License: CC BY 4.0
arXiv:2404.02008v1 [physics.flu-dyn] 02 Apr 2024

Singularity formation of vortex sheets in 2D Euler equations using the characteristic mapping method

Julius Bergmann\aff1,2 \corresp julius.bergmann .at. univ-amu.fr    Thibault Maurel–Oujia\aff1    Xi–Yuan (Bruce) Yin\aff3    Jean–Christophe Nave\aff4    Kai Schneider\aff1 \aff1Institut de Mathématiques de Marseille, Aix–Marseille Université, CNRS, Marseille, France \aff2Institut für Strömungsmechanik und Technische Akustik, Technische Universität Berlin, Berlin, Germany \aff3LMFA–CNRS, Ecole Centrale de Lyon, Lyon, France \aff4Department of Mathematics and Statistics, McGill University, Montréal, Québec, Canada
Abstract

The goal of this numerical study is to get insight into singular solutions of the two-dimensional (2D) Euler equations for non-smooth initial data, in particular for vortex sheets. To this end high resolution computations of vortex layers in 2D incompressible Euler flows are performed using the characteristic mapping method (CMM). This semi-Lagrangian method evolves the flow map using the gradient-augmented level set method (GALS). The semi-group structure of the flow map allows its decomposition into sub-maps (each over a finite time interval), and thus the precision can be controlled by choosing appropriate remapping times. Composing the flow map yields exponential resolution in linear time, a unique feature of CMM, and thus fine scale flow structures can be resolved in great detail. Here the roll-up process of vortex layers is studied varying the thickness of the layer showing its impact on the growth of palinstrophy and possible blow up of absolute vorticity. The curvature of the vortex sheet shows a singular-like behavior. The self-similar structure of the vortex core is investigated in the vanishing thickness limit. Conclusions on the non-uniqueness of weak solutions of 2D Euler for non-smooth initial data are drawn and the presence of flow singularities is revealed tracking them in the complex plane.

keywords:

1 Introduction


The emergence of coherent vorticity is ubiquitous in high Reynolds number turbulence, as for example in shear layers or in wall bounded flow in form of thin boundary layers detaching and traveling into the bulk flow. The dynamics of thin vortex layers in the limit of vanishing viscosity is of tremendous interest for understanding the formation of singularities in the incompressible 2D Euler equations with non-smooth initial data. In contrast to 2D Euler with smooth initial conditions, for which results on global regularity and uniqueness of the solution are well known (Yudovich, 1963), it was shown by Székelyhidi (2011) that for vortex sheet initial data infinitely many non-stationary weak solutions exist, which moreover conserve energy. For a review on the mathematics of turbulence we refer to (Majda & Bertozzi, 2001; Bardos & Titi, 2013) and for 2D vortex layers we refer to the detailed introduction of Caflisch et al. (2022).

In a recent work Caflisch et al. (2022) showed by means of direct numerical simulation using a classical pseudo-spectral method that 2D vortex layers may have complex singularities and after the roll up of the layer that the vortex cores may have unbounded vorticity in the limit of infinite Reynolds numbers. We recall that Caflisch et al. (2022) did almost exclusively viscous computations by solving Navier–Stokes using Fourier pseudo-spectral methods and coupling the vortex layer thickness with the viscosity, i.e. the inverse Reynolds number. Some computations for the inviscid case, i.e. for 2D Euler, were likewise presented in Caflisch et al. (2017) and Caflisch et al. (2022). However, the inviscid computations were limited to minimum layer thickness values (δ=0.0141𝛿0.0141\delta=0.0141italic_δ = 0.0141). Note that smaller values did not work due to the necessary resolution requirements, however they also used larger values (cf. table 3 in Caflisch et al. (2022)). The motion of an inviscid vortex sheet, i.e. for vanishing thickness, is governed by the Birkhoff–Rott model equation (Birkhoff, 1962; Rott, 1956) which develops a singularity in finite time starting from smooth initial data (Moore & Stuart, 1979). Perturbations grow due to the Kelvin–Helmholtz instability and the vortex sheet does roll up. Regularized simulations using a vortex blob method have been also performed in Caflisch et al. (2022) and compared with Navier–Stokes computations.

The vanishing viscosity limit of 2D Navier–Stokes in the presence of boundaries was likewise studied in Nguyen Van Yen et al. (2011, 2018). Dipole-wall collisions were simulated and the existence of a Reynolds-independent energy dissipation rate was shown. In this context Prandtl’s classical boundary layer argument was complemented, which states that both the boundary-layer thickness and dissipation rate are proportional to Re1/2.𝑅superscript𝑒12Re^{-1/2}.italic_R italic_e start_POSTSUPERSCRIPT - 1 / 2 end_POSTSUPERSCRIPT . However for detaching boundary layers, Kato’s scaling was shown to be more appropriate than Prandtl’s scaling, which implies that the boundary layer scales with Re1𝑅superscript𝑒1Re^{-1}italic_R italic_e start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT. Some reviews of 2D flows with walls can be found in Clercx & Van Heijst (2017) and mathematical analysis of weak solutions of the 2D Navier–Stokes equations in bounded domains, in the vanishing viscosity limit, in Constantin et al. (2019).

State of the art for solving Navier–Stokes or Euler equations numerically with high precision are pseudo-spectral methods (Hussaini & Zang, 1987; Ishihara et al., 2009) which have been also extensively used for investigating nearly singular solutions of the 3D Euler equations (Hou & Li, 2007; Gibbon et al., 2008). A Cauchy–Lagrange method for computing 2D Euler flows was proposed in Podvigina et al. (2016). This semi-Lagrangian method exploits the time-analyticity of fluid particle trajectories and was shown to be more efficient than pseudo-spectral computations. However, detailed singularity studies have not been reported so far. An even more powerful tool for solving the incompressible Euler equations is the Characterisitic Mapping Method (CMM). Evolving the flow map with a Gradient Augmented Level Set Method, developed in (Nave et al., 2010; Seibold et al., 2012; Chidyagwai et al., 2011), one can decompose the long time deformation into short time sub-maps due to the semi-group structure of the flow map. This yields a numerical scheme with exponential resolution in linear time developed for linear advection in Mercier et al. (2020) and 2D Euler in (Yin et al., 2021), allowing to capture the exponential growth of vorticity gradients. The implementation of the method has global third order convergence in space and time and its efficiency has been demonstrated in comparison with spectral and Cauchy–Lagrange methods in (Yin et al., 2021). More recently an extension to 3D incompressible Euler flows has been proposed in (Yin et al., 2023). The compositional adaptivity of CMM is an essential feature which allows detailed insight into the small scales of the solution without using prohibitive numerical resolutions.

The goal of the present paper is to revisit the vortex layer computations of Caflisch et al. (2022) in the inviscid case in more depth using CMM and to extend the range to smaller δ𝛿\deltaitalic_δ-values and longer times. As presented in Caflisch et al. (2017), for too thick vortex layers the vortex merging process starts to interfere with the formation of the two vortex blobs with spiral arms. Here we compute solutions of 2D incompressible Euler flows and study the dynamics of these extremely thin vortex layers in the vanishing thickness limit and investigate possible singularities. The aim is to get some insights into the non-uniqueness of week solutions and about possible singularities. Curvature and vortex strength of the vortex center-line are analyzed and a temporal and spatial normalization unveils the dynamics for vanishing thickness limit, as well as the investigation of singularities in the complex plane. The palinstrophy growth and energy spectra show and distinguish the impact of both the forming vortices and vortex merger process. Thanks to the high resolution capabilities of CMM we get insight into the fine scale structure of vortex cores and their dynamics in Euler flows.

The remainder of the paper is organized as follows. Set-up and initial conditions are discussed in section 2. A short description of the characteristic mapping method for solving the 2D incompressible Euler equations is given in section 3. Section 4 gives an overview on the performed computations and numerical results are then presented in section 5. A singularity analysis in the complex plane is performed in section 6. Finally, conclusions and perspectives for future work are given in section 7.

2 Governing equations and initial condition

We consider inviscid flow in a 2π2𝜋2\pi2 italic_π periodic domain Ω=[0,2π]×[0,2π]Ω02𝜋02𝜋\Omega=[0,2\pi]\times[0,2\pi]roman_Ω = [ 0 , 2 italic_π ] × [ 0 , 2 italic_π ] in the plane, governed by the 2D incompressible Euler equations. Starting point is the vorticity transport equation,

tω+(𝒖)ω=0subscript𝑡𝜔𝒖𝜔0\partial_{t}\omega+({\bm{u}}\cdot\nabla)\omega=0∂ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_ω + ( bold_italic_u ⋅ ∇ ) italic_ω = 0 (1)

where the vorticity is defined as ω=×𝒖𝜔𝒖\omega=\nabla\times{\bm{u}}italic_ω = ∇ × bold_italic_u and 𝒖𝒖\bm{u}bold_italic_u is the incompressible velocity, satisfying 𝒖=0𝒖0\nabla\cdot\bm{u}=0∇ ⋅ bold_italic_u = 0. The curl operator is invertible and the velocity can be computed from the vorticity, 𝒖=×Δ1ω𝒖superscriptΔ1𝜔\bm{u}=-\nabla\times\Delta^{-1}\omegabold_italic_u = - ∇ × roman_Δ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT italic_ω using the Biot–Savart operator. Equation 1 is completed with a suitable initial condition ω0(𝒙)=ω(𝒙,t=0)subscript𝜔0𝒙𝜔𝒙𝑡0\omega_{0}({\bm{x}})=\omega({\bm{x}},t=0)italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( bold_italic_x ) = italic_ω ( bold_italic_x , italic_t = 0 ), here a regularized vortex sheet and where 𝒙=(x,y)𝒙𝑥𝑦\bm{x}=(x,y)bold_italic_x = ( italic_x , italic_y ) with (x,y)Ω𝑥𝑦Ω(x,y)\in\Omega( italic_x , italic_y ) ∈ roman_Ω as the cardinal directions. All variables have been non-dimensionalized, similar to Caflisch et al. (2022), as follows,

𝒙=𝒙*1λ𝒙superscript𝒙1𝜆\bm{x}=\bm{x}^{*}\frac{1}{\lambda}bold_italic_x = bold_italic_x start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG italic_λ end_ARG (2a)
t=t*Γλ2𝑡superscript𝑡Γsuperscript𝜆2t=t^{*}\frac{\Gamma}{\lambda^{2}}italic_t = italic_t start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT divide start_ARG roman_Γ end_ARG start_ARG italic_λ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG (2b)
𝒖=𝒖*λΓ𝒖superscript𝒖𝜆Γ\bm{u}=\bm{u^{*}}\frac{\lambda}{\Gamma}bold_italic_u = bold_italic_u start_POSTSUPERSCRIPT bold_* end_POSTSUPERSCRIPT divide start_ARG italic_λ end_ARG start_ARG roman_Γ end_ARG (2c)
𝝎=𝝎*λ2Γ𝝎superscript𝝎superscript𝜆2Γ\bm{\omega}=\bm{\omega^{*}}\frac{\lambda^{2}}{\Gamma}bold_italic_ω = bold_italic_ω start_POSTSUPERSCRIPT bold_* end_POSTSUPERSCRIPT divide start_ARG italic_λ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG roman_Γ end_ARG (2d)

With the domain length Lx=Ly=2πsubscript𝐿𝑥subscript𝐿𝑦2𝜋L_{x}=L_{y}=2\piitalic_L start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT = italic_L start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT = 2 italic_π we have for the spatial scaling λ=Lx/(2π)=1𝜆subscript𝐿𝑥2𝜋1\lambda=L_{x}/(2\pi)=1italic_λ = italic_L start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT / ( 2 italic_π ) = 1. The circulation ΓΓ\Gammaroman_Γ equals 2π2𝜋2\pi2 italic_π, as shown below. The regularized vortex sheet initial condition proposed in Caflisch et al. (2022) reads,

ω0(x,y)=12πδexp((yπϕ(x))22δ2)subscript𝜔0𝑥𝑦12𝜋𝛿superscript𝑦𝜋italic-ϕ𝑥22superscript𝛿2\displaystyle\omega_{0}(x,y)=\frac{1}{\sqrt{2\pi}\delta}\exp\left(-\frac{(y-% \pi-\phi(x))^{2}}{2\delta^{2}}\right)italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_x , italic_y ) = divide start_ARG 1 end_ARG start_ARG square-root start_ARG 2 italic_π end_ARG italic_δ end_ARG roman_exp ( - divide start_ARG ( italic_y - italic_π - italic_ϕ ( italic_x ) ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_δ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) (3)

where δ𝛿\deltaitalic_δ is the thickness parameter and ϕitalic-ϕ\phiitalic_ϕ a perturbation function. The initial field corresponds to a vorticity line with Gaussian cross section of thickness δ𝛿\deltaitalic_δ centered around a perturbation function ϕ(x)italic-ϕ𝑥\phi(x)italic_ϕ ( italic_x ) shown in figure 1. The profile for ϕ(x)italic-ϕ𝑥\phi(x)italic_ϕ ( italic_x ) is sinusoidal with ϕ(x)=sin(x)/2italic-ϕ𝑥𝑥2\phi(x)=\sin(x)/2italic_ϕ ( italic_x ) = roman_sin ( italic_x ) / 2 and (x,y)Ω𝑥𝑦Ω(x,y)\in\Omega( italic_x , italic_y ) ∈ roman_Ω. The position of the center line ϕitalic-ϕ\phiitalic_ϕ is the curve of maximum vorticity, which oscillates around y=π𝑦𝜋y=\piitalic_y = italic_π. For δ0𝛿0\delta\rightarrow 0italic_δ → 0 this vortex layer converges towards a vortex sheet used in Caflisch et al. (2022) as initial condition for the Birkhoff–Rott equation.

Refer to caption
Refer to caption
(a) Initial condition ω0subscript𝜔0\omega_{0}italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT for δ0.032𝛿0.032\delta\approx 0.032italic_δ ≈ 0.032
Refer to caption
(b) Energy spectrum for different δ𝛿\deltaitalic_δ-values
Figure 1: Initial condition for δ0.032𝛿0.032\delta\approx 0.032italic_δ ≈ 0.032 and energy spectra for different δ𝛿\deltaitalic_δ-values.

