artificial chemotaxis of self-phoretic active colloids: collective … · phoretic colloids near...

8
Articial Chemotaxis of Self-Phoretic Active Colloids: Collective Behavior Published as part of the Accounts of Chemical Research special issue Fundamental Aspects of Self-Powered Nano- and Micromotors. Holger Stark* Technische Universitä t Berlin, Institute of Theoretical Physics, Hardenbergstrasse 36, D-10623 Berlin, Germany CONSPECTUS: Microorganisms use chemotaxis, regulated by internal complex chemical pathways, to swim along chemical gradients to nd better living conditions. Articial microswimmers can mimic such a strategy by a pure physical process called diusiophoresis, where they drift and orient along the gradient in a chemical density eld. Similarly, for other forms of taxis in nature such as photo- or thermotaxis the phoretic counterpart exists. In this Account, we concentrate on the chemotaxis of self-phoretic active colloids. They are driven by self-electro- and diusiophoresis at the particle surface and thereby acquire a swimming speed. During this process, they also produce nonuniform chemical elds in their surroundings through which they interact with other colloids by translational and rotational diusiophoresis. In combination with active motion, this gives rise to eective phoretic attraction and repulsion and thereby to diverse emergent collective behavior. A particular appealing example is dynamic clustering in dilute suspensions rst reported by a group from Lyon. A subtle balance of attraction and repulsion causes very dynamic clusters, which form and resolve again. This is in stark contrast to the relatively static clusters of motility-induced phase separation at larger densities. To treat chemotaxis in active colloids conned to a plane, we formulate two Langevin equations for position and orientation, which include translational and rotational diusiophoretic drift velocities. The colloids are chemical sinks and develop their long-range chemical proles instantaneously. For dense packings, we include screening of the chemical elds. We present a state diagram in the two diusiophoretic parameters governing translational, as well as rotational, drift and, thereby, explore the full range of phoretic attraction and repulsion. The identied states range from a gaslike phase over dynamic clustering states 1 and 2, which we distinguish through their cluster size distributions, to dierent types of collapsed states. The latter include a full chemotactic collapse for translational phoretic attraction. Turning it into an eective repulsion, with increasing strength rst the collapsed cluster starts to uctuate at the rim, then oscillates, and ultimately becomes a static collapsed cloud. We also present a state diagram without screening. Finally, we summarize how the famous KellerSegel model derives from our Langevin equations through a multipole expansion of the full one-particle distribution function in position and orientation. The KellerSegel model gives a continuum equation for treating chemotaxis of microorganisms on the level of their spatial density. Our theory is extensible to mixtures of active and passive particles and allows to include a dipolar correction to the chemical eld resulting from the dipolar symmetry of Janus colloids. 1. INTRODUCTION In nature microorganisms have developed the ability to sense their environment to direct their motion. An important behavioral response is called taxis. Microorganisms can sense and move along the gradient of an external stimulus or eld and thereby are able to nd better living conditions. 1 The most prominent example is chemotaxis, where the gradient is formed by the density of a chemical species, 2 and many micro- organisms employ this strategy. 26 Other forms of taxis found for living organisms are gravitaxis, 7 rheotaxis, 8 magnetotaxis 9 phototaxis, 10 or thermotaxis. 11 To implement a taxis strategy, microorganisms change their swimming mode in response to eld gradients. For example, bacteria modify their tumble rate using chemical sensors and an internal biochemical signaling pathway 1215 or light sensing algae alter the breast stroke of their agella, which is triggered by light receptors. 10 The rapidly evolving eld of articially designed micro- swimmers 16,17 is also driven by the idea of mimicking dierent taxis strategies, now based on pure physical principles, to control their swimming paths and explore a wide eld of intriguing applications. Indeed, in colloidal systems phoretic motion induced by eld gradients (phoresis) is well established. 18 Combined with activity, colloids exhibit novel Received: June 3, 2018 Published: October 16, 2018 Article pubs.acs.org/accounts Cite This: Acc. Chem. Res. 2018, 51, 2681-2688 © 2018 American Chemical Society 2681 DOI: 10.1021/acs.accounts.8b00259 Acc. Chem. Res. 2018, 51, 26812688 Downloaded via TU BERLIN on November 30, 2018 at 23:12:08 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Upload: others

Post on 20-Oct-2020

7 views

Category:

Documents


0 download

TRANSCRIPT

  • Artificial Chemotaxis of Self-Phoretic Active Colloids: CollectiveBehaviorPublished as part of the Accounts of Chemical Research special issue “Fundamental Aspects of Self-PoweredNano- and Micromotors”.

    Holger Stark*

    Technische Universitaẗ Berlin, Institute of Theoretical Physics, Hardenbergstrasse 36, D-10623 Berlin, Germany

    CONSPECTUS: Microorganisms use chemotaxis, regulated by internal complexchemical pathways, to swim along chemical gradients to find better living conditions.Artificial microswimmers can mimic such a strategy by a pure physical process calleddiffusiophoresis, where they drift and orient along the gradient in a chemical density field.Similarly, for other forms of taxis in nature such as photo- or thermotaxis the phoreticcounterpart exists.In this Account, we concentrate on the chemotaxis of self-phoretic active colloids. Theyare driven by self-electro- and diffusiophoresis at the particle surface and thereby acquire aswimming speed. During this process, they also produce nonuniform chemical fields intheir surroundings through which they interact with other colloids by translational androtational diffusiophoresis. In combination with active motion, this gives rise to effectivephoretic attraction and repulsion and thereby to diverse emergent collective behavior. Aparticular appealing example is dynamic clustering in dilute suspensions first reported by agroup from Lyon. A subtle balance of attraction and repulsion causes very dynamicclusters, which form and resolve again. This is in stark contrast to the relatively static clusters of motility-induced phaseseparation at larger densities.To treat chemotaxis in active colloids confined to a plane, we formulate two Langevin equations for position and orientation,which include translational and rotational diffusiophoretic drift velocities. The colloids are chemical sinks and develop theirlong-range chemical profiles instantaneously. For dense packings, we include screening of the chemical fields.We present a state diagram in the two diffusiophoretic parameters governing translational, as well as rotational, drift and,thereby, explore the full range of phoretic attraction and repulsion. The identified states range from a gaslike phase overdynamic clustering states 1 and 2, which we distinguish through their cluster size distributions, to different types of collapsedstates. The latter include a full chemotactic collapse for translational phoretic attraction. Turning it into an effective repulsion,with increasing strength first the collapsed cluster starts to fluctuate at the rim, then oscillates, and ultimately becomes a staticcollapsed cloud. We also present a state diagram without screening. Finally, we summarize how the famous Keller−Segel modelderives from our Langevin equations through a multipole expansion of the full one-particle distribution function in position andorientation. The Keller−Segel model gives a continuum equation for treating chemotaxis of microorganisms on the level of theirspatial density.Our theory is extensible to mixtures of active and passive particles and allows to include a dipolar correction to the chemicalfield resulting from the dipolar symmetry of Janus colloids.

