Relevance of the Basset history term for Lagrangian particle dynamics

Julio Urizarna-Carasa    Daniel Ruprecht [email protected] Lehrstuhl Computational Mathematics, Institut für Mathematik, Technische Universität Hamburg, Hamburg, Germany    Alexandra von Kameke [email protected] Heinrich-Blasius-Institute, Faculty of Engineering and Computer Science, Hamburg University of Applied Sciences, Hamburg, Germany    Kathrin Padberg-Gehle [email protected]. Applied Mathematics, Institute for Mathematics and its Didactics, Leuphana University Lüneburg, Lüneburg, Germany
(July 1, 2024)
Abstract

The movement of small but finite spherical particles in a fluid can be described by the Maxey-Riley equation (MRE) if they are too large to be considered passive tracers. The MRE contains an integral "history term" modeling wake effects, which causes the force acting on a particle at some given time to depend on its full past trajectory. The history term causes complications in the numerical solution of the MRE and is therefore often neglected, despite both numerical and experimental evidence that its effects are generally not negligible. By numerically computing trajectories with and without the history term of a large number of particles in different flow fields, we investigate its impact on the large-scale Lagrangian dynamics of simulated particles. We show that for moderate to large Stokes numbers, ignoring the history term leads to significant differences in clustering patterns. Furthermore, we compute finite-time Lyapunov exponents and show that, even for small particles, the differences in the resulting scalar field from ignoring the BHT can be significant, in particular if the underlying flow is turbulent.

preprint: AIP/123-QED

Neglecting the Basset history term in the Maxey-Riley equation can cause significant distortions not only of the trajectories of individual particles but also of the macroscopic Lagrangian dynamics.

I Introduction

Fluid motion is all around us, in the ocean, in the atmosphere, in industrial installations and in everyday live. Oftentimes, the fluid carries other materials with it, be it plankton, buoys, dust, particulate matter or any other type of small particulate pieces of immiscible material that does not dissolve in the carrier liquid. If the volume fraction of the particulate pieces is low, it can be assumed that they move in isolation and do not collide or affect each other’s movement. Furthermore, if their mass fraction is low, the effect they have on the overall flow field of the carrier phase can be neglected. Thorough discussions of the many different aspects of the mathematical description of the motion of inertial particles and real world applications can be found in the literature [1].

The equation of motion for inertial but small particles that emerges from an inspection of the acting forces is called the Maxey-Riley equation (MRE) [2]. The MRE accounts for the Basset history force, which models the influence of the particle’s past accelerations on its present motion. The equation is valid for particles of intermediate size that are too large to be considered passive tracers but not so large that they significantly disturb the fluid or that surface effects become important. The alternative is using a fully coupled fluid-structure interaction model, which, while detailed and realistic, requires massive computational effort and access to a powerful high-performance computing system [3].

Mathematically, the MRE is a second order integro-differential equation. The integral or Basset history term (BHT), which captures the history force, is difficult to handle numerically due to its non-local nature and is thus typically ignored [4, 5]. However, depending on the size and density of the particle, ignoring the history term changes the simulated trajectories significantly [6, 7, 8]. The importance of the BHT for matching simulations of particle trajectories to experiments has also been confirmed: Candelier et al. [9] investigate the impact of the BHT experimentally to elucidate how a particle is ejected out of a vortex flow and find that “calculations without history force overestimate particle ejection”. Similar observations have been made in simulations[10].

In many applications, one is interested in the dynamics of ensembles of particles rather than in single particle trajectories. Such a macroscopic view on Lagrangian particle dynamics is particularly relevant in the context of studying transport and mixing processes. Lagrangian coherent structures (LCS) form the time-dependent skeleton of the flow, encompassing regions of stretching and folding that enhance or mitigate particle transport[11]. Many different computational approaches have been developed over the past almost three decades to identify coherent flow structures, such as LCS or coherent sets, and to study their dynamics, including bifurcations [12, 13, 14]. Finite-time Lyapunov exponents (FTLE) are often heuristically used to highlight regions of different dynamical behavior. Inertial particles are known to interact with the underlying flow skeleton and tend to concentrate along different coherent structures, for example vortices, depending on their material properties such as size and density[15, 16]. There are several studies that compare the FTLE fields for ideal tracers that exactly follow the underlying flow with those computed for inertial particles. They found that the corresponding flow structures can be crucially different [17, 18]. However, these studies did not consider the history term and the specific influence of the BHT on the LCS has not yet been studied. Therefore, this paper investigates the differences in flow structures for particle dynamics simulated with and without Basset term. For this, we make use of recent mathematical developments in the numerical solution of the Maxey-Riley equation with history term[6, 19], which has opened the door to the simulation of many particles[20].

I.1 Contributions

Our paper investigates the impact of neglecting the BHT on the macroscopic dynamics of simulated Lagrangian particles. We retain the assumption that the flow influences the particle motion but that there is no impact on the background flow, allowing us to use the MRE as model. We study the resulting flow patterns in a controlled setting for three example systems, the double gyre [21], the Bickley jet [22], and a Faraday flow, a two-dimensional, fully turbulent, experimentally measured flow [23, 24].

We compare particle dispersion and finite time Lyapunov exponents (FTLE) for particle trajectories computed with and without history term with different Stokes numbers and densities. Our analysis demonstrates that, even for moderate Stokes numbers of unity or more, the particle dispersion patterns revealed by the FTLE are noticeably changed by ignoring the history term. In particular, our results suggest that simulations without BHT produce results that are comparable to flows with an effective Stokes number that is larger than used value. This is very much in accordance with previous findings [9, 10]. Our results also suggest that care must be taken when attempting to predict flow regimes, or sudden changes thereof, that involve inertial particle dynamics. Important examples would be the oil spill in the Gulf of Mexico, ocean search-and-rescue [25], or the splitting of the polar vortex.

I.2 Outline

Section II introduces the the Maxey-Riley equation, its numerical solution and the computation of finite-time Lyapunov exponents. Section III describes the three flow fields used in our study. In Section IV we present the results of our investigation. First we analyse and compare the final particle positions and their preferential concentration. Second, we present FTLE fields for a range of different parameter configurations. Conclusions and a summary can be found in Section V.

II Background

In this section, we introduce the Maxey-Riley equation (MRE) and its numerical treatment as well as the equation and approximations used to obtain the finite-time Lyapunov exponents (FTLE).

II.1 The Maxey-Riley Equation (MRE)

Throughout this work, we consider (i) spherical particles with a small radius compared to the flow dimensions, which guarantees a small particle Reynolds number [4], (ii) one-way interactions between particles and fluid, so that the flow influences the particle dynamics but not the other way round [6] and, (iii) no particle collisions. Under these assumptions, particle trajectories can be modeled using the MRE in nondimensionalized form

𝒗˙=bold-˙𝒗absent\displaystyle\bm{\dot{v}}=overbold_˙ start_ARG bold_italic_v end_ARG = 1RD𝒖Dt+limit-from1𝑅𝐷𝒖𝐷𝑡\displaystyle\;\frac{1}{R}\frac{D\bm{u}}{Dt}\;+divide start_ARG 1 end_ARG start_ARG italic_R end_ARG divide start_ARG italic_D bold_italic_u end_ARG start_ARG italic_D italic_t end_ARG + (1a)
1RS(𝒗𝒖)+limit-from1𝑅𝑆𝒗𝒖\displaystyle\;-\frac{1}{RS}\left(\bm{v}-\bm{u}\right)\;+- divide start_ARG 1 end_ARG start_ARG italic_R italic_S end_ARG ( bold_italic_v - bold_italic_u ) + (1b)
1R3πS{1t(𝒗(t0)𝒖(t0))+t0t𝒗(s)𝒖(s)ts𝑑s},1𝑅3𝜋𝑆1𝑡𝒗subscript𝑡0𝒖subscript𝑡0superscriptsubscriptsubscript𝑡0𝑡𝒗𝑠𝒖𝑠𝑡𝑠differential-d𝑠\displaystyle\;-\frac{1}{R}\sqrt{\frac{3}{\pi S}}\left\{\frac{1}{\sqrt{t}}% \left(\bm{v}(t_{0})-\bm{u}(t_{0})\right)+\int_{t_{0}}^{t}\frac{\bm{v}(s)-\bm{u% }(s)}{\sqrt{t-s}}ds\right\},- divide start_ARG 1 end_ARG start_ARG italic_R end_ARG square-root start_ARG divide start_ARG 3 end_ARG start_ARG italic_π italic_S end_ARG end_ARG { divide start_ARG 1 end_ARG start_ARG square-root start_ARG italic_t end_ARG end_ARG ( bold_italic_v ( italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) - bold_italic_u ( italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) ) + ∫ start_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT divide start_ARG bold_italic_v ( italic_s ) - bold_italic_u ( italic_s ) end_ARG start_ARG square-root start_ARG italic_t - italic_s end_ARG end_ARG italic_d italic_s } , (1c)

as adapted by Prasath et al. [6] from the original paper by Maxey and Riley [2], where 𝒗:=𝒙˙(t)assign𝒗bold-˙𝒙𝑡\bm{v}:=\bm{\dot{x}}(t)bold_italic_v := overbold_˙ start_ARG bold_italic_x end_ARG ( italic_t ) is the particle’s absolute velocity, 𝒙(t)𝒙𝑡\bm{x}(t)bold_italic_x ( italic_t ) the particle’s position and 𝒖:=𝒖(𝒙(t),t)assign𝒖𝒖𝒙𝑡𝑡\bm{u}:=\bm{u}(\bm{x}(t),t)bold_italic_u := bold_italic_u ( bold_italic_x ( italic_t ) , italic_t ) the value of the Eulerian velocity field at the particle’s position. Note that we only consider particles moving in a plane so that 𝒙(t),𝒗(t)2𝒙𝑡𝒗𝑡superscript2\bm{x}(t),\bm{v}(t)\in\mathbb{R}^{2}bold_italic_x ( italic_t ) , bold_italic_v ( italic_t ) ∈ blackboard_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT throughout. Further,

