Mechanisms of unstable blowup in a quadratic nonlinear Schrödinger equation

Jonathan Jaquette 111Department of Mathematical Sciences, New Jersey Institute of Technology; Newark, NJ, USA; [email protected] 222Department of Mathematics and Statistics, Boston University; Boston, MA, USA
Abstract

In the work Cho et al. [Jpn. J. Ind. Appl. Math. 33 (2016): 145-166] the authors conjecture that the quadratic nonlinear Schrödinger equation (NLS) iut=uxx+u2𝑖subscript𝑢𝑡subscript𝑢𝑥𝑥superscript𝑢2iu_{t}=u_{xx}+u^{2}italic_i italic_u start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT = italic_u start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT + italic_u start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT for x𝕋𝑥𝕋x\in\mathbb{T}italic_x ∈ blackboard_T is globally well-posed for real initial data. While this may be true in a generic sense, we present evidence that this conjecture is false. Moreover we conjecture that the set of real initial data which blows up under the NLS dynamics is the stable manifold of the zero equilibrium for the nonlinear heat equation ut=uxx+u2subscript𝑢𝑡subscript𝑢𝑥𝑥superscript𝑢2u_{t}=u_{xx}+u^{2}italic_u start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT = italic_u start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT + italic_u start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT.


1 Introduction

Determining whether solutions to a differential equation can blow up in finite time is important for understanding the scope of validity for a given model. The blowup of solutions can be easy to detect through direct numerical simulation if it occurs for a large open set of initial data, but detection may be difficult if the blowup set is unstable and has finite co-dimension. Indeed, unstable objects are by definition sensitive to perturbation, and numerically simulating a blowup solution to a PDE causes both the temporal and spatial resolution to diverge as the blowup time is approached. Whether blowup occurs is an open question for important PDEs such as the Navier-Stokes equations [Fef00], and the development of techniques to find elusive blowup solutions is an active area of research [Pro22, WLGSB23, Hou23].

A classical PDE for studying blowup is given by the quadratic nonlinear heat equation:

utsubscript𝑢𝑡\displaystyle u_{t}italic_u start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT =uxx+u2absentsubscript𝑢𝑥𝑥superscript𝑢2\displaystyle=u_{xx}+u^{2}= italic_u start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT + italic_u start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (1)

On a periodic domain, if the average of the initial data is positive, a0=𝕋u>0subscript𝑎0subscript𝕋𝑢0a_{0}=\int_{\mathbb{T}}u>0italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = ∫ start_POSTSUBSCRIPT blackboard_T end_POSTSUBSCRIPT italic_u > 0, then the solution blows up. However, blowup becomes less generic when (1) is considered as a complex PDE as has been studied in models of vortex stretching [CLM85, GNSY13, HL08]. For a simpler example, consider the complex ODE and its solutions:

ztsubscript𝑧𝑡\displaystyle z_{t}italic_z start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT =z2,absentsuperscript𝑧2\displaystyle=z^{2},= italic_z start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , z(t)𝑧𝑡\displaystyle z(t)italic_z ( italic_t ) =z01+z0tabsentsubscript𝑧01subscript𝑧0𝑡\displaystyle=\frac{z_{0}}{1+z_{0}t}= divide start_ARG italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG 1 + italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_t end_ARG

for initial condition z(0)=z0𝑧0subscript𝑧0z(0)=z_{0}\in\mathbb{C}italic_z ( 0 ) = italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∈ blackboard_C. While the solution will blow up in forward (backwards) time if z0subscript𝑧0z_{0}italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is a positive (negative) real number, conversely if z0\subscript𝑧0\z_{0}\in\mathbb{C}\backslash\mathbb{R}italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∈ blackboard_C \ blackboard_R then it is globally well-posed.

To robustly analyze blowup solutions, one can make a self-similar ansatz for scaling the magnitude and spatial dependence of a solution as it approach blowup time [EF08, QS19]. This self-similar change of variables renormalizes the blowup problem into one of studying the self-similar dynamics of a new PDE produced by the change of variables. In the simplest case, a blowup profile can be equated with an equilibrium to a new PDE. However the self-similar dynamics need not be simple, and may in fact be periodic, or even chaotic [EF08, CM18].

An alternate methodology for understanding blowup was developed in Masuda’s work [Mas83, Mas84], where he considers continuing the solution in the complex plane of time around the blowup point, and shows for close to constant initial data there is a branching singularity. In this work, solving equation (1) in complex time is equivalent to solving the family of quadratic evolution equations:

utsubscript𝑢𝑡\displaystyle u_{t}italic_u start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT =eiθ(xxu+u2)absentsuperscript𝑒𝑖𝜃subscript𝑥𝑥𝑢superscript𝑢2\displaystyle=e^{i\theta}\left(\partial_{xx}u+u^{2}\right)= italic_e start_POSTSUPERSCRIPT italic_i italic_θ end_POSTSUPERSCRIPT ( ∂ start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT italic_u + italic_u start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) (2)

for θ[π/2,π/2]𝜃𝜋2𝜋2\theta\in[-\pi/2,\pi/2]italic_θ ∈ [ - italic_π / 2 , italic_π / 2 ] and for u:𝕋1:𝑢superscript𝕋1u:\mathbb{T}^{1}\to\mathbb{C}italic_u : blackboard_T start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT → blackboard_C. Here θ=0𝜃0\theta=0italic_θ = 0 corresponds to the original nonlinear heat equation, θ=π/4𝜃𝜋4\theta=\pi/4italic_θ = italic_π / 4 is complex Ginzberg-Landau like, and θ=π/2𝜃𝜋2\theta=\pi/2italic_θ = italic_π / 2 is a nonlinear Schrödinger equation (NLS).

This approach was investigated numerically by [COS16], showing that large, not close-to-constant initial data also has a branching singularity. Furthermore, they observed that numerical solutions generically converge to zero, and posed the following conjecture:

Conjecture 1.1 ([COS16]).

The nonlinear Schrödinger equation:

iut𝑖subscript𝑢𝑡\displaystyle iu_{t}italic_i italic_u start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT =uxx+u2absentsubscript𝑢𝑥𝑥superscript𝑢2\displaystyle=u_{xx}+u^{2}= italic_u start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT + italic_u start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (3)

is globally well-posed for any real initial data, small or large.

Beyond convergence to zero, this nonlinear Schrödinger equation exhibits rich dynamical structure, possessing nontrivial equilibria, homoclinic orbits, heteroclinic orbits, and periodic orbits [JLT22a, FMS23, Jaq22, JLT22b]. Recently Fiedler and Stuke [FS24] studied (3) within the context of the 1-parameter family of PDEs:

utsubscript𝑢𝑡\displaystyle u_{t}italic_u start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT =eiθ(uxx+6u2λ)absentsuperscript𝑒𝑖𝜃subscript𝑢𝑥𝑥6superscript𝑢2𝜆\displaystyle=e^{i\theta}\left(u_{xx}+6u^{2}-\lambda\right)= italic_e start_POSTSUPERSCRIPT italic_i italic_θ end_POSTSUPERSCRIPT ( italic_u start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT + 6 italic_u start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_λ ) (4)

In short, this work demonstrates that real eternal solutions are not complex entire. They prove there exists real initial data to (4) which for the heat equation (θ=0)𝜃0(\theta=0)( italic_θ = 0 ) is a heteroclinic orbit existing globally in time, however for the NLS (θ=±π2(\theta=\pm\tfrac{\pi}{2}( italic_θ = ± divide start_ARG italic_π end_ARG start_ARG 2 end_ARG) it blows up in finite time. However, this analysis requires λ>0𝜆0\lambda>0italic_λ > 0 and does not resolve Conjecture 1.1.

In this paper we present evidence which suggests that Conjecture 1.1 is false. Furthermore we present an alternative in Conjecture 1.2, that blowup from real initial data occurs on a codimension-1 manifold.

Conjecture 1.2.

Let 𝒲s(0)C(𝕋1,)superscript𝒲𝑠0𝐶superscript𝕋1\mathcal{W}^{s}(0)\subseteq C(\mathbb{T}^{1},\mathbb{R})caligraphic_W start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT ( 0 ) ⊆ italic_C ( blackboard_T start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT , blackboard_R ) denote the stable manifold of the 00 equilibrium for the nonlinear heat equation (1). Initial data u0C(𝕋1,)subscript𝑢0𝐶superscript𝕋1u_{0}\in C(\mathbb{T}^{1},\mathbb{R})italic_u start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∈ italic_C ( blackboard_T start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT , blackboard_R ) is globally well-posed under (3) if and only if u0𝒲s(0)subscript𝑢0superscript𝒲𝑠0u_{0}\notin\mathcal{W}^{s}(0)italic_u start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∉ caligraphic_W start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT ( 0 ).

It is a well-known fact that if 𝒲𝒲\mathcal{W}caligraphic_W is the stable manifold of an ODE x˙=f(x)˙𝑥𝑓𝑥\dot{x}=f(x)over˙ start_ARG italic_x end_ARG = italic_f ( italic_x ), then 𝒲𝒲\mathcal{W}caligraphic_W is also the unstable manifold of x˙=f(x)˙𝑥𝑓𝑥\dot{x}=-f(x)over˙ start_ARG italic_x end_ARG = - italic_f ( italic_x ). While the trajectories may differ, the stable manifold of (1) remains invariant under the dynamics of (2) for all θ𝜃\thetaitalic_θ. If |θ|<π/2𝜃𝜋2|\theta|<\pi/2| italic_θ | < italic_π / 2 it will remain a stable manifold. But for θ=±π/2𝜃plus-or-minus𝜋2\theta=\pm\pi/2italic_θ = ± italic_π / 2 then 𝒲𝒲\mathcal{W}caligraphic_W becomes a submanifold of the center manifold of (3), and exhibits secular growth of solutions. In Section 3 we give a heuristic characterization of the real initial data which blows up versus globally well-posed.

In Section 4 we use self-similar variables to analyze the numerical blowup of real initial data, and compare it to the known case of blowup for monochromatic initial data [Jaq22]. While the two blowup profiles appear similar, they appear to obey different scaling rates. All source code is available at [Jaq24].

2 Evidence of blowup from real initial data

Numerical evidence has been suggestive that Conjecture 1.1 may be true in a generic sense, i.e. for “almost all” real initial data. Indeed, for real initial data which is close-to-constant (and arbitrarily large!), it has been shown that the solution will limit to zero in both forward and backwards time:

Theorem 2.1 (Theorem 1.3 [JLT22a]).

