Mass crystals and cage effect in vorticity crystals

Jean-Rรฉgis Angilella111email: Jean-Regis.Angilella@unicaen.fr Normandie Universitรฉ, UNICAEN, UNIROUEN, ABTE, ESIX Cherbourg, Caen 14000, France.
Abstract

We study the motion of tiny heavy inertial particles advected by a two dimensional inviscid fluid flow composed of N๐‘Nitalic_N identical point vortices regularly placed on a ring, and forming a crystal. In the limit of weak particle inertia, we show asymptotically that, in the reference frame of the crystal, inertial particles have N๐‘Nitalic_N asymptotically stable equilibrium positions located outside the crystal, in agreement with numerical observations by Ravichandran et al. (Sฤdhanฤ 42, 2017). In addition to these โ€satelliteโ€ attracting points, we observe that for Nโ‰ฅ3๐‘3N\geq 3italic_N โ‰ฅ 3 the center of the ring, though degenerate, is a stable equilibrium position for inertial particles. This creates a kind of cage effect, where inclusions slowly drift towards the center under the effect of the surrounding vortices. This cage effect is observed to persist even at larger Stokes numbers, in contrast with the satellite attracting points that vanish when the Stokes number is above some critical value.

inertial particles ; inviscid fluids ; vortical flows

Vortex crystals are two-dimensional flow structures appearing in inviscid fluids, where point vortices place themselves naturally at the tips of polygonal structures that rotate as a solid body Aref et al. (2002). They were observed experimentally four decades ago in superfluid Helium Yarmchuk et al. (1979), then in pure electron columns (Fine et al., 1995; Durkin and Fajans, 2000a, b), floating disks (Grzybowski et al., 2000; Grzybowski and Whitesides, 2002), and more recently on the poles of Jupiter Adriani et al. (2018). Both experiments and simulations showed that crystals can emerge from random vortex distributions and persist even in the presence of weak viscous effects Schecter et al. (1999); Jin and Dubin (2000); Siegelman et al. (2022). Shortly after the initial disordered state, vortices advect each other in a chaotic manner, and produce vortex layers creating a diffuse background Schecter et al. (1999). Successive vortex mergers lead to intense vortices which eventually converge towards a stable configuration in rigid rotation.

The influence of this self-organization on the transport of inertial particles suspended in the fluid has received little attention so far. However, it can be conjectured that the emergence of such flow structures will certainly affect the distribution of particles, as these objects are known to be very sensitive to the distribution of vorticity. In strong contrast with turbulent flows however, vortex crystals induce steady flows in the reference frame rotating with the crystal. In the case where two point vortices rotate around each other, their centres forming a rigid segment, it has been shown that heavy inertial particles could be trapped by attracting points at rest in the frame of the segment Angilella (2010); Ravichandran et al. (2014). This property has been observed to persist in the case where more than two vortices interact to form a crystal Ravichandran et al. (2017). The goal of the present work is to address this question theoretically, in the case of heavy inertial particles transported at low particle Reynolds number in an inviscid potential flow composed of identical vortices regularly placed on a ring. It will be shown that equilibrium points always exist in the frame of the crystal for inertial particles, in the limit of weak inertia, and that these points are hyperbolic (i.e. the real parts of the eigenvalues of the particle flow gradient are non-zero there). Some of these points will be shown to be asymptotically stable, so that the particle cloud at long time takes the form of a mass crystal at rest in the reference frame of the vortex crystal. These trapping points being located away from the center of vorticity, they will be referred to as satellite points. In addition, in the present case where no vortex is placed at the center of vorticity, we will show that this point, though degenerate (it is a center point with pure imaginary eigenvalues), is asymptotically stable and can attract a significant number of particles. Inclusions trapped there will then be surrounded by the vortices and maintained within a kind of cage. In the laboratory frame, these particles will appear as almost motionless at long times, while vortices rotate around them. Both the satellite attracting points and the cage effect will be studied asymptotically, in the limit of small Stokes numbers. We will then show some numerical simulations, involving up to seven vortices, which confirm theoretical predictions.

Prior studying the motion of inclusions in vortex crystals, we recall some basic considerations about equilibrium positions of inertial particles (Babiano et al., 2000; Haller and Sapsis, 2008; Cartwright et al., 2010). We assume particles are much heavier than the fluid, do not interact, and have a small particle Reynolds number. Under these conditions, the non-dimensional motion equation of an inertial particle with time-dependent position ๐—โข(t)๐—๐‘ก\mathbf{X}(t)bold_X ( italic_t ), transported in a steady flow with velocity field ๐ฎโข(๐—)๐ฎ๐—\mathbf{u(X)}bold_u ( bold_X ), assuming a linear drag and neglecting the added mass and Boussinesq-Basset forces, is of the form:

๐—ยจ=1Sโขtโข(๐ฎโข(๐—)โˆ’๐—ห™)+๐…โข(๐—,๐—ห™)ยจ๐—1๐‘†๐‘ก๐ฎ๐—ห™๐—๐…๐—ห™๐—\ddot{\mathbf{X}}=\frac{1}{St}(\mathbf{u}(\mathbf{X})-\dot{\mathbf{X}})+% \mathbf{F}(\mathbf{X},\dot{\mathbf{X}})overยจ start_ARG bold_X end_ARG = divide start_ARG 1 end_ARG start_ARG italic_S italic_t end_ARG ( bold_u ( bold_X ) - overห™ start_ARG bold_X end_ARG ) + bold_F ( bold_X , overห™ start_ARG bold_X end_ARG ) (1)

where Sโขt๐‘†๐‘กStitalic_S italic_t is the Stokes number, that is the non-dimensional response time of the inclusion. In the case of spherical particles with mass mpsubscript๐‘š๐‘m_{p}italic_m start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT and radius rpsubscript๐‘Ÿ๐‘r_{p}italic_r start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT, the Stokes number reads Sโขt=ฮฉ0โขmp/(6โขฯ€โขฮผโขrp)๐‘†๐‘กsubscriptฮฉ0subscript๐‘š๐‘6๐œ‹๐œ‡subscript๐‘Ÿ๐‘St=\Omega_{0}m_{p}/(6\pi\mu\,r_{p})italic_S italic_t = roman_ฮฉ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_m start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT / ( 6 italic_ฯ€ italic_ฮผ italic_r start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ), where ฮผ๐œ‡\muitalic_ฮผ is the dynamic viscosity of the fluid and 1/ฮฉ01subscriptฮฉ01/\Omega_{0}1 / roman_ฮฉ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is a characteristic flow time which will be defined below.

