unknowns - arxivcolloids. soon after, the seminal work of asakura and oo-sawa (ao) showed that, in a...

10
Hunting Mermaids in Real Space: Known Knowns, Known Unknowns and Unknown Unknowns C. Patrick Royall 1, 2, 3 1 School of Chemistry, Cantock’s Close, University of Bristol, BS8 1TS, UK 2 Centre for Nanoscience and Quantum Information, Bristol BS8 1FD, UK 3 HH Wills Physics Laboratory, University of Bristol, Bristol BS8 1TL, UK We review efforts to realise so-called mermaid (or short-ranged attraction/long ranged repulsion) interactions in 3d real space. The repulsive and attractive contributions to these interactions in charged colloids and colloid-polymer mixtures, may be accurately realised, by comparing particle- resolved studies with colloids to computer simulation. However, when we review work where these interactions have been combined, despite early indications of behaviour consistent with predictions, closer analysis reveals that in the non-aqueous systems used for particle-resolved studies, the idea of summing the attractive and repulsive components leads to wild deviations with experiment. We suggest that the origin lies in the weak ion dissociation in these systems with low dielectric constant solvents. Ultimately this leads even to non-centro-symmetric interactions and a new level of complexity in these systems. INTRODUCTION Colloids provide important models for liquids and solids, and among their properties that lead to this is the simplicity of their interactions, which may often be treated as being spherically symmetric [1, 2]. Perhaps the simplest of these models is the hard sphere, which was fa- mously demonstrated in experiments with sterically stab- lised colloids [3], challenging though it may be realise per- fectly hard interactions in practise, as the colloids always carry some electrostatic charge [4]. Rather earlier than the quasi-hard spheres, the Derjaguin-Landau-Verwey- Overbeek (DLVO) theory [5] provided the theoretical foundations for a tunable, long-range repulsion between colloids. Soon after, the seminal work of Asakura and Oo- sawa (AO) showed that, in a solution of non-adsorbing polymers, colloids experience a tuneable attraction due to the polymer degrees of freedom [6, 7]. These two ap- proaches provide a framework by which attractions, and repulsions, between colloids may be manipulated. The combination of both the electrostatic repulsion of DLVO and the AO attractions suggests that colloids may be tuned to have “mermaid” interactions, so-called ow- ing to the “attractive head” and “repulsive tail”, Fig. 1 [8]. Also known as short–ranged attraction–long–ranged repulsion (SALR) systems, these exhibit a rich and ex- citing phase behaviour, since the competing interactions lead to a complex energy landscape [9–13]. Under such competing interactions, mesophases are predicted, such as lamellae, gyroid phases and clusters [9, 14–17]. The clusters formed may themselves order into exotic phases such as cluster crystals [18–20] and co-existing cluster fluids [21]. One might imagine then, that given the tuneablity of colloidal systems and that these competing interactions exhibit such a rich phase behavior, then colloidal sys- tems, imaged in the glory of 3d real space [1] would lend themselves to the realisation of the exotic phases thus predicted by simulation and theory. Yet no ordered phase in a system with competing interactions has ever been observed in 3d real space, and the reasons underly- ing this paradox form the subject of this short review. We emphasise the 3d real space here, because stripe- like lamellar phases and large clusters have indeed been found in 2d systems on an air-water interface, which are well-described by mermaid interactions [14, 22–24] and other approaches, such as using tilted rotating electric fields hold considerable promise [25]. Given that ordered phases have been observed in 2d experiments, here we focus on 3d particle-resolved studies, that is to say work done using confocal microcopy with density- and refrac- tive index-matched systems [1]. Before proceeding, we note that one of the interesting features of the “mermaid” potential is that it may be interpreted as a basic model for ionic systems [26], cement [27] and globular proteins such as lyzozyme [10, 28, 29] whose phase behaviour can be compared with colloids [28–31]. We shall therefore make connection to work on protein systems closely connected to the colloids where appropriate. In no sense is this short review intended to be com- prehensive. We have chosen to focus on our own field, particle-resolved studies of colloids. We humbly beg the learned reader for forgiveness if, particularly outside this field, we have neglected to mention relevant work, or in- deed if we are to interpret work in a manner that seems at odds with the prevailing view of that field. This paper is organised as follows. In section we dis- cuss the principles by which well-known interactions be- tween colloids may be tuned such that a mermaid-like potential may be realised. The two contributions – the attractive head and repulsive tail – are described in a little more detail in sections and respectively. In sec- tion we outline the experimental systems suitable for realising such an interaction in real space in 3d, using particle-resolved studies. We then review some experi- arXiv:1805.00560v1 [cond-mat.soft] 1 May 2018

Upload: others

Post on 02-Aug-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Unknowns - arXivcolloids. Soon after, the seminal work of Asakura and Oo-sawa (AO) showed that, in a solution of non-adsorbing polymers, colloids experience a tuneable attraction due

Hunting Mermaids in Real Space: Known Knowns, Known Unknowns and UnknownUnknowns

C. Patrick Royall1, 2, 3

1School of Chemistry, Cantock’s Close, University of Bristol, BS8 1TS, UK2Centre for Nanoscience and Quantum Information, Bristol BS8 1FD, UK3HH Wills Physics Laboratory, University of Bristol, Bristol BS8 1TL, UK

We review efforts to realise so-called mermaid (or short-ranged attraction/long ranged repulsion)interactions in 3d real space. The repulsive and attractive contributions to these interactions incharged colloids and colloid-polymer mixtures, may be accurately realised, by comparing particle-resolved studies with colloids to computer simulation. However, when we review work where theseinteractions have been combined, despite early indications of behaviour consistent with predictions,closer analysis reveals that in the non-aqueous systems used for particle-resolved studies, the ideaof summing the attractive and repulsive components leads to wild deviations with experiment.We suggest that the origin lies in the weak ion dissociation in these systems with low dielectricconstant solvents. Ultimately this leads even to non-centro-symmetric interactions and a new levelof complexity in these systems.

INTRODUCTION

Colloids provide important models for liquids andsolids, and among their properties that lead to this isthe simplicity of their interactions, which may often betreated as being spherically symmetric [1, 2]. Perhaps thesimplest of these models is the hard sphere, which was fa-mously demonstrated in experiments with sterically stab-lised colloids [3], challenging though it may be realise per-fectly hard interactions in practise, as the colloids alwayscarry some electrostatic charge [4]. Rather earlier thanthe quasi-hard spheres, the Derjaguin-Landau-Verwey-Overbeek (DLVO) theory [5] provided the theoreticalfoundations for a tunable, long-range repulsion betweencolloids. Soon after, the seminal work of Asakura and Oo-sawa (AO) showed that, in a solution of non-adsorbingpolymers, colloids experience a tuneable attraction dueto the polymer degrees of freedom [6, 7]. These two ap-proaches provide a framework by which attractions, andrepulsions, between colloids may be manipulated.