    1. INTRODUCTION

    In nature microorganisms have developed the ability to sensetheir environment to direct their motion. An importantbehavioral response is called taxis. Microorganisms can senseand move along the gradient of an external stimulus or fieldand thereby are able to find better living conditions.1 The mostprominent example is chemotaxis, where the gradient is formedby the density of a chemical species,2 and many micro-organisms employ this strategy.2−6 Other forms of taxis foundfor living organisms are gravitaxis,7 rheotaxis,8 magnetotaxis9

    phototaxis,10 or thermotaxis.11 To implement a taxis strategy,microorganisms change their swimming mode in response tofield gradients. For example, bacteria modify their tumble rate

    using chemical sensors and an internal biochemical signalingpathway12−15 or light sensing algae alter the breast stroke oftheir flagella, which is triggered by light receptors.10

    The rapidly evolving field of artificially designed micro-swimmers16,17 is also driven by the idea of mimicking differenttaxis strategies, now based on pure physical principles, tocontrol their swimming paths and explore a wide field ofintriguing applications. Indeed, in colloidal systems phoreticmotion induced by field gradients (phoresis) is wellestablished.18 Combined with activity, colloids exhibit novel

    Received: June 3, 2018Published: October 16, 2018

    Article

    pubs.acs.org/accountsCite This: Acc. Chem. Res. 2018, 51, 2681−2688

    © 2018 American Chemical Society 2681 DOI: 10.1021/acs.accounts.8b00259Acc. Chem. Res. 2018, 51, 2681−2688

    Dow

    nloa

    ded

    via

    TU

    BE

    RL

    IN o

    n N

    ovem

    ber

    30, 2

    018

    at 2

    3:12

    :08

    (UT

    C).

    Se

    e ht

    tps:

    //pub

    s.ac

    s.or

    g/sh

    arin

    ggui

    delin

    es f

    or o

    ptio

    ns o

    n ho

    w to

    legi

    timat

    ely

    shar

    e pu

    blis

    hed

    artic

    les.

    https://pubs.acs.org/page/achre4/self-powered-nano-micromotors.htmlhttps://pubs.acs.org/page/achre4/self-powered-nano-micromotors.htmlpubs.acs.org/accountshttp://pubs.acs.org/action/showCitFormats?doi=10.1021/acs.accounts.8b00259http://dx.doi.org/10.1021/acs.accounts.8b00259

  • complex behavior while they sense temperature gradients,19−22

    gravitational fields,23 perform phototaxis,24 or show diffusio-phoresis,25−29 where they drift and orient along chemicalgradients.In this Account, we concentrate on how chemical fields

    influence the collective motion of self-phoretic activecolloids,28,30−33 which produce nonuniform chemical fieldsby consuming chemical reactants and generating products.Prominently, they show dynamic clustering25−27,29,34 orstimulate the assembly of passive colloids.35 In suspensionsof active AgCl particles a colloidal collapse and clusteroscillations were observed.36,37 Continuum theories reveal awealth of pattern formation.28,38,39 A detailed diffusiophoreticand hydrodynamic modeling reproduces the motion of self-phoretic colloids near boundaries with step-like topography.40

    On the molecular level enzymes are established as nanomotors,which undergo chemotaxis.41−44 Hydrodynamic modeling ofchemical nanomotors either of Janus type or sphere dimersshow different kinds of clustering45−47 or orientational orderwhen pinned to a substrate.48 Finally, also emulsion dropletsexhibit chemotaxis.49

    This Account summarizes our work on the chemotaxis ofself-phoretic active colloids.27,29 It was very much inspired byexperiments of ref 25. Gold particles half covered withplatinum, which catalyzes the reaction of H2O2 into waterand oxygen, self-propel through a combination of self-diffusio-and electrophoresis. They create nonuniform chemical fields intheir surroundings with chemical gradients, along whichneighboring particles drift (translational diffusiophoresis) orreorient (rotational diffusiophoresis). In combination with self-propulsion this gives rise to an effective phoretic attraction andrepulsion. A subtle balance of these effective interactions thencauses dynamic clustering in dilute suspensions, where clustersform and resolve again.25,26 This is in stark contrast to therelatively static clusters of motility-induced phase separation atlarger densities.50,51

    In the following we summarize our particle-based Langevinmodel for addressing collective motion of self-phoretic activecolloids. In formulating the model, we wanted to concentrateon the essential features. Exploring the full range of phoreticattraction and repulsion, we illustrate the collective dynamicsof self-phoretic colloids ranging from a gas-like state, overdynamic clustering 1 and 2, to different types of collapsedstates including a full chemotactic collapse. The latter can berationalized by the Keller−Segel equation,52,53 which derivesfrom our Langevin model.

    2. MODELTo explore the collective dynamics of self-phoretic colloidsinduced by diffusiophoretic or chemotactic interactions, we setup a simplified model system. We consider a colloidalmonolayer close to a bounding plate with fluid in the infinitehalf-space above it. The active colloids are Janus particlespartially covered by a catalyst, which catalyzes a chemicalreaction in the surrounding fluid. In general, through acombined process of self-electro- and diffusiophoresis, thecolloids move with a velocity v0 along the unit vector e, whichgives a direction fixed in the particle. In the following, weassume that v0 and e are not changed by the presence of othernearby particles. Instead of taking into account all reactantsand products of the chemical reaction, we simply consider theself-phoretic colloid as a chemical sink. It consumes a chemicaland thereby produces a nonuniform chemical field with

    concentration c, in which nearby colloids perform translationaland rotational diffusiophoretic motion. Hydrodynamic inter-actions, which have a reduced range close to a no-slip surface,are neglected against these effective interactions. The influenceof hydrodynamic flow has been studied in different systemsand is reviewed in ref 17.2.1. Translational and Rotational Diffusiophoresis

    The molecules or solutes of the chemical field interact with thesurface of a colloid. This results in a body force on the fluid,which influences fluid pressure. Ultimately a gradient in theconcentration c along the particle surface causes a pressuregradient, which then drives a slip velocity vs = ζ∇c along thecolloidal surface, where the slip velocity coefficient ζ dependson the surface interaction potential between the chemical andthe colloidal surface. Averaging vs over the particle surfacegives the translational diffusiophoretic velocity18