The term ๐…๐…\mathbf{F}bold_F contains various terms corresponding to body forces, hydrodynamic forces or fictitious forces (if the reference frame is non-inertial). Equilibrium positions, if any, correspond to

๐šฝโข(๐—,Sโขt)โ‰ก๐ฎโข(๐—)+Sโขtโข๐…โข(๐—,๐ŸŽ)=๐ŸŽ.๐šฝ๐—๐‘†๐‘ก๐ฎ๐—๐‘†๐‘ก๐…๐—00\mathbf{\Phi}(\mathbf{X},St)\equiv\mathbf{u}(\mathbf{X})+St\,\mathbf{F}(% \mathbf{X},{\mathbf{0}})={\mathbf{0}}.bold_ฮฆ ( bold_X , italic_S italic_t ) โ‰ก bold_u ( bold_X ) + italic_S italic_t bold_F ( bold_X , bold_0 ) = bold_0 . (2)

We assume that the fluid flow has a stagnation point at some position ๐—eโขq0superscriptsubscript๐—๐‘’๐‘ž0\mathbf{X}_{eq}^{0}bold_X start_POSTSUBSCRIPT italic_e italic_q end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT: ๐ฎโข(๐—eโขq0)=๐ŸŽ๐ฎsuperscriptsubscript๐—๐‘’๐‘ž00\mathbf{u}(\mathbf{X}_{eq}^{0})=\mathbf{0}bold_u ( bold_X start_POSTSUBSCRIPT italic_e italic_q end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT ) = bold_0. Then, by virtue of the implicit functions theorem, if the gradient of ๐šฝ๐šฝ\mathbf{\Phi}bold_ฮฆ at (๐—eโขq0,0)superscriptsubscript๐—๐‘’๐‘ž00(\mathbf{X}_{eq}^{0},0)( bold_X start_POSTSUBSCRIPT italic_e italic_q end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT , 0 ) is invertible, that is if the gradient of ๐ฎ๐ฎ\mathbf{u}bold_u at ๐—eโขq0superscriptsubscript๐—๐‘’๐‘ž0\mathbf{X}_{eq}^{0}bold_X start_POSTSUBSCRIPT italic_e italic_q end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT is invertible, there exists a continuously differentiable function ๐—eโขqโข(Sโขt)subscript๐—๐‘’๐‘ž๐‘†๐‘ก\mathbf{X}_{eq}(St)bold_X start_POSTSUBSCRIPT italic_e italic_q end_POSTSUBSCRIPT ( italic_S italic_t ), defined for Sโขt๐‘†๐‘กStitalic_S italic_t in the vicinity of 0, and satisfying ๐šฝโข(๐—eโขqโข(Sโขt),Sโขt)=๐ŸŽ๐šฝsubscript๐—๐‘’๐‘ž๐‘†๐‘ก๐‘†๐‘ก0\mathbf{\Phi}(\mathbf{X}_{eq}(St),St)=\mathbf{0}bold_ฮฆ ( bold_X start_POSTSUBSCRIPT italic_e italic_q end_POSTSUBSCRIPT ( italic_S italic_t ) , italic_S italic_t ) = bold_0, and ๐—eโขqโข(0)=๐—eโขq0subscript๐—๐‘’๐‘ž0superscriptsubscript๐—๐‘’๐‘ž0\mathbf{X}_{eq}(0)=\mathbf{X}_{eq}^{0}bold_X start_POSTSUBSCRIPT italic_e italic_q end_POSTSUBSCRIPT ( 0 ) = bold_X start_POSTSUBSCRIPT italic_e italic_q end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT. In other words, a single-valued continuous equilibrium branch persists, inertial particles have equilibrium positions that depend continuously on Sโขt๐‘†๐‘กStitalic_S italic_t, and which coincides with ๐—eโขq0superscriptsubscript๐—๐‘’๐‘ž0\mathbf{X}_{eq}^{0}bold_X start_POSTSUBSCRIPT italic_e italic_q end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT in the limit of small inertia. This shows that equilibrium positions of inertial particles exist in the vicinity of any fluid equilibrium position with non-zero eigenvalues, in the limit of small Stokes numbers. In the following, we will study the position and stability of these points in the case of a vortex crystal.

We consider the case where the flow is a vortex crystal composed of N๐‘Nitalic_N identical point vortices located at ๐ฑn=aโข(cosโก(nโข2โขฯ€/N),sinโก(nโข2โขฯ€/N))subscript๐ฑ๐‘›๐‘Ž๐‘›2๐œ‹๐‘๐‘›2๐œ‹๐‘\mathbf{x}_{n}=a(\cos(n2\pi/N),\sin(n2\pi/N))bold_x start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = italic_a ( roman_cos ( italic_n 2 italic_ฯ€ / italic_N ) , roman_sin ( italic_n 2 italic_ฯ€ / italic_N ) ), with n=0,..,Nโˆ’1n=0,..,N-1italic_n = 0 , . . , italic_N - 1. We will limit ourselves to the case Nโ‰ค7๐‘7N\leq 7italic_N โ‰ค 7 (Thomsonโ€™s heptagon), as a larger number of vortices placed on a ring might lead to unstable flows Mertz (1978). These vortices form a rigid body (a regular polygon with radius a๐‘Žaitalic_a) that rotates with respect to the laboratory frame, with a constant angular velocity ฮฉ0=(Nโˆ’1)โขฮ“/(4โขฯ€โขa2)subscriptฮฉ0๐‘1ฮ“4๐œ‹superscript๐‘Ž2\Omega_{0}=(N-1)\,{\Gamma}/{(4\pi a^{2})}roman_ฮฉ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = ( italic_N - 1 ) roman_ฮ“ / ( 4 italic_ฯ€ italic_a start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ), where ฮ“ฮ“\Gammaroman_ฮ“ is the strength of vortices. Without loss of generality we assume ฮ“>0ฮ“0\Gamma>0roman_ฮ“ > 0, so that the crystal rotates in the counter-clockwise direction (ฮฉ0>0subscriptฮฉ00\Omega_{0}>0roman_ฮฉ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT > 0). In the frame rotating with vortices, the flow is steady and its velocity field is given by ๐ฎ=(โˆ‚ฯˆ/โˆ‚y,โˆ’โˆ‚ฯˆ/โˆ‚x)๐ฎ๐œ“๐‘ฆ๐œ“๐‘ฅ\mathbf{u}=(\partial\psi/\partial y,-\partial\psi/\partial x)bold_u = ( โˆ‚ italic_ฯˆ / โˆ‚ italic_y , - โˆ‚ italic_ฯˆ / โˆ‚ italic_x ), where the streamfunction ฯˆ๐œ“\psiitalic_ฯˆ reads:

ฯˆ=12โขฮฉ0โข|๐ฑ|2โˆ’ฮ“4โขฯ€โขโˆ‘nlnโก(|๐ฑโˆ’๐ฑn|2)๐œ“12subscriptฮฉ0superscript๐ฑ2ฮ“4๐œ‹subscript๐‘›superscript๐ฑsubscript๐ฑ๐‘›2\psi=\frac{1}{2}\Omega_{0}|\mathbf{x}|^{2}-\frac{\Gamma}{4\pi}\sum_{n}\ln(|% \mathbf{x}-\mathbf{x}_{n}|^{2})italic_ฯˆ = divide start_ARG 1 end_ARG start_ARG 2 end_ARG roman_ฮฉ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT | bold_x | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - divide start_ARG roman_ฮ“ end_ARG start_ARG 4 italic_ฯ€ end_ARG โˆ‘ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT roman_ln ( | bold_x - bold_x start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) (3)

with ๐ฑ=(x,y)๐ฑ๐‘ฅ๐‘ฆ\mathbf{x}=(x,y)bold_x = ( italic_x , italic_y ). The corresponding streamlines for N=๐‘absentN=italic_N = 2, 3, 4 and 7 are shown in Fig. 1. In addition to the circular streamlines around vortices, the flow is characterized by N๐‘Nitalic_N anticyclonic cells rotating around elliptic equilibrium positions of fluid points, that will be denoted ๐—eโขq0superscriptsubscript๐—๐‘’๐‘ž0\mathbf{X}_{eq}^{0}bold_X start_POSTSUBSCRIPT italic_e italic_q end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT throughout the paper (the superscript 0 indicating that they correspond to particles with Sโขt=0๐‘†๐‘ก0St=0italic_S italic_t = 0).