The combination of both the electrostatic repulsion ofDLVO and the AO attractions suggests that colloids maybe tuned to have “mermaid” interactions, so-called ow-ing to the “attractive head” and “repulsive tail”, Fig. 1[8]. Also known as short–ranged attraction–long–rangedrepulsion (SALR) systems, these exhibit a rich and ex-citing phase behaviour, since the competing interactionslead to a complex energy landscape [9–13]. Under suchcompeting interactions, mesophases are predicted, suchas lamellae, gyroid phases and clusters [9, 14–17]. Theclusters formed may themselves order into exotic phasessuch as cluster crystals [18–20] and co-existing clusterfluids [21].

One might imagine then, that given the tuneablity ofcolloidal systems and that these competing interactionsexhibit such a rich phase behavior, then colloidal sys-tems, imaged in the glory of 3d real space [1] wouldlend themselves to the realisation of the exotic phases

thus predicted by simulation and theory. Yet no orderedphase in a system with competing interactions has everbeen observed in 3d real space, and the reasons underly-ing this paradox form the subject of this short review.We emphasise the 3d real space here, because stripe-like lamellar phases and large clusters have indeed beenfound in 2d systems on an air-water interface, which arewell-described by mermaid interactions [14, 22–24] andother approaches, such as using tilted rotating electricfields hold considerable promise [25]. Given that orderedphases have been observed in 2d experiments, here wefocus on 3d particle-resolved studies, that is to say workdone using confocal microcopy with density- and refrac-tive index-matched systems [1].

Before proceeding, we note that one of the interestingfeatures of the “mermaid” potential is that it may beinterpreted as a basic model for ionic systems [26], cement[27] and globular proteins such as lyzozyme [10, 28, 29]whose phase behaviour can be compared with colloids[28–31]. We shall therefore make connection to work onprotein systems closely connected to the colloids whereappropriate.

In no sense is this short review intended to be com-prehensive. We have chosen to focus on our own field,particle-resolved studies of colloids. We humbly beg thelearned reader for forgiveness if, particularly outside thisfield, we have neglected to mention relevant work, or in-deed if we are to interpret work in a manner that seemsat odds with the prevailing view of that field.

This paper is organised as follows. In section we dis-cuss the principles by which well-known interactions be-tween colloids may be tuned such that a mermaid-likepotential may be realised. The two contributions – theattractive head and repulsive tail – are described in alittle more detail in sections and respectively. In sec-tion we outline the experimental systems suitable forrealising such an interaction in real space in 3d, usingparticle-resolved studies. We then review some experi-

arX

iv:1

805.

0056

0v1

[co

nd-m

at.s

oft]

1 M

ay 2

018

Page 2: Unknowns - arXivcolloids. Soon after, the seminal work of Asakura and Oo-sawa (AO) showed that, in a solution of non-adsorbing polymers, colloids experience a tuneable attraction due

2

FIG. 1: A “mermaid”, or SALR potential. The “attractivehead” leads to condensation, but the “repulsive tail” opposesthis effect, leading to competing interactions and a complexenergy landscape [32].

ments which set out to measure the two contributionsto the mermaid potential in section . Given these con-tributions, we then consider attempts to realise actualmermaid-type potentials in section . The details of theinteractions in some of these papers are considered insection , which leads us to address the observation ofqualitative breakdown of the idea that one can sum theattractive and repulsive components of a mermaid poten-tial in section . We present our conclusions and providesome points for future directions in section .

HOW TO MAKE A MERMAID: INTERACTIONSBETWEEN COLLOIDS

Attractive Head

As alluded to in the introduction, in order to re-alise mermaid interactions, one seeks a short-rangedattraction and a long-ranged repulsion. There aremany ways to induce interactions between colloids[1, 2]. Tuneable attractions which have been imple-mented in particle-resolved studies of colloids range fromdepletion/Asakura-Oosawa [33], dipolar (rotating field in2d) [34], critical Casimir [35]. Other mechanisms in-clude tuning stablisation against van der Waals attrac-tions [36].

In the case of mermaid-type interactions, the Asakura-Oosawa or depletion mechanism has usually been used.For polymers that are substantially smaller than thecolloids, the resulting mixture can be described by anAsakura-Oosawa (AO) model, which treats the polymer

molecules as an ideal gas with hard interactions with thecolloids [37–40]. The AO effective interaction potentialbetween two colloids can be written as:

βuAO(r) =

∞ for r < σπ(2RG)3zPR

6(1+q)3

q3 ×{1− 3r

2(1+q)σ + r3

2(1+q)3σ3 } for σ ≤ r < σ

+(2RG)

0 for r ≥ σ + (2RG)

(1)where β is 1/kBT . The polymer fugacity zPR is equal tothe number density ρPR of ideal polymers in a reservoir atthe same chemical potential as the colloid-polymer mix-ture. The polymer-colloid size ratio q = 2RG/σ whereRG is the polymer radiius of gyration and σ is the col-loid diameter.

Repulsive Tail

Like the mechanisms for attraction noted above, arange of methods have been used in particle-resolvedstudies of colloids to yield tuneable, long-range repul-sions. In addition to electrostatic (DLVO) [5] repulsions,tuneable magnetic dipolar interactions have been demon-strated [41], and electric dipolar interactions are possible(also in 2d). Pertinent to attempts to realise mermaid in-teractions are the electrostatic interactions. In its linear-Poisson-Boltzmann (DLVO) form, the electrostatic inter-action between two colloids takes a Yukawa form.

βuyuk(r) =

{∞ for r < σ

βεyukexp(−κ(r−σ))

r/σ for r ≥ σ(2)

where r is the center–to–center separation of the two col-loids. The contact potential is given by

βεyuk =Z2

(1 + κσ/2)2λBσ

(3)

where Z is the colloid charge, κ is the inverse Debyescreening length and λB is the Bjerrum length. The in-verse Debye screening length is given by

κ =√

4πλBρion (4)

where ρion is the number density of small monovalentions. Note that here a (monovalent) salt ion pair wouldcount as two ions.

PARTICLE-RESOLVED STUDIES

To understand more about how we might realisemermaid-type potentials with particle resolved studies,

Page 3: Unknowns - arXivcolloids. Soon after, the seminal work of Asakura and Oo-sawa (AO) showed that, in a solution of non-adsorbing polymers, colloids experience a tuneable attraction due

3

we need to consider the particular experimental modelsystems used. A more detailed discussion may be foundin ref. [1] and with a particular focus on the interactionsbetween the particles in ref. [4], so here we briefly notethe salient points.

