Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Dissipation-driven emergence of a soliton condensate in a nonlinear electrical transmission line

Loic Fache Univ. Lille, CNRS, UMR 8523 - PhLAM - Physique des Lasers Atomes et Molécules, F-59 000 Lille, France    Hervé Damart Univ. Lille, CNRS, UMR 8523 - PhLAM - Physique des Lasers Atomes et Molécules, F-59 000 Lille, France    François Copie Univ. Lille, CNRS, UMR 8523 - PhLAM - Physique des Lasers Atomes et Molécules, F-59 000 Lille, France    Thibault Bonnemain Department of Mathematics, King’s College, London, United Kingdom    Thibault Congy Department of Mathematics, Physics and Electrical Engineering, Northumbria University, Newcastle upon Tyne, NE1 8ST, United Kingdom    Giacomo Roberti Department of Mathematics, Physics and Electrical Engineering, Northumbria University, Newcastle upon Tyne, NE1 8ST, United Kingdom    Pierre Suret Univ. Lille, CNRS, UMR 8523 - PhLAM - Physique des Lasers Atomes et Molécules, F-59 000 Lille, France    Gennady El Department of Mathematics, Physics and Electrical Engineering, Northumbria University, Newcastle upon Tyne, NE1 8ST, United Kingdom    Stéphane Randoux stephane.randoux@univ-lille.fr Univ. Lille, CNRS, UMR 8523 - PhLAM - Physique des Lasers Atomes et Molécules, F-59 000 Lille, France
(July 3, 2024)
Abstract

We present an experimental study on the perturbed evolution of Korteweg-deVries soliton gases in a weakly dissipative nonlinear electrical transmission line. The system’s dynamics reveal that an initially dense, fully randomized, soliton gas evolves into a coherent macroscopic state identified as a soliton condensate through nonlinear spectral analysis. The emergence of the soliton condensate is driven by the spatial rearrangement of the systems’s eigenmodes and by the proliferation of new solitonic states due to nonadiabatic effects, a phenomenon not accounted for by the existing hydrodynamic theories.

In the seminal article from 1965, Zabusky and Kruskal discovered stability and interaction properties of solitons in numerical simulations of the Korteweg-deVries (KdV) equation which they had obtained as a long-wave approximation of the Fermi–Pasta–Ulam-Tsingou oscillator chain Zabusky and Kruskal (1965). Two years later, Gardner et al. introduced the inverse scattering transform (IST) method, which provides an analytical framework to solve initial value problems for the KdV equation with rapidly decaying initial conditions Gardner et al. (1967). IST theory was then generalized to other nonlinear dispersive equations of physical relevance, such as the one-dimensional nonlinear Schrödinger equation (1D-NLSE) Zakharov and Shabat (1972); Novikov et al. (1984); Yang (2010), which is a universal model for the description of many physical systems.

Solitons play a fundamental role in nonlinear physics due to their remarkable property of retaining shape, amplitude, and velocity upon interactions with other solitons Newell (1985); Miura (1976); Drazin and Johnson (1989); Copie et al. (2023). This is reflected in IST theory by the fact that solitons represent normal modes of integrable nonlinear partial differential equations. The IST formalism can be seen as a canonical transformation which carries the physical coordinates, in which the integrable equations are originally given, to new (IST) coordinates which are action-angle variables. Within the IST formalism, the amplitude and velocity of each individual soliton are encoded into a discrete eigenvalue ηksubscript𝜂𝑘\eta_{k}italic_η start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT, which represents an action variable that is time-invariant – the so-called isospectrality property of integrable evolution Newell (1985); Miura (1976); Drazin and Johnson (1989).

From the earliest days of IST theory, it was understood that the decomposition of the nonlinear field into a basis of normal modes represents a powerful tool applicable to the study of perturbed soliton evolutions relevant to physical applications where purely integrable dynamics are unlikely Karpman and Maslov (1977); Kaup et al. (1978); Kivshar and Malomed (1989). The evolution of a KdV soliton under the action of weak dissipation is a prototypical example where the effects of a small perturbation on an integrable equation have been studied extensively Madsen and Mei (1969); Grimshaw (1970); Johnson (1973); Kaup et al. (1978); Kodama and Ablowitz (1981); Kivshar and Malomed (1989); Grimshaw et al. (1994); Johnson (1994); Ko and Kuehl (1978); Knickerbocker and Newell (1980, 1985); El and Grimshaw (2002).

On the other hand, the question of emergent, large-scale behaviors in integrable or nearly-integrable many-body systems, such as soliton gases (SGs), is one of the main focuses of present-day experimental and theoretical physics Schemmer et al. (2019); Castro-Alvaredo et al. (2016); De Nardis et al. (2018); Ruggiero et al. (2020); Redor et al. (2019); Marcucci et al. (2019); Suret et al. (2020, 2023); Fache et al. (2024); Mossman et al. (2024); Vecchio et al. (2020); Congy et al. (2024). The out-of-equilibrium evolution of SGs can be investigated using the complementary tools of the nonlinear spectral kinetic theory of SGs Zakharov (1971); El and Kamchatnov (2005); El and Tovbis (2020) and generalized hydrodynamics (GHD), the hydrodynamic theory for many-body quantum and classical integrable systems Doyon et al. (2018); Doyon (2020); Castro-Alvaredo et al. (2016); Bertini et al. (2016); Bonnemain and Doyon (2024). The correspondence between these two hydrodynamic theories has been recently established in ref. Bonnemain et al. (2022) for KdV SGs.

In this Letter, we describe experiments that explore the perturbed evolution of KdV SGs in a weakly dissipative nonlinear medium. We conduct our experiments in an electrical nonlinear transmission line (NLTL), where diffusion plays a perturbative role, and the dynamics are accurately captured by the KdV-Burgers (KdVB) equation. Our experiments reveal that an initially dense KdV SG, manifested as a strongly oscillating random field, slowly evolves into a macroscopically coherent structure consisting of a broad smooth nonlinear rarefaction wave connected at its leading edge to an oscillating diffusive-dispersive shock wave El et al. (2017). From the perspective of the kinetic theory of SG, the formed composite nonlinear wave represents a soliton condensate, a critically dense SG whose macroscopic properties are dominated by soliton interactions while the individual soliton dynamics are not discernible El and Tovbis (2020); Congy et al. (2023). Soliton condensates have peculiar spectral properties which we identify by analyzing the IST eigenfunctions (or normal modes) composing the resulting wavefield. As was shown in Congy et al. (2023) the integrable KdV SG dynamics do not permit a spontaneous formation of a soliton condensate from a regular (non-condensate) SG evolution. Contrastingly, as we demonstrate, the formation of a soliton condensate from a non-condensate initial state in a perturbed integrable system becomes possible owing to a complex nonadiabatic process of the generation of numerous new solitonic modes, a phenomenon not accounted for by the existing hydrodynamic or kinetic theories of SG.

Refer to caption
Figure 1: Experiments. (a) Schematic representation of the NLTL. (b) Space-time plot showing the emergence of the soliton condensate in the NLTL. (c) Measured SG at t=0𝑡0t=0italic_t = 0 (blue line), t=175𝑡175t=175italic_t = 175 (orange line), t=420𝑡420t=420italic_t = 420 (green line). (d) Time evolution of the discrete eigenvalues measured during the evolution of the SG. The red flow is associated to newly generated eigenvalues arising from the excitation of the continuous spectrum. (e) Time evolution of the number N(t)𝑁𝑡N(t)italic_N ( italic_t ) of discrete eigenvalues measured during the evolution of the SG. The left (resp. right) inset shows the discrete IST spectrum at t=0𝑡0t=0italic_t = 0 (resp. t=420𝑡420t=420italic_t = 420). The red line in (e) is obtained from numerical simulations of the experiment with ϵ=0.1italic-ϵ0.1\epsilon=0.1italic_ϵ = 0.1.

Our experimental setup, shown schematically in Fig. 1(a), is an NLTL with an architecture similar to that used in previous investigations of KdV soliton propagation Remoissenet (1996); Ricketts and Ham (2018); Jäger (1982); Elizondo-Decanini et al. (2015). It consists of a ladder network of 160160160160 identical lumped inductor-varactor sections with inductance L𝐿Litalic_L and varactor capacitance C(V)𝐶𝑉C(V)italic_C ( italic_V ). The varactors are commercial components (Varicap) for which C(V)𝐶𝑉C(V)italic_C ( italic_V ) is a linearly-decreasing function of the applied voltage V𝑉Vitalic_V. The varactors have a small series resistance R𝑅Ritalic_R, which results into frequency-dependent NLTL losses. Hence, propagation in the NLTL in the long-wavelength limit is not described by the KdV equation but by the KdVB equation Ricketts and Ham (2018):

ut+6uux+uxxx=ϵuxx,subscript𝑢𝑡6𝑢subscript𝑢𝑥subscript𝑢𝑥𝑥𝑥italic-ϵsubscript𝑢𝑥𝑥u_{t}+6uu_{x}+u_{xxx}=\epsilon u_{xx},italic_u start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT + 6 italic_u italic_u start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT + italic_u start_POSTSUBSCRIPT italic_x italic_x italic_x end_POSTSUBSCRIPT = italic_ϵ italic_u start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT , (1)

where u(x,t)𝑢𝑥𝑡u(x,t)italic_u ( italic_x , italic_t ) represents the normalized electrical voltage, see Supplemental material for details about the experimental setup and the normalization. The strength ϵitalic-ϵ\epsilonitalic_ϵ of the Burgers term in Eq. (1) is determined by the resistance R𝑅Ritalic_R of the varactors. In our experiments, the value of ϵitalic-ϵ\epsilonitalic_ϵ is 0.10.10.10.1.

The initial condition u0(x)=u(x,t=0)subscript𝑢0𝑥𝑢𝑥𝑡0u_{0}(x)=u(x,t=0)italic_u start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_x ) = italic_u ( italic_x , italic_t = 0 ) used in our experiment is a dense SG numerically synthesized as a N𝑁Nitalic_N-soliton solution of the KdV equation using the Darboux transformation method Congy et al. (2023); Liao and Huang (1996); Perego (2024), see Supplemental material for details. The discrete IST spectrum of the initial condition is composed of N=200𝑁200N=200italic_N = 200 discrete eigenvalues ηksubscript𝜂𝑘\eta_{k}italic_η start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT that are randomly distributed between 0.20.20.20.2 and 0.60.60.60.6 in a uniform way, see Fig. 1(e). As shown in Fig. 1(c), the initial field measured at the first LC𝐿𝐶LCitalic_L italic_C cell of the NLTL is a random oscillating field in which individual solitons cannot be discerned due to their strong overlap and interactions.

Measuring the voltage along the NLTL, we have built a space-time plot revealing the evolution of the SG up to t=420𝑡420t=420italic_t = 420, see Fig. 1(b). In the absence of the diffusion term (ϵ=0italic-ϵ0\epsilon=0italic_ϵ = 0), the dense SG initially localized in space would have evolved into a rank-ordered expanding soliton train, see ref. Karpman (1967) and Fig. 3(g). Contrastingly, we observe that the presence of the small diffusion term prevents the dense SG from reaching a diluted state where individual solitons separate, see Fig. 1(b)(c). Physically, weak dissipation promotes soliton interactions, leading to the emergence at long time (t400similar-to𝑡400t\sim 400italic_t ∼ 400) of a broad, smooth wavefield with a coherent oscillatory structure identified as a diffusive-dispersive shock wave El et al. (2017); Ricketts and Ham (2018).