β:=ρpρf,assign𝛽subscript𝜌𝑝subscript𝜌𝑓\displaystyle\beta:=\frac{\rho_{p}}{\rho_{f}},italic_β := divide start_ARG italic_ρ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG start_ARG italic_ρ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT end_ARG , R:=1+2β3,assign𝑅12𝛽3\displaystyle R:=\frac{1+2\beta}{3},italic_R := divide start_ARG 1 + 2 italic_β end_ARG start_ARG 3 end_ARG , S:=13a2νT,assign𝑆13superscript𝑎2𝜈𝑇\displaystyle S:=\frac{1}{3}\frac{a^{2}}{\nu T},italic_S := divide start_ARG 1 end_ARG start_ARG 3 end_ARG divide start_ARG italic_a start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_ν italic_T end_ARG , (2)

where ρpsubscript𝜌𝑝\rho_{p}italic_ρ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT, ρfsubscript𝜌𝑓\rho_{f}italic_ρ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT correspond to the particle and fluid densities, a𝑎aitalic_a is the particle’s radius, ν𝜈\nuitalic_ν its kinematic viscosity and T𝑇Titalic_T the time scale of the flow. The effective density ratio R𝑅Ritalic_R controls buoyancy and inertia and accounts for the added mass effect that arises because the particle accelerates surrounding fluid and carries it along its path. Heavier particles that are denser than the carrier fluid with β>1𝛽1\beta>1italic_β > 1, where also R>1𝑅1R>1italic_R > 1, are less susceptible to the fluid forces than lighter particles, i.e. β<1𝛽1\beta<1italic_β < 1 with R<1𝑅1R<1italic_R < 1, compare for the particle acceleration and first term in (1a).

The nondimensional parameter S𝑆Sitalic_S in (1) characterizes the particle radius with respect to the flow dimensions. We refer to S𝑆Sitalic_S as the Stokes number since it compares the particle relaxation timescale τp=a2νsubscript𝜏𝑝superscript𝑎2𝜈\tau_{p}=\frac{a^{2}}{\nu}italic_τ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = divide start_ARG italic_a start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_ν end_ARG to a flow timescale T𝑇Titalic_T. Often, T𝑇Titalic_T is chosen to be the time scale of the mean flow or eddy turnover time but setting it properly requires some care [26]. We compute LCS for values of S=0.1𝑆0.1S=0.1italic_S = 0.1, S=1𝑆1S=1italic_S = 1 and S=10𝑆10S=10italic_S = 10. This is in line with values studied in other papers [27, 7].

However, note that the Stokes numbers in both references correspond to the experimental Stokes number

Sexp=S23ρpρf=29a2νρpρf1T=29a2L2ρpρfULν=29a2L2ρpρfRefsubscript𝑆𝑒𝑥𝑝𝑆23subscript𝜌𝑝subscript𝜌𝑓29superscript𝑎2𝜈subscript𝜌𝑝subscript𝜌𝑓1𝑇29superscript𝑎2superscript𝐿2subscript𝜌𝑝subscript𝜌𝑓𝑈𝐿𝜈29superscript𝑎2superscript𝐿2subscript𝜌𝑝subscript𝜌𝑓𝑅subscript𝑒𝑓S_{exp}=S\cdot\frac{2}{3}\frac{\rho_{p}}{\rho_{f}}=\frac{2}{9}\frac{a^{2}}{\nu% }\frac{\rho_{p}}{\rho_{f}}\frac{1}{T}=\frac{2}{9}\frac{a^{2}}{L^{2}}\frac{\rho% _{p}}{\rho_{f}}\frac{UL}{\nu}=\frac{2}{9}\frac{a^{2}}{L^{2}}\frac{\rho_{p}}{% \rho_{f}}Re_{f}italic_S start_POSTSUBSCRIPT italic_e italic_x italic_p end_POSTSUBSCRIPT = italic_S ⋅ divide start_ARG 2 end_ARG start_ARG 3 end_ARG divide start_ARG italic_ρ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG start_ARG italic_ρ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT end_ARG = divide start_ARG 2 end_ARG start_ARG 9 end_ARG divide start_ARG italic_a start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_ν end_ARG divide start_ARG italic_ρ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG start_ARG italic_ρ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT end_ARG divide start_ARG 1 end_ARG start_ARG italic_T end_ARG = divide start_ARG 2 end_ARG start_ARG 9 end_ARG divide start_ARG italic_a start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG divide start_ARG italic_ρ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG start_ARG italic_ρ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT end_ARG divide start_ARG italic_U italic_L end_ARG start_ARG italic_ν end_ARG = divide start_ARG 2 end_ARG start_ARG 9 end_ARG divide start_ARG italic_a start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG divide start_ARG italic_ρ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG start_ARG italic_ρ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT end_ARG italic_R italic_e start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT (3)

with U𝑈Uitalic_U, L𝐿Litalic_L and Ref𝑅subscript𝑒𝑓Re_{f}italic_R italic_e start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT being the flow velocity and timescales and the corresponding flow Reynolds number. The experimential Stokes number is slightly different from the non-dimensional S𝑆Sitalic_S we prescribe here. Experimentalists tend to use Sexpsubscript𝑆expS_{\textrm{exp}}italic_S start_POSTSUBSCRIPT exp end_POSTSUBSCRIPT, which also incorporates the density ratio, to examine how closely the particle follows the flow [28]. This Stokes number arises as a prefactor of the Stokes drag term, not mingled with the added mass term, in the non-dimensionalized Maxey-Riley equation and is widely used in the literature [29]. It allows to account for non-Stokesian drag when Re>1𝑅𝑒1Re>1italic_R italic_e > 1 by using the expression Sexp=fa2νρpρf1Tsubscript𝑆exp𝑓superscript𝑎2𝜈subscript𝜌𝑝subscript𝜌𝑓1𝑇S_{\textrm{exp}}=f\cdot\frac{a^{2}}{\nu}\frac{\rho_{p}}{\rho_{f}}\frac{1}{T}italic_S start_POSTSUBSCRIPT exp end_POSTSUBSCRIPT = italic_f ⋅ divide start_ARG italic_a start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_ν end_ARG divide start_ARG italic_ρ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG start_ARG italic_ρ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT end_ARG divide start_ARG 1 end_ARG start_ARG italic_T end_ARG, where f𝑓fitalic_f is a factor that depends on the particle Reynolds number f=f(Rep)𝑓𝑓𝑅subscript𝑒𝑝f=f(Re_{p})italic_f = italic_f ( italic_R italic_e start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ) [30]. For our density ratios of β=ρp/ρf=2/3𝛽subscript𝜌𝑝subscript𝜌𝑓23\beta=\rho_{p}/\rho_{f}=2/3italic_β = italic_ρ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT / italic_ρ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT = 2 / 3 and β=4/3𝛽43\beta=4/3italic_β = 4 / 3, a value of Sexp=10subscript𝑆exp10S_{\textrm{exp}}=10italic_S start_POSTSUBSCRIPT exp end_POSTSUBSCRIPT = 10 as studied by Olivieri et al. [7] corresponds to a Stokes number as used in this article of S=22.50𝑆22.50S=22.50italic_S = 22.50 and S=11.25𝑆11.25S=11.25italic_S = 11.25 respectively.

The MRE is derived from Newton’s second law for the particle movement and a velocity field approximation of the influence of the particle on the local flow from potential flow theory under cree** flow assumption. This implies that the validity of the formula is only guaranteed for low particle Reynolds numbers with low relative velocities. Terms on the right hand side of equation (1) correspond to different forces acting on the particle. Term (1a) are the forces acting on a particle in an unperturbed fluid lumped up with the added mass effect. Term (1b) is the Stokes drag. Term (1c), the so-called Basset History Term (BHT), accounts for the effects of the lagging boundary layer around the sphere [31].

Since the aim of our work is to examine the influence of the history force, we neglect the so-called Faxen corrections that model the influence of the particle’s finite size and scale with the curvature of the velocity field. However, for larger particles, the Faxen corrections should be considered as they have been shown to play a crucial role in shear flows [32] where they cause fluctuations in the particle’s velocity and thus a relative velocity between particle and fluid flow. In (1), since we assume that the particle initially has same velocity as the fluid, only particles with buoyancy R1𝑅1R\neq 1italic_R ≠ 1 will produce non-zero relative velocities and cause the trajectories of particles with and without history force to deviate.

II.2 Numerical solution of the MRE

The BHT is made up of a history integral with a singular kernel. This term changes the MRE from an ordinary differential into an integro-differential equation that is not easily solvable either analytically or numerically. For this reason, the term is often omitted [33], modified [34, 35, 36, 37] or approximated [38, 39, 40, 41]. Several numerical approximations, based on quadrature schemes, were obtained by Van Hinsberg et al. [41] and Daitche [42]. However these schemes become storage-intensive for large time grids. In 2019, Prasath et al. showed that the MRE can be written "as a forced, time-dependent Robin boundary condition of the one-dimensional diffusion equation" [6] and provided a storage-efficient numerical scheme. In this paper, trajectories with BHT are calculated with the second order finite-difference, IMEX solver (FD2 + IMEX2) by Urizarna-Carasa et al. [19] that relies on Prasath’s reformulation. The scheme uses the second order spatial discretization provided by Koleva [43] coupled with a second order Midpoint IMEX method given by Ascher et al. [44] to solve the reformulated MRE. Trajectories without BHT are calculated with the explicit adaptive Runge-Kutta method of order 5(4) provided by the solve_ivp solver of Python’s SciPy [45] library, using a relative and absolute tolerance of 108superscript10810^{-8}10 start_POSTSUPERSCRIPT - 8 end_POSTSUPERSCRIPT. For verification, trajectories without BHT were compared against the Leap-Frog method, whereas trajectories with BHT were compared against Daitche’s 3rdsuperscript3𝑟𝑑3^{rd}3 start_POSTSUPERSCRIPT italic_r italic_d end_POSTSUPERSCRIPT order method [42]. All figures shown in this paper can be reproduced using the provided code [46].