Particle resolved studies uses relatively large colloids(often 3000 nm diameter), so that sedimentation can bea major problem. This means that the particles must bedispersed in a density-matching solvent, which is usuallya mixture of two solvents, one with a density larger thanand one with a density smaller than the particles. Thesecond requirement is that the solvent has the same re-fractive index as the colloids, enabling high-resolution 3doptical imaging with confocal microscopy. We note thatone elegant means to meet these criteria is to use microgelparticles, which are essentially densely cross linked poly-mers. Like (linear) polymers, these swell, such that thevast majority of the material inside the particle is solvent.This means that good density- and index-matching areintrinsic to the system. However, with the odd notableexception [42], most work on particle-resolved studies in3d has focussed on solid particles.

Among systems with solid colloidal particles, thoseusing poly-methyl methacrylate (PMMA) particles (thesame material as in the original hard sphere work ofPusey and van Megen [3]) have dominated the field [1].Now the solvents originally used for light-scattering stud-ies which predated particle-resolved studies, and usedsmaller (typically between 200 and 400 nm diameter) par-ticles had a very low dielectric constant of around two.There was some flexibility of solvent choice, as these par-ticles were small enough that density matching was notrequired. In any case, it appears that the main deviationof hard-sphere like behaviour came from the steric stabil-isation, which induces a slight degree of softness [4, 43].

The larger particles required for particle-resolved stud-ies necessitated density matching solvents, in the form ofhalogenated solvents such as cyclohexyl bromide. Thedensity-matching solvents typically used have a some-what larger dielectric constant of for example 5.37 inthe case of the density-matching mixture of cis-decalinand cyclohexyl bromide [44, 45]. The change in Bjer-rum length (the interaction range over which two ele-mentary charges have an energy of interaction equal tothe thermal energy kBT ) from around 30 nm to 8 nmhad significant implications for the degree of ionic dis-sociation: very little in the older, dielectric constant twosolvent based systems, but the newer model systems suit-able for particle-resolved studies exhibited enough iondissociation that the electrostatics, while weak comparedto aqueous system were nevertheless strong enough thatthe phase behaviour exhibited wild deviations from hardspheres, with “low-density crystals” at volume fractionsφ ∼ 0.01 [44, 46–48]. In other words, the increase in sizeof the colloids, to 3000 nm for particle-resolved studies,necessitating the use of a density-matching solvent, led

FIG. 2: Radial distribution functions for charged colloids atvarious volume fractions (here denoted η). Circles are de-termined from coordinate data from particle-resolved studies.These are compared with simulation data using the Yukawapotential (solid lines) and primitive model (top, dashed).Data offset for clarity [45].

to a fundamental change in the behaviour of the system:the particles exhibited significant repulsions, acting overdistances up to tens of microns, ideal for realising mer-maid type interactions.

It is worth nothing that even for these density-matching solvents, the dielectric constant can be tuned.While cyclohexyl bromide (and its relative cycloheptylbromide) remain the most popular, combinations involv-ing carbon tetrachloride [33] and tetrachloro ethylene [49]lead to density matching solvents with rather lower di-electric constants. The lower dielectric constant wouldthen suppress ion dissociation, leading to a reduction incharging, as exploited by Klix et al. (section ) [50].

The nature of the charging in these low-dielectric con-stant systems is complex and poorly understood [44, 47,50]. This leads to behaviour that under certain condi-tions deviates wildly from expectations. Sadly, we shallsee that precisely those conditions required for mermaid-type behaviour, i.e. where the particles are close together(short range attraction) and far apart (long-range repul-sion) correspond to such deviations from the expectedbehaviour.

ATTRACTIONS AND REPULSIONS INPARTICLE-RESOLVED STUDIES

Before we explore successes and, as we shall see, moreexplicitly, failures, to realise mermaid potentials in realspace, let us first consider the components of the inter-action – the attractive head and the repulsive tail. It ispossible to measure interactions between colloidal par-ticles and glass walls with total internal reflection mi-croscopy [51, 52], and between pairs of colloidal particleswith optical tweezers [53]. Optical tweezers were used

Page 4: Unknowns - arXivcolloids. Soon after, the seminal work of Asakura and Oo-sawa (AO) showed that, in a solution of non-adsorbing polymers, colloids experience a tuneable attraction due

4

to measure the AO attraction [54], however the samemethod [55] obtained spurious results for the related bi-nary hard sphere system, which also exhibits depletion[56, 57].

While determining the interactions is an importantstep, demonstrating the potential of a system to exhibitan exotic phase behaviour (presuming it were able toequilibrate) can raise major questions as to whether asystem can in fact be described by a simple interaction.In addition to the issues of equilibration (leading oftento disordered non-equilibrium states, sometimes termed“junk” [58–61]), the interactions of charged colloids forexample are intrinsically density-dependent, due to thefact that the counter-ion contribution to the electrostaticscreening term is itself dependent on the colloid concen-tration, and this effect has been observed in experiments[62, 63]. More drastic effects can also be observed, likelydue to counter-ion condensation leading to unusual phasebehaviour in the form of re-entrant melting [45]. Otherdeviations from the expectations of Eq. 2 include many-body interactions (i.e. a breakdown of pairwise addivity)[62] and non-centro-symmetric interactions in colloidalcrystals [64]. Other than these last two observations, allof these effects are consistent with the Yukawa interac-tion, albeit with state-dependent interaction parameters.And significant though the observations of Refs. [62] and[64] are, the magnitude of the deviation from Yukawa be-haviour is not large.

Attempts to directly compare the results of particle-resolved studies experiments with charged colloids haveusually resulted in behaviour consistent with a Yukawadescription [45, 47, 62, 65]. Such direct comparison typ-ically exploits the fact that, for an isotropic, pairwiseadditive system, the radial distribution function g(r) isuniquely determined by the pair interaction [66]. Such astatement is true in principle, but often in the case of adense fluid, the g(r) can be rather insensitive to the pairinteraction – an observation than underlies the idea thatthe hard core (which may be an effective hard core inthe case of charged colloids [67]) dominates the structureof such systems [68]. Nevertheless, under typical exper-imental conditions (Fig. 2), an accuracy of around 20%is possible in the parameters εyuk and κ that determinethe Yukawa interaction Eq. 2 [45, 47].

Figure 2 shows the success of the Yukawa descrip-tion. The parameters obtained are close to those ofthe primitive model, developed in computer simulationby Vladimir Lobaskin and Per Linse [69] and here im-plemented by Antti-Pekka Hynninen for a much highercharge asymmetry between ions and colloids [70, 71]. Wesee that the parameters in the full primitive model case(top line, Fig, 2) is very similar to the value of the Yukawamodel (second line down from the top, Fig, 2). In short,we conclude that the Yukawa model can provide a gooddescription of the long-ranged repulsion between chargedcolloids, in particle-resolved studies.