Using the perspective of the IST method, we can gain insights into the features observed in the experiment. It is well known that the IST spectrum of the solutions of the KdV equation (Eq. (1) with ϵ=0italic-ϵ0\epsilon=0italic_ϵ = 0) can be obtained by solving the linear Schrödinger equation for an auxiliary function ψ(x,t)𝜓𝑥𝑡\psi(x,t)italic_ψ ( italic_x , italic_t )

ψxx+(u(x,t)λ)ψ=0,subscript𝜓𝑥𝑥𝑢𝑥𝑡𝜆𝜓0\psi_{xx}+(u(x,t)-\lambda)\psi=0,italic_ψ start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT + ( italic_u ( italic_x , italic_t ) - italic_λ ) italic_ψ = 0 , (2)

where the KdV field u(x,t)𝑢𝑥𝑡u(x,t)italic_u ( italic_x , italic_t ) plays the role of a potential with t𝑡titalic_t being a parameter Newell (1985); Miura (1976); Drazin and Johnson (1989). For the fundamental soliton solution of the KdV equation uk(x,t)=2ηk2sech2(ηk(x4ηk2tx0k))subscript𝑢𝑘𝑥𝑡2superscriptsubscript𝜂𝑘2superscriptsech2subscript𝜂𝑘𝑥4superscriptsubscript𝜂𝑘2𝑡subscript𝑥0𝑘u_{k}(x,t)=2\eta_{k}^{2}\,\text{sech}^{2}(\eta_{k}(x-4\eta_{k}^{2}t-x_{0k}))italic_u start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_x , italic_t ) = 2 italic_η start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT sech start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_η start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_x - 4 italic_η start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_t - italic_x start_POSTSUBSCRIPT 0 italic_k end_POSTSUBSCRIPT ) ), Eq. (2) admits only one eigenmode that is parameterized by one real discrete eigenvalue λk=ηk2subscript𝜆𝑘superscriptsubscript𝜂𝑘2\lambda_{k}=-\eta_{k}^{2}italic_λ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = - italic_η start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT and by one normalized real eigenfunction ψk(x,t)subscript𝜓𝑘𝑥𝑡\psi_{k}(x,t)italic_ψ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_x , italic_t ) that is localized in space (|ψk(x,t)|0subscript𝜓𝑘𝑥𝑡0|\psi_{k}(x,t)|\rightarrow 0| italic_ψ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_x , italic_t ) | → 0 for |x|𝑥|x|\rightarrow\infty| italic_x | → ∞ and +ψk2(x,t)𝑑x=1superscriptsubscriptsuperscriptsubscript𝜓𝑘2𝑥𝑡differential-d𝑥1\int_{-\infty}^{+\infty}\psi_{k}^{2}(x,t)\,dx=1∫ start_POSTSUBSCRIPT - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + ∞ end_POSTSUPERSCRIPT italic_ψ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_x , italic_t ) italic_d italic_x = 1) Drazin and Johnson (1989). Importantly, any N𝑁Nitalic_N-soliton solution of the KdV equation with purely discrete IST spectrum can be represented using the following reconstruction formula Gardner et al. ; Miura (1976); Newell (1985)

u(x,t)=4k=1Nηkψk2(x,t).𝑢𝑥𝑡4superscriptsubscript𝑘1𝑁subscript𝜂𝑘superscriptsubscript𝜓𝑘2𝑥𝑡u(x,t)=4\sum_{k=1}^{N}\eta_{k}\,\psi_{k}^{2}(x,t).italic_u ( italic_x , italic_t ) = 4 ∑ start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT italic_η start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_ψ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_x , italic_t ) . (3)

Using a perturbative approach, where the Burgers term in Eq. (1) is considered to be small, we computed the discrete eigenvalues ηksubscript𝜂𝑘\eta_{k}italic_η start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT associated with the experimentally measured field u(x,t)𝑢𝑥𝑡u(x,t)italic_u ( italic_x , italic_t ) by solving Eq. (2) at various times t𝑡titalic_t. If the dynamics were governed by the KdV equation, one would expect an isospectral evolution, wherein the N=200𝑁200N=200italic_N = 200 eigenvalues specified by our initial condition would remain unchanged over time. Instead, we observe that the eigenvalues distribute along two distinct “flows”, see Fig. 1(d). In the first flow (black lines in Fig. 1(d)), the eigenvalues associated with the eigenmodes composing the initial SG slowly decay with time, as expected from the soliton perturbation theory Karpman and Maslov (1977); Kaup et al. (1978); Kivshar and Malomed (1989). The second flow (red lines in Fig. 1(d)) consists of the newly formed discrete eigenvalues progressively emerging during the evolution. These new eigenvalues correspond to the generation of small-amplitude solitons whose number N𝑁Nitalic_N grows with time, see Fig. 1(e) showing that N𝑁Nitalic_N grows from N(t=0)=200𝑁𝑡0200N(t=0)=200italic_N ( italic_t = 0 ) = 200 to N(t=420)=225𝑁𝑡420225N(t=420)=225italic_N ( italic_t = 420 ) = 225.

Refer to caption
Figure 2: Experiments. (a) Space-time evolution of a solitary wave in the NLTL. (b) Measured solitary wave at t=0𝑡0t=0italic_t = 0 (blue line), t=5.8𝑡5.8t=5.8italic_t = 5.8 (orange line), t=16.6𝑡16.6t=16.6italic_t = 16.6 (green line). (c) Time evolution of the discrete eigenvalues measured during the evolution of the solitary wave. The red lines in (c) are obtained from numerical simulations of the experiment (ϵ=0.1italic-ϵ0.1\epsilon=0.1italic_ϵ = 0.1). (g) Numerical simulations showing the oscillatory evolution of the masses MDSsubscript𝑀𝐷𝑆M_{DS}italic_M start_POSTSUBSCRIPT italic_D italic_S end_POSTSUBSCRIPT and MCSsubscript𝑀𝐶𝑆M_{CS}italic_M start_POSTSUBSCRIPT italic_C italic_S end_POSTSUBSCRIPT of the discrete and continous spectra.

The described complex process of the proliferation of new soliton states represents a counterpart of a simpler phenomenon observed in the individual evolution of a solitary wave in the NLTL. As seen in Fig. 2(a)(b), the amplitude and velocity of the soliton, taken as the initial condition, slowly decay during the propagation in the NLTL. This gradual evolution of the solitary wave is accompanied by the formation of a trailing shelf, which is analogous to a prominent feature of the propagation of solitary waves in shallow water with slowly varying depth Knickerbocker and Newell (1985); Kaup et al. (1978); Newell (1985). Nonlinear spectral analysis, performed by solving Eq. (2) for the measured field u(x,t)𝑢𝑥𝑡u(x,t)italic_u ( italic_x , italic_t ), reveals that the discrete eigenvalue linked to the initial condition gradually decays over time, while two new discrete eigenvalues are created during the shelf formation, as seen in Fig. 2(c).

This is a non-adiabatic process because, although the soliton’s amplitude changes slowly, the shelf’s extent does not, resulting in an O(1)𝑂1O(1)italic_O ( 1 ) mass of the shelf despite its small amplitude. Consequently, the solitary wave excites radiative modes (associated with the continuous IST spectrum), potentially leading to the creation of new solitonic eigenmodes Kaup et al. (1978); Newell (1985); Wright (1980); El and Grimshaw (2002). In our experiment described by the KdVB equation, the excitation of the continuous spectrum by the solitary wave is observed through opposite oscillations between the masses MDSsubscript𝑀𝐷𝑆M_{DS}italic_M start_POSTSUBSCRIPT italic_D italic_S end_POSTSUBSCRIPT and MCSsubscript𝑀𝐶𝑆M_{CS}italic_M start_POSTSUBSCRIPT italic_C italic_S end_POSTSUBSCRIPT of the discrete and continuous spectra, such that the total mass M=u(x,t)dx=MDS+MCS𝑀𝑢𝑥𝑡differential-d𝑥subscript𝑀𝐷𝑆subscript𝑀𝐶𝑆M=\int u(x,t){\rm d}x=M_{DS}+M_{CS}italic_M = ∫ italic_u ( italic_x , italic_t ) roman_d italic_x = italic_M start_POSTSUBSCRIPT italic_D italic_S end_POSTSUBSCRIPT + italic_M start_POSTSUBSCRIPT italic_C italic_S end_POSTSUBSCRIPT is conserved Wright (1980), see Fig. 2(d) and Supplemental Material for details.

Refer to caption
Figure 3: Numerical simulations. (a) SG used as initial condition in the experiment. (b) Spatial distribution of the discrete eigenmodes represented as ϕk(x)=ηk+sψk(x)subscriptitalic-ϕ𝑘𝑥subscript𝜂𝑘𝑠subscript𝜓𝑘𝑥\phi_{k}(x)=\eta_{k}+s\,\psi_{k}(x)italic_ϕ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_x ) = italic_η start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT + italic_s italic_ψ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_x ), where ηksubscript𝜂𝑘\eta_{k}italic_η start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT (resp. ψk(x)subscript𝜓𝑘𝑥\psi_{k}(x)italic_ψ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_x )) (k=1,..200𝑘1..200k=1,..200italic_k = 1 , ..200) represent the eigenvalues (resp. eigenfunctions) computed by solving Eq. (2). (c) DOS of the SG shown in (a). (d), (e), (f) Same as (a), (b), (c) but for a simulation of Eq. (1) at t=420𝑡420t=420italic_t = 420 with ϵ=0.1italic-ϵ0.1\epsilon=0.1italic_ϵ = 0.1. (g), (h), (i) Same as (a), (b), (c) but for a simulation of Eq. (1) at t=4000𝑡4000t=4000italic_t = 4000 with ϵ=0italic-ϵ0\epsilon=0italic_ϵ = 0. (j), (k), (k) Same as (a), (b), (c) but for a simulation of Eq. (1) at t=4000𝑡4000t=4000italic_t = 4000 with ϵ=2.103italic-ϵsuperscript2.103\epsilon=2.10^{-3}italic_ϵ = 2.10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT. The inset in (f) (resp. in (l)) shows the DOS measured at x=500𝑥500x=500italic_x = 500 (resp. x=2900𝑥2900x=2900italic_x = 2900).