II.3 Finite-time Lyapunov exponents (FTLE)

Finite-time Lyapunov exponents (FTLE) are a measure of the exponential growth of an infinitesimal perturbation in the initial condition under the action of the dynamical system. FTLE fields have become a popular, heuristic tool in nonlinear dynamics to highlight organizing structures in phase space such as invariant manifolds of hyperbolic objects or Lagrangian coherent structures in time-dependent flows[21, 11]. Typically FTLE have large values for repelling LCS when computed in forward time, and for attracting ones when computed in backward time. Close to zero values are frequently observed for elliptic motion, such as in the center of vortices. While more sophisticated approaches have been developed to identify LCS [11], we use FTLE as a convenient and meaningful scalar field that allows to visualize regions of different dynamical behavior.

The expression to compute the FTLE of a flow field at a position 𝒙𝒙\bm{x}bold_italic_x is

σt0t(𝒙):=1|tt0|ln(λmax(Δ)),assignsuperscriptsubscript𝜎subscript𝑡0𝑡𝒙1𝑡subscript𝑡0subscript𝜆𝑚𝑎𝑥Δ\displaystyle\sigma_{t_{0}}^{t}(\bm{x}):=\frac{1}{\lvert t-t_{0}\rvert}\ln% \left(\sqrt{\lambda_{max}(\Delta)}\right),italic_σ start_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT ( bold_italic_x ) := divide start_ARG 1 end_ARG start_ARG | italic_t - italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT | end_ARG roman_ln ( square-root start_ARG italic_λ start_POSTSUBSCRIPT italic_m italic_a italic_x end_POSTSUBSCRIPT ( roman_Δ ) end_ARG ) , (4)

where 𝒙=(x,y)T𝒙superscript𝑥𝑦𝑇\bm{x}=(x,y)^{T}bold_italic_x = ( italic_x , italic_y ) start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT is the position vector, λmax(Δ)subscript𝜆𝑚𝑎𝑥Δ\lambda_{max}(\Delta)italic_λ start_POSTSUBSCRIPT italic_m italic_a italic_x end_POSTSUBSCRIPT ( roman_Δ ) is the largest eigenvalue of the Cauchy-Green deformation tensor Δ:=(DΦt0t)DΦt0tassignΔsuperscript𝐷superscriptsubscriptΦsubscript𝑡0𝑡𝐷superscriptsubscriptΦsubscript𝑡0𝑡\Delta:=\left(D\Phi_{t_{0}}^{t}\right)^{*}D\Phi_{t_{0}}^{t}roman_Δ := ( italic_D roman_Φ start_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT italic_D roman_Φ start_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT,and DΦt0t𝐷superscriptsubscriptΦsubscript𝑡0𝑡D\Phi_{t_{0}}^{t}italic_D roman_Φ start_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT is the deformation gradient of the flow map. The flow map Φt0tsuperscriptsubscriptΦsubscript𝑡0𝑡\Phi_{t_{0}}^{t}roman_Φ start_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT is evaluated by integrating (1) using either solve_ivp or the FD2 + IMEX2 scheme described above. DΦt0t𝐷superscriptsubscriptΦsubscript𝑡0𝑡D\Phi_{t_{0}}^{t}italic_D roman_Φ start_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT can either be obtained by solving a variational equation or by approximating the derivative numerically using finite-differences. The latter is straightforward to implement and thus used in most cases, even when a kinematic model is available.

FTLE methods have also been applied in the context of inertial particle dynamics as described by the Maxey-Riley equation [17, 16, 18] and perturbations both in positions and velocities have been taken into account to define inertial finite-time Lyapunv exponents (iFTLE). In that case not only partial derivatives with respect to positions and velocities would make up the deformation gradient of the flow map, resulting in a 4×4444\times 44 × 4 matrix in the case of a two-dimensional flow velocity field. Since our aim is to study the influence of the Basset history term on particle dynamics and transport properties by comparing flow structures computed with and without history term we neglect perturbations in the initial velocities and the flow map only considers the particle positions. Therefore, in our study, DΦt0t𝐷superscriptsubscriptΦsubscript𝑡0𝑡D\Phi_{t_{0}}^{t}italic_D roman_Φ start_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT is a 2×2222\times 22 × 2 deformation gradient.

Numerical particle trajectories can only be calculated at discrete points in space. Thus, the entries of the deformation gradient of a particle in an interior point of the domain (position 𝒙i,jsubscript𝒙𝑖𝑗\bm{x}_{i,j}bold_italic_x start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT in the stencil shown in Figure 1) are approximated by the second order finite-difference scheme [17]

DΦt0t|1,1\displaystyle D\Phi_{t_{0}}^{t}\rvert_{1,1}italic_D roman_Φ start_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT | start_POSTSUBSCRIPT 1 , 1 end_POSTSUBSCRIPT xi+1,j(tend)xi1,j(tend)xi+1,j(t0)xi1,j(t0),absentsubscript𝑥𝑖1𝑗subscript𝑡𝑒𝑛𝑑subscript𝑥𝑖1𝑗subscript𝑡𝑒𝑛𝑑subscript𝑥𝑖1𝑗subscript𝑡0subscript𝑥𝑖1𝑗subscript𝑡0\displaystyle\approx\frac{x_{i+1,j}(t_{end})-x_{i-1,j}(t_{end})}{x_{i+1,j}(t_{% 0})-x_{i-1,j}(t_{0})},≈ divide start_ARG italic_x start_POSTSUBSCRIPT italic_i + 1 , italic_j end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT italic_e italic_n italic_d end_POSTSUBSCRIPT ) - italic_x start_POSTSUBSCRIPT italic_i - 1 , italic_j end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT italic_e italic_n italic_d end_POSTSUBSCRIPT ) end_ARG start_ARG italic_x start_POSTSUBSCRIPT italic_i + 1 , italic_j end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) - italic_x start_POSTSUBSCRIPT italic_i - 1 , italic_j end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) end_ARG , (5a)
DΦt0t|1,2\displaystyle D\Phi_{t_{0}}^{t}\rvert_{1,2}italic_D roman_Φ start_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT | start_POSTSUBSCRIPT 1 , 2 end_POSTSUBSCRIPT xi,j+1(tend)xi,j1(tend)yi,j+1(t0)yi,j1(t0),absentsubscript𝑥𝑖𝑗1subscript𝑡𝑒𝑛𝑑subscript𝑥𝑖𝑗1subscript𝑡𝑒𝑛𝑑subscript𝑦𝑖𝑗1subscript𝑡0subscript𝑦𝑖𝑗1subscript𝑡0\displaystyle\approx\frac{x_{i,j+1}(t_{end})-x_{i,j-1}(t_{end})}{y_{i,j+1}(t_{% 0})-y_{i,j-1}(t_{0})},≈ divide start_ARG italic_x start_POSTSUBSCRIPT italic_i , italic_j + 1 end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT italic_e italic_n italic_d end_POSTSUBSCRIPT ) - italic_x start_POSTSUBSCRIPT italic_i , italic_j - 1 end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT italic_e italic_n italic_d end_POSTSUBSCRIPT ) end_ARG start_ARG italic_y start_POSTSUBSCRIPT italic_i , italic_j + 1 end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) - italic_y start_POSTSUBSCRIPT italic_i , italic_j - 1 end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) end_ARG , (5b)
DΦt0t|2,1\displaystyle D\Phi_{t_{0}}^{t}\rvert_{2,1}italic_D roman_Φ start_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT | start_POSTSUBSCRIPT 2 , 1 end_POSTSUBSCRIPT yi+1,j(tend)yi1,j(tend)xi+1,j(t0)xi1,j(t0),absentsubscript𝑦𝑖1𝑗subscript𝑡𝑒𝑛𝑑subscript𝑦𝑖1𝑗subscript𝑡𝑒𝑛𝑑subscript𝑥𝑖1𝑗subscript𝑡0subscript𝑥𝑖1𝑗subscript𝑡0\displaystyle\approx\frac{y_{i+1,j}(t_{end})-y_{i-1,j}(t_{end})}{x_{i+1,j}(t_{% 0})-x_{i-1,j}(t_{0})},≈ divide start_ARG italic_y start_POSTSUBSCRIPT italic_i + 1 , italic_j end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT italic_e italic_n italic_d end_POSTSUBSCRIPT ) - italic_y start_POSTSUBSCRIPT italic_i - 1 , italic_j end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT italic_e italic_n italic_d end_POSTSUBSCRIPT ) end_ARG start_ARG italic_x start_POSTSUBSCRIPT italic_i + 1 , italic_j end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) - italic_x start_POSTSUBSCRIPT italic_i - 1 , italic_j end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) end_ARG , (5c)
DΦt0t|2,2\displaystyle D\Phi_{t_{0}}^{t}\rvert_{2,2}italic_D roman_Φ start_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT | start_POSTSUBSCRIPT 2 , 2 end_POSTSUBSCRIPT yi,j+1(tend)yi,j1(tend)yi,j+1(t0)yi,j1(t0).absentsubscript𝑦𝑖𝑗1subscript𝑡𝑒𝑛𝑑subscript𝑦𝑖𝑗1subscript𝑡𝑒𝑛𝑑subscript𝑦𝑖𝑗1subscript𝑡0subscript𝑦𝑖𝑗1subscript𝑡0\displaystyle\approx\frac{y_{i,j+1}(t_{end})-y_{i,j-1}(t_{end})}{y_{i,j+1}(t_{% 0})-y_{i,j-1}(t_{0})}.≈ divide start_ARG italic_y start_POSTSUBSCRIPT italic_i , italic_j + 1 end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT italic_e italic_n italic_d end_POSTSUBSCRIPT ) - italic_y start_POSTSUBSCRIPT italic_i , italic_j - 1 end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT italic_e italic_n italic_d end_POSTSUBSCRIPT ) end_ARG start_ARG italic_y start_POSTSUBSCRIPT italic_i , italic_j + 1 end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) - italic_y start_POSTSUBSCRIPT italic_i , italic_j - 1 end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) end_ARG . (5d)

At the boundaries, where values outside the initial domain are not available, a first order approximation is used. Following Shadden et al. [21], all FTLE plots in this paper show σt0t(𝒙)|tt0|superscriptsubscript𝜎subscript𝑡0𝑡𝒙𝑡subscript𝑡0\sigma_{t_{0}}^{t}(\bm{x})\lvert t-t_{0}\rvertitalic_σ start_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT ( bold_italic_x ) | italic_t - italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT |.

