15
J. Chem. Phys. 150, 044902 (2019); https://doi.org/10.1063/1.5065656 150, 044902 © 2019 Author(s). Phoretic and hydrodynamic interactions of weakly confined autophoretic particles Cite as: J. Chem. Phys. 150, 044902 (2019); https://doi.org/10.1063/1.5065656 Submitted: 11 October 2018 . Accepted: 03 January 2019 . Published Online: 25 January 2019 Eva Kanso , and Sébastien Michelin

Phoretic and hydrodynamic interactions of weakly confined ... · Phoretic particles self-propel using self-generated physico-chemical gradients at their surface. Within a suspension,

  • Upload
    others

  • View
    3

  • Download
    0

Embed Size (px)

Citation preview

  • J. Chem. Phys. 150, 044902 (2019); https://doi.org/10.1063/1.5065656 150, 044902

    © 2019 Author(s).

    Phoretic and hydrodynamic interactions ofweakly confined autophoretic particlesCite as: J. Chem. Phys. 150, 044902 (2019); https://doi.org/10.1063/1.5065656Submitted: 11 October 2018 . Accepted: 03 January 2019 . Published Online: 25 January 2019

    Eva Kanso , and Sébastien Michelin

    http://oasc12039.247realmedia.com/RealMedia/ads/click_lx.ads/www.aip.org/pt/adcenter/pdfcover_test/L-37/1858055942/x01/AIP-PT/MB_JCPArticleDL_WP_0818/large-banner.jpg/434f71374e315a556e61414141774c75?xhttps://doi.org/10.1063/1.5065656https://doi.org/10.1063/1.5065656https://aip.scitation.org/author/Kanso%2C+Evahttp://orcid.org/0000-0003-0336-585Xhttps://aip.scitation.org/author/Michelin%2C+S%C3%A9bastienhttp://orcid.org/0000-0002-9037-7498https://doi.org/10.1063/1.5065656https://aip.scitation.org/action/showCitFormats?type=show&doi=10.1063/1.5065656http://crossmark.crossref.org/dialog/?doi=10.1063%2F1.5065656&domain=aip.scitation.org&date_stamp=2019-01-25

  • The Journal ofChemical Physics ARTICLE scitation.org/journal/jcp

    Phoretic and hydrodynamic interactionsof weakly confined autophoretic particles

    Cite as: J. Chem. Phys. 150, 044902 (2019); doi: 10.1063/1.5065656Submitted: 11 October 2018 • Accepted: 3 January 2019 •Published Online: 25 January 2019

    Eva Kanso1,a) and Sébastien Michelin2,b)

    AFFILIATIONS1Aerospace and Mechanical Engineering, University of Southern California, Los Angeles, California 90089-1191, USA2LadHyX—Département de Mécanique, Ecole Polytechnique—CNRS, 91128 Palaiseau, France

    Note: This article is part of the Special Topic “Chemical Physics of Active Matter” in J. Chem. Phys.a)Electronic mail: [email protected])Electronic mail: [email protected]

    ABSTRACTPhoretic particles self-propel using self-generated physico-chemical gradients at their surface. Within a suspension, they inter-act hydrodynamically by setting the fluid around them into motion and chemically by modifying the chemical background seenby their neighbours. While most phoretic systems evolve in confined environments due to buoyancy effects, most models focuson their interactions in unbounded flows. Here, we propose a first model for the interaction of phoretic particles in Hele-Shawconfinement and show that in this limit, hydrodynamic and phoretic interactions share not only the same scaling but also thesame form, albeit in opposite directions. In essence, we show that phoretic interactions effectively reverse the sign of the inter-actions that would be obtained for swimmers interacting purely hydrodynamically. Yet, hydrodynamic interactions cannot beneglected as they significantly impact the magnitude of the interactions. This model is then used to analyse the behavior of asuspension. The suspension exhibits swirling and clustering collective modes dictated by the orientational interactions betweenparticles, similar to hydrodynamic swimmers, but here governed by the surface properties of the phoretic particle; the reversalin the sign of the interaction tends to slow down the swimming motion of the particles.

    Published under license by AIP Publishing. https://doi.org/10.1063/1.5065656

    I. INTRODUCTION

    To self-propel autonomously at the microscopic scale,biological and synthetic micro-swimmers must overcome theviscous resistance of the surrounding fluid to create asymmet-ric and non-reciprocal flow fields in their immediate vicinity.1Beyond their individual self-propulsion, understanding theirinteractions and collective behavior is fascinating researchersacross disciplines, particularly because their small scale sug-gest simpler interaction routes than larger and more complexorganisms or systems.

    While biological swimmers mainly rely on the actuationof flexible appendages such as flagella or cilia,2,3 artificialmicroswimmers fall within two main categories. Externally-actuated swimmers respond to an external force or torqueapplied by a magnetic,4 electric,5 or acoustic field6 at themacroscopic level. In contrast, autophoretic (or fuel-based)swimmers exploit the chemical and electrical properties of

    their surface to generate slip flows within a thin inter-action layer in response to local self-generated gradientsof their physico-chemical environment. This ability to turnsuch gradients into fluid motion is known as phoresis,7and this can arise from concentration of a diffusing solutespecies (diffusiophoresis), temperature (thermophoresis), orelectric potential (electrophoresis). Popular experimentalrealizations of such systems include metallic or bi-metalliccolloids catalyzing the decomposition of hydrogen perox-ide solutions8–12 or other redox reactions,13 as well asheat-releasing particles in binary mixtures.24 It should benoted that in many aspects, active droplets, which achieveself-propulsion through self-generated Marangoni flows,14,15can also be considered as examples of synthetic fuel-basedswimmers.

    Because their individual self-propulsion is based solelyon the interaction with their immediate microscopic envi-ronment, these systems have recently been intensely stud-

    J. Chem. Phys. 150, 044902 (2019); doi: 10.1063/1.5065656 150, 044902-1

    Published under license by AIP Publishing

    https://scitation.org/journal/jcphttps://doi.org/10.1063/1.5065656https://www.scitation.org/action/showCitFormats?type=show&doi=10.1063/1.5065656https://crossmark.crossref.org/dialog/?doi=10.1063/1.5065656&domain=aip.scitation.org&date_stamp=2019-January-25https://doi.org/10.1063/1.5065656https://orcid.org/0000-0003-0336-585Xhttps://orcid.org/0000-0002-9037-7498mailto:[email protected]:[email protected]://doi.org/10.1063/1.5065656

  • The Journal ofChemical Physics ARTICLE scitation.org/journal/jcp

    ied, experimentally and numerically, as canonical examples ofactive matter to analyse collective dynamics at the micronscale, demonstrating complex or chaotic behavior as wellas clustering.16–20 Chemically-patterned systems (e.g., Janusparticles) have played a central role in these investigations.Their chemical polarity establishes the chemical gradientrequired for propulsion and can be obtained using a pas-sive colloid partially coated with a catalytic9,12,21–23 or heat-absorbing layer,24,25 bi-metallic swimmers,8,10,26,27 or activematerial encapsulated in a passive colloid.11 Beyond the quali-tative understanding of the link between their asymmetry andtheir self-propulsion, the details of the competing physico-chemical mechanisms at the heart of each system are still thefocus of ongoing investigations.28,29

    All such systems share a common feature: they gener-ate flow fields and motility from the dynamics of physico-chemical fields (which will be taken in the following as theconcentration of a chemical solute for simplicity and gener-ality) that can diffuse and potentially be advected by the fluidflows. This provides two distinct interaction routes betweenindividual swimmers. Like their biological counterparts, theirself-propulsion sets the surrounding fluid into motion whichinfluences the trajectory of their neighbors. But, just asbacteria and other microorganisms perform chemotaxis inresponse to different chemical nutrient or waste compounds,these systems can also exhibit a chemotactic behavior andrespond to the chemical signals left by the other particles.30,31Understanding the role of each interaction route32–34 as wellas their interplay and competition35 is the focus of activeresearch.

    A major difficulty in the quantitative rationalization ofexperimental results on collective dynamics of such particleslies in their geometric environment. Most phoretic swimmersare either denser or lighter than the surrounding fluid, andbuoyancy forces effectively confine them to the immediatevicinity of a solid wall or a free-surface.10,11,36 This confine-ment and reduced-dimensionality can have profound conse-quences on their collective behavior.37 Yet, the vast majorityof existing models consider their evolution in a 3D unboundedenvironment.28,38–40 Recent and successful attempts havedemonstrated the ability of confining boundaries to signif-icantly influence the dynamics of single particles,41–44 evenproviding controlling strategies for guidance.26,45,46 Becausethey can profoundly affect the flow field or chemical con-centration generated by the particle, confining boundariesalso significantly modify the interactions among them.37,47As an example, their ability to screen differently the viscousand potential components of the flow field generated by theswimmer48 allows confining boundaries to significantly alterthe clustering dynamics of such particles; these modificationswere recently shown to depend fundamentally on the numberand nature of the confining surfaces.49

    The goal of the present work is to analyse the fundamentalrole of confinement on the interactions of many phoretic par-ticles and, in particular, on the relative weight of the hydrody-namic and chemical (phoretic) coupling. To this end, we modela dilute suspension of weakly-confined particles in a Hele-Shaw cell, effectively assuming a separation of three length

    scales: the size of the swimmers, the depth of the confiningchamber, and the typical distance between two swimmers.Because of this separation of the three length scales, confine-ment profoundly modifies the interaction dynamics, which isdriven by the hydrodynamic and chemical fields screened bythe presence of the confining walls, while the individual self-propulsion remains essentially unchanged at leading order.This effective decoupling of self-propulsion and interactiondynamics, which are normally intrinsically linked as they arisefrom the same slip distribution at the particles’ surface, pro-vides some fundamental insight on the influence of boundarieson the latter.