Fix a complex scalar z0=reiϕsubscript𝑧0𝑟superscript𝑒𝑖italic-ϕz_{0}=re^{i\phi}\in\mathbb{C}italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = italic_r italic_e start_POSTSUPERSCRIPT italic_i italic_ϕ end_POSTSUPERSCRIPT ∈ blackboard_C and a function w0:𝕋1:subscript𝑤0superscript𝕋1w_{0}:\mathbb{T}^{1}\to\mathbb{C}italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT : blackboard_T start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT → blackboard_C on a 2π/ω2𝜋𝜔2\pi/\omega2 italic_π / italic_ω periodic torus having summable Fourier coefficients, that is w0(x)=kakeikωxsubscript𝑤0𝑥subscript𝑘subscript𝑎𝑘superscript𝑒𝑖𝑘𝜔𝑥w_{0}(x)=\sum_{k\in\mathbb{Z}}a_{k}e^{ik\omega x}italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_x ) = ∑ start_POSTSUBSCRIPT italic_k ∈ blackboard_Z end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_i italic_k italic_ω italic_x end_POSTSUPERSCRIPT for a={ak}1𝑎subscript𝑎𝑘superscript1a=\{a_{k}\}\in\ell^{1}italic_a = { italic_a start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT } ∈ roman_ℓ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT. Let u(t)𝑢𝑡u(t)italic_u ( italic_t ) be the solution of (3) with initial data u0=z0+w0subscript𝑢0subscript𝑧0subscript𝑤0u_{0}=z_{0}+w_{0}italic_u start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, and suppose:

a1<eπ/2|z0|.subscriptnorm𝑎superscript1superscript𝑒𝜋2subscript𝑧0\|a\|_{\ell^{1}}<e^{-\pi/2}|z_{0}|.∥ italic_a ∥ start_POSTSUBSCRIPT roman_ℓ start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT end_POSTSUBSCRIPT < italic_e start_POSTSUPERSCRIPT - italic_π / 2 end_POSTSUPERSCRIPT | italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT | .

If 0ϕπ0italic-ϕ𝜋0\leq\phi\leq\pi0 ≤ italic_ϕ ≤ italic_π then limtu(t)=0subscript𝑡𝑢𝑡0\lim_{t\to-\infty}u(t)=0roman_lim start_POSTSUBSCRIPT italic_t → - ∞ end_POSTSUBSCRIPT italic_u ( italic_t ) = 0. If πϕ2π𝜋italic-ϕ2𝜋\pi\leq\phi\leq 2\piitalic_π ≤ italic_ϕ ≤ 2 italic_π then limt+u(t)=0subscript𝑡𝑢𝑡0\lim_{t\to+\infty}u(t)=0roman_lim start_POSTSUBSCRIPT italic_t → + ∞ end_POSTSUBSCRIPT italic_u ( italic_t ) = 0.

To prove Conjecture 1.1 one could hope to show that all real initial data limits to zero, however this is not necessary for proving global well-posedness. For example, solutions could limit to a nontrivial equilibrium or a periodic orbit. In [JLT22a] it is shown via computer assisted proofs that there exist at least two distinct families of equilibria to (3), with heteroclinic orbits linking the trivial and nontrivial equilibria. In [FS24] it is shown that the complex equilibria to (1) are given by the Weierstrass elliptic functions, and in [Jaq22] it is shown that small initial data supported on positive Fourier modes will yield periodic orbits of locked frequency.

In light of Theorem 2.1, to look for possible counter examples to Conjecture 1.1 one must consider initial data which is close to zero average. To that end, consider the 1-parameter family of real initial data:

u30(x;A)subscript𝑢30𝑥𝐴\displaystyle u_{30}(x;A)italic_u start_POSTSUBSCRIPT 30 end_POSTSUBSCRIPT ( italic_x ; italic_A ) =30cos(2πx)+Aabsent302𝜋𝑥𝐴\displaystyle=30\cos(2\pi x)+A= 30 roman_cos ( 2 italic_π italic_x ) + italic_A (5)

Note that due to the even nonlinearity and even initial data, the Fourier series of the solution can be expressed as a cosine series with complex, time-varying coefficients:

u(t,x)=nan(t)e2πinωx=a0(t)+2n=1an(t)cos(2πnωx)𝑢𝑡𝑥subscript𝑛subscript𝑎𝑛𝑡superscript𝑒2𝜋𝑖𝑛𝜔𝑥subscript𝑎0𝑡2superscriptsubscript𝑛1subscript𝑎𝑛𝑡2𝜋𝑛𝜔𝑥\displaystyle u(t,x)=\sum_{n\in\mathbb{Z}}a_{n}(t)e^{\frac{2\pi in}{\omega}x}=% a_{0}(t)+2\sum_{n=1}^{\infty}a_{n}(t)\cos(\tfrac{2\pi n}{\omega}x)italic_u ( italic_t , italic_x ) = ∑ start_POSTSUBSCRIPT italic_n ∈ blackboard_Z end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_t ) italic_e start_POSTSUPERSCRIPT divide start_ARG 2 italic_π italic_i italic_n end_ARG start_ARG italic_ω end_ARG italic_x end_POSTSUPERSCRIPT = italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_t ) + 2 ∑ start_POSTSUBSCRIPT italic_n = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_t ) roman_cos ( divide start_ARG 2 italic_π italic_n end_ARG start_ARG italic_ω end_ARG italic_x ) (6)

By Theorem 2.1 if |A|>30eπ2144.3𝐴30superscript𝑒𝜋2144.3|A|>30e^{\tfrac{\pi}{2}}\approx 144.3| italic_A | > 30 italic_e start_POSTSUPERSCRIPT divide start_ARG italic_π end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT ≈ 144.3, then the solution is guaranteed to exist for all t𝑡t\in\mathbb{R}italic_t ∈ blackboard_R and converge to zero in both forward and backwards time. In Figure 1 is depicted the solution of the NLS (3) over the time interval t[0,1]𝑡01t\in[0,1]italic_t ∈ [ 0 , 1 ], taking initial data in (5) with integers 150A150150𝐴150-150\leq A\leq 150- 150 ≤ italic_A ≤ 150. This computation was performed using 256 Fourier modes and a time step of h=104superscript104h=10^{-4}italic_h = 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT with an exponential integrator [DHT14].

Almost all initial data converges to zero, however close to A=5𝐴5A=-5italic_A = - 5 the solutions take a long time to decay. For each solution which does decay to zero, Theorem 2.1 enables us to identify when it enters a so called “trap** region” where it is guaranteed to converge to zero. This is denoted in Figure 1 by the dashed gray line. While a neighborhood of initial data with A𝐴Aitalic_A about 55-5- 5 does decay, it seems to take asymptotically long to enter the trap** region. This provides us with a more robust measure of identifying solutions which do not converge to the zero equilibrium. Moreover, there appears to be a codimension-1 manifold for which initial data to the NLS does not converge to zero.

Refer to caption
Figure 1: The norm u30(t;A)Lsubscriptnormsubscript𝑢30𝑡𝐴superscript𝐿\|u_{30}(t;A)\|_{L^{\infty}}∥ italic_u start_POSTSUBSCRIPT 30 end_POSTSUBSCRIPT ( italic_t ; italic_A ) ∥ start_POSTSUBSCRIPT italic_L start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT for u30(x;A)subscript𝑢30𝑥𝐴u_{30}(x;A)italic_u start_POSTSUBSCRIPT 30 end_POSTSUBSCRIPT ( italic_x ; italic_A ) as in (5). The dashed gray lines represent where the solution reaches the trap** region given by Theorem 2.1.

Informed by Conjecture 1.2 we selected A30=5.3070235subscript𝐴305.3070235A_{30}=-5.3070235italic_A start_POSTSUBSCRIPT 30 end_POSTSUBSCRIPT = - 5.3070235 in (5), and the solution of this initial value problem under the dynamics of the NLS (3) is plotted in Figure 2. This computation used a 4096 Fourier mode truncation and a time step of h=107superscript107h=10^{-7}italic_h = 10 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT. We selected A30subscript𝐴30A_{30}italic_A start_POSTSUBSCRIPT 30 end_POSTSUBSCRIPT using a bisection method so that u30subscript𝑢30u_{30}italic_u start_POSTSUBSCRIPT 30 end_POSTSUBSCRIPT would be on the stable manifold 𝒲(0)𝒲0\mathcal{W}(0)caligraphic_W ( 0 ) for the nonlinear heat equation (1). We also note that nearby values of A𝐴Aitalic_A (e.g. ±1%absentplus-or-minuspercent1\approx\pm 1\%≈ ± 1 %) also appear to blowup under the dynamics of (3).

Overall, the trajectory appears to oscillate while steadily growing larger. As shown in Figure 2 (b), the Lsuperscript𝐿L^{\infty}italic_L start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT norm of the solution maintains regular oscillations of fixed period yet steadily grows larger in amplitude. In Figure 2 (c), we display the relative composition of each Fourier modes to the L2superscript𝐿2L^{2}italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT norm. By Parseval’s identity u(t)L22=a(t)2=|a0(t)|2+2n=1|an(t)|2superscriptsubscriptnorm𝑢𝑡superscript𝐿22subscriptnorm𝑎𝑡superscript2superscriptsubscript𝑎0𝑡22superscriptsubscript𝑛1superscriptsubscript𝑎𝑛𝑡2\|u(t)\|_{L^{2}}^{2}=\|a(t)\|_{\ell^{2}}=|a_{0}(t)|^{2}+2\sum_{n=1}^{\infty}|a% _{n}(t)|^{2}∥ italic_u ( italic_t ) ∥ start_POSTSUBSCRIPT italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = ∥ italic_a ( italic_t ) ∥ start_POSTSUBSCRIPT roman_ℓ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUBSCRIPT = | italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_t ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 2 ∑ start_POSTSUBSCRIPT italic_n = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT | italic_a start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_t ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, and we plot in Figure 2(c) the ratio of the nthsuperscript𝑛thn^{\text{th}}italic_n start_POSTSUPERSCRIPT th end_POSTSUPERSCRIPT mode:

En(t):={|an|2a(t)2if n=02|an|2a(t)2if n0assignsubscript𝐸𝑛𝑡casessuperscriptsubscript𝑎𝑛2subscriptnorm𝑎𝑡superscript2if 𝑛02superscriptsubscript𝑎𝑛2subscriptnorm𝑎𝑡superscript2if 𝑛0\displaystyle E_{n}(t):=\begin{dcases}\frac{|a_{n}|^{2}}{\|a(t)\|_{\ell^{2}}}&% \mbox{if }n=0\\ \frac{2|a_{n}|^{2}}{\|a(t)\|_{\ell^{2}}}&\mbox{if }n\neq 0\end{dcases}italic_E start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_t ) := { start_ROW start_CELL divide start_ARG | italic_a start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG ∥ italic_a ( italic_t ) ∥ start_POSTSUBSCRIPT roman_ℓ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUBSCRIPT end_ARG end_CELL start_CELL if italic_n = 0 end_CELL end_ROW start_ROW start_CELL divide start_ARG 2 | italic_a start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG ∥ italic_a ( italic_t ) ∥ start_POSTSUBSCRIPT roman_ℓ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUBSCRIPT end_ARG end_CELL start_CELL if italic_n ≠ 0 end_CELL end_ROW (7)
Refer to caption
Figure 2: Solution to the NLS in (3) with initial data (5) with A30=5.3070235subscript𝐴305.3070235A_{30}=-5.3070235italic_A start_POSTSUBSCRIPT 30 end_POSTSUBSCRIPT = - 5.3070235: (a) Space time plot of the solution; (b) Inverse norm of the solution 1/u(t)L1subscriptnorm𝑢𝑡superscript𝐿1/\|u(t)\|_{L^{\infty}}1 / ∥ italic_u ( italic_t ) ∥ start_POSTSUBSCRIPT italic_L start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT; (c) Relative proportions Ensubscript𝐸𝑛E_{n}italic_E start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT of the Fourier modes, see (7).