The energy spectrum of the initial condition describes the initial frequencies present in the flow field. It is defined as:

E(k,t)=12k1/2|𝒌|<k+1/2|𝒖^(𝒌,t)|2𝐸𝑘𝑡12subscript𝑘12𝒌𝑘12superscript^𝒖𝒌𝑡2\displaystyle E(k,t)=\frac{1}{2}\sum_{k-1/2\leq|{\bm{k}}|<k+1/2}|\widehat{\bm{% u}}({\bm{k},t})|^{2}italic_E ( italic_k , italic_t ) = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ∑ start_POSTSUBSCRIPT italic_k - 1 / 2 ≤ | bold_italic_k | < italic_k + 1 / 2 end_POSTSUBSCRIPT | over^ start_ARG bold_italic_u end_ARG ( bold_italic_k , italic_t ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (4)

Here ^^\widehat{\cdot}over^ start_ARG ⋅ end_ARG denotes the 2D Fourier transform. The initial condition, a vortex line regularized with a Gaussian cross section, does in the limit of δ0𝛿0\delta\rightarrow 0italic_δ → 0 approach a vortex line, i.e. a Dirac distribution. The enstrophy spectrum thus exhibits a k0superscript𝑘0k^{0}italic_k start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT scaling and consequently the energy spectrum yields a k2superscript𝑘2k^{-2}italic_k start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT scaling, using the relation E(k,t)=k2Z(k,t)𝐸𝑘𝑡superscript𝑘2𝑍𝑘𝑡E(k,t)=k^{-2}Z(k,t)italic_E ( italic_k , italic_t ) = italic_k start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT italic_Z ( italic_k , italic_t ). For large δ𝛿\deltaitalic_δ values the regularization becomes more and more visible resulting in a faster, i.e. exponential decay for large wave-numbers, as illustrated in figure 0(b). For viscous flow simulations, considered in Caflisch et al. (2022), δ𝛿\deltaitalic_δ was related via δ=Re1/2𝛿𝑅superscript𝑒12\delta=Re^{-1/2}italic_δ = italic_R italic_e start_POSTSUPERSCRIPT - 1 / 2 end_POSTSUPERSCRIPT, where Re=Γ/ν𝑅𝑒Γ𝜈Re=\Gamma/\nuitalic_R italic_e = roman_Γ / italic_ν is the circulation based Reynolds number with Γ=ω(𝒙,t=0)𝑑𝒙Γ𝜔𝒙𝑡0differential-d𝒙\Gamma=\int\omega({\bm{x}},t=0)d{\bm{x}}roman_Γ = ∫ italic_ω ( bold_italic_x , italic_t = 0 ) italic_d bold_italic_x being the initial circulation of the vortex layer and ν𝜈\nuitalic_ν being the kinematic viscosity. Note that in the present study ν𝜈\nuitalic_ν vanishes. For the given initial condition re-shifted to Ωs=[π,π]×[π,π]subscriptΩ𝑠𝜋𝜋𝜋𝜋\Omega_{s}=[-\pi,\pi]\times[-\pi,\pi]roman_Ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = [ - italic_π , italic_π ] × [ - italic_π , italic_π ], the initial circulation is

Ωsω0(𝒙)𝑑𝒙subscriptsubscriptΩ𝑠subscript𝜔0𝒙differential-d𝒙\displaystyle\int_{\Omega_{s}}\omega_{0}(\bm{x})d\bm{x}∫ start_POSTSUBSCRIPT roman_Ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( bold_italic_x ) italic_d bold_italic_x =ππππ12πδexp((yϕ(x))22δ2)𝑑x𝑑yabsentsuperscriptsubscript𝜋𝜋superscriptsubscript𝜋𝜋12𝜋𝛿superscript𝑦italic-ϕ𝑥22superscript𝛿2differential-d𝑥differential-d𝑦\displaystyle=\int_{-\pi}^{\pi}\int_{-\pi}^{\pi}\frac{1}{\sqrt{2\pi}\delta}% \exp{\left(-\frac{(y-\phi(x))^{2}}{2\delta^{2}}\right)}dxdy= ∫ start_POSTSUBSCRIPT - italic_π end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_π end_POSTSUPERSCRIPT ∫ start_POSTSUBSCRIPT - italic_π end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_π end_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG square-root start_ARG 2 italic_π end_ARG italic_δ end_ARG roman_exp ( - divide start_ARG ( italic_y - italic_ϕ ( italic_x ) ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_δ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) italic_d italic_x italic_d italic_y (5)
=ππ(πϕ(x))/2δ(πϕ(x))/2δ1πexp(v2)𝑑v𝑑xabsentsuperscriptsubscript𝜋𝜋superscriptsubscript𝜋italic-ϕ𝑥2𝛿𝜋italic-ϕ𝑥2𝛿1𝜋superscript𝑣2differential-d𝑣differential-d𝑥\displaystyle=\int_{-\pi}^{\pi}\int_{(-\pi-\phi(x))/2\delta}^{(\pi-\phi(x))/2% \delta}\frac{1}{\sqrt{\pi}}\exp{\left(-v^{2}\right)}dvdx= ∫ start_POSTSUBSCRIPT - italic_π end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_π end_POSTSUPERSCRIPT ∫ start_POSTSUBSCRIPT ( - italic_π - italic_ϕ ( italic_x ) ) / 2 italic_δ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_π - italic_ϕ ( italic_x ) ) / 2 italic_δ end_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG square-root start_ARG italic_π end_ARG end_ARG roman_exp ( - italic_v start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) italic_d italic_v italic_d italic_x
=ππ12[erf(v)](πϕ(x))/2δ(πϕ(x))/2δ𝑑x=2π+o(exp(δ2)).absentsuperscriptsubscript𝜋𝜋12superscriptsubscriptdelimited-[]erf𝑣𝜋italic-ϕ𝑥2𝛿𝜋italic-ϕ𝑥2𝛿differential-d𝑥2𝜋𝑜superscript𝛿2\displaystyle=\int_{-\pi}^{\pi}\frac{1}{2}\Bigl{[}{\rm{erf}}(v)\Bigr{]}_{(-\pi% -\phi(x))/2\delta}^{(\pi-\phi(x))/2\delta}dx=2\pi+o(\exp(-\delta^{-2})).= ∫ start_POSTSUBSCRIPT - italic_π end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_π end_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG 2 end_ARG [ roman_erf ( italic_v ) ] start_POSTSUBSCRIPT ( - italic_π - italic_ϕ ( italic_x ) ) / 2 italic_δ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_π - italic_ϕ ( italic_x ) ) / 2 italic_δ end_POSTSUPERSCRIPT italic_d italic_x = 2 italic_π + italic_o ( roman_exp ( - italic_δ start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT ) ) .

The evaluation limits (±πϕ(x))/2δplus-or-minus𝜋italic-ϕ𝑥2𝛿(\pm\pi-\phi(x))/2\delta( ± italic_π - italic_ϕ ( italic_x ) ) / 2 italic_δ arise from the truncation of the Gaussian profile by the periodic box (on the upper and lower sides of which there is technically a discontinuity).

Here erf(z)=2π0zexp(v2)𝑑verf𝑧2𝜋superscriptsubscript0𝑧superscript𝑣2differential-d𝑣{\rm erf}(z)=\frac{2}{\sqrt{\pi}}\int_{0}^{z}\exp(-v^{2})dvroman_erf ( italic_z ) = divide start_ARG 2 end_ARG start_ARG square-root start_ARG italic_π end_ARG end_ARG ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT roman_exp ( - italic_v start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) italic_d italic_v is the error function which rapidly approaches 1111 for z𝑧z\rightarrow\inftyitalic_z → ∞, i.e. for δ0𝛿0\delta\rightarrow 0italic_δ → 0. Hence the error function can be approximated by 1111 for all investigated thickness values δ𝛿\deltaitalic_δ to a certainty far below machine precision, justified by the expansion erf(x)=1ex21π(1x12x3+34x5158x7)+o(x8ex2)erf𝑥1superscript𝑒superscript𝑥21𝜋1𝑥12superscript𝑥334superscript𝑥5158superscript𝑥7𝑜superscript𝑥8superscript𝑒superscript𝑥2{\rm erf}(x)=1-e^{-x^{2}}\frac{1}{\sqrt{\pi}}\left(\frac{1}{x}-\frac{1}{2x^{3}% }+\frac{3}{4x^{5}}-\frac{15}{8x^{7}}\right)+o(x^{-8}e^{-x^{2}})roman_erf ( italic_x ) = 1 - italic_e start_POSTSUPERSCRIPT - italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG square-root start_ARG italic_π end_ARG end_ARG ( divide start_ARG 1 end_ARG start_ARG italic_x end_ARG - divide start_ARG 1 end_ARG start_ARG 2 italic_x start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG + divide start_ARG 3 end_ARG start_ARG 4 italic_x start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT end_ARG - divide start_ARG 15 end_ARG start_ARG 8 italic_x start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPT end_ARG ) + italic_o ( italic_x start_POSTSUPERSCRIPT - 8 end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT ) (Abramowitz & Stegun, 1964). The term [erf(v)](πϕ(x))/δ(πϕ(x))/δsuperscriptsubscriptdelimited-[]erf𝑣𝜋italic-ϕ𝑥𝛿𝜋italic-ϕ𝑥𝛿[{\rm{erf}}(v)]_{(-\pi-\phi(x))/\delta}^{(\pi-\phi(x))/\delta}[ roman_erf ( italic_v ) ] start_POSTSUBSCRIPT ( - italic_π - italic_ϕ ( italic_x ) ) / italic_δ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_π - italic_ϕ ( italic_x ) ) / italic_δ end_POSTSUPERSCRIPT can therefore approximated to 2 with error of order o(exp((π12)2δ2))𝑜superscript𝜋122superscript𝛿2o(\exp(-(\pi-\frac{1}{2})^{2}\delta^{-2}))italic_o ( roman_exp ( - ( italic_π - divide start_ARG 1 end_ARG start_ARG 2 end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_δ start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT ) ) using that |ϕ(x)|12italic-ϕ𝑥12|\phi(x)|\leq\frac{1}{2}| italic_ϕ ( italic_x ) | ≤ divide start_ARG 1 end_ARG start_ARG 2 end_ARG.

For incompressible and inviscid flow, energy E(t)𝐸𝑡E(t)italic_E ( italic_t ) and enstrophy Z(t)𝑍𝑡Z(t)italic_Z ( italic_t ) remain constant over time while the palinstrophy does increase super exponentially (Ayala & Protas, 2014). Those quantities are defined via

E(t)𝐸𝑡\displaystyle E(t)italic_E ( italic_t ) =12Ω|𝒖(𝒙,t)|2𝑑𝒙,absent12subscriptΩsuperscript𝒖𝒙𝑡2differential-d𝒙\displaystyle=\frac{1}{2}\int_{\Omega}|\bm{u}(\bm{x},t)|^{2}d\bm{x}\,,= divide start_ARG 1 end_ARG start_ARG 2 end_ARG ∫ start_POSTSUBSCRIPT roman_Ω end_POSTSUBSCRIPT | bold_italic_u ( bold_italic_x , italic_t ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_d bold_italic_x , (6)
Z(t)𝑍𝑡\displaystyle Z(t)italic_Z ( italic_t ) =12Ω|ω(𝒙,t)|2𝑑𝒙,absent12subscriptΩsuperscript𝜔𝒙𝑡2differential-d𝒙\displaystyle=\frac{1}{2}\int_{\Omega}|\omega(\bm{x},t)|^{2}d\bm{x}\,,= divide start_ARG 1 end_ARG start_ARG 2 end_ARG ∫ start_POSTSUBSCRIPT roman_Ω end_POSTSUBSCRIPT | italic_ω ( bold_italic_x , italic_t ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_d bold_italic_x , (7)
P(t)𝑃𝑡\displaystyle P(t)italic_P ( italic_t ) =12Ω|ω(𝒙,t)|2𝑑𝒙.absent12subscriptΩsuperscript𝜔𝒙𝑡2differential-d𝒙\displaystyle=\frac{1}{2}\int_{\Omega}|\nabla\omega(\bm{x},t)|^{2}d\bm{x}\,.= divide start_ARG 1 end_ARG start_ARG 2 end_ARG ∫ start_POSTSUBSCRIPT roman_Ω end_POSTSUBSCRIPT | ∇ italic_ω ( bold_italic_x , italic_t ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_d bold_italic_x . (8)

With varying thickness δ𝛿\deltaitalic_δ, the initial energy E(t=0)𝐸𝑡0E(t=0)italic_E ( italic_t = 0 ) was found to increase linearly with decreasing vortex sheet thickness with a limit value of 1.52absent1.52\approx 1.52≈ 1.52 for δ0𝛿0\delta\rightarrow 0italic_δ → 0 (figure 2), The initial enstrophy Z(t=0)𝑍𝑡0Z(t=0)italic_Z ( italic_t = 0 ) is shown below to scale with δ1superscript𝛿1\delta^{-1}italic_δ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT and the initial palinstrophy P(t=0)𝑃𝑡0P(t=0)italic_P ( italic_t = 0 ) with δ3superscript𝛿3\delta^{-3}italic_δ start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT, both values thus go to infinity with vanishing vortex sheet thickness. Similar to the circulation, for the shifted domain ΩssubscriptΩ𝑠\Omega_{s}roman_Ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT we get for the initial enstrophy:

12Ωs|ω0(𝒙)|2𝑑𝒙12subscriptsubscriptΩ𝑠superscriptsubscript𝜔0𝒙2differential-d𝒙\displaystyle\frac{1}{2}\int_{\Omega_{s}}|\omega_{0}(\bm{x})|^{2}d\bm{x}divide start_ARG 1 end_ARG start_ARG 2 end_ARG ∫ start_POSTSUBSCRIPT roman_Ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT end_POSTSUBSCRIPT | italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( bold_italic_x ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_d bold_italic_x =12ππππ12πδ2exp((yϕ(x))2δ2)𝑑x𝑑yabsent12superscriptsubscript𝜋𝜋superscriptsubscript𝜋𝜋12𝜋superscript𝛿2superscript𝑦italic-ϕ𝑥2superscript𝛿2differential-d𝑥differential-d𝑦\displaystyle=\frac{1}{2}\int_{-\pi}^{\pi}\int_{-\pi}^{\pi}\frac{1}{2\pi\delta% ^{2}}\exp{\left(-\frac{(y-\phi(x))^{2}}{\delta^{2}}\right)}dxdy= divide start_ARG 1 end_ARG start_ARG 2 end_ARG ∫ start_POSTSUBSCRIPT - italic_π end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_π end_POSTSUPERSCRIPT ∫ start_POSTSUBSCRIPT - italic_π end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_π end_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG 2 italic_π italic_δ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG roman_exp ( - divide start_ARG ( italic_y - italic_ϕ ( italic_x ) ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_δ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) italic_d italic_x italic_d italic_y (9)
=12ππ(πϕ(x))/δ(πϕ(x))/δ12πδexp(v2)𝑑v𝑑xabsent12superscriptsubscript𝜋𝜋superscriptsubscript𝜋italic-ϕ𝑥𝛿𝜋italic-ϕ𝑥𝛿12𝜋𝛿superscript𝑣2differential-d𝑣differential-d𝑥\displaystyle=\frac{1}{2}\int_{-\pi}^{\pi}\int_{(-\pi-\phi(x))/\delta}^{(\pi-% \phi(x))/\delta}\frac{1}{2\pi\delta}\exp{\left(-v^{2}\right)}dvdx= divide start_ARG 1 end_ARG start_ARG 2 end_ARG ∫ start_POSTSUBSCRIPT - italic_π end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_π end_POSTSUPERSCRIPT ∫ start_POSTSUBSCRIPT ( - italic_π - italic_ϕ ( italic_x ) ) / italic_δ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_π - italic_ϕ ( italic_x ) ) / italic_δ end_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG 2 italic_π italic_δ end_ARG roman_exp ( - italic_v start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) italic_d italic_v italic_d italic_x
=12ππ12πδ[erf(v)](πϕ(x))/δ(πϕ(x))/δ𝑑x=π2δ+o(exp(δ2)).absent12superscriptsubscript𝜋𝜋12𝜋𝛿superscriptsubscriptdelimited-[]erf𝑣𝜋italic-ϕ𝑥𝛿𝜋italic-ϕ𝑥𝛿differential-d𝑥𝜋2𝛿𝑜superscript𝛿2\displaystyle=\frac{1}{2}\int_{-\pi}^{\pi}\frac{1}{2\sqrt{\pi}\delta}\Bigl{[}{% \rm{erf}}(v)\Bigr{]}_{(-\pi-\phi(x))/\delta}^{(\pi-\phi(x))/\delta}dx=\frac{% \sqrt{\pi}}{2\delta}+o(\exp(-\delta^{-2})).= divide start_ARG 1 end_ARG start_ARG 2 end_ARG ∫ start_POSTSUBSCRIPT - italic_π end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_π end_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG 2 square-root start_ARG italic_π end_ARG italic_δ end_ARG [ roman_erf ( italic_v ) ] start_POSTSUBSCRIPT ( - italic_π - italic_ϕ ( italic_x ) ) / italic_δ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_π - italic_ϕ ( italic_x ) ) / italic_δ end_POSTSUPERSCRIPT italic_d italic_x = divide start_ARG square-root start_ARG italic_π end_ARG end_ARG start_ARG 2 italic_δ end_ARG + italic_o ( roman_exp ( - italic_δ start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT ) ) .