    ζ ζ ∇= [⟨ ⟩ − ⟨ ⊗ − ⟩] cv 1 n n 1(3 )/2D (1)

    where ∇c is evaluated at the particle center, n is the localsurface normal, and ⟨...⟩ means average over the particlesurface. For particles with uniform surface properties but alsofor half-coated Janus colloids the quadrupolar term in eq 1vanishes and we obtain the diffusiophoretic drift velocity,which we use in the following

    ζ ∇= − cvD tr (2)

    Here, ζtr ≔ −⟨ζ⟩ is the translational diffusiophoreticparameter.The slip velocity field also causes a diffusiophoretic

    rotational velocity18

    ω ζ ζ∇ ∇= ⟨ ⟩ × = − ×a

    c cn e9

    4 iD rot (3)

    where the rotational diffusiophoretic parameter ζrot is definedby (9/4a)⟨ζn⟩ = −ζrote. Spherical particles with a uniformsurface (constant ζ) do not rotate. However, for half-coatedJanus particles ωD is nonzero, as long as the solutes interactdifferently with the two sides, and symmetry dictates thatswimming direction e is parallel to ⟨ζn⟩.Since the slip velocity coefficient ζ can be controlled by

    choosing appropriate materials for the Janus colloids and theircaps, but also by the geometry and number of the capscovering the colloidal surface,28 we expect the phoreticparameters ζrot and ζtr to be tunable to either positive ornegative values. This opens the possibility to induce andexplore a variety of collective colloidal dynamics. For example,since each active particle acts as a chemical sink in our setting,neighboring particles will drift toward the sink for positivetranslational diffusiophoretic parameter ζtr in eq 2, whichcorresponds to an effective attraction, while ζtr < 0 gives rise toan effective repulsion. Similarly, a positive rotational parameterζrot in eq 3 rotates the swimming direction of an active colloidtoward a neighboring chemical sink and the colloid movestoward the sink. Hence, rotational phoresis also acts like anattractive colloidal interaction while it becomes repulsive forζrot < 0.2.2. Langevin Dynamics and Chemical Field

    At the micron scale inertia is negligible and position ri andorientation ei of the ith Janus colloid obey overdampedLangevin equations. Adding up the deterministic translationalvelocities from self-propulsion and diffusiophoretic drift (v0ei +

    Accounts of Chemical Research Article

    DOI: 10.1021/acs.accounts.8b00259Acc. Chem. Res. 2018, 51, 2681−2688

    2682

    http://dx.doi.org/10.1021/acs.accounts.8b00259

  • vD) and using the kinematic relation for the orientation vector,ei̇ = ωD × ei, we obtain

    ξζ ∇̇ = − +v cr e r( )i i i i0 tr (4)

    μζ ∇̇ = − − ⊗ + ×ce 1 e e r e( ) ( )i i i i i irot (5)

    The additional vectors represent translational (ξi) androtational (μi) white noise of thermal origin with zero meanand respective time correlation functions ⟨ξi(t) ⊗ ξi(t′)⟩ =2Dtr1δ(t − t′) and ⟨μi(t)⊗μi(t′)⟩ = 2Drot1δ(t − t′), where Dtrand Drot are the translational and rotational diffusioncoefficients, respectively. All the results presented in thisreview are generated by a two-dimensional version of eqs 4 and5, since the Janus colloids move in a monolayer. An effectivehard-core repulsion between the colloids is implemented.Whenever they overlap during the simulations, we separatethem along the line connecting their centers to the point ofcontact. Finally, hydrodynamic flow fields are not includedhere. Close to a bounding plane they are less long-ranged andwe also assume that chemical interactions dominate.According to our model assumption the colloids consume a

    chemical with rate k, which typically diffuses much faster onthe micron scale than the Janus colloids swim. Therefore, whenthe colloids move, they carry around with them a staticconcentration field c, which obeys the Poisson equation [for adiscussion of this approximation and its impact, see refs 38 and39]

    ∑ δ∇= − −=

    D c k r r0 ( )i

    N

    ic2

    1 (6)

    where we approximate the Janus colloids as point-like chemicalsinks. We neglect here higher-order contributions that wouldbreak the radial symmetry of the concentration field because ofthe dipolar character of the Janus colloid. The solution is givenby

    ∑π

    = −| − |=

    c hckh

    Dr

    r r( )

    41

    i

    N

    i2D 0

    c 1 (7)

    Here, c0 is the concentration field far away from the colloids,and we have multiplied with the thickness of the colloidalmonolayer h = 2a to obtain a two-dimensional density in theplane of the monolayer. Note that according to our model thechemical field still diffuses in an infinite three-dimensional half-space. A no-flux boundary condition at the bounding surfacedoes not change the principal 1/r dependence in eq 7.If the chemotactic attraction between the active colloids is

    sufficiently strong, they form compact clusters, where thechemical substance cannot diffuse freely between the colloids.We roughly take this into account by implementing a screenedchemical field, whenever a colloid is surrounded by six closelypacked neighbors with distances below rs = 2a(1 + ϵ). Then,we replace the algebraic decay 1/r = 1/|r − ri| in eq 7 byexp[−(r − ξ)/ξ]/r, where we introduce the screening length ξ≔ rs. We set ϵ = 0.3 but have checked that varying ϵ by 50%does not change the results. [In the Supporting Information ofref 26, the authors also mention a cutoff of the particleattraction at distances larger than three particle diameters. So,in particular, dynamic clustering should also be visible forlarger screening lengths.]