While the spacetime plot in Figure 2(a) appears quite complex, the relative energy of each spatial mode in Figure 2(c) paints a clearer picture. As the solution evolves in time, higher spatial frequencies are progressively excited with increasing temporal oscillations. Furthermore, these oscillations grow ever more complex and almost fractal-like.

3 Secular growth of solutions along a center manifold

3.1 Dynamics of the PDE’s Galerkin truncation

To investigate the initial growth phase of the solutions we consider finite Galerkin truncations of (3). Consider a function given as a 2π2𝜋2\pi2 italic_π-periodic cosine series as in (6), where for ease of analysis in this section we take ω=2π𝜔2𝜋\omega=2\piitalic_ω = 2 italic_π. Then the N𝑁Nitalic_N-mode truncation of (2) yields the following dynamics on the Fourier modes:

eiθa˙nsuperscript𝑒𝑖𝜃subscript˙𝑎𝑛\displaystyle e^{-i\theta}\dot{a}_{n}italic_e start_POSTSUPERSCRIPT - italic_i italic_θ end_POSTSUPERSCRIPT over˙ start_ARG italic_a end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT =n2an+n1+n1=nNn1,n2Na|n1|a|n2|absentsuperscript𝑛2subscript𝑎𝑛subscriptsubscript𝑛1subscript𝑛1𝑛formulae-sequence𝑁subscript𝑛1subscript𝑛2𝑁subscript𝑎subscript𝑛1subscript𝑎subscript𝑛2\displaystyle=-n^{2}a_{n}+\sum_{\begin{subarray}{c}n_{1}+n_{1}=n\\ -N\leq n_{1},n_{2}\leq N\end{subarray}}a_{|n_{1}|}a_{|n_{2}|}= - italic_n start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT + ∑ start_POSTSUBSCRIPT start_ARG start_ROW start_CELL italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_n end_CELL end_ROW start_ROW start_CELL - italic_N ≤ italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_n start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ≤ italic_N end_CELL end_ROW end_ARG end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT | italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT | end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT | italic_n start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT | end_POSTSUBSCRIPT (8)

By the cosine ansatz we have an=ansubscript𝑎𝑛subscript𝑎𝑛a_{n}=a_{-n}italic_a start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = italic_a start_POSTSUBSCRIPT - italic_n end_POSTSUBSCRIPT, thereby this defines a complex ODE on N+1superscript𝑁1\mathbb{C}^{N+1}blackboard_C start_POSTSUPERSCRIPT italic_N + 1 end_POSTSUPERSCRIPT. Furthermore the system has an equilibrium at 0N+10superscript𝑁10\in\mathbb{C}^{N+1}0 ∈ blackboard_C start_POSTSUPERSCRIPT italic_N + 1 end_POSTSUPERSCRIPT whose linearization has eigenvalues λn=eiθn2subscript𝜆𝑛superscript𝑒𝑖𝜃superscript𝑛2\lambda_{n}=-e^{i\theta}n^{2}italic_λ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = - italic_e start_POSTSUPERSCRIPT italic_i italic_θ end_POSTSUPERSCRIPT italic_n start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT for 0nN0𝑛𝑁0\leq n\leq N0 ≤ italic_n ≤ italic_N.

Define 𝒲(0)𝒲0\mathcal{W}(0)caligraphic_W ( 0 ) as the unique invariant manifold tangent to {0}×N0superscript𝑁\{0\}\times\mathbb{C}^{N}{ 0 } × blackboard_C start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT. Selecting different values of θ𝜃\thetaitalic_θ will induce different internal dynamics on 𝒲𝒲\mathcal{W}caligraphic_W, however 𝒲𝒲\mathcal{W}caligraphic_W will remain an invariant manifold for any choice of θ𝜃\thetaitalic_θ. For example, if the real component of eiθsuperscript𝑒𝑖𝜃-e^{i\theta}- italic_e start_POSTSUPERSCRIPT italic_i italic_θ end_POSTSUPERSCRIPT is negative, then 𝒲𝒲\mathcal{W}caligraphic_W is a submanifold of the equilibrium’s stable set, and trajectories on 𝒲𝒲\mathcal{W}caligraphic_W will exponentially approach the zero equilibrium in forward time.

More generally if \mathcal{M}caligraphic_M is an invariant complex manifold of any ODE x˙=f(x)˙𝑥𝑓𝑥\dot{x}=f(x)over˙ start_ARG italic_x end_ARG = italic_f ( italic_x ), then \mathcal{M}caligraphic_M will also be an invariant manifold of the system x˙=μf(x)˙𝑥𝜇𝑓𝑥\dot{x}=\mu f(x)over˙ start_ARG italic_x end_ARG = italic_μ italic_f ( italic_x ) for scalars μ𝜇\mu\in\mathbb{C}italic_μ ∈ blackboard_C. The most well know usage of this fact is the case μ=1𝜇1\mu=-1italic_μ = - 1. That is, the stable manifold of an equilibrium in forward time is also the unstable manifold of the equilibrium in backwards time. While trajectories on \mathcal{M}caligraphic_M remain invariant when μ𝜇\mu\in\mathbb{R}italic_μ ∈ blackboard_R, this is not the case when μ𝜇\muitalic_μ is complex.

Refer to caption
Figure 3: Phase portraits of the dynamics to (9) with the invariant manifold 𝒲𝒲\mathcal{W}caligraphic_W tangent to {0}×0\{0\}\times\mathbb{C}{ 0 } × blackboard_C depicted in orange. (a) When θ=0𝜃0\theta=0italic_θ = 0 real initial data remains invariant, and the stable manifold 𝒲𝒲\mathcal{W}caligraphic_W divides trajectories which converge to zero, from those which/that blowup in finite time. (b) When θ=π/2𝜃𝜋2\theta=\pi/2italic_θ = italic_π / 2, the invariant manifold 𝒲𝒲\mathcal{W}caligraphic_W is foliated by periodic orbits.

For a simple illustration consider the dynamics of (8) in the case N=1𝑁1N=1italic_N = 1, where we obtain the system of complex ODEs:

eiθa˙0superscript𝑒𝑖𝜃subscript˙𝑎0\displaystyle e^{-i\theta}\dot{a}_{0}italic_e start_POSTSUPERSCRIPT - italic_i italic_θ end_POSTSUPERSCRIPT over˙ start_ARG italic_a end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT =a02+2a12absentsuperscriptsubscript𝑎022superscriptsubscript𝑎12\displaystyle=a_{0}^{2}+2a_{1}^{2}= italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 2 italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (9)
eiθa˙1superscript𝑒𝑖𝜃subscript˙𝑎1\displaystyle e^{-i\theta}\dot{a}_{1}italic_e start_POSTSUPERSCRIPT - italic_i italic_θ end_POSTSUPERSCRIPT over˙ start_ARG italic_a end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT =a1+2a0a1absentsubscript𝑎12subscript𝑎0subscript𝑎1\displaystyle=-a_{1}+2a_{0}a_{1}= - italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + 2 italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT

The case θ=0𝜃0\theta=0italic_θ = 0 is analogous to the nonlinear heat equation (1), see Figure 3(a). Real initial data is invariant, and if a0>0subscript𝑎00a_{0}>0italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT > 0 then there is finite time blowup. Furthermore the manifold 𝒲𝒲\mathcal{W}caligraphic_W divides solutions which blowup in finite time from those which converge algebraically to zero. Furthermore if π/2<θ<π/2𝜋2𝜃𝜋2-\pi/2<\theta<\pi/2- italic_π / 2 < italic_θ < italic_π / 2 then solutions on 𝒲𝒲\mathcal{W}caligraphic_W will exponentially decay.

In contrast, consider the case θ=±π2𝜃plus-or-minus𝜋2\theta=\pm\frac{\pi}{2}italic_θ = ± divide start_ARG italic_π end_ARG start_ARG 2 end_ARG analogous to the NLS in (3), see Figure 3(b). In this case the eigenvalues {λ0,λ1}={0,i}subscript𝜆0subscript𝜆10minus-or-plus𝑖\{\lambda_{0},\lambda_{1}\}=\{0,\mp i\}{ italic_λ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT } = { 0 , ∓ italic_i } are purely imaginary and the manifold 𝒲𝒲\mathcal{W}caligraphic_W is a submanifold of the center manifold. Furthermore 𝒲(0)𝒲0\mathcal{W}(0)caligraphic_W ( 0 ) is foliated by a Lyapunov family of periodic orbits [Hen73]. In Figure 2 we can similarly observe periodic behavior for short time scales (0<t<0.250𝑡0.250<t<0.250 < italic_t < 0.25). However, like with Figure 2(c), we will see in the dynamics of (8) for larger values of N𝑁Nitalic_N that there is a secular drift of solutions along 𝒲𝒲\mathcal{W}caligraphic_W, whereby the lower modes progressively excite the higher modes.

3.2 Parameterizing the manifold 𝒲𝒲\mathcal{W}caligraphic_W

The existence of stable, unstable, and center manifolds associated to equilibria has been established for a wide variety of dynamical systems, such as ODEs, PDEs, and DDEs [SY02]. However even if such a manifold is unique, its representation via a coordinate chart is not unique, nor is there just one method available for computing an approximation for the manifold.

To analyze and compute the invariant manifold 𝒲𝒲\mathcal{W}caligraphic_W associated with (8) we shall use the parameterization method [CFdlL03a, CFdlL03b, CFDLL05], which has had great success analyzing the dynamics of PDEs [RJ19, JLT22a, JLT22b, GJT22, JH22, OVTF23]. This approach may be seen to be in contrast to the graph approach, where one fixes the representation of the invariant manifold as graph over the particular eigenspace and the internal dynamics need to be solved for. Instead, the parameterization method fixes the internal dynamics and solves for the coordinate chart as a map into the entire phase space.