Respectively, the initial palinstrophy can be computed analytically with the gradient of the initial vorticity:

ω0=(ϕ(x)/δ1/δ)yϕ(x)δω0(x,y)subscript𝜔0matrixsuperscriptitalic-ϕ𝑥𝛿1𝛿𝑦italic-ϕ𝑥𝛿subscript𝜔0𝑥𝑦\nabla\omega_{0}=-\left(\begin{matrix}-\phi^{\prime}(x)/\delta\\ 1/\delta\end{matrix}\right)\frac{y-\phi(x)}{\delta}\omega_{0}(x,y)∇ italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = - ( start_ARG start_ROW start_CELL - italic_ϕ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_x ) / italic_δ end_CELL end_ROW start_ROW start_CELL 1 / italic_δ end_CELL end_ROW end_ARG ) divide start_ARG italic_y - italic_ϕ ( italic_x ) end_ARG start_ARG italic_δ end_ARG italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_x , italic_y ) (10)

We thus obtain

ΩS|ω0|2𝑑𝒙subscriptsubscriptΩ𝑆superscriptsubscript𝜔02differential-d𝒙\displaystyle\int_{\Omega_{S}}|\nabla\omega_{0}|^{2}d\bm{x}∫ start_POSTSUBSCRIPT roman_Ω start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT end_POSTSUBSCRIPT | ∇ italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_d bold_italic_x =12πδ3ππ(πϕ(x))/δ(πϕ(x))/δ(1+(ϕ(x))2)v2exp(v2)𝑑v𝑑xabsent12𝜋superscript𝛿3superscriptsubscript𝜋𝜋superscriptsubscript𝜋italic-ϕ𝑥𝛿𝜋italic-ϕ𝑥𝛿1superscriptsuperscriptitalic-ϕ𝑥2superscript𝑣2superscript𝑣2differential-d𝑣differential-d𝑥\displaystyle=\frac{1}{2\pi\delta^{3}}\int_{-\pi}^{\pi}\int_{(-\pi-\phi(x))/% \delta}^{(\pi-\phi(x))/\delta}\left(1+(\phi^{\prime}(x))^{2}\right)v^{2}\exp(-% v^{2})dvdx= divide start_ARG 1 end_ARG start_ARG 2 italic_π italic_δ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG ∫ start_POSTSUBSCRIPT - italic_π end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_π end_POSTSUPERSCRIPT ∫ start_POSTSUBSCRIPT ( - italic_π - italic_ϕ ( italic_x ) ) / italic_δ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_π - italic_ϕ ( italic_x ) ) / italic_δ end_POSTSUPERSCRIPT ( 1 + ( italic_ϕ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_x ) ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) italic_v start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_exp ( - italic_v start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) italic_d italic_v italic_d italic_x (11)
=18πδ3ππ(1+cos2(x)4)[πerf(v)2vexp(v2)](πϕ(x))/δ(πϕ(x))/δ𝑑xabsent18𝜋superscript𝛿3superscriptsubscript𝜋𝜋1superscript2𝑥4superscriptsubscriptdelimited-[]𝜋erf𝑣2𝑣superscript𝑣2𝜋italic-ϕ𝑥𝛿𝜋italic-ϕ𝑥𝛿differential-d𝑥\displaystyle=\frac{1}{8\pi\delta^{3}}\int_{-\pi}^{\pi}\left(1+\frac{\cos^{2}(% x)}{4}\right)\Bigl{[}\sqrt{\pi}{\rm{erf}}(v)-2v\exp(-v^{2})\Bigr{]}_{(-\pi-% \phi(x))/\delta}^{(\pi-\phi(x))/\delta}dx= divide start_ARG 1 end_ARG start_ARG 8 italic_π italic_δ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG ∫ start_POSTSUBSCRIPT - italic_π end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_π end_POSTSUPERSCRIPT ( 1 + divide start_ARG roman_cos start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_x ) end_ARG start_ARG 4 end_ARG ) [ square-root start_ARG italic_π end_ARG roman_erf ( italic_v ) - 2 italic_v roman_exp ( - italic_v start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) ] start_POSTSUBSCRIPT ( - italic_π - italic_ϕ ( italic_x ) ) / italic_δ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_π - italic_ϕ ( italic_x ) ) / italic_δ end_POSTSUPERSCRIPT italic_d italic_x
=9π32δ3+o(exp(δ2)).absent9𝜋32superscript𝛿3𝑜superscript𝛿2\displaystyle=\frac{9\sqrt{\pi}}{32\delta^{3}}+o(\exp(-\delta^{-2})).= divide start_ARG 9 square-root start_ARG italic_π end_ARG end_ARG start_ARG 32 italic_δ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG + italic_o ( roman_exp ( - italic_δ start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT ) ) .

We bound the size of this error using maxϕ(x)italic-ϕ𝑥\max\phi(x)roman_max italic_ϕ ( italic_x ), by erf((π1/2)/δ)=1o(exp((π1/2)/δ)2))\rm{erf}((\pi-1/2)/\delta)=1-o(\exp(-(\pi-1/2)/\delta)^{2}))roman_erf ( ( italic_π - 1 / 2 ) / italic_δ ) = 1 - roman_o ( roman_exp ( - ( italic_π - 1 / 2 ) / italic_δ ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) ) and similarly for the 2uexp(u2)2𝑢superscript𝑢22u\exp(-u^{2})2 italic_u roman_exp ( - italic_u start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) term. This then gives the initial palinstrophy P(t=0)=9π32δ3+o(exp(δ2))𝑃𝑡09𝜋32superscript𝛿3𝑜superscript𝛿2P(t=0)=\frac{9\sqrt{\pi}}{32\delta^{3}}+o(\exp(-\delta^{-2}))italic_P ( italic_t = 0 ) = divide start_ARG 9 square-root start_ARG italic_π end_ARG end_ARG start_ARG 32 italic_δ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG + italic_o ( roman_exp ( - italic_δ start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT ) ).

The numerical values for normalized initial enstrophy Z(t=0)δ=π20.886𝑍𝑡0𝛿𝜋20.886Z(t=0)\cdot\delta=\frac{\sqrt{\pi}}{2}\approx 0.886italic_Z ( italic_t = 0 ) ⋅ italic_δ = divide start_ARG square-root start_ARG italic_π end_ARG end_ARG start_ARG 2 end_ARG ≈ 0.886 and palinstrophy P(t=0)δ3=9π320.499𝑃𝑡0superscript𝛿39𝜋320.499P(t=0)\cdot\delta^{3}=\frac{9\sqrt{\pi}}{32}\approx 0.499italic_P ( italic_t = 0 ) ⋅ italic_δ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT = divide start_ARG 9 square-root start_ARG italic_π end_ARG end_ARG start_ARG 32 end_ARG ≈ 0.499 encountered in the simulations are consistent with the analytically derived values up to machine precision (1016superscript101610^{-16}10 start_POSTSUPERSCRIPT - 16 end_POSTSUPERSCRIPT). The same applies for the initial circulation, matching the value 2π2𝜋2\pi2 italic_π.

Refer to caption
Figure 2: Initial energy E(t=0)𝐸𝑡0E(t=0)italic_E ( italic_t = 0 ) as a function of the sheet thickness δ𝛿\deltaitalic_δ showing a linear decrease with a value of 1.52absent1.52\approx 1.52≈ 1.52 for δ0𝛿0\delta\rightarrow 0italic_δ → 0.

3 Characteristic mapping for 2D Euler

Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Figure 3: Illustration of the pullback operation by the characteristic map. Figure (a) shows the initial vorticity, figure (b) overlays onto figure (a) the backward map, and figure (c) shows the time t𝑡titalic_t vorticity obtained by pullback using the relabelling symmetry.

In this section, we briefly recall the characteristic mapping method (CMM) for solving the 2D incompressible Euler equations. For details and numerical implementation we refer the reader to Yin et al. (2021). The numerical results in this paper are computed using a GPU Cuda implementation of the method. The open source code is available on GitHub (Yadav et al., 2023).

The CMM for the 2D incompressible Euler equations is based on the 2D scalar vorticity formulation (equation 1) coupled with the CMM for linear transport. We recall that the characteristic map χBsubscript𝜒𝐵\chi_{B}italic_χ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT is the backward Lagrangian flow map and can be thought of as the back-to-label operator for the advection of Lagrangian particles. Indeed, for any particle trajectory γ(t)𝛾𝑡\gamma(t)italic_γ ( italic_t ) given by,

ddtγ(t)=𝒖(γ(t),t)𝑑𝑑𝑡𝛾𝑡𝒖𝛾𝑡𝑡\frac{d}{dt}\gamma(t)=\bm{u}(\gamma(t),t)divide start_ARG italic_d end_ARG start_ARG italic_d italic_t end_ARG italic_γ ( italic_t ) = bold_italic_u ( italic_γ ( italic_t ) , italic_t ) (12)

with initial condition γ(0)=γ0𝛾0subscript𝛾0\gamma(0)=\gamma_{0}italic_γ ( 0 ) = italic_γ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, we have that χBsubscript𝜒𝐵\chi_{B}italic_χ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT satisfies,

γ0=χB(γ(t),t).subscript𝛾0subscript𝜒𝐵𝛾𝑡𝑡\gamma_{0}=\chi_{B}(\gamma(t),t).italic_γ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = italic_χ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT ( italic_γ ( italic_t ) , italic_t ) . (13)

It follows by the method of characteristics, that any scalar quantity ϕitalic-ϕ\phiitalic_ϕ advected by the same velocity 𝒖𝒖\bm{u}bold_italic_u satisfies ϕ(𝒙,t)=ϕ0(χB(𝒙,t))italic-ϕ𝒙𝑡subscriptitalic-ϕ0subscript𝜒𝐵𝒙𝑡\phi(\bm{x},t)=\phi_{0}(\chi_{B}(\bm{x},t))italic_ϕ ( bold_italic_x , italic_t ) = italic_ϕ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_χ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT ( bold_italic_x , italic_t ) ), this is known as the relabelling symmetry.

In the vorticity equations, ω𝜔\omegaitalic_ω is a scalar-valued quantity advected by 𝒖𝒖\bm{u}bold_italic_u and thus has the above relabelling symmetry. The velocity field 𝒖𝒖\bm{u}bold_italic_u is in turn obtained from the Biot–Savart law. The coupling of the vorticity equations with the characteristic map yields the following governing equations for the numerical method:

tχB+(𝒖)χB=0subscript𝑡subscript𝜒𝐵𝒖subscript𝜒𝐵0\displaystyle\partial_{t}\chi_{B}+({\bm{u}}\cdot\nabla)\chi_{B}=0∂ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_χ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT + ( bold_italic_u ⋅ ∇ ) italic_χ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT = 0 (14a)
ω(𝒙,t)=ω0(χB(𝒙,t))𝜔𝒙𝑡subscript𝜔0subscript𝜒𝐵𝒙𝑡\displaystyle\omega({\bm{x}},t)=\omega_{0}(\chi_{B}({\bm{x}},t))italic_ω ( bold_italic_x , italic_t ) = italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_χ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT ( bold_italic_x , italic_t ) ) (14b)
𝒖=×Δ1ω𝒖superscriptΔ1𝜔\displaystyle{\bm{u}}=-\nabla\times\Delta^{-1}\omegabold_italic_u = - ∇ × roman_Δ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT italic_ω (14c)

Numerically, the advection equation for the map (14a) is discretized through the Gradient-Augmented Level-Set method (Nave et al., 2010; Seibold et al., 2012; Chidyagwai et al., 2011) which is a semi-Lagrangian method using Hermite cubic spatial interpolation with Runge–Kutta time-stepping schemes. The vorticity evaluation step (14b) is performed on a fixed fine grid by interpolation of the Hermite cubic representation of χBsubscript𝜒𝐵\chi_{B}italic_χ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT followed by a direct evaluation of the initial condition ω0subscript𝜔0\omega_{0}italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. Figure 3 illustrates the computation of the vorticity by pullback on the initial condition. Lastly, the velocity field is defined through the Biot–Savart law (14c) as the curl of the Hermite interpolant of the stream function, which we obtain from a spectral solver for the Poisson equation using Fast Fourier Transforms. The time dependence of the velocity field is obtained from a Lagrange extrapolation in time using stream function data from the three most recent time steps.

This approach has several desirable numerical properties for the problem investigated here. Firstly, the use of the advection solution operator χBsubscript𝜒𝐵\chi_{B}italic_χ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT maintains a back-to-label symmetry of the vorticity field. This ensures a non-dissipative numerical method, since there exists a coordinate transform which transports the numerical solution to the exact one. This property is beneficial to the study of singularity formation since artificial viscosity is eliminated as potential regularization mechanism preventing blow-up. Secondly, the characteristic maps benefit from the structure of the Lie-groups of volume-preserving diffeomorphisms. Therefore, a long-time map can be represented as the composition of multiple short-time sub-maps. For a time subdivision 0<t1<t2<<tm1<T0subscript𝑡1subscript𝑡2subscript𝑡𝑚1𝑇0<t_{1}<t_{2}<\ldots<t_{m-1}<T0 < italic_t start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT < italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT < … < italic_t start_POSTSUBSCRIPT italic_m - 1 end_POSTSUBSCRIPT < italic_T, the sub-map decomposition is an adaptive and multi-resolution representation of the full map χBsubscript𝜒𝐵\chi_{B}italic_χ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT given by

χB(,T)=χ[t1,0]χ[t2,t1]χ[T,tm1]subscript𝜒𝐵𝑇subscript𝜒subscript𝑡10subscript𝜒subscript𝑡2subscript𝑡1subscript𝜒𝑇subscript𝑡𝑚1\chi_{B}(\cdot,T)=\chi_{[t_{1},0]}\circ\chi_{[t_{2},t_{1}]}\circ\dots\circ\chi% _{[T,t_{m-1}]}italic_χ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT ( ⋅ , italic_T ) = italic_χ start_POSTSUBSCRIPT [ italic_t start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , 0 ] end_POSTSUBSCRIPT ∘ italic_χ start_POSTSUBSCRIPT [ italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_t start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ] end_POSTSUBSCRIPT ∘ ⋯ ∘ italic_χ start_POSTSUBSCRIPT [ italic_T , italic_t start_POSTSUBSCRIPT italic_m - 1 end_POSTSUBSCRIPT ] end_POSTSUBSCRIPT (15)

where χ[ti,ti1]subscript𝜒subscript𝑡𝑖subscript𝑡𝑖1\chi_{[t_{i},t_{i-1}]}italic_χ start_POSTSUBSCRIPT [ italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_t start_POSTSUBSCRIPT italic_i - 1 end_POSTSUBSCRIPT ] end_POSTSUBSCRIPT represents the backwards map in the time interval [ti1,ti]subscript𝑡𝑖1subscript𝑡𝑖[t_{i-1},t_{i}][ italic_t start_POSTSUBSCRIPT italic_i - 1 end_POSTSUBSCRIPT , italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ]. On each time sub-interval, the spatial deformation is comparatively small and can be well resolved on a coarser grid, the remapping step is triggered dynamically as the numerical resolution of the sub-map coarse grids are depleted. The full map obtained from the map composition in (15) retains all scales formed by each sub-map.

