoustic waves by a gridded plate and a metal disk
,”
J. Plasma Phys.
49
,
331
339
(
1993
).
33.
P. O.
Dovner
,
A. I.
Eriksson
,
R.
Boström
, and
B.
Holback
, “
Freja multiprobe observations of electrostatic solitary structures
,”
Geophys. Res. Lett.
21
,
1827
1830
, https://doi.org/10.1029/94GL00886 (
1994
).
34.
Y.
Nakamura
,
H.
Bailung
, and
P. K.
Shukla
, “
Observation of ion-acoustic shocks in a dusty plasma
,”
Phys. Rev. Lett.
83
,
1602
1605
(
1999
).
35.
C.
Niemann
,
S. H.
Glenzer
,
J.
Knight
,
L.
Divol
,
E. A.
Williams
,
G.
Gregori
,
B. I.
Cohen
,
C.
Constantin
,
D. H.
Froula
,
D. S.
Montgomery
, and
R. P.
Johnson
, “
Observation of the parametric two-ion decay instability with Thomson scattering
,”
Phys. Rev. Lett.
93
,
045004
(
2004
).
36.
C. M.
Roach
,
J. W.
Connor
, and
S.
Janjua
, “
Trapped particle precession in advanced tokamaks
,”
Plasma Phys. Controlled Fusion
37
,
679
698
(
1995
).
37.
R.
Hubbard
,
D.
Gordon
,
J.
Cooley
,
B.
Hafizi
,
T.
Jones
,
D.
Kaganovich
,
P.
Sprangle
,
A.
Ting
,
A.
Zigler
, and
J.
Dexter
, “
Trapping and acceleration of nonideal injected electron bunches in laser Wakefield accelerators
,”
IEEE Trans. Plasma Sci.
33
,
712
722
(
2005
).
38.
G. I.
Stegeman
and
M.
Segev
, “
Optical spatial solitons and their interactions: universality and diversity
,”
Science
286
,
1518
1523
(
1999
).
39.
L.
Schott
, “
Propagation of ion acoustic waves close to a plasma boundary
,”
J. Appl. Phys.
59
,
1390
1391
(
1986
).
40.
K.
Mizuno
,
Y.
Nagatani
,
K.
Yamashita
, and
M.
Matsukawa
, “
Propagation of two longitudinal waves in a cancellous bone with the closed pore boundary
,”
J. Acoust. Soc. Am.
130
,
EL122
EL127
(
2011
).
41.
H. H.
Kuehl
, “
Reflection of an ion-acoustic soliton by plasma inhomogeneities
,”
Phys. Fluids
26
,
1577
1583
(
1983
).
42.
P.
Chatterjee
,
M.
Kr. Ghorui
, and
C. S.
Wong
, “
Head-on collision of dust-ion-acoustic soliton in quantum pair-ion plasma
,”
Phys. Plasmas
18
,
103710
(
2011
).
43.
T.
Chapman
,
S.
Brunner
,
J. W.
Banks
,
R. L.
Berger
,
B. I.
Cohen
, and
E. A.
Williams
, “
New insights into the decay of ion waves to turbulence, ion heating, and soliton generation
,”
Phys. Plasmas
21
,
042107
(
2014
).
44.
H. C.
Bandulet
,
C.
Labaune
,
K.
Lewis
, and
S.
Depierreux
, “
Thomson-scattering study of the subharmonic decay of ion-acoustic waves driven by the Brillouin instability
,”
Phys. Rev. Lett.
93
,
035002
(
2004
).
45.
V. O.
Jensen
and
J. P.
Lynov
, “
Energy properties of ion acoustic waves in stable and unstable plasmas
,”
Phys. Fluids
22
,
1133
1140
(
1979
).
46.
A. J.
Anastassiades
and
E.
Sideris
, “
Energy losses due to an ion-acoustic wave instability in a plasma
,”
J. Appl. Phys.
51
,
5675
5677
(
1980
).
47.
A.
Sen
,
S.
Tiwari
,
S.
Mishra
, and
P.
Kaw
, “
Nonlinear wave excitations by orbiting charged space debris objects
,”
Adv. Space Res.
56
,
429
435
(
2015
).
48.
A.
Sen
,
R.
Mukherjee
,
S. K.
Yadav
,
C.
Crabtree
, and
G.
Ganguli
, “
Electromagnetic pinned solitons for space debris detection
,”
Phys. Plasmas
30
,
012301
(
2023
).
49.
S.
Kumar Tiwari
and
A.
Sen
, “
Wakes and precursor soliton excitations by a moving charged object in a plasma
,”
Phys. Plasmas
23
,
022301
(
2016
).
50.
S. K.
Tiwari
,
A.
Das
,
P.
Kaw
, and
A.
Sen
, “
Observation of sharply peaked solitons in dusty plasma simulations
,”
New J. Phys.
14
,
063008
(
2012
).
51.
S.
Jaiswal
,
P.
Bandyopadhyay
, and
A.
Sen
, “
Experimental observation of precursor solitons in a flowing complex plasma
,”
Phys. Rev. E
93
,
041201
(
2016
).