To review the parameterization method we follow the presentation in [HCF+16]. Consider F:nn:𝐹superscript𝑛superscript𝑛F:\mathbb{C}^{n}\to\mathbb{C}^{n}italic_F : blackboard_C start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT → blackboard_C start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT and the differential equation given by:

z˙=eiθF(z)˙𝑧superscript𝑒𝑖𝜃𝐹𝑧\dot{z}=e^{i\theta}F(z)over˙ start_ARG italic_z end_ARG = italic_e start_POSTSUPERSCRIPT italic_i italic_θ end_POSTSUPERSCRIPT italic_F ( italic_z )

having an equilibrium zsubscript𝑧z_{*}italic_z start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT. Let Ln×d𝐿superscript𝑛𝑑L\in\mathbb{C}^{n\times d}italic_L ∈ blackboard_C start_POSTSUPERSCRIPT italic_n × italic_d end_POSTSUPERSCRIPT denote a matrix whose d𝑑ditalic_d-columns are eigenvectors of the linearization DF(z)𝐷𝐹subscript𝑧DF(z_{*})italic_D italic_F ( italic_z start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT ), and let Vn𝑉superscript𝑛V\subseteq\mathbb{C}^{n}italic_V ⊆ blackboard_C start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT denote the d𝑑ditalic_d-dimensional subspace spanned by the column vectors of L𝐿Litalic_L. The goal is to look for a parameterization z=W(σ):dn:𝑧𝑊𝜎superscript𝑑superscript𝑛z=W(\sigma):\mathbb{C}^{d}\to\mathbb{C}^{n}italic_z = italic_W ( italic_σ ) : blackboard_C start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT → blackboard_C start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT of the invariant manifold 𝒲𝒲\mathcal{W}caligraphic_W tangent to V𝑉Vitalic_V at zsubscript𝑧z_{*}italic_z start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT. The internal dynamics on the manifold are described by a vector field σ˙=eiθf(σ)˙𝜎superscript𝑒𝑖𝜃𝑓𝜎\dot{\sigma}=e^{i\theta}f(\sigma)over˙ start_ARG italic_σ end_ARG = italic_e start_POSTSUPERSCRIPT italic_i italic_θ end_POSTSUPERSCRIPT italic_f ( italic_σ ) on dsuperscript𝑑\mathbb{C}^{d}blackboard_C start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT with f(0)=0𝑓00f(0)=0italic_f ( 0 ) = 0, and the invariance equation is given by:

eiθF(W(σ))superscript𝑒𝑖𝜃𝐹𝑊𝜎\displaystyle e^{i\theta}F(W(\sigma))italic_e start_POSTSUPERSCRIPT italic_i italic_θ end_POSTSUPERSCRIPT italic_F ( italic_W ( italic_σ ) ) =DW(σ)eiθf(σ)absent𝐷𝑊𝜎superscript𝑒𝑖𝜃𝑓𝜎\displaystyle=DW(\sigma)e^{i\theta}f(\sigma)= italic_D italic_W ( italic_σ ) italic_e start_POSTSUPERSCRIPT italic_i italic_θ end_POSTSUPERSCRIPT italic_f ( italic_σ ) (10)

Note that the complex phase eiθsuperscript𝑒𝑖𝜃e^{i\theta}italic_e start_POSTSUPERSCRIPT italic_i italic_θ end_POSTSUPERSCRIPT cancels and disappears completely!

The parameterization method seeks to write W𝑊Witalic_W and f𝑓fitalic_f as power series in σ=(σ1,,σd)𝜎subscript𝜎1subscript𝜎𝑑\sigma=(\sigma_{1},\dots,\sigma_{d})italic_σ = ( italic_σ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , italic_σ start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ):

W(σ)𝑊𝜎\displaystyle W(\sigma)italic_W ( italic_σ ) =z+k1Wk(σ),absentsubscript𝑧subscript𝑘1subscript𝑊𝑘𝜎\displaystyle=z_{*}+\sum_{k\geq 1}W_{k}(\sigma),= italic_z start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT + ∑ start_POSTSUBSCRIPT italic_k ≥ 1 end_POSTSUBSCRIPT italic_W start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_σ ) , f(σ)=k1fk(σ)𝑓𝜎subscript𝑘1subscript𝑓𝑘𝜎\displaystyle f(\sigma)=\sum_{k\geq 1}f_{k}(\sigma)italic_f ( italic_σ ) = ∑ start_POSTSUBSCRIPT italic_k ≥ 1 end_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_σ ) (11)

where each Wk(σ)subscript𝑊𝑘𝜎W_{k}(\sigma)italic_W start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_σ ) (respectively fksubscript𝑓𝑘f_{k}italic_f start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT) is a homogeneous polynomial in σ𝜎\sigmaitalic_σ of degree k𝑘kitalic_k with coefficients in nsuperscript𝑛\mathbb{C}^{n}blackboard_C start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT (respectively with coefficients in dsuperscript𝑑\mathbb{C}^{d}blackboard_C start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT), that is:

Wk(σ)subscript𝑊𝑘𝜎\displaystyle W_{k}(\sigma)italic_W start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_σ ) =k1++kd=kW(k1kd)σ1k1σdkd,absentsubscriptsubscript𝑘1subscript𝑘𝑑𝑘subscript𝑊subscript𝑘1subscript𝑘𝑑superscriptsubscript𝜎1subscript𝑘1superscriptsubscript𝜎𝑑subscript𝑘𝑑\displaystyle=\sum_{k_{1}+\dots+k_{d}=k}W_{(k_{1}\dots k_{d})}\sigma_{1}^{k_{1% }}\cdots\sigma_{d}^{k_{d}},= ∑ start_POSTSUBSCRIPT italic_k start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + ⋯ + italic_k start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT = italic_k end_POSTSUBSCRIPT italic_W start_POSTSUBSCRIPT ( italic_k start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT … italic_k start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ) end_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ⋯ italic_σ start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_POSTSUPERSCRIPT , W(k1kd)subscript𝑊subscript𝑘1subscript𝑘𝑑\displaystyle W_{(k_{1}\dots k_{d})}italic_W start_POSTSUBSCRIPT ( italic_k start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT … italic_k start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ) end_POSTSUBSCRIPT nabsentsuperscript𝑛\displaystyle\in\mathbb{C}^{n}∈ blackboard_C start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT (12)

To enforce the invariant manifold 𝒲𝒲\mathcal{W}caligraphic_W being tangent to the eigenspace V𝑉Vitalic_V we take W1(σ)=Lσsubscript𝑊1𝜎𝐿𝜎W_{1}(\sigma)=L\sigmaitalic_W start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_σ ) = italic_L italic_σ and f1(σ)=ΛLσsubscript𝑓1𝜎subscriptΛ𝐿𝜎f_{1}(\sigma)=\Lambda_{L}\sigmaitalic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_σ ) = roman_Λ start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT italic_σ. The higher order terms in the power series may be obtained by matching the order-k𝑘kitalic_k terms in (10), yielding the cohomological equations:

DF(z)Wk(σ)DWk(σ)ΛLσLfk(σ)=Ek(σ)𝐷𝐹subscript𝑧subscript𝑊𝑘𝜎𝐷subscript𝑊𝑘𝜎subscriptΛ𝐿𝜎𝐿subscript𝑓𝑘𝜎subscript𝐸𝑘𝜎\displaystyle DF(z_{*})W_{k}(\sigma)-DW_{k}(\sigma)\Lambda_{L}\sigma-Lf_{k}(% \sigma)=-E_{k}(\sigma)italic_D italic_F ( italic_z start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT ) italic_W start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_σ ) - italic_D italic_W start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_σ ) roman_Λ start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT italic_σ - italic_L italic_f start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_σ ) = - italic_E start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_σ ) (13)

where:

Ek(σ)=[F(W<k(σ))]k[DW<k(σ)f<k(σ)]ksubscript𝐸𝑘𝜎subscriptdelimited-[]𝐹subscript𝑊absent𝑘𝜎𝑘subscriptdelimited-[]𝐷subscript𝑊absent𝑘𝜎subscript𝑓absent𝑘𝜎𝑘E_{k}(\sigma)=[F(W_{<k}(\sigma))]_{k}-[DW_{<k}(\sigma)f_{<k}(\sigma)]_{k}italic_E start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_σ ) = [ italic_F ( italic_W start_POSTSUBSCRIPT < italic_k end_POSTSUBSCRIPT ( italic_σ ) ) ] start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT - [ italic_D italic_W start_POSTSUBSCRIPT < italic_k end_POSTSUBSCRIPT ( italic_σ ) italic_f start_POSTSUBSCRIPT < italic_k end_POSTSUBSCRIPT ( italic_σ ) ] start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT

is the order-k𝑘kitalic_k error term, depending only on coefficients whose order is strictly less that k𝑘kitalic_k. Thus through (13) the coefficients Wksubscript𝑊𝑘W_{k}italic_W start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT may be solved recursively for orders k2𝑘2k\geq 2italic_k ≥ 2, subject to the choice of coefficients fksubscript𝑓𝑘f_{k}italic_f start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT.

The simplest dynamics one may conjugate to is linear dynamics σ˙=ΛLσ˙𝜎subscriptΛ𝐿𝜎\dot{\sigma}=\Lambda_{L}\sigmaover˙ start_ARG italic_σ end_ARG = roman_Λ start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT italic_σ wherein fk=0subscript𝑓𝑘0f_{k}=0italic_f start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = 0 for k2𝑘2k\geq 2italic_k ≥ 2. Then for each k=(k1,,kd)𝑘subscript𝑘1subscript𝑘𝑑\vec{k}=(k_{1},\dots,k_{d})over→ start_ARG italic_k end_ARG = ( italic_k start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , italic_k start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ) equation (13) reduces to the linear system:

(DF(z)kλIn)Wk=Ek𝐷𝐹subscript𝑧𝑘𝜆subscript𝐼𝑛subscript𝑊𝑘subscript𝐸𝑘(DF(z_{*})-\vec{k}\cdot\lambda I_{n})W_{\vec{k}}=-E_{\vec{k}}( italic_D italic_F ( italic_z start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT ) - over→ start_ARG italic_k end_ARG ⋅ italic_λ italic_I start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) italic_W start_POSTSUBSCRIPT over→ start_ARG italic_k end_ARG end_POSTSUBSCRIPT = - italic_E start_POSTSUBSCRIPT over→ start_ARG italic_k end_ARG end_POSTSUBSCRIPT

This can be solved for |k|2𝑘2|k|\geq 2| italic_k | ≥ 2 so long as the matrix (DF(z)kλIn)𝐷𝐹subscript𝑧𝑘𝜆subscript𝐼𝑛(DF(z_{*})-\vec{k}\cdot\lambda I_{n})( italic_D italic_F ( italic_z start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT ) - over→ start_ARG italic_k end_ARG ⋅ italic_λ italic_I start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) is invertible. This fails to occur when there is an internal resonance, which is to say when:

kλ=λi𝑘𝜆subscript𝜆𝑖\vec{k}\cdot\lambda=\lambda_{i}over→ start_ARG italic_k end_ARG ⋅ italic_λ = italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT

for kd𝑘superscript𝑑\vec{k}\in\mathbb{N}^{d}over→ start_ARG italic_k end_ARG ∈ blackboard_N start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT and |k|2𝑘2|k|\geq 2| italic_k | ≥ 2. This can only occur if the eigenvalues (λ1λd)subscript𝜆1subscript𝜆𝑑(\lambda_{1}\dots\lambda_{d})( italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT … italic_λ start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ) are rationally dependent.