    After reviewing the fundamental mechanisms of self-propulsion of spherical Janus particles, the hydrodynamicand chemical signatures of individual particles as well as theresulting drifts in external flows and concentration gradientsin Sec. II, we derive the leading order far-field hydrodynamicand chemical signatures of an individual particle in a Hele-Shaw environment in Sec. III. In both settings, the fundamentalscalings of these signatures and interaction drifts are clearlyidentified and demonstrate the fundamental role of confine-ment in setting these interactions. Section IV presents theequations of motion for N interacting particles, and Sec. Vfinally applies this model to the dynamics of a dilute sus-pension. We finally discuss our results and present someperspectives in Sec. VI.

    II. JANUS PARTICLES IN UNBOUNDED DOMAINSWe first analyse the individual motion, signature, and

    interactions of chemically-active spherical particles in theabsence of any confinement. The particles consideredthroughout this paper are spherical and chemically active withradius a. Their self-propulsion along self-generated gradientsof a physico-chemical field results from their polar chem-ical activity and their sensitivity to the field.28,38,39 In thefollowing, we focus for simplicity on neutral diffusiophore-sis for which the driving field is the concentration of a solutespecies produced or consumed at the surface of the particle.The principles and quantitative results can be easily adaptedto other phoretic mechanisms such as electrophoresis orthermophoresis.

    Before focusing on the interactions of such particles inconfined environments, we first review their self-propulsionand interactions in the canonical context of unbounded flows.Most of the analyses presented in this section therefore sum-marizes some classical results on the self-propulsion,38,39hydrodynamic coupling,50 and phoretic interactions30,31 ofsuch particles in bulk flows, and their derivation is briefly out-lined here so as to emphasize their physical origin and rela-tive scaling, which is critical to the later understanding of thescreening effects in Sec. III.

    The chemically-active spherical particles have radius a,and their physico-chemical properties are characterized bychemical activity A(xs) (i.e., their ability to modify the soluteconcentration) and mobility M(xs) (i.e., their ability to drive asurface slip flow from a local concentration gradient), whichdepend on the position xs along the surface of the particle.

    J. Chem. Phys. 150, 044902 (2019); doi: 10.1063/1.5065656 150, 044902-2

    Published under license by AIP Publishing

    https://scitation.org/journal/jcp

  • The Journal ofChemical Physics ARTICLE scitation.org/journal/jcp

    FIG. 1. Schematics of a single auto-phoretic Janus particle of radius a in uncon-fined 3D space, with self-propelled velocity Up, and a population of particles inweak Hele-Shaw confinement, where a � h � L. The unit vector p indicatingthe orientation of the Janus particle is directed orthogonally to the plane that delin-eates the two sides of the Janus particle, depicted in red and blue, respectively.Specifically, p is oriented from red to blue.

    Axisymmetric spherical particles are considered with an axisof symmetry p (see Fig. 1).

    A. Self-propulsion of isolated auto-phoretic particlesThe concentration field C(r) around the particle rela-

    tive to its background value satisfies the following system ofequations:

    D∇2C = 0, for r ≥ a, (1)

    C |r→∞ = 0, D(n · ∇C)��r=a = −A(µ). (2)

    Here, r is the position vector measured from the particlecenter, r = ‖r‖ is its magnitude, and µ = cos θ = p · r/rin spherical polar coordinates given by (r, θ). This prob-lem can be canonically solved for the concentrationfield38

    C(r,µ) =∞∑

    m=0

    aAmD(m + 1)

    (ar

    )m+1Lm(µ), (3)

    where Am = 2m+12 ∫1−1 A Lmdµ are the Legendre moments of the

    axisymmetric activity distribution A(µ), with Lm(µ) being themth Legendre polynomial.

    The gradient in chemical concentration induces a slipvelocity at the surface of the phoretic particle given by7

    uslip = M(µ)(I − nn) · ∇C |r=a, (4)

    where M(µ) is the axisymmetric particle mobility, I is the iden-tity tensor, and nn refers to the dyadic product of the vectorn with itself. This chemically-induced slip serves as a forcingboundary condition to solve for the Stokes flow around theforce-free and torque-free phoretic particle

    η∇2u − ∇p = 0, ∇ · u = 0, for r ≥ a, (5)

    subject to boundary conditions

    u |r→∞ = 0, u |r=a = U + uslip. (6)

    In Eq. (5), u is the fluid velocity field, p is the pressure field, ηis the dynamic viscosity, and U = Up is the swimming veloc-ity of the particle. The relative flow velocity at the surface ispurely tangential; therefore, the flow field u in the lab frame isgenerically given by51,52

    u = − 1r3∂ψ

    ∂µr − 1

    r(1 − µ2)∂ψ

    ∂r

    ( rrr2− I

    )· p, (7)

    where ψ is the stream function

    ψ(r,µ) =Ua2(1 − µ2)

    2r̄+∑n≥2

    (2n + 1)a2αn2n(n + 1)

    (1 − µ2)L′n(µ)[r̄−n − r2−n

    ],

    (8)

    with r̄ = r/a. The prime in L′n denotes the derivative withrespect to µ. Substituting Eq. (8) into Eq. (7), one gets that thefirst term in ψ corresponds to a source dipole [u ∼ (a/r)3 forr � a] and the term of order n in the infinite sum includesa potential singularity (gradient of a source dipole) and a vis-cous singularity (gradient of a Stokeslet). For example, n = 2includes a source quadrupole and force dipole, n = 3 a sourceoctopole and a force quadrupole, and so on. The intensityof these singularities is defined in terms of αn, given forn ≥ 2 by

    αn = −1

    2D

    ∞∑m,p=0

    AmMpm + 1

    ∫ 1−1

    (1 − µ2)L′mL′nLpdµ, (9)

    where Mp =2p+1

    2 ∫1−1 M Lpdµ are the Legendre moments of M(µ).

    The magnitude U of the swimming velocity is given by38

    U = −∞∑

    m=1

    mAm2m + 1

    (Mm−12m − 1 −

    Mm+12m + 3

    ). (10)

    The physico-chemical properties of the axisymmetricparticle are set by its activity and mobility distributions A(µ)and M(µ). For a generic particle, these two functions are arbi-trary (they can be positive or negative). A specific exampleis that of a hemispherical Janus particle, as shown in Fig. 1,with the chemical properties of the front half given by (Af ,Mf ) and of the back half given by (Ab, Mb). The first two Leg-endre moments of the activity and mobility distributions aregiven by A0 = (Af + Ab)/2, A1 = 3(Af − Ab)/4, and M0 = (Mf+ Mb)/2, M1 = 3(Mf −Mb)/4, respectively, whereas all non-zeroeven moments are identically zero, that is to say, A2n = M2n = 0for n , 0. The swimming velocity U and α2 can be foundanalytically38,40

    U = −A1M03D

    =(Ab − Af )(Mf + Mb)

    8D(11)

    and

    α2 = −16κA1M1

    9D= −

    κ(Mf −Mb)(Af − Ab)D

    , (12)

    with κ being a numerical constant, defined as

    κ =34

    ∞∑m=1

    2m + 1m + 1

    ∫ 10

    Lmdµ

    ∫ 10µ(1 − µ2)L′mdµ

    (13)

    such that κ ≈ 0.0872.

    B. Hydrodynamic and chemical signaturesThe chemical and hydrodynamic signatures of a Janus

    particle in an unbounded three-dimensional (3D) domain areobtained by keeping the two most dominant terms in Eqs. (3)and (7), respectively. The chemical signature consists of (i) a

    J. Chem. Phys. 150, 044902 (2019); doi: 10.1063/1.5065656 150, 044902-3

    Published under license by AIP Publishing

    https://scitation.org/journal/jcp

  • The Journal ofChemical Physics ARTICLE scitation.org/journal/jcp

    source of solute proportional to the net production rate A0and (ii) a source dipole proportional to A1

    C =1

    (4πa2A0

    D

    )1r

    +1

    (2πa3A1

    D

    ) (p · rr3

    )+ O

    (a3

    r3

    ). (14)

    The hydrodynamic signature consists of (i) a force dipole(or stresslet) proportional to α2, (ii) a source dipole pro-portional to U, and (iii) a force quadrupole proportional toα3

    u =1

    8πη∇

    (Ir

    +rrr3

    ): A − 1

    4π∇

    ( rr3

    )· B

    +1

    8πη∇∇

    (Ir

    +rrr3

    ) ... C + O(a4r4

    ), (15)

    where the coefficients are given by

    A =103πηa2α2(3pp − I), B = 2πa3Up, (16)

    C =7

    24πηa3α3

    [pI + (pI)T12 + Ip − 18ppp

    ]. (17)

    Here, (·)T12 is the transpose over the first two indices [i.e.,(C)T12ijk = Cjik].

    Note that we consistently carry the development ofC and u up to O(a3/r3) and O(a4/r4), respectively, sincethe chemical and hydrodynamic drifts on neighboringparticles will be driven by ∇C and u, respectively. The leading-order terms in the fluid velocity field u(r) and the chemi-cal concentration field C(r) are summarized in Table I. Wehighlight that the axisymmetric Janus particle swims in thep-direction at speed U given in Eq. (11), but it does not reori-ent under the influence of its own chemical activity. Theresults presented in Table I emphasize that the dominanthydrodynamic (force dipole) and chemical (source) interac-tions result in particle drift following the same algebraic decayin 1/r2 in the far-field limit. This underscores the necessity toaccount for both types of interactions, in contrast with sim-plifying assumptions regularly made in existing studies wherehydrodynamic interactions are neglected,53,54 mostly on thegrounds that the force dipole vanishes for hemispheric Janusparticles, which is only correct for the very specific caseof a uniform mobility,40 which is likely difficult to achieveexperimentally.