𝒙i1,j(t0)subscript𝒙𝑖1𝑗subscript𝑡0\bm{x}_{i-1,j}(t_{0})bold_italic_x start_POSTSUBSCRIPT italic_i - 1 , italic_j end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT )𝒙i,jsubscript𝒙𝑖𝑗\bm{x}_{i,j}bold_italic_x start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT𝒙i+1,j(t0)subscript𝒙𝑖1𝑗subscript𝑡0\bm{x}_{i+1,j}(t_{0})bold_italic_x start_POSTSUBSCRIPT italic_i + 1 , italic_j end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT )𝒙i,j1(t0)subscript𝒙𝑖𝑗1subscript𝑡0\bm{x}_{i,j-1}(t_{0})bold_italic_x start_POSTSUBSCRIPT italic_i , italic_j - 1 end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT )𝒙i,j+1(t0)subscript𝒙𝑖𝑗1subscript𝑡0\bm{x}_{i,j+1}(t_{0})bold_italic_x start_POSTSUBSCRIPT italic_i , italic_j + 1 end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT )𝒙i1,j(t0+Δt)subscript𝒙𝑖1𝑗subscript𝑡0Δ𝑡\bm{x}_{i-1,j}(t_{0}+\Delta t)bold_italic_x start_POSTSUBSCRIPT italic_i - 1 , italic_j end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + roman_Δ italic_t )𝒙i,jsubscript𝒙𝑖𝑗\bm{x}_{i,j}bold_italic_x start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT𝒙i+1,j(t0+Δt)subscript𝒙𝑖1𝑗subscript𝑡0Δ𝑡\bm{x}_{i+1,j}(t_{0}+\Delta t)bold_italic_x start_POSTSUBSCRIPT italic_i + 1 , italic_j end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + roman_Δ italic_t )𝒙i,j1(t0+Δt)subscript𝒙𝑖𝑗1subscript𝑡0Δ𝑡\bm{x}_{i,j-1}(t_{0}+\Delta t)bold_italic_x start_POSTSUBSCRIPT italic_i , italic_j - 1 end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + roman_Δ italic_t )𝒙i,j+1(t0+Δt)subscript𝒙𝑖𝑗1subscript𝑡0Δ𝑡\bm{x}_{i,j+1}(t_{0}+\Delta t)bold_italic_x start_POSTSUBSCRIPT italic_i , italic_j + 1 end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + roman_Δ italic_t )ΔtΔ𝑡\Delta troman_Δ italic_t
Figure 1: Stencil used for the calculation of the FTLE σt0t(𝒙i,j)superscriptsubscript𝜎subscript𝑡0𝑡subscript𝒙𝑖𝑗\sigma_{t_{0}}^{t}(\bm{x}_{i,j})italic_σ start_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT ( bold_italic_x start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT ) at position 𝒙i,jsubscript𝒙𝑖𝑗\bm{x}_{i,j}bold_italic_x start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT and an example of the translation of the particles in the stencil after a time ΔtΔ𝑡\Delta troman_Δ italic_t.

III Flow fields

This section introduces the three flow fields that will serve as test cases for our study. Our choice of two-dimensional velocity fields includes the frequently studied kinematic double gyre and Bickley jet as well as a Faraday flow obtained from experimental measurements.

In all three fields, particle trajectories are integrated from t0=0subscript𝑡00t_{0}=0italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0 until tend=10subscript𝑡𝑒𝑛𝑑10t_{end}=10italic_t start_POSTSUBSCRIPT italic_e italic_n italic_d end_POSTSUBSCRIPT = 10 with a timestep of Δt=0.01Δ𝑡0.01\Delta t=0.01roman_Δ italic_t = 0.01. Particles always start with zero relative velocity, that is 𝒗(t0)=𝒖(𝒙(t0),t0)𝒗subscript𝑡0𝒖𝒙subscript𝑡0subscript𝑡0\bm{v}(t_{0})=\bm{u}(\bm{x}(t_{0}),t_{0})bold_italic_v ( italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) = bold_italic_u ( bold_italic_x ( italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) , italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ).

III.0.1 Double gyre

The double gyre is a popular example in the study of the FTLEs. It is a simplified model of a pattern that occurs in geophysical flows [21]. The velocity flow field is obtained from the streamfunction

ψ(x,y,t)=Asin(πf(x,t))sin(πy),𝜓𝑥𝑦𝑡𝐴𝜋𝑓𝑥𝑡𝜋𝑦\displaystyle\psi(x,y,t)=A\sin(\pi f(x,t))\sin(\pi y),italic_ψ ( italic_x , italic_y , italic_t ) = italic_A roman_sin ( italic_π italic_f ( italic_x , italic_t ) ) roman_sin ( italic_π italic_y ) , (6)

where

f(x,t)𝑓𝑥𝑡\displaystyle f(x,t)italic_f ( italic_x , italic_t ) =a(t)x2+b(t)x,absent𝑎𝑡superscript𝑥2𝑏𝑡𝑥\displaystyle=a(t)x^{2}+b(t)x,= italic_a ( italic_t ) italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_b ( italic_t ) italic_x , (7a)
a(t)𝑎𝑡\displaystyle a(t)italic_a ( italic_t ) =εsin(ωt),absent𝜀𝜔𝑡\displaystyle=\varepsilon\sin(\omega t),= italic_ε roman_sin ( italic_ω italic_t ) , (7b)
b(t)𝑏𝑡\displaystyle b(t)italic_b ( italic_t ) =12εsin(ωt),absent12𝜀𝜔𝑡\displaystyle=1-2\varepsilon\sin(\omega t),= 1 - 2 italic_ε roman_sin ( italic_ω italic_t ) , (7c)

with A=0.1𝐴0.1A=0.1italic_A = 0.1, ε=0.25𝜀0.25\varepsilon=0.25italic_ε = 0.25, ω=π5𝜔𝜋5\omega=\frac{\pi}{5}italic_ω = divide start_ARG italic_π end_ARG start_ARG 5 end_ARG.

We consider 20301203012030120301 particles starting in the spatial domain M=[0,2]×[0,1]𝑀0201M=[0,2]\times[0,1]italic_M = [ 0 , 2 ] × [ 0 , 1 ]. Particles are distributed in 201201201201 vertical lines by 101101101101 horizontal lines, so that adjacent particles are separated by a distance of 0.010.010.010.01. Ideal passive tracers cannot leave the domain M𝑀Mitalic_M but inertial particles can cross the domain walls. Particles leaving the domain boundaries will continue being affected by the flow field, since the double gyre is defined for all point in 2superscript2\mathbb{R}^{2}blackboard_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT.

III.0.2 Bickley Jet

As our second example, we consider the Bickley jet proposed by Rypina et al.  [22]. It is defined by the streamfunction

Ψ(x,y,t)=Ψ𝑥𝑦𝑡absent\displaystyle\Psi(x,y,t)=roman_Ψ ( italic_x , italic_y , italic_t ) = U0Ltanh(yL)+limit-fromsubscript𝑈0𝐿𝑦𝐿\displaystyle-U_{0}L\tanh\left(\frac{y}{L}\right)+- italic_U start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_L roman_tanh ( divide start_ARG italic_y end_ARG start_ARG italic_L end_ARG ) + (8a)
+i=13AiU0Lsech2(yL)cos(kixσit),superscriptsubscript𝑖13subscript𝐴𝑖subscript𝑈0𝐿superscriptsech2𝑦𝐿subscript𝑘𝑖𝑥subscript𝜎𝑖𝑡\displaystyle+\sum_{i=1}^{3}A_{i}U_{0}L\operatorname{sech}^{2}\left(\frac{y}{L% }\right)\cos(k_{i}x-\sigma_{i}t),+ ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_U start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_L roman_sech start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( divide start_ARG italic_y end_ARG start_ARG italic_L end_ARG ) roman_cos ( italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_x - italic_σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_t ) , (8b)

and serves as an idealized model of stratospheric flow. We use same the parameter values as Hadjighasem et al. [47, 48], i.e.  U0=5.414subscript𝑈05.414U_{0}=5.414italic_U start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 5.414, A1=0.0075subscript𝐴10.0075A_{1}=0.0075italic_A start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 0.0075, A2=0.15subscript𝐴20.15A_{2}=0.15italic_A start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 0.15, A3=0.3subscript𝐴30.3A_{3}=0.3italic_A start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = 0.3, L=1.770𝐿1.770L=1.770italic_L = 1.770, c1/U0=0.1446subscript𝑐1subscript𝑈00.1446c_{1}/U_{0}=0.1446italic_c start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT / italic_U start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.1446, c2/U0=0.205subscript𝑐2subscript𝑈00.205c_{2}/U_{0}=0.205italic_c start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT / italic_U start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.205, c3/U0=0.461subscript𝑐3subscript𝑈00.461c_{3}/U_{0}=0.461italic_c start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT / italic_U start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.461, k1=2/resubscript𝑘12subscript𝑟𝑒k_{1}=2/r_{e}italic_k start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 2 / italic_r start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT, k2=4/resubscript𝑘24subscript𝑟𝑒k_{2}=4/r_{e}italic_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 4 / italic_r start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT, k3=6/resubscript𝑘36subscript𝑟𝑒k_{3}=6/r_{e}italic_k start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = 6 / italic_r start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT where re=6.371subscript𝑟𝑒6.371r_{e}=6.371italic_r start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT = 6.371 as well as σi=cikisubscript𝜎𝑖subscript𝑐𝑖subscript𝑘𝑖\sigma_{i}=c_{i}k_{i}italic_σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_c start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, i=1,2,3𝑖123i=1,2,3italic_i = 1 , 2 , 3. For our choice of parameters and when considered on a cylinder, the system exhibits a meandering central jet and three regular vortices on each side of the jet.