Returning to the dynamics resulting from the Galerkin truncation in (8), recall that the zero equilibrium has eigenvalues λn=eiθn2subscript𝜆𝑛superscript𝑒𝑖𝜃superscript𝑛2\lambda_{n}=-e^{i\theta}n^{2}italic_λ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = - italic_e start_POSTSUPERSCRIPT italic_i italic_θ end_POSTSUPERSCRIPT italic_n start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT for integers 0nN0𝑛𝑁0\leq n\leq N0 ≤ italic_n ≤ italic_N. Thus, resonances occur whenever we can write a square integer as a sum of other square integers:

m2=n=1m1knn2superscript𝑚2superscriptsubscript𝑛1𝑚1subscript𝑘𝑛superscript𝑛2m^{2}=\sum_{n=1}^{m-1}k_{n}n^{2}italic_m start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = ∑ start_POSTSUBSCRIPT italic_n = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m - 1 end_POSTSUPERSCRIPT italic_k start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_n start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT

for non-negative integers knsubscript𝑘𝑛k_{n}italic_k start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT with k1++km12subscript𝑘1subscript𝑘𝑚12k_{1}+\dots+k_{m-1}\geq 2italic_k start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + ⋯ + italic_k start_POSTSUBSCRIPT italic_m - 1 end_POSTSUBSCRIPT ≥ 2. This happens abundantly often! For example m2=m212superscript𝑚2superscript𝑚2superscript12m^{2}=m^{2}\cdot 1^{2}italic_m start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = italic_m start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⋅ 1 start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT or m2=(m2n2)12+1n2superscript𝑚2superscript𝑚2superscript𝑛2superscript121superscript𝑛2m^{2}=(m^{2}-n^{2})\cdot 1^{2}+1\cdot n^{2}italic_m start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = ( italic_m start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_n start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) ⋅ 1 start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 1 ⋅ italic_n start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT for any 1<n<m1𝑛𝑚1<n<m1 < italic_n < italic_m. While resonances are an obstruction to conjugating the internal dynamics of 𝒲𝒲\mathcal{W}caligraphic_W to a linear system, a parameterization may still be obtained for more nontrivial internal dynamics.

For an example, consider the N=3𝑁3N=3italic_N = 3 Galerkin truncation of (8) given by:

eiθa˙0superscript𝑒𝑖𝜃subscript˙𝑎0\displaystyle e^{-i\theta}\dot{a}_{0}italic_e start_POSTSUPERSCRIPT - italic_i italic_θ end_POSTSUPERSCRIPT over˙ start_ARG italic_a end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT =a02+2a12+2a22+2a32absentsuperscriptsubscript𝑎022superscriptsubscript𝑎122superscriptsubscript𝑎222superscriptsubscript𝑎32\displaystyle=a_{0}^{2}+2a_{1}^{2}+2a_{2}^{2}+2a_{3}^{2}= italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 2 italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 2 italic_a start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 2 italic_a start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT
eiθa˙1superscript𝑒𝑖𝜃subscript˙𝑎1\displaystyle e^{-i\theta}\dot{a}_{1}italic_e start_POSTSUPERSCRIPT - italic_i italic_θ end_POSTSUPERSCRIPT over˙ start_ARG italic_a end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT =a1+2a0a1+2a1a2+2a2a3absentsubscript𝑎12subscript𝑎0subscript𝑎12subscript𝑎1subscript𝑎22subscript𝑎2subscript𝑎3\displaystyle=-a_{1}+2a_{0}a_{1}+2a_{1}a_{2}+2a_{2}a_{3}= - italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + 2 italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + 2 italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + 2 italic_a start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT (14)
eiθa˙2superscript𝑒𝑖𝜃subscript˙𝑎2\displaystyle e^{-i\theta}\dot{a}_{2}italic_e start_POSTSUPERSCRIPT - italic_i italic_θ end_POSTSUPERSCRIPT over˙ start_ARG italic_a end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT =4a2+2a0a2+a12+2a1a3absent4subscript𝑎22subscript𝑎0subscript𝑎2superscriptsubscript𝑎122subscript𝑎1subscript𝑎3\displaystyle=-4a_{2}+2a_{0}a_{2}+a_{1}^{2}+2a_{1}a_{3}= - 4 italic_a start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + 2 italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 2 italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT
eiθa˙3superscript𝑒𝑖𝜃subscript˙𝑎3\displaystyle e^{-i\theta}\dot{a}_{3}italic_e start_POSTSUPERSCRIPT - italic_i italic_θ end_POSTSUPERSCRIPT over˙ start_ARG italic_a end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT =9a3+2a0a3+2a1a2absent9subscript𝑎32subscript𝑎0subscript𝑎32subscript𝑎1subscript𝑎2\displaystyle=-9a_{3}+2a_{0}a_{3}+2a_{1}a_{2}= - 9 italic_a start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT + 2 italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT + 2 italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT

Again, we define 𝒲(0)𝒲0\mathcal{W}(0)caligraphic_W ( 0 ) as the unique invariant manifold tangent to {0}×30superscript3\{0\}\times\mathbb{C}^{3}{ 0 } × blackboard_C start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT. The equilibrium’s nonzero eigenvalues are eiθλ=(1,4,9)superscript𝑒𝑖𝜃𝜆149e^{-i\theta}\lambda=(-1,-4,-9)italic_e start_POSTSUPERSCRIPT - italic_i italic_θ end_POSTSUPERSCRIPT italic_λ = ( - 1 , - 4 , - 9 ), and we obtain the resonances:

(4,0,0)λ400𝜆\displaystyle(4,0,0)\cdot\lambda( 4 , 0 , 0 ) ⋅ italic_λ =λ2,absentsubscript𝜆2\displaystyle=\lambda_{2},= italic_λ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , (1,2,0)λ120𝜆\displaystyle(1,2,0)\cdot\lambda( 1 , 2 , 0 ) ⋅ italic_λ =λ3,absentsubscript𝜆3\displaystyle=\lambda_{3},= italic_λ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT , (5,1,0)λ510𝜆\displaystyle(5,1,0)\cdot\lambda( 5 , 1 , 0 ) ⋅ italic_λ =λ3,absentsubscript𝜆3\displaystyle=\lambda_{3},= italic_λ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT , (9,0,0)λ900𝜆\displaystyle(9,0,0)\cdot\lambda( 9 , 0 , 0 ) ⋅ italic_λ =λ3absentsubscript𝜆3\displaystyle=\lambda_{3}= italic_λ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT

One can deal with resonances by adjusting the function f𝑓fitalic_f, that one conjugates the dynamics to [vdBMJR16]. In our code [Jaq24] we calculate a parameterization W𝑊Witalic_W such that the dynamics are conjugate to:

μ1σ˙1superscript𝜇1subscript˙𝜎1\displaystyle\mu^{-1}\dot{\sigma}_{1}italic_μ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT over˙ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT =σ1absentsubscript𝜎1\displaystyle=-\sigma_{1}= - italic_σ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT
μ1σ˙2superscript𝜇1subscript˙𝜎2\displaystyle\mu^{-1}\dot{\sigma}_{2}italic_μ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT over˙ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT =4σ2+13σ14absent4subscript𝜎213superscriptsubscript𝜎14\displaystyle=-4\sigma_{2}+\tfrac{1}{3}\sigma_{1}^{4}= - 4 italic_σ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + divide start_ARG 1 end_ARG start_ARG 3 end_ARG italic_σ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT (15)
μ1σ˙3superscript𝜇1subscript˙𝜎3\displaystyle\mu^{-1}\dot{\sigma}_{3}italic_μ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT over˙ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT =9σ3σ1σ22+1924σ15σ2+1181σ19absent9subscript𝜎3subscript𝜎1superscriptsubscript𝜎221924superscriptsubscript𝜎15subscript𝜎21181superscriptsubscript𝜎19\displaystyle=-9\sigma_{3}-\sigma_{1}\sigma_{2}^{2}+\tfrac{19}{24}\sigma_{1}^{% 5}\sigma_{2}+\tfrac{11}{81}\sigma_{1}^{9}= - 9 italic_σ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT - italic_σ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + divide start_ARG 19 end_ARG start_ARG 24 end_ARG italic_σ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT italic_σ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + divide start_ARG 11 end_ARG start_ARG 81 end_ARG italic_σ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 9 end_POSTSUPERSCRIPT

where μ=eiθ𝜇superscript𝑒𝑖𝜃\mu=e^{i\theta}italic_μ = italic_e start_POSTSUPERSCRIPT italic_i italic_θ end_POSTSUPERSCRIPT. The specific coefficients for the higher order terms, such as 13σ1413superscriptsubscript𝜎14\frac{1}{3}\sigma_{1}^{4}divide start_ARG 1 end_ARG start_ARG 3 end_ARG italic_σ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT and 1924σ15σ21924superscriptsubscript𝜎15subscript𝜎2\tfrac{19}{24}\sigma_{1}^{5}\sigma_{2}divide start_ARG 19 end_ARG start_ARG 24 end_ARG italic_σ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT italic_σ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, are not able to be determined in advance and requires one to first solve the cohomological equation (13) for all of the lower order terms. While nonlinear, the internal dynamics of 𝒲𝒲\mathcal{W}caligraphic_W determined by (15) is integrable, and for initial data σ(0)=(γ1,γ2,γ3)𝜎0subscript𝛾1subscript𝛾2subscript𝛾3\sigma(0)=(\gamma_{1},\gamma_{2},\gamma_{3})italic_σ ( 0 ) = ( italic_γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_γ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_γ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ) its solution is given by:

σ1(t)subscript𝜎1𝑡\displaystyle\sigma_{1}(t)italic_σ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_t ) =γ1eμt,absentsubscript𝛾1superscript𝑒𝜇𝑡\displaystyle=\gamma_{1}e^{-\mu t},= italic_γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - italic_μ italic_t end_POSTSUPERSCRIPT ,
σ2(t)subscript𝜎2𝑡\displaystyle\sigma_{2}(t)italic_σ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_t ) =(γ2+μt3γ14)e4μt,absentsubscript𝛾2𝜇𝑡3superscriptsubscript𝛾14superscript𝑒4𝜇𝑡\displaystyle=\left(\gamma_{2}+\frac{\mu t}{3}\gamma_{1}^{4}\right)e^{-4\mu t},= ( italic_γ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + divide start_ARG italic_μ italic_t end_ARG start_ARG 3 end_ARG italic_γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT ) italic_e start_POSTSUPERSCRIPT - 4 italic_μ italic_t end_POSTSUPERSCRIPT , (16)
σ3(t)subscript𝜎3𝑡\displaystyle\sigma_{3}(t)italic_σ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ( italic_t ) =(γ3μtγ1γ22+(19μt24μ2t23)γ15γ2+(11μt81+19μ2t2144μ3t327)γ19)e9μtabsentsubscript𝛾3𝜇𝑡subscript𝛾1superscriptsubscript𝛾2219𝜇𝑡24superscript𝜇2superscript𝑡23superscriptsubscript𝛾15subscript𝛾211𝜇𝑡8119superscript𝜇2superscript𝑡2144superscript𝜇3superscript𝑡327superscriptsubscript𝛾19superscript𝑒9𝜇𝑡\displaystyle=\left(\gamma_{3}-\mu t\gamma_{1}\gamma_{2}^{2}+\left(\frac{19\mu t% }{24}-\frac{\mu^{2}t^{2}}{3}\right)\gamma_{1}^{5}\gamma_{2}+\left(\frac{11\mu t% }{81}+\frac{19\mu^{2}t^{2}}{144}-\frac{\mu^{3}t^{3}}{27}\right)\gamma_{1}^{9}% \right)e^{-9\mu t}= ( italic_γ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT - italic_μ italic_t italic_γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_γ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + ( divide start_ARG 19 italic_μ italic_t end_ARG start_ARG 24 end_ARG - divide start_ARG italic_μ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_t start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 3 end_ARG ) italic_γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT italic_γ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + ( divide start_ARG 11 italic_μ italic_t end_ARG start_ARG 81 end_ARG + divide start_ARG 19 italic_μ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_t start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 144 end_ARG - divide start_ARG italic_μ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT italic_t start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG start_ARG 27 end_ARG ) italic_γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 9 end_POSTSUPERSCRIPT ) italic_e start_POSTSUPERSCRIPT - 9 italic_μ italic_t end_POSTSUPERSCRIPT