Using nonlinear spectral analysis, we now examine how these nonadiabatic effects impact the evolution of the perturbed KdV SG. Fig. 3(a) shows the SG used as initial condition in our experiment. The associated distribution of the discrete eigenvalues ηksubscript𝜂𝑘\eta_{k}italic_η start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT and eigenfunctions ψk(x)subscript𝜓𝑘𝑥\psi_{k}(x)italic_ψ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_x ) parameterizing this SG is plotted in a single representation showing ϕk(x)=ηk+sψk(x)subscriptitalic-ϕ𝑘𝑥subscript𝜂𝑘𝑠subscript𝜓𝑘𝑥\phi_{k}(x)=\eta_{k}+s\,\psi_{k}(x)italic_ϕ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_x ) = italic_η start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT + italic_s italic_ψ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_x ) (k=1,..200𝑘1..200k=1,..200italic_k = 1 , ..200), see Fig. 3(b). The eigenvalues and eigenfunctions are computed by solving numerically Eq. (2) for u(x,t)=u0(x)𝑢𝑥𝑡subscript𝑢0𝑥u(x,t)=u_{0}(x)italic_u ( italic_x , italic_t ) = italic_u start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_x ). We introduced a scaling factor, 0<s10𝑠much-less-than10<s\ll 10 < italic_s ≪ 1, to prevent overlap among the eigenfunctions in Fig. 3(b).

As shown in the insets of Fig. 3(b), the degree of localization in space of the eigenfunctions composing the random initial condition u0(x)subscript𝑢0𝑥u_{0}(x)italic_u start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_x ) depends of their associated eigenvalues. Nonlinear discrete eigenmodes with the largest eigenvalues (ηk0.6similar-tosubscript𝜂𝑘0.6\eta_{k}\sim 0.6italic_η start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ∼ 0.6) are more localized in space than the modes with the smallest eigenvalues (ηk0.2similar-tosubscript𝜂𝑘0.2\eta_{k}\sim 0.2italic_η start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ∼ 0.2). Taking u0(x)subscript𝑢0𝑥u_{0}(x)italic_u start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_x ) as the initial condition and solving numerically Eq. (1) for ϵ=0.1italic-ϵ0.1\epsilon=0.1italic_ϵ = 0.1, we obtain the field shown in Fig. 3(d), in good quantitative agreement with the experimental measurements. Fig. 3(e) reveals that the distributions of the eigenvalues and of the eigenfunctions composing the field have been profoundly modified by the nonlinear evolution. In addition to the creation of new eigenfunctions with spectral parameters between 00 and 0.20.20.20.2, all the eigenfunctions composing the wavefield at t=420𝑡420t=420italic_t = 420 have lost their random structure and adopted similar uniform oscillatory shape.

This large-scale behavior is in sharp contrast with that observed in the absence of diffusion, for ϵ=0italic-ϵ0\epsilon=0italic_ϵ = 0. As shown in Fig. 3(g), for a pure KdV evolution, a rank-ordered soliton train forms at long evolution time, as a result of an isospectral evolution where the eigenfunctions composing the nonlinear wave field become spatially localized as solitonic peaks, see Fig. 3(h). An intermediate situation between the experimental case (ϵ=0.1italic-ϵ0.1\epsilon=0.1italic_ϵ = 0.1 in Fig. 3(d)) and the purely integrable case (ϵ=0italic-ϵ0\epsilon=0italic_ϵ = 0 in Fig. 3(g)) is observed for a very small diffusion parameter (ϵ=2.103italic-ϵsuperscript2.103\epsilon=2.10^{-3}italic_ϵ = 2.10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT). In this situation, the SG exhibits a slowed-down expansion, in which a random soliton train propagates over a non-zero background, see Fig. 3(j). Fig. 3(k) shows that the associated eigenfunctions are divided into two classes: the localized modes and the delocalized modes, the latter originating from the excitation of the continuous spectrum.

Now we show that the squared eigenfunctions ψk2(x,t)superscriptsubscript𝜓𝑘2𝑥𝑡\psi_{k}^{2}(x,t)italic_ψ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_x , italic_t ) are directly associated with the density of states (DOS) f(ηk;x,t)𝑓subscript𝜂𝑘𝑥𝑡f(\eta_{k};x,t)italic_f ( italic_η start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ; italic_x , italic_t ) in the SG spectral kinetic theory, revealing their fundamental role in this framework. The SG spectral kinetic theory is a statistical theory describing the transport in space and time of the DOS. The DOS f(η;x,t)𝑓𝜂𝑥𝑡f(\eta;x,t)italic_f ( italic_η ; italic_x , italic_t ) is defined as the joint distribution over the spectral eigenvalues and the soliton positions, so that f(η;x,t)dηdx𝑓𝜂𝑥𝑡d𝜂d𝑥f(\eta;x,t){\rm d}\eta{\rm d}xitalic_f ( italic_η ; italic_x , italic_t ) roman_d italic_η roman_d italic_x represents the number of soliton states found at time t𝑡titalic_t in the element of the phase space [η,η+dη]×[x,x+dx]𝜂𝜂d𝜂𝑥𝑥d𝑥[\eta,\eta+{\rm d}\eta]\times[x,x+{\rm d}x][ italic_η , italic_η + roman_d italic_η ] × [ italic_x , italic_x + roman_d italic_x ] El and Tovbis (2020); El (2021); Suret et al. (2020, 2023); Fache et al. (2024). In particular, knowledge of the DOS provides the statistical moments of the field u(x,t)𝑢𝑥𝑡u(x,t)italic_u ( italic_x , italic_t ), including its mean value El (2021); Bonnemain et al. (2022):

u(x,t)=401ηf(η;x,t)dη.delimited-⟨⟩𝑢𝑥𝑡4superscriptsubscript01𝜂𝑓𝜂𝑥𝑡differential-d𝜂\langle u\rangle(x,t)=4\int_{0}^{1}\eta f(\eta;x,t){\rm d}\eta.⟨ italic_u ⟩ ( italic_x , italic_t ) = 4 ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT italic_η italic_f ( italic_η ; italic_x , italic_t ) roman_d italic_η . (4)

The brackets in Eq. (4) denote ensemble averaging over a large number of realizations of the SG.

By comparing Eq. (3) with Eq. (4), one can infer that the DOS can be computed using the equality

f(ηk;x,t)dη=j=kk+Δkψj2(x,t),𝑓subscript𝜂𝑘𝑥𝑡d𝜂superscriptsubscript𝑗𝑘𝑘Δ𝑘delimited-⟨⟩superscriptsubscript𝜓𝑗2𝑥𝑡f(\eta_{k};x,t){\rm d}\eta=\sum_{j=k}^{k+\Delta k}\langle\psi_{j}^{2}(x,t)\rangle,italic_f ( italic_η start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ; italic_x , italic_t ) roman_d italic_η = ∑ start_POSTSUBSCRIPT italic_j = italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k + roman_Δ italic_k end_POSTSUPERSCRIPT ⟨ italic_ψ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_x , italic_t ) ⟩ , (5)

where k𝑘kitalic_k and k+Δk𝑘Δ𝑘k+\Delta kitalic_k + roman_Δ italic_k are integer indices bounding the discrete summation over the squared eigenfunctions associated with eigenvalues between ηksubscript𝜂𝑘\eta_{k}italic_η start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT and ηk+dηsubscript𝜂𝑘d𝜂\eta_{k}+{\rm d}\etaitalic_η start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT + roman_d italic_η. Equation (5) provides a local expression for the DOS of a SG, akin to formulas used in other branches of physics such as electron microscopy García de Abajo and Kociak (2008); Tersoff and Hamann (1985).

Fig. 3(c), (f), (i), (l) show the DOS of the SGs from our numerical simulations. These plots were created using Eq. (5) and 1000100010001000 realizations of each SG shown in Fig. 3(a), (d), (g), (j). Fig. 3(c) shows that the DOS of the initial SG is not uniform in space. The fact that the KdV evolution leads to the emergence of a rank-ordered soliton train (Fig. 3(g)), with eigenfunctions becoming localized in space (Fig. 3(h)), translates into the DOS being localized around the curve shown in Fig. 3(i).

Conversely, the SG formed in the experiment under the perturbative influence of diffusion exhibits a very specific DOS, locally described by the Weyl distribution f(x,η)=η/(πλ12(x)η2)𝑓𝑥𝜂𝜂𝜋superscriptsubscript𝜆12𝑥superscript𝜂2f(x,\eta)=\eta/(\pi\sqrt{\lambda_{1}^{2}(x)-\eta^{2}})italic_f ( italic_x , italic_η ) = italic_η / ( italic_π square-root start_ARG italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_x ) - italic_η start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) Congy et al. (2023), with the function λ1(x)subscript𝜆1𝑥\lambda_{1}(x)italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_x ) being slowly modulated in space, see Fig. 3(f). This DOS corresponds to the peculiar type of SG known as a genus zero soliton condensate in the spectral kinetic theory of SG Congy et al. (2023); Kuijlaars and Tovbis (2021); Gelash et al. (2021); El and Tovbis (2020). Here this soliton condensate can be seen as the SG of minimal entropy or of zero variance Bonnemain et al. (2022).

For a diffusion strength smaller than the experimental one (ϵ=2.103italic-ϵsuperscript2.103\epsilon=2.10^{-3}italic_ϵ = 2.10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT), the DOS at a given position x>0𝑥0x>0italic_x > 0 is supported on two disjoint spectral intervals, see Fig. 3(l). The corresponding coherent oscillatory structure, identified above as a dispersive-diffusive shock wave, is also characterized by the soliton condensate DOS within the SG spectral framework, however, now it is the genus one condensate, see Supplemental Material and Congy et al. (2023) for details. As shown in Congy et al. (2023), the transitions from a non-condensate to the condensate state and vice versa are not possible in an integrable system. Our experiment demonstrates that soliton condensate can spontaneously emerge in a nearly integrable physical system due to the presence of a perturbative dissipative effect. This observation suggests the possibility of performing further experiments equivalent of Joule expansion, by evolving a SG according to KdVB equation and suddenly setting the dissipation parameter ϵitalic-ϵ\epsilonitalic_ϵ to zero.

In summary, we have reported the emergence of a soliton condensate in a NLTL, a nearly integrable physical system described by the KdVB equation. This phenomenon arises from the dissipation-driven rearrangement of solitonic eigenmodes within the system, accompanied by the excitation of the continuous spectrum, resulting in the formation of new soliton states. The coupling between solitons and continuous spectrum radiation has been recently highlighted in deep-water models less restrictive than the integrable 1D-NLSE Gelash et al. (2024). Notably, current hydrodynamic theories like GHD have yet to incorporate the excitation of the continuous spectrum by a statistical ensemble of solitons. We anticipate that our work will inspire further research in this area.

Acknowledgements.
This work has been partially supported by the Agence Nationale de la Recherche through the StormWave (ANR-21-CE30-0009) and SOGOOD (ANR-21-CE30-0061) projects, the LABEX CEMPI project (ANR-11-LABX-0007), the Ministry of Higher Education and Research, Hauts de France council and European Regional Development Fund (ERDF) through the Nord-Pas de Calais Regional Research Council and the European Regional Development Fund (ERDF) through the Contrat de Projets Etat-Région (CPER Photonics for Society P4S). The authors would like to thank the Centre d’Etudes et de Recherche Lasers et Application (CERLA) for technical support and the Isaac Newton Institute for Mathematical Sciences for support and hospitality during the programme “Dispersive hydrodynamics: mathematics, simulation and experiments, with applications in nonlinear waves”. TB was supported by the Engineering and Physical Sciences Research Council (EPSRC) under grant EP/W010194/1. GE’s and GR’s work was also supported by the EPSRC, grant number EP/W032759/1. GR thanks the Simons Foundation for partial support.