In our study, 16281162811628116281 initial conditions are chosen in the rectangle M=[0,20]×[4,4]𝑀02044M=[0,20]\times[-4,4]italic_M = [ 0 , 20 ] × [ - 4 , 4 ]. Particles are distributed within 81818181 horizontal lines by 201201201201 vertical lines, so that again there is a separation of 0.010.010.010.01 among adjacent particles. No periodic boundary conditions are imposed, so particles are allowed to leave the initial domain and disperse over 2superscript2\mathbb{R}^{2}blackboard_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. In contrast to other representations of the Bickley jet, this field does not thus represent a cylinder.

III.0.3 Faraday flow

The Faraday flow field is an experimentally-measured flow field. It is obtained from the surface of a 2 mmtimes2millimeter2\text{\,}\mathrm{mm}start_ARG 2 end_ARG start_ARG times end_ARG start_ARG roman_mm end_ARG-thick layer of distilled water in a cylindrical container of 290 mmtimes290millimeter290\text{\,}\mathrm{mm}start_ARG 290 end_ARG start_ARG times end_ARG start_ARG roman_mm end_ARG diameter as it is vertically shaken with a monochromatic sinusoidal signal at a frequency of 50 Hztimes50hertz50\text{\,}\mathrm{Hz}start_ARG 50 end_ARG start_ARG times end_ARG start_ARG roman_Hz end_ARG at a measured forcing acceleration of a=1.6g𝑎1.6𝑔a=1.6gitalic_a = 1.6 italic_g which corresponds to a supercriticality of ϵ=0.04italic-ϵ0.04\epsilon=0.04italic_ϵ = 0.04  [23]. This movement produces the so-called Faraday waves, quasi-standing waves over the fluid’s surface, and a turbulent space- and time-dependent 2D-velocity field which exhibits all characteristics of two-dimensional turbulence [49]. Energy is injected into the flow at a scale of half the Faraday wavelength which corresponds to the spatial scale of the smallest occurring vortices. In contrast to three-dimensional turbulence, energy is passed from this scale upwards to larger scales and an inverse energy cascade is forming, distributing the energy from the forcing scale up to the system size. Therefore, above half the Faraday wavelength all vortex sizes occur, leading to effective turbulent diffusion of tracers in the flow above the forcing scale [50]. The velocity field is measured using particle image velocimetry (PIV) with floating white, hollow, glass microspheres (diameter of 70μm70𝜇𝑚70\mu m70 italic_μ italic_m, density of 0.15g/cm30.15𝑔𝑐superscript𝑚30.15g/cm^{3}0.15 italic_g / italic_c italic_m start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT, Fibre Glast) to visualize the horizontal motion. The tracer particles are chosen such that they follow the fluid motion as closely as possible. A non-ionic surfactant (1%percent11\%1 % Polysorbate 80) is used in 10%percent1010\%10 % or less solids solution to ensure that the particles do not aggregate and sink. The particles are recorded with a high-speed camera (pco.dimax HS2 and Carl Zeiss Makro-Planar T2.8/100mm𝑇2.8100𝑚𝑚T*2.8/100mmitalic_T ∗ 2.8 / 100 italic_m italic_m lens) and triggered at a frequency of 400 Hztimes400hertz400\text{\,}\mathrm{Hz}start_ARG 400 end_ARG start_ARG times end_ARG start_ARG roman_Hz end_ARG. The details of the experimental setup are similar to those given by Colombi et al. [24, 23]. From the particle images, velocity data is obtained using PIVview (PIVTEC GmbH, Germany) and MATLAB. One set of data for the Faraday flow consists of discrete velocity values for both horizontal and velocity components, vx𝑣𝑥vxitalic_v italic_x and vy𝑣𝑦vyitalic_v italic_y, at 115×8611586115\times 86115 × 86 spatial grid points over an area of roughly 73.90×55.42mm273.9055.42𝑚superscript𝑚273.90\times 55.42mm^{2}73.90 × 55.42 italic_m italic_m start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT as results from a length conversion via calibration. These measurements were taken for 1056105610561056 succesive time steps, resulting in a total measurement time of 42.24 stimes42.24second42.24\text{\,}\mathrm{s}start_ARG 42.24 end_ARG start_ARG times end_ARG start_ARG roman_s end_ARG. When needed, intermediate velocity values are interpolated in space with SciPy’s bivariate rectangular Spline [45]. Interpolation in time is carried out using linear interpolation.

Simulated particle trajectories are computed for 40401404014040140401 particles initialized in the rectangle M[0,0.07]×[0,0.05]𝑀00.0700.05M\approx[0,0.07]\times[0,0.05]italic_M ≈ [ 0 , 0.07 ] × [ 0 , 0.05 ]. Particles are distributed along 201201201201 vertical and 201201201201 horizontal lines and are allowed to leave the initial domain, since no boundary condition is imposed. The flow field’s velocity and its derivatives are set to 00 beyond the boundaries, so that particles behave as relaxing particles after they leave the initial domain and are unable to return.

IV Numerical results

In this section, we analyze the impact of the history term first on the final distribution of particles and then on the FTLE. The flow fields we consider are a double gyre, the Bickley jet and an experimentally measured Faraday flow as described in Section III.

IV.1 Particle clustering

For the double gyre and Bickley jet we show the final positions of all particles, first for ligher-than-fluid and then for heavier-than-fluid particles. We omit clustering plots for the Faraday flow as they do not provide significant additional insight, but they can easily be generated using the accompanying code. Finally, for all three flow fields, we state the relative average difference in the final position of particles computed with and without BHT

d:=1Ni=1N(xi(tend)xinobht(tend)21Nj=1Nxj(tend)xj(t0)2).assign𝑑1𝑁superscriptsubscript𝑖1𝑁subscriptnormsubscript𝑥𝑖subscript𝑡endsuperscriptsubscript𝑥𝑖nobhtsubscript𝑡end21𝑁superscriptsubscript𝑗1𝑁subscriptnormsubscript𝑥𝑗subscript𝑡endsubscript𝑥𝑗subscript𝑡02d:=\frac{1}{N}\sum_{i=1}^{N}\left(\frac{\left\|x_{i}(t_{\text{end}})-x_{i}^{% \text{nobht}}(t_{\text{end}})\right\|_{2}}{\frac{1}{N}\sum_{j=1}^{N}\left\|x_{% j}(t_{\text{end}})-x_{j}(t_{0})\right\|_{2}}\right).italic_d := divide start_ARG 1 end_ARG start_ARG italic_N end_ARG ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT ( divide start_ARG ∥ italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT end end_POSTSUBSCRIPT ) - italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT nobht end_POSTSUPERSCRIPT ( italic_t start_POSTSUBSCRIPT end end_POSTSUBSCRIPT ) ∥ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG start_ARG divide start_ARG 1 end_ARG start_ARG italic_N end_ARG ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT ∥ italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT end end_POSTSUBSCRIPT ) - italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) ∥ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG ) . (9)

in Table 1. Here, N𝑁Nitalic_N is the number of particles, xi(tend)subscript𝑥𝑖subscript𝑡endx_{i}(t_{\text{end}})italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT end end_POSTSUBSCRIPT ) the final position of a particle computed with the full MRE and xinobht(tend)superscriptsubscript𝑥𝑖nobhtsubscript𝑡endx_{i}^{\text{nobht}}(t_{\text{end}})italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT nobht end_POSTSUPERSCRIPT ( italic_t start_POSTSUBSCRIPT end end_POSTSUBSCRIPT ) the final position of the same particle computed without BHT.

The average relative difference in final positions computed with and without BHT for the double gyre and Bickley jet with S=0.1𝑆0.1S=0.1italic_S = 0.1 are less than 10% with standard deviations of similar order. However, for the more turbulent Faraday flow that contains flow structures spanning a large range scales, even for S=0.1𝑆0.1S=0.1italic_S = 0.1 we see a significant effect from the BHT. This can likely be explained by the turbulent characteristics of the Faraday flow which, due to its turbulence in space and the large fluctuations in time, has more separatrices where nearby particles get caught up in very different fates. Further, the lowest spatial scale in the Faraday flow, half the Faraday wavelength λf/25mmsubscript𝜆𝑓25𝑚𝑚\lambda_{f}/2\approx 5mmitalic_λ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT / 2 ≈ 5 italic_m italic_m is smaller when compared to the simulated particle sizes. For instance calculating the size of the simulated particle for S=0.1𝑆0.1S=0.1italic_S = 0.1 via

dp=Sβ92(λf/2)2Refsubscript𝑑𝑝𝑆𝛽92superscriptsubscript𝜆𝑓22𝑅subscript𝑒𝑓d_{p}=\sqrt{S\beta\frac{9}{2}\frac{(\lambda_{f}/2)^{2}}{Re_{f}}}italic_d start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = square-root start_ARG italic_S italic_β divide start_ARG 9 end_ARG start_ARG 2 end_ARG divide start_ARG ( italic_λ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT / 2 ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_R italic_e start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT end_ARG end_ARG (10)

results in dp500μmsubscript𝑑𝑝500𝜇𝑚d_{p}\approx 500\mu mitalic_d start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ≈ 500 italic_μ italic_m to 800μm800𝜇𝑚800\mu m800 italic_μ italic_m for β𝛽\betaitalic_β set to 2/3232/32 / 3 or 4/3434/34 / 3 respectively, which is only a factor of ten away from the injection scale λf/2subscript𝜆𝑓2\lambda_{f}/2italic_λ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT / 2. The approximated Reynolds number of the Faraday flow is Ref100𝑅subscript𝑒𝑓100Re_{f}\approx 100italic_R italic_e start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ≈ 100.

For all three flow fields, for Stokes numbers of S=1𝑆1S=1italic_S = 1 and S=10𝑆10S=10italic_S = 10, the computed final position are substantially different when ignoring the BHT. Furthermore, the standard deviations suggest that there are large differences between particles in how ignoring the BHT affects their trajectories. It is clear that the BHT does not have the same effect on all trajectories but that there is a more complex process at play.