    2.3. Essential Parameters

    Rescaling the dimensions of all quantities in the equations ofmotion can be used to reduce the number of systemparameters and thereby reveals the essential parameters. Onepossibility is to set the rescaled diffusion constants, whichdetermine the noise strengths in the Langevin equations 4 and5, to one. This is achieved when rescaling time by tr = 1/(2Drot) and length by = =l D D a/ 2.33r tr rot . The thermaldiffusion coefficients in a bulk fluid would give

    =D D a/ 1.15tr rot , while the experiments of ref 25 measurelr = 1.79a for colloids moving close to a bottom wall. Thehigher value used for all reported results in this Account doesnot change the qualitative behavior of our system.Ultimately, one obtains four essential system parameters: the

    Peclet number = v D DPe /(2 )0 tr rot , the rescaled translational

    diffusiophoretic parameter ζ π ζ→kh D D D/(8 ) /tr c rot tr3

    tr, therescaled rotational diffusiophoretic parameter ζrotkh/(8πDcDtr)→ ζrot, and the area fraction σ defined as the ratio of projectedarea of all Janus colloids to the area of the simulation box. Notethat the factor kh/(4πDc) from eq 7 is already subsumed intothe rescaled diffusiophoretic parameters, when using ∇c2D(ri)in the Langevin eqs 4 and 5.2.4. Numerical Implementation

    The Langevin eqs 4 and 5 are solved by a typical Euler schemein a two-dimensional square simulation box. Always 800particles are used and different area fractions σ are realized byadjusting the size of the simulation box. To implement thehard-core interactions mentioned earlier, a sufficiently smalltime step has to be chosen, which makes them the numericallymost expensive part of the simulations. Therefore, a neighborlist is implemented such that the search for overlappingparticles could be restricted to eight immediate neighbors, atmost. Furthermore, due to the small translational steps of thecolloids it is sufficient to update the chemical concentrationfield every 50th time step, which further saves computationaltime.Since only bulk properties are of interest here, boundary

    conditions are implemented, which keep the colloids awayfrom the boundary of the simulation box. Thus, wheneverhitting the boundary, the particles are reflected into a randomlychosen direction.

    3. RESULTS AND DISCUSSION

    3.1. Overview: State Diagram

    Figure1 shows a typical state diagram at low area fractions ofthe colloids for the two phoretic parameters ζtr and ζrot. It isroughly divided by a diagonal line, which separates collapsedstates, where all colloids form one cluster with differentproperties, from states with varying cluster size. In particular,for ζtr and ζrot both positive meaning that translational androtational phoretic motion act like an effective attractionbetween the colloids, a sharp transition from a gaslike to acollapsed state occurs, where all particles are packed into onesingle static cluster (see snapshot in Figure 2, bottom right).This behavior is reminiscent of the chemotactic collapse inbacterial systems.52,53 Making both phoretic interactionsrepulsive (ζtr < 0 and ζrot < 0), the gaslike state is realized,where small transient clusters are possible.From here two directions are possible. On the one hand,

    inreasing ζtr to sufficiently large ζtr > 0, so that translational

    Accounts of Chemical Research Article

    DOI: 10.1021/acs.accounts.8b00259Acc. Chem. Res. 2018, 51, 2681−2688

    2683

    http://dx.doi.org/10.1021/acs.accounts.8b00259

  • diffusiophoresis acts attractive, dynamic clustering occurs intwo different states 1 and 2. Motile clusters form that stronglyfluctuate in shape and size and may ultimately dissolve again(see snapshots in Figure 2, top right and bottom left). Thestrongly dynamic clusters form due to a delicate balance of theattractive translational phoresis (holding the clusters together)and the fact that active colloids turn and swim away from thecluster (effective repulsion). This behavior is decisivelydifferent from the motility-induced phase separation of hard-core active particles.50,51 On the other hand, keeping ζtr < 0(effective repulsion) but making ζrot sufficiently large so thatactive colloids orient and therefore swim toward each other(effective attraction), different types of collapsed states occur.With decreasing ζtr first the collapsed cluster fluctuates at the

    rim due to particles leaving and rejoining it. Then the clusterdevelops nearly regular oscillations and, finally, for stronglynegative ζtr a static collapsed cloud appears, where colloids donot touch and their mutual distance increases toward the rimof the cloud.It is possible to derive the famous Keller−Segel equation

    from the Langevin eqs 4 and 5 (see section 3.5 and ref 27). Asin bacterial systems it here predicts a chemotactic collapse ofthe active colloids at 8πσ(ζtr + ζrotPe)/(1 + 2Pe

    2) = b, where bis a positive constant. In the ζrot−ζtr state diagram, the linearseparation line between the gaslike and collapsed state has anegative slope and it is shifted along the vertical to positivevalues. Thus, it gives the right trend for the state diagram ofFigure 1. The Keller−Segel equation is a mean-field equation,it only contains the density of the colloids. Therefore, it cannotdescribe subtle states such as dynamic clustering 1 and 2. Forexample, if ζtr is sufficiently large so that the phoretic driftvelocity vD exceeds the swim speed v0, the active colloids willalways stay attached to a cluster, regardless how negative ζrotbecomes. Hence, the separation line in Figure 1 deviates fromthe simple straight line.Finally, the state diagram of Figure 1 does not change

    qualitatively with Pe and σ, as long as Pe is well above one andσ sufficiently low so that phase separation does not occur.In the following two sections we characterize the dynamic

    clustering states and collapsed states in detail, report on thestate diagram, when screening of the chemical field in densecolloid clusters is switched off, and comment on how to derivethe Keller−Segel equation from our Langevin model.3.2. Gas Phase and Dynamic Clustering

    3.2.1. Signature of Dynamic Clustering. Dynamicclustering states are distinguished from the gaslike state bythe occurrence of larger cluster and a pronounced increase incluster size between states 1 and 2. This is already pictured inFigure 1, where the dynamic clustering state 1 shows anincrease of the mean cluster size Nc in a small region. To bettercharacterize these states and justify, why there are twoclustering states, Figure 3 shows the cluster size distributionfor increasing ζtr for fixed rotational phoretic parameter ζrot =−0.38. The curves until ζtr = 15.4 are well fitted by

    = −β−P n c n n n( ) exp( / )0 0 (8)

    where the cutoff size n0 is a measure how large the clusters canbecome. While for pure steric interaction (ζtr = 0, blue curve)

    Figure 1. Full state diagram ζtr versus ζrot at Pe = 19 and surfacefraction σ = 0.05. The mean cluster size Nc for the gaslike anddynamic-clustering state 1 are color-coded. A full discussion isprovided in the main text. Reprinted with permission from ref 29.Copyright 2015 European Physical Journal.

    Figure 2. Snapshots of colloid configurations for increasing ζtrans atζrot = −0.38 and Pe = 19. Top left: Gas-like state. Top right: Dynamicclustering 1. Bottom left: Dynamic clustering 2. Bottom right:Collapsed state. Nc is the mean cluster size. Reprinted with permissionfrom ref 27. Copyright 2014 American Physical Society.

    Figure 3. Cluster size distributions P(n) for increasing translationalphoretic parameter ζtr at Pe = 19 and ζrot = −0.38. The transitionbetween dynamic clustering states 1 and 2 occurs between the red andgreen curves. Inset: mean cluster size Nc versus ζtr. The transition isindicated. Reprinted with permission from ref 27. Copyright 2014American Physical Society.

