Tunable Rheology of Dense Soft Deformable Colloids

  • Upload
    jro84

  • View
    217

  • Download
    0

Embed Size (px)

Citation preview

  • 7/25/2019 Tunable Rheology of Dense Soft Deformable Colloids

    1/14

    Tunable rheology of dense soft deformable colloids

    Dimitris Vlassopoulos a,b, Michel Cloitre c

    a FORTH, Institute of Electronic Structure & Laser, Heraklion 70013, Crete, Greeceb University of Crete, Department of Materials Science & Technology, Heraklion 71003, Crete, Greecec Matire Molle et Chimie (UMR 7167, ESPCI-CNRS), ESPCI ParisTech, 10 Rue Vauquelin, 75005 Paris, France

    a b s t r a c ta r t i c l e i n f o

    Article history:

    Received 24 August 2014

    Received in revised form 26 September 2014Accepted 29 September 2014

    Available online 7 October 2014

    Keywords:

    Soft colloids

    Viscoelasticity

    Microgels

    Star polymers

    Glass/Jamming

    Depletion

    Particle elasticity

    Viscosity

    Yield stress/strain

    Flow curves

    In the last two decades, advances in synthetic, experimental and modeling/simulation methodologies have

    considerably enhanced our understanding of colloidal suspension rheology and put theeld at the forefront of

    soft matter research. Recent accomplishments include the ability to tailor the ow of colloidal materials via

    controlled changes of particle microstructure and interactions. Whereas hard sphere suspensions have been

    the most widely studied colloidal system, there is no richer type of particles than soft colloids in this respect.

    Yet,despitethe remarkable progress in theeld,many outstandingchallenges remain in ourquest to linkparticle

    microstructure to macroscopic properties and eventually design appropriate soft composites. Addressing

    them will provide the route towards novel responsive systems with hierarchical structures and multiple

    functionalities. Here we discuss the key structural and rheological parameters which determine the tunable

    rheology of dense soft deformable colloids. We restrict our discussion to non-crystallizing suspensions of

    spherical particles without electrostatic or enthalpic interactions.

    2014 Elsevier Ltd. All rights reserved.

    1. Introduction

    1.1. Classication of soft colloids

    Looking at the rich landscape of soft matter systems it is a prime

    challenge to identify and classify the different types of nanoparticles

    which have been investigated in reasonable depth and at the same

    time are representatives of basic features in terms of phase behavior

    and macroscopic properties. A reasonably comprehensive, albeit non-

    exhaustive list of soft colloids includes emulsions and nanoemulsions,

    surfactant onionsand liposomes, microgels,grafted coreshell particles,

    block copolymer micelles, dendrimers, and star polymers[18]. Soft-

    ness can be appreciated from different perspectives: particle elasticity,

    diversity of soft interactions, and particle volume fraction.

    1.2. Particle elasticity

    Theimportance of theelasticityof Brownian objectcan be quantied

    by the non-dimensional parameter = F/kTwhich represents the

    ratio of theelastic free energy, F, to thethermalenergy kT[911]. Poly-

    mer coils have a free energy of entropic origin withF kTindicating

    that 1; they are the most deformable objects available. On the

    other hand, spherical elastic particles with modulus Eand radiusRare

    characterized byF=ER3 and=ER3/kT. Thus, innitely rigid hard

    spheres correspond to the limit . Besides hard spheres, colloidal

    star polymers and microgels are archetypical representatives of two

    important classes of soft particles. Star polymers are made of a large

    number of arms, f, each with the degree of polymerization Na,

    which are attached to a central core. Following the DaoudCotton

    model [12], each arm can be viewed as a succession of blobs whose vol-

    ume V(r) r3f3/2 increaseswith distance rfromthe core. Due to theto-

    pological stretching of the arms, the elastic modulus at distance rscales

    as:E kT/V(r). At the periphery of the star, we haveEkTR-3f3/2 so that

    f3/2. For stars with a few hundreds of arms, is of the order of a few

    thousands. Microgel particles are crosslinked polymeric networks swol-

    len by solvent[4,13,14]. Their modulus depends on many parameters

    such as the crosslink density, the solvent quality, the presence of ions,

    and the network architecture. For neutral microgels, a reasonable esti-

    mate isE kT/V,whereV NxV0(Nx: number of statistical units be-

    tween crosslinks; V0: volume of a statistical unit). For submicron

    microgels (typically, R = 100 nm; V0 = 1 n m3; Nx = 100), we estimate

    = 104. Nanoemulsion and emulsion droplets are soft colloidal objects

    used in many applications. Here the elastic energyis associated with the

    surface energy of the interface (typically oil/water) which resists defor-

    mation [15]:F=R2 (with the interfacial tension). For R= 100 nm

    and= 103 N/m, we estimate 106. For completeness, let us briey

    mention multilamellar vesicles which exhibit strong analogies with

    Current Opinion in Colloid & Interface Science 19 (2014) 561574

    E-mail addresses: [email protected](D. Vlassopoulos), [email protected]

    (M. Cloitre).

    http://dx.doi.org/10.1016/j.cocis.2014.09.007

    1359-0294/ 2014 Elsevier Ltd. All rights reserved.