The movement of the vortex sheet center line, parametrized by 𝒙(θ,t)𝒙𝜃𝑡\bm{x}(\theta,t)bold_italic_x ( italic_θ , italic_t ), is tracked by following the trajectories of Lagrangian particles in the vortex core. The initial position corresponds to the perturbation function ϕ(x)italic-ϕ𝑥\phi(x)italic_ϕ ( italic_x ), i.e. 𝒙(θ,t=0)=(θ,ϕ(θ)+π),θ[0,2π]formulae-sequence𝒙𝜃𝑡0𝜃italic-ϕ𝜃𝜋𝜃02𝜋\bm{x}(\theta,t=0)=(\theta,\phi(\theta)+\pi),\quad\theta\in[0,2\pi]bold_italic_x ( italic_θ , italic_t = 0 ) = ( italic_θ , italic_ϕ ( italic_θ ) + italic_π ) , italic_θ ∈ [ 0 , 2 italic_π ]. The initial vortex center curve is discretized using NPsubscript𝑁𝑃N_{P}italic_N start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT uniformly distributed sample particles with θn=n 2π/Npsubscript𝜃𝑛𝑛2𝜋subscript𝑁𝑝\theta_{n}=n\,2\pi/N_{p}italic_θ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = italic_n 2 italic_π / italic_N start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT. A sufficiently high number of points NP106subscript𝑁𝑃superscript106N_{P}\geq 10^{6}italic_N start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT ≥ 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT ensures that all dynamics of the material line is captured even under strong elongation effects.
Their time evolution under the numerical velocity field computed from the CMM is then tracked individually using standard explicit Runge–Kutta methods. We must note here that the velocity field obtained from the CMM is only C0superscript𝐶0C^{0}italic_C start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT in space since it is defined as the curl of the Hermite cubic interpolant of the stream function. The size of the discontinuities in higher derivatives depends on the grid size Nψsubscript𝑁𝜓N_{\psi}italic_N start_POSTSUBSCRIPT italic_ψ end_POSTSUBSCRIPT used for the stream function with the jumps in ithsuperscript𝑖𝑡i^{th}italic_i start_POSTSUPERSCRIPT italic_t italic_h end_POSTSUPERSCRIPT derivative scaling like D4ψNψ4+isubscriptnormsuperscript𝐷4𝜓superscriptsubscript𝑁𝜓4𝑖\|D^{4}\psi\|_{\infty}N_{\psi}^{-4+i}∥ italic_D start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT italic_ψ ∥ start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT italic_ψ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 4 + italic_i end_POSTSUPERSCRIPT. This introduces a negligible error in the Lsuperscript𝐿L^{\infty}italic_L start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT error for the curve position and derivatives but can be a source of noise for the regularity analysis.

4 Performed computations

Several runs with successively decreasing δ𝛿\deltaitalic_δ-values were executed on state-of-the art graphics cards, maximizing the available usable memory. All used parameters are summarized in table 1. Two grid sizes were used: A coarser one for the description of the flow map χ𝜒\chiitalic_χ and velocity 𝒖𝒖\bm{u}bold_italic_u and one for the initial vorticity ω0subscript𝜔0\omega_{0}italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT defined for each sub-map as well as the discrete vorticity used in the Biot–Savart law. The merit of this lies in the enhanced smoothness of the velocity field, obviating the need for a highly detailed representation. Additionally, fine scales of the flow captured by the flow map are retained from composition of the individual sub-maps. The time-step ΔtΔ𝑡\Delta troman_Δ italic_t is set after a Courant–Friedrich–Lewis (CFL) number of 1/3131/31 / 3 of the coarse map to neglect its influence, however the semi-Lagrangian method can easily deal with large CFL numbers >1absent1>1> 1 as well. The incompressibility threshold of δinc,bsubscript𝛿𝑖𝑛𝑐𝑏\delta_{inc,b}italic_δ start_POSTSUBSCRIPT italic_i italic_n italic_c , italic_b end_POSTSUBSCRIPT ensures the volume preservation of the sub-maps, i.e. we monitor that |detχB|δinc,Bsubscript𝜒𝐵subscript𝛿𝑖𝑛𝑐𝐵|\det\nabla\chi_{B}|\leq\delta_{inc,B}| roman_det ∇ italic_χ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT | ≤ italic_δ start_POSTSUBSCRIPT italic_i italic_n italic_c , italic_B end_POSTSUBSCRIPT. The parameter of the local stencil size ϵmsubscriptitalic-ϵ𝑚\epsilon_{m}italic_ϵ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT defines the distance at which spatial derivatives for the creation of the Hermite interpolants with the GALS method are computed, together with its corresponding map update stencil order. The filter size kLPsubscript𝑘𝐿𝑃k_{LP}italic_k start_POSTSUBSCRIPT italic_L italic_P end_POSTSUBSCRIPT defines the cut-off frequency for the low-pass filter and was disabled by setting it to high frequencies of negligible energy. The fluid and embedded particle time schemes are chosen to adapted versions of the Runge–Kutta scheme with improved efficiency for the CM method. At last, NPsubscript𝑁𝑃N_{P}italic_N start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT is the number of equidistant particles embedded at the vortex center line. Further in-depth analysis and explanation of all used parameters are reported in Yin et al. (2021); Bergmann (2022).

Refer to caption
(a) t=5𝑡5t=5italic_t = 5 and δ0.032𝛿0.032\delta\approx 0.032italic_δ ≈ 0.032
Refer to caption
Refer to caption
(b) Close-up with overlayed center line
Refer to caption
Refer to caption
(c) Close-up of local palinstrophy
Figure 4: Developed vortices in global field and close-up of upper vortex with overlayed center-line (a,b). After an initial phase two symmetric vortices start to form with center with condensed region of high vorticity and spiral arms. Figure (c) shows corresponding local palinstrophy |ω(𝒙,t)|2superscript𝜔𝒙𝑡2|\nabla\omega(\bm{x},t)|^{2}| ∇ italic_ω ( bold_italic_x , italic_t ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT.
Refer to caption
(a) Results obtained with CMM
Refer to caption
(b) Results from Caflisch et al. (2022)
Figure 5: Comparison between our results and Caflisch et al. (2022) with δ0.014𝛿0.014\delta\approx 0.014italic_δ ≈ 0.014 and at t=2.85𝑡2.85t=2.85italic_t = 2.85. The CMM results were adapted to the shifted region ΩssubscriptΩ𝑠\Omega_{s}roman_Ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT for comparison purposes. They show excellent agreement. Figure adapted from Caflisch et al. (2022), printed with their permission.

The parameters were specifically chosen to maximize the coarse grid size for the given GPU memory, which in return leads to lower growth of incompressibility error and therefore more accurate flow representation over time. With too low grid resolution, a simulation with small thickness δ𝛿\deltaitalic_δ leads to premature emergence of Kelvin–Helmholtz (KH) instabilities along the line of high vorticity. This was observed to be suppressed with increased grid size and therefore attributed to numerically induced errors. For the highest available grid settings on the NVIDIA A100 cards, a value of δ0.007𝛿0.007\delta\approx 0.007italic_δ ≈ 0.007 was the lowest observed value where perturbations are bounded and do not dominate the flow behavior. In simulations with smaller vortex sheet thickness the flow is completely dominated by the emerging KH-vortices before any roll-up process starts to occur. In total, simulations for six different δ𝛿\deltaitalic_δ-values have been performed successfully ranging from δ0.007=1/2104𝛿0.00712superscript104\delta\approx 0.007=1/\sqrt{2\cdot 10^{4}}italic_δ ≈ 0.007 = 1 / square-root start_ARG 2 ⋅ 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT end_ARG to δ0.045=1/5102𝛿0.04515superscript102\delta\approx 0.045=1/\sqrt{5\cdot 10^{2}}italic_δ ≈ 0.045 = 1 / square-root start_ARG 5 ⋅ 10 start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG. The values were placed with 1/δ21superscript𝛿21/\delta^{2}1 / italic_δ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT in order to ensure comparability to the Navier–Stokes results from Caflisch et al. (2022) by their overall dynamics, as they did for their inviscid results.

Parameter V100 A100 Parameter V100 A100
Ncoarsesubscript𝑁coarseN_{\text{coarse}}italic_N start_POSTSUBSCRIPT coarse end_POSTSUBSCRIPT & Nψsubscript𝑁𝜓N_{\psi}italic_N start_POSTSUBSCRIPT italic_ψ end_POSTSUBSCRIPT 8192819281928192 12288122881228812288 Nfinesubscript𝑁fineN_{\text{fine}}italic_N start_POSTSUBSCRIPT fine end_POSTSUBSCRIPT & Nωsubscript𝑁𝜔N_{\omega}italic_N start_POSTSUBSCRIPT italic_ω end_POSTSUBSCRIPT 12288122881228812288 24576245762457624576
ΔtfluidΔsubscript𝑡fluid\Delta t_{\text{fluid}}roman_Δ italic_t start_POSTSUBSCRIPT fluid end_POSTSUBSCRIPT 1/245761245761/245761 / 24576 1/368641368641/368641 / 36864 δinc,bsubscript𝛿𝑖𝑛𝑐𝑏\delta_{inc,b}italic_δ start_POSTSUBSCRIPT italic_i italic_n italic_c , italic_b end_POSTSUBSCRIPT 103superscript10310^{-3}10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT 103superscript10310^{-3}10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT
ϵmsubscriptitalic-ϵ𝑚\epsilon_{m}italic_ϵ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT 104superscript10410^{-4}10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT 51055superscript1055\cdot 10^{-5}5 ⋅ 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT kLPsubscript𝑘𝐿𝑃k_{LP}italic_k start_POSTSUBSCRIPT italic_L italic_P end_POSTSUBSCRIPT 4096409640964096 12288122881228812288
Fluid time scheme RK4Mod RK3Mod Map update stencil 4th order 4th order
NPsubscript𝑁𝑃N_{P}italic_N start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT 106superscript10610^{6}10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT 107superscript10710^{7}10 start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPT Particle time scheme RK4Mod RK3Mod
Table 1: Simulation parameters for the different runs on Nvidia V100 and A100 machines. Given are the resolution of coarse and fine grids Ncoarsesubscript𝑁𝑐𝑜𝑎𝑟𝑠𝑒N_{coarse}italic_N start_POSTSUBSCRIPT italic_c italic_o italic_a italic_r italic_s italic_e end_POSTSUBSCRIPT and Nfinesubscript𝑁𝑓𝑖𝑛𝑒N_{fine}italic_N start_POSTSUBSCRIPT italic_f italic_i italic_n italic_e end_POSTSUBSCRIPT, time-step ΔtfluidΔsubscript𝑡fluid\Delta t_{\text{fluid}}roman_Δ italic_t start_POSTSUBSCRIPT fluid end_POSTSUBSCRIPT, incompressibility threshold δinc,bsubscript𝛿𝑖𝑛𝑐𝑏\delta_{inc,b}italic_δ start_POSTSUBSCRIPT italic_i italic_n italic_c , italic_b end_POSTSUBSCRIPT, local stencil size ϵmsubscriptitalic-ϵ𝑚\epsilon_{m}italic_ϵ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT for the GALS method with its stencil order, low-pass filter size kLPsubscript𝑘𝐿𝑃k_{LP}italic_k start_POSTSUBSCRIPT italic_L italic_P end_POSTSUBSCRIPT, time scheme order for fluid and particle advection and number of particles.

All runs were performed until the map-stack of emerging sub-maps depleted the CPU RAM of the available compute nodes. The computations were computed on the IDRIS supercomputer. Up to 43 convoluted maps for A100 runs and 117 convoluted maps for V100 runs were captured. Details on the executed run architecture and final observed time together with the amount of captured windings of the emerging vortex structures are given in table 2. The settings ensure that energy and enstrophy are conserved to an error of 2105absent2superscript105\approx 2\cdot 10^{-5}≈ 2 ⋅ 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT, see figure 6, which is supported by the error of the CM-method being not dissipative in nature for transported quantities.

δ𝛿absent\delta\approxitalic_δ ≈ 0.0045 0.007 0.01 0.0141 0.022 0.032 0.045
1/δ21superscript𝛿21/\delta^{2}1 / italic_δ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT 5×1045superscript1045\times 10^{4}5 × 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT 2×1042superscript1042\times 10^{4}2 × 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT 1×1041superscript1041\times 10^{4}1 × 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT 5×1035superscript1035\times 10^{3}5 × 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT 2×1032superscript1032\times 10^{3}2 × 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT 1×1031superscript1031\times 10^{3}1 × 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT 5×1025superscript1025\times 10^{2}5 × 10 start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT
Run on A100 A100 V100 V100 V100 V100 V100
Final time 1.24 2.29 2.86 3.46 4.37 5.82 7.59
Captured windings 0 3 5 6 6 7 7
Table 2: Overview on the used vorticity layer thickness values δ𝛿\deltaitalic_δ including the obtained final time and the number of windings, for which computations were performed.

Simulations with successively decreasing thickness values δ𝛿\deltaitalic_δ reveal the flow behavior for the limit δ0𝛿0\delta\rightarrow 0italic_δ → 0. An example for the flow dynamics at t=5𝑡5t=5italic_t = 5 and with δ0.032𝛿0.032\delta\approx 0.032italic_δ ≈ 0.032 is depicted in figure 4. Interestingly, two small vortices start to roll up along a common center each. These have a compressed, round-shaped vortex core of high vorticity and two elongated spiral arms. Eventually, both vortices start to attract each other and will eventually merge into a single stable vortex for long times. With decreasing sheet thickness the roll-up process is triggered earlier and the emerging structures are smaller in scale. Figure 3(b) shows that the vortex core is round with distortion towards the second vortex due to the mutual attraction. The local palinstrophy captured in figure 3(c), is very pronounced in the spiral arms and shows the spiraling structures also present within the vortex core itself.
The acquired results are comparable with inviscid flow fields reported by Caflisch et al. (2017) and Caflisch et al. (2022). The depicted vortex in figure 4(a) is in excellent agreement with figure 14d of Caflisch et al. (2022), given in figure 4(b). It is important to mention that the original results are depicted in non-equal aspect ratio, due to which the vortex core appears more elliptic.

Refer to caption
(a) Relative energy error E(t)E(t=0)1𝐸𝑡𝐸𝑡01\frac{E(t)}{E(t=0)}-1divide start_ARG italic_E ( italic_t ) end_ARG start_ARG italic_E ( italic_t = 0 ) end_ARG - 1
Refer to caption
(b) Relative enstrophy error Z(t)Z(t=0)1𝑍𝑡𝑍𝑡01\frac{Z(t)}{Z(t=0)}-1divide start_ARG italic_Z ( italic_t ) end_ARG start_ARG italic_Z ( italic_t = 0 ) end_ARG - 1
Figure 6: Relative energy and enstrophy error to the initial condition t=0𝑡0t=0italic_t = 0 for simulations with different δ𝛿\deltaitalic_δ-values over time. Both quantities are well conserved up to 21052superscript1052\cdot 10^{-5}2 ⋅ 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT.
Refer to caption
(a) Scaling of palinstrophy growth at early times shown in log-log.
Refer to caption
(b) Palinstrophy growth at later time shown in log-lin.
Figure 7: Initial palinstrophy growth shows a t2superscript𝑡2t^{2}italic_t start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT-scaling (a). At later times an exponential palinstrophy growth is found which is due to the formation of rolled-up vortices (b). The steepness of the slope increases for smaller δ𝛿\deltaitalic_δ-values and exponential growth starts at earlier times.

5 Analysis of the results

In the following, the numerical results will be analyzed and discussed. Comparison of global quantities as well as the material line uncover the different flow dynamics. Normalization in space and time with respective scaling laws give insight into dynamics for the non-smooth initial condition when δ0𝛿0\delta\rightarrow 0italic_δ → 0.
The palinstrophy P𝑃Pitalic_P over time, shown in figure 7, exhibits two different growth stages. At early times, all performed simulations show an algebraic growth following a t2superscript𝑡2t^{2}italic_t start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT-scaling. This initial growth period is later on overtaken by a strong exponential growth once vortex structures emerge. The steepness of the growth increases with decreasing vortex sheet thickness δ𝛿\deltaitalic_δ. For the slope s𝑠sitalic_s of the exponential growth with P(t)exp(st)proportional-to𝑃𝑡𝑠𝑡P(t)\propto\exp(st)italic_P ( italic_t ) ∝ roman_exp ( italic_s italic_t ), a relation of sδ0.77𝑠superscript𝛿0.77s\approx\delta^{-0.77}italic_s ≈ italic_δ start_POSTSUPERSCRIPT - 0.77 end_POSTSUPERSCRIPT was found. This means that in the limit of δ0𝛿0\delta\rightarrow 0italic_δ → 0 palinstrophy diveres. This is consistent with results in the literature for Navier–Stokes where exponential growth of palinstrophy was derived, see e.g. Lesieur (2008). For small times a powerlaw behavior was predicted in Ayala & Protas (2014) and estimates of the maximum palinstrophy growth were given. The lower limit of the investigated δ𝛿\deltaitalic_δ-values, i.e. for δ0.0045𝛿0.0045\delta\approx 0.0045italic_δ ≈ 0.0045, experiences artifacts in form of numerically accelerated secondary Kelvin–Helmholtz instabilities along the material line, resulting in the premature and irregular steep palinstrophy growth.