Average relative distance and standard deviation
Flow field S𝑆Sitalic_S R=7/9𝑅79R=7/9italic_R = 7 / 9 R=11/9𝑅119R=11/9italic_R = 11 / 9
Double gyre 0.10.10.10.1 0.03±0.06plus-or-minus0.030.060.03\pm 0.060.03 ± 0.06 0.03±0.05plus-or-minus0.030.050.03\pm 0.050.03 ± 0.05
1111 0.34±0.38plus-or-minus0.340.380.34\pm 0.380.34 ± 0.38 0.53±0.63plus-or-minus0.530.630.53\pm 0.630.53 ± 0.63
10101010 0.65±0.75plus-or-minus0.650.750.65\pm 0.750.65 ± 0.75 0.44±0.56plus-or-minus0.440.560.44\pm 0.560.44 ± 0.56
Bickley jet 0.10.10.10.1 0.06±0.11plus-or-minus0.060.110.06\pm 0.110.06 ± 0.11 0.07±0.13plus-or-minus0.070.130.07\pm 0.130.07 ± 0.13
1111 0.18±0.36plus-or-minus0.180.360.18\pm 0.360.18 ± 0.36 0.14±0.23plus-or-minus0.140.230.14\pm 0.230.14 ± 0.23
10101010 0.35±0.37plus-or-minus0.350.370.35\pm 0.370.35 ± 0.37 0.29±0.33plus-or-minus0.290.330.29\pm 0.330.29 ± 0.33
Faraday flow 0.10.10.10.1 0.29±0.45plus-or-minus0.290.450.29\pm 0.450.29 ± 0.45 0.40±0.52plus-or-minus0.400.520.40\pm 0.520.40 ± 0.52
1111 1.06±0.76plus-or-minus1.060.761.06\pm 0.761.06 ± 0.76 1.20±0.73plus-or-minus1.200.731.20\pm 0.731.20 ± 0.73
10101010 2.88±1.59plus-or-minus2.881.592.88\pm 1.592.88 ± 1.59 2.65±1.45plus-or-minus2.651.452.65\pm 1.452.65 ± 1.45
Table 1: Average relative distance (9) and standard deviation between the final position when computing the trajectories of particles with and without Basset history term.

IV.1.1 Double gyre

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 2: Particle positions at final time tend=10subscript𝑡𝑒𝑛𝑑10t_{end}=10italic_t start_POSTSUBSCRIPT italic_e italic_n italic_d end_POSTSUBSCRIPT = 10 for the double gyre calculated without (top) and with (bottom) BHT. All particles have the same effective density ratio R=7/9𝑅79R=\nicefrac{{7}}{{9}}italic_R = / start_ARG 7 end_ARG start_ARG 9 end_ARG. Stokes numbers increase from left to right: S=0.1𝑆0.1S=0.1italic_S = 0.1 (left), S=1𝑆1S=1italic_S = 1 (center) and S=10𝑆10S=10italic_S = 10 (right).
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 3: Particle positions at final time tend=10subscript𝑡𝑒𝑛𝑑10t_{end}=10italic_t start_POSTSUBSCRIPT italic_e italic_n italic_d end_POSTSUBSCRIPT = 10 for the double gyre calculated without (top) and with (bottom) BHT. All particles have the same effective density ratio R=11/9𝑅119R=\nicefrac{{11}}{{9}}italic_R = / start_ARG 11 end_ARG start_ARG 9 end_ARG. Stokes numbers increase from left to right: S=0.1𝑆0.1S=0.1italic_S = 0.1 (left), S=1𝑆1S=1italic_S = 1 (center) and S=10𝑆10S=10italic_S = 10 (right).

Figure 2 shows the positions of particles in the double gyre at the end of the simulation for Stokes numbers S=0.1𝑆0.1S=0.1italic_S = 0.1 (left), S=1𝑆1S=1italic_S = 1 (middle) and S=10𝑆10S=10italic_S = 10 (right) for particles that are lighter than the fluid with R=7/9𝑅79R=7/9italic_R = 7 / 9. Figure 3 shows the same for denser-than-fluid particles with R=11/9𝑅119R=11/9italic_R = 11 / 9. Upper figures show final positions when computing particle trajectories without history term while lower figures use the full MRE. Table 1 shows the average relative distance (9) between the final position of particles computed with and without BHT.

For the lighter-than-fluid particle with R=7/9𝑅79R=7/9italic_R = 7 / 9 and a small Stokes number of S=0.1𝑆0.1S=0.1italic_S = 0.1, there is little visible impact from the history term. However, for S=1𝑆1S=1italic_S = 1, neglecting the BHT leads to much more pronounced clustering, although the patterns that form are still broadly similar. For S=10𝑆10S=10italic_S = 10, finally, clustering patterns also show significant qualitative differences. The main reason is that ignoring the BHT causes a large number of particles to escape the double gyre (note the different scaling of the y-axis in the right compared to the middle and left figure). This agrees with the findings of Candelier et al. [9], who found that ignoring the BHT when simulating particles in a vortex flow leads to an overestimation of ejection. This could be problematic in many cases, one example being sorting particles by Stokes number using their different separation characteristics [51].

The picture looks similar for the denser-than-fluid particle with R=11/9𝑅119R=11/9italic_R = 11 / 9. For S=0.1𝑆0.1S=0.1italic_S = 0.1, there is little difference between clustering patterns computed with and without BHT. Again, a visible difference emerges for S=1𝑆1S=1italic_S = 1 although for the denser-than-fluid particles the BHT leads to slightly more focused concentration patterns. The “tentacles” formed by esca** particles are much narrower when including the BHT, suggesting again that without BHT more particles are ejected from the domain. For S=10𝑆10S=10italic_S = 10 we also observe the expected accentuated ejection for the denser-than-fluid particles. The patterns formed by the particles are neater and tidier when the BHT is included, in contrast to the more erratic and dispersed patterns obtained when it is ignored. This is also in accordance with observations that neglecting the history terms leads to more complex dynamics[10].

IV.1.2 Bickley jet

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 4: Particle positions at final time tend=10subscript𝑡𝑒𝑛𝑑10t_{end}=10italic_t start_POSTSUBSCRIPT italic_e italic_n italic_d end_POSTSUBSCRIPT = 10 for the Bickley jet calculated without (top) and with (bottom) BHT. All particles have the same effective density ratio R=7/9𝑅79R=\nicefrac{{7}}{{9}}italic_R = / start_ARG 7 end_ARG start_ARG 9 end_ARG. Stokes numbers increase from left to right: S=0.1𝑆0.1S=0.1italic_S = 0.1 (left), S=1𝑆1S=1italic_S = 1 (center) and S=10𝑆10S=10italic_S = 10 (right).
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 5: Particle positions at final time tend=10subscript𝑡𝑒𝑛𝑑10t_{end}=10italic_t start_POSTSUBSCRIPT italic_e italic_n italic_d end_POSTSUBSCRIPT = 10 for the Bickley jet calculated without (top) and with (bottom) BHT. All particles have the same effective density ratio R=11/9𝑅119R=\nicefrac{{11}}{{9}}italic_R = / start_ARG 11 end_ARG start_ARG 9 end_ARG. Stokes numbers increase from left to right: S=0.1𝑆0.1S=0.1italic_S = 0.1 (left), S=1𝑆1S=1italic_S = 1 (center) and S=10𝑆10S=10italic_S = 10 (right).

Figures 4 shows the positions of particles in the Bickley jet at the end of the simulation for Stokes numbers S=0.1𝑆0.1S=0.1italic_S = 0.1 (left), S=1𝑆1S=1italic_S = 1 (middle) and S=10𝑆10S=10italic_S = 10 (right) for particles with R=7/9𝑅79R=7/9italic_R = 7 / 9 that are lighter than the fluid. Figure 5 shows the same for denser-than-fluid particles with R=11/9𝑅119R=11/9italic_R = 11 / 9.

As for the double gyre, the BHT has little impact for S=0.1𝑆0.1S=0.1italic_S = 0.1, both for the lighter- and denser-than-fluid particle. For lighter-than-fluid particles and S=1𝑆1S=1italic_S = 1, neglecting the BHT leads again to stronger clustering in the centers of the vortices. For the denser-than-fluid particles, this effect is reversed and ignoring the BHT leads to stronger ejection. As in the case of the double gyre, we see a lot more particles escape from the central jet into the lower and upper regions when ignoring the BHT (note the changed scaling of the y-axis in the plots for S=10𝑆10S=10italic_S = 10). Again, ignoring the BHT produces more irregular patterns, especially for larger S𝑆Sitalic_S.

IV.2 Finite-time Lyapunov exponents

This section shows colormap plots of the FTLE for all three flow fields. For every flow field, we show two sets of nine plots, each set showing the FTLE results obtained without and with history term as well as the difference between the two respective fields in percent for Stokes numbers S=0.1𝑆0.1S=0.1italic_S = 0.1, S=1𝑆1S=1italic_S = 1, S=10𝑆10S=10italic_S = 10 and density ratios R=7/9𝑅79R=7/9italic_R = 7 / 9 and R=11/9𝑅119R=11/9italic_R = 11 / 9 . The relative difference between FTLEs Δσt0,relativetΔsuperscriptsubscript𝜎subscript𝑡0𝑟𝑒𝑙𝑎𝑡𝑖𝑣𝑒𝑡\Delta\sigma_{t_{0},relative}^{t}roman_Δ italic_σ start_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_r italic_e italic_l italic_a italic_t italic_i italic_v italic_e end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT is calculated from

Δσt0,relativet:=100σt0,Historytσt0,Stokestσt0,Historytl(D)assignΔsuperscriptsubscript𝜎subscript𝑡0𝑟𝑒𝑙𝑎𝑡𝑖𝑣𝑒𝑡100superscriptsubscript𝜎subscript𝑡0𝐻𝑖𝑠𝑡𝑜𝑟𝑦𝑡superscriptsubscript𝜎subscript𝑡0𝑆𝑡𝑜𝑘𝑒𝑠𝑡subscriptnormsuperscriptsubscript𝜎subscript𝑡0𝐻𝑖𝑠𝑡𝑜𝑟𝑦𝑡subscript𝑙𝐷\displaystyle\Delta\sigma_{t_{0},relative}^{t}:=100\;\frac{\sigma_{t_{0},% History}^{t}-\sigma_{t_{0},Stokes}^{t}}{||\sigma_{t_{0},History}^{t}||_{l_{% \infty}(D)}}roman_Δ italic_σ start_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_r italic_e italic_l italic_a italic_t italic_i italic_v italic_e end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT := 100 divide start_ARG italic_σ start_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_H italic_i italic_s italic_t italic_o italic_r italic_y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT - italic_σ start_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_S italic_t italic_o italic_k italic_e italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT end_ARG start_ARG | | italic_σ start_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_H italic_i italic_s italic_t italic_o italic_r italic_y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT | | start_POSTSUBSCRIPT italic_l start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT ( italic_D ) end_POSTSUBSCRIPT end_ARG (11)

where DΩ×[0,T]𝐷Ω0𝑇D\subset\Omega\times[0,T]italic_D ⊂ roman_Ω × [ 0 , italic_T ] and Ω2Ωsuperscript2\Omega\subseteq\mathbb{R}^{2}roman_Ω ⊆ blackboard_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT is the spatial domain. The differences are signed and negative values correspond to areas where neglecting the BHT will increase FTLE while positive areas mean that neglecting the BHT will lead to smaller FTLE.