Refer to caption
Figure 1: Streamlines of the flow in the frame rotating with the crystal, when N๐‘Nitalic_N = 2, 3, 4 and 7. Lengths are set non-dimensional by means of the radius a๐‘Žaitalic_a of the circumscribed circle containing vortices. Vortices have a positive sign and are located at (cosโก(nโข2โขฯ€/N),sinโก(nโข2โขฯ€/N))๐‘›2๐œ‹๐‘๐‘›2๐œ‹๐‘(\cos(n2\pi/N),\sin(n2\pi/N))( roman_cos ( italic_n 2 italic_ฯ€ / italic_N ) , roman_sin ( italic_n 2 italic_ฯ€ / italic_N ) ), n=0,..,Nโˆ’1n=0,..,N-1italic_n = 0 , . . , italic_N - 1. In each case, the N๐‘Nitalic_N recirculation cells are anticyclonic. Note that, when N>2๐‘2N>2italic_N > 2, an additional anticyclonic cell exists around the center-of-vorticity located at (x,y)=(0,0)๐‘ฅ๐‘ฆ00(x,y)=(0,0)( italic_x , italic_y ) = ( 0 , 0 ).

These anticyclonic cells will play a key role in the trapping of inertial particles: the Coriolis force acting on these particles will be directed towards the interior of the cell. Note also that, as soon as N>2๐‘2N>2italic_N > 2, another anticyclonic recirculation cell exists around the center-of-vorticity (0,0)00(0,0)( 0 , 0 ), which is also an elliptic equilibrium position of fluid points. The equilibrium positions of fluid points are fixed points of the system ๐—ห™=๐ฎโข(๐—)ห™๐—๐ฎ๐—\dot{\mathbf{X}}=\mathbf{u}(\mathbf{X})overห™ start_ARG bold_X end_ARG = bold_u ( bold_X ). Their stability can be studied by means of the eigenvalues ฮผ๐œ‡\muitalic_ฮผ of the gradient of ๐ฎ๐ฎ\mathbf{u}bold_u, which satisfy

ฮผ2=(โˆ‚uโˆ‚x)2+(โˆ‚uโˆ‚y)2โˆ’2โขโˆ‚uโˆ‚yโ‰กIโข(๐—).superscript๐œ‡2superscript๐‘ข๐‘ฅ2superscript๐‘ข๐‘ฆ22๐‘ข๐‘ฆ๐ผ๐—\mu^{2}=(\frac{\partial u}{\partial x})^{2}+(\frac{\partial u}{\partial y})^{2% }-2\frac{\partial u}{\partial y}\equiv I(\mathbf{X}).italic_ฮผ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = ( divide start_ARG โˆ‚ italic_u end_ARG start_ARG โˆ‚ italic_x end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + ( divide start_ARG โˆ‚ italic_u end_ARG start_ARG โˆ‚ italic_y end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 2 divide start_ARG โˆ‚ italic_u end_ARG start_ARG โˆ‚ italic_y end_ARG โ‰ก italic_I ( bold_X ) . (4)

Here we chose 1/ฮฉ01subscriptฮฉ01/\Omega_{0}1 / roman_ฮฉ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT as time unit, and made use of the fact that the flow is incompressible and irrotational in the laboratory frame, that is โˆ‚uโˆ‚x+โˆ‚vโˆ‚y=0๐‘ข๐‘ฅ๐‘ฃ๐‘ฆ0\frac{\partial u}{\partial x}+\frac{\partial v}{\partial y}=0divide start_ARG โˆ‚ italic_u end_ARG start_ARG โˆ‚ italic_x end_ARG + divide start_ARG โˆ‚ italic_v end_ARG start_ARG โˆ‚ italic_y end_ARG = 0 and โˆ‚vโˆ‚xโˆ’โˆ‚uโˆ‚y=โˆ’2๐‘ฃ๐‘ฅ๐‘ข๐‘ฆ2\frac{\partial v}{\partial x}-\frac{\partial u}{\partial y}=-2divide start_ARG โˆ‚ italic_v end_ARG start_ARG โˆ‚ italic_x end_ARG - divide start_ARG โˆ‚ italic_u end_ARG start_ARG โˆ‚ italic_y end_ARG = - 2. Using both expressions we could re-write the gradient of ๐ฎ๐ฎ\mathbf{u}bold_u in terms of derivatives of the x๐‘ฅxitalic_x-velocity only, and obtain its characteristic polynomial (4). Elliptic equilibrium points ๐—eโขq0superscriptsubscript๐—๐‘’๐‘ž0\mathbf{X}_{eq}^{0}bold_X start_POSTSUBSCRIPT italic_e italic_q end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT are characterized by pure imaginary eigenvalues, so that I0โ‰กIโข(๐—eโขq0)<0subscript๐ผ0๐ผsuperscriptsubscript๐—๐‘’๐‘ž00I_{0}\equiv I(\mathbf{X}_{eq}^{0})<0italic_I start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT โ‰ก italic_I ( bold_X start_POSTSUBSCRIPT italic_e italic_q end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT ) < 0.