References

Supplemental material for :
”Dissipation-driven emergence of a soliton condensate in a nonlinear electrical transmission line”

Loic Fache,1 Hervé Damart,1 François Copie,1 Thibault Bonnemain, 2, Giacomo Roberti,3, Thibault Congy,3 Pierre Suret,1 Gennady El,3 Stéphane Randoux1

1 Univ. Lille, CNRS, UMR 8523 - PhLAM -Physique des Lasers Atomes et Molécules, F-59 000 Lille, France

2 Department of Mathematics, King’s College, London, United Kingdom

3 Department of Mathematics, Physics and Electrical Engineering, Northumbria University, Newcastle upon Tyne, NE1 8ST, United Kingdom

The purpose of this Supplemental Material is to provide some details about the experimental setup and about the experimental methododology. All equations, figures, reference numbers within this document are prepended with “S” to distinguish them from corresponding numbers in the Letter.

I Description of the experimental setup and of its associated dynamical equations

Refer to caption
Figure S4: (a) Schematic representation of the experimental setup consisting into a ladder network of 160160160160 lumped inductor-varactor sections. The inductors have a constant inductance L15μsimilar-to-or-equals𝐿15𝜇L\simeq 15\muitalic_L ≃ 15 italic_μH while the varactors have a capacitance that decreases with the applied voltage. The varactors have also a small but non-negligible series resistance R5Ωsimilar-to-or-equals𝑅5ΩR\simeq 5\,\Omegaitalic_R ≃ 5 roman_Ω. (b) Photograph of the experimental setup. (c) Typical characteristic of the varicap used in the NLTL showing the evolution of the capacitance C𝐶Citalic_C as a function of the applied voltage. The characteristic is nearly linear for positive voltages below 3similar-toabsent3\sim 3∼ 3 V. For this range of voltage, it can be approximated by C(Vn)=C0(12bVn)𝐶subscript𝑉𝑛subscript𝐶012𝑏subscript𝑉𝑛C(V_{n})=C_{0}(1-2bV_{n})italic_C ( italic_V start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) = italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( 1 - 2 italic_b italic_V start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) with C0=140subscript𝐶0140C_{0}=140italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 140 pF and b=0.154𝑏0.154b=0.154italic_b = 0.154 V-1.

Our experimental setup is the electrical nonlinear transmission line (NLTL) schematically shown in Fig. S1(a). It consists into a ladder network of 160160160160 lumped inductor-varactor sections with each individual inductance (KEMET SBCP-47HY150B) being L15μsimilar-to-or-equals𝐿15𝜇L\simeq 15\muitalic_L ≃ 15 italic_μ H and each varactor capacitance C(Vn)𝐶subscript𝑉𝑛C(V_{n})italic_C ( italic_V start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) being a decreasing function of the voltage Vnsubscript𝑉𝑛V_{n}italic_V start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT applied to it. The varactors are commercial components (Varicap: NXP-BB201). Fig. S1(c) shows the measured capacitance C(Vn)𝐶subscript𝑉𝑛C(V_{n})italic_C ( italic_V start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) as the function of the voltage Vnsubscript𝑉𝑛V_{n}italic_V start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT. For positive voltages below 3similar-toabsent3\sim 3∼ 3 V, the capacitance decreases approximately linearly with the voltage and the curve plotted in Fig. S1(c) can be well approximated by C(Vn)=C0(12bVn)𝐶subscript𝑉𝑛subscript𝐶012𝑏subscript𝑉𝑛C(V_{n})=C_{0}(1-2bV_{n})italic_C ( italic_V start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) = italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( 1 - 2 italic_b italic_V start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) with C0=140subscript𝐶0140C_{0}=140italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 140 pF and b=0.154𝑏0.154b=0.154italic_b = 0.154 V-1. Finally and importantly, the varactors have a small but non-negligible resistance R5Ωsimilar-to𝑅5ΩR\sim 5\,\Omegaitalic_R ∼ 5 roman_Ω.

The theoretical derivation of the dynamical equations describing the NLTL shown in Fig. S1(a) has been reported in details in ref. Ricketts and Ham (2018). For the sake of clarity and completeness, we report here the main steps showing that nonlinear propagation in the NLTL in the long-wavelength limit is described by the KdV-Burgers equation. The nodal equations for the voltages and currents shown in Fig. S1(a) read:

InIn1=Q(Vn)t,subscript𝐼𝑛subscript𝐼𝑛1𝑄subscript𝑉𝑛𝑡I_{n}-I_{n-1}=\frac{\partial Q(V_{n})}{\partial t},italic_I start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT - italic_I start_POSTSUBSCRIPT italic_n - 1 end_POSTSUBSCRIPT = divide start_ARG ∂ italic_Q ( italic_V start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) end_ARG start_ARG ∂ italic_t end_ARG , (S1)
LInt=[Vn1+RQ(Vn1)t][Vn+RQ(Vn)t],𝐿subscript𝐼𝑛𝑡delimited-[]subscript𝑉𝑛1𝑅𝑄subscript𝑉𝑛1𝑡delimited-[]subscript𝑉𝑛𝑅𝑄subscript𝑉𝑛𝑡L\frac{\partial I_{n}}{\partial t}=\left[V_{n-1}+R\frac{\partial Q(V_{n-1})}{% \partial t}\right]-\left[V_{n}+R\frac{\partial Q(V_{n})}{\partial t}\right],italic_L divide start_ARG ∂ italic_I start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_t end_ARG = [ italic_V start_POSTSUBSCRIPT italic_n - 1 end_POSTSUBSCRIPT + italic_R divide start_ARG ∂ italic_Q ( italic_V start_POSTSUBSCRIPT italic_n - 1 end_POSTSUBSCRIPT ) end_ARG start_ARG ∂ italic_t end_ARG ] - [ italic_V start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT + italic_R divide start_ARG ∂ italic_Q ( italic_V start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) end_ARG start_ARG ∂ italic_t end_ARG ] , (S2)
LIn+1t=[Vn+RQ(Vn)t][Vn+1+RQ(Vn+1)t],𝐿subscript𝐼𝑛1𝑡delimited-[]subscript𝑉𝑛𝑅𝑄subscript𝑉𝑛𝑡delimited-[]subscript𝑉𝑛1𝑅𝑄subscript𝑉𝑛1𝑡L\frac{\partial I_{n+1}}{\partial t}=\left[V_{n}+R\frac{\partial Q(V_{n})}{% \partial t}\right]-\left[V_{n+1}+R\frac{\partial Q(V_{n+1})}{\partial t}\right],italic_L divide start_ARG ∂ italic_I start_POSTSUBSCRIPT italic_n + 1 end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_t end_ARG = [ italic_V start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT + italic_R divide start_ARG ∂ italic_Q ( italic_V start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) end_ARG start_ARG ∂ italic_t end_ARG ] - [ italic_V start_POSTSUBSCRIPT italic_n + 1 end_POSTSUBSCRIPT + italic_R divide start_ARG ∂ italic_Q ( italic_V start_POSTSUBSCRIPT italic_n + 1 end_POSTSUBSCRIPT ) end_ARG start_ARG ∂ italic_t end_ARG ] , (S3)

Combining Eqs. (S1), (S2), (S3), we obtain:

L2Q(Vn)t2=Vn1+Vn+12Vn+Rt[Q(Vn+1)+Q(Vn1)2Q(Vn)]𝐿superscript2𝑄subscript𝑉𝑛superscript𝑡2subscript𝑉𝑛1subscript𝑉𝑛12subscript𝑉𝑛𝑅𝑡delimited-[]𝑄subscript𝑉𝑛1𝑄subscript𝑉𝑛12𝑄subscript𝑉𝑛L\frac{\partial^{2}Q(V_{n})}{\partial t^{2}}=V_{n-1}+V_{n+1}-2V_{n}+R\frac{% \partial}{\partial t}\left[Q(V_{n+1})+Q(V_{n-1})-2Q(V_{n})\right]italic_L divide start_ARG ∂ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_Q ( italic_V start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) end_ARG start_ARG ∂ italic_t start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG = italic_V start_POSTSUBSCRIPT italic_n - 1 end_POSTSUBSCRIPT + italic_V start_POSTSUBSCRIPT italic_n + 1 end_POSTSUBSCRIPT - 2 italic_V start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT + italic_R divide start_ARG ∂ end_ARG start_ARG ∂ italic_t end_ARG [ italic_Q ( italic_V start_POSTSUBSCRIPT italic_n + 1 end_POSTSUBSCRIPT ) + italic_Q ( italic_V start_POSTSUBSCRIPT italic_n - 1 end_POSTSUBSCRIPT ) - 2 italic_Q ( italic_V start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) ] (S4)

Changing Vn(t)subscript𝑉𝑛𝑡V_{n}(t)italic_V start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_t ) into V(n,t)𝑉𝑛𝑡V(n,t)italic_V ( italic_n , italic_t ) and Taylor-expanding the voltages V(n,t)𝑉𝑛𝑡V(n,t)italic_V ( italic_n , italic_t ) as

Vn±1=V±Vn+12!2Vn2±13!3Vn3+14!4Vn4±O(n5),subscript𝑉plus-or-minus𝑛1plus-or-minusplus-or-minusplus-or-minus𝑉𝑉𝑛12superscript2𝑉superscript𝑛213superscript3𝑉superscript𝑛314superscript4𝑉superscript𝑛4𝑂superscript𝑛5V_{n\pm 1}=V\pm\frac{\partial V}{\partial n}+\frac{1}{2!}\frac{\partial^{2}V}{% \partial n^{2}}\pm\frac{1}{3!}\frac{\partial^{3}V}{\partial n^{3}}+\frac{1}{4!% }\frac{\partial^{4}V}{\partial n^{4}}\pm O(n^{5}),italic_V start_POSTSUBSCRIPT italic_n ± 1 end_POSTSUBSCRIPT = italic_V ± divide start_ARG ∂ italic_V end_ARG start_ARG ∂ italic_n end_ARG + divide start_ARG 1 end_ARG start_ARG 2 ! end_ARG divide start_ARG ∂ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_V end_ARG start_ARG ∂ italic_n start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ± divide start_ARG 1 end_ARG start_ARG 3 ! end_ARG divide start_ARG ∂ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT italic_V end_ARG start_ARG ∂ italic_n start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG + divide start_ARG 1 end_ARG start_ARG 4 ! end_ARG divide start_ARG ∂ start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT italic_V end_ARG start_ARG ∂ italic_n start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT end_ARG ± italic_O ( italic_n start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT ) , (S5)

we obtain

Vn+1+Vn12Vn=2V(n,t)n2+1124V(n,t)n4+O(n5)subscript𝑉𝑛1subscript𝑉𝑛12subscript𝑉𝑛superscript2𝑉𝑛𝑡superscript𝑛2112superscript4𝑉𝑛𝑡superscript𝑛4𝑂superscript𝑛5V_{n+1}+V_{n-1}-2V_{n}=\frac{\partial^{2}V(n,t)}{\partial n^{2}}+\frac{1}{12}% \frac{\partial^{4}V(n,t)}{\partial n^{4}}+O(n^{5})italic_V start_POSTSUBSCRIPT italic_n + 1 end_POSTSUBSCRIPT + italic_V start_POSTSUBSCRIPT italic_n - 1 end_POSTSUBSCRIPT - 2 italic_V start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = divide start_ARG ∂ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_V ( italic_n , italic_t ) end_ARG start_ARG ∂ italic_n start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + divide start_ARG 1 end_ARG start_ARG 12 end_ARG divide start_ARG ∂ start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT italic_V ( italic_n , italic_t ) end_ARG start_ARG ∂ italic_n start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT end_ARG + italic_O ( italic_n start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT ) (S6)