For R=1𝑅1R=1italic_R = 1 the FTLE fields are independent of S𝑆Sitalic_S and, since we do not consider any effects that induce non-zero relative velocities, they agree independent of whether the BHT is present or not. We confirmed that our numerical solutions reproduce this behavior correctly but do not show any figures.

IV.2.1 Double gyre

Figure 6 shows the FTLE without BHT (left), with BHT (center) and the difference between the two (right) for lighter-than-fluid particles with R=7/9𝑅79R=7/9italic_R = 7 / 9. From top to bottom the Stokes number changes from S=0.1𝑆0.1S=0.1italic_S = 0.1 to S=1𝑆1S=1italic_S = 1 and S=10𝑆10S=10italic_S = 10.

In line with the clustering plots, the impact of the BHT for S=0.1𝑆0.1S=0.1italic_S = 0.1 is small. The FTLE show no visible differences and agree to within a few percent. For S=1𝑆1S=1italic_S = 1, the FTLE still look qualitatively similar although some differences emerge. In particular, the region of strong divergence between the gyres shifts it position. When the Stokes number becomes even larger, at S=10𝑆10S=10italic_S = 10, we start to see substantial qualitative differences as well. Without BHT, a number of filaments of strong divergence appear that are not present in the FTLE with history term. There are now substantial differences between the two FTLE fields of up to ±100plus-or-minus100\pm 100± 100% in some regions.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 6: FTLEs for light particles (R=7/9𝑅79R=\nicefrac{{7}}{{9}}italic_R = / start_ARG 7 end_ARG start_ARG 9 end_ARG) and three Stokes numbers, S=0.1𝑆0.1S=0.1italic_S = 0.1 (top), S=1𝑆1S=1italic_S = 1 (center), S=10𝑆10S=10italic_S = 10 (bottom), moving in the Double Gyre with periodic Boundary conditions. Left and center columns show the FTLEs computed with trajectories calculated without (left) and with (right) BHT. Right column shows the relative difference in the FTLE.

Figure 7 shows the same plots but for heavier-than-fluid particles with R=11/9𝑅119R=11/9italic_R = 11 / 9. As for the light particle, there is little difference between the FTLE for S=0.1𝑆0.1S=0.1italic_S = 0.1. For S=1𝑆1S=1italic_S = 1, however, the differences between BHT and no BHT are more pronounced. While still qualitatively similar, the regions of strong separation between and around the gyres look noticeably different. Without BHT, there are three well separated filaments of strong divergence between the gyres which are much closer together and almost fused when the BHT is used. More generally, the red, high-FTLE filaments around the gyres shift away from the center of the domain when including the history term.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 7: FTLEs for heavy particles (R=11/9𝑅119R=\nicefrac{{11}}{{9}}italic_R = / start_ARG 11 end_ARG start_ARG 9 end_ARG) and three Stokes numbers, S=0.1𝑆0.1S=0.1italic_S = 0.1 (top), S=1𝑆1S=1italic_S = 1 (center), S=10𝑆10S=10italic_S = 10 (bottom), moving in the Double Gyre with periodic Boundary conditions. Left and center columns show the FTLEs computed with trajectories calculated without (left) and with (right) BHT. Right column shows the relative difference in the FTLE.

IV.2.2 Bickley jet

Figures 8 and 9 shows the FTLE (left and center) and difference between FTLE with and without BHT (right) for the Bickley jet. Again, for both lighter- and heavier-than-fluid particles, the differences for S=0.1𝑆0.1S=0.1italic_S = 0.1 are mostly small, although a few localized regions emerge where shifts in the position of divergence zones around the vortices lead to differences of over 50505050%.

This effect becomes more pronounced for S=1𝑆1S=1italic_S = 1 where the FTLE in the jet are now showing differences up to 100100100100%. For lighter-than-fluid particles, knot-like objects structures emerge in regions of high curvature within the jet for the case S=1𝑆1S=1italic_S = 1 without BHT. While this looks like a numerical artifact on first sight, it can be well explained by the fact that particles initialized in the jet may leave it due to the inertial effects and may enter the neighborhood of one of surrounding vortices. This is also visible in the corresponding final particle positions: compare the central panels in Figure 4. Note that the high-FTLE structures for S=1𝑆1S=1italic_S = 1 computed without BHT are similar those computed for the higher Stokes number S=10𝑆10S=10italic_S = 10 with BHT.

Again, for S=10𝑆10S=10italic_S = 10, the differences increase and ignoring the BHT leads to a strong overestimation of FTLE of up to 100100100100% in some parts. Even with BHT we now see the emergence of knot-like structures in the jet that were already observed for S=1𝑆1S=1italic_S = 1 without BHT.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 8: FTLEs for light particles (R=7/9𝑅79R=\nicefrac{{7}}{{9}}italic_R = / start_ARG 7 end_ARG start_ARG 9 end_ARG) and three Stokes numbers, S=0.1𝑆0.1S=0.1italic_S = 0.1 (top), S=1𝑆1S=1italic_S = 1 (center), S=10𝑆10S=10italic_S = 10 (bottom), moving in the Bickley Jet. Left and center columns show the FTLEs computed with trajectories calculated without (left) and with (right) BHT. Right column shows the relative difference in the FTLE.
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 9: FTLEs for heavy particles (R=11/9𝑅119R=\nicefrac{{11}}{{9}}italic_R = / start_ARG 11 end_ARG start_ARG 9 end_ARG) and three Stokes numbers, S=0.1𝑆0.1S=0.1italic_S = 0.1 (top), S=1𝑆1S=1italic_S = 1 (center), S=10𝑆10S=10italic_S = 10 (bottom), moving in the Bickley Jet. Left and center columns show the FTLEs computed with trajectories calculated without (left) and with (right) BHT. Right column shows the relative difference in the FTLE.

IV.2.3 Faraday flow

Finally, figures 10 and 11 show the FTLE and differences between FTLE for the Faraday flow. For both R=7/9𝑅79R=7/9italic_R = 7 / 9 and R=11/9𝑅119R=11/9italic_R = 11 / 9 we see locally concentrated significant quantitative differences already for S=0.1𝑆0.1S=0.1italic_S = 0.1. The effect is more pronounced for the heavier-than-fluid particles where substantial regions with differences in the FTLE between 50505050 and 100100100100% arise. Table 1 substantiates this observations, showing that, even on average, final particle positions are very different with and without BHT.

For S=1𝑆1S=1italic_S = 1 and both lighter and heavier particles, there are much larger zones with high, positive FTLEs without BHT than with. The effect is even stronger for S=10𝑆10S=10italic_S = 10 where throughout most of the domain the FTLE without BHT are larger than +55+5+ 5 whereas with BHT significant areas with smaller FTLE of around 23232-32 - 3 remain. In both cases, the FTLE are overestimated by up to 100100100100% when ignoring the BHT. This suggests that ignoring the BHT may lead to unrealistically effective mixing in simulations, because particles trajectories diverge too rapidly.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 10: FTLEs for light particles (R=7/9𝑅79R=\nicefrac{{7}}{{9}}italic_R = / start_ARG 7 end_ARG start_ARG 9 end_ARG) and three Stokes numbers, S=0.1𝑆0.1S=0.1italic_S = 0.1 (top), S=1𝑆1S=1italic_S = 1 (center), S=10𝑆10S=10italic_S = 10 (bottom), moving in the Faraday flow. Left and center columns show the FTLEs computed with trajectories calculated without (left) and with (right) BHT. Right column shows the relative difference in the FTLE.
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 11: FTLEs for heavy particles (R=11/9𝑅119R=\nicefrac{{11}}{{9}}italic_R = / start_ARG 11 end_ARG start_ARG 9 end_ARG) and three Stokes numbers, S=0.1𝑆0.1S=0.1italic_S = 0.1 (top), S=1𝑆1S=1italic_S = 1 (center), S=10𝑆10S=10italic_S = 10 (bottom), moving in the Faraday Flow. Left and center columns show the FTLEs computed with trajectories calculated without (left) and with (right) BHT. Right column shows the relative difference in the FTLE.

V Conclusions

We investigate how the Basset history term (BHT) in the Maxey-Riley equation changes the Lagrangian dynamics of simulated inertial particles. To this end, we solve the nondimensional Maxey-Riley equation with and without BHT for thousands of particles in three flow fields, a double gyre, the Bickley jet and an experimentally measured Faraday flow. We compare the clustering of the particles and finite-time Lyapunov exponents for trajectories computed with and without BHT.

For the double gyre and Bickley jet and a Stokes number of S=0.1𝑆0.1S=0.1italic_S = 0.1 we see little difference between the dynamics computed with and without BHT. Since the Faraday flow is more turbulent, even for S=0.1𝑆0.1S=0.1italic_S = 0.1 ignoring the BHT already has a visible impact on the FTLE field. For S=1𝑆1S=1italic_S = 1 and even more so for S=10𝑆10S=10italic_S = 10, these significant difference also emerge for the double gyre and the Bickley jet. In line with previous findings, ignoring the BHT leads to an overestimation of ejection of particles from vortices in the flow. It also shifts the positions of areas of strong divergence, even when the overall patterns still look similar.