    Accounts of Chemical Research Article

    DOI: 10.1021/acs.accounts.8b00259Acc. Chem. Res. 2018, 51, 2681−2688

    2684

    http://dx.doi.org/10.1021/acs.accounts.8b00259

  • and small ζtr (gray curve) an exponential decay is predominant,the distributions for larger ζtr (orange and red curves) followfirst the power law before they are cut off by the exponential.The exponent β = 2.1 ± 0.1 gradually decreases for morenegative ζrot and the fit is robust against increasing particlenumber.27 So there is a smooth transition from the gaslike tothe dynamic clustering state 1.Further increasing ζtr, the distribution P(n) develops an

    inflection point (green curves), which marks the onset of thedynamic clustering state 2. At the transition the mean clustersize strongly increases as the inset indicates. Now, the sum oftwo power-law-exponential curves have to be used for fittingthe distributions,

    = − + −β β− −P n c n n n c n n n( ) exp( / ) exp( / )1 1 2 21 2 (9)with β1 = 2.1 ± 0.2 and β2 ≈ 1.5. Both dynamic clusteringstates are observed for all negative ζrot. However, turning offscreening of the chemical field within clusters, dynamicclustering is less pronounced and the clustering state 2 doesnot occur at all.27 Finally, all the exponents βi decrease forlarger area fraction σ and the sharp increase of Nc at thetransition from state 1 to 2 vanishes (see Figure 4). In

    literature similar cluster-size distributions including thetransition indicated by the occurrence of an inflection pointwere observed in experiments on gliding bacteria, when thebacterial density was varied.54 While this system shows purehard-core interactions causing nematic alignment, we insteadvary the strength of the diffusiophoretic coupling.3.2.2. Dynamic Clustering for Varying Pećlet Num-

    ber. In Figure 5, we plot the state diagram ζtr versus Pećletnumber Pe. Dynamic clustering becomes more pronounced forlarger ζtr and then also occurs at larger Pe, which is necessaryto establish the delicate balance between translational phoreticattraction (ζtr > 0) and effective repulsion (ζrot < 0), where theactive colloids swim away from the clusters. However, forconstant ζtr large clusters disappear with increasing Pe. This isin contrast to the experiments with diffusiophoretic coupling,which showed a linear scaling of the mean cluster size with Pe:Nc ∼ Pe, when increasing fuel concentration.25To rationalize the experimental scaling prediction, we note

    that increasing the fuel concentration c0 means higher reactionrate k on the colloidal surface and therefore more self-activitybut also larger phoretic forces, which is encoded in the reduced

    phoretic parameters ζtr, ζrot ∝ k introduced in section 2.3.Assuming Michaelis−Menten kinetics for the reaction rate inthe linear regime well before saturation, k ∼ c0, we not onlyfind Pe ∝ c055 but also ζtr ∝ ζrot ∝ c0. Therefore, varying c0defines a straight line in the ζtr−ζrot−Pe parameter space. Inthis three-dimensional space the dynamic clustering states 1and 2 are separated by a plane. We choose different lines,which always hit the transition plane and plot in Figure 6 the

    mean cluster size Nc versus Pe along the lines. The blue andpurple curves show the strong increase of Nc when theclustering state 2 is entered, since the respective lines arenearly normal to the transition plane. Tilting the lines more,the increase of Nc becomes smoother. In particular, the greengraph shows an almost linear increase of Nc in the range Pe =10−20. Thus, Figure 6 demonstrates that the relation of clustersize and swimming speed might take different forms dependingon the relation between fuel concentration c0, activity Pe, andphoretic strengths ζtr and ζrot.

    3.3. Clustering States

    To classify the different collapsed states in Figure 1 and todescribe the hexagonal order in the N-particle cluster, weintroduce the global 6-fold bond orientational parameter

    Figure 4. Mean cluster size Nc plotted against ζtr for different arealfractions σ. The dashed line indicates the transition between dynamicclustering states 1 and 2. The rotational diffusiophoretic parameter isζrot = −0.38. Reprinted with permission from ref 29. Copyright 2015European Physical Journal.

    Figure 5. State diagram ζtr versus Pe at ζrot = −0.38 and surfacefraction σ = 0.05. The mean cluster size Nc for the gaslike anddynamic-clustering state 1 are color-coded. Reprinted with permissionfrom ref 27. Copyright 2014 American Physical Society.

    Figure 6. Mean cluster size Nc versus Pe for different lines in theζtr−ζrot−Pe parameter space. The lines are defined via a para-metrization with x ∈ [0, 1], where Pe varies as in the experiments of,25Pe = 9.5 + 11.5x, ζtr = 4.8 + 16.6x, and ζrot = −0.16 − ζ0x. Theparameter ζ0 defines the different graphs. The transition betweenclustering states 1 and 2 roughly occurs at the intersection of the fittedtwo straight lines. Reprinted with permission from ref 27. Copyright2014 American Physical Society.

    Accounts of Chemical Research Article

    DOI: 10.1021/acs.accounts.8b00259Acc. Chem. Res. 2018, 51, 2681−2688

    2685

    http://dx.doi.org/10.1021/acs.accounts.8b00259

  • ≔ ∈ [ ]

    ≔ α

    =

    qN

    q

    q e

    10, 1

    with16

    k

    Nk

    k

    j

    i

    61

    6( )

    6( ) 6

    k

    kj

    6( ) (10)

    The second sum goes over the set k6( ) of six nearest neighbors

    of particle k and αkj is the angle between the line connectingparticle k to j and some prescribed axis.17 The local bondparameter q6

    (k) becomes one if all six nearest neighbors form aregular hexagon around the central colloid and the global orderparameter is one in a hexagonal lattice.The temporal evolution of q6 is plotted in Figure 7 for

    several ζtr at constant ζrot = 4.5. At positive ζtr the order

    parameter is nearly constant in time indicating a staticcrystalline cluster, while q6 < 1 results from the colloids atthe rim of the cluster, which do not have six neighbors on ahexagon. At ζtr = 0 and especially for negative ζtr, whereparticles effectively repel each other due to translationalphoretic motion, the order parameter shows increasinglystrong fluctuations. The cluster fluctuates close to the rim,where particles leave and rejoin it frequently (ζtr = −6.4 inFigure 7). A video of the fluctuating collapsed state is attachedto ref 29. Further decreasing ζtr the order parameter developsnearly regular oscillations, where the cluster oscillates betweendensely packed and a cloud of confined colloids (ζtr = −12.8 inFigure 7). When the dense packing develops, the diffusiopho-retic interaction becomes strongly screened and the particleslose their orientations toward the cluster center. Repulsion dueto translational phoresis takes over and the cluster expandsuntil the particles are oriented back to the center. This initiatesthe collapse of the cloud and the cycle starts again. Thepulsating cluster is visualized in a video attached to ref 29. Thepower spectrum of q6 shows a broad peak at a nonzerofrequency and confirms the regular oscillations.29 Finally,further decreasing ζtr, the oscillations abruptly stop and a staticcollapsed cloud forms, where the strong effective repulsionprevents direct contact between the particles (ζtr = −16.0 inFigure 7). In this state the hexagonal bond order is small sinceq6 ≈ 0.35 is close to the value q6 = 1/3 of uniformly distributedparticles.