Note that if π/2<θ<π/2𝜋2𝜃𝜋2-\pi/2<\theta<\pi/2- italic_π / 2 < italic_θ < italic_π / 2 then Re(μ)<0𝑅𝑒𝜇0Re(-\mu)<0italic_R italic_e ( - italic_μ ) < 0 whereby all of these solutions exponentially decay.

However if θ=±π/2𝜃plus-or-minus𝜋2\theta=\pm\pi/2italic_θ = ± italic_π / 2 then μ=±i𝜇plus-or-minus𝑖\mu=\pm iitalic_μ = ± italic_i, whereby the solutions in (16) are oscillatory with secular drift exciting the higher modes. Indeed, while |σ1(t)|subscript𝜎1𝑡|\sigma_{1}(t)|| italic_σ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_t ) | stays bounded for all t𝑡titalic_t, the higher modes grow like |σ2(t)|=𝒪(t|γ1|4)subscript𝜎2𝑡𝒪𝑡superscriptsubscript𝛾14|\sigma_{2}(t)|=\mathcal{O}(t|\gamma_{1}|^{4})| italic_σ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_t ) | = caligraphic_O ( italic_t | italic_γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT ) and |σ3(t)|=𝒪(t3|γ1|9)subscript𝜎3𝑡𝒪superscript𝑡3superscriptsubscript𝛾19|\sigma_{3}(t)|=\mathcal{O}(t^{3}|\gamma_{1}|^{9})| italic_σ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ( italic_t ) | = caligraphic_O ( italic_t start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT | italic_γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 9 end_POSTSUPERSCRIPT ). Hence small initial data in the lowest modes requires a long time before the higher modes are excited to a comparable level.

Refer to caption
Figure 4: Solution to the four-mode Galerkin truncation in (3.2) with initial data (17) (a): The inverse norm 1/u(N)(t)L1subscriptnormsuperscript𝑢𝑁𝑡superscript𝐿1/\|u^{(N)}(t)\|_{L^{\infty}}1 / ∥ italic_u start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT ( italic_t ) ∥ start_POSTSUBSCRIPT italic_L start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ; (b) Relative proportions Ensubscript𝐸𝑛E_{n}italic_E start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT of the Fourier modes, see (7).

Like with any power series, the parameterization of 𝒲𝒲\mathcal{W}caligraphic_W is only valid on the power series’ radius of convergence. (Based on our computation, the radius of convergence seems to be about 0.800.800.800.80.) While the explicit solutions given in (16) never blow up, it may be possible for a solution on 𝒲𝒲\mathcal{W}caligraphic_W to blowup after it leaves the local coordinate chart. To investigate this, we consider the following initial condition selected on the invariant manifold 𝒲𝒲\mathcal{W}caligraphic_W:

(σ1σ2σ3)matrixsubscript𝜎1subscript𝜎2subscript𝜎3\displaystyle\left(\begin{matrix}\sigma_{1}\\ \sigma_{2}\\ \sigma_{3}\end{matrix}\right)( start_ARG start_ROW start_CELL italic_σ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL italic_σ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL italic_σ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT end_CELL end_ROW end_ARG ) =( 0.43006549172907950.07398732057014827  0.00530826265454094),absentmatrix0.43006549172907950.073987320570148270.00530826265454094\displaystyle=\left(\begin{matrix}\;0.4300654917290795\\ -0.07398732057014827\\ \;\;0.00530826265454094\end{matrix}\right),= ( start_ARG start_ROW start_CELL 0.4300654917290795 end_CELL end_ROW start_ROW start_CELL - 0.07398732057014827 end_CELL end_ROW start_ROW start_CELL 0.00530826265454094 end_CELL end_ROW end_ARG ) , W(σ)𝑊𝜎\displaystyle W(\sigma)italic_W ( italic_σ ) =(0.223014092570049420.500)absentmatrix0.223014092570049420.500\displaystyle=\left(\begin{matrix}-0.22301409257004942\\ 0.5\\ 0\\ 0\end{matrix}\right)= ( start_ARG start_ROW start_CELL - 0.22301409257004942 end_CELL end_ROW start_ROW start_CELL 0.5 end_CELL end_ROW start_ROW start_CELL 0 end_CELL end_ROW start_ROW start_CELL 0 end_CELL end_ROW end_ARG ) (17)

When integrated under (3.2) with θ=π/2𝜃𝜋2\theta=\pi/2italic_θ = italic_π / 2 this trajectory undergoes a large growth in norm, see Figure 4. Like with Figure 2, one can see that there is a steady cascade pum** energy into the higher modes. This initial condition does not appear to completely blowup, as in the later stage of trajectory 1/u(t)L1subscriptnorm𝑢𝑡superscript𝐿1/\|u(t)\|_{L^{\infty}}1 / ∥ italic_u ( italic_t ) ∥ start_POSTSUBSCRIPT italic_L start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT grows rapidly. However, such a non-blowup may also be seen to be consistent with the existence of an unstable blowup profile. Note also that as our computation of 𝒲𝒲\mathcal{W}caligraphic_W was with a finite Taylor series, our the initial condition was not exactly on the manifold 𝒲𝒲\mathcal{W}caligraphic_W. (For our 20th order Taylor approximation of W𝑊Witalic_W, we estimate the error of the initial point to be on the order of 105superscript10510^{-5}10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT.) While one could get a more precise initial point on the manifold by choosing a smaller initial condition σ𝜎\sigmaitalic_σ, this would conversely increase the integration time and the associated errors.

4 Apparent self-similar blowup profiles

The finite Fourier truncation model offers a heuristic explanation for why initial data from the stable manifold of the heat equation (1) will grow larger, and exhibit a seemingly periodic forward cascade to the higher modes. However this analysis is localized at the zero equilibrium. While it is suggestive of how small initial data grows to become finitely large in norm, it does not explain how large initial data may grow without bound and blowup.

To focus on the dynamics of blowup we consider larger initial data. With consideration to Conjecture 1.2, we fix the following real initial data:

u300(x)subscript𝑢300𝑥\displaystyle u_{300}(x)italic_u start_POSTSUBSCRIPT 300 end_POSTSUBSCRIPT ( italic_x ) =300cos(2πx)+A300,absent3002𝜋𝑥subscript𝐴300\displaystyle=300\cos(2\pi x)+A_{300},= 300 roman_cos ( 2 italic_π italic_x ) + italic_A start_POSTSUBSCRIPT 300 end_POSTSUBSCRIPT , A300subscript𝐴300\displaystyle A_{300}italic_A start_POSTSUBSCRIPT 300 end_POSTSUBSCRIPT =189.286840601635,absent189.286840601635\displaystyle=-189.286840601635,= - 189.286840601635 , (18)

see Figure 5. We used computational parameters of 4096 Fourier mode truncation and a time step of h=107superscript107h=10^{-7}italic_h = 10 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT. For comparison we also consider monochromatic initial data:

u300mc(x)superscriptsubscript𝑢300𝑚𝑐𝑥\displaystyle u_{300}^{mc}(x)italic_u start_POSTSUBSCRIPT 300 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m italic_c end_POSTSUPERSCRIPT ( italic_x ) =300e2πix,absent300superscript𝑒2𝜋𝑖𝑥\displaystyle=300e^{2\pi ix},= 300 italic_e start_POSTSUPERSCRIPT 2 italic_π italic_i italic_x end_POSTSUPERSCRIPT , (19)

see Figure 6. As 300>6(2π)2236.93006superscript2𝜋2236.9300>6(2\pi)^{2}\approx 236.9300 > 6 ( 2 italic_π ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ≈ 236.9, the initial data in (19) is guaranteed to blowup as per [Jaq22].

Theorem 4.1 (Theorem 1.5[Jaq22]).

Consider (3) with initial data u0(x)=Aeiωxsubscript𝑢0𝑥𝐴superscript𝑒𝑖𝜔𝑥u_{0}(x)=Ae^{i\omega x}italic_u start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_x ) = italic_A italic_e start_POSTSUPERSCRIPT italic_i italic_ω italic_x end_POSTSUPERSCRIPT with A𝐴A\in\mathbb{C}italic_A ∈ blackboard_C, ω>0𝜔0\omega>0italic_ω > 0.

  1. (a)

    If |A|3ω2𝐴3superscript𝜔2|A|\leq 3\omega^{2}| italic_A | ≤ 3 italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, then u𝑢uitalic_u is a smooth periodic solution with period 2πω22𝜋superscript𝜔2\frac{2\pi}{\omega^{2}}divide start_ARG 2 italic_π end_ARG start_ARG italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG.

  2. (b)

    If |A|6ω2𝐴6superscript𝜔2|A|\geq 6\omega^{2}| italic_A | ≥ 6 italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, then u(t)L2subscriptnorm𝑢𝑡superscript𝐿2\|u(t)\|_{L^{2}}∥ italic_u ( italic_t ) ∥ start_POSTSUBSCRIPT italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUBSCRIPT blows up in finite time |T|2πω2subscript𝑇2𝜋superscript𝜔2|T_{*}|\leq\frac{2\pi}{\omega^{2}}| italic_T start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT | ≤ divide start_ARG 2 italic_π end_ARG start_ARG italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG.