    C. Hydrodynamic and phoretic drift in externalflows and chemical gradients

    When the particle is placed in an external and possi-bly non-uniform flow u∞(r), it will translate and reorient. Fora spherical particle, the translation and rotational drifts aregiven at the leading order by Faxen’s laws50

    Ud = u∞(r0), Ωd =12ω∞(r0), (18)

    where ω∞ = ∇ × u∞ is the external vorticity field and r0 isthe position of the particle’s centroid. Non-spherical particlesare also sensitive to the local rate of strain due to their non-isotropy, and thin elongated particles would reorient in theprincipal strain direction.55

    The existence of an external non-uniform concentrationfield C∞(r), generated, for example, by the particle’s neighbors,induces an additional slip velocity on the particle’s boundarythat leads to a phoretic drift and reorientation.7,56 To computethe phoretic drift velocities, one needs to solve the followingsolute diffusion problem:

    D∇2Cd = 0, for r ≥ a, (19)

    Cd��r→∞ = C∞(r) ≈ C0∞ + G∞ · r +12H∞ : rr + . . . , (20)

    D(n · ∇Cd)��r=a = 0. (21)

    Here, we expanded the chemical field C∞ in a Taylor seriesabout the particle position r0, with G∞ = ∇C∞ |r0 and H∞= ∇∇C∞ |r0 evaluated at the particle location. The solution toEq. (19) is uniquely obtained as57

    Cd(r) = C0∞ + (G∞ · r)[1 +

    12

    (ar

    )3]+

    12

    (H∞ : rr)[1 +

    23

    (ar

    )5]+ · · · . (22)

    The additional slip velocity on the boundary of the particle isgiven by

    uslip = M(r)(I − nn) · ∇Cd

    = M(r)(I − nn) ·[32G∞ +

    53H∞ · n

    ]+ · · · . (23)

    TABLE I. Scaling laws of the chemical and hydrodynamic fields created by an isolated self-propelled phoretic particle in anunbounded 3D domain based on Eqs. (14) and (15), respectively. Here, A0, A1 and M0, M1 represent the surface activity andmotility of the particle, respectively, D is the molecular diffusivity of the solute, a is the particle size, and r is the distance fromthe particle at which these fields are evaluated.

    Hydrodynamic signature Chemical signature

    u(r) C(r)

    Force dipole Source dipole Force quadrupole Source Source dipole

    Intensity a2A1M1

    Da3

    A1M0D

    a3α3 a2A0D

    a3A1D

    Decay rate1

    r21

    r31

    r31r

    1r2

    J. Chem. Phys. 150, 044902 (2019); doi: 10.1063/1.5065656 150, 044902-4

    Published under license by AIP Publishing

    https://scitation.org/journal/jcp

  • The Journal ofChemical Physics ARTICLE scitation.org/journal/jcp

    TABLE II. Scaling laws of the chemical and hydrodynamic drift in an unbounded 3D domain created by a phoretic particle ata distance L from the particle of interest.

    Hydrodynamic drift Chemical drift

    Force dipole Source dipole Force quadrupole Source Source dipole

    Translationala2

    L2A1M1

    Da3

    L3A1M0

    Da3α3

    L3a2

    L2A0M0

    Da3

    L3A1M0

    Ddrift Ud

    Rotational1L

    a2

    L2A1M1

    D1L

    a3

    L3A1M0

    D1L

    a3α3L3

    1a

    a2

    L2A0M1

    D1a

    a3

    L3A1M1

    Ddrift Ωd

    When the background concentration field C∞ is generated byother particles located at a distance L far enough from the par-ticle of interest, the contribution of the second term involvingH∞ = ∇∇C∞ |r0 is subdominant by a factor a/L; we thus omit itin the following analysis.

    Determining the translation and rotation velocities of aforce- and torque-free particle for a given prescribed slip dis-tribution is a now-classical linear fluid dynamics problem.58For a spherical particle, the reciprocal theorem for Stokes’flow around an isolated sphere provides the drift translationaland rotational velocities as7

    Ucd = −〈uslip(r)

    〉= −3G∞

    2· 〈M(r)(I − nn)〉, (24)

    Ωcd = −3

    2a

    〈n × uslip(r)

    〉= − 9

    4a〈M(r)n

    〉 ×G∞, (25)where 〈. . .〉 denotes the spatial average taken over the parti-cle’s surface. These expressions can be simplified further foraxisymmetric particles noting that31

    〈M(r)nn〉 = M03

    I +M215

    (3pp − I), (26)

    〈M(r)n〉 = M1p3

    (27)

    so that the translational and rotational drift velocities canfinally be rewritten as (with G∞ = ∇C∞ |r0 )

    Ucd = −M0∇C∞ |r0 +M210

    (3pp − I) · ∇C∞ |r0 , (28)

    Ωcd = −3M14a

    p × ∇C∞ |r0 . (29)

    For the particular case of a hemispheric Janus particle, alleven modes A2n and M2n (for n > 0) are zero due to symmetry,and one finally obtains

    Ucd = −M0∇C∞ |r0 , Ωcd = −

    3M14a

    p × ∇C∞ |r0 . (30)

    The phoretic interactions include a reorientation to alignthe particle with (respectively, against) the chemical gradientif M1 < 0 (respectively, M1 > 0) [Eq. (29)], as well as a transla-tional drift with components both along the chemical gradientand in the particle’s direction [Eq. (28)]. These different contri-butions provide different modes of chemotaxis to the catalyticcolloids that were recently analyzed in detail.30,31 The trans-lational velocity scales as Ucd ∼ M0G and the reorientation

    velocity scales as Ωcd ∼ M1G/a, where G is the characteristicsolute concentration gradient created by the other particlesat the location of the particle considered. The reorientationtime scale is τcrotation = 1/Ω

    cd ∼ a/M1G.

    We now briefly discuss the relative magnitude of thesechemical and hydrodynamic effects. When the backgroundchemical gradient experienced by a given particle is dueto a second particle located at a distance L, this gradi-ent scales as G ∼ (A0/D)(a/L)2 at leading order, if theseparticles are net sources or sinks of solute (A0 , 0) [seeEq. (14)]. Consequently, the chemical drift velocity scales asUcd ∼ (M0A0/D)(a/L)

    2 and the reorientation time scale isτcrotation = 1/Ω

    cd ∼ a(D/A0M1)(L/a)

    2. If the net productionrate vanishes (A0 = 0), the gradients are weaker, scaling asG ∼ (A1/D)(a/L)3, and therefore Ucd ∼ (M0A1/D)(a/L)

    3 andτcrotation ∼ a(D/A1M1)(L/a)

    3.These scalings for chemical interactions between parti-

    cles can be compared to their hydrodynamic counterparts.The leading order translational hydrodynamic drift scales asUhd ∼ (M1A1/D)(a/L)

    2 [see Eq. (15)], and the reorientation timescales as τhrotation ∼ L(D/A1M1/)(L/a)

    2. A summary of the scalingof the translational and rotational velocities due to hydrody-namic and phoretic drifts is given in Table II.

    When the particles act as net sources or sinks of solute(A0 , 0), the chemical and hydrodynamic drift velocities areof the same order Ucd ∼ U

    hd, but chemical rotations act much

    faster than hydrodynamic rotations τcrotation � τhrotation. When

    the chemical signature of the particle is a source dipole only(A0 = 0), the hydrodynamic drift is dominant Ucd � U

    hd, whereas

    chemical and hydrodynamic rotations are of the same orderτcrotation ∼ τ

    hrotation.

    III. WEAKLY CONFINED JANUS PARTICLESIN HELE-SHAW CELLS

    Section II established that in an unbounded domain (i) thedynamics of individual Janus particles is the superposition oftheir self-propulsion and the translational and rotational driftsassociated with the background concentration of solute (thelatter is not related to their activity) and (ii) only the latterplays a role in their re-orientation when the particles are netsources or sinks of solute.

    J. Chem. Phys. 150, 044902 (2019); doi: 10.1063/1.5065656 150, 044902-5

    Published under license by AIP Publishing

    https://scitation.org/journal/jcp

  • The Journal ofChemical Physics ARTICLE scitation.org/journal/jcp

    We investigate here how these results and associatedscalings are modified when the Janus particles are confinedin a Hele-Shaw cell consisting of two no-slip walls that arechemically-inert, with no activity or mobility, and separatedby a distance h [Fig. 1(b)]. We consider the joint limit of (i) weakconfinement, where the particle radius a is much smaller thanthe gap h between the walls, namely, a � h, and (ii) Hele-Shawinteractions, where the typical distance L between particles ismuch greater than the cell depth leading to dilute suspensionswith h � L. The first condition implies that the chemical andhydrodynamic drift of a particle in response to the local chem-ical and hydrodynamic fields are obtained in the same wayas those of a particle in unbounded flow. The second condi-tion is important to determine these local fields, given a dilutesuspension of particles in the Hele-Shaw cell, as discussednext.

    A. Chemical signature in Hele-Shaw confinementWithin the Hele-Shaw cell and outside of the individual

    particles, the solute concentration is governed by Laplace’sequation with no-flux condition at the confining walls

    D∇2C = 0, with ∂C∂z

    �����z=±h/2= 0. (31)

    Here, we consider the Cartesian coordinates (x, y, z) suchthat the (x, y)-plane is located in the middle of the Hele-Shaw cell, parallel to the bounding walls, and z is along itsdepth, and we introduce the corresponding inertial frame(ex, ey, ez) as depicted in Fig. 1. We non-dimensionalize thelength in the z-direction by h and in the (x, y)-plane by L with� = h/L � 1, and we seek a solution to Eq. (31) in this form:C = C0 + �2C1 + · · · . We immediately obtain from Eq. (31)that C0 is independent of z. That is to say, at the leadingorder, the solute concentration is homogeneous across thechannel depth and its leading order gradient is horizontal. Adirect but fundamental consequence of this homogeneity isthat all Janus particles will be forced to align parallel to the (x,y)-plane, which justifies a quasi-2D approach to the presentproblem. The presence of a net source of solute in 2D con-finement would induce a logarithmic far-field singular behav-ior in the concentration field. For a well-posed problem, oneneeds either to consider two different populations of parti-cles (net producers and net consumers in similar proportions)or a single population of dipolar particles. We focus on thelatter.