    3.4. State Diagram without Screening

    Some of the states in the full state diagram of Figure 1discussed so far depend on the screening of the chemical field,which is implemented when the colloids are densely packed.To demonstrate its influence, we show in Figure 8 the state

    diagram in the absence of any screening for the sameparameters. A comparison reveals that without screeningdynamic clustering 1 is less pronounced producing on averagesmaller clusters and the dynamic clustering state 2 disappearscompletely. The rich phenomenology of the collapsed states atnegative ζtr is replaced by a single core−corona state, where adensely packed core is surrounded by a cloud of colloids (seeFigure 9). The extension of the core decreases, when ζtrbecomes more negative.

    3.5. Relation to Keller−Segel ModelIn the end, we illustrate here how to derive the Keller-Segelequation starting from our model for the chemically interactingactive particles. More details are presented in ref 27. TheLangevin eqs 4 and 5 are equivalent to the Smoluchowskiequation for the full one-particle distribution function P(e, r,t), where we neglect direct interactions between the colloids:

    ζ

    ζ

    ∇ ∇ ∇ ∇

    ∂ = − · + · +

    + ∂ [ ∂ · ] + ∂φ φ φ

    P t v Pe P c D P

    c P D P

    e r

    e

    ( , , ) ( ) ( )

    ( ) ( )t 0 tr tr

    2

    rot rot2

    (11)

    To formulate the Smoluchowski equation, we used e = (sin φ,cos φ) and rewrote eq 5 to ∂tφi = ζrot∂φ ei·∇c + μi. One thenderives dynamic equations for the colloidal density P0(r,t) =∫ P(e, r, t)dφ, as well as the polar [P1(r,t) = ∫ e P dφ ] and

    Figure 7. Time evolution of the bond orientational parameter q6 fordifferent ζtr. Further parameters are Pe = 19, ζrot = 4.5, and σ = 0.05.Reprinted with permission from ref 29. Copyright 2015 EuropeanPhysical Journal.

    Figure 8. Full state diagram ζtr versus ζrot at Pe = 19 and σ = 0.05 inthe absence of screening. Reprinted with permission from ref 29.Copyright 2015 European Physical Journal.

    Figure 9. Snapshots of the colloidal configuration in the core−coronastate: (a) ζtr = −6.4 and (b) ζtr = −16. Other parameters are Pe = 19and ζrot = 4.1. Reprinted with permission from ref 29. Copyright 2015European Physical Journal.

    Accounts of Chemical Research Article

    DOI: 10.1021/acs.accounts.8b00259Acc. Chem. Res. 2018, 51, 2681−2688

    2686

    http://dx.doi.org/10.1021/acs.accounts.8b00259

  • the nematic ∫ φ= ⊗ −( )t PP r e e( , ) d12 2ÄÇÅÅÅÅÅÅ

    ÉÖÑÑÑÑÑÑ order parame-

    ter. The resulting hierarchy of coupled equations is closed bysetting P3 to zero. Neglecting time derivatives of P1 and P2 ontime scales much larger than the rotational diffusion time 1/Drot and also higher-order spatial derivatives, one arrivesultimately at the Keller−Segel equation for the colloidaldensity

    ζ ∇ ∇ ∇∂ = +P P c D P( )t 0 eff 0 eff2

    0 (12)

    with renormalized phoretic coefficient and effective transla-tional diffusion constant:

    ζ ζζ

    = + = +v

    DD D

    vD2

    and2eff tr

    rot 0

    roteff tr

    02

    rot (13)

    From the Keller−Segel equation the condition for thechemotactic collapse mentioned in section 3.1 can be derived.

    4. CONCLUSIONSThis Account reviews our work on self-phoretic active colloidsinteracting by self-generated chemical gradients and therebymimicking chemotaxis well-known from the biological world.The combination of translational and rotational diffusiopho-retic drift velocities with self-propulsion gives rise to effectivephoretic attraction and repulsion. Exploration of the full rangeof the drift velocities reveals that a variety of dynamic statesoccur that range from a gas-like state, over dynamic clustering1 and 2, to different types of collapsed states including a fullchemotactic collapse. The latter can be rationalized by theKeller−Segel equation, which derives from our Langevinmodel.We shortly discuss one point. In our state diagram in Figure

    1, the dynamic clustering states only appear in a narrow region,while in the experiments dynamic clustering seems to be ageneric feature.25,26 One reason could be that furtherattractive/repulsive forces between the active colloids arepresent in the experiments, which are not included in ourmodel. Indeed, in simulations with a Lennard-Jones potentialdynamic clustering with cluster size distributions similar to theones discussed in this article are reported.56,57

    Possible extensions of the Langevin model presented in thisarticle are mixtures of active and passive particles. Anexperimental realization was published recently in ref 35.Furthermore, Janus particles do not just produce a monopoledisturbance in the chemical environment but also a dipolarcontribution due to their polar character. We are currentlyexploring the consequences of such a contribution. Ultimately,a full solution of the chemical field with appropriate boundaryconditions at the colloid surface is needed for being able totreat particles, which are really close to each other. However,this will also directly influence the swimming speed of thecolloids. In addition, a full hydrodynamic treatment needs tobe included, which has been addressed in recent publica-tions.45−47 Finally, the slip velocity coefficient ζ in eq 1determines the two diffusiophoretic parameters ζrot and ζtr. Itstrongly depends on material properties. So, as a challenge tothe fabrication of self-phoretic colloids, one can ask if it ispossible to purposely tune the phoretic parameters by choosingappropriate materials, catalysts, cap sizes, and physicalmechanisms for the phoretic process?All this will help to further develop the idea of how artificial

    microswimmers can mimic biological taxis strategies to control

    their swimming paths and explore a wide field of intriguingapplications.