In particular, the determinant of โˆ‡๐ฎโˆ‡๐ฎ\nabla\mathbf{u}โˆ‡ bold_u at ๐—eโขq0superscriptsubscript๐—๐‘’๐‘ž0\mathbf{X}_{eq}^{0}bold_X start_POSTSUBSCRIPT italic_e italic_q end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT is equal to |I0|subscript๐ผ0|I_{0}|| italic_I start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT | and is non-zero. The implicit functions theorem mentioned above implies the existence of the branch ๐—eโขqโข(Sโขt)subscript๐—๐‘’๐‘ž๐‘†๐‘ก\mathbf{X}_{eq}(St)bold_X start_POSTSUBSCRIPT italic_e italic_q end_POSTSUBSCRIPT ( italic_S italic_t ), for inertial particles transported in this flow in the limit of small Stokes number Sโขt๐‘†๐‘กStitalic_S italic_t. For a given Stokes number, the stability of the branch ๐—eโขqโข(Sโขt)subscript๐—๐‘’๐‘ž๐‘†๐‘ก\mathbf{X}_{eq}(St)bold_X start_POSTSUBSCRIPT italic_e italic_q end_POSTSUBSCRIPT ( italic_S italic_t ) emerging from ๐—eโขq0superscriptsubscript๐—๐‘’๐‘ž0\mathbf{X}_{eq}^{0}bold_X start_POSTSUBSCRIPT italic_e italic_q end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT can be investigated by linearising the particle motion equation (1), in the vicinity of (๐—eโขqโข(Sโขt),๐—ห™=๐ŸŽ)subscript๐—๐‘’๐‘ž๐‘†๐‘กห™๐—0(\mathbf{X}_{eq}(St),\dot{\mathbf{X}}=\mathbf{0})( bold_X start_POSTSUBSCRIPT italic_e italic_q end_POSTSUBSCRIPT ( italic_S italic_t ) , overห™ start_ARG bold_X end_ARG = bold_0 ), with ๐…=๐—โˆ’2โข๐ณ^ร—๐—ห™๐…๐—2^๐ณห™๐—\mathbf{F}=\mathbf{X}-2\hat{\mathbf{z}}\times\dot{\mathbf{X}}bold_F = bold_X - 2 over^ start_ARG bold_z end_ARG ร— overห™ start_ARG bold_X end_ARG corresponding to the non-dimensional centrifugal and Coriolis forces (๐ณ^^๐ณ\hat{\mathbf{z}}over^ start_ARG bold_z end_ARG is the direct unit vector perpendicular to the (x,y)๐‘ฅ๐‘ฆ(x,y)( italic_x , italic_y ) plane). The system is then recast as a dynamical system with 4 degrees-of-freedom (๐—,๐—ห™)๐—ห™๐—(\mathbf{X},\dot{\mathbf{X}})( bold_X , overห™ start_ARG bold_X end_ARG ). Exploiting the fact that Sโขtโ‰ช1much-less-than๐‘†๐‘ก1St\ll 1italic_S italic_t โ‰ช 1, one obtains that the eigenvalues of the dynamical system (1) at the equilibrium point have the following real parts:

Rโขeโข(ฮป1)๐‘…๐‘’subscript๐œ†1\displaystyle Re(\lambda_{1})italic_R italic_e ( italic_ฮป start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) =\displaystyle== Rโขeโข(ฮป3)=โˆ’(1+I0)โขSโขt,๐‘…๐‘’subscript๐œ†31subscript๐ผ0๐‘†๐‘ก\displaystyle Re(\lambda_{3})=-(1+I_{0})\,St,italic_R italic_e ( italic_ฮป start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ) = - ( 1 + italic_I start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) italic_S italic_t , (5)
Rโขeโข(ฮป2)๐‘…๐‘’subscript๐œ†2\displaystyle Re(\lambda_{2})italic_R italic_e ( italic_ฮป start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) =\displaystyle== Rโขeโข(ฮป4)=โˆ’1Sโขt+(1+I0)โขSโขt๐‘…๐‘’subscript๐œ†41๐‘†๐‘ก1subscript๐ผ0๐‘†๐‘ก\displaystyle Re(\lambda_{4})=-\frac{1}{St}+(1+I_{0})\,Stitalic_R italic_e ( italic_ฮป start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ) = - divide start_ARG 1 end_ARG start_ARG italic_S italic_t end_ARG + ( 1 + italic_I start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) italic_S italic_t (6)

plus terms of order Oโข(Sโขt2)๐‘‚๐‘†superscript๐‘ก2O(St^{2})italic_O ( italic_S italic_t start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ). Hence, if 1+I0โ‰ 01subscript๐ผ001+I_{0}\not=01 + italic_I start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT โ‰  0, and even if the fluid stagnation point is degenerate (non-hyperbolic), the neighbouring equilibrium point of inertial particles is not, since all real parts are non-zero in the limit of small Stokes numbers. It is hyperbolic and its stability can be derived by examining the sign of these real parts. In the limit where Sโขtโ‰ช1much-less-than๐‘†๐‘ก1St\ll 1italic_S italic_t โ‰ช 1, the real part of ฮป2subscript๐œ†2\lambda_{2}italic_ฮป start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT and ฮป4subscript๐œ†4\lambda_{4}italic_ฮป start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT is always strictly negative. If 1+I0>01subscript๐ผ001+I_{0}>01 + italic_I start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT > 0, the real part of ฮป1subscript๐œ†1\lambda_{1}italic_ฮป start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and ฮป3subscript๐œ†3\lambda_{3}italic_ฮป start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT is also strictly negative and the equilibrium point ๐—eโขqโข(Sโขt)subscript๐—๐‘’๐‘ž๐‘†๐‘ก\mathbf{X}_{eq}(St)bold_X start_POSTSUBSCRIPT italic_e italic_q end_POSTSUBSCRIPT ( italic_S italic_t ) is asymptotically stable and attracts particles released in its basin of attraction. For the four cases of figure 1, the position of the fluid elliptic stagnation points in the anticyclonic cells could easily be found (exactly for N๐‘Nitalic_N=2 and 4), and one gets: 1+I0=1/41subscript๐ผ0141+I_{0}=1/41 + italic_I start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 1 / 4 (N=2๐‘2N=2italic_N = 2), 1+I0โ‰ƒ0.121similar-to-or-equals1subscript๐ผ00.1211+I_{0}\simeq 0.1211 + italic_I start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT โ‰ƒ 0.121 (N=3๐‘3N=3italic_N = 3), 1+I0=11/4โˆ’71subscript๐ผ011471+I_{0}=11/4-\sqrt{7}1 + italic_I start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 11 / 4 - square-root start_ARG 7 end_ARG (N=4๐‘4N=4italic_N = 4), and 1+I0โ‰ƒ0.335similar-to-or-equals1subscript๐ผ00.3351+I_{0}\simeq 0.3351 + italic_I start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT โ‰ƒ 0.335 (N=7๐‘7N=7italic_N = 7). These values being positive, we conclude that inertial particles must be trapped in the vicinity of these points, provided their Stokes number is sufficiently small. Since the elliptic points in the anticyclonic cells are all located at |๐—|>1๐—1|\mathbf{X}|>1| bold_X | > 1, that is outside the ring formed by the vortices, the attracting points ๐—eโขqโข(Sโขt)subscript๐—๐‘’๐‘ž๐‘†๐‘ก\mathbf{X}_{eq}(St)bold_X start_POSTSUBSCRIPT italic_e italic_q end_POSTSUBSCRIPT ( italic_S italic_t ) will be called satellite attracting points in the following.