With the material line 𝒙(θ,t)𝒙𝜃𝑡\bm{x}(\theta,t)bold_italic_x ( italic_θ , italic_t ), given by the individual particle positions, we define following (Caflisch et al., 2022) several quantities, the arc length s(θ)𝑠𝜃s(\theta)italic_s ( italic_θ ), curvature κc(θ)subscript𝜅𝑐𝜃\kappa_{c}(\theta)italic_κ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_θ ) and true vortex strength γc(θ)subscript𝛾𝑐𝜃\gamma_{c}(\theta)italic_γ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_θ ). These are used to analyze the material line dynamics and are defined respectively as,

s(θ)𝑠𝜃\displaystyle s(\theta)italic_s ( italic_θ ) =02π|𝒙θ|𝑑θ,absentsuperscriptsubscript02𝜋subscript𝒙𝜃differential-d𝜃\displaystyle=\int_{0}^{2\pi}|\bm{x}_{\theta}|d\theta\,,= ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 italic_π end_POSTSUPERSCRIPT | bold_italic_x start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT | italic_d italic_θ , (16)
κc(θ)subscript𝜅𝑐𝜃\displaystyle\kappa_{c}(\theta)italic_κ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_θ ) =(xθyθθyθxθθ)/(xθ2+yθ2)3/2,absentsubscript𝑥𝜃subscript𝑦𝜃𝜃subscript𝑦𝜃subscript𝑥𝜃𝜃superscriptsuperscriptsubscript𝑥𝜃2superscriptsubscript𝑦𝜃232\displaystyle=(x_{\theta}y_{\theta\theta}-y_{\theta}x_{\theta\theta})/(x_{% \theta}^{2}+y_{\theta}^{2})^{3/2}\,,= ( italic_x start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT italic_y start_POSTSUBSCRIPT italic_θ italic_θ end_POSTSUBSCRIPT - italic_y start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT italic_θ italic_θ end_POSTSUBSCRIPT ) / ( italic_x start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_y start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 3 / 2 end_POSTSUPERSCRIPT , (17)
γc(θ)subscript𝛾𝑐𝜃\displaystyle\gamma_{c}(\theta)italic_γ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_θ ) =|𝒙θ|1.absentsuperscriptsubscript𝒙𝜃1\displaystyle=|\bm{x}_{\theta}|^{-1}\,.= | bold_italic_x start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT . (18)

with the notation 𝒙θ=𝒙θsubscript𝒙𝜃𝒙𝜃\bm{x}_{\theta}=\frac{\partial\bm{x}}{\partial\theta}bold_italic_x start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT = divide start_ARG ∂ bold_italic_x end_ARG start_ARG ∂ italic_θ end_ARG and correspondingly for the other derivatives. Numerically the derivatives were computed using central 4th order finite differences over the particle positions for both the first and second derivative. Both the elongation of the spiral arms and contraction around the vortex core from figure 4 can be found in the evolution of the arc length of the material line in figure 7(a). The overall arc length is both increased by the emerging two vortices as well as the global vortex merging process with the effect intensifying strongly over time especially around the vortex core position. The almost vertical lines correspond to strong stretching of the material line, while the almost horizontal part correspond to extreme compression of the particles. Our results for the Euler case are much more pronounced but comparable with those shown in Caflisch et al. (2022) figure 12b for Navier–Stokes.

Refer to caption
(a) s(θ)𝑠𝜃s(\theta)italic_s ( italic_θ ) for δ0.032𝛿0.032\delta\approx 0.032italic_δ ≈ 0.032
Refer to caption
(b) κcsubscript𝜅𝑐\kappa_{c}italic_κ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT for δ0.014𝛿0.014\delta\approx 0.014italic_δ ≈ 0.014
Refer to caption
(c) γcsubscript𝛾𝑐\gamma_{c}italic_γ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT for δ0.014𝛿0.014\delta\approx 0.014italic_δ ≈ 0.014
Figure 8: Arc length of the material curve for δ=0.032𝛿0.032\delta=0.032italic_δ = 0.032 with vortex center as vertical dotted line (a). Curvature κcsubscript𝜅𝑐\kappa_{c}italic_κ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT (b) and true vortex strength γcsubscript𝛾𝑐\gamma_{c}italic_γ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT (b) over arc length s(θ)𝑠𝜃s(\theta)italic_s ( italic_θ ) around vortex center for δ=0.014𝛿0.014\delta=0.014italic_δ = 0.014.

The curvature shown in figure 7(b) showcases the strength of the winding process, forming two peaks of opposite sign around the vortex cores. These are situated at the edge of the coalesced vortex centers from which at longer times the spiral arms start. Involving second order derivatives the curvature experiences high-frequency oscillations from numerical derivation of the particle positions. A Gaussian filter with standard deviation of 3104×Np3superscript104subscript𝑁𝑝3\cdot 10^{-4}\times N_{p}3 ⋅ 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT × italic_N start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT was used to mitigate this effect. In the center of the developed vortex the true vortex strength reaches a peak showcasing strong convergence towards the vortex core. Before any spiral arms are formed all particles are compressed towards the vortex core (with γc>1subscript𝛾𝑐1\gamma_{c}>1italic_γ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT > 1). Once the vortices start to rotate, two spiral arms will form and elongation occurs (γc<1subscript𝛾𝑐1\gamma_{c}<1italic_γ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT < 1) outside the vortex core. Both quantities reach maximum peak values at different times, depicted by the green curve in figure  7(c), and then start to disperse.

Refer to caption
(a) Curvature κc(θ)subscript𝜅𝑐𝜃\kappa_{c}(\theta)italic_κ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_θ )
Refer to caption
(b) True vortex strength γc(θ)subscript𝛾𝑐𝜃\gamma_{c}(\theta)italic_γ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_θ )
Figure 9: Temporal re-scaling for the maximum values of the curvature κssubscript𝜅𝑠\kappa_{s}italic_κ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT and true vortex strength γssubscript𝛾𝑠\gamma_{s}italic_γ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT as well as temporal re-scaling for the vortex turnover with empirically determined coefficients. The critical times tssubscript𝑡𝑠t_{s}italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, given in figure (b), were found to be decreasing with lower δ𝛿\deltaitalic_δ-values.
Refer to caption
(a) Vortex turnover
Refer to caption
(b) Rescaled palinstrophy growth in log-log.
Figure 10: Temporal re-scaling for the vortex turnover (a) and palinstrophy growth (b) with empirically determined coefficients. The critical times tssubscript𝑡𝑠t_{s}italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT are given in figure (a).

Tracking the maximum value of curvature and vortex strength over time enables comparisons between different δ𝛿\deltaitalic_δ-values, shown in figure 9. Here, the curves have been matched empirically for the given scaling laws. The curvature was found to empirically scale with δ0.9superscript𝛿0.9\delta^{-0.9}italic_δ start_POSTSUPERSCRIPT - 0.9 end_POSTSUPERSCRIPT and the true vortex strength with δ0.31superscript𝛿0.31\delta^{-0.31}italic_δ start_POSTSUPERSCRIPT - 0.31 end_POSTSUPERSCRIPT, for which an explanation was not found. Nonetheless, this scaling supports the conjecture of Caflisch et al. (2022) that only for the limit of δ0𝛿0\delta\rightarrow 0italic_δ → 0 the curvature as well as true vortex strength go towards infinity for finite time. For this limit the vortex core coalesces to an individual point with infinite curvature.
Interestingly, over time both quantities show the same behavior with an initial steep increase to an overall maximum, decreasing again with oscillations. As the observed dispersion effect appears similar for different vortex sheet thicknesses δ𝛿\deltaitalic_δ it is interpreted as a flow behavior rather than a result of numerical artifacts. It can be explained by the flow evolution. At first the vorticity sheet close to the vortex core starts to roll up. Two regions are formed based on the true vortex strength: One close to the center of the vortex were the material line with surrounding region of vorticity is condensed and one outside where the material line is strongly elongated from the formation of the spiral arms due to the rotation of the vortex core. At some point the vortex core is reformed into a circular shape and rotates with quasi-constant speed due to the resulting velocity from the Biot–Savart law.

Refer to caption
(a) Re-scaled by δ0.9superscript𝛿0.9\delta^{-0.9}italic_δ start_POSTSUPERSCRIPT - 0.9 end_POSTSUPERSCRIPT
Refer to caption
(b) Re-scaled by δ1superscript𝛿1\delta^{-1}italic_δ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, as done in (Caflisch et al., 2022)
Figure 11: Spatial re-scaling with different factors for the fourth time that 𝒙(θ,t)/θ0𝒙𝜃𝑡𝜃0\partial\bm{x}(\theta,t)/\partial\theta\approx 0∂ bold_italic_x ( italic_θ , italic_t ) / ∂ italic_θ ≈ 0 in the y𝑦yitalic_y direction for different δ𝛿\deltaitalic_δ-values.

This is also visible in figure 9(a). Here the occurrence of the vanishing gradient is plotted over re-scaled time, where each occurrence marks when the gradient of the material line at the spiral center becomes zero, i.e. when either the x𝑥xitalic_x- or y𝑦yitalic_y-component of the derivative of the material line position 𝒙(θ,t)θ𝒙𝜃𝑡𝜃\frac{\partial{\bm{x}}(\theta,t)}{\partial\theta}divide start_ARG ∂ bold_italic_x ( italic_θ , italic_t ) end_ARG start_ARG ∂ italic_θ end_ARG vanishes. As the vortex turnover over re-scaled time occurs in constant periods (linear growth in figure 9(a)), the vortex rotates with constant speed. It is similar to a solid body rotation and portrays a stable vortex core.
Additionally, the palinstrophy growth was also found to be re-scalable in time using tssubscript𝑡𝑠t_{s}italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT and delta𝑑𝑒𝑙𝑡𝑎deltaitalic_d italic_e italic_l italic_t italic_a, as shown in figure 9(b). In all graphs of figures 9 and 10, temporal scaling has been used. As for different δ𝛿\deltaitalic_δ-values the results are similar, those four criteria were used to match them and determine an empirical relation. The ansatz was taken from Caflisch et al. (2022) using (tts)δ1𝑡subscript𝑡𝑠superscript𝛿1(t-t_{s})\delta^{-1}( italic_t - italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) italic_δ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, where tssubscript𝑡𝑠t_{s}italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT is the singularity time and δ1superscript𝛿1\delta^{-1}italic_δ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT as re-scaling factor. The flow scales linearly with decreasing δ𝛿\deltaitalic_δ-values. The critical times are not equal to ts=1.505subscript𝑡𝑠1.505t_{s}=1.505italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = 1.505, the value of the Birkhoff–Rott equation but were observed to be slightly larger. They are closely constant for larger δ𝛿\deltaitalic_δ-values but changed once it was decreased, possibly going towards the critical time of the BR-equation for vanishing vortex sheet thickness δ𝛿\deltaitalic_δ. The linear scaling with δ1superscript𝛿1\delta^{-1}italic_δ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT in comparison to the scaling of δ2/3superscript𝛿23\delta^{-2/3}italic_δ start_POSTSUPERSCRIPT - 2 / 3 end_POSTSUPERSCRIPT for viscous flow reported by Caflisch et al. (2022) results from different flow dynamics. With no viscosity the whole dynamics of the condensed vortex core is an accumulation of vorticity present in the initial vortex sheets, which scales directly with the thickness δ𝛿\deltaitalic_δ.

Refer to caption
(a) δ=0.045𝛿0.045\delta=0.045italic_δ = 0.045
Refer to caption
(b) δ=0.01𝛿0.01\delta=0.01italic_δ = 0.01
Figure 12: Energy spectra E(k)𝐸𝑘E(k)italic_E ( italic_k ) defined in equation 4 for different δ𝛿\deltaitalic_δ-values. Both start with an initial scaling of k2superscript𝑘2k^{-2}italic_k start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT with exponential decay for large k𝑘kitalic_k as given in figure 0(b).

The material curves for different δ𝛿\deltaitalic_δ-values have strong spatial self-similarity. However, while Caflisch et al. (2022) reported a scaling of δ1superscript𝛿1\delta^{-1}italic_δ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, this did not match our computed results. These show an empirical scaling of δ0.9superscript𝛿0.9\delta^{-0.9}italic_δ start_POSTSUPERSCRIPT - 0.9 end_POSTSUPERSCRIPT, which is slightly different, as shown in figure 11. The similarity scaling is valid for all reported simulations and all times around the vortex cores and only breaks down with the vortex merging process for later times (observed for δ0.045𝛿0.045\delta\approx 0.045italic_δ ≈ 0.045 for t>7𝑡7t>7italic_t > 7).
The characteristic of the initial energy spectra, defined in equation 4, have already been explained in section 2. For larger δ𝛿\deltaitalic_δ-values the initial profile with exponential decay experiences a shift of energy to finer scales (figure 11(a)). In comparison, the results with lower δ𝛿\deltaitalic_δ-values with initially more pronounced k2superscript𝑘2k^{-2}italic_k start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT Dirac-scaling show only little shift in energy from larger to finer scales (figure 11(b)). However, for both simulations the two vortices have already emerged for the captured time and performed several windings, while the results for figure 11(b) did not yet observe any developed distortion from the vortex merging process. Eventually, once this global process continues the energy spectra are expected to behave similarly as to that for larger δ𝛿\deltaitalic_δ-values. The formation of the local vortices exhibits therefore only little impact on the energy spectra.
The two vortices which form are quite unstable once perturbed to sufficient degree. With ongoing global merging process with increasing influence on the flow field, the shape of the vortex core becomes more and more elliptic (figure 13). This brings an imbalance to the velocity field and secondary vortices form, breaking down the regularity. The individual parts of the spiral arm become compressed into filament structures and secondary vortices form.

Refer to caption
(a) t=5𝑡5t=5italic_t = 5
Refer to caption
(b) t=6.2𝑡6.2t=6.2italic_t = 6.2
Refer to caption
Refer to caption
(c) t=7.4𝑡7.4t=7.4italic_t = 7.4
Figure 13: Material line (in black) and vorticity for δ=0.032𝛿0.032\delta=0.032italic_δ = 0.032 at different times. The initially more stable spiral arms experience distortions and filament structures start to emerge with ongoing global vortex merger process.

6 Singularity analysis

The complex singularity of an analytic function can be analyzed using the Fourier transform. The width of the analyticity strip can be obtained by considering the asymptotic behavior of the Fourier spectrum governed by Laplace’s formula. To obtain information about singularities outside the analyticity strip the Borel–Polya–van der Hoeven (BPH) method has been proposed (Pauls & Frisch, 2007).

6.1 Borel–Polya–van der Hoeven method

The Borel–Polya–van der Hoeven (BPH) method (Pauls & Frisch, 2007) is a numerical tool for finding complex singularities of single variable functions by combining the information on singularities obtained from Borel transforms for Taylor series (Pólya, 1933) with numerical techniques for asymptotic interpolation (van der Hoeven, 2006). This method has been successfully applied to find singularities for the 1D Burgers equations (Pauls & Frisch, 2007) and has been used to analyze potential singularity formation in the 2D Euler vortex sheet problem (Caflisch et al., 2022). The following is a summary of the method along with some small modifications implemented to adapt to our case.