(a) (b)

PR

PR

PR PR

FIG. 3: Radial distribution functions g(r) of colloid-polymermixtures at various polymer concentrations. Monte-Carlosimulations with polymer reservoir volume fraction φP , ac-cording to Eq. (1) (solid lines), are compared to the exper-imental results (circles). Dashed lines correspond to the re-lation g(r) ≈ exp(−βuAO(r)). Monte-Carlo simulations con-sider experimental resolution and polydispersity [57].

In the case of attractions, we can employ the samestrategy in the case of a system exhibiting Asakura-Osawa interactions, such as a colloid-polymer mixture[72, 73]. The Asakura-Oosawa interaction is rathershorter-ranged, and many-body effects are small, and in-deed vanish if the size ratio q < 0.1547 [37] and hardto detect for q = 0.25 [74]. One issue is that, becausethe interaction is shorter-ranged, particle tracking errorsare more of an issue, so they tend to be comparable insize to the structure of the interaction and thus to theresulting g(r). Such errors can be mimicked by addingGaussian-distributed noise to coordinates generated bysimulation. This leads to a good agreement between ex-periment and simulation as shown in Fig. 3 [57]. There-fore we conclude that the Asakura-Oosawa model is alsowell-represented in colloidal systems for particle-resolvedstudies. It has also been noted that generic short-rangedattractions give similar behaviour [75], which has also beseen in the correspondence of the square-well attractionand colloid-polymer mixtures [76, 77].

PUTTING IT ALL TOGETHER: MERMAIDS INREAL SPACE?

We have seen above that it is possible to realise, withreasonable accuracy, the two components of mermaidtype interactions - the long-range repulsion and the short-range attraction. Let us now consider what happenswhen the two are combined – that is to say, a suspen-sion of charged colloids has polymers added, such thatthe system exhibits a depletion interaction in addition tothe long-ranged electrostatic repulsion.

Page 5: Unknowns - arXivcolloids. Soon after, the seminal work of Asakura and Oo-sawa (AO) showed that, in a solution of non-adsorbing polymers, colloids experience a tuneable attraction due

5

FIG. 4: Indications of Mermaid-like behaviour in the formof elongated clusters. A confocal micrograph of the clustersin a sample with volume fraction φ = 0.086 and polymerconcentration cp = 3 mg cm?3 q = 0.021. Here the colloidshad a diameter of 1320 nm. Note the spacing between themonomers, indicating a significant strength and range of therepulsive interactions. Modified with permission from [78].

Early work, particularly that of the Edinburgh group(Fig. 4) appeared very promising: the colloids were foundto cluster, and to form rather elongated clusters, unlikethe more spherical clusters that would be expected in thecase of systems without the long-ranged repulsions [29,78]. A little before, inspired by the analogy with atomicnuclei (strong nuclear force versus electrostatic repulsion)such elongated clusters had in fact been predicted [79, 80].Further work followed, with “Bernal Spirals” found in asimilar system, but one where the colloid concentrationwas high enough that it percolated, i.e. formed a gel [81].Simulations using reasonable values for the interactionsreproduced the behaviour observed in the experiments,at least qualitatively, in the sense that elongated clusterswere formed [10, 82]. It seemed only a matter of timebefore the particle-resolved studies would deliver orderedphases predicted from simulation, such as lamellae [14].But no such ordered phase has ever been seen, and wedevote the remainder of this article to exploring why thismight be.

INTERACTIONS IN THE MERMAID SYSTEMS

We now consider the interaction parameters quotedin experimental realisations of Mermaid-type systems.Campbell et al. [81] report clusters and Bernal spiralsin a system in which they measured the colloid chargein a dilute suspension to be Z = 140 e per 1.5 µm di-ameter colloid, where e is the elementary charge. Ac-cording to Eq. 3, this maps to a Yukawa contact po-tential βεyuk = 35. Now such a repulsion strength ex-ceeds the kind of attraction strengths typically acces-

sible to the Asakura-Oosawa potential, at least for thepolymer-colloid size ratio in question. Using a similarDebye length and reasonable values of the attractive wellfor the AO attraction, Malins et al. found only very lim-ited clustering at βεyuk = 5, corresponding to a colloidcharge of Z = 47 e and expected none at higher Yukawacontact potentials. In another study on gels in systems ofcompeting interactions, Dibble et al. [83] quote a similarvalue of Z = 165 e per µm diameter colloid. MoreoverSedgwick et al. [78] report a charge of Z < 103 e intheir study of clustering. Although not strictly inconsis-tent, this seems rather higher than the values for whichclustering is expected.

Analysis of these studies paints a picture of anoma-lously strong repulsions, which would be a struggle forthe AO attractions to overcome. One interesting case oc-curred when the range if the electrostatic interaction wasreduced by the presence of salt [84]. While not exactly amermaid potential as here the repulsions had a compa-rable range to the attractions, simple addition of Eqs. 2and 1 gave an accurate description of the system. Otherthan the work of Kohl et al. [85], which considered simi-lar parameters, as far as we are aware, no other work hassucceeded in finding quantitative agreement with simpleaddition of the AO attraction and electrostatic repulsion.We emphasise that this is the same system as used in theother experiments (to all intents and purposes, the par-ticular PMMA synthesis run is different, but this doesnot affect the qualitative behaviour). The only differ-ence is that (presumably) the higher ionic strength cor-responding to the shorter Debye length means that thereare sufficient ions to suppress the effects we discuss laterin section .

Worse was to follow. The numbers quoted above sug-gest that while the repulsions seem anomalously strong,i.e. too strong for clusters to form, the difference was notwild, i.e. less than an order of magnitude. This changedwith the work of Klix et al. [86]. Figure 5 shows a g(r) fit-ted with results from a simulation following the Yukawamodel (Eq. 2), much like those in Fig. 2. However,the volume fraction was very much lower, and requiringneutrality of the overall system (by balancing the colloidcharge with counter-ions) places constraints on the De-bye length through Eq. 4. The estimate for the contactpotential was a staggering βεyuk ≈ 1000. It is hard toimagine how clusters might form in this system, yet, asFig. 6 shows, indeed polymer-induced depletion interac-tions nevertheless led to clustering.