    We consider the solute concentration field generated bya single Janus particle located at r0 and oriented horizontallysuch that p · ez = 0. In an unbounded flow domain, the con-centration generated by this chemical dipole is obtained bysetting A0 = 0 in Eq. (14), which we rewrite as

    C =1

    (2πa3A1

    D

    )p · (r − r0)‖r − r0 ‖3

    , (32)

    where 2πa3A1/D is the intensity of the chemical dipole inthe unbounded domain. When the particle is located withina Hele-Shaw channel, at r0 = δez, where δ is the vertical posi-tion of the particle relative to the central plane (|δ| < h/2), the

    concentration within the channel gap can be reconstructedusing the method of images by superimposing identical dipoleslocated at vertical positions z+n = δ + 2nh and z−n = −δ + (2n + 1)h(n ∈ Z). To this end, we write r = x + zez, where x = xex + yey.The concentration field due to the chemical dipole and itsinfinite image system is given by

    C =1

    (2πa3A1

    D

    )(p · x)

    ×∞∑

    n=−∞

    1

    ‖r − z+nez ‖3/2

    +1

    ‖r − z−nez ‖3/2· (33)

    In the Hele-Shaw limit ‖x‖ � h, the leading order concentra-tion field is obtained using Riemann’s sum as

    CH−S =1

    (2πa3A1

    D

    )(p · x)h‖x‖2

    ∞∑m=−∞

    h/‖x‖1 +

    (mh‖x‖

    )23/2

    =1

    (2πa3A1

    Dh

    )(p · x)‖x‖2

    · (34)

    This concentration field corresponds to a two-dimensionalchemical dipole of intensity 2πa3A1/hD. It is homogeneouswithin the channel depth, i.e., it is independent of the particle’svertical position δ.

    B. Hydrodynamic signature in Hele-Shawconfinement

    The far-field hydrodynamic signature of a Janus particlein an unbounded fluid domain is given by Eq. (15). The lead-ing order term of the velocity field is that of a force dipole (orStokeslet dipole) whose velocity field decays as 1/r2. The dom-inant correction to the leading order term includes a potentialhorizontal source dipole and force quadrupole, both decayingas 1/r3.

    We now consider a Janus particle that is confinedbetween the two no-slip surfaces. We are interested in itshydrodynamic signature in the (far-field) Hele-Shaw limit, atdistances L much greater than the channel depth h. By lin-earity of the Stokes equations, the velocity field produced bythe confined particle is the sum of the velocity fields pro-duced by the confined singularities: force dipole, potentialsource dipole, and force quadrupole. The effect of confine-ment between two rigid walls on a force singularity (Stokeslet)is analyzed at length by Liron and Mochon.59 They showedthat in the far-field (L � h), a Stokeslet oriented along thehorizontal direction (parallel to the confining walls) inducesan exponentially-decaying velocity field in the z-direction,whereas in the (x, y)-plane, its dominant behavior correspondsto a two-dimensional source dipole. The direction of thesource dipole is that of the original Stokeslet, and its strengthdepends in a parabolic way on its placement between the twowalls. Mathematically, for a Stokeslet of unit strength locatedin a plane z = δ between the two walls such that

    ust =1

    8πη(Ir

    +rrr3

    ) · p, (35)

    J. Chem. Phys. 150, 044902 (2019); doi: 10.1063/1.5065656 150, 044902-6

    Published under license by AIP Publishing

    https://scitation.org/journal/jcp

  • The Journal ofChemical Physics ARTICLE scitation.org/journal/jcp

    the leading order far-field flow in Hele-Shaw confinement isgiven by [see Eq. (51) in Ref. 59]

    uH−Sst = −3h

    2πη

    (14− δ

    2

    h2

    ) (14− z

    2

    h2

    ) (I‖x‖2

    − 2 xx‖x‖4

    )· p. (36)

    Therefore, the far-field flow generated by a horizontal forcedipole corresponding to the first term in Eq. (15),

    ufd =1

    8πη∇

    (Ir

    +rrr3

    ): A, (37)

    becomes, when confined between two walls, that of a two-dimensional source quadrupole decaying as 1/‖x‖3

    uH−Sfd = −3h

    2πη

    (14− δ

    2

    h2

    ) (14− z

    2

    h2

    )× ∇

    (I‖x‖2

    − 2 xx‖x‖4

    ): A (38)

    where A = 10πηa2α2(pp − I/3) [see Eq. (16)] and α2 is given inEq. (12). Substituting into Eq. (38) and evaluating the resultingexpression at δ = z = 0, the leading order flow field generatedby a horizontal force dipole is obtained as

    uH−Sfd =5ha2κA1M1

    9D

    [∇

    (I‖x‖2

    − 2 xx‖x‖4

    )]: (3pp − I). (39)

    We now examine the effect of confinement on the flowgenerated by a 3D source dipole. We recall that the sourcedipole can be seen either as a potential flow solution to theStokes equations, associated with Laplace’s equation,60 or asa degenerate force quadrupole. Following the latter approach,the source dipole contribution in the unconfined domain cor-responding to the second term in Eq. (15) can be rewritten asthe Laplacian of a Stokeslet

    usd = −1

    4π∇( r

    r3

    )· B = − 1

    8π∇2

    (Ir

    +rrr3

    )· B, (40)

    where B = 2πa3Up is given by Eq. (16) and the last equality fol-lows directly by differentiation [e.g., Kim and Karrila,50 Chap.2, Eq. (2.12)]. The leading order velocity field associated withthe confined potential singularity is then obtained by consid-ering the dominant contribution to the Laplacian of Eq. (36)and is given by

    uH−Ssd = −3πh

    (14− δ

    2

    h2

    ) (I‖x‖2

    − 2 xx‖x‖4

    )· B. (41)

    Evaluating at δ = 0 and substituting the expression for U fromEq. (11), we get that

    uH−Ssd = −3a3U

    2h

    (I‖x‖2

    − 2 xx‖x‖4

    )· p

    =a3A1M0

    2hD

    (I‖x‖2

    − 2 xx‖x‖4

    )· p. (42)

    It is important to note that in unbounded 3D domains, thecontribution of the source dipole was a higher order correc-tion to the contribution of the force dipole. The situation isreversed in Hele-Shaw confinement: the source dipole decaysas 1/‖x‖2, whereas the force dipole decays as 1/‖x‖3. Mean-while, the force quadrupole decays as 1/‖x‖4. Ignoring thelatter, the leading-order terms in the flow field created bya self-propelled phoretic particle placed horizontally in themid-plane of the channel in the Hele-Shaw limit is computedby substituting Eqs. (39) and (42) into Eq. (15)

    uH−S =a3A1M0

    2hD

    (I‖x‖2

    − 2 xx‖x‖4

    )· p

    +5ha2κA1M1

    9D

    [∇

    (I‖x‖2

    − 2 xx‖x‖4

    )]: (3pp − I). (43)

    The leading-order terms in the hydrodynamic and chem-ical signatures, uH−S(x) and CH−S(x), of a self-propelledphoretic particle in the Hele-Shaw limit are summarized inTable III.

    C. Hydrodynamic and phoretic interactionsunder weak confinement

    Under weak confinement a � h, the hydrodynamic andphoretic drifts of individual particles—which result from thechemical and hydrodynamic fields immediately around it—canbe determined as in the unbounded case. The hydrodynamicand phoretic drift velocities are therefore given by Eqs. (18)and (30), respectively, where u∞ and ∇C∞ are the velocity fieldand concentration gradient created by other phoretic parti-cles, respectively, and should now be evaluated for uH−S(x)and CH−S(x) in the Hele-Shaw limit obtained in Eqs. (34)and (43).

    Using Eqs. (30) and (34), the phoretic drifts of a particlewith orientation p0 created by a particle located at relative

    TABLE III. Scaling laws of the chemical and hydrodynamic signatures induced by a confined self-propelled phoretic particlein a Hele-Shaw cell of width h such that a � h and h � L. Parameter values are defined in Table I.

    Hydrodynamic signature Chemical signature

    uH−S(x) CH−S(x)

    Force dipole Source dipole Force quadrupole Source Source dipole

    Intensity ha2A1M1

    Da3

    hA1M0

    Da3α3

    h. . .

    a3

    hA1D

    Decay rate1

    L31

    L21

    L4. . .

    1L

    J. Chem. Phys. 150, 044902 (2019); doi: 10.1063/1.5065656 150, 044902-7

    Published under license by AIP Publishing

    https://scitation.org/journal/jcp

  • The Journal ofChemical Physics ARTICLE scitation.org/journal/jcp

    TABLE IV. Scaling laws of the chemical and hydrodynamic drift under confinement created by a phoretic particle at a distanceL from the particle of interest.

    Hydrodynamic drift Chemical drift

    Force dipole Source dipole Force quadrupole Source Source dipole

    Translationalha2

    L3A1M1

    Da3

    hL2A1M0

    Da3α3hL4

    . . .a3

    hL2A1M0

    Ddrift Ud

    Rotational . . . . . . . . . . . .1a

    a3

    hL2A1M1

    Ddrift Ωd

    position x with orientation p are given by

    Ucd = −a3M0A1

    Dh

    [I‖x‖2

    − 2xx‖x‖4

    ]· p, (44)

    Ωcd = −3a2A1M1

    4Dhp0 ×

    [(I‖x‖2

    − 2xx‖x‖4

    )· p

    ], (45)

    and, respectively, scale as Ucd ∼ (a3/hL2)(A1M0/D) and Ωcd

    ∼ (a3/hL2)(M1A1/aD). Similarly, the velocity of translationaldrift due to hydrodynamic interactions is computed fromEqs. (18) and (43)

    Uhd =a3A1M0

    2Dh

    (I‖x‖2

    − 2xx‖x‖4

    )· p

    +5ha2κA1M1

    9D

    [∇

    (I‖x‖2

    − 2xx‖x‖4

    )]: (3pp − I). (46)

    For spherical particles, the rotational hydrodynamic driftarises only from the vorticity field, as expressed explicitly inEq. (18). Since the flow field in Eq. (43) is potential, the vorticityfield is identically zero, and therefore there is no hydrody-namic rotational drift in the Hele-Shaw limit. A summary ofthe phoretic and hydrodynamic drift velocities in this limit isgiven in Table IV.