Given a complex function f(Z)𝑓𝑍f(Z)italic_f ( italic_Z ) with formal power series

f(Z)=n=0anZn,𝑓𝑍superscriptsubscript𝑛0subscript𝑎𝑛superscript𝑍𝑛f(Z)=\sum_{n=0}^{\infty}a_{n}Z^{n},italic_f ( italic_Z ) = ∑ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_Z start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT , (19)

its Borel transform and Borel–Laplace transform are given respectively by

fB(ξ)=n=0ann!ξn,fBL(Z)=1Zf(1Z)=n=0anZn+1,formulae-sequencesuperscript𝑓𝐵𝜉superscriptsubscript𝑛0subscript𝑎𝑛𝑛superscript𝜉𝑛superscript𝑓𝐵𝐿𝑍1𝑍𝑓1𝑍superscriptsubscript𝑛0subscript𝑎𝑛superscript𝑍𝑛1f^{B}(\xi)=\sum_{n=0}^{\infty}\frac{a_{n}}{n!}\xi^{n},\quad\quad f^{BL}(Z)=% \frac{1}{Z}f\left(\frac{1}{Z}\right)=\sum_{n=0}^{\infty}\frac{a_{n}}{Z^{n+1}},italic_f start_POSTSUPERSCRIPT italic_B end_POSTSUPERSCRIPT ( italic_ξ ) = ∑ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT divide start_ARG italic_a start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT end_ARG start_ARG italic_n ! end_ARG italic_ξ start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT , italic_f start_POSTSUPERSCRIPT italic_B italic_L end_POSTSUPERSCRIPT ( italic_Z ) = divide start_ARG 1 end_ARG start_ARG italic_Z end_ARG italic_f ( divide start_ARG 1 end_ARG start_ARG italic_Z end_ARG ) = ∑ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT divide start_ARG italic_a start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT end_ARG start_ARG italic_Z start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT end_ARG , (20)

so named since fBLsuperscript𝑓𝐵𝐿f^{BL}italic_f start_POSTSUPERSCRIPT italic_B italic_L end_POSTSUPERSCRIPT is formally the Laplace transform of fBsuperscript𝑓𝐵f^{B}italic_f start_POSTSUPERSCRIPT italic_B end_POSTSUPERSCRIPT.

The BPH method is built on Pólya’s theorem which is based on the observation that for an=cnsubscript𝑎𝑛superscript𝑐𝑛a_{n}=c^{n}italic_a start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = italic_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT for some complex number c𝑐citalic_c, the Borel transform is the exponential

fB(ξ)=n=0(cξ)nn!=ecξ,superscript𝑓𝐵𝜉superscriptsubscript𝑛0superscript𝑐𝜉𝑛𝑛superscript𝑒𝑐𝜉f^{B}(\xi)=\sum_{n=0}^{\infty}\frac{(c\xi)^{n}}{n!}=e^{c\xi},italic_f start_POSTSUPERSCRIPT italic_B end_POSTSUPERSCRIPT ( italic_ξ ) = ∑ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT divide start_ARG ( italic_c italic_ξ ) start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_ARG start_ARG italic_n ! end_ARG = italic_e start_POSTSUPERSCRIPT italic_c italic_ξ end_POSTSUPERSCRIPT , (21)

while the Borel–Laplace transform is a simple pole at c𝑐citalic_c:

fBL(Z)=n=0cnZn+1=1Zc.superscript𝑓𝐵𝐿𝑍superscriptsubscript𝑛0superscript𝑐𝑛superscript𝑍𝑛11𝑍𝑐f^{BL}(Z)=\sum_{n=0}^{\infty}\frac{c^{n}}{Z^{n+1}}=\frac{1}{Z-c}.italic_f start_POSTSUPERSCRIPT italic_B italic_L end_POSTSUPERSCRIPT ( italic_Z ) = ∑ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT divide start_ARG italic_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_ARG start_ARG italic_Z start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT end_ARG = divide start_ARG 1 end_ARG start_ARG italic_Z - italic_c end_ARG . (22)

Writing fBsuperscript𝑓𝐵f^{B}italic_f start_POSTSUPERSCRIPT italic_B end_POSTSUPERSCRIPT in polar coordinates using ξ=reiθ𝜉𝑟superscript𝑒𝑖𝜃\xi=re^{-i\theta}italic_ξ = italic_r italic_e start_POSTSUPERSCRIPT - italic_i italic_θ end_POSTSUPERSCRIPT (reversed phase parametrization for convenience), and c=|c|eiϕ𝑐𝑐superscript𝑒𝑖italic-ϕc=|c|e^{i\phi}italic_c = | italic_c | italic_e start_POSTSUPERSCRIPT italic_i italic_ϕ end_POSTSUPERSCRIPT, we have that

ln(fB(r,θ))=|c|r(cos(ϕθ)+isin(ϕθ)),superscript𝑓𝐵𝑟𝜃𝑐𝑟italic-ϕ𝜃𝑖italic-ϕ𝜃\ln(f^{B}(r,\theta))=|c|r(\cos(\phi-\theta)+i\sin(\phi-\theta)),roman_ln ( italic_f start_POSTSUPERSCRIPT italic_B end_POSTSUPERSCRIPT ( italic_r , italic_θ ) ) = | italic_c | italic_r ( roman_cos ( italic_ϕ - italic_θ ) + italic_i roman_sin ( italic_ϕ - italic_θ ) ) , (23)

and therefore rln(|fB(r,θ)|)subscript𝑟superscript𝑓𝐵𝑟𝜃\partial_{r}\ln(|f^{B}(r,\theta)|)∂ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT roman_ln ( | italic_f start_POSTSUPERSCRIPT italic_B end_POSTSUPERSCRIPT ( italic_r , italic_θ ) | ) is maximized at θ=ϕ𝜃italic-ϕ\theta=\phiitalic_θ = italic_ϕ to value |c|𝑐|c|| italic_c | thereby revealing the position of the pole.

For fBLsuperscript𝑓𝐵𝐿f^{BL}italic_f start_POSTSUPERSCRIPT italic_B italic_L end_POSTSUPERSCRIPT given by a linear combination of multiple isolated poles fBL(Z)=C1Zc1+C2Zc2++CmZcmsuperscript𝑓𝐵𝐿𝑍subscript𝐶1𝑍subscript𝑐1subscript𝐶2𝑍subscript𝑐2subscript𝐶𝑚𝑍subscript𝑐𝑚f^{BL}(Z)=\frac{C_{1}}{Z-c_{1}}+\frac{C_{2}}{Z-c_{2}}+\dots+\frac{C_{m}}{Z-c_{% m}}italic_f start_POSTSUPERSCRIPT italic_B italic_L end_POSTSUPERSCRIPT ( italic_Z ) = divide start_ARG italic_C start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG italic_Z - italic_c start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG + divide start_ARG italic_C start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG start_ARG italic_Z - italic_c start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG + ⋯ + divide start_ARG italic_C start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT end_ARG start_ARG italic_Z - italic_c start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT end_ARG, the supporting function σ(θ)𝜎𝜃\sigma(\theta)italic_σ ( italic_θ ) is given by

σ(θ):=limrrln(|fB(r,θ)|)=maxj=1,2,,m|cj|cos(ϕjθ).assign𝜎𝜃subscript𝑟subscript𝑟superscript𝑓𝐵𝑟𝜃subscript𝑗12𝑚subscript𝑐𝑗subscriptitalic-ϕ𝑗𝜃\sigma(\theta):=\lim_{r\to\infty}\partial_{r}\ln(|f^{B}(r,\theta)|)=\max_{j=1,% 2,\ldots,m}|c_{j}|\cos(\phi_{j}-\theta).italic_σ ( italic_θ ) := roman_lim start_POSTSUBSCRIPT italic_r → ∞ end_POSTSUBSCRIPT ∂ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT roman_ln ( | italic_f start_POSTSUPERSCRIPT italic_B end_POSTSUPERSCRIPT ( italic_r , italic_θ ) | ) = roman_max start_POSTSUBSCRIPT italic_j = 1 , 2 , … , italic_m end_POSTSUBSCRIPT | italic_c start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT | roman_cos ( italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT - italic_θ ) . (24)

The curve σ(θ)eiθ𝜎𝜃superscript𝑒𝑖𝜃\sigma(\theta)e^{i\theta}italic_σ ( italic_θ ) italic_e start_POSTSUPERSCRIPT italic_i italic_θ end_POSTSUPERSCRIPT then describes the convex hull of the set of singularities. This also means that only the singularities of fBLsuperscript𝑓𝐵𝐿f^{BL}italic_f start_POSTSUPERSCRIPT italic_B italic_L end_POSTSUPERSCRIPT furthest from the origin are identified from the vertices of the convex hull. Since the Borel–Laplace transform ffBLmaps-to𝑓superscript𝑓𝐵𝐿f\mapsto f^{BL}italic_f ↦ italic_f start_POSTSUPERSCRIPT italic_B italic_L end_POSTSUPERSCRIPT involves the change of variable Z1/Zmaps-to𝑍1𝑍Z\mapsto 1/Zitalic_Z ↦ 1 / italic_Z, we have that the complex reciprocals of these singularities are the singularities of f𝑓fitalic_f closest to the origin.

This method is then applied to estimate the location of complex plane singularities for a real valued periodic function u(x)𝑢𝑥u(x)italic_u ( italic_x ) on S1[π,π)similar-tosuperscript𝑆1𝜋𝜋S^{1}\sim[-\pi,\pi)italic_S start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ∼ [ - italic_π , italic_π ). We write u𝑢uitalic_u in terms of Fourier series and make the analytic extension in some strip around the real axis

u(z)=ku^keikz𝑢𝑧subscript𝑘subscript^𝑢𝑘superscript𝑒𝑖𝑘𝑧u(z)=\sum_{k\in\mathbb{Z}}\hat{u}_{k}e^{ikz}italic_u ( italic_z ) = ∑ start_POSTSUBSCRIPT italic_k ∈ blackboard_Z end_POSTSUBSCRIPT over^ start_ARG italic_u end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_i italic_k italic_z end_POSTSUPERSCRIPT (25)

which is decomposed as the sum of two functions

u+(z)=k>0u^keikz,superscript𝑢𝑧subscript𝑘0subscript^𝑢𝑘superscript𝑒𝑖𝑘𝑧\displaystyle u^{+}(z)=\sum_{k>0}\hat{u}_{k}e^{ikz},italic_u start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT ( italic_z ) = ∑ start_POSTSUBSCRIPT italic_k > 0 end_POSTSUBSCRIPT over^ start_ARG italic_u end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_i italic_k italic_z end_POSTSUPERSCRIPT , (26a)
u(z)=k>0u^keikz.superscript𝑢𝑧subscript𝑘0subscript^𝑢𝑘superscript𝑒𝑖𝑘𝑧\displaystyle u^{-}(z)=\sum_{k>0}\hat{u}_{-k}e^{-ikz}.italic_u start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( italic_z ) = ∑ start_POSTSUBSCRIPT italic_k > 0 end_POSTSUBSCRIPT over^ start_ARG italic_u end_ARG start_POSTSUBSCRIPT - italic_k end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - italic_i italic_k italic_z end_POSTSUPERSCRIPT . (26b)

We note that u+superscript𝑢u^{+}italic_u start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT is analytic in the upper half plane, i.e. poles are contained in the lower half plane. Similarly, all poles of usuperscript𝑢u^{-}italic_u start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT are in the upper half plane. The change of variables Z=ez𝑍superscript𝑒𝑧Z=e^{-z}italic_Z = italic_e start_POSTSUPERSCRIPT - italic_z end_POSTSUPERSCRIPT and Z=eiz𝑍superscript𝑒𝑖𝑧Z=e^{-iz}italic_Z = italic_e start_POSTSUPERSCRIPT - italic_i italic_z end_POSTSUPERSCRIPT are made to u+superscript𝑢u^{+}italic_u start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT and usuperscript𝑢u^{-}italic_u start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT respectively to yield

f1(Z)=u+(ilnZ),subscript𝑓1𝑍superscript𝑢𝑖𝑍\displaystyle f_{1}(Z)=u^{+}(-i\ln Z),italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_Z ) = italic_u start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT ( - italic_i roman_ln italic_Z ) , (27a)
f2(Z)=u(ilnZ),subscript𝑓2𝑍superscript𝑢𝑖𝑍\displaystyle f_{2}(Z)=u^{-}(i\ln Z),italic_f start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_Z ) = italic_u start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( italic_i roman_ln italic_Z ) , (27b)

where we use the [π,π)𝜋𝜋[-\pi,\pi)[ - italic_π , italic_π ) branch cut. We note now that both f1subscript𝑓1f_{1}italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and f2subscript𝑓2f_{2}italic_f start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT are analytic inside the unit disk corresponding to the image of the upper and respectively lower half planes under the change of coordinate. Applying the BPH method on f1subscript𝑓1f_{1}italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and f2subscript𝑓2f_{2}italic_f start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT then reveals the poles closest to the unit circle and hence closest to the real axis after the logarithmic change of coordinate.

The algorithm of the BPH method is summarized as follows. For a function f𝑓fitalic_f given as a truncated power series in (19) with coefficients ansubscript𝑎𝑛a_{n}italic_a start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT, we compute its Borel transform fB(ξ)superscript𝑓𝐵𝜉f^{B}(\xi)italic_f start_POSTSUPERSCRIPT italic_B end_POSTSUPERSCRIPT ( italic_ξ ) on a grid in polar coordinates given by ξj,k=rjeiθksubscript𝜉𝑗𝑘subscript𝑟𝑗superscript𝑒𝑖subscript𝜃𝑘\xi_{j,k}=r_{j}e^{-i\theta_{k}}italic_ξ start_POSTSUBSCRIPT italic_j , italic_k end_POSTSUBSCRIPT = italic_r start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - italic_i italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_POSTSUPERSCRIPT. The grid points are given by

rj=r0+jRM,θk=2πkKformulae-sequencesubscript𝑟𝑗subscript𝑟0𝑗𝑅𝑀subscript𝜃𝑘2𝜋𝑘𝐾r_{j}=r_{0}+\frac{jR}{M},\quad\quad\theta_{k}=\frac{2\pi k}{K}italic_r start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + divide start_ARG italic_j italic_R end_ARG start_ARG italic_M end_ARG , italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = divide start_ARG 2 italic_π italic_k end_ARG start_ARG italic_K end_ARG (28)

for j=0,1,2,,M1𝑗012𝑀1j=0,1,2,\ldots,M-1italic_j = 0 , 1 , 2 , … , italic_M - 1 and k=0,1,2,,K1𝑘012𝐾1k=0,1,2,\ldots,K-1italic_k = 0 , 1 , 2 , … , italic_K - 1 with an appropriate choice of r0subscript𝑟0r_{0}italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and R𝑅Ritalic_R to be discussed later.