    ■ AUTHOR INFORMATIONCorresponding Author

    *E-mail: [email protected]

    Holger Stark: 0000-0002-6388-5390Notes

    The author declares no competing financial interest.

    Biography

    Holger Stark received his Ph.D. from University of Stuttgart andpursued postdoctoral studies at the University of Pennsylvania inPhiladelphia. After staying as a Heisenberg fellow at the University ofKonstanz and a group leader at the Max-Planck-Institute forDynamics and Self-Organization in Göttingen, he became a Professorof Theoretical Physics at the Technical University Berlin. His researchinterests lie in the areas of nonequilibrium statistical physics of softmatter and biological systems.

    ■ ACKNOWLEDGMENTSThe author thanks Oliver Pohl and Julian Stürmer forcollaborating on the topic and acknowledges funding fromthe DFG within the research training group GRK 1558 and thepriority program SPP 1726, project number STA 352/11. Theauthor also thanks Julian Stürmer for preparing the picture forthe Conspectus.

    ■ REFERENCES(1) Dusenberg, D. B. Living at Micro Scale; Harvard University Press:Cambridge, USA, 2009; Chapter 14.(2) Berg, H.; Brown, D. Chemotaxis in Escherichia coli analysed bythree-dimensional tracking. Nature 1972, 239, 500.(3) Grimm, A. C.; Harwood, C. S. Chemotaxis of Pseudomonas spp.to the polyaromatic hydrocarbon naphthalene. Appl. Environ.Microbiol. 1997, 63, 4111−4115.(4) Homma, M.; Oota, H.; Kojima, S.; Kawagishi, I.; Imae, Y.Chemotactic responses to an attractant and a repellent by the polarand lateral flagellar systems of Vibrio alginolyticus. Microbiology 1996,142, 2777−2783.(5) Amselem, G.; Theves, M.; Bae, A.; Bodenschatz, E.; Beta, C. Astochastic description of Dictyostelium chemotaxis. PLoS One 2012,7, e37213.(6) Pohl, O.; Hintsche, M.; Alirezaeizanjani, Z.; Seyrich, M.; Beta,C.; Stark, H. Inferring the Chemotactic Strategy of P. putida and E.coli Using Modified Kramers-Moyal Coefficients. PLoS Comput. Biol.2017, 13, e1005329.(7) Had̈eri, D.-P.; Rosumi, A.; Schaf̈er, J.; Hemmersbach, R.Gravitaxis in the flagellate Euglena gracilisis controlled by an activegravireceptor. J. Plant Physiol. 1995, 146, 474.(8) Marcos, H. C. F.; Powers, T. R.; Stocker, R.; et al. Bacterialrheotaxis. Proc. Natl. Acad. Sci. U. S. A. 2012, 109, 4780−4785.(9) Blakemore, R. Magnetotactic bacteria. Science 1975, 190, 377−379.(10) Garcia, X.; Rafaï, S.; Peyla, P. Light Control of the Flow ofPhototactic Microswimmer Suspensions. Phys. Rev. Lett. 2013, 110,138106.(11) Mori, I.; Ohshima, Y. Neural regulation of thermotaxis inCaenorhabditis elegans. Nature 1995, 376, 344−348.(12) Alon, U.; Surette, M. G.; Barkai, N.; Leibler, S. Robustness inbacterial chemotaxis. Nature 1999, 397, 168−171.

    Accounts of Chemical Research Article

    DOI: 10.1021/acs.accounts.8b00259Acc. Chem. Res. 2018, 51, 2681−2688

    2687

    mailto:[email protected]://orcid.org/0000-0002-6388-5390http://dx.doi.org/10.1021/acs.accounts.8b00259

  • (13) Korobkova, E.; Emonet, T.; Vilar, J. M.; Shimizu, T. S.; Cluzel,P. From molecular noise to behavioural variability in a singlebacterium. Nature 2004, 428, 574−578.(14) Sourjik, V. Receptor clustering and signal processing in E. colichemotaxis. Trends Microbiol. 2004, 12, 569−576.(15) Berg, H. E. coli in Motion; Springer, New York, 2004.(16) Bechinger, C.; Di Leonardo, R.; Löwen, H.; Reichhardt, C.;Volpe, G.; Volpe, G. Active particles in complex and crowdedenvironments. Rev. Mod. Phys. 2016, 88, 045006.(17) Zöttl, A.; Stark, H. Topical Review: Emergent behavior inactive colloids. J. Phys.: Condens. Matter 2016, 28, 253001.(18) Anderson, J. L. Colloidal Transport by Interfacial Forces. Annu.Rev. Fluid Mech. 1989, 21, 61−99.(19) Golestanian, R. Collective Behavior of Thermally ActiveColloids. Phys. Rev. Lett. 2012, 108, 038303.(20) Cohen, J. A.; Golestanian, R. Emergent cometlike swarming ofoptically driven thermally active colloids. Phys. Rev. Lett. 2014, 112,068302.(21) Braun, M.; Würger, A.; Cichos, F. Trapping of single nano-objects in dynamic temperature fields. Phys. Chem. Chem. Phys. 2014,16, 15207.(22) Bickel, T.; Zecua, G.; Würger, A. Polarization of active Janusparticles. Phys. Rev. E 2014, 89, 050303.(23) Ten Hagen, B.; Kümmel, F.; Wittkowski, R.; Takagi, D.;Löwen, H.; Bechinger, C. Gravitaxis of asymmetric self-propelledcolloidal particles. Nat. Commun. 2014, 5, 4829.(24) Lozano, C.; ten Hagen, B.; Löwen, H.; Bechinger, C. Phototaxisof synthetic microswimmers in optical landscapes. Nat. Commun.2016, 7, 12828.(25) Theurkauff, I.; Cottin-Bizonne, C.; Palacci, J.; Ybert, C.;Bocquet, L. Dynamic Clustering in Active Colloidal Suspensions withChemical Signaling. Phys. Rev. Lett. 2012, 108, 268303.(26) Palacci, J.; Sacanna, S.; Steinberg, A. P.; Pine, D. J.; Chaikin, P.M. Living Crystals of Light-Activated Colloidal Surfers. Science 2013,339, 936.(27) Pohl, O.; Stark, H. Dynamic clustering and chemotacticcollapse of self-phoretic active particles. Phys. Rev. Lett. 2014, 112,238303.(28) Saha, S.; Golestanian, R.; Ramaswamy, S. Clusters, asters, andcollective oscillations in chemotactic colloids. Phys. Rev. E 2014, 89,062316.(29) Pohl, O.; Stark, H. Self-phoretic active particles interacting bydiffusiophoresis: A numerical study of the collapsed state and dynamicclustering. Eur. Phys. J. E: Soft Matter Biol. Phys. 2015, 38, 93.(30) Wang, W.; Duan, W.; Ahmed, S.; Sen, A.; Mallouk, T. E. FromOne to Many: Dynamic Assembly and Collective Behavior of Self-Propelled Colloidal Motors. Acc. Chem. Res. 2015, 48, 1938−1946.(31) Banigan, E. J.; Marko, J. F. Self-propulsion and interactions ofcatalytic particles in a chemically active medium. Phys. Rev. E: Stat.Phys., Plasmas, Fluids, Relat. Interdiscip. Top. 2016, 93, 012611.(32) Yan, W.; Brady, J. F. The behavior of active diffusiophoreticsuspensions: An accelerated Laplacian dynamics study. J. Chem. Phys.2016, 145, 134902.(33) Moran, J. L.; Posner, J. D. Phoretic Self-Propulsion. Annu. Rev.Fluid Mech. 2017, 49, 511.(34) Ginot, F.; Theurkauff, I.; Detcheverry, F.; Ybert, C.; Cottin-Bizonne, C. Aggregation-fragmentation and individual dynamics ofactive clusters. Nat. Commun. 2018, 9, 696.(35) Singh, D. P.; Choudhury, U.; Fischer, P.; Mark, A. G. Non-Equilibrium Assembly of Light-Activated Colloidal Mixtures. Adv.Mater. 2017, 29, 1701328.(36) Ibele, M.; Mallouk, T. E.; Sen, A. Schooling Behavior of Light-Powered Autonomous Micromotors in Water. Angew. Chem., Int. Ed.2009, 48, 3308.(37) Ibele, M. E.; Lammert, P. E.; Crespi, V. H.; Sen, A. Emergent,collective oscillations of self-mobile particles and patterned surfacesunder redox conditions. ACS Nano 2010, 4, 4845.