Substituting the charge-voltage relation of the varactor Q(Vn)=C0(VnbVn2)𝑄subscript𝑉𝑛subscript𝐶0subscript𝑉𝑛𝑏superscriptsubscript𝑉𝑛2Q(V_{n})=C_{0}(V_{n}-bV_{n}^{2})italic_Q ( italic_V start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) = italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_V start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT - italic_b italic_V start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) in Eq. (S4), we find that propagation in the NLTL in the long-wavelength limit is described at leading order by the following partial differential equation Ricketts and Ham (2018):

Vttb(V2)tt=v02(Vnn+Vnnnn12)+RLVnnt,subscript𝑉𝑡𝑡𝑏subscriptsuperscript𝑉2𝑡𝑡superscriptsubscript𝑣02subscript𝑉𝑛𝑛subscript𝑉𝑛𝑛𝑛𝑛12𝑅𝐿subscript𝑉𝑛𝑛𝑡V_{tt}-b(V^{2})_{tt}=v_{0}^{2}\left(V_{nn}+\frac{V_{nnnn}}{12}\right)+\frac{R}% {L}V_{nnt},italic_V start_POSTSUBSCRIPT italic_t italic_t end_POSTSUBSCRIPT - italic_b ( italic_V start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) start_POSTSUBSCRIPT italic_t italic_t end_POSTSUBSCRIPT = italic_v start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_V start_POSTSUBSCRIPT italic_n italic_n end_POSTSUBSCRIPT + divide start_ARG italic_V start_POSTSUBSCRIPT italic_n italic_n italic_n italic_n end_POSTSUBSCRIPT end_ARG start_ARG 12 end_ARG ) + divide start_ARG italic_R end_ARG start_ARG italic_L end_ARG italic_V start_POSTSUBSCRIPT italic_n italic_n italic_t end_POSTSUBSCRIPT , (S7)

where v0=1/LC0subscript𝑣01𝐿subscript𝐶0v_{0}=1/\sqrt{LC_{0}}italic_v start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 1 / square-root start_ARG italic_L italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG. The subscripts in Eq. (S7) denote derivation with respect to n𝑛nitalic_n and t𝑡titalic_t. Following ref. Ricketts and Ham (2018) a small dimensionless parameter α𝛼\alphaitalic_α is introduced for scaling linear and nonlinear effects. Changing V𝑉Vitalic_V into αV𝛼𝑉\alpha Vitalic_α italic_V, R/L𝑅𝐿R/Litalic_R / italic_L into α1/2R/Lsuperscript𝛼12𝑅𝐿\alpha^{1/2}R/Litalic_α start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT italic_R / italic_L and introducing the following new variables:

{s=α1/2(n+(α1)v0t),τ=α3/2t,cases𝑠superscript𝛼12𝑛𝛼1subscript𝑣0𝑡otherwise𝜏superscript𝛼32𝑡otherwise\begin{cases}s=\alpha^{1/2}(n+(\alpha-1)v_{0}t),\\ \tau=\alpha^{3/2}t,\end{cases}{ start_ROW start_CELL italic_s = italic_α start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT ( italic_n + ( italic_α - 1 ) italic_v start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_t ) , end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL italic_τ = italic_α start_POSTSUPERSCRIPT 3 / 2 end_POSTSUPERSCRIPT italic_t , end_CELL start_CELL end_CELL end_ROW

we obtain the following equation that describes the nonlinear propagation of waves in the NLTL in physical variables

Vτ+v0Vs+bv0VVs+v0243Vs3=R2L2Vs2.𝑉𝜏subscript𝑣0𝑉𝑠𝑏subscript𝑣0𝑉𝑉𝑠subscript𝑣024superscript3𝑉superscript𝑠3𝑅2𝐿superscript2𝑉superscript𝑠2\frac{\partial V}{\partial\tau}+v_{0}\frac{\partial V}{\partial s}+bv_{0}V% \frac{\partial V}{\partial s}+\frac{v_{0}}{24}\frac{\partial^{3}V}{\partial s^% {3}}=\frac{R}{2L}\frac{\partial^{2}V}{\partial s^{2}}.divide start_ARG ∂ italic_V end_ARG start_ARG ∂ italic_τ end_ARG + italic_v start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT divide start_ARG ∂ italic_V end_ARG start_ARG ∂ italic_s end_ARG + italic_b italic_v start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_V divide start_ARG ∂ italic_V end_ARG start_ARG ∂ italic_s end_ARG + divide start_ARG italic_v start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG 24 end_ARG divide start_ARG ∂ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT italic_V end_ARG start_ARG ∂ italic_s start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG = divide start_ARG italic_R end_ARG start_ARG 2 italic_L end_ARG divide start_ARG ∂ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_V end_ARG start_ARG ∂ italic_s start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG . (S8)

Considering that Vs1v0Vτsimilar-tosubscript𝑉𝑠1subscript𝑣0subscript𝑉𝜏V_{s}\sim-\frac{1}{v_{0}}V_{\tau}italic_V start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ∼ - divide start_ARG 1 end_ARG start_ARG italic_v start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG italic_V start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT, the equation describing the space-time evolution of the voltage along the LC oscillator chain reads:

Vs+1v0Vτbc0VVτ124v03VτττR2Lc03Vττ=0.subscript𝑉𝑠1subscript𝑣0subscript𝑉𝜏𝑏subscript𝑐0𝑉subscript𝑉𝜏124superscriptsubscript𝑣03subscript𝑉𝜏𝜏𝜏𝑅2𝐿superscriptsubscript𝑐03subscript𝑉𝜏𝜏0V_{s}+\frac{1}{v_{0}}V_{\tau}-\frac{b}{c_{0}}VV_{\tau}-\frac{1}{24v_{0}^{3}}V_% {\tau\tau\tau}-\frac{R}{2Lc_{0}^{3}}V_{\tau\tau}=0.italic_V start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT + divide start_ARG 1 end_ARG start_ARG italic_v start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG italic_V start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT - divide start_ARG italic_b end_ARG start_ARG italic_c start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG italic_V italic_V start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT - divide start_ARG 1 end_ARG start_ARG 24 italic_v start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG italic_V start_POSTSUBSCRIPT italic_τ italic_τ italic_τ end_POSTSUBSCRIPT - divide start_ARG italic_R end_ARG start_ARG 2 italic_L italic_c start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG italic_V start_POSTSUBSCRIPT italic_τ italic_τ end_POSTSUBSCRIPT = 0 . (S9)

Finally introducing the following normalized space and time variables together with the normalized field u(x,t)𝑢𝑥𝑡u(x,t)italic_u ( italic_x , italic_t )

{t=13b3V03s,x=2v0bV0(τsv0),u=VV0,cases𝑡13superscript𝑏3superscriptsubscript𝑉03𝑠otherwise𝑥2subscript𝑣0𝑏subscript𝑉0𝜏𝑠subscript𝑣0otherwise𝑢𝑉subscript𝑉0otherwise\begin{cases}t=\frac{1}{3}\sqrt{b^{3}V_{0}^{3}}s,\\ x=-2v_{0}\sqrt{bV_{0}}\left(\tau-\frac{s}{v_{0}}\right),\\ u=\frac{V}{V_{0}},\end{cases}{ start_ROW start_CELL italic_t = divide start_ARG 1 end_ARG start_ARG 3 end_ARG square-root start_ARG italic_b start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG italic_s , end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL italic_x = - 2 italic_v start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT square-root start_ARG italic_b italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ( italic_τ - divide start_ARG italic_s end_ARG start_ARG italic_v start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ) , end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL italic_u = divide start_ARG italic_V end_ARG start_ARG italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG , end_CELL start_CELL end_CELL end_ROW (S10)

we obtain the KdV-Burgers equation:

ut+6uux+uxxx=ϵuxx,subscript𝑢𝑡6𝑢subscript𝑢𝑥subscript𝑢𝑥𝑥𝑥italic-ϵsubscript𝑢𝑥𝑥u_{t}+6uu_{x}+u_{xxx}=\epsilon u_{xx},italic_u start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT + 6 italic_u italic_u start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT + italic_u start_POSTSUBSCRIPT italic_x italic_x italic_x end_POSTSUBSCRIPT = italic_ϵ italic_u start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT , (S11)

where ϵ=6R/(Lv0bV0)italic-ϵ6𝑅𝐿subscript𝑣0𝑏subscript𝑉0\epsilon=6R/(Lv_{0}\sqrt{bV_{0}})italic_ϵ = 6 italic_R / ( italic_L italic_v start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT square-root start_ARG italic_b italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ), V0subscript𝑉0V_{0}italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT being the value of the voltage used for normalization.

Fig. S2 compares the space-time diagram recorded in the experiment (left column) with the space-time diagram converted to dimensionless units (right column) using the transformations specified by Eq. (S10). The initial condition is identical to the one used in the Letter: it is a soliton gas (SG) parameterized by 200200200200 discrete eigenvalues ηksubscript𝜂𝑘\eta_{k}italic_η start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT that are randomly distributed between 0.20.20.20.2 and 0.60.60.60.6 in an uniform way. After the step of the numerical spectral synthesis of the SG (see Sec. II), it is generated by a programmable electric waveform generator having a bandwidth of 100100100100 MHz and a sampling rate of 1111 GS/s. The SG has an initial duration of 28μsimilar-toabsent28𝜇\sim 28\mu∼ 28 italic_μs. The typical timescale associated with voltage fluctuations, which are linked to the randomness of the SG, is around 150150150150 ns. The experimental space-diagram is recorded by measuring the voltage every 5555 LC cells using an oscilloscope having a bandwidth of 200200200200 MHz.

The normalized time interval over which the evolution of the SG is observed is as large as 420420420420. Such an extended evolution time is achieved not in a single pass through the NLTL, but in 6666 passes. Each pass corresponds to a normalized time interval Δt=70Δ𝑡70\Delta t=70roman_Δ italic_t = 70, which is equivalent to physical propagation through the 160160160160 LC cells composing the NLTL. To achieve multiple passes in the NLTL, the electrical voltage recorded at the final LC cell after each pass is used as the initial condition for the subsequent pass.

Due to the small series resistance of the inductors (0.09Ωsimilar-toabsent0.09Ω\sim 0.09\,\Omega∼ 0.09 roman_Ω), there is slight damping in the NLTL. Consequently, the mass, defined as M=u(x,t)𝑑x𝑀𝑢𝑥𝑡differential-d𝑥M=\int u(x,t)dxitalic_M = ∫ italic_u ( italic_x , italic_t ) italic_d italic_x, is not perfectly conserved in a single pass. In the experiment, the mass decays by 5%similar-toabsentpercent5\sim 5\%∼ 5 % over Δt=70Δ𝑡70\Delta t=70roman_Δ italic_t = 70. This mass decay is compensated for after each pass by a slight amplification of the signal recorded at the final LC cell before it is reinjected in the NLTL. This compensation results in small periodic oscillations of the mass, as shown in Fig. S2(c). However, the mass is well conserved on average over the entire evolution of the SG between t=0𝑡0t=0italic_t = 0 and t=420𝑡420t=420italic_t = 420.