In addition to these satellite points, another equilibrium point exists for particles in these flows. Indeed, in the vicinity of the fluid stagnation point located at the center-of-vorticity (0,0)00(0,0)( 0 , 0 ), one can check that the non-dimensional streamfunction reads ฯˆโ‰ƒ12โขr2+Oโข(rN)similar-to-or-equals๐œ“12superscript๐‘Ÿ2๐‘‚superscript๐‘Ÿ๐‘\psi\simeq\frac{1}{2}r^{2}+O(r^{N})italic_ฯˆ โ‰ƒ divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_O ( italic_r start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT ), leading to I0=โˆ’1subscript๐ผ01I_{0}=-1italic_I start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = - 1 when N>2๐‘2N>2italic_N > 2. Eq. (5) therefore implies that Rโขeโข(ฮป1)=Rโขeโข(ฮป3)=Oโข(Sโขt2)๐‘…๐‘’subscript๐œ†1๐‘…๐‘’subscript๐œ†3๐‘‚๐‘†superscript๐‘ก2Re(\lambda_{1})=Re(\lambda_{3})=O(St^{2})italic_R italic_e ( italic_ฮป start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) = italic_R italic_e ( italic_ฮป start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ) = italic_O ( italic_S italic_t start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ), so that the linear analysis does not allow to conclude about the stability of the equilibrium branch near the center. Particles rotate around the center, while they can drift towards the interior or the exterior of the cell under the effect of the non-linear terms Oโข(rN)๐‘‚superscript๐‘Ÿ๐‘O(r^{N})italic_O ( italic_r start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT ) appearing in the streamfunction. To study the effect of these terms, we consider an inertial particle injected very near (0,0)00(0,0)( 0 , 0 ), i.e. |๐—โข(0)|=ฮตโ‰ช1๐—0๐œ€much-less-than1|\mathbf{X}(0)|=\varepsilon\ll 1| bold_X ( 0 ) | = italic_ฮต โ‰ช 1 and re-scale the positions by ฮต๐œ€\varepsilonitalic_ฮต, setting xโˆ—=x/ฮตsubscript๐‘ฅ๐‘ฅ๐œ€x_{*}=x/\varepsilonitalic_x start_POSTSUBSCRIPT โˆ— end_POSTSUBSCRIPT = italic_x / italic_ฮต, yโˆ—=y/ฮตsubscript๐‘ฆ๐‘ฆ๐œ€y_{*}=y/\varepsilonitalic_y start_POSTSUBSCRIPT โˆ— end_POSTSUBSCRIPT = italic_y / italic_ฮต. This leads to

ฯˆโˆ—=ฯˆฮต2=ฯˆ0โˆ—+ฮตNโˆ’2โขฯˆNโˆ’2โˆ—+Oโข(ฮต2โขNโˆ’2)superscript๐œ“๐œ“superscript๐œ€2superscriptsubscript๐œ“0superscript๐œ€๐‘2superscriptsubscript๐œ“๐‘2๐‘‚superscript๐œ€2๐‘2\psi^{*}=\frac{\psi}{\varepsilon^{2}}=\psi_{0}^{*}+\varepsilon^{N-2}\psi_{N-2}% ^{*}+O(\varepsilon^{2N-2})italic_ฯˆ start_POSTSUPERSCRIPT โˆ— end_POSTSUPERSCRIPT = divide start_ARG italic_ฯˆ end_ARG start_ARG italic_ฮต start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG = italic_ฯˆ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT โˆ— end_POSTSUPERSCRIPT + italic_ฮต start_POSTSUPERSCRIPT italic_N - 2 end_POSTSUPERSCRIPT italic_ฯˆ start_POSTSUBSCRIPT italic_N - 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT โˆ— end_POSTSUPERSCRIPT + italic_O ( italic_ฮต start_POSTSUPERSCRIPT 2 italic_N - 2 end_POSTSUPERSCRIPT ) (7)

where ฯˆ0โˆ—=(xโˆ—2+yโˆ—2)/2superscriptsubscript๐œ“0superscriptsubscript๐‘ฅ2superscriptsubscript๐‘ฆ22\psi_{0}^{*}=(x_{*}^{2}+y_{*}^{2})/2italic_ฯˆ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT โˆ— end_POSTSUPERSCRIPT = ( italic_x start_POSTSUBSCRIPT โˆ— end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_y start_POSTSUBSCRIPT โˆ— end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) / 2 and ฯˆNโˆ’2โˆ—โข(xโˆ—,yโˆ—)superscriptsubscript๐œ“๐‘2subscript๐‘ฅsubscript๐‘ฆ\psi_{N-2}^{*}(x_{*},y_{*})italic_ฯˆ start_POSTSUBSCRIPT italic_N - 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT โˆ— end_POSTSUPERSCRIPT ( italic_x start_POSTSUBSCRIPT โˆ— end_POSTSUBSCRIPT , italic_y start_POSTSUBSCRIPT โˆ— end_POSTSUBSCRIPT ) is a polynomial of degree N๐‘Nitalic_N: ฯˆNโˆ’2โˆ—=xโˆ—3โˆ’3โขxโˆ—โขyโˆ—2superscriptsubscript๐œ“๐‘2superscriptsubscript๐‘ฅ33subscript๐‘ฅsuperscriptsubscript๐‘ฆ2\psi_{N-2}^{*}=x_{*}^{3}-3x_{*}y_{*}^{2}italic_ฯˆ start_POSTSUBSCRIPT italic_N - 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT โˆ— end_POSTSUPERSCRIPT = italic_x start_POSTSUBSCRIPT โˆ— end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT - 3 italic_x start_POSTSUBSCRIPT โˆ— end_POSTSUBSCRIPT italic_y start_POSTSUBSCRIPT โˆ— end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (if N=3๐‘3N=3italic_N = 3), ฯˆNโˆ’2โˆ—=2/3โขxโˆ—4+2/3โขyโˆ—4โˆ’4โขxโˆ—2โขyโˆ—2superscriptsubscript๐œ“๐‘223superscriptsubscript๐‘ฅ423superscriptsubscript๐‘ฆ44superscriptsubscript๐‘ฅ2superscriptsubscript๐‘ฆ2\psi_{N-2}^{*}=2/3\,{x_{*}}^{4}+2/3\,{y_{*}}^{4}-4\,{x_{*}}^{2}{y_{*}}^{2}italic_ฯˆ start_POSTSUBSCRIPT italic_N - 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT โˆ— end_POSTSUPERSCRIPT = 2 / 3 italic_x start_POSTSUBSCRIPT โˆ— end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT + 2 / 3 italic_y start_POSTSUBSCRIPT โˆ— end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT - 4 italic_x start_POSTSUBSCRIPT โˆ— end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_y start_POSTSUBSCRIPT โˆ— end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT(if N=4๐‘4N=4italic_N = 4), etc.

Refer to caption
Figure 2: Clouds of particles in the laboratory frame, advected by 3 vortices (first column), 4 vortices (middle column) and 7 vortices (third column), after two turns (upper row) and 12 turns (middle row) of the crystal. White circles are vortex centers, and black dots are particles. The Stokes number of particles is Sโขt=0.062๐‘†๐‘ก0.062St=0.062italic_S italic_t = 0.062 for N=3๐‘3N=3italic_N = 3 and 4 vortices, and Sโขt=0.021๐‘†๐‘ก0.021St=0.021italic_S italic_t = 0.021 for the case N=7๐‘7N=7italic_N = 7. The last row shows the initial positions of the particles trapped by satellite points (blue dots) and by the center point (red dots). White squares indicate the position of the degenerate fluid stagnation points ๐—eโขq0superscriptsubscript๐—๐‘’๐‘ž0\mathbf{X}_{eq}^{0}bold_X start_POSTSUBSCRIPT italic_e italic_q end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT in anticyclonic cells.