    (38) Liebchen, B.; Marenduzzo, D.; Pagonabarraga, I.; Cates, M. E.Clustering and Pattern Formation in Chemorepulsive Active Colloids.Phys. Rev. Lett. 2015, 115, 258301.(39) Liebchen, B.; Marenduzzo, D.; Cates, M. E. PhoreticInteractions Generically Induce Dynamic Clusters and Wave Patternsin Active Colloids. Phys. Rev. Lett. 2017, 118, 268001.(40) Simmchen, J.; Katuri, J.; Uspal, W. E.; Popescu, M. N.;Tasinkevych, M.; Sańchez, S. Topographical pathways guide chemicalmicroswimmers. Nat. Commun. 2016, 7, 10598.(41) Sengupta, S.; Dey, K. K.; Muddana, H. S.; Tabouillot, T.; Ibele,M. E.; Butler, P. J.; Sen, A. Enzyme Molecules as Nanomotors. J. Am.Chem. Soc. 2013, 135, 1406−1414.(42) Dey, K. K.; Das, S.; Poyton, M. F.; Sengupta, S.; Butler, P. J.;Cremer, P. S.; Sen, A. Chemotactic Separation of Enzymes. ACS Nano2014, 8, 11941−11949.(43) Duan, W.; Wang, W.; Das, S.; Yadav, V.; Mallouk, T. E.; Sen, A.Synthetic Nano- and Micromachines in Analytical Chemistry:Sensing, Migration, Capture, Delivery, and Separation. Annu. Rev.Anal. Chem. 2015, 8, 311−333.(44) Zhao, X.; Palacci, H.; Yadav, V.; Spiering, M. M.; Gilson, M. K.;Butler, P. J.; Hess, H.; Benkovic, S. J.; Sen, A. Substrate-drivenchemotactic assembly in an enzyme cascade. Nat. Chem. 2017, 10,311−317.(45) Huang, M.-J.; Schofield, J.; Kapral, R. A microscopic model forchemically-powered Janus motors. Soft Matter 2016, 12, 5581−5589.(46) Huang, M.-J.; Schofield, J.; Kapral, R. Chemotactic andhydrodynamic effects on collective dynamics of self-diffusiophoreticJanus motors. New J. Phys. 2017, 19, 125003.(47) Colberg, P. H.; Kapral, R. Many-body dynamics of chemicallypropelled nanomotors. J. Chem. Phys. 2017, 147, 064910.(48) Robertson, B.; Stark, H.; Kapral, R. Collective orientationaldynamics of pinned chemically-propelled nanorotors. Chaos 2018, 28,045109.(49) Jin, C.; Krüger, C.; Maass, C. C. Chemotaxis andautochemotaxis of self-propelling droplet swimmers. Proc. Natl.Acad. Sci. U. S. A. 2017, 114, 5089−5094.(50) Cates, M. E.; Tailleur, J. Motility-Induced Phase Separation.Annu. Rev. Condens. Matter Phys. 2015, 6, 219−244.(51) Bialke,́ J.; Speck, T.; Löwen, H. Active colloidal suspensions:Clustering and phase behavior. J. Non-Cryst. Solids 2015, 407, 367−375.(52) Keller, E. F.; Segel, L. A. J. Theor. Biol. 1970, 26, 399−415.(53) Masoud, H.; Shelley, M. J. Collective surfing of chemicallyactive particles. Phys. Rev. Lett. 2014, 112, 128304.(54) Peruani, F.; Starruß, J.; Jakovljevic, V.; Søgaard-Andersen, L.;Deutsch, A.; Bar̈, M. Phys. Rev. Lett. 2012, 108, 098102.(55) Moran, J. L.; Wheat, P. M.; Posner, J. Locomotion ofelectrocatalytic nanomotors due to reaction induced chargeautoelectrophoresis. Phys. Rev. E 2010, 81, 065302.(56) Mognetti, B. M.; Šaric,́ A.; Angioletti-Uberti, S.; Cacciuto, A.;Valeriani, C.; Frenkel, D. Living Clusters and Crystals from Low-Density Suspensions of Active Colloids. Phys. Rev. Lett. 2013, 111,245702.(57) Alarcoń, F.; Valeriani, C.; Pagonabarraga, I. Morphology ofclusters of attractive dry and wet self-propelled spherical particlesuspensions. Soft Matter 2017, 13, 814.

    Accounts of Chemical Research Article

    DOI: 10.1021/acs.accounts.8b00259Acc. Chem. Res. 2018, 51, 2681−2688

    2688

    http://dx.doi.org/10.1021/acs.accounts.8b00259