Not only did the system cluster (and gel) upon ad-dition of polymer, it aged by emission of particles fromthe cluster, as Fig. 6 shows. Overall the cluster sizethroughout the system fell measurably. The electrostaticrepulsions held the system in a glassy state, with peculiarsub-diffusive dynamics, even at volume fractions as lowas φ ≈ 0.01 [86]. One possible explanation of this odd be-haviour was that somehow the charge was acquired after

Page 6: Unknowns - arXivcolloids. Soon after, the seminal work of Asakura and Oo-sawa (AO) showed that, in a solution of non-adsorbing polymers, colloids experience a tuneable attraction due

6

0 5 10 15 20 25 30 350.0

0.5

1.0

1.5

2.0

g(r)

r/

FIG. 5: Radial distribution functions. Dashed red line corre-sponds to a colloid charge Z = 400 e, solid cyan to Z = 800 e.We assumed that the Debye screening length was dominatedby the colloidal counterions, in other words the system is closeto the salt-free limit. This leads to a fitting which dependssolely upon Z. Lower values of Z gave poor fits, higher valuesof Z led to crystallization [86].

the colloids had clustered or gelled. This would explainthe aging behaviour, but it still seems odd that the clus-ters and gels remained even somewhat intact under suchmassive electrostatic repulsions.

QUALITATIVE BREAKDOWN OF THEYUKAWA DESCRIPTION: ION CONDENSATION

Klix et al. [50] also considered the case when the col-loid charge was very weak, comparable to values usedin computer simulation studies [10, 87]. This they ef-fected by tuning the dielectric constant of the solvent tobe close to two. Here they again found clustering, andconsidered each cluster as a separate system, which wasshown to be reasonable for the parameters of the system,notably that the interactions between the clusters weresmall [88], so the energy landscale of each cluster couldbe considered in isolation, allowing an analogy to atomicand molecular systems [89, 90]. With careful mapping ofthe interaction parameters to computer simulation, Ma-lins et al. [87] found that upon increasing the attrac-tion strength, almost all four-membered clusters formedtetrahedra, five-membered triangular biprisms, while six-membered clusters had competing populations of octa-hedra and polytetrahedra, as also found in experimentson “sticky spheres” (with no long-range repulsion) [91].However in the experiments with the mermaid-type sys-tem, the yield of tetrahedra was only 20%, with the sameholding for the triangular biprisms and polytetrahedra[50].

Conductivity measurements suggested that upon ad-

dition of polymer (and clustering), the colloid chargedropped significantly [50]. This observation was consis-tent with previous work which had shown that the colloidcharge drops strongly upon increasing the volume frac-tion in the absence of polymer, i.e. a purely repulsivesystem [45], so that one could even imagine the clustersas being locally at high volume fraction (and thus hav-ing a lower charge). This is even hinted at in imagessuch as Fig. 4 where the monomers are well-separated,indicating a strong, long-ranged repulsion.

Returning to the work of Klix et al. [50], noting thatthe Bjerrum length, at some 23 nm was so large that ionscould interact with multiple binding sites on adjacent col-loids in a cluster, through an extension of the Primitivemodel to an explicit site-binding model such that thecolloid charge is represented through charging sites onthe surface of the particle, the authors suggested thation condensation between colloids could lead to signifi-cant charge asymmetry (Fig. 7). This is caused by ionspreferentially condensing around contact points betweentwo particles. Such anisotropic ion condensation wouldlead to a breakdown in the spherically symmetric chargedistribution around the colloid implicit in Eq. 2. Thiswould then suggest an energy barrier sufficient to pre-vent the particles forming the tetrahedra (and triangularbiprisms and polytetrahedra). This argument requiredthat the colloid dynamics were comparable to those ofthe small counter- and co-ions. Usually this is absolutelynot the case, but the ionic concentration in this systemwas so low that the time taken for the ions to diffuse theirseparation was on the 0.1 s timescale, comparable to thatof the colloids. Thus the case was made for a breakdownin the Yukawa description in the repulsions preventingthe system reaching its ground state, for clusters of fouror more particles [50].

DISCUSSION AND CONCLUSIONS

We have seen that while the components of mer-maid potentials, the short range attraction and thelong-ranged repulsion can be accurately obtained in3d particle-resolved studies, their combination remainshighly problematic. As described in section . The shortrange attraction is well-captured by colloid-polymer mix-tures, while the long-ranged repulsion is found in chargedcolloids, particularly in the low dielectric constant sol-vents characteristic of particle-resolved studies, whereweak ion dissociation leads to very long Debye screen-ing lengths, up to tens of microns.

However, putting these together leads to a breakdownin the idea that a simple summation of the attractiveand repulsive components will describe the system. Inmost of the work which addresses mermaid-type interac-tions (section ), the colloid-colloid repulsion seems to beanomalously high. In one case the Yukawa contact po-

Page 7: Unknowns - arXivcolloids. Soon after, the seminal work of Asakura and Oo-sawa (AO) showed that, in a solution of non-adsorbing polymers, colloids experience a tuneable attraction due

7

t=0 t=1.16 t=2.32 t=3.48 t=4.64 t=5.8

FIG. 6: Aging mechanism of a cluster glassy state. An emission process from a 5-membered cluster to a 4-membered cluster,as shown by arrows at a volume fraction of φ = 0.051 and polymer concentration cp = 5.158 g/l. Time t is expressed in unitsof the structural relaxation time divided by 1000. Particles are 1.95 µm in diameter. [86].

FIG. 7: Simulation snapshot of the explicit site PrimitiveModel. Here the separation between the colloid surfaces isset to h = 0.05σ. Binding sites on the colloid surface and freeions

are shown in red and cyan, respectively (actual size). Noteenhanced condensation of ions between colloids. [50].

tential is some 1000 kBT, wildly in excess of that achiev-able by the Asakura-Oosawa attraction. Furthermore,the large Bjerrum length in some cases can lead to asym-metry in the interactions, i.e. a breakdown of the DLVOpicture of treating the electrostatic repulsion as a Yukawainteraction. It is our opinion that it is challenging to re-alise the kind of ordered phases, such as lamellae and gy-roid phases, in the systems used for 3d particle-resolvedstudies, based on polymethyl methacrylate colloids in lowdielectric constant solvents.

This observation begs the question of what systemsmight prove more amenable to such ordering. We notedthe early 2d work [24], in which ordering into lamellae wasseen. Now the interactions at interfaces are notoriouslycomplex, and in any case we are mainly interested in 3dsystems. One possibility would be use a system wherethe electrostatic interactions are better understood, forexample an aqueous system. However, in aqueous sys-tems, the Debye length is typically much smaller, as theionic strength is typically much higher. It would be pos-sible to use a solvent of intermediate dielectric constant,in the hope that the Debye length would still be longenough that the electrostatic repulsions could be long-ranged [92–94]. Alternatively, smaller particles might al-

leviate the need to match the density, as the sedimen-tation would be very much reduced by, say, an order ofmagnitude drop in the particle diameter. This wouldthen perhaps provide a fruitful route to realising the kindof structures predicted for mermaid potentials, control-lably, in 3d real space.