Then we calculate the change of the streamfunction ฮ”โขฯˆโˆ—=ฯˆโˆ—โข(๐—โข(T))โˆ’ฯˆโˆ—โข(๐—โข(0))ฮ”superscript๐œ“superscript๐œ“๐—๐‘‡superscript๐œ“๐—0\Delta\psi^{*}=\psi^{*}(\mathbf{X}(T))-\psi^{*}(\mathbf{X}(0))roman_ฮ” italic_ฯˆ start_POSTSUPERSCRIPT โˆ— end_POSTSUPERSCRIPT = italic_ฯˆ start_POSTSUPERSCRIPT โˆ— end_POSTSUPERSCRIPT ( bold_X ( italic_T ) ) - italic_ฯˆ start_POSTSUPERSCRIPT โˆ— end_POSTSUPERSCRIPT ( bold_X ( 0 ) ) during one period T=2โขฯ€๐‘‡2๐œ‹T=2\piitalic_T = 2 italic_ฯ€ of the rotation of the crystal. Negative values of this quantity will indicate that the particle drifts towards the ring center (as ฯˆโˆ—superscript๐œ“\psi^{*}italic_ฯˆ start_POSTSUPERSCRIPT โˆ— end_POSTSUPERSCRIPT increases with the distance to (0,0)00(0,0)( 0 , 0 ), when ฮต๐œ€\varepsilonitalic_ฮต is sufficiently small). To achieve this we consider a fixed value of ฮต๐œ€\varepsilonitalic_ฮต and assume that ฮต2โขNโˆ’2โ‰ชSโขtโ‰ชฮตmuch-less-thansuperscript๐œ€2๐‘2๐‘†๐‘กmuch-less-than๐œ€\varepsilon^{2N-2}\ll St\ll\varepsilonitalic_ฮต start_POSTSUPERSCRIPT 2 italic_N - 2 end_POSTSUPERSCRIPT โ‰ช italic_S italic_t โ‰ช italic_ฮต. On perturbing Eq. (1) we get, for the re-scaled variables:

๐—ห™โˆ—โ‰ƒ๐ฎโˆ—+Sโขtโข(โˆ’(โˆ‡๐ฎโˆ—)โข๐ฎโˆ—+๐—โˆ—โˆ’2โข๐ณ^ร—๐ฎโˆ—)+Oโข(Sโขt2)similar-to-or-equalssubscriptห™๐—subscript๐ฎ๐‘†๐‘กโˆ‡subscript๐ฎsubscript๐ฎsubscript๐—2^๐ณsubscript๐ฎ๐‘‚๐‘†superscript๐‘ก2\dot{\mathbf{X}}_{*}\simeq\mathbf{u}_{*}+St\,(-(\nabla\mathbf{u}_{*})\mathbf{u% }_{*}+\mathbf{X}_{*}-2\hat{\mathbf{z}}\times\mathbf{u}_{*})+O(St^{2})overห™ start_ARG bold_X end_ARG start_POSTSUBSCRIPT โˆ— end_POSTSUBSCRIPT โ‰ƒ bold_u start_POSTSUBSCRIPT โˆ— end_POSTSUBSCRIPT + italic_S italic_t ( - ( โˆ‡ bold_u start_POSTSUBSCRIPT โˆ— end_POSTSUBSCRIPT ) bold_u start_POSTSUBSCRIPT โˆ— end_POSTSUBSCRIPT + bold_X start_POSTSUBSCRIPT โˆ— end_POSTSUBSCRIPT - 2 over^ start_ARG bold_z end_ARG ร— bold_u start_POSTSUBSCRIPT โˆ— end_POSTSUBSCRIPT ) + italic_O ( italic_S italic_t start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) (8)

where ๐ฎโˆ—=โˆ‡ฯˆโˆ—ร—๐ณ^subscript๐ฎโˆ‡superscript๐œ“^๐ณ\mathbf{u}_{*}=\nabla\psi^{*}\times\hat{\mathbf{z}}bold_u start_POSTSUBSCRIPT โˆ— end_POSTSUBSCRIPT = โˆ‡ italic_ฯˆ start_POSTSUPERSCRIPT โˆ— end_POSTSUPERSCRIPT ร— over^ start_ARG bold_z end_ARG. We then obtain

ฮ”โขฯˆโˆ—=Sโขtโขโˆซ0Tโˆ‡ฯˆโˆ—โข(๐—โข(t))โ‹…๐—ห™โˆ—โข๐‘‘tฮ”superscript๐œ“๐‘†๐‘กsuperscriptsubscript0๐‘‡โ‹…โˆ‡subscript๐œ“๐—๐‘กsubscriptห™๐—differential-d๐‘ก\Delta\psi^{*}=St\,\int_{0}^{T}\nabla\psi_{*}(\mathbf{X}(t))\cdot\dot{\mathbf{% X}}_{*}dtroman_ฮ” italic_ฯˆ start_POSTSUPERSCRIPT โˆ— end_POSTSUPERSCRIPT = italic_S italic_t โˆซ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT โˆ‡ italic_ฯˆ start_POSTSUBSCRIPT โˆ— end_POSTSUBSCRIPT ( bold_X ( italic_t ) ) โ‹… overห™ start_ARG bold_X end_ARG start_POSTSUBSCRIPT โˆ— end_POSTSUBSCRIPT italic_d italic_t
=Sโขtโขโˆซ0T[โˆ‡ฯˆโˆ—โ‹…๐—โˆ—โˆ’2โข|โˆ‡ฯˆโˆ—|2โˆ’โˆ‡ฯˆโˆ—โ‹…(โˆ‡๐ฎโˆ—)โข๐ฎโˆ—]โข๐‘‘t.absent๐‘†๐‘กsuperscriptsubscript0๐‘‡delimited-[]โˆ‡โ‹…superscript๐œ“subscript๐—2superscriptโˆ‡superscript๐œ“2โ‹…โˆ‡superscript๐œ“โˆ‡subscript๐ฎsubscript๐ฎdifferential-d๐‘ก=St\,\int_{0}^{T}[\nabla\psi^{*}\cdot\mathbf{X}_{*}-2|\nabla\psi^{*}|^{2}-% \nabla\psi^{*}\cdot(\nabla\mathbf{u}_{*})\mathbf{u}_{*}]dt.= italic_S italic_t โˆซ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT [ โˆ‡ italic_ฯˆ start_POSTSUPERSCRIPT โˆ— end_POSTSUPERSCRIPT โ‹… bold_X start_POSTSUBSCRIPT โˆ— end_POSTSUBSCRIPT - 2 | โˆ‡ italic_ฯˆ start_POSTSUPERSCRIPT โˆ— end_POSTSUPERSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - โˆ‡ italic_ฯˆ start_POSTSUPERSCRIPT โˆ— end_POSTSUPERSCRIPT โ‹… ( โˆ‡ bold_u start_POSTSUBSCRIPT โˆ— end_POSTSUBSCRIPT ) bold_u start_POSTSUBSCRIPT โˆ— end_POSTSUBSCRIPT ] italic_d italic_t . (9)