Refer to caption
Figure S5: (a) Space-time diagram recorded in the NLTL by using a 200200200200 MHz oscilloscope. The voltage is measured every 5555 LC cells. n𝑛nitalic_n-index indicates the LC cell at which the voltage is probed. (b) Normalized experimental space-time diagram computed by using Eq. (S10). (c) Time evolution of the measured mass M(t)=u(x,t)𝑑x𝑀𝑡𝑢𝑥𝑡differential-d𝑥M(t)=\int u(x,t)dxitalic_M ( italic_t ) = ∫ italic_u ( italic_x , italic_t ) italic_d italic_x showing small oscillations due to the weak damping and its compensation at each pass in the NLTL. (d) Time evolution of the voltage measured at n=0𝑛0n=0italic_n = 0 and n=960𝑛960n=960italic_n = 960, after 6666 passes in the NLTL. (e) Normalized field u(x,t)𝑢𝑥𝑡u(x,t)italic_u ( italic_x , italic_t ) at t=0𝑡0t=0italic_t = 0 and t420similar-to𝑡420t\sim 420italic_t ∼ 420.

II Nonlinear spectral synthesis of soliton gases using the Darboux method

The algorithm to generate N-soliton solution of the KdV equation with N=200𝑁200N=200italic_N = 200 is based on the Darboux method. The Darboux transform is a recursive transformation scheme where the fundamental soliton solution of the KdV equation is used as a building block for the construction of higher-order N𝑁Nitalic_N-soliton solutions through the iterative addition of new discrete eigenvalues ηksubscript𝜂𝑘\eta_{k}italic_η start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT. The algorithm for the generation of KdV SG has been developed in ref. Congy et al. (2023). Refering the reader to ref. Congy et al. (2023); Liao and Huang (1996) for more details about the mathematical description of the Darboux machinary, we only provide here the algorithmic recipe to generate the N-soliton solution in a practical way.

- The first step consists is generating an ensemble of N𝑁Nitalic_N discrete eigenvalues ηksubscript𝜂𝑘\eta_{k}italic_η start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT (k=1N𝑘1𝑁k=1-Nitalic_k = 1 - italic_N) ranked in ascending order, from the smallest to the largest. An ensemble of N𝑁Nitalic_N position parameters x0ksubscript𝑥0𝑘x_{0k}italic_x start_POSTSUBSCRIPT 0 italic_k end_POSTSUBSCRIPT must also be created. For generating the SG used as initial condition in our experiment, the position parameters x0ksubscript𝑥0𝑘x_{0k}italic_x start_POSTSUBSCRIPT 0 italic_k end_POSTSUBSCRIPT are randomly and uniformely distributed over the interval [25,25]2525[-25,25][ - 25 , 25 ]. The spectral parameters ηksubscript𝜂𝑘\eta_{k}italic_η start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT associated with this SG are randomly distributed between 0.20.20.20.2 and 0.60.60.60.6 in an uniform way.

- A numerical box of size L𝐿Litalic_L and a given time t𝑡titalic_t must then be defined is such a way that the quantities:

Θk(x,t)=ηk(xx0k4ηk2t)subscriptΘ𝑘𝑥𝑡subscript𝜂𝑘𝑥subscript𝑥0𝑘4superscriptsubscript𝜂𝑘2𝑡\Theta_{k}(x,t)=\eta_{k}(x-x_{0k}-4\eta_{k}^{2}t)roman_Θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_x , italic_t ) = italic_η start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_x - italic_x start_POSTSUBSCRIPT 0 italic_k end_POSTSUBSCRIPT - 4 italic_η start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_t ) (S12)

can be computed for x[L/2,L/2]𝑥𝐿2𝐿2x\in[-L/2,L/2]italic_x ∈ [ - italic_L / 2 , italic_L / 2 ] for all the values of k𝑘kitalic_k (k=1N𝑘1𝑁k=1-Nitalic_k = 1 - italic_N).

- For each k𝑘kitalic_k between 1111 and N𝑁Nitalic_N, compute :

q1(x,t,ηk)=ηktanh(Θk(x,t))forkoddsubscript𝑞1𝑥𝑡subscript𝜂𝑘subscript𝜂𝑘subscriptΘ𝑘𝑥𝑡forkodd\displaystyle q_{1}(x,t,\eta_{k})=\eta_{k}\tanh(\Theta_{k}(x,t))\quad\mathrm{% for\,\,k\,\,odd}italic_q start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_x , italic_t , italic_η start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) = italic_η start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT roman_tanh ( roman_Θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_x , italic_t ) ) roman_for roman_k roman_odd (S13)
q1(x,t,ηk)=ηk(tanh(Θk(x,t)))1forkevensubscript𝑞1𝑥𝑡subscript𝜂𝑘subscript𝜂𝑘superscriptsubscriptΘ𝑘𝑥𝑡1forkeven\displaystyle q_{1}(x,t,\eta_{k})=\eta_{k}(\tanh(\Theta_{k}(x,t)))^{-1}\quad% \mathrm{for\,\,k\,\,even}italic_q start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_x , italic_t , italic_η start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) = italic_η start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( roman_tanh ( roman_Θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_x , italic_t ) ) ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_for roman_k roman_even

- For Nk2𝑁𝑘2N\geq k\geq 2italic_N ≥ italic_k ≥ 2, compute:

qk(x,t,ηk)=ηk2ηk12qk1(x,t,ηk)qk1(x,t,ηk1)qk1(x,t,ηk1)subscript𝑞𝑘𝑥𝑡subscript𝜂𝑘superscriptsubscript𝜂𝑘2superscriptsubscript𝜂𝑘12subscript𝑞𝑘1𝑥𝑡subscript𝜂𝑘subscript𝑞𝑘1𝑥𝑡subscript𝜂𝑘1subscript𝑞𝑘1𝑥𝑡subscript𝜂𝑘1q_{k}(x,t,\eta_{k})=\frac{\eta_{k}^{2}-\eta_{k-1}^{2}}{q_{k-1}(x,t,\eta_{k})-q% _{k-1}(x,t,\eta_{k-1})}-q_{k-1}(x,t,\eta_{k-1})italic_q start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_x , italic_t , italic_η start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) = divide start_ARG italic_η start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_η start_POSTSUBSCRIPT italic_k - 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_q start_POSTSUBSCRIPT italic_k - 1 end_POSTSUBSCRIPT ( italic_x , italic_t , italic_η start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) - italic_q start_POSTSUBSCRIPT italic_k - 1 end_POSTSUBSCRIPT ( italic_x , italic_t , italic_η start_POSTSUBSCRIPT italic_k - 1 end_POSTSUBSCRIPT ) end_ARG - italic_q start_POSTSUBSCRIPT italic_k - 1 end_POSTSUBSCRIPT ( italic_x , italic_t , italic_η start_POSTSUBSCRIPT italic_k - 1 end_POSTSUBSCRIPT ) (S14)

- For Nk2𝑁𝑘2N\geq k\geq 2italic_N ≥ italic_k ≥ 2, compute iteratively

uk(x,t)=uk1(x,t)2ηk2+2qk(x,t,ηk)2subscript𝑢𝑘𝑥𝑡subscript𝑢𝑘1𝑥𝑡2superscriptsubscript𝜂𝑘22subscript𝑞𝑘superscript𝑥𝑡subscript𝜂𝑘2u_{k}(x,t)=-u_{k-1}(x,t)-2\eta_{k}^{2}+2q_{k}(x,t,\eta_{k})^{2}italic_u start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_x , italic_t ) = - italic_u start_POSTSUBSCRIPT italic_k - 1 end_POSTSUBSCRIPT ( italic_x , italic_t ) - 2 italic_η start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 2 italic_q start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_x , italic_t , italic_η start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (S15)

by starting the iteration from u1(x,t)=2η12sech2(Θ1(x,t))subscript𝑢1𝑥𝑡2superscriptsubscript𝜂12superscriptsech2subscriptΘ1𝑥𝑡u_{1}(x,t)=-2\eta_{1}^{2}\,\text{sech}^{2}(\Theta_{1}(x,t))italic_u start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_x , italic_t ) = - 2 italic_η start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT sech start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( roman_Θ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_x , italic_t ) ). The N-soliton solution u(x,t)𝑢𝑥𝑡u(x,t)italic_u ( italic_x , italic_t ) of the KDV equation associated with the distribution of the spectral parameters ηksubscript𝜂𝑘\eta_{k}italic_η start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT initially chosen is obtained by changing uN(x,t)subscript𝑢𝑁𝑥𝑡u_{N}(x,t)italic_u start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT ( italic_x , italic_t ) into uN(x,t)subscript𝑢𝑁𝑥𝑡-u_{N}(x,t)- italic_u start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT ( italic_x , italic_t ).

III Solitary Wave Propagation, Mass Conservation, and Continuous Spectrum Excitation in the KdVB Equation

The question of the evolution of a KdV soliton under the influence of dissipative perturbations has been extensively studied in the literature Madsen and Mei (1969); Grimshaw (1970); Johnson (1973); Kaup et al. (1978); Kodama and Ablowitz (1981); Kivshar and Malomed (1989); Grimshaw et al. (1994); Johnson (1994); Ko and Kuehl (1978); Knickerbocker and Newell (1980, 1985); El and Grimshaw (2002). In particular, the concept that new discrete eigenvalues can be generated due to the excitation of continuous spectrum radiation has been discussed in references Kaup et al. (1978); Wright (1980); Newell (1985) in the context of solitary wave propagation in shallow water with slowly varying depth. In this section, we present numerical simulations to describe the long-term evolution of a solitary wave, aiming to synthesize and illustrate the nonadiabatic features previously identified in the literature within the framework of KdVB dynamics.