Acknowledgments

In addition to Per Linse, we dedicate this article alsoto Antti-Pekka Hynninen, whose brilliance and humourwas an inspiration to many. We are grateful to GeorgeStell who coined the term “mermaid potential”. Wethank Patrick Charbonneau for his insightful commentsregarding Junk experiments, John Russo for help withthe graphics, Bob Evans for his insight into the mermaidstate and Paul Bartlett, Alfons van Blaaderen, AlinaCiach, Marjolein Dijkstra, Jeroen van Duijneveldt, Mir-jam Leunissen, Wilson Poon, Willem Kegel, Chris Klix,Francesco Sciortino, Richard Sear, Hajime Tanaka andNigel Wilding for constructive discussions over the years.The author acknowledges the Royal Society, the JapanSociety of the Promotion of Science (JSPS), EuropeanResearch Council (ERC consolidator grant NANOPRS,project number 617266) for financial support and EPSRCgrant code EP/H022333/1 for the provision of a confo-cal microscope, and the Shipwrights Arms for access toa mermaid.

[1] A. Ivlev, H Lowen, G. E. Morfill, and C. P. Royall.Complex Plasmas and Colloidal Dispersions: Particle-resolved Studies of Classical Liquids and Solids. WorldScientific Publishing Co., Singapore Scientific, 2012.

[2] C. Likos. Effective interactions in soft condensed matterphysics. Phys. Rep., 348:267–439, 2001.

[3] P.N. Pusey and W. van Megen. Phase behaviour of con-centrated suspensions of nearly hard colloidal spheres.Nature, 320:340, 1986.

[4] C. P. Royall, W. C. K. Poon, and E. R. Weeks. In searchof colloidal hard spheres. Soft Matter, 9:17–27, 2013.

Page 8: Unknowns - arXivcolloids. Soon after, the seminal work of Asakura and Oo-sawa (AO) showed that, in a solution of non-adsorbing polymers, colloids experience a tuneable attraction due

8

[5] E.J.W. Verwey and J. Overbeek. Theory of the Stabilityof Lyophobic Colloids. Elsevier, Amsterdam, 1948.

[6] S. Asakura and F. Oosawa. On interaction between 2bodies immersed in a solution of macromolecules. J.Chem. Phys., 22(7):1255–1256, 1954.

[7] S. Asakura and F. Oosawa. Interaction between particlessuspended in solutions of macromolecules. J. Poly. Sci.,33:183–192, 1958.

[8] D. Fusco and P. Charbonneau. Soft matter perspectiveon protein crystal assembly. Coll. Surf. B.: Biointerfaces,137:22–31, 2016.

[9] A. Ciach. Universal sequence of ordered structuresobtained from mesoscopic description of self-assembly.Phys. Rev. E, 78:061505, 2008.

[10] F. Sciortino, S. Mossa, E. Zaccarelli, and P. Tartaglia.Equilibrium cluster phases and low-density arrested dis-ordered states: The role of short- range attraction andlong-range repulsion. Phys. Rev. Lett., 93:055701, 2004.

[11] F. Sciortino and P. Tartaglia. Glassy colloidal systems.Adv. in Phys., 54(6):471–524, 2005.

[12] G. Tarjus, S .A. Kivelson, Z. Nussinov, and P. Viot. Thefrustration-based approach of supercooled liquids and theglass transition: a review and critical assessment. J.Phys.: Condens. Matter, 17:R1143–R1182, 2005.

[13] M. Tarzia and A. Coniglio. Lamellar order, microphasestructures, and glassy phase in a field theoretic model forcharged colloids. Phys. Rev. E, 75:011410, 2007.

[14] A. J. Archer, D. Pini, R. Evans, and L. Reatto. Modelcolloidal fluid with competing interactions: Bulk and in-terfacial properties. J. Chem. Phys., 126:014104, 2007.

[15] A. Ciach and W. T. Gozdz. Mesoscopic descriptionof network-forming clusters of weakly charged colloids.Condens. Matter Phys., 13:23603, 2010.

[16] M. Edelmann and R. Roth. Gyroid phase of fluids withspherically symmetric competing interactions. Phys. Rev.E, 93:621146, 2016.

[17] Y. Zhuang, K. Zhang, and P. Charbonneau. Equilibriumphase behavior of a continuous-space microphase former.Phys. Rev. Lett., 116:098301, 2016.

[18] D. A. Lenz, R. Blaak, C. N. Likos, and B. M. Mladek. Mi-croscopically resolved simulations prove the existence ofsoft cluster crystals. Phys. Rev. Lett., 109:228301, 2012.

[19] B. M. Mladek, P. Charbonneau, and D. Frenkel. Phasecoexistence of cluster crystals: Beyond the gibbs phaserule. Phys. Rev. Lett., 99:235702, 2007.

[20] B. M. Mladek, D. Gottwald, G. Kahl, M. Neumann, andC. N. Likos. Formation of polymorphic cluster phases fora class of models of purely repulsive soft spheres. Phys.Rev. Lett., 96:045701, 2006.

[21] M. B. Sweatman, R. Fartaria, and L. Lue. Cluster forma-tion in fluids with competing short-range and long-rangeinteractions. J. Chem. Phys., 140:124508, 2014.

[22] F. Ghezzi and J. C. Earnshaw. Formation of meso-structures in colloidal monolayers. J. Phys.: Condens.Matter, 9:L517–L523, 1997.

[23] A. D. Law, Auriol M., D. Smith, T. S. Horozov, andD. M. A. Buzza. Self-assembly of two-dimensional col-loidal clusters by tuning the hydrophobicity, composi-tion, and packing geometry. Phys. Rev. Lett., 110:138301,2013.

[24] R. P. Sear, S.-W. Chung, G. Markovich, W. M. Gel-bart, and J. R. Heath. Spontaneous patterning of quan-tum dots at the air-water interface. Phys. Rev. E,59:R6255(R), 1999.

[25] A. T. Pham, Y. Zhuang, P Detwiler, J. E. S. Socolar,P. Charbonneau, and B. B. Yellen. Phase diagram andaggregation dynamics of a monolayer of paramagneticcolloids. Phys. Rev. E, 95:052607, 2017.

[26] A. Ciach and G. Stell. Effect of competition betweencoulomb and dispersion forces on phase transitions inionic systems. J Chem. Phys., 114:3617–3630, 2001.

[27] K. Ioannidou, L. Kanduc, M. Li, Frenkel D., J. Dobnikar,and E. Del Gado. The crucial effect of early-stage gelationon the mechanical properties of cement hydrates. NatureComm., 7:12106, 2016.

[28] P. Kowalczyk, A. Ciach, P. A. Gauden, and A. P. Terzyk.Equilibrium clusters in concentrated lysozyme protein so-lutions. J. Coll. Interf. Sci., 363:579–584, 2010.

[29] A. Stradner, H. Sedgwick, F. Cardinaux, W.-C.K. Poon,S.U. Egelhaaf, and P. Schurtenberger. Equilibrium clus-ter formation in concentrated protein solutions and col-loids. Nature, 432:492–495, 2004.