    A direct comparison of the translational drift velocities Ucdand Uhd shows that the contributions of the hydrodynamic andchemical source dipoles follow the same scaling, while thatdue to the hydrodynamic force quadrupole is always subdom-inant (see Table IV). Comparing the velocities due to hydro-dynamic and chemical source dipoles to that induced by theforce dipole, the following three regimes can arise depend-ing on the relative magnitude of a/h versus h/L or on therelative magnitude of the range of the interactions L versush2/a:

    • For far-field interactions L � h2/a, the translationaldrift velocity Ud is governed at the leading order bythe hydrodynamic and chemical source dipoles. Theinfluence of the force dipole is negligible.

    • For short-range interactions L � h2/a, the translationaldrift due to the hydrodynamic force dipole is dominant.The chemical drift is negligible and so is the role of thehydrodynamic source dipole.

    • When L ∼ h2/a, the drifts induced by the hydrodynamicforce dipole and source dipole and the chemical sourcedipole are all of the same order.

    Effectively, if the particles’ density is small enough (dilutesystems) that the typical distance between particles satisfiesL � h2/a, then the leading-order interactions of the differ-ent particles can be written solely in terms of source dipoles.Higher particle densities require more care to account for theeffect of the force dipole. In all regimes, one can readily ver-ify that the rotational drift due to the chemical source dipoleis always dominant. Starting from this insight, we next formu-late equations of motion governing the far-field interaction ofmultiple phoretic Janus particles in each of these regimes.

    IV. FAR-FIELD INTERACTIONS OF CONFINEDAUTO-PHORETIC PARTICLES

    Based on the results obtained in Sec. III, we formulate aself-consistent description for the dynamics of N autophoreticparticles in weak confinement (a � h) and dilute suspensions(h � L).

    Let particle j be located at xj with orientation given bythe unit vector pj (j = 1, . . ., N). For simplicity, we considerthat all particles are located on the midplane of the chan-nel (δ = 0). The translational motion of particle j is due (i) toits self-propulsion at speed U as a result of its own chemi-cal activity and mobility property, (ii) the hydrodynamic driftgenerated by the motion of its neighbors, and (iii) the chemi-cal drift resulting from the particle’s own mobility in responseto the chemical activity of its neighbors. The particle’s orien-tation pj also changes in response to these flow and chemicaldisturbances. Given a � h, the resulting equations of motionof particle j are given at the leading order by the particle’sbehavior in unbounded flows

    ẋj = Upj + u(xj) + µc∇C(xj), (47)

    ṗj = (I − pjpj) ·[w(xj) · pj + νc∇C(xj)

    ]. (48)

    Here, U is the self-propulsion velocity given in Eq. (11), µc andνc are the chemical translational mobility coefficients, and w= (∇u − ∇uT)/2 is the anti-symmetric (vorticity) componentof the local velocity gradient. In Sec. II, we obtained that, fora spherical Janus particle, U, µc, and νc are given by Eq. (30),which we rewrite for convenience as

    J. Chem. Phys. 150, 044902 (2019); doi: 10.1063/1.5065656 150, 044902-8

    Published under license by AIP Publishing

    https://scitation.org/journal/jcp

  • The Journal ofChemical Physics ARTICLE scitation.org/journal/jcp

    U = −A1M03D

    , µc = −M0, νc = −3M14a

    . (49)

    In Sec. III, we found that the rotational drift is alwaysdominated by the chemical component. Therefore, in the fol-lowing, we will neglect the hydrodynamic term in the orien-tation equation. We also obtained expressions for the localhydrodynamic and chemical fields generated by other par-ticles. Specifically, the chemical concentration C is given byEq. (34) and the hydrodynamic flow field u by Eq. (43). Here,we drop the H-S superscript with the understanding that allquantities are in Hele-Shaw confinement. Put together, we getthat, at the leading order,

    C(x) =a3A1Dh

    ∑k

    pk · (x − xk)‖x − xk ‖2

    , (50)

    ∇C(x) = −a3A1Dh

    ∑k

    G(x, xk) · pk, (51)

    and

    u(x) = − a3A1M02Dh

    ∑k

    G(x, xk) · pk

    − 5ha2κA1M19

    ∑k

    ∇G(x, xk) : (3pkpk − I), (52)

    with

    G(x, xk) =2(x − xk)(x − xk)‖x − xk ‖4

    − I‖x − xk ‖2

    · (53)

    We evaluate the above expressions at particle j andsubstitute the result into Eqs. (47) and (48), noting that the sec-ond term in the hydrodynamic flow field [Eq. (52)] should onlybe included for particles when |(xj − xk)| / h2/a. For dilute sus-pensions, where the volume fraction of particles is such that|(xj − xk)| � h2/a, the contribution of the force dipole is sub-dominant. In this case, we note that the hydrodynamic andchemical translational drifts on particle j take the exact sameform given by u(xj) and −M0∇C(xj) = −2u(xj), but act in oppositedirections, with the latter being dominant (and exactly twiceas large). The equations of motion for the phoretic particlesthen simplify to

    ẋj = −A1M0

    3Dpj +

    a3A1M02Dh

    ∑k,j

    G(xj, xk) · pk, (54)

    ṗj =3a2M1A1

    4Dh(I − pjpj)

    ∑k,j

    G(xj, xk) · pk. (55)

    We note that these equations take a particularly simple form:they are equivalent to the interaction of hydrodynamic sourcedipoles with a reversed hydrodynamic drift, leading to a nega-tive effective hydrodynamic mobility coefficient. That is to say,particles tend to drift in the opposite direction of the local flowdue to the dominance of the chemical drift.

    It is convenient to rewrite Eqs. (54) and (55) in non-dimensional form. To this end, we assume without anyloss of generality that A1M0 > 0 so that the direction of

    self-propulsion is q = −p (the case A1M0 < 0 could be treatedsimilarly); we then use the characteristic length scale a anda characteristic time scale 3Da/A1M0 based on the nominalswimming velocity. Equations (54) and (55) become

    ẋj = qj + µ∑k,j

    G(xj, xk) · qk, (56)

    q̇j = ν(I − qjqj)∑k,j

    G(xj, xk) · qk, (57)

    respectively, where µ = −3a/2h and ν = 9aM1/4hM0 are thetranslational and rotational mobility, respectively, with theunderstanding that all variables in Eqs. (56) and (57) are non-dimensional (here and thereafter µ is the translational mobilitycoefficient not to be confused with µ = cos θ in Sec. II). Froma chemical point of view, the sign of ν is directly related to thereorientation of the particle in the direction of or opposed toits chemical drift.

    It should be noted here that the mobility coefficientµ = −3a/2h is negative. That is, in contrast with the standardmicro-swimmers,61–64 weakly confined phoretic particles driftin the opposite direction to the local flow induced by otherparticles, instead following their chemical drift. This, in turn, isthe result of phoretic and hydrodynamics interactions havingthe same dependence but opposite behavior in this Hele-Shawlimit.

    V. COLLECTIVE DYNAMICS AND PAIR INTERACTIONSIn Sec. IV, the leading order dynamics for phoretic par-

    ticles in weak Hele-Shaw confinement was shown to takethe form of interacting potential dipoles, similarly to othercategories of confined micro-swimmers, but for a reversetranslational mobility.

    A. Suspension dynamicsTo analyze the interplay between this negative transla-

    tional motility and the orientational dynamics within a suspen-sion of phoretic particles, we numerically solve Eqs. (56) and(57) for a population of particles in a doubly-periodic domain.We account for the doubly infinite system of images using theWeiestrass-zeta function.62,65,66 To avoid collision, we intro-duce a Lennard-Jones repulsion potential to the translationalequation [Eq. (56)].62 The resulting steric forces act locallyand decay rapidly such that they do not affect the long-rangechemical and hydrodynamic interactions among the particles.We use a standard time stepping algorithm to solve for theevolution of a population of N = 100 particles that are ini-tially spatially distributed at random orientations in a doubly-periodic square domain of size D/a = 19.5.62 The aspect ratiois fixed to h/a = 5 so that the translational mobility coeffi-cient is µ = −0.3 and the orientational coefficient is varied,ν ∈ [−5, 5].

    Three distinct types of global behaviors emerge depend-ing on the sign of the rotational mobility ν: (i) a swirlingbehavior where particles form transient chains that emerge,break, and rearrange elsewhere for ν > 0, (ii) random particlemotions for ν = 0, and (iii) aggregation and clustering for ν < 0.

    J. Chem. Phys. 150, 044902 (2019); doi: 10.1063/1.5065656 150, 044902-9

    Published under license by AIP Publishing

    https://scitation.org/journal/jcp

  • The Journal ofChemical Physics ARTICLE scitation.org/journal/jcp

    Representative simulations are shown in Fig. 2. These col-lective phenomena are reminiscent to those observed in themotion of hydrodynamic dipoles,62,67 but bear distinctive fea-tures due to the negative translational motility, as discussednext.

    The large-scale phenomena are dictated by the orienta-tional dynamics of the particles. The swirling phenomenon canbe readily understood by examining the orientational inter-action of two particles, as depicted schematically in Fig. 3(a).For ν > 0, a particle aligns with the drift created by anotherparticle. That is to say, it reorients towards the tangent tothe streamlines of the potential dipole created by the nearbyparticle, causing it naturally to follow that particle. As aresult, particles of positive ν tend to form chains. However,given the negative motility coefficient µ < 0, as particlesalign into such chain-like structures, their forward transla-tional motion is slowed down by the dipolar flow field oftheir neighbors, resulting in a decrease in the particles’ veloci-ties, as evidenced from the probability distribution function inFig. 2(a).