Refer to caption
Figure S6: Numerical simulations of Eq. (S16) with ϵ=0.1italic-ϵ0.1\epsilon=0.1italic_ϵ = 0.1. The initial condition is a soliton given by u(x,t=0)=2η12sech2(η1x)𝑢𝑥𝑡02superscriptsubscript𝜂12𝑠𝑒𝑐superscript2subscript𝜂1𝑥u(x,t=0)=2\eta_{1}^{2}\,sech^{2}(\eta_{1}\,x)italic_u ( italic_x , italic_t = 0 ) = 2 italic_η start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_s italic_e italic_c italic_h start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_η start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_x ) with η1=0.707subscript𝜂10.707\eta_{1}=0.707italic_η start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 0.707. (a) Field u(x,t)𝑢𝑥𝑡u(x,t)italic_u ( italic_x , italic_t ) numerically computed at t=0𝑡0t=0italic_t = 0 (blue line) and at t=100𝑡100t=100italic_t = 100 (orange line). (b) Space-time plot showing the evolution of the solitary wave. (c) Corresponding time evolution of the mass M𝑀Mitalic_M, of the momentum P𝑃Pitalic_P and of the energy E𝐸Eitalic_E. (d) Time evolution of the discrete eigenvalues successively created during the time evolution of the solitary wave. (e) Time evolution of the discrete masses defined as mi=4ηisubscript𝑚𝑖4subscript𝜂𝑖m_{i}=4\eta_{i}italic_m start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 4 italic_η start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and of the discrete mass MDS=i=154ηisubscript𝑀𝐷𝑆superscriptsubscript𝑖154subscript𝜂𝑖M_{DS}=\sum\limits_{i=1}^{5}4\eta_{i}italic_M start_POSTSUBSCRIPT italic_D italic_S end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT 4 italic_η start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. The dashed black line represents the total mass defined as M=u(x,t)𝑑x𝑀𝑢𝑥𝑡differential-d𝑥M=\int u(x,t)dxitalic_M = ∫ italic_u ( italic_x , italic_t ) italic_d italic_x. (f) Time evolution of the discrete mass MDSsubscript𝑀𝐷𝑆M_{DS}italic_M start_POSTSUBSCRIPT italic_D italic_S end_POSTSUBSCRIPT (black line) and of the mass of the continuous spectrum defined as Mcs=1π+log(1|b(ξ)|2|a(ξ)|2)𝑑ξsubscript𝑀𝑐𝑠1𝜋superscriptsubscript1superscript𝑏𝜉2superscript𝑎𝜉2differential-d𝜉M_{cs}=\frac{1}{\pi}\int_{-\infty}^{+\infty}\log\left(1-\frac{|b(\xi)|^{2}}{|a% (\xi)|^{2}}\right)d\xiitalic_M start_POSTSUBSCRIPT italic_c italic_s end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG italic_π end_ARG ∫ start_POSTSUBSCRIPT - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + ∞ end_POSTSUPERSCRIPT roman_log ( 1 - divide start_ARG | italic_b ( italic_ξ ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG | italic_a ( italic_ξ ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) italic_d italic_ξ. The red line is the total mass defined as M=MDS+MCS𝑀subscript𝑀𝐷𝑆subscript𝑀𝐶𝑆M=M_{DS}+M_{CS}italic_M = italic_M start_POSTSUBSCRIPT italic_D italic_S end_POSTSUBSCRIPT + italic_M start_POSTSUBSCRIPT italic_C italic_S end_POSTSUBSCRIPT according to Eq. (S22).

As in the Letter, we consider the KdVB equation

ut+6uux+uxxx=ϵuxx,subscript𝑢𝑡6𝑢subscript𝑢𝑥subscript𝑢𝑥𝑥𝑥italic-ϵsubscript𝑢𝑥𝑥u_{t}+6uu_{x}+u_{xxx}=\epsilon u_{xx},italic_u start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT + 6 italic_u italic_u start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT + italic_u start_POSTSUBSCRIPT italic_x italic_x italic_x end_POSTSUBSCRIPT = italic_ϵ italic_u start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT , (S16)

and its associated Schrödinger equation in the IST theory

ψxx+(u(x,t)λ)ψ=0.subscript𝜓𝑥𝑥𝑢𝑥𝑡𝜆𝜓0\psi_{xx}+(u(x,t)-\lambda)\psi=0.italic_ψ start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT + ( italic_u ( italic_x , italic_t ) - italic_λ ) italic_ψ = 0 . (S17)

The mass M𝑀Mitalic_M of the field u(x,t)𝑢𝑥𝑡u(x,t)italic_u ( italic_x , italic_t ) is defined as :

M=+u(x,t)𝑑x.𝑀superscriptsubscript𝑢𝑥𝑡differential-d𝑥M=\int_{-\infty}^{+\infty}u(x,t)dx.italic_M = ∫ start_POSTSUBSCRIPT - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + ∞ end_POSTSUPERSCRIPT italic_u ( italic_x , italic_t ) italic_d italic_x . (S18)

The mass M𝑀Mitalic_M is a conserved quantity both in KdV and in KdVB equations: dM/dt=0𝑑𝑀𝑑𝑡0dM/dt=0italic_d italic_M / italic_d italic_t = 0. The momentum P𝑃Pitalic_P is defined as :

P=+u2(x,t)𝑑x.𝑃superscriptsubscriptsuperscript𝑢2𝑥𝑡differential-d𝑥P=\int_{-\infty}^{+\infty}u^{2}(x,t)dx.italic_P = ∫ start_POSTSUBSCRIPT - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + ∞ end_POSTSUPERSCRIPT italic_u start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_x , italic_t ) italic_d italic_x . (S19)

The momentum P𝑃Pitalic_P is a conserved quantity in KdV equation (dP/dt=0𝑑𝑃𝑑𝑡0dP/dt=0italic_d italic_P / italic_d italic_t = 0) but it is not a conserved quantity in KdVB equation. For the KdVB equation, it is easy to show that :

dPdt=2ϵ+ux2(x,t)𝑑x,𝑑𝑃𝑑𝑡2italic-ϵsuperscriptsubscriptsuperscriptsubscript𝑢𝑥2𝑥𝑡differential-d𝑥\frac{dP}{dt}=-2\epsilon\int_{-\infty}^{+\infty}u_{x}^{2}(x,t)dx,divide start_ARG italic_d italic_P end_ARG start_ARG italic_d italic_t end_ARG = - 2 italic_ϵ ∫ start_POSTSUBSCRIPT - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + ∞ end_POSTSUPERSCRIPT italic_u start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_x , italic_t ) italic_d italic_x , (S20)

which means that the rate at which this quantity decays is determined by the gradients uxsubscript𝑢𝑥u_{x}italic_u start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT of the field in space. The energy E𝐸Eitalic_E is defined as :

E=+(u3(x,t)ux2(x,t)2)𝑑x.𝐸superscriptsubscriptsuperscript𝑢3𝑥𝑡superscriptsubscript𝑢𝑥2𝑥𝑡2differential-d𝑥E=\int_{-\infty}^{+\infty}\left(u^{3}(x,t)-\frac{u_{x}^{2}(x,t)}{2}\right)dx.italic_E = ∫ start_POSTSUBSCRIPT - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + ∞ end_POSTSUPERSCRIPT ( italic_u start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ( italic_x , italic_t ) - divide start_ARG italic_u start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_x , italic_t ) end_ARG start_ARG 2 end_ARG ) italic_d italic_x . (S21)

The energy is a conserved quantity in KdV but not in KdVB. In KdVB, the energy changes in time but does not necessarily decrease.

In the IST formalism, the mass M𝑀Mitalic_M and the momentum P𝑃Pitalic_P have the following expressions:

M=4kηk+1π+log(1|b(ξ)|2|a(ξ)|2)𝑑ξ𝑀4subscript𝑘subscript𝜂𝑘1𝜋superscriptsubscript1superscript𝑏𝜉2superscript𝑎𝜉2differential-d𝜉M=4\sum_{k}\eta_{k}+\frac{1}{\pi}\int_{-\infty}^{+\infty}\log\left(1-\frac{|b(% \xi)|^{2}}{|a(\xi)|^{2}}\right)d\xiitalic_M = 4 ∑ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_η start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT + divide start_ARG 1 end_ARG start_ARG italic_π end_ARG ∫ start_POSTSUBSCRIPT - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + ∞ end_POSTSUPERSCRIPT roman_log ( 1 - divide start_ARG | italic_b ( italic_ξ ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG | italic_a ( italic_ξ ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) italic_d italic_ξ (S22)
P=163kηk34π+ξ2log(1|b(ξ)|2|a(ξ)|2)𝑑ξ𝑃163subscript𝑘superscriptsubscript𝜂𝑘34𝜋superscriptsubscriptsuperscript𝜉21superscript𝑏𝜉2superscript𝑎𝜉2differential-d𝜉P=\frac{16}{3}\sum_{k}\eta_{k}^{3}-\frac{4}{\pi}\int_{-\infty}^{+\infty}\xi^{2% }\log\left(1-\frac{|b(\xi)|^{2}}{|a(\xi)|^{2}}\right)d\xiitalic_P = divide start_ARG 16 end_ARG start_ARG 3 end_ARG ∑ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_η start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT - divide start_ARG 4 end_ARG start_ARG italic_π end_ARG ∫ start_POSTSUBSCRIPT - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + ∞ end_POSTSUPERSCRIPT italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_log ( 1 - divide start_ARG | italic_b ( italic_ξ ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG | italic_a ( italic_ξ ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) italic_d italic_ξ (S23)

where ηksubscript𝜂𝑘\eta_{k}italic_η start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT are the discrete eigenvalues that can be computed by solving Eq. (S17) for the potential u(x,t)𝑢𝑥𝑡u(x,t)italic_u ( italic_x , italic_t ). ρ(ξ)=b(ξ)/a(ξ)𝜌𝜉𝑏𝜉𝑎𝜉\rho(\xi)=b(\xi)/a(\xi)italic_ρ ( italic_ξ ) = italic_b ( italic_ξ ) / italic_a ( italic_ξ ) is the reflection coefficient defined in the IST theory, see e.g. ref. Kaup et al. (1978); Wright (1980); Newell (1985). In Eq. (S22) (resp. Eq. (S23)), the discrete summation over the ηksubscript𝜂𝑘\eta_{k}italic_η start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT (resp. ηk3superscriptsubscript𝜂𝑘3\eta_{k}^{3}italic_η start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT) is related to the discrete part of the spectrum. On the other hand the integral term in Eq. (S22) (resp. (S23)) represents the contribution to the mass (resp. to the momentum) of the continuous part of the spectrum that is associated with the radiative modes.

Fig. S3 shows numerical simulations of Eq. (S16) with ϵ=0.1italic-ϵ0.1\epsilon=0.1italic_ϵ = 0.1 for an initial condition being under the form of a soliton given by u(x,t=0)=2η12sech2(η1x)𝑢𝑥𝑡02superscriptsubscript𝜂12𝑠𝑒𝑐superscript2subscript𝜂1𝑥u(x,t=0)=2\eta_{1}^{2}\,sech^{2}(\eta_{1}\,x)italic_u ( italic_x , italic_t = 0 ) = 2 italic_η start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_s italic_e italic_c italic_h start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_η start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_x ) with η1=0.707subscript𝜂10.707\eta_{1}=0.707italic_η start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 0.707, like in the experiment shown in Fig. 2 of the Letter. The maximum evolution time reached in the numerical simulation shown in Fig. S3 is 100100100100, which is significantly larger than the experimental evolution time (t16similar-to-or-equals𝑡16t\simeq 16italic_t ≃ 16). The presence of diffusion described by the term on the right-hand side of Eq. (S16) induces a slow decay in the soliton’s amplitude and a gradual decrease in its velocity, as illustrated in Fig. S3(a)(b). This evolution is accompanied by the formation of a shelf and of small oscillations behind the solitary wave. The mass M𝑀Mitalic_M is conserved during the evolution while the momentum P𝑃Pitalic_P and the energy E𝐸Eitalic_E decrease over time, see Fig. S3(c).