Refer to caption
Figure 5: Numerical solution of (3) using initial data from (18). (a) Real and imaginary components of u(t,x)𝑢𝑡𝑥u(t,x)italic_u ( italic_t , italic_x ) when t=0.0743𝑡0.0743t=0.0743italic_t = 0.0743 (s=8.95𝑠8.95s=8.95italic_s = 8.95), depicted with the direction of movement of the blowup point(s); (b) Inverse norm of solution 1/u(t)L1subscriptnorm𝑢𝑡superscript𝐿1/\|u(t)\|_{L^{\infty}}1 / ∥ italic_u ( italic_t ) ∥ start_POSTSUBSCRIPT italic_L start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT; (c-d) Real and imaginary components of U𝑈Uitalic_U from (20) using self-similar coordinates with scaling (α,β)=(1,1/2)𝛼𝛽112(\alpha,\beta)=(1,1/2)( italic_α , italic_β ) = ( 1 , 1 / 2 ).
Refer to caption
Figure 6: Numerical solution of (3) using initial data from (19). (a) Real and imaginary components of u(t,x)𝑢𝑡𝑥u(t,x)italic_u ( italic_t , italic_x ) when t=0.004𝑡0.004t=0.004italic_t = 0.004 (s=7.47𝑠7.47s=7.47italic_s = 7.47), depicted with the direction of movement of the blowup point; (b) Inverse norm of solution 1/u(t)L1subscriptnorm𝑢𝑡superscript𝐿1/\|u(t)\|_{L^{\infty}}1 / ∥ italic_u ( italic_t ) ∥ start_POSTSUBSCRIPT italic_L start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT; (c-d) Real and imaginary components of U𝑈Uitalic_U from (20) using self-similar coordinates with scaling (α,β)=(2,1)𝛼𝛽21(\alpha,\beta)=(2,1)( italic_α , italic_β ) = ( 2 , 1 ).

As predicted by Conjecture 1.2 and Theorem 4.1, the numerical solutions of the initial data (18) and (19) appear to blowup. Plotted in Figures 5-6(a) are the numerical solutions after a period of time. While the monochromatic initial data yields a single blowup point moving to the right, the real initial data produces two blowup points moving in opposite directions.

Moreover, the solutions appear to undergo self-similar blowup. In a general context, this occurs when a blowup solution u𝑢uitalic_u to a given PDE can be regularized using a self-similar change of variables into a new solution U𝑈Uitalic_U with nontrivial limiting behavior [EF08, QS19]. For a given blowup time T𝑇Titalic_T and pair of scaling parameters (α,β)𝛼𝛽(\alpha,\beta)( italic_α , italic_β ) define:

u(t,x)𝑢𝑡𝑥\displaystyle u(t,x)italic_u ( italic_t , italic_x ) =1(Tt)αU(s,y)absent1superscript𝑇𝑡𝛼𝑈𝑠𝑦\displaystyle=\frac{1}{(T-t)^{\alpha}}U(s,y)= divide start_ARG 1 end_ARG start_ARG ( italic_T - italic_t ) start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT end_ARG italic_U ( italic_s , italic_y ) (20)

using the self-similar coordinates:

y𝑦\displaystyle yitalic_y =xξ(t)(Tt)βabsent𝑥𝜉𝑡superscript𝑇𝑡𝛽\displaystyle=\frac{x-\xi(t)}{(T-t)^{\beta}}= divide start_ARG italic_x - italic_ξ ( italic_t ) end_ARG start_ARG ( italic_T - italic_t ) start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT end_ARG s𝑠\displaystyle sitalic_s =log(Tt)absent𝑇𝑡\displaystyle=-\log(T-t)= - roman_log ( italic_T - italic_t ) (21)

where ξ(t)𝜉𝑡\xi(t)italic_ξ ( italic_t ) denotes the moving location of the spatial blowup point. Note that if α>0𝛼0\alpha>0italic_α > 0 and lim infsU(s)L>0subscriptlimit-infimum𝑠subscriptnorm𝑈𝑠superscript𝐿0\liminf_{s\to\infty}\|U(s)\|_{L^{\infty}}>0lim inf start_POSTSUBSCRIPT italic_s → ∞ end_POSTSUBSCRIPT ∥ italic_U ( italic_s ) ∥ start_POSTSUBSCRIPT italic_L start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT > 0, then u(t)𝑢𝑡u(t)italic_u ( italic_t ) blows up as tT𝑡𝑇t\to Titalic_t → italic_T. While the dynamics of u𝑢uitalic_u are singular as tT𝑡𝑇t\to Titalic_t → italic_T, ideally the dynamics of U𝑈Uitalic_U are well behaved as s𝑠s\to\inftyitalic_s → ∞.

In Figures 5-6(b) the inverse norm 1/u(t)L1subscriptnorm𝑢𝑡superscript𝐿1/\|u(t)\|_{L^{\infty}}1 / ∥ italic_u ( italic_t ) ∥ start_POSTSUBSCRIPT italic_L start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT can be seen to approach zero (thus blowing up). The solution with real initial data first decreases in norm and oscillates to a small extent before it begins a path towards blowup. In contrast, the solution with monochromatic initial data immediately increases in norm and accelerates in the later stage.

To estimate the scaling parameter α𝛼\alphaitalic_α, we fit the data in Figures 5-6(b) to the equation y(t)=C0|Tt|α𝑦𝑡subscript𝐶0superscript𝑇𝑡𝛼y(t)=C_{0}|T-t|^{\alpha}italic_y ( italic_t ) = italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT | italic_T - italic_t | start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT. For the solution with real initial data, fitting the data over the time interval [0.070,0.074]0.0700.074[0.070,0.074][ 0.070 , 0.074 ] yielded α=1.1457𝛼1.1457\alpha=1.1457italic_α = 1.1457 with an R2superscript𝑅2R^{2}italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT value of 0.99990.99990.99990.9999. For the solution with monochromatic initial data, fitting the data over the time interval [0.038,0.045]0.0380.045[0.038,0.045][ 0.038 , 0.045 ] yielded α=2.0098𝛼2.0098\alpha=2.0098italic_α = 2.0098 with an R2superscript𝑅2R^{2}italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT value of 0.99920.99920.99920.9992. Note that even though the two blowup profiles in Figures 5-6(a) look qualitatively similar, they appear to obey distinct rates of blowup. Additionally, while the R2superscript𝑅2R^{2}italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT values are suggestive of a good fit, the residual errors are not normally distributed, indicative of systematic bias. Practically speaking, we observe that fit parameters are quite sensitive to the time range over which the data is fit. For the case of real initial data, the non-integer value of α𝛼\alphaitalic_α may also be suggestive of a logarithmic correction term being needed in the norm scaling, as is the case with the nonlinear heat equation [BK94, MZ97].

To investigate whether these numerical solutions obey self-similar scaling, we perform a change of variables in (3) using the self-similar coordinates, resulting in the following equations that govern the self-similar dynamics:

i(αU+sU+βyyU)𝑖𝛼𝑈subscript𝑠𝑈𝛽𝑦subscript𝑦𝑈\displaystyle i\left(\alpha U+\partial_{s}U+\beta y\partial_{y}U\right)italic_i ( italic_α italic_U + ∂ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT italic_U + italic_β italic_y ∂ start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT italic_U ) =e(2β1)syyU+e(β1)siξ˙yU+e(α1)sU2absentsuperscript𝑒2𝛽1𝑠subscript𝑦𝑦𝑈superscript𝑒𝛽1𝑠𝑖˙𝜉subscript𝑦𝑈superscript𝑒𝛼1𝑠superscript𝑈2\displaystyle=e^{(2\beta-1)s}\partial_{yy}U+e^{(\beta-1)s}i\dot{\xi}\partial_{% y}U+e^{(\alpha-1)s}U^{2}= italic_e start_POSTSUPERSCRIPT ( 2 italic_β - 1 ) italic_s end_POSTSUPERSCRIPT ∂ start_POSTSUBSCRIPT italic_y italic_y end_POSTSUBSCRIPT italic_U + italic_e start_POSTSUPERSCRIPT ( italic_β - 1 ) italic_s end_POSTSUPERSCRIPT italic_i over˙ start_ARG italic_ξ end_ARG ∂ start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT italic_U + italic_e start_POSTSUPERSCRIPT ( italic_α - 1 ) italic_s end_POSTSUPERSCRIPT italic_U start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (22)

For consistency in scaling, balance laws suggest that α=2β𝛼2𝛽\alpha=2\betaitalic_α = 2 italic_β. Rounding to the closest integer our statistical fits for α𝛼\alphaitalic_α, we obtain scaling parameters (α,β)=(1,1/2)𝛼𝛽112(\alpha,\beta)=(1,1/2)( italic_α , italic_β ) = ( 1 , 1 / 2 ) for the real initial data and (α,β)=(2,1)𝛼𝛽21(\alpha,\beta)=(2,1)( italic_α , italic_β ) = ( 2 , 1 ) for the monochromatic initial data. Making the change of variables into these self-similar coordinates, we plot the solutions U(s,y)𝑈𝑠𝑦U(s,y)italic_U ( italic_s , italic_y ) in Figure 5-6(c-d) for the real and imaginary components of the solutions.

In Figure 5-6(c-d) the self-similar solutions U(s,y)𝑈𝑠𝑦U(s,y)italic_U ( italic_s , italic_y ) appear to be decently regularized. For smaller values of s𝑠sitalic_s, periodic copies of the blowup profile may be observed to diverge away from y=0𝑦0y=0italic_y = 0, as would be expected. For larger values of s𝑠sitalic_s however the blowup profile appears to fade, a likely result of an imprecise selection of the blowup time, and the significant numerical error which accumulate as the blowup time is approached. While the solution with real initial data appears to grow according to a power law starting at t0.06𝑡0.06t\geq 0.06italic_t ≥ 0.06, the blowup profile qualitatively changes. This may be most prominently seen in Figure 5(d) when comparing the imaginary component of U(s,y)𝑈𝑠𝑦U(s,y)italic_U ( italic_s , italic_y ) over the regions s[6,7]𝑠67s\in[6,7]italic_s ∈ [ 6 , 7 ] and s[8,10]𝑠810s\in[8,10]italic_s ∈ [ 8 , 10 ], and may be indicative of non-trivial self-similar dynamics.