At each grid point, the Borel transform is given by a N𝑁Nitalic_N-truncated sum of (20). If we select the number of rays K𝐾Kitalic_K to be equal to the number of power series terms N𝑁Nitalic_N, we get that

fB(ξj,k)=n=0N1anrjnn!ei2πknN=n=0N1anrjnn!eiθkn,superscript𝑓𝐵subscript𝜉𝑗𝑘superscriptsubscript𝑛0𝑁1subscript𝑎𝑛superscriptsubscript𝑟𝑗𝑛𝑛superscript𝑒𝑖2𝜋𝑘𝑛𝑁superscriptsubscript𝑛0𝑁1subscript𝑎𝑛superscriptsubscript𝑟𝑗𝑛𝑛superscript𝑒𝑖subscript𝜃𝑘𝑛f^{B}(\xi_{j,k})=\sum_{n=0}^{N-1}a_{n}\frac{r_{j}^{n}}{n!}e^{-i\frac{2\pi kn}{% N}}=\sum_{n=0}^{N-1}a_{n}\frac{r_{j}^{n}}{n!}e^{-i\theta_{k}n},italic_f start_POSTSUPERSCRIPT italic_B end_POSTSUPERSCRIPT ( italic_ξ start_POSTSUBSCRIPT italic_j , italic_k end_POSTSUBSCRIPT ) = ∑ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N - 1 end_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT divide start_ARG italic_r start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_ARG start_ARG italic_n ! end_ARG italic_e start_POSTSUPERSCRIPT - italic_i divide start_ARG 2 italic_π italic_k italic_n end_ARG start_ARG italic_N end_ARG end_POSTSUPERSCRIPT = ∑ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N - 1 end_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT divide start_ARG italic_r start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_ARG start_ARG italic_n ! end_ARG italic_e start_POSTSUPERSCRIPT - italic_i italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_n end_POSTSUPERSCRIPT , (29)

which, for fixed j𝑗jitalic_j, is the expression for the discrete Fourier transforms of the sequence {bn}n=0N1superscriptsubscriptsubscript𝑏𝑛𝑛0𝑁1\{b_{n}\}_{n=0}^{N-1}{ italic_b start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT } start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N - 1 end_POSTSUPERSCRIPT with

bn=anrjnn!=an(rjexp(lnΓ(n+1)n))n=exp(nlnrj+lnanlnΓ(n+1)).subscript𝑏𝑛subscript𝑎𝑛superscriptsubscript𝑟𝑗𝑛𝑛subscript𝑎𝑛superscriptsubscript𝑟𝑗Γ𝑛1𝑛𝑛𝑛subscript𝑟𝑗subscript𝑎𝑛Γ𝑛1b_{n}=a_{n}\frac{r_{j}^{n}}{n!}=a_{n}\left(r_{j}\exp\left(\frac{-\ln\Gamma(n+1% )}{n}\right)\right)^{n}=\exp\left(n\ln r_{j}+\ln a_{n}-\ln\Gamma(n+1)\right).italic_b start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = italic_a start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT divide start_ARG italic_r start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_ARG start_ARG italic_n ! end_ARG = italic_a start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_r start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT roman_exp ( divide start_ARG - roman_ln roman_Γ ( italic_n + 1 ) end_ARG start_ARG italic_n end_ARG ) ) start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT = roman_exp ( italic_n roman_ln italic_r start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT + roman_ln italic_a start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT - roman_ln roman_Γ ( italic_n + 1 ) ) . (30)

FFT algorithms are used to compute the discrete Fourier transform in order to reduce round-off errors and improve speed and the log-Gamma function is used to avoid numerical overflow in the computation of the larger n𝑛nitalic_n terms.

Then for each fixed θksubscript𝜃𝑘\theta_{k}italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT, σ(θk)𝜎subscript𝜃𝑘\sigma(\theta_{k})italic_σ ( italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) can be estimated from finite difference on ln(|fB(r,θ)|)superscript𝑓𝐵𝑟𝜃\ln(|f^{B}(r,\theta)|)roman_ln ( | italic_f start_POSTSUPERSCRIPT italic_B end_POSTSUPERSCRIPT ( italic_r , italic_θ ) | ). Higher order asymptotic interpolation could be used but would require high precision arithmetic. For double precision computations, we opted for simple finite difference. From the supporting function σ(θ)𝜎𝜃\sigma(\theta)italic_σ ( italic_θ ), we can identify SBL={y1*,y2*,,ys*}superscript𝑆𝐵𝐿subscriptsuperscript𝑦1subscriptsuperscript𝑦2subscriptsuperscript𝑦𝑠S^{BL}=\{y^{*}_{1},y^{*}_{2},\ldots,y^{*}_{s}\}italic_S start_POSTSUPERSCRIPT italic_B italic_L end_POSTSUPERSCRIPT = { italic_y start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_y start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , … , italic_y start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT }, the vertices of the convex hull of the singularities of fBLsuperscript𝑓𝐵𝐿f^{BL}italic_f start_POSTSUPERSCRIPT italic_B italic_L end_POSTSUPERSCRIPT which we recall lie within the unit complex disk. The choice of the values of r0subscript𝑟0r_{0}italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and R𝑅Ritalic_R needs to be adjusted carefully. Since along the ray of each yi*superscriptsubscript𝑦𝑖y_{i}^{*}italic_y start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT, the Borel transform is expected to grow as fB(ξ)exp(yi*ξ)similar-tosuperscript𝑓𝐵𝜉superscriptsubscript𝑦𝑖𝜉f^{B}(\xi)\sim\exp(y_{i}^{*}\xi)italic_f start_POSTSUPERSCRIPT italic_B end_POSTSUPERSCRIPT ( italic_ξ ) ∼ roman_exp ( italic_y start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT italic_ξ ), in order to capture the exponential growth, one should pick r0subscript𝑟0r_{0}italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT larger than 1/|yi*|1superscriptsubscript𝑦𝑖1/|y_{i}^{*}|1 / | italic_y start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT |. The maximum radius R𝑅Ritalic_R should be picked as large as possible until the effects of finite precision arithmetic start dominating.

Applying the inverse Borel–Laplace transform gives us the set of singularities S={Z1*,Z2*,,Zs*}𝑆subscriptsuperscript𝑍1subscriptsuperscript𝑍2subscriptsuperscript𝑍𝑠S=\{Z^{*}_{1},Z^{*}_{2},\ldots,Z^{*}_{s}\}italic_S = { italic_Z start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_Z start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , … , italic_Z start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT } for f𝑓fitalic_f, where Zj*=1/yj*subscriptsuperscript𝑍𝑗1superscriptsubscript𝑦𝑗Z^{*}_{j}=1/y_{j}^{*}italic_Z start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = 1 / italic_y start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT, which are now the singularities of f𝑓fitalic_f closest to the unit circle. The locations of the singularities zi*subscriptsuperscript𝑧𝑖z^{*}_{i}italic_z start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT of u±superscript𝑢plus-or-minusu^{\pm}italic_u start_POSTSUPERSCRIPT ± end_POSTSUPERSCRIPT are then obtained by performing the inverse transform zi*=ilnZsubscriptsuperscript𝑧𝑖minus-or-plus𝑖𝑍z^{*}_{i}=\mp i\ln Zitalic_z start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = ∓ italic_i roman_ln italic_Z. This means that if one has an a priori estimate for the radius of analyticity δ𝛿\deltaitalic_δ of u𝑢uitalic_u, the value r0subscript𝑟0r_{0}italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT should be chosen larger than eδsuperscript𝑒𝛿e^{\delta}italic_e start_POSTSUPERSCRIPT italic_δ end_POSTSUPERSCRIPT.

Refer to caption
(a)
Refer to caption
(b)
Figure 14: Evolution of the complex plane singularities for the true vortex strength (a) and curvature (b) for various values of δ𝛿\deltaitalic_δ from t=0.8𝑡0.8t=0.8italic_t = 0.8 to 1.61.61.61.6. The position of the singularity at time t=1.5𝑡1.5t=1.5italic_t = 1.5 is marked by the circles. The singularity positions are computed using the BPH method using radii in [1,2.4]12.4[1,2.4][ 1 , 2.4 ]. The use of larger radii was not possible due to the presence of noise in the data.

We applied the BPH method to analyze the singularity formation in the vortex core curve γ𝛾\gammaitalic_γ. As in (Caflisch et al., 2022), we compute the closest complex singularities for the vortex strength function σ𝜎\sigmaitalic_σ and the curvature function κ𝜅\kappaitalic_κ given by

σ=|γ˙|1,𝜎superscript˙𝛾1\displaystyle\sigma=|\dot{\gamma}|^{-1},italic_σ = | over˙ start_ARG italic_γ end_ARG | start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT , (31)
κ=|γ˙×γ¨||γ˙|3.𝜅˙𝛾¨𝛾superscript˙𝛾3\displaystyle\kappa=\frac{|\dot{\gamma}\times\ddot{\gamma}|}{|\dot{\gamma}|^{3% }}.italic_κ = divide start_ARG | over˙ start_ARG italic_γ end_ARG × over¨ start_ARG italic_γ end_ARG | end_ARG start_ARG | over˙ start_ARG italic_γ end_ARG | start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG . (32)

Since the CMM relies on Hermite cubic interpolation for the stream function which yields a C0superscript𝐶0C^{0}italic_C start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT velocity field with small discontinuities in the derivative (order of h3superscript3h^{3}italic_h start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT where here h1e9similar-to1𝑒9h\sim 1e-9italic_h ∼ 1 italic_e - 9), the evolved vortex curve γ𝛾\gammaitalic_γ is also only C0superscript𝐶0C^{0}italic_C start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT. The lack of regularity shows up as noise in the large wave-numbers of γ^^𝛾\hat{\gamma}over^ start_ARG italic_γ end_ARG. To perform the BPH analysis on this data, a smoothing kernel exp(1×107k4)1superscript107superscript𝑘4\exp(-1\times 10^{-7}k^{4})roman_exp ( - 1 × 10 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT italic_k start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT ) in Fourier space and a truncation to the 10000 first wave-numbers is applied to the curve as pre-processing. The results are shown in figure 14, where the position of the complex singularities are drawn for times between 0.80.80.80.8 and 1.61.61.61.6. The general trajectories of the complex plane singularities before time t=1.5𝑡1.5t=1.5italic_t = 1.5 are in agreement with the results found for the viscous case in (Caflisch et al., 2022). We observe that all singularities move towards the real axis as time advances until t=1.5𝑡1.5t=1.5italic_t = 1.5 (marked with circles) after which they seem to move away in some cases. The meaning of this is not clear from these tests as the accuracy of the singularity analysis is very limited due to the limited available arithmetic precision. We cannot exclude the possibility that the BPH method is detecting another complex singularity which overtakes the one tracked in the above figures resulting in a transition to a new singularity which is closest to the real axis.

7 Conclusions

The flow of vortex layers governed by the incompressible 2D Euler equations has been computed for successively decreasing layer thickness using the characteristic mapping method. This semi-Lagrangian method features exponential resolution in linear time and thus allows to capture the exponential growth of the vorticity gradients. Our results agree with pseudo-spectral computations of Caflisch et al. (2022) and go even beyond their reported results. In particular, the range of results for vanishing vortex sheets thickness was extended down to δ0.007𝛿0.007\delta\approx 0.007italic_δ ≈ 0.007 and longer captured simulations enabled further analysis of the emerging vortex structures. Energy and enstrophy are conserved to a high degree thanks to the non-dissipative feature of the CM method. The palinstrophy shows at the beginning a super-exponential growth, which is intercepted and dominated by a stronger exponential growth once the two vortices form. The dynamics of the center-line of the vortex sheet can be described by measures of the arc length, curvature and vortex strength. The arc length and vortex strength showcase, that the emerging vortices have a region of very strong compression around the vortex centers and of elongation in the spiral arms that form. These form a peak in compression directly at the center and scale with decreasing vortex sheet thickness δ𝛿\deltaitalic_δ. For the limit of δ0𝛿0\delta\rightarrow 0italic_δ → 0 the vortex strength forms a singularity at the center of rotation where it is compressed into it with vortex strength tending to infinity. The curvature displays similar singular behavior. Instead of forming maximum values directly at the vortex center, it forms two at the edge of the condensed vortex blobs. These are of opposite sign and describe the transition from the vortex blob formation to the spiral arms and additionally scale with vortex sheet thickness δ𝛿\deltaitalic_δ, again going to infinity for the limit of vanishing δ𝛿\deltaitalic_δ-values, further supporting the formation of singularity for non-smooth initial data suggested in Caflisch et al. (2022). These quantities also show strong self-similarity in time and space. With the help of the curvature, vortex strength and turnover time, the temporal rescaling of the vortex dynamics was unveiled. In fact, after initial build up these vortices rotate with constant speed, similar to a solid body rotation. The maximum values for the curvature and vortex strength also do not observe monotonous growth over time, but after a steep increase they oscillate around a common value. The temporal rescaling was not found to be consistent with that of the BR-equation and similarities are clearly present. In space, the formed vortices show strong self-similarity even for many turnover events and over a large range of δ𝛿\deltaitalic_δ-values. The observed scaling slightly differs from that found in Caflisch et al. (2022) in the viscous case, i.e. for Navier–Stokes. The energy spectra show for small δ𝛿\deltaitalic_δ values a power law scaling with slope close to 22-2- 2. Simulations for longer times show, that at some point the vortex merger starts to distort the round-shaped vortex blobs, this leads to instability in the roll-up process and secondary vortices start to form, eventually breaking down the spiral structure by continuous filamentation.

In future work we will apply CMM to compute fine scale structures of vorticity gradients in 2D Euler. The transport equation of the vorticty gradients contains a source term for vorticity gradient stretching, similar to the vortex stretching source term in 3D Euler. The latter can be solved likewise by CMM, similar to what has proposed in Yin et al. (2023).

Another challenging perspective is applying CMM to study Euler flows in 3D and to investigate numerically possible singularities. First results presenting low resolution computations for the Kerr (1993); Hou & Li (2007) initial condition can be found in Yin et al. (2023). High resolution 3D computations considering different flow configurations, like Kerr (1993); Hou & Li (2007) and more recently the one by Moffatt & Kimura (2019a, b), will be published in forthcoming work.


Acknowledgements

The authors acknowledge partial funding from the Agence Nationale de la Recherche (ANR), grant ANR-20-CE46-0010-01. This work was granted access to the HPC resources of IDRIS under the allocation 2022-91664 made by GENCI. Center de Calcul Intensif d’Aix-Marseille is acknowledged for granting access to its high performance computing resources. Part of this research was finalized while KS was visiting the Institute for Mathematical and Statistical Innovation (IMSI) at University of Chicago, which is supported by the National Science Foundation (Grant No. DMS-1929348). JCN Acknowledges partial support from the NSERC Discovery Program.


 



Appendix

In the following we present numerical validation of the open access CMM cuda code (Yadav et al., 2023) similar to what has been done in Yin et al. (2021) for the Matlab implementation.

Validation of convergence order in space and time for Cuda-code (Yadav et al., 2023)


All convergence tests presented use the same parameters as in Yin et al. (2021) in order to make the results comparable. The 4-mode-flow was used as initial condition with,

ω0(x,y)=cos(x)+cos(y)+0.6cos(2x)+0.2cos(3x).subscript𝜔0𝑥𝑦𝑥𝑦0.62𝑥0.23𝑥\displaystyle\omega_{0}(x,y)=\cos(x)+\cos(y)+0.6\cos(2x)+0.2\cos(3x)\;.italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_x , italic_y ) = roman_cos ( italic_x ) + roman_cos ( italic_y ) + 0.6 roman_cos ( 2 italic_x ) + 0.2 roman_cos ( 3 italic_x ) . (33)

In table 3, all numerical parameters for the reference simulations can be found for which the error analyses have been carried out. The grid for the flow map χ𝜒\chiitalic_χ (Ncoarsesubscript𝑁𝑐𝑜𝑎𝑟𝑠𝑒N_{coarse}italic_N start_POSTSUBSCRIPT italic_c italic_o italic_a italic_r italic_s italic_e end_POSTSUBSCRIPT), initial vorticity ω0subscript𝜔0\omega_{0}italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT (Nfinesubscript𝑁𝑓𝑖𝑛𝑒N_{fine}italic_N start_POSTSUBSCRIPT italic_f italic_i italic_n italic_e end_POSTSUBSCRIPT) and sampling of the vorticity for the FFT (Nω)N_{\omega})italic_N start_POSTSUBSCRIPT italic_ω end_POSTSUBSCRIPT ) were chosen of same size and the velocity field uses a grid with increased values to reduce the influence of the non-smoothness of the velocity-gradients. The time-step ΔtΔ𝑡\Delta troman_Δ italic_t is set to a Courant–Friedrich–Lewis (CFL) number of 2222. The size ϵmsubscriptitalic-ϵ𝑚\epsilon_{m}italic_ϵ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT defines the stencil size for the GALS method. In the presented computational study no low-pass filtering and no remapping was used in order to capture the sub-map error correctly. All simulations were run until a final time of t=1𝑡1t=1italic_t = 1. For further understanding of the impact of parameters, the readers are directed to (Yin et al., 2021; Bergmann, 2022).