    When the orientational interactions are suppressed(ν = 0), particles do not change orientation, except to avoidcollisions when other particles are sufficiently close to trig-ger steric interactions. In other words, the long-range inter-actions among particles can at most slow down the particletranslational velocities due to the negative motility coefficient,without introducing bias in the particles’ orientation and posi-tion relative to each other. The reduction in the translational

    FIG. 3. Schematic depiction of the orientation interaction of two phoretic particles:(a) for ν > 0, a particle reorients to align with the dipolar flow field created by theother particle, leading the two particles to “tail gate” each other; (b) for ν < 0, itreorients opposite to the dipolar field created by the other particle, leading the twoparticles to aggregate head on.

    velocity depends on the particles’ location, which is initiallyrandom and remains random under subsequent interactionsbetween particles [see Fig. 2(b)]. As a result, the translational

    FIG. 2. Collective behavior of auto-phoretic particles in doubly periodic domain exhibiting (a) chain-like formation and swirling behavior for ν = 5; (b) advection and stericinteraction for ν = 0; and (c) aggregation for ν = −5. Snapshots are shown at t = 1750, 1000, and 15, respectively. The unit vector p is directed from red to blue, as indicatedin Fig. 1. Parameter values are set to N = 100, U = 1, a = 1, and µ = −0.3. The domain size is D = 19.5. The bottom row shows the probability distributions of the particlevelocities in the three cases. In all cases, the average velocity is smaller than the self-propelled velocity of an individual particle U = 1.

    J. Chem. Phys. 150, 044902 (2019); doi: 10.1063/1.5065656 150, 044902-10

    Published under license by AIP Publishing

    https://scitation.org/journal/jcp

  • The Journal ofChemical Physics ARTICLE scitation.org/journal/jcp

    velocities follow a somewhat broad-band distribution of valuesbelow U = 1 (i.e., their self-propulsion velocity), as depicted inthe bottom row of Fig. 2(b).

    The clustering behavior observed for ν < 0 can beexplained by recalling that, in that case, particles reorient inthe opposite direction of the local drift resulting from thechemical and hydrodynamic interactions with the other par-ticles. Therefore, a particle travels in the opposite directionto the streamlines created by a nearby particle, as illustratedin Fig. 3(b), which leads to aggregation. The particles beginto aggregate at a relatively short time scale, as indicated bythe time of the snapshot in Fig. 2(c), and lead to the formationof clusters. Some of these clusters are unstable, while othersform stable aggregates that attract each other and slow downconsiderably as they coalesce to form larger aggregates. Thevelocity distribution function on the bottom row of Fig. 2(c)has a strong peak around zero velocity because most particlesare attached to stable clusters, with a few particles moving atunit speed.

    B. Pair interactions and stabilityThe orientational interactions underlying these large-

    scale phenomena can be examined analytically in more detailin the special case of pairs of phoretic particles. In particular,the observations that for ν > 0, particles tend to follow eachother and form quasi-stable chainlike structures, whereas forν < 0, particles tend to collide head-on and form clustersare rooted in two-particle interactions: for ν < 0, chainlikestructures are unstable while side-by-side motions that leadto head-on collisions are unstable for ν > 0, as discussednext.

    First, we rewrite Eqs. (56) and (57) for two particles to getthat

    ẋ1 = q1 +µ

    d2

    (2dd − I

    d2

    )· q2, (58)

    ẋ2 = q2 +µ

    d2

    (2ddd2− I

    )· q1, (59)

    q̇1 =ν

    d2(1 − q1q1) ·

    (2ddd2− I

    )· q2, (60)

    q̇2 =ν

    d2(1 − q2q2) ·

    (2ddd2− I

    )· q1, (61)

    where d = x2 − x1 is the relative position vector and d = ‖d‖ isthe relative distance between the two particles. We reformu-late these equations in terms of d and the relative orientationangles αj of qj (j = 1, 2) with respect to d. To this end, we intro-duce the unit vector e = d/d ≡ (cos θ, sin θ), where θ is theorientation of d in the fixed inertial frame. The relative anglesαj of qj with respect to e satisfy the identities e · qj = cosαj ande × qj = sinαj. The global translational velocity of the two par-ticles can be omitted here due to the fact that the system isinvariant under translational symmetry. Expressing Eq. (61) interms of (d, θ, α1, α2), we get that

    ḋ =(1 − µ

    d2

    )(cosα2 − cosα1), (62)

    θ̇ =

    (1d

    d3

    )(sinα2 − sinα1), (63)

    α̇1 = α̇2 = −ν

    d2sin(α1 + α2) −

    (1d

    d3

    )(sinα2 − sinα1). (64)

    This leads to the surprising result that the relative orientationα2 − α1 of the two swimmers is a constant of motion. Further-more, θ does not influence (d, α1, α2) because the system isinvariant under rotational symmetry; we could thus solve for(d, α1, α2) independently. It is more convenient to introduceδ = (α2 − α1)/2 and γ = (α1 + α2)/2 such that

    ḋ =(1 − µ

    d2

    )sinγ sin δ, (65)

    δ̇ = 0, (66)

    γ̇ = − νd2

    sin 2γ − 2(

    1d

    d3

    )cosγ sin δ. (67)

    There are two configurations that lead to relative equilibria ofthis system of equations (δ̇ = γ̇ = 0): (i) follower configuration:δ = 0, γ = 0, and (ii) Side-by-side configuration: δ = 0, γ = ±π/2,as depicted in Fig. 4.

    In the follower configuration, both particles have the sameorientation aligned with their relative distance e. The solu-tion corresponds to a translation of the two particles at thesame velocity (1 + 2µ/d2o)e, where do is constant. LinearizingEq. (67) around the equilibrium (do, 0, 0) provides at leadingorder

    ḋ = 0, δ̇ = 0, γ̇ = −2 νd2oγ − 2*

    ,

    1do

    d3o+-δ, (68)

    and this equilibrium is linearly unstable for ν < 0. For ν ≥ 0,the system is neutrally stable and weakly nonlinear analysisshould be performed to analyze its stability. These findingsare consistent with the observations that chainlike structures,which are reminiscent of the follower configuration, are not

    FIG. 4. Relative equilibria of pairs of particles in (a) follower configuration and (b)side-by-side configuration, both at a constant separation distance do. The followerconfiguration is unstable for ν < 0, while the side-by-side configuration is unstablefor ν > 0.

    J. Chem. Phys. 150, 044902 (2019); doi: 10.1063/1.5065656 150, 044902-11

    Published under license by AIP Publishing

    https://scitation.org/journal/jcp

  • The Journal ofChemical Physics ARTICLE scitation.org/journal/jcp

    observed for ν < 0. Physically speaking, the follower configu-ration is unstable for ν < 0 because particles tend to align inthe opposite direction to the ambient flow. Therefore, a slightperturbation away from the equilibrium, say of the followerparticle, would cause it to move opposite to the streamlinesof the leader particle, which drives it further away from theequilibrium.

    In the side-by-side configuration, both particles are paral-lel to each other and perpendicular to e. They undergo a trans-lational motion with velocity (1 − µ/d2o)e, where do is constant.Linearizing around (do, 0, ±π/2) provides at leading order

    ḋ = ±(1 − µ

    d2o

    )δ, δ̇ = 0, γ̇ = 2

    ν

    d2oγ, (69)

    and this equilibrium is linearly unstable for ν > 0 and neu-trally stable otherwise. These results are consistent with theintuitive analysis presented in Fig. 3. Side-by-side particlespointing in the same direction are unstable for ν > 0 becausethey will want to reorient in the direction of the dipolar fieldof the other particle.

    Taken together, the results from the numerical simula-tions for a population of weakly-confined particles and thelinear stability analysis for a pair of particles indicate that thecollective behavior is dominated by the orientational dynam-ics and by the orientational mobility coefficient ν. The negativetranslational mobility coefficient slows down the particles butdoes not affect the modes of collective behavior.

    VI. CONCLUSIONS

    In this work, a first-principle approach was presented forderiving the equations governing the interactions of a popu-lation of auto-phoretic Janus particles under weak Hele-Shawconfinement. The goal was to analyse the effects of confine-ment on the interactions of many phoretic particles, and, inparticular, on the relative weight of the hydrodynamic andchemical (phoretic) couplings. Both effects take the same formwithin this particular limit, but act in opposite directions, withthe magnitude of phoretic interactions being exactly twiceas large and, therefore, driving the collective dynamics ofweakly-confined Janus particles. Yet, hydrodynamic interac-tions are not negligible and in fact account for a reduction by50% in the effective translational and rotational drifts whencompared to pure phoretic interactions. This analysis furtherprovides a detailed insight on the relative weight of hydrody-namics and phoretic drift by obtaining precise and compara-tive scalings for the two routes of interactions in unboundedand confined geometries.

    In the Hele-Shaw limit considered, the leading orderdynamics takes the form of interacting potential dipoles, sim-ilarly to other categories of confined micro-swimmers62–64but with a reverse translational mobility due to the phoreticcoupling. The reverse translational motility slows down theparticles but does not affect the emergent collective modes,which are governed by the sign of the orientational mobil-ity coefficient ν. Particles that align with the drift createdby the other particles (ν > 0) exhibit global swirling andchaotic-like behavior, while particles that align opposite to the

    induced drift (ν < 0) tend to aggregate and form stationaryclusters.