[30] F. Cardinaux, T. Gibaud, A Stradner, and P. Schurten-berger. Interplay between spinodal decomposition andglass formation in proteins exhibiting short-range attrac-tions. Phys. Rev. Lett., 99:118301, 2007.

[31] A. M. Kulkarni, N. M. Dixit, and C. F. Zukoski. Ergodicand non-ergodic phase transitions in globular protein sus-pensions. Faraday Disc., 123:37–50, 2002.

[32] We gratefully acknowlege the Shipwrights Arms, Helford,Cornwall, UK for access to a mermaid.

[33] E. H. A. de Hoog, W. K. Kegel, A. van Blaaderen, andH. N. W. Lekkerkerker. Direct observation of crystal-isation and aggregation in a phase-separating colloid-polymer suspension. Phys. Rev. E., 64:021407, 2001.

[34] N. Elsner, D. R. E. Snoswell, C. P. Royall, and B. V.Vincent. Simple models for 2d tunable colloidal crystalsin rotating ac electric fields. J. Chem. Phys., 130:154901,2009.

[35] S. Nguyen, V. D. Faber, Z. Hu, G. H. Wegdam, andP. Schall. Controlling colloidal phase transitions withcritical casimir forces. Nature Comm., 4:1584, 2013.

[36] A. Vrij, M. H. G. M. Penders, P. W. Rouw, C. G.de Kruif, J. K. G. Dhont, and H. N. W. Smits,C. Lekkerkerker. Phase-transition phenomena in colloidalsystems with attractive and repulsive particle interac-tions. Faraday Discussions, 90:31–40, 1990.

[37] M. Dijkstra, J. M. Brader, and R. Evans. Phase be-haviour and structure of model colloid-polymer mixtures.J. Phys.: Condens. Matter, 11:10079–10106, 1999.

[38] M. Dijkstra, R. van Roij, and R. Evans. Effective inter-actions, structure, and isothermal compressibility of col-loidal suspensions. J. Chem. Phys, 113:4799–4807, 2000.

[39] J. A. Long, D. W. J. Osmond, and B. Vincent. Theequilibrium aspects of weak flocculation. J. Coll. Interf.Sci., 42:545–553, 1973.

[40] A. Vrij. Polymers at interfaces and interactions in col-loidal dispersions. Pure Appl. Chem., 48(4):471–483,1976.

[41] K. Zahn, J. M. Mendez-Alcaraz, and G. Maret. Hydro-dynamic interactions may enhance the self-diffusion ofcolloidal particles. Phys. Rev. Lett., 79:175–179, 1997.

[42] A. M. Alsayed, M. F. Islam, J. Zhang, Collings P. J.,and A. Yodh. Premelting at defects within bulk colloidalcrystals. Science, 309:1207–1210, 2005.

[43] G. Bryant, S. R. Williams, L. Qian, I. K. Snook, E. Perez,and F. Pincet. How hard is a colloidal ‘hard-sphere’ in-teraction? Phys. Rev. E., 66:060501, 2002.

Page 9: Unknowns - arXivcolloids. Soon after, the seminal work of Asakura and Oo-sawa (AO) showed that, in a solution of non-adsorbing polymers, colloids experience a tuneable attraction due

9

[44] M.E. Leunissen. Manipulating Colloids with Charge andElectric Fields. PhD thesis, Utrecht Universiteit, 2006.

[45] C. P. Royall, M. E. Leunissen, A-P. Hyninnen, M. Di-jkstra, and A. van Blaaderen. Re-entrant melting andfreezing in a model system of charged colloids. J. Chem.Phys., 124:244706, 2006.

[46] I. Rios de Anda, A. Statt, F. Turci, and C. P. Roy-all. Low-density crystals in charged colloids: Comparisonwith yukawa theory. Contrib. Plasma Phys., 55:172–179,2015.

[47] C. P. Royall, M. E. Leunissen, and A. van Blaaderen. Anew colloidal model system to study long-range interac-tions quantitatively in real space. J. Phys.: Condens.Matter, 15(48):S3581–S3596, December 2003.

[48] A Yethiraj and As van Blaaderen. A colloidal modelsystem with an inteaction tunable from hard sphere tosoft and dipolar. Nature, 421:513–517, 2003.

[49] T. Ohtsuka, C. P. Royall, and H. Tanaka. Local structureand dynamics in colloidal fluids and gels. Europhys. Lett.,84:46002, 2008.

[50] C. L. Klix, K. Murata, H. Tanaka, S. Williams, A. Ma-lins, and C. P. Royall. Novel kinetic trapping in chargedcolloidal clusters due to self-induced surface charge orga-nization. Sci. Rep., 3:2072, 2013.

[51] C. Bechinger, D. Rudhardt, P. Leiderer, R. Roth, andS. Dietrich. Understanding depletion forces beyond en-tropy. Phys. Rev. Lett., 83:3960–3963, 1999.

[52] M. Piech and J.Y. Walz. Direct measurement of depletionand structural forces in polydisperse, charged systems. J.Coll. Interf. Sci., 253:117–129, 2002.

[53] J. C. Crocker and D. G. Grier. Microscopic measure-ment of the pair interaction potential of charge-stablizedcolloids. Phys. Rev. Lett., 81:352–355, 1994.

[54] R. Verma, J.C. Crocker, T.C. Lubensky, and A.G. Yodh.Entropic colloidal interactions in concentrated dna solu-tions. Phys. Rev. Lett., 81:4004–4007, 1998.

[55] J.C. Crocker, J.A. Matteo, A.D. Dinsmore, and A.G.Yodh. Entropic attraction and repulsion in binary col-loids probed with a line optical tweezer. Phys. Rev. Lett.,83:4352, 1999.

[56] R. Roth, R. Evans, and S. Dietrich. Depletion potentialin hard-sphere mixtures: Theory and applications. Phys.Rev. E., 62:5360–5377, 2000.

[57] C. P. Royall, A. A. Louis, and H. Tanaka. Measuringcolloidal interactions with confocal microscopy. J. Chem.Phys., 127(4):044507, 2007.

[58] P. Charbonneau. Personal communciation.[59] W. C. K. Poon. The physics of a model colloid-polymer

mixture. J. Phys.: Condens. Matter, 14(33):R859–R880,August 2002.

[60] L. Ramos and L. Cipelletti. Slow dynamics in glassy softmatter. J. Phys.: Condens. Matter, 17(6):R253–R285,2005.

[61] E. Zaccarelli. Colloidal gels: Equilibrium and non-equilibrium routes. J. Phys.: Condens. Matter,19:323101, 2007.