Generally, ignoring the BHT overestimates FTLE and leads to Lagrangian dynamics that would only be seen for a larger Stokes number if the BHT was considered. This also makes sense mathematically: with the BHT, the full history of the particle influences the forces acting on a particle at some given time. Without BHT, only the instantaneous forces due to the material derivative of the flow field and Stokes drag act on the particle. Therefore, without BHT, the forces from one time step to the next will change more drastically whereas the BHT adds a form of “mathematical inertia” since the integral over the particle’s past trajectory changes more slowly and the acting forces and thus the particle’s speed and direction will change less rapidly.

In conclusion, our analysis backs up previous studies [9, 10, 6] that demonstrate that the Basset history term cannot safely be ignored even when simulating comparatively small particles. It confirms that the differences matter not only at the level of individual particle trajectories but that also the larger scale Lagrangian dynamics change potentially significantly if the BHT is neglected. In particular, the overestimation of FTLE could lead to unrealistically efficient mixing in simulations when the BHT is ignored. Given that recent advances in numerical mathematics now allow for the efficient solution of the full MRE [42, 6, 19], we argue that simulations of inertial particles should routinely include the history term.

Acknowledgements.
This project is funded by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) – SFB 1615 – 503850735.

Data Availability Statement

AVAILABILITY OF DATA STATEMENT OF DATA AVAILABILITY
Data openly available in a public repository that issues datasets with DOIs The data that support the findings of this study are openly available at https://doi.org/10.5281/zenodo.12557890.

References

  • Maxey and Dent [2017] M. R. Maxey and G. L. Dent, in Collective Dynamics of Particles, Vol. 576, edited by C. Marchioli (Springer International Publishing, Cham, 2017) pp. 1–38, series Title: CISM International Centre for Mechanical Sciences.
  • Maxey and Riley [1983] M. R. Maxey and J. J. Riley, The Physics of Fluids 26, 883 (1983).
  • Costa et al. [2020] P. Costa, L. Brandt, and F. Picano, Journal of Fluid Mechanics 891, E2 (2020).
  • Michaelides [1992] E. E. Michaelides, Physics of Fluids A: Fluid Dynamics 4, 1579 (1992).
  • Farazmand and Haller [2015] M. Farazmand and G. Haller, Nonlinear Analysis: Real World Applications 22, 98 (2015).
  • Prasath et al. [2019] S. G. Prasath, V. Vasan, and R. Govindarajan, Journal of Fluid Mechanics 868, 428 (2019).
  • Olivieri et al. [2014] S. Olivieri, F. Picano, G. Sardina, D. Iudicone, and L. Brandt, Physics of Fluids 26, 041704 (2014).
  • Daitche and Tél [2014] A. Daitche and T. Tél, New Journal of Physics 16, 073008 (2014).
  • Candelier et al. [2004] F. Candelier, J. R. Angilella, and M. Souhar, Physics of Fluids 16, 1765 (2004).
  • Guseva et al. [2013] K. Guseva, U. Feudel, and T. Tél, Phys. Rev. E 88, 042909 (2013).
  • Haller [2015] G. Haller, Annual Review of Fluid Mechanics 47, 137 (2015).
  • Allshouse and Peacock [2015] M. R. Allshouse and T. Peacock, Chaos 25, 097617 (2015).
  • Hadjighasem et al. [2017] A. Hadjighasem, M. Farazmand, D. Blazevski, G. Froyland, and G. Haller, Chaos 27, 053104 (2017).
  • Badza et al. [2023] A. Badza, T. W. Mattner, and S. Balasuriya, Physica D: Nonlinear Phenomena 444, 133580 (2023).
  • Sapsis and Haller [2010] T. Sapsis and G. Haller, Chaos: An Interdisciplinary Journal of Nonlinear Science 20, 017515 (2010).
  • Sudharsan et al. [2016] M. Sudharsan, S. L. Brunton, and J. J. Riley, Phys. Rev. E 93, 033108 (2016).
  • Garaboa-Paz and Pérez-Muñuzuri [2015] D. Garaboa-Paz and V. Pérez-Muñuzuri, Nonlinear Processes in Geophysics 22, 571 (2015).
  • Günther and Theisel [2017] T. Günther and H. Theisel, IEEE Transactions on Visualization and Computer Graphics 23, 970 (2017).
  • Urizarna-Carasa et al. [2024] J. Urizarna-Carasa, L. Schlegel, and D. Ruprecht, arXiv preprint arXiv:2403.13515  (2024).
  • Haller [2019] G. Haller, Journal of Fluid Mechanics 874, 1–4 (2019).
  • Shadden et al. [2005] S. C. Shadden, F. Lekien, and J. E. Marsden, Physica D: Nonlinear Phenomena 212, 271 (2005).
  • Rypina et al. [2007] I. Rypina, M. G. Brown, F. J. Beron-Vera, H. Koçak, M. J. Olascoaga, and I. Udovydchenkov, Journal of the Atmospheric Sciences 64, 3595 (2007).
  • Colombi et al. [2021] R. Colombi, M. Schlüter, and A. v. Kameke, Experiments in fluids 62, 1 (2021).
  • Colombi et al. [2022] R. Colombi, N. Rohde, M. Schlüter, and A. von Kameke, Fluids 7, 148 (2022).
  • Beron-Vera et al. [2019] F. J. Beron-Vera, M. J. Olascoaga, and P. Miron, Physics of Fluids 31, 096602 (2019).
  • Huilier [2021] D. G. F. Huilier, Fluids 610.3390/fluids6040145 (2021).
  • Daitche [2015] A. Daitche, Journal of Fluid Mechanics 782, 567–593 (2015).
  • Bourgoin [2017] M. Bourgoin, Some aspects of the collective dynamics of particles in turbulent flows, in Collective Dynamics of Particles: From Viscous to Turbulent Flows, edited by C. Marchioli (Springer International Publishing, Cham, 2017) p. 67–97.
  • Ouellette et al. [2008] N. T. Ouellette, P. J. J. O’Malley, and J. P. Gollub, Phys. Rev. Lett. 101, 174504 (2008).
  • Hofmann et al. [2022] S. Hofmann, C. Weiland, J. Fitschen, A. von Kameke, M. Hoffmann, and M. Schlüter, Chemical Engineering Journal 449, 137549 (2022).
  • Langlois et al. [2015] G. P. Langlois, M. Farazmand, and G. Haller, Journal of nonlinear science 25, 1225 (2015).
  • Chong et al. [2013] K. Chong, S. D. Kelly, S. Smith, and J. D. Eldredge, Physics of Fluids 25, 033602 (2013).
  • Cummins et al. [2020] C. P. Cummins, O. J. Ajayi, F. V. Mehendale, R. Gabl, and I. M. Viola, Physics of Fluids 32, 083302 (2020).
  • Lovalenti and Brady [1993] P. M. Lovalenti and J. F. Brady, Journal of Fluid Mechanics 256, 561 (1993).
  • Mei [1994] R. Mei, Journal of Fluid Mechanics 270, 133 (1994).
  • Dorgan and Loth [2007] A. Dorgan and E. Loth, International journal of multiphase flow 33, 833 (2007).
  • Moreno-Casas and Bombardelli [2016] P. A. Moreno-Casas and F. A. Bombardelli, Environmental Fluid Mechanics 16, 193 (2016).
  • Klinkenberg et al. [2014] J. Klinkenberg, H. De Lange, and L. Brandt, Meccanica 49, 811 (2014).
  • Elghannay and Tafti [2016] H. A. Elghannay and D. K. Tafti, International Journal of Multiphase Flow 85, 284 (2016).
  • Parmar et al. [2018] M. Parmar, S. Annamalai, S. Balachandar, and A. Prosperetti, Journal of fluid mechanics 844, 970 (2018).
  • Van Hinsberg et al. [2011] M. Van Hinsberg, J. ten Thije Boonkkamp, and H. J. Clercx, Journal of Computational Physics 230, 1465 (2011).
  • Daitche [2013] A. Daitche, Journal of Computational Physics 254, 93 (2013).
  • Koleva [2005] M. N. Koleva, in International Conference on Large-Scale Scientific Computing (Springer, 2005) pp. 509–517.
  • Ascher et al. [1997] U. M. Ascher, S. J. Ruuth, and R. J. Spiteri, Applied Numerical Mathematics 25, 151 (1997).
  • Virtanen et al. [2020] P. Virtanen, R. Gommers, T. E. Oliphant, M. Haberland, T. Reddy, D. Cournapeau, E. Burovski, P. Peterson, W. Weckesser, J. Bright, S. J. van der Walt, M. Brett, J. Wilson, K. J. Millman, N. Mayorov, A. R. J. Nelson, E. Jones, R. Kern, E. Larson, C. J. Carey, İ. Polat, Y. Feng, E. W. Moore, J. VanderPlas, D. Laxalde, J. Perktold, R. Cimrman, I. Henriksen, E. A. Quintero, C. R. Harris, A. M. Archibald, A. H. Ribeiro, F. Pedregosa, P. van Mulbregt, and SciPy 1.0 Contributors, Nature Methods 17, 261 (2020).
  • Urizarna-Carasa [2024] J. Urizarna-Carasa, JulioUri/Basset-History-Term-Influence: Repo ready for submission (2024), https://doi.org/10.5281/zenodo.12557890.
  • Hadjighasem et al. [2016] A. Hadjighasem, D. Karrasch, H. Teramoto, and G. Haller, Physical Review E 93, 063107 (2016).
  • Padberg-Gehle and Schneide [2017] K. Padberg-Gehle and C. Schneide, Nonlinear Processes in Geophysics 24, 661 (2017).
  • von Kameke et al. [2011] A. von Kameke, F. Huhn, G. Fernández-García, A. P. Muñuzuri, and V. Pérez-Muñuzuri, Phys. Rev. Lett. 107, 074502 (2011).
  • von Kameke et al. [2013] A. von Kameke, F. Huhn, A. P. Muñuzuri, and V. Pérez-Muñuzuri, Phys. Rev. Lett. 110, 088302 (2013).
  • Tallapragada and Ross [2008] P. Tallapragada and S. D. Ross, Phys. Rev. E 78, 036308 (2008).