    These collective modes were previously analyzed, albeitfor micro-swimmers with positive translational mobilitycoefficient, showing that the transition from swirling to clus-tering and aggregation as ν decreases from positive to nega-tive occurs systematically over the phase space consisting ofν and the particles’ area fraction Φ (with Φ being the ratioof the area of all particles to the area of the doubly peri-odic domain) taken to lie in the dilute to semi-dilute rangeΦ ∈ [0.1, 0.3].62 Therefore, we expect the global swirling andaggregation modes in the weakly confined auto-phoretic par-ticles, and the transition between them as ν decreases, to berobust to changes in the number of particles in this rangeof Φ. Furthermore, these global modes were shown to berobust to rotational Brownian noise for a range of rotationaldiffusion coefficients with Péclet numbers of order 1.62 Wethus expect rotational diffusion in this range of Péclet num-bers to have a small effect on the collective modes of weaklyconfined auto-phoretic particles. In fact, the framework pre-sented here purposely neglects the influence of thermal fluc-tuations and the Brownian nature of the dynamics of smallJanus colloids, as our focus was on understanding the role ofconfinement in screening and tuning each route of determin-istic interactions between chemically active colloids. Futurestudies will address in detail how such screening is poten-tially influenced by stochastic fluctuations in the particles’dynamics.

    These findings, albeit in the context of a simplified model,may have profound implications on understanding and con-trolling the collective behavior of active films by auto-phoreticparticles. They demonstrate that, in weak Hele-Shaw confine-ment, the emergent phase is controllable by the surface prop-erties of the individual Janus particles. The surface chemistrydictate the ability of a Janus particle to drive surface slip fromlocal concentration gradients, which, in turn, dictates the signand value of ν. Therefore, the mobility and chemical activity (Mand A) can be viewed as control parameters to systematicallyand predictably engineer active films with distinct emergentproperties, from spontaneous large-scale swirling motions tostationary clusters and aggregates.

    ACKNOWLEDGMENTSThis work was partially supported by a visiting scholar

    position from the Laboratoire d’Hydrodynamique LadHyX,Département de Mécanique, Ecole Polytechnique and by theNational Science Foundation via the NSF INSPIRE (Grant No.16-08744 to E.K.). This work was also supported by the Euro-pean Research Council (ERC) under the European Union’sHorizon 2020 research and innovation program (Grant Agree-ment No. 714027 to S.M.).

    REFERENCES1E. Lauga and T. R. Powers, “The hydrodynamics of swimming microorgan-isms,” Rep. Prog. Phys. 72(9), 096601 (2009).2C. Brennen and H. Winnet, “Fluid mechanics of propulsion by cilia andflagella,” Annu. Rev. Fluid. Mech. 9, 339–398 (1977).

    J. Chem. Phys. 150, 044902 (2019); doi: 10.1063/1.5065656 150, 044902-12

    Published under license by AIP Publishing

    https://scitation.org/journal/jcphttps://doi.org/10.1088/0034-4885/72/9/096601https://doi.org/10.1146/annurev.fl.09.010177.002011

  • The Journal ofChemical Physics ARTICLE scitation.org/journal/jcp

    3E. Lauga, “Bacterial hydrodynamics,” Annu. Rev. Fluid. Mech. 48, 105–130(2016).4R. Dreyfus, J. Baudry, M. L. Roper, M. Fermigier, H. A. Stone, and J. Bibette,“Microscopic artificial swimmers,” Nature 437, 862–865 (2005).5B. Antoine, J.-B. Caussin, N. Desreumaux, O. Dauchot, and D. Bartolo,“Emergence of macroscopic directed motion in populations of motile col-loids,” Nature 503(7474), 95–98 (2013).6W. Wang, L. A. Castro, M. Hoyos, and T. E. Mallouk, “Autonomous motionof metallic microrods propelled by ultrasound,” ACS Nano 6(7), 6122–6132(2012).7J. L. Anderson, “Colloid transport by interfacial forces,” Annu. Rev. Fluid.Mech. 21, 61–99 (1989).8W. F. Paxton, K. C. Kistler, C. C. Olmeda, A. Sen, S. K. St. Angelo, Y. Cao,T. E. Mallouk, P. E. Lammert, and V. H. Crespi, “Catalytic nanomotors:Autonomous movement of striped nanorods,” J. Am. Chem. Soc. 126(41),13424–13431 (2004).9J. R. Howse, R. A. L. Jones, A. J. Ryan, T. Gough, R. Vafabakhsh, andR. Golestanian, “Self-motile colloidal particles: From directed propulsion torandom walk,” Phys. Rev. Lett. 99(4), 048102 (2007).10I. Theurkauff, C. Cottin-Bizonne, J. Palacci, C. Ybert, and L. Bocquet,“Dynamic clustering in active colloidal suspensions with chemical signal-ing,” Phys. Rev. Lett. 108, 268303 (2012).11J. Palacci, S. Sacanna, A. P. Steinberg, D. J. Pine, and P. M. Chaikin,“Living crystals of light-activated colloidal surfers,” Science 339, 936–940(2013).12A. Brown and W. Poon, “Ionic effects in self-propelled Pt-coated Janusswimmers,” Soft Matter 10, 4016–4027 (2014).13M. Ibele, T. E. Mallouk, and A. Sen, “Schooling behavior of light-poweredautonomous micromotors in water,” Angew. Chem., Int. Ed. 48(18), 3308–3312 (2009).14Z. Izri, M. N. van der Linden, S. Michelin, and O. Dauchot, “Self-propulsionof pure water droplets by spontaneous Marangoni-stress-driven motion,”Phys. Rev. Lett. 113, 248302 (2014).15S. Herminghaus, C. C. Maass, C. Krüger, S. Thutupalli, L. Goehring, andC. Bahr, “Interfacial mechanisms in active emulsions,” Soft Matter 10, 7008(2014).16I. Buttinoni, J. Bialké, F. Kümmel, H. Löwen, C. Bechinger, and T. Speck,“Dynamical clustering and phase separation in suspensions of self-propelled colloidal particles,” Phys. Rev. Lett. 110, 238301 (2013).17P. Colberg and R. Kapral, “Many-body dynamics of chemically propellednanomotors,” J. Chem. Phys. 147, 064910 (2017).18P. Ilien, R. Golestanian, and A. Sen, “‘Fuelled’ motion: Phoretic motilityand collective behaviour of active colloids,” Chem. Soc. Rev. 46, 5508–5518(2017).19F. Ginot, I. Theurkauff, F. Detcheverry, C. Ybert, and C. Cottin-Bizonne,“Aggregation-fragmentation and individual dynamics of active clusters,”Nat. Commun. 9, 696 (2018).20A. Varma, T. D. Montenegro-Johnson, and S. Michelin, “Clustering-induced self-propulsion of isotropic phoretic particles,” Soft Matter 14, 7155(2018).21H. Ke, S. Ye, R. L. Carroll, and K. Showalter, “Motion analysis of self-propelled Pt-silica particles in hydrogen peroxide solutions,” J. Phys.Chem. A 114, 5462–5467 (2010).22L. F. Valadares, Y.-G. Tao, N. S. Zacharia, V. Kitaev, F. Galembeck, R. Kapral,and G. A. Ozin, “Catalytic nanomotors: Self-propelled sphere dimers,” Small6, 565–572 (2010).23S. J. Ebbens and J. R. Howse, “Direct observation of the direction of motionfor spherical catalytic swimmers,” Langmuir 27, 12293–12296 (2011).24G. Volpe, I. Buttinoni, D. Vogt, H.-J. Kümmerer, and C. Bechinger,“Microswimmers in patterned environments,” Soft Matter 7, 8810–8815(2011).25L. Baraban, R. Streubel, D. Makarov, L. Han, D. Karnaushenko, O. G.Schmidt, and G. Cuniberti, “Fuel-free locomotion of Janus motors:Magnetically-induced thermophoresis,” ACS Nano 7, 1360–1367 (2013).