[62] M. Brunner, C. Bechinger, W. Strepp, V. Lobaskin, andH. H. von Gruenberg. Density-dependent pair interac-tions in 2d. Europhys. Lett., 58:926–965, 2002.

[63] L. F. Rojas, C. Urban, P. Schurtenberger, T. Gisler, andH. H. von Gruenberg. Reappearance of structure in col-loidal suspensions. Europhys. Lett., 60:802–808, 2002.

[64] D. Reinke, H. Stark, H.-H. von Gruenberg, A. B.Schofield, G. Maret, and U. Gasser. Noncentral forces in

crystals of charged colloids. Phys. Rev. Lett., 98:038301,2007.

[65] M. F. Hsu, E. R. Dufresne, and D. A. Weitz. Chargestabilization in nonpolar solvents. Langmuir, 21:4881–4887, 2005.

[66] J.-P. Hansen and I.R. Macdonald. Theory of Simple Liq-uids. Academic, London, 4th edition, 2013.

[67] J. A. Barker and D. Henderson. What is ”liquid”? under-standing the states of matter. Rev. Mod. Phys., 48:587–671, 1976.

[68] John D. Weeks, David Chandler, and Hans C. Ander-sen. Role of Repulsive Forces in Determining the Equi-librium Structure of Simple Liquids. J. Chem. Phys.,54(12):5237, January 1971.

[69] V. Lobaskin and P. Linse. Simulation of an asym-metric electrolyte with charge asymmetry 60:1 usinghard-sphere and soft-sphere models. J. Chem. Phys.,111:4300–4309, 1999.

[70] A.-P. Hynninen. Phase Behavior of Charged Colloids andthe Effect of External Fields. PhD thesis, Utrecht Uni-versity, 2005.

[71] A.-P. Hynninen and M. Dijkstra. Melting line of chargedcolloids from primitive model simulations. J. Chem.Phys., 123:244902, 2005.

[72] E. K. Hobbie. Metastability and depletion-driven aggre-gation. Phys. Rev. Lett., 89:3996–4001, 1998.

[73] E. K. Hobbie. Depletion-driven phase separation and re-versible aggregation in confined colloidal mixtures. Lang-muir, 15:8807–8812, 1999.

[74] J. Taffs, A Malins, S. R. Williams, and C. P. Royall. Astructural comparison of models of colloid-polymer mix-tures. J. Phys.: Condens. Matter, 22:104119, 2010.

[75] G. Foffi, C. De Michele, F. Sciortino, and P. Tartaglia.Scaling of dynamics with the range of interactionin short-range attractive colloids. Phys. Rev. Lett.,94:078301, Feb 2005.

[76] D. Richard, J. E. Hallet, T. Speck, and Royall C. P. Cou-pling between criticality and gelation in ”sticky” spheres:A structural analysis. ArXiV, page 1802.09481, 2018.

[77] C. P. Royall, S. R. Williams, and H. Tanaka. Vitrifi-cation and gelation in sticky spheres. J. Chem. Phys.,148:044501, 2018.

[78] H. Sedgwick, S. U. Egelhaaf, and W. C. K. Poon. Clustersand gels in systems of sticky particles. J. Phys.: Condens.Matter, 16:S4913–S4922, 2004.

[79] J. Groenewold and W. K. Kegel. Colloidal cluster phases,gelation and nuclear matter. J. Phys.: Condens. Matter,16:S4877–S4886, 2004.

[80] J. Groenewold and W.K. Kegel. Anomalously largeequilibrium clusters of colloids. J. Phys. Chem. B.,105:11702–11709, 2001.

[81] A. I. Campbell, V. J. Anderson, J. S. van Duijneveldt,and P. Bartlett. Dynamical arrest in attractive colloids:The effect of long-range repulsion. Phys. Rev. Lett.,94(20):208301, May 2005.

[82] F. Sciortino, P. Tartaglia, and E. Zaccarelli. One-dimensional cluster growth and branching gels in col-loidal systems with short-range depletion attracion andscreened electrostatic repulsion. J. Phys. Chem. B.,109:21942–21953, 2005.

[83] C. J. Dibble, M. Kogan, and M. J. Solomon. Structureand dynamics of colloidal depletion gels: Coincidence oftransitions and heterogeneity. Phys. Rev. E., 74:041403,2006.

Page 10: Unknowns - arXivcolloids. Soon after, the seminal work of Asakura and Oo-sawa (AO) showed that, in a solution of non-adsorbing polymers, colloids experience a tuneable attraction due

10

[84] C.P. Royall, D.G.A.L. Aarts, and H. Tanaka. Fluid struc-ture in colloid-polymer mixtures: the competition be-tween electrostatics and depletion. J. Phys.: Condens.Matter, 17:S3401–S3408, 2005.

[85] M. Kohl, R.F. Capellmann, M. Laurati, S.U. Egelhaaf,and M. Schmiedeberg. Directed percolation identifiedas equilibrium pre-transition towards non-equilibrium ar-rested gel states. Nature Comm., 7:11817, 2016.

[86] C. L. Klix, C. P. Royall, and H. Tanaka. Structural anddynamical features of multiple metastable glassy statesin a colloidal system with competing interactions. Phys.Rev. Lett., 104:165702, 2010.

[87] A. Malins, S. R. Williams, J. Eggers, H. Tanaka, andC. P. Royall. Geometric frustration in small colloidalclusters. J. Phys.: Condens. Matter, 21:425103, 2009.

[88] A. Malins, S. R. Williams, J. Eggers, H. Tanaka, andC. P. Royall. The effect of inter-cluster interactions onthe structure of colloidal clusters. J. Non-Cryst. Solids,357:760–766, 2010.

[89] J. P. K. Doye, D. J. Wales, and R. S. Berry. The effect of

the range of the potential on the structures of clusters.J. Chem. Phys., 103(10):4234–4249, September 1995.

[90] D. J. Wales. Energy Landscapes: Applications to Clus-ters, Biomolecules and Glasses. Cambridge UniversityPress, Cambridge, 2004.

[91] G. Meng, N. Arkus, M. P. Brenner, and V. N. Manoha-ran. The free-energy landscape of clusters of attractivehard spheres. Science, 327:560–563, 2010.

[92] J. Dong, M. Meissner, J. Eggers, A. M. Seddon, and C. P.Royall. Opposed flow focusing: evidence of a second orderjetting transition. submitted, 2018.

[93] T. E. Kodger, R. E. Guerra, and J. Sprakel. Precise col-loids with tunable interactions for confocal microscopy.Sci. Rep., 5:14635, 2015.

[94] C. Zhang, N. Gnan, T. G. Mason, E. Zaccarelli, andF. Scheffold. Dynamical and structural signatures of theglass transition in emulsions. J. Stat. Mech.: Theory andExperiment, page 094003, 2016.