52.
A. S.
Truitt
and
C. M.
Hartzell
, “
Simulating plasma solitons from orbital debris using the forced Korteweg–de Vries equation
,”
J. Spacecr. Rockets
57
,
876
897
(
2020
).
53.
M.
Nopoush
and
H.
Abbasi
, “
Hybrid (particle in cell-fluid) simulation of ion-acoustic soliton generation including super-thermal and trapped electrons
,”
Phys. Plasmas
18
,
082106
(
2011
).
54.
V.
Dharodi
,
A.
Kumar
, and
A.
Sen
, “
Signatures of an energetic charged body streaming in a plasma
,”
Phys. Rev. E
107
,
025207
(
2023
).
55.
X.
Qi
,
Y.-X.
Xu
,
W.-S.
Duan
,
L.-Y.
Zhang
, and
L.
Yang
, “
Particle-in-cell simulation of the head-on collision between two ion acoustic solitary waves in plasmas
,”
Phys. Plasmas
21
,
082118
(
2014
).
56.
S.
Sharma
,
S.
Sengupta
, and
A.
Sen
, “
Particle-in-cell simulation of large amplitude ion-acoustic solitons
,”
Phys. Plasmas
22
,
022115
(
2015
).
57.
X.
Qi
,
Y.-X.
Xu
,
X.-Y.
Zhao
,
L.-Y.
Zhang
,
W.-S.
Duan
, and
L.
Yang
, “
Application of particle-in-cell simulation to the description of ion acoustic solitary waves
,”
IEEE Trans. Plasma Sci.
43
,
3815
3820
(
2015
).
58.
B.
Kakad
,
A.
Kakad
, and
Y.
Omura
, “
Nonlinear evolution of ion acoustic solitary waves in space plasmas: Fluid and particle-in-cell simulations
,”
J. Geophys. Res. Space Phys.
119
,
5589
5599
, https://doi.org/10.1002/2014JA019798 (
2014
).
59.
E.
Ott
and
R. N.
Sudan
, “
Nonlinear theory of ion acoustic waves with Landau damping
,”
Phys. Fluids
12
,
2388
2394
(
1969
).
60.
J. D.
Meiss
and
P. J.
Morrison
, “
Nonlinear electron Landau damping of ion-acoustic solitons
,”
Phys. Fluids
26
,
983
989
(
1983
).
61.
A.
Myers
,
A.
Almgren
,
L. D.
Amorim
,
J.
Bell
,
L.
Fedeli
,
L.
Ge
,
K.
Gott
,
D. P.
Grote
,
M.
Hogan
,
A.
Huebl
,
R.
Jambunathan
,
R.
Lehe
,
C.
Ng
,
M.
Rowan
,
O.
Shapoval
,
M.
Thévenet
,
J. L.
Vay
,
H.
Vincenti
,
E.
Yang
,
N.
Zaïm
,
W.
Zhang
,
Y.
Zhao
, and
E.
Zoni
, “
Porting WarpX to GPU-accelerated platforms
,”
Parallel Comput.
108
,
102833
(
2021
).
62.
J. L.
Vay
,
A.
Almgren
,
J.
Bell
,
L.
Ge
,
D. P.
Grote
,
M.
Hogan
,
O.
Kononenko
,
R.
Lehe
,
A.
Myers
,
C.
Ng
,
J.
Park
,
R.
Ryne
,
O.
Shapoval
,
M.
Thévenet
, and
W.
Zhang
, “
Warp-X: A new exascale computing platform for beam–plasma simulations
,”
Nucl. Instrum. Phys. Res. Sect. A
909
,
476
479
(
2018
).
63.
J. P.
Keener
and
D. W.
McLaughlin
, “
Solitons under perturbations
,”
Phys. Rev. A
16
,
777
790
(
1977
).
64.
V. I.
Karpman
and
E. M.
Maslov
, “
Perturbation theory for solitons
,”
Zh. Eksp. Teor. Fiz.
73
,
537
559
(
1977
). See http://jetp.ras.ru/cgi-bin/dn/e_046_02_0281.pdf.
65.
S.
Watanabe
, “
Dissipative effect on formation of ion acoustic solitons
,”
J. Phys. Soc. Jpn.
43
,
1054
1059
(
1977
).
66.
H.
Schamel
, “
Role of trapped particles and waves in plasma solitons-theory and application
,”
Phys. Scr.
20
,
306
316
(
1979
).
67.
S. M.
Hosseini Jenab
and
F.
Spanier
, “
Study of trapping effect on ion-acoustic solitary waves based on a fully kinetic simulation approach
,”
Phys. Plasmas
23
,
102306
(
2016
).
68.
P.
Bernhardt
,
L.
Scott
,
A.
Howarth
,
V.
Foss
, and
G. J.
Morales
, “
Modeling of plasma wave generation by orbiting space objects for proximity detection
,” in
Proceedings of the Advanced Maui Optical and Space Surveillance (AMOS) Technologies Conference
(
Maui Economic Development Board, Inc.
,
2023
), p.
74
.