    26M. S. Davies Wykes, X. Zhong, J. Tong, T. Adachi, Y. Liu, M. D. Ward,M. J. Shelley, and J. Zhang, “Guiding microscale swimmers using teardrop-shaped posts,” Soft Matter 13, 4681 (2017).27Y. Wang, R. M. Hernandez, D. J. Bartlett, Jr., J. M. Bingham, T. R. Kline,A. Sen, and T. E. Mallouk, “Bipolar electrochemical mechanism for thepropulsion of catalytic nanomotors in hydrogen peroxide solutions,” Lang-muir 22, 10451–10456 (2006).28J. L. Moran and J. D. Posner, “Phoretic self-propulsion,” Annu. Rev. Fluid.Mech. 49, 511–540 (2017).29Y. Ibrahim, R. Golestanian, and T. B. Liverpool, “Multiple phoretic mech-anisms in the self-propulsion of a Pt-insulator Janus swimmers,” J. FluidMech. 828, 318 (2017).30S. Saha, R. Golestanian, and S. Ramaswamy, “Clusters, asters, and col-lective oscillations in chemotactic colloids,” Phys. Rev. E 89, 062316(2014).31M. Tătulea-Codrean and E. Lauga, “Artificial chemotaxis of phoreticswimmers: Instantaneous and long-time behaviour,” J. Fluid Mech. 856,921–957 (2018).32R. Soto and R. Golestanian, “Self-assembly of catalytically active colloidalmolecules: Tailoring activity through surface chemistry,” Phys. Rev. Lett. 112,068301 (2014).33B. Liebchen, D. Marenduzzo, I. Pagonabarraga, and M. E. Cates, “Clus-tering and pattern formation in chemorepulsive active colloids,” Phys. Rev.Lett. 115, 258301 (2015).34B. Liebchen, D. Marenduzzo, and M. E. Cates, “Phoretic interactionsgenerically induce dynamic clusters and wave patterns in active colloids,”Phys. Rev. Lett. 118, 268001 (2017).35M.-J. Huang, J. Schofield, and R. Kapral, “Chemotactic and hydrodynamiceffects on collective dynamics of self-diffusiophoretic Janus motors,” NewJ. Phys. 19, 125003 (2017).36J. Palacci, C. Cottin-Bizonne, C. Ybert, and L. Bocquet, “Sedimentationand effective temperature of active colloidal suspensions,” Phys. Rev. Lett.105(8), 088304 (2010).37C. Krüger, C. Bahr, S. Herminghaus, and C. C. Maass, “Dimensionality mat-ters in the collective behaviour of active emulsions,” Eur. Phys. J. E 39, 64–72(2016).38R. Golestanian, T. B. Liverpool, and A. Ajdari, “Designing phoretic micro-and nano-swimmers,” New J. Phys. 9, 126 (2007).39F. Jülicher and J. Prost, “Generic theory of colloidal transport,” Eur. Phys.J. E 29(1), 27–36 (2009).40S. Michelin and E. Lauga, “Phoretic self-propulsion at finite Péclet num-bers,” J. Fluid Mech. 747, 572–604 (2014).41W. E. Uspal, M. N. Popescu, S. Dietrich, and M. Tasinkevych, “Self-propulsion of a catalytically active particle near a planar wall: From reflec-tion to sliding and hovering,” Soft Matter 11, 434–438 (2015).42W. E. Uspal, M. N. Popescu, S. Dietrich, and M. Tasinkevych, “Guidingcatalytically active particles with chemically patterned surfaces,” Phys. Rev.Lett. 117, 048002 (2016).43A. Mozaffari, N. Sharifi-Mood, J. Koplik, and C. Maldarelli, “Self-diffusiophoretic colloidal propulsion near a solid boundary,” Phys. Fluids 28,053107 (2016).44P. Malgaretti, M. N. Popescu, and S. Dietrich, “Self-diffusiophoresisinduced by fluid interfaces,” Soft Matter 14, 1375 (2018).45S. Das, A. Garg, A. I. Campbell, J. Howse, A. Sen, D. l Velegol, R. Golestanian,and S. J. Ebbens, “Boundaries can steer active Janus spheres,” Nat. Commun.6, 8999 (2015).46J. Simmchen, J. Katuri, W. E. Uspal, M. N. Popescu, M. Tasinkevych, andS. Sánchez, “Topographical pathways guide chemical microswimmers,” Nat.Commun. 7, 10598 (2016).47A. Dominguez, P. Malgaretti, M. N. Popescu, and S. Dietrich, “Effec-tive interaction between active colloids and fluid interfaces induced byMarangoni flows,” Phys. Rev. Lett. 116, 078301 (2016).48J. R. Blake and A. T. Chwang, “Fundamental singularities of viscous flow.I. The image system in the vicinity of a stationary no-slip boundary,” J. Eng.Math. 8, 23–29 (1974).

    J. Chem. Phys. 150, 044902 (2019); doi: 10.1063/1.5065656 150, 044902-13

    Published under license by AIP Publishing

    https://scitation.org/journal/jcphttps://doi.org/10.1146/annurev-fluid-122414-034606https://doi.org/10.1038/nature04090https://doi.org/10.1038/nature12673https://doi.org/10.1021/nn301312zhttps://doi.org/10.1146/annurev.fl.21.010189.000425https://doi.org/10.1146/annurev.fl.21.010189.000425https://doi.org/10.1021/ja047697zhttps://doi.org/10.1103/physrevlett.99.048102https://doi.org/10.1103/physrevlett.108.268303https://doi.org/10.1126/science.1230020https://doi.org/10.1039/c4sm00340chttps://doi.org/10.1002/anie.200804704https://doi.org/10.1103/physrevlett.113.248302https://doi.org/10.1039/c4sm00550chttps://doi.org/10.1103/physrevlett.110.238301https://doi.org/10.1063/1.4997572https://doi.org/10.1039/c7cs00087ahttps://doi.org/10.1038/s41467-017-02625-7https://doi.org/10.1039/c8sm00690chttps://doi.org/10.1021/jp101193uhttps://doi.org/10.1021/jp101193uhttps://doi.org/10.1002/smll.200901976https://doi.org/10.1021/la2033127https://doi.org/10.1039/c1sm05960bhttps://doi.org/10.1021/nn305726mhttps://doi.org/10.1039/c7sm00203chttps://doi.org/10.1021/la0615950https://doi.org/10.1021/la0615950https://doi.org/10.1146/annurev-fluid-122414-034456https://doi.org/10.1146/annurev-fluid-122414-034456https://doi.org/10.1017/jfm.2017.502https://doi.org/10.1017/jfm.2017.502https://doi.org/10.1103/physreve.89.062316https://doi.org/10.1017/jfm.2018.718https://doi.org/10.1103/physrevlett.112.068301https://doi.org/10.1103/physrevlett.115.258301https://doi.org/10.1103/physrevlett.115.258301https://doi.org/10.1103/physrevlett.118.268001https://doi.org/10.1088/1367-2630/aa958chttps://doi.org/10.1088/1367-2630/aa958chttps://doi.org/10.1103/physrevlett.105.088304https://doi.org/10.1140/epje/i2016-16064-yhttps://doi.org/10.1088/1367-2630/9/5/126https://doi.org/10.1140/epje/i2008-10446-8https://doi.org/10.1140/epje/i2008-10446-8https://doi.org/10.1017/jfm.2014.158https://doi.org/10.1039/c4sm02317jhttps://doi.org/10.1103/physrevlett.117.048002https://doi.org/10.1103/physrevlett.117.048002https://doi.org/10.1063/1.4948398https://doi.org/10.1039/c7sm02347bhttps://doi.org/10.1038/ncomms9999https://doi.org/10.1038/ncomms10598https://doi.org/10.1038/ncomms10598https://doi.org/10.1103/physrevlett.116.078301https://doi.org/10.1007/bf02353701https://doi.org/10.1007/bf02353701

  • The Journal ofChemical Physics ARTICLE scitation.org/journal/jcp

    49S. Thutupalli, D. Geyer, R. Singh, R. Adhikari, and H. A. Stone, “Flow-induced phase separation of active particles is controlled by boundaryconditions,” Proc. Natl. Acad. Sci. U. S. A. 115(21), 5403–5408 (2018).50S. Kim and S. J. Karrila, Microhydrodynamics (Butterworth-Heinemann,1991).51J. R. Blake, “A spherical envelope approach to ciliary propulsion,” J. FluidMech. 46, 199–208 (1971).52S. Michelin and E. Lauga, “Optimal feeding is optimal swimming for allPéclet numbers,” Phys. Fluids 23(10), 101901 (2011).53R. Soto and R. Golestanian, “Self-assembly of active colloidal moleculeswith dynamic function,” Phys. Rev. E 91, 052304 (2015).54B. Liebchen and H. Löwen, “Which interactions dominate in active col-loids?,” e-print arXiv:1808.07389v1 (2018).55G. B. Jeffery, “The motion of ellipsoidal particles immersed in a viscousfluid,” Proc. R. Soc. London, Ser. A 102(715), 161–179 (1922).56O. Pohl and H. Stark, “Dynamic clustering and chemotactic collapse ofself-phoretic active particles,” Phys. Rev. Lett. 112, 238303 (2014).57J. D. Jackson, Classical Electrodynamics (John Wiley & Sons, New York,1962).58H. A. Stone and A. D. T. Samuel, “Propulsion of microorganisms by surfacedistorsions,” Phys. Rev. Lett. 77, 4102 (1996).

    59N. Liron and S. Mochon, “Stokes flow for a stokeslet between two parallelflat plates,” J. Eng. Math. 10(4), 287–303 (Oct 1976).60S. E. Spagnolie and E. Lauga, “Hydrodynamics of self-propulsion near aboundary: Predictions and accuracy of far-field approximations,” J. FluidMech. 700, 105–147 (2012).61N. Desreumaux, N. Florent, E. Lauga, and D. Bartolo, “Active and drivenhydrodynamic crystals,” Eur. Phys. J. E 35(8), 68 (2012).62A. C. H. Tsang and E. Kanso, “Flagella-induced transitions in the col-lective behavior of confined microswimmers,” Phys. Rev. E 90, 021001(2014).63A. C. H. Tsang and E. Kanso, “Circularly confined microswimmers exhibitmultiple global patterns,” Phys. Rev. E 91, 043008 (2015).64A. C. H. Tsang and E. Kanso, “Density shock waves in confined microswim-mers,” Phys. Rev. Lett. 116, 048101 (2016).65A. C. H. Tsang and E. Kanso, “Dipole interactions in doubly periodicdomains,” J. Nonlinear Sci. 23, 971–991 (2013).66R. Coquereaux, A. Grossmann, and B. E. Lautrup, “Iterative method forcalculation of the weierstrass elliptic function,” IMA J. Numer. Anal. 10, 119–128 (1990).67E. Kanso and A. C. H. Tsang, “Dipole models of self-propelled bodies,” FluidDyn. Res. 46, 061407 (2014).

    J. Chem. Phys. 150, 044902 (2019); doi: 10.1063/1.5065656 150, 044902-14

    Published under license by AIP Publishing

    https://scitation.org/journal/jcphttps://doi.org/10.1073/pnas.1718807115https://doi.org/10.1017/s002211207100048xhttps://doi.org/10.1017/s002211207100048xhttps://doi.org/10.1063/1.3642645https://doi.org/10.1103/physreve.91.052304http://arxiv.org/abs/1808.07389v1https://doi.org/10.1098/rspa.1922.0078https://doi.org/10.1103/physrevlett.112.238303https://doi.org/10.1103/physrevlett.77.4102https://doi.org/10.1007/bf01535565https://doi.org/10.1017/jfm.2012.101https://doi.org/10.1017/jfm.2012.101https://doi.org/10.1140/epje/i2012-12068-yhttps://doi.org/10.1103/physreve.90.021001https://doi.org/10.1103/physreve.91.043008https://doi.org/10.1103/physrevlett.116.048101https://doi.org/10.1007/s00332-013-9174-5https://doi.org/10.1093/imanum/10.1.119https://doi.org/10.1088/0169-5983/46/6/061407https://doi.org/10.1088/0169-5983/46/6/061407