Alternatively, the simplest blowup scenario would be the so-called Type-I blowup [EF08, QS19], whereby U𝑈Uitalic_U converges to an equilibrium with scaling parameters (α,β)=(1,1/2)𝛼𝛽112(\alpha,\beta)=(1,1/2)( italic_α , italic_β ) = ( 1 , 1 / 2 ) and (22) simplifies to:

i(U+sU+y2yU)𝑖𝑈subscript𝑠𝑈𝑦2subscript𝑦𝑈\displaystyle i\left(U+\partial_{s}U+\tfrac{y}{2}\partial_{y}U\right)italic_i ( italic_U + ∂ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT italic_U + divide start_ARG italic_y end_ARG start_ARG 2 end_ARG ∂ start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT italic_U ) =yyU+es/2iξ˙yU+U2absentsubscript𝑦𝑦𝑈superscript𝑒𝑠2𝑖˙𝜉subscript𝑦𝑈superscript𝑈2\displaystyle=\partial_{yy}U+e^{-s/2}i\dot{\xi}\partial_{y}U+U^{2}= ∂ start_POSTSUBSCRIPT italic_y italic_y end_POSTSUBSCRIPT italic_U + italic_e start_POSTSUPERSCRIPT - italic_s / 2 end_POSTSUPERSCRIPT italic_i over˙ start_ARG italic_ξ end_ARG ∂ start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT italic_U + italic_U start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (23)

In this case, as limses/2=0subscript𝑠superscript𝑒𝑠20\lim_{s\to\infty}e^{-s/2}=0roman_lim start_POSTSUBSCRIPT italic_s → ∞ end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - italic_s / 2 end_POSTSUPERSCRIPT = 0, an equilibrium solution would satisfy the following second order ODE:

00\displaystyle 0 =yyUi2yyUiU+U2,absentsubscript𝑦𝑦𝑈𝑖2𝑦subscript𝑦𝑈𝑖𝑈superscript𝑈2\displaystyle=\partial_{yy}U-\tfrac{i}{2}y\partial_{y}U-iU+U^{2},= ∂ start_POSTSUBSCRIPT italic_y italic_y end_POSTSUBSCRIPT italic_U - divide start_ARG italic_i end_ARG start_ARG 2 end_ARG italic_y ∂ start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT italic_U - italic_i italic_U + italic_U start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , limy±|U(y)|=0subscript𝑦plus-or-minus𝑈𝑦0\displaystyle\lim_{y\to\pm\infty}|U(y)|=0roman_lim start_POSTSUBSCRIPT italic_y → ± ∞ end_POSTSUBSCRIPT | italic_U ( italic_y ) | = 0 (24)

Despite our efforts, we have not found a nontrivial solution to (24).

5 Conclusion

While Conjecture 1.1 may be true in a generic sense, we have presented numerical evidence of real initial data to (3) which blows up in finite time. By tracking solutions to the 1-parameter family of initial data in (5), we were able to identify initial conditions leading to blowup. While tracking the Lsuperscript𝐿L^{\infty}italic_L start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT norm of solutions proved sufficient, measuring when solutions entered a trap** region of zero provided a much more robust method of identification.

As previously mentioned, the blowup solution identified in Section 2 has two distinct qualitative features: (i) on shorter time scales, the solution periodic oscillates with progressive excitement of the higher modes; (ii) on longer time scales, the solution steadily grows, eventually leading to blowup. In Section 3 we provide a heuristic explanation for the mechanisms behind feature (i). Namely, we use the parameterization method to analyze a submanifold 𝒲𝒲\mathcal{W}caligraphic_W of the center manifold of 00. While solutions on 𝒲𝒲\mathcal{W}caligraphic_W do oscillate, a secular drift due to a resonance of eigenvalues induces a cascade whereby the lower modes excite the higher modes.

This analysis leads us to propose Conjecture 1.2, that real initial data to (3) will blowup if and only if it lies on the stable manifold of 00 for the nonlinear heat equation (1). We note that the finite Galerkin truncation model only has a finite number of eigenvalue resonances, and is thus amenable to the parameterization method. While empirical evidence supports Conjecture 1.2, we cannot hope to prove it using the parameterization method due to the infinite number of resonances in the full PDE.

To analyze the later stage of blowup in (3) we performed a self-similar analysis in Section 4, comparing blowup solutions starting from both real and monochromatic initial data. While both solutions appear to exhibit self-similar blowup, key questions remain. For example, what is the exact scaling rate for the solution starting with real initial data, and is a logarithmic correction term required? And does the self-similar solution limit to an equilibrium or some more complicated dynamical object, such as a periodic orbit or a chaotic attractor?

Acknowledgments

The author would like to thank Panayotis Kevrekidis and Javier Gómez-Serrano for fruitful discussions regarding this work.

References

  • [BK94] Jean Bricmont and Antti Kupiainen. Universality in blow-up for nonlinear heat equations. Nonlinearity, 7(2):539, 1994.
  • [CFdlL03a] Xavier Cabré, Ernest Fontich, and Rafael de la Llave. The parameterization method for invariant manifolds I: manifolds associated to non-resonant subspaces. Indiana University mathematics journal, pages 283–328, 2003.
  • [CFdlL03b] Xavier Cabré, Ernest Fontich, and Rafael de la Llave. The parameterization method for invariant manifolds II: regularity with respect to parameters. Indiana University mathematics journal, pages 329–360, 2003.
  • [CFDLL05] Xavier Cabré, Ernest Fontich, and Rafael De La Llave. The parameterization method for invariant manifolds III: overview and applications. Journal of Differential Equations, 218(2):444–515, 2005.
  • [CLM85] Peter Constantin, Peter D Lax, and Andrew Majda. A simple one-dimensional model for the three-dimensional vorticity equation. Communications on pure and applied mathematics, 38(6):715–724, 1985.
  • [CM18] Ciro S Campolina and Alexei A Mailybaev. Chaotic blowup in the 3d incompressible Euler equations on a logarithmic lattice. Physical review letters, 121(6):064501, 2018.
  • [COS16] C.-H. Cho, H. Okamoto, and M. Shōji. A blow-up problem for a nonlinear heat equation in the complex plane of time. Japan Journal of Industrial and Applied Mathematics, 33(1):145–166, Feb 2016.
  • [DHT14] Tobin A Driscoll, Nicholas Hale, and Lloyd N Trefethen. Chebfun guide, 2014.
  • [EF08] Jens Eggers and Marco A Fontelos. The role of self-similarity in singularities of partial differential equations. Nonlinearity, 22(1):R1, 2008.
  • [Fef00] Charles L Fefferman. Existence and smoothness of the Navier-stokes equation. The millennium prize problems, 57:67, 2000.
  • [FMS23] Yue Feng, Georg Maierhofer, and Katharina Schratz. Long-time error bounds of low-regularity integrators for nonlinear Schrödinger equations. arXiv preprint arXiv:2302.00383, 2023.
  • [FS24] Bernold Fiedler and Hannes Stuke. Real eternal PDE solutions are not complex entire: a quadratic parabolic example. arXiv preprint arXiv:2403.06490, 2024.
  • [GJT22] Jorge Gonzalez, JD Mireles James, and Necibe Tuncer. Finite element approximation of invariant manifolds by the parameterization method. Partial Differential Equations and Applications, 3(6):75, 2022.
  • [GNSY13] Jong-Shenq Guo, Hirokazu Ninomiya, Masahiko Shimojo, and Eiji Yanagida. Convergence and blow-up of solutions for a complex-valued heat equation with a quadratic nonlinearity. Transactions of the American Mathematical Society, 365(5):2447–2467, 2013.
  • [HCF+16] Alex Haro, Marta Canadell, Jordi-Lluis Figueras, Alejandro Luque, and Josep-Maria Mondelo. The parameterization method for invariant manifolds. Applied mathematical sciences, 195, 2016.
  • [Hen73] Jacques Henrard. Lyapunov’s center theorem for resonant equilibrium. Journal of Differential Equations, 14(3):431–441, 1973.
  • [HL08] Thomas Y Hou and Congming Li. Dynamic stability of the three-dimensional axisymmetric Navier-Stokes equations with swirl. Communications on Pure and Applied Mathematics: A Journal Issued by the Courant Institute of Mathematical Sciences, 61(5):661–697, 2008.
  • [Hou23] Thomas Y Hou. Potentially singular behavior of the 3d Navier–Stokes equations. Foundations of Computational Mathematics, 23(6):2251–2299, 2023.
  • [Jaq22] Jonathan Jaquette. Quasiperiodicity and blowup in integrable subsystems of nonconservative nonlinear Schrödinger equations. Journal of Dynamics and Differential Equations, pages 1–25, 2022.
  • [Jaq24] Jonathan Jaquette. Matlab code of “Mechanisms of unstable blowup in a quadratic nonlinear Schrödinger equation”. https://github.com/JCJaquette/Unstable-Blowup-NLS, 2024.
  • [JH22] Shobhit Jain and George Haller. How to compute invariant manifolds and their reduced dynamics in high-dimensional finite element models. Nonlinear dynamics, 107(2):1417–1450, 2022.
  • [JLT22a] Jonathan Jaquette, Jean-Philippe Lessard, and Akitoshi Takayasu. Global dynamics in nonconservative nonlinear Schrödinger equations. Advances in Mathematics, 398:108234, 2022.
  • [JLT22b] Jonathan Jaquette, Jean-Philippe Lessard, and Akitoshi Takayasu. Singularities and heteroclinic connections in complex-valued evolutionary equations with a quadratic nonlinearity. Communications in Nonlinear Science and Numerical Simulation, 107:106188, 2022.
  • [Mas83] Kyûya Masuda. Blow-up of solutions of some nonlinear diffusion equations. In Hiroshi Fujita, Peter D. Lax, and Gilbert Strang, editors, Nonlinear Partial Differential Equations in Applied Science; Proceedings of The U.S.-Japan Seminar, Tokyo, 1982, volume 81 of North-Holland Mathematics Studies, pages 119 – 131. North-Holland, 1983.
  • [Mas84] Kyûya Masuda. Analytic solutions of some nonlinear diffusion equations. Mathematische Zeitschrift, 187(1):61–73, Mar 1984.
  • [MZ97] Frank Merle and Hatem Zaag. Stability of the blow-up profile for equations of the type ut=Δu+|u|p1usubscript𝑢𝑡Δ𝑢superscript𝑢𝑝1𝑢u_{t}=\Delta u+|u|^{p-1}uitalic_u start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT = roman_Δ italic_u + | italic_u | start_POSTSUPERSCRIPT italic_p - 1 end_POSTSUPERSCRIPT italic_u. Duke Mathematical Journal, 86(1):143 – 195, 1997.
  • [OVTF23] Andrea Opreni, Alessandra Vizzaccaro, Cyril Touzé, and Attilio Frangi. High-order direct parametrisation of invariant manifolds for model order reduction of finite element structures: application to generic forcing terms and parametrically excited systems. Nonlinear Dynamics, 111(6):5401–5447, 2023.
  • [Pro22] Bartosz Protas. Systematic search for extreme and singular behaviour in some fundamental models of fluid mechanics. Philosophical Transactions of the Royal Society A, 380(2225):20210035, 2022.
  • [QS19] Pavol Quittner and Philippe Souplet. Superlinear parabolic problems. Springer, 2019.
  • [RJ19] Christian Reinhardt and JD Mireles James. Fourier–Taylor parameterization of unstable manifolds for parabolic partial differential equations: Formalism, implementation and rigorous validation. Indagationes Mathematicae, 30(1):39–80, 2019.
  • [SY02] George R Sell and Yuncheng You. Dynamics of evolutionary equations, volume 143. Springer Science & Business Media, 2002.
  • [vdBMJR16] Jan Bouwe van den Berg, Jason D Mireles James, and Christian Reinhardt. Computing (un) stable manifolds with validated error bounds: non-resonant and resonant spectra. Journal of Nonlinear Science, 26:1055–1095, 2016.
  • [WLGSB23] Yongji Wang, C-Y Lai, Javier Gómez-Serrano, and Tristan Buckmaster. Asymptotic self-similar blow-up profile for three-dimensional axisymmetric Euler equations using neural networks. Physical Review Letters, 130(24):244002, 2023.