Fig. S3(d) shows that the amplitude η1subscript𝜂1\eta_{1}italic_η start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT of the discrete eigenvalue linked to the initial condition gradually decays over time, while 4444 new discrete eigenvalues are created during the shelf formation. Fig. S3(d) shows the evolution of the masses mi=4ηisubscript𝑚𝑖4subscript𝜂𝑖m_{i}=4\eta_{i}italic_m start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 4 italic_η start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT of each of the 4444 discrete eigenmodes, see Eq. (S22). Of course the discrete masses mi(t)=4ηi(t)subscript𝑚𝑖𝑡4subscript𝜂𝑖𝑡m_{i}(t)=4\eta_{i}(t)italic_m start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) = 4 italic_η start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) (i=14)𝑖14(i=1-4)( italic_i = 1 - 4 ) follow the same time evolution as the discrete eigenvalues ηi(t)subscript𝜂𝑖𝑡\eta_{i}(t)italic_η start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ). The interesting point is that the sum of the discrete masses MDS(t)=i=184ηi(t)subscript𝑀𝐷𝑆𝑡superscriptsubscript𝑖184subscript𝜂𝑖𝑡M_{DS}(t)=\sum\limits_{i=1}^{8}4\eta_{i}(t)italic_M start_POSTSUBSCRIPT italic_D italic_S end_POSTSUBSCRIPT ( italic_t ) = ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT 4 italic_η start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) is not constant but slightly fluctuates above the constant value of the total mass M=u(x,t)𝑑x𝑀𝑢𝑥𝑡differential-d𝑥M=\int u(x,t)dxitalic_M = ∫ italic_u ( italic_x , italic_t ) italic_d italic_x, see the full and dotted black lines in Fig. S3(e).

The decay in time of η1subscript𝜂1\eta_{1}italic_η start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT is associated to a decrease in mass, which must be compensated by an increase to preserve the fact that the total mass M=u(x,t)𝑑x𝑀𝑢𝑥𝑡differential-d𝑥M=\int u(x,t)dxitalic_M = ∫ italic_u ( italic_x , italic_t ) italic_d italic_x must be conserved Kivshar and Malomed (1989). The missing mass is provided by the generation of the new discrete eigenvalue η2subscript𝜂2\eta_{2}italic_η start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT. However the decay of η1(t)subscript𝜂1𝑡\eta_{1}(t)italic_η start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_t ) and the growing of η2(t)subscript𝜂2𝑡\eta_{2}(t)italic_η start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_t ) can never be balanced in such an exact way that the mass associated with these two discrete eigenvalues is constant in time : m1,2(t)=m1(t)+m2(t)=4(η1(t)+η2(t))ctesubscript𝑚12𝑡subscript𝑚1𝑡subscript𝑚2𝑡4subscript𝜂1𝑡subscript𝜂2𝑡𝑐𝑡𝑒m_{1,2}(t)=m_{1}(t)+m_{2}(t)=4(\eta_{1}(t)+\eta_{2}(t))\neq cteitalic_m start_POSTSUBSCRIPT 1 , 2 end_POSTSUBSCRIPT ( italic_t ) = italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_t ) + italic_m start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_t ) = 4 ( italic_η start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_t ) + italic_η start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_t ) ) ≠ italic_c italic_t italic_e. For the total mass conservation condition dM/dt=0𝑑𝑀𝑑𝑡0dM/dt=0italic_d italic_M / italic_d italic_t = 0 to be fulfilled, there is no choice but for the mass Mcssubscript𝑀𝑐𝑠M_{cs}italic_M start_POSTSUBSCRIPT italic_c italic_s end_POSTSUBSCRIPT associated with the radiative modes to fluctuate in the same way as Mdsubscript𝑀𝑑M_{d}italic_M start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT but with an opposite sign (MCS=1π+log(1|b(ξ)|2|a(ξ)|2)𝑑ξsubscript𝑀𝐶𝑆1𝜋superscriptsubscript1superscript𝑏𝜉2superscript𝑎𝜉2differential-d𝜉M_{CS}=\frac{1}{\pi}\int_{-\infty}^{+\infty}\log\left(1-\frac{|b(\xi)|^{2}}{|a% (\xi)|^{2}}\right)d\xiitalic_M start_POSTSUBSCRIPT italic_C italic_S end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG italic_π end_ARG ∫ start_POSTSUBSCRIPT - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + ∞ end_POSTSUPERSCRIPT roman_log ( 1 - divide start_ARG | italic_b ( italic_ξ ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG | italic_a ( italic_ξ ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) italic_d italic_ξ). This is illustrated in Figure S3(f), which shows that the negative mass fluctuations of the continuous spectrum exactly compensate for the positive fluctuations of the discrete mass MDSsubscript𝑀𝐷𝑆M_{DS}italic_M start_POSTSUBSCRIPT italic_D italic_S end_POSTSUBSCRIPT, so that the total mass M=MDS+MCS𝑀subscript𝑀𝐷𝑆subscript𝑀𝐶𝑆M=M_{DS}+M_{CS}italic_M = italic_M start_POSTSUBSCRIPT italic_D italic_S end_POSTSUBSCRIPT + italic_M start_POSTSUBSCRIPT italic_C italic_S end_POSTSUBSCRIPT is constant over time.

IV Soliton condensates of genus 0 and genus 1

The notion of soliton condensate as the ‘densest possible soliton gas’ for a given spectral support was introduced in El and Tovbis (2020) in the context of the focusing NLS equation. A detailed theory of soliton condensates for the KdV equation has been developed in Congy et al. (2023). Here, we refer the reader to these papers and only provide the mathematical expressions for the densities of state (DOS) of soliton condensates of genus 00 and 1111. We also show that these mathematical expressions fit very well the DOS found in numerical simulations reported in Fig. 3 of the Letter.

Refer to caption
Figure S7: (a) The blue crosses represents the spectral distribution shown in the inset of Fig. 3(f). The orange line fits the blue crosses with the model (S24) with λ1=0.533subscript𝜆10.533\lambda_{1}=0.533italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 0.533. (b) The blue crosses represents the spectral distribution shown in the inset of Fig. 3(l). The orange line fits the blue crosses with the model (S25) with λ1=0.131subscript𝜆10.131\lambda_{1}=0.131italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 0.131, λ2=0.295subscript𝜆20.295\lambda_{2}=0.295italic_λ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 0.295, λ3=0.359subscript𝜆30.359\lambda_{3}=0.359italic_λ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = 0.359.

The DOS of a KdV soliton condensate of genus 00 is given by the so-called Weyl distribution that reads:

f(η)=ηπλ12η2;η[0,λ1].formulae-sequence𝑓𝜂𝜂𝜋superscriptsubscript𝜆12superscript𝜂2𝜂0subscript𝜆1f(\eta)=\frac{\eta}{\pi\sqrt{\lambda_{1}^{2}-\eta^{2}}};\qquad\eta\in[0,% \lambda_{1}].italic_f ( italic_η ) = divide start_ARG italic_η end_ARG start_ARG italic_π square-root start_ARG italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_η start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_ARG ; italic_η ∈ [ 0 , italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ] . (S24)

Fig. S4(a) shows that the DOS of the SG measured at t=420𝑡420t=420italic_t = 420 in the electrical NLTL (ϵ=0.1italic-ϵ0.1\epsilon=0.1italic_ϵ = 0.1) is very well fitted by the Weyl distribution with λ1=0.533subscript𝜆10.533\lambda_{1}=0.533italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 0.533, see also Fig. 3(f) in the Letter.

The DOS of a KdV soliton condensate of genus 1111 lives on a spectral support not composed of one band but of two bands with their endpoints being given by the three real parameters λ1subscript𝜆1\lambda_{1}italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, λ2subscript𝜆2\lambda_{2}italic_λ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, λ3subscript𝜆3\lambda_{3}italic_λ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT. The DOS of the soliton condensate of genus 1 reads

f(η)=iη(η2w2)π(η2λ12)(η2λ22)(η2λ32);η[0,λ1][λ2,λ3]formulae-sequence𝑓𝜂𝑖𝜂superscript𝜂2superscript𝑤2𝜋superscript𝜂2superscriptsubscript𝜆12superscript𝜂2superscriptsubscript𝜆22superscript𝜂2superscriptsubscript𝜆32𝜂0subscript𝜆1subscript𝜆2subscript𝜆3f(\eta)=\frac{i\eta(\eta^{2}-w^{2})}{\pi\sqrt{(\eta^{2}-\lambda_{1}^{2})(\eta^% {2}-\lambda_{2}^{2})(\eta^{2}-\lambda_{3}^{2})}};\qquad\eta\in[0,\lambda_{1}]% \cup[\lambda_{2},\lambda_{3}]italic_f ( italic_η ) = divide start_ARG italic_i italic_η ( italic_η start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_w start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) end_ARG start_ARG italic_π square-root start_ARG ( italic_η start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) ( italic_η start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_λ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) ( italic_η start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_λ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) end_ARG end_ARG ; italic_η ∈ [ 0 , italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ] ∪ [ italic_λ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ] (S25)

with w2=λ32(λ32λ12)χ(m)superscript𝑤2superscriptsubscript𝜆32superscriptsubscript𝜆32superscriptsubscript𝜆12𝜒𝑚w^{2}=\lambda_{3}^{2}-(\lambda_{3}^{2}-\lambda_{1}^{2})\chi(m)italic_w start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = italic_λ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - ( italic_λ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) italic_χ ( italic_m ), χ(m)=E(m)/K(m)𝜒𝑚𝐸𝑚𝐾𝑚\chi(m)=E(m)/K(m)italic_χ ( italic_m ) = italic_E ( italic_m ) / italic_K ( italic_m ) and m=(λ22λ12)/(λ32λ12)𝑚superscriptsubscript𝜆22superscriptsubscript𝜆12superscriptsubscript𝜆32superscriptsubscript𝜆12m=(\lambda_{2}^{2}-\lambda_{1}^{2})/(\lambda_{3}^{2}-\lambda_{1}^{2})italic_m = ( italic_λ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) / ( italic_λ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ). E(m)𝐸𝑚E(m)italic_E ( italic_m ) (resp. K(m)𝐾𝑚K(m)italic_K ( italic_m )) is the complete elliptic integral of the second (resp. first) kind.

Fig. S4(b) shows that the DOS of the SG measured at (x=2900𝑥2900x=2900italic_x = 2900, t=4000𝑡4000t=4000italic_t = 4000) in the numerical simulations of the KdvB equation with small diffusion (ϵ=2.103italic-ϵsuperscript2.103\epsilon=2.10^{-3}italic_ϵ = 2.10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT) is remarkably well fitted by Eq. (S25) with λ1=0.131subscript𝜆10.131\lambda_{1}=0.131italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 0.131, λ2=0.295subscript𝜆20.295\lambda_{2}=0.295italic_λ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 0.295, λ3=0.359subscript𝜆30.359\lambda_{3}=0.359italic_λ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = 0.359, see also Fig. 3(l) in the Letter. It was conjectured and numerically verified in Congy et al. (2023) that realizations of the genus 1111 condensate with DOS (S25) almost surely coincide with periodic wave solutions of KdV parameterized by λ1subscript𝜆1\lambda_{1}italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, λ2subscript𝜆2\lambda_{2}italic_λ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT and λ3subscript𝜆3\lambda_{3}italic_λ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT. These solutions locally describe diffusive-dispersive shock waves of the KdVB equation El et al. (2017).