Name Value Name Value
Ncoarsesubscript𝑁coarseN_{\text{coarse}}italic_N start_POSTSUBSCRIPT coarse end_POSTSUBSCRIPT 1024102410241024 Nfinesubscript𝑁fineN_{\text{fine}}italic_N start_POSTSUBSCRIPT fine end_POSTSUBSCRIPT 1024102410241024
Nψsubscript𝑁𝜓N_{\psi}italic_N start_POSTSUBSCRIPT italic_ψ end_POSTSUBSCRIPT 2048204820482048 Nωsubscript𝑁𝜔N_{\omega}italic_N start_POSTSUBSCRIPT italic_ω end_POSTSUBSCRIPT 1024102410241024
hfluidsubscriptfluidh_{\text{fluid}}italic_h start_POSTSUBSCRIPT fluid end_POSTSUBSCRIPT 1/51215121/5121 / 512
ϵmsubscriptitalic-ϵ𝑚\epsilon_{m}italic_ϵ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT 103superscript10310^{-3}10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT Initial condition 4-mode-flow
Fluid time scheme RK3 Map update stencil 4th order
Table 3: Settings of the reference simulation (Yin et al., 2021).

The errors of four quantities were computed to examine the convergence order. Those are the flow map and vorticity error in Lsubscript𝐿L_{\infty}italic_L start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT-norm and the energy and enstrophy conservation error in L2subscript𝐿2L_{2}italic_L start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT-norm. All the errors were evaluated by sampling the map on a uniform 20482superscript204822048^{2}2048 start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT-grid and computing the vorticity and velocity respectively on this map.

Map error =χref(,tn)χ(,tn)absentsubscriptnormsubscript𝜒𝑟𝑒𝑓subscript𝑡𝑛𝜒subscript𝑡𝑛\displaystyle=||\chi_{ref}(\cdot,t_{n})-\chi(\cdot,t_{n})||_{\infty}= | | italic_χ start_POSTSUBSCRIPT italic_r italic_e italic_f end_POSTSUBSCRIPT ( ⋅ , italic_t start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) - italic_χ ( ⋅ , italic_t start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) | | start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT (34)
Vorticity error =ωref(,tn)ω(,tn)absentsubscriptnormsubscript𝜔𝑟𝑒𝑓subscript𝑡𝑛𝜔subscript𝑡𝑛\displaystyle=||\omega_{ref}(\cdot,t_{n})-\omega(\cdot,t_{n})||_{\infty}= | | italic_ω start_POSTSUBSCRIPT italic_r italic_e italic_f end_POSTSUBSCRIPT ( ⋅ , italic_t start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) - italic_ω ( ⋅ , italic_t start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) | | start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT (35)
Energy error =ω(,tn)22ω(,t0)22absentsubscriptnorm𝜔superscriptsubscript𝑡𝑛22subscriptnorm𝜔superscriptsubscript𝑡022\displaystyle=||\omega(\cdot,t_{n})^{2}||_{2}-||\omega(\cdot,t_{0})^{2}||_{2}= | | italic_ω ( ⋅ , italic_t start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT | | start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT - | | italic_ω ( ⋅ , italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT | | start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT (36)
Enstrophy error =𝒖(,tn)22𝒖(,t0)22absentsubscriptnorm𝒖superscriptsubscript𝑡𝑛22subscriptnorm𝒖superscriptsubscript𝑡022\displaystyle=||\bm{u}(\cdot,t_{n})^{2}||_{2}-||\bm{u}(\cdot,t_{0})^{2}||_{2}= | | bold_italic_u ( ⋅ , italic_t start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT | | start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT - | | bold_italic_u ( ⋅ , italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT | | start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT (37)

According to Yin et al. (2021), an error bound for the characteristic map is given by

~n=𝒪(Δx2min(Δt,Δx2Δt1)+Δts+Δtp)superscript~𝑛𝒪Δsuperscript𝑥2Δ𝑡Δsuperscript𝑥2Δsuperscript𝑡1Δsuperscript𝑡𝑠Δsuperscript𝑡𝑝\displaystyle\tilde{\mathcal{E}}^{n}=\mathcal{O}\left(\Delta x^{2}\min(\Delta t% ,\Delta x^{2}\Delta t^{-1})+\Delta t^{s}+\Delta t^{p}\right)over~ start_ARG caligraphic_E end_ARG start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT = caligraphic_O ( roman_Δ italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_min ( roman_Δ italic_t , roman_Δ italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Δ italic_t start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ) + roman_Δ italic_t start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT + roman_Δ italic_t start_POSTSUPERSCRIPT italic_p end_POSTSUPERSCRIPT ) (38)

Here s𝑠sitalic_s and p𝑝pitalic_p are the orders of the s-stage Runge–Kutta scheme and the used order of Lagrange time interpolation for the velocity, respectively. Values were set as p=s𝑝𝑠p=sitalic_p = italic_s to balance computation requirements with achieved numerical accuracy. Due to the bi-cubic spatial Hermite interpolation used, errors of order 𝒪(Δx3)𝒪Δsuperscript𝑥3\mathcal{O}(\Delta x^{3})caligraphic_O ( roman_Δ italic_x start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ) for the convergence order in space are expected (figure 15). One order is reduced to the theoretical 4th order as the velocity to advect the map using the GALS framework is sampled as a first order derivative from the stream function. The Cuda code is capable of including first to fourth order Runge–Kutta scheme. An example for convergence in time with third order scheme paired with third order Lagrange-interpolation of the velocity is depicted in figure 15. Again, all quantities converge with the expected order.

Refer to caption
(a) Convergence errors in space for fluid flow
Refer to caption
(b) Convergence errors in time for fluid flow
Refer to caption
(c) Convergence errors in time for fluid particles
Figure 15: Convergence errors from equation 34-37 in space and time for the flow variables and equation 40 for the fluid particles. Expected convergence rates are plotted with dotted lines.

The evolution of fluid particles used in the manuscript to track the material line of the vortex sheet, is computed using Lagrangian point particles, where the velocity is set to that of the fluid flow,

d𝒙pdt𝑑subscript𝒙𝑝𝑑𝑡\displaystyle\frac{d\bm{x}_{p}}{dt}divide start_ARG italic_d bold_italic_x start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG start_ARG italic_d italic_t end_ARG =𝒖absent𝒖\displaystyle=\bm{u}= bold_italic_u (39)

Here 𝒙psubscript𝒙𝑝\bm{x}_{p}bold_italic_x start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT is the position of the point particle and 𝒖𝒖\bm{u}bold_italic_u the fluid velocity. For time integration we tested different numerical schemes (cf. figure 14(c)), for spatial interpolation of the velocity at the particle positions bi-cubic Hermite interpolation is applied. Similarly to the map error quantities, the particle position error can be defined as

Particle error=𝒙p,ref(,tn)𝒙p(,tn)Particle errorsubscriptnormsubscript𝒙𝑝𝑟𝑒𝑓subscript𝑡𝑛subscript𝒙𝑝subscript𝑡𝑛\displaystyle\text{Particle error}=||\bm{x}_{p,ref}(\cdot,t_{n})-\bm{x}_{p}(% \cdot,t_{n})||_{\infty}Particle error = | | bold_italic_x start_POSTSUBSCRIPT italic_p , italic_r italic_e italic_f end_POSTSUBSCRIPT ( ⋅ , italic_t start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) - bold_italic_x start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ( ⋅ , italic_t start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) | | start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT (40)

The particles where initially scattered with uniform random distribution over the computational domain. Due to the volume preservation property of the incompressible flow, their distribution will remain uniform as well. All fluid parameters were kept similarly to the reference computation (table 3) and 106superscript10610^{6}10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT particles were deployed. The tested time-stepping methods show the expected convergence orders. Merely fourth order deployed time-stepping schemes are reduced to third order, which is due to the third order convergence of the fluid velocity (figure 14(c)).

References

  • Abramowitz & Stegun (1964) Abramowitz, M. & Stegun, I.A. 1964 Handbook of Mathematical Functions with Formulas, Graphs, and Mathematical Tables, 9th edn. New York: Dover.
  • Ayala & Protas (2014) Ayala, D. & Protas, B. 2014 Maximum palinstrophy growth in 2d incompressible flows. J. Fluid Mech. 742, 340–367.
  • Bardos & Titi (2013) Bardos, C.W. & Titi, E.S. 2013 Mathematics and turbulence: where do we stand? J. of Turb. 14 (3), 42–76, arXiv: https://doi.org/10.1080/14685248.2013.771838.
  • Bergmann (2022) Bergmann, J. 2022 Investigation of mixing and particle transport in 2d incompressible Euler flows using the characteristic mapping method. Master’s thesis, I2M, Aix–Marseille Université.
  • Birkhoff (1962) Birkhoff, G.D. 1962 Helmholtz and Taylor instability. In Proceedings of Symposia in Applied Mathematics, , vol. XII, pp. 55–76. Amer. Math. Soc.
  • Caflisch et al. (2017) Caflisch, R.E., Gargano, F., Sammartino, M. & Sciacca, V. 2017 Regularized Euler-α𝛼\alphaitalic_α motion of an infinite array of vortex sheets. Bolletino Unione Mat. Ital. 10, 113–141.
  • Caflisch et al. (2022) Caflisch, R.E., Gargano, F., Sammartino, M. & Sciacca, V. 2022 Complex singularity analysis for vortex layer flows. J. Fluid Mech. 932, A21.
  • Chidyagwai et al. (2011) Chidyagwai, Prince, Nave, Jean-Christophe, Rosales, Rodolfo Ruben & Seibold, Benjamin 2011 A comparative study of the efficiency of jet schemes. arXiv preprint arXiv:1104.0542 .
  • Clercx & Van Heijst (2017) Clercx, H.J.H. & Van Heijst, G.J.F. 2017 Dissipation of coherent structures in confined two-dimensional turbulence. Phys. Fluids 29 (11).
  • Constantin et al. (2019) Constantin, P., Lopes Filho, M.C., Nussenzveig Lopes, H. J. & Vicol, V. 2019 Vorticity measures and the inviscid limit. Archive for Rational Mechanics and Analysis 234, 575–593.
  • Gibbon et al. (2008) Gibbon, J.D., Bustamante, M. & Kerr, R.M. 2008 The three-dimensional Euler equations: singular or non-singular? Nonlinearity 21 (8), T123.
  • van der Hoeven (2006) van der Hoeven, J. 2006 Algorithms for asymptotic interpolation. J. of Symb. Comp. .
  • Hou & Li (2007) Hou, T.Y. & Li, R. 2007 Computing nearly singular solutions using pseudo-spectral methods. J. Comp. Phys. 226 (1), 379–397.
  • Hussaini & Zang (1987) Hussaini, M.Y. & Zang, T.A. 1987 Spectral methods in fluid dynamics. Annu. Rev. Fluid Mech. 19 (1), 339–367, arXiv: https://doi.org/10.1146/annurev.fl.19.010187.002011.
  • Ishihara et al. (2009) Ishihara, T., Gotoh, T. & Kaneda, Y. 2009 Study of high–Reynolds number isotropic turbulence by direct numerical simulation. Annu. Rev. Fluid Mech. 41 (1), 165–180, arXiv: https://doi.org/10.1146/annurev.fluid.010908.165203.
  • Kerr (1993) Kerr, R.M. 1993 Evidence for a singularity of the three‐dimensional, incompressible Euler equations. Phys. Fluids A: Fluid Dyn. 5 (7), 1725–1746, arXiv: https://pubs.aip.org/aip/pof/article-pdf/5/7/1725/12783847/1725_1_online.pdf.
  • Lesieur (2008) Lesieur, Marcel 2008 Two-dimensional turbulence. Turbulence in Fluids: Fourth Revised and Enlarged Edition pp. 311–348.
  • Majda & Bertozzi (2001) Majda, A.J. & Bertozzi, A.L. 2001 Vorticity and Incompressible Flow. Cambridge Texts in Applied Mathematics . Cambridge University Press.
  • Mercier et al. (2020) Mercier, Olivier, Yin, Xi-Yuan & Nave, Jean-Christophe 2020 The characteristic mapping method for the linear advection of arbitrary sets. SIAM Journal on Scientific Computing 42 (3), A1663–A1685.
  • Moffatt & Kimura (2019a) Moffatt, H.K. & Kimura, Y. 2019a Towards a finite-time singularity of the Navier–Stokes equations part 1. derivation and analysis of dynamical system. J. Fluid Mech. 861, 930–967.
  • Moffatt & Kimura (2019b) Moffatt, H.K. & Kimura, Y. 2019b Towards a finite-time singularity of the Navier–Stokes equations. part 2. vortex reconnection and singularity evasion. J. Fluid Mech. 870, R1.
  • Moore & Stuart (1979) Moore, D.W. & Stuart, J.T. 1979 The spontaneous appearance of a singularity in the shape of an evolving vortex sheet. Proc. R. Soc. A: Math. Phys. Eng. Sci. 365 (1720), 105–119, arXiv: https://royalsocietypublishing.org/doi/pdf/10.1098/rspa.1979.0009.
  • Nave et al. (2010) Nave, J.-C., Rosales, R.R. & Seibold, B. 2010 A gradient-augmented level set method with an optimally local, coherent advection scheme. J. Comput. Phys. 229 (10), 3802–3827.
  • Nguyen Van Yen et al. (2018) Nguyen Van Yen, N., Waidmann, M., Klein, R., Farge, M. & Schneider, K. 2018 Energy dissipation caused by boundary layer instability at vanishing viscosity. J. Fluid Mech. 849, 676–717.
  • Nguyen Van Yen et al. (2011) Nguyen Van Yen, R., Farge, M. & Schneider, K. 2011 Energy dissipating structures produced by walls in two-dimensional flows at vanishing viscosity. Phys. Rev. Lett. 106 (18), 184502.
  • Pauls & Frisch (2007) Pauls, W. & Frisch, U. 2007 A Borel transform method for locating singularities of Taylor and Fourier series. J. Stat. Phys. 127 (6), 1095–1119.
  • Podvigina et al. (2016) Podvigina, O., Zheligovsky, V. & Frisch, U. 2016 The Cauchy–Lagrangian method for numerical analysis of Euler flow. J. Comput. Phys. 306, 320–342.
  • Pólya (1933) Pólya, G. 1933 Untersuchungen über Lücken und Singularitäten von Potenzreihen. Ann. Math. 34 (4), 731–777.
  • Rott (1956) Rott, N. 1956 Diffraction of a weak shock with vortex generation. J. Fluid Mech. 1 (1), 111–128.
  • Seibold et al. (2012) Seibold, Benjamin, Rosales, Rodolfo R. & Nave, Jean-Christophe 2012 Jet schemes for advection problems. Discrete and Continuous Dynamical Systems - B 17 (4), 1229–1259.
  • Székelyhidi (2011) Székelyhidi, L. 2011 Weak solutions to the incompressible Euler equations with vortex sheet initial data. C. R. Math. 349 (19), 1063–1066.
  • Yadav et al. (2023) Yadav, B., Maurel Oujia, T., Saber, N., Bergmann, J. & Krah, P. 2023 CMM cuda code. https://github.com/CharacteristicMappingMethod/cmm-turbulence.
  • Yin et al. (2021) Yin, X.-Y., Mercier, O., Yadav, B., Schneider, K. & Nave, J.-C. 2021 A characteristic mapping method for the two-dimensional incompressible Euler equations. J. Comput. Phys. 424, 109781.
  • Yin et al. (2023) Yin, X.-Y., Schneider, K. & Nave, J.-C. 2023 A characteristic mapping method for the three-dimensional incompressible Euler equations. J. Comput. Phys. 477, 111876.
  • Yudovich (1963) Yudovich, V.I. 1963 Non-stationary flow of an ideal incompressible liquid. USSR Comput. Math. & Math. Phys. 3 (6), 1407–1456.