The first term in the integral (9) is the contribution of the centrifugal force, the second one is the contribution of the Coriolis force, and the last one is the contribution of the drag due to the flow induced by vortices. As expected, the second term is always negative because the Coriolis force acts towards the right-hand-side of ๐—ห™ห™๐—\dot{\mathbf{X}}overห™ start_ARG bold_X end_ARG here, whereas the streamfunction ฯˆโˆ—superscript๐œ“\psi^{*}italic_ฯˆ start_POSTSUPERSCRIPT โˆ— end_POSTSUPERSCRIPT increases towards the left-hand-side of streamlines. These integrals are calculated by approximating the particle trajectory by the fluid point trajectory within integrands: ๐—โˆ—โข(t)=๐—fโข(t)+Oโข(Sโขt)subscript๐—๐‘กsubscript๐—๐‘“๐‘ก๐‘‚๐‘†๐‘ก\mathbf{X}_{*}(t)={\mathbf{X}}_{f}(t)+O(St)bold_X start_POSTSUBSCRIPT โˆ— end_POSTSUBSCRIPT ( italic_t ) = bold_X start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ( italic_t ) + italic_O ( italic_S italic_t ) with ๐—ห™fโ‰ƒ๐ฎโˆ—0+ฮตNโˆ’2โข๐ฎNโˆ’2โˆ—similar-to-or-equalssubscriptห™๐—๐‘“subscript๐ฎabsent0superscript๐œ€๐‘2subscriptsuperscript๐ฎ๐‘2\dot{\mathbf{X}}_{f}\simeq\mathbf{u}_{*0}+\varepsilon^{N-2}\mathbf{u}^{*}_{N-2}overห™ start_ARG bold_X end_ARG start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT โ‰ƒ bold_u start_POSTSUBSCRIPT โˆ— 0 end_POSTSUBSCRIPT + italic_ฮต start_POSTSUPERSCRIPT italic_N - 2 end_POSTSUPERSCRIPT bold_u start_POSTSUPERSCRIPT โˆ— end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_N - 2 end_POSTSUBSCRIPT. The solution is approximated in the form ๐—f=(cosโกt,โˆ’sinโกt)+ฮตโข๐—1โข(t)+ฮต2โข๐—2โข(t)+eโขtโขcsubscript๐—๐‘“๐‘ก๐‘ก๐œ€subscript๐—1๐‘กsuperscript๐œ€2subscript๐—2๐‘ก๐‘’๐‘ก๐‘{\mathbf{X}}_{f}=(\cos t,-\sin t)+\varepsilon{\mathbf{X}}_{1}(t)+\varepsilon^{% 2}{\mathbf{X}}_{2}(t)+etcbold_X start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT = ( roman_cos italic_t , - roman_sin italic_t ) + italic_ฮต bold_X start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_t ) + italic_ฮต start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT bold_X start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_t ) + italic_e italic_t italic_c, where the first term corresponds to the motion of a fluid point when vortices are at infinity, and ๐—iโข(t)subscript๐—๐‘–๐‘ก{\mathbf{X}}_{i}(t)bold_X start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) are to be found by injecting this expansion into the motion equation of fluid points. In the case N=3๐‘3N=3italic_N = 3, we get

ฮ”โขฯˆโˆ—=โˆ’36โขฯ€โขSโขtโขฮต2ฮ”superscript๐œ“36๐œ‹๐‘†๐‘กsuperscript๐œ€2\Delta\psi^{*}=-36\,\pi\,St\,\varepsilon^{2}roman_ฮ” italic_ฯˆ start_POSTSUPERSCRIPT โˆ— end_POSTSUPERSCRIPT = - 36 italic_ฯ€ italic_S italic_t italic_ฮต start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (10)

plus terms of order Sโขt2๐‘†superscript๐‘ก2St^{2}italic_S italic_t start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. We conclude that particles released in the vicinity of ๐—=๐ŸŽ๐—0\mathbf{X}=\mathbf{0}bold_X = bold_0, i.e. within the anticyclonic cell around the center, will slowly drift towards the center. The set of vortices maintain the particles trapped in a kind of โ€cageโ€. We note also that the drift of particles towards the center is very slow: back to the original scales (i.e. when lengths are reduced by the radius of the vortex crystal a๐‘Žaitalic_a), Eq. (10) reads ฮ”โขฯˆ=โˆ’36โขฯ€โขSโขtโข|๐—โข(0)|4ฮ”๐œ“36๐œ‹๐‘†๐‘กsuperscript๐—04\Delta\psi=-36\,\pi\,St\,|\mathbf{X}(0)|^{4}roman_ฮ” italic_ฯˆ = - 36 italic_ฯ€ italic_S italic_t | bold_X ( 0 ) | start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT. Therefore, the change of streamfunction at each period is very small since |๐—โข(0)|โ‰ช1much-less-than๐—01|\mathbf{X}(0)|\ll 1| bold_X ( 0 ) | โ‰ช 1.

These theoretical results suggest that particles, in a system of N๐‘Nitalic_N vortices located on a ring, can be trapped by N๐‘Nitalic_N satellite attracting points located outside the ring, and by an additional attracting point located at the center-of-vorticity. To check these results we have performed numerical simulations (in the laboratory frame with coordinates (X,Y)๐‘‹๐‘Œ(X,Y)( italic_X , italic_Y )) involving N๐‘Nitalic_N point vortices moving under their mutual influence and initially placed on a ring with unit radius. In addition, 10000 inertial particles have been released within the flow domain, and their positions have been computed by solving Eq. (1). It should be noted that when N=3๐‘3N=3italic_N = 3 the critical Stokes number, below which equilibrium points do not exist, can be obtained analytically from Eq. (2). After a few manipulations using polar coordinates we show that the non-dimensional radius r=|๐—eโขq|๐‘Ÿsubscript๐—๐‘’๐‘žr=|\mathbf{X}_{eq}|italic_r = | bold_X start_POSTSUBSCRIPT italic_e italic_q end_POSTSUBSCRIPT | of equilibrium positions satisfies either r=0๐‘Ÿ0r=0italic_r = 0 or r6โˆ’6โขr4/(1+Sโขt2)+9โขr2/(1+Sโขt2)โˆ’1=0superscript๐‘Ÿ66superscript๐‘Ÿ41๐‘†superscript๐‘ก29superscript๐‘Ÿ21๐‘†superscript๐‘ก210r^{6}-{6}r^{4}/(1+St^{2})+{9}r^{2}/(1+St^{2})-1=0italic_r start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT - 6 italic_r start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT / ( 1 + italic_S italic_t start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) + 9 italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / ( 1 + italic_S italic_t start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) - 1 = 0. This is a cubic equation in r2superscript๐‘Ÿ2r^{2}italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, the discriminant of which is zero when the Stokes number is equal to the critical value:

Sโขtc=1/2โข2โข48โขAโˆ’24/3โขโ€‰192/3โขA+280Aโˆ’4โˆ’2โขA๐‘†subscript๐‘ก๐‘12248๐ดsuperscript243superscript1923๐ด280๐ด42๐ดSt_{c}=1/2\,\sqrt{2\,{\frac{\sqrt{48\,A-{2}^{4/3}\,{19}^{2/3}A+280}}{\sqrt{A}}% }-4-2\,A}italic_S italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = 1 / 2 square-root start_ARG 2 divide start_ARG square-root start_ARG 48 italic_A - 2 start_POSTSUPERSCRIPT 4 / 3 end_POSTSUPERSCRIPT 19 start_POSTSUPERSCRIPT 2 / 3 end_POSTSUPERSCRIPT italic_A + 280 end_ARG end_ARG start_ARG square-root start_ARG italic_A end_ARG end_ARG - 4 - 2 italic_A end_ARG

where A=(24+24/3โขโ€‰192/3)1/2๐ดsuperscript24superscript243superscript192312A=({24+{2}^{4/3}\,{19}^{2/3}})^{1/2}italic_A = ( 24 + 2 start_POSTSUPERSCRIPT 4 / 3 end_POSTSUPERSCRIPT 19 start_POSTSUPERSCRIPT 2 / 3 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT, leading to Sโขtcโ‰ˆ0.206๐‘†subscript๐‘ก๐‘0.206St_{c}\approx 0.206italic_S italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT โ‰ˆ 0.206 in the three-vortex system, in agreement with Fig. 3 of Ref. Ravichandran et al. (2017) (up to a 2โขฯ€2๐œ‹2\pi2 italic_ฯ€ factor due to a different definition of the Stokes number). For N=4๐‘4N=4italic_N = 4 or more, no analytical expression can be found, and one has to use a numerical algorithm to approach Sโขtc๐‘†subscript๐‘ก๐‘St_{c}italic_S italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, leading to Sโขtcโ‰ˆ0.15๐‘†subscript๐‘ก๐‘0.15St_{c}\approx 0.15italic_S italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT โ‰ˆ 0.15 (N=4๐‘4N=4italic_N = 4) and Sโขtcโ‰ˆ0.05๐‘†subscript๐‘ก๐‘0.05St_{c}\approx 0.05italic_S italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT โ‰ˆ 0.05 (N=7๐‘7N=7italic_N = 7) (see also Fig. 3 of Ref. Ravichandran et al. (2017)). The first column of Fig. 2 shows the resulting clouds of particles, when N=3๐‘3N=3italic_N = 3 and Sโขt=0.3โขSโขtc๐‘†๐‘ก0.3๐‘†subscript๐‘ก๐‘St=0.3St_{c}italic_S italic_t = 0.3 italic_S italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. It confirms the existence of the three satellite attracting points, as well as trapping near the center-of-vorticity. The second and third columns of Fig. 2 correspond to N=4๐‘4N=4italic_N = 4 and N=7๐‘7N=7italic_N = 7 respectively, with Stokes number chosen arbitrarily, such that Sโขt<Sโขtc๐‘†๐‘ก๐‘†subscript๐‘ก๐‘St<St_{c}italic_S italic_t < italic_S italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. As expected, some particles are trapped by satellite points, or by the center point. Trapping by this point is very slow and long-term simulations (not shown here) would show tiny spots of particles near the satellite points, with a large cloud of particles slowly decaying near the center. The last row of Fig. 2 shows the initial positions of trapped particles, which give a view of the basin of attraction of the satellite points (blue dots), together with the basin of the center point (red dots) for N=๐‘absentN=italic_N = 3, 4 and 7. The size of the basin of the central point significantly increases with N๐‘Nitalic_N, as the size of the anticyclonic cell grows with N๐‘Nitalic_N and most trapped particles come from the interior of this cell. We note that these basins are not fractal, as expected since the dynamics is non-chaotic, but they have a spiral structure near vortices. Therefore, blue and red basins might be highly intricated near vortices.

These analyses show that inertial particles might eventually form a mass crystal moving as a rigid-body with the vorticity crystal. In addition, another attracting point exists near the center of the vortex crystal, in high contrast with the 2-vortex system where this point does not exist Angilella (2010); Ravichandran et al. (2014). These observations could reveal interesting features of particle dynamics in either laboratory of geophysical flows, where vortex crystals have been shown to emerge. Further analyses about particles lighter than the fluid, or with a density of the order of the fluid density, should also reveal interesting particle dynamics. Finally, the case where a vortex is present at the center-of-vorticity, which is known to increase the stability of the crystal, is also of interest for particle transport and will be the next step of this work.

References

  • Aref et al. (2002) H. Aref, P. K. Newton, M. A. Stremler, T. Tokieda, and D. L. Vainchtein, TAM Reports 1008 (2002).
  • Yarmchuk et al. (1979) E. Yarmchuk, M. Gordon, and R. Packard, Physical Review Letters 43, 214 (1979).
  • Fine et al. (1995) K. Fine, A. Cass, W. Flynn, and C. Driscoll, Physical Review Letters 75, 3277 (1995).
  • Durkin and Fajans (2000a) D. Durkin and J. Fajans, Physics of fluids 12, 289 (2000a).
  • Durkin and Fajans (2000b) D. Durkin and J. Fajans, Physical Review Letters 85, 4052 (2000b).
  • Grzybowski et al. (2000) B. A. Grzybowski, H. A. Stone, and G. M. Whitesides, Nature 405, 1033 (2000).
  • Grzybowski and Whitesides (2002) B. A. Grzybowski and G. M. Whitesides, The Journal of Physical Chemistry B 106, 1188 (2002).
  • Adriani et al. (2018) A. Adriani, A. Mura, G. Orton, C. Hansen, F. Altieri, M. Moriconi, J. Rogers, G. Eichstรคdt, T. Momary, A. P. Ingersoll, et al., Nature 555, 216 (2018).
  • Schecter et al. (1999) D. A. Schecter, D. H. Dubin, K. Fine, and C. Driscoll, Physics of Fluids 11, 905 (1999).
  • Jin and Dubin (2000) D. Z. Jin and D. Dubin, Physics of plasmas 7, 1719 (2000).
  • Siegelman et al. (2022) L. Siegelman, W. R. Young, and A. P. Ingersoll, Proceedings of the National Academy of Sciences 119, e2120486119 (2022).
  • Angilella (2010) J. R. Angilella, Physica D 239, 1789 (2010).
  • Ravichandran et al. (2014) S. Ravichandran, P. Perlekar, and R. Govindarajan, Physics of Fluids 26 (2014).
  • Ravichandran et al. (2017) S. Ravichandran, P. Deepu, and R. Govindarajan, Sฤdhanฤ 42, 597 (2017).
  • Babiano et al. (2000) A. Babiano, J. Cartwright, O. Piro, and A. Provenzale, Phys. Rev. Let. 84, 5764 (2000).
  • Haller and Sapsis (2008) G. Haller and T. Sapsis, Physica D 237, 573 (2008).
  • Cartwright et al. (2010) J. Cartwright, U. Feudel, G. Karolyi, A. De Moura, O. Piro, and T. Tel, pp. 51โ€“87 (2010), non-linear Dynamics and Chaos: Advances and Perspectives. Thiel, M. Ed. Springer-Verlag Berlin Heidelberg.
  • Mertz (1978) G. J. Mertz, The Physics of Fluids 21, 1092 (1978).