    Contents lists available atScienceDirect

    Current Opinion in Colloid & Interface Science

    j o u r n a l h o m e p a g e : w w w . e l s e v i e r . c o m / l o c a t e / c o c i s

    http://dx.doi.org/10.1016/j.cocis.2014.09.007http://dx.doi.org/10.1016/j.cocis.2014.09.007http://dx.doi.org/10.1016/j.cocis.2014.09.007mailto:[email protected]:[email protected]://dx.doi.org/10.1016/j.cocis.2014.09.007http://www.sciencedirect.com/science/journal/13590294http://www.elsevier.com/locate/cocishttp://www.elsevier.com/locate/cocishttp://www.sciencedirect.com/science/journal/13590294http://dx.doi.org/10.1016/j.cocis.2014.09.007mailto:[email protected]:[email protected]://dx.doi.org/10.1016/j.cocis.2014.09.007http://crossmark.crossref.org/dialog/?doi=10.1016/j.cocis.2014.09.007&domain=pdf
  • 7/25/2019 Tunable Rheology of Dense Soft Deformable Colloids

    2/14

    emulsion droplets, the effective surface elasticity resulting from a com-

    bination of the bending and compressive elasticity of the smectic layers

    constituting the particles [4,16]. In the context of this discussion, recent

    efforts to characterize the elasticity of individual particles using various

    techniques like micropipette aspiration, AFM, and microuidic ow ap-

    pear very promising[1723].

    Fig. 1summarizes the classication of soft colloids based on the

    above approach. Clearly, colloidal particles span the softness parameter

    space from ultrasoft polymer coils and star polymers to quasi-hardspheres, indicating that softness can be tuned at will during synthesis

    or preparation to meet the requirements of various applications. Natu-

    rally, our classication is somewhat simplied and does not account

    for the variationsin parameters like star functionality, crosslinkdensity,

    and interfacial tension (always in the absence of charges or enthalpicef-

    fects). For example, star polymers are the simplest representatives of a

    wider class of particles with heterogeneous internal microstructure. It

    includes dendrimers [8,24,25], block copolymer micelles resulting

    from the dissolution of diblock or multiblock copolymers in a selective

    solvent[2633], and polymer-grafted (or polymer-adsorbed) particles

    [11,3436]. The latter nd many applications as they provide an exqui-

    site way to stabilize colloidal particles and tailor their rheology. Further

    complications to our classication arise from more complex structures,

    such as for example coreshell microgels consisting of a solid core cov-

    ered with a network-like shell [3740]and block copolymer micelles

    with a multi-compartmented coreshell microstructure [29,32]. Despite

    these issues, the classication ofFig. 1offers a comprehensive descrip-

    tion of the important features governing the behavior of soft colloids.

    A useful, albeit rough parameter for coreshell particle's softness is

    the fraction of the shell layer [34]: s = L/(L + rc), where L is its thickness

    andrcis the core radius; for hard spheres, stars (forf= 100 arms andNa= 100 statistical units per arm) and polymers, s is about 0, 0.95

    and 1, respectively. In all these systems it is possible to de-swell the

    outer soft layer by adjusting the physicochemical conditions (pH, tem-

    perature, solvent quality, addition of small non-adsorbing polymers)

    so that the system switches from soft-like to hard-like behavior.

    1.3. Particle interactions

    The softness of particles is linked to the softness of interactions

    which can be characterized by the form of the repulsive pair potential

    between two particles. For instance, star polymers or, equivalently, par-

    ticles covered with long end-grafted polymer chains (such as block

    copolymer micelles) are characterized by coarse-grained ultrasoft po-

    tentials which exhibit long-ranged Yukawa-like repulsion at large

    center-to-center distances and logarithmic repulsion at short distances

    (Fig. 1)[41]. The amplitude of the potential and its range of repulsion

    depend on the number of arms. The logarithmic repulsion accounts

    for the microscopic interactions which develop when the arms come

    into contact, interpenetrate, and deform at short distances. The number

    of arms,f, is the unique control parameter for tuning the starstar inter-

    actionfrom hard-sphere-like to polymer-like (Fig. 1), the potential scal-

    ing withf3/2 as the normalized elastic free energy[41].This potential

    has been very successful in describing the structure of stars and star-like micelles, including the crystallization and glass transition, as con-

    rmed experimentally[5,27,30,4144]. On the other hand, particles

    without dangling chains like microgels or emulsion droplets can be

    thought of as effective hard spheres up to the point of contact, and as

    elastically deformable spheres at shorter distances. When the particles

    come into contact they develop facets through which they exert normal

    repulsive forces which are well modeled using Hertz theory [4547].

    The resulting potential is accurate up to particle deformation of about

    15% but extensionshave been proposed to account for the large particle

    overlaps that take place under ow[48]. These potentials successfully

    apply to dense microgel suspensions or concentrated emulsions

    [4550]. Hence, the key features of soft interactions that distinguish

    them from their hard-sphere counterparts are their wide-ranging and

    the less-steep potential at contact and beyond, which explain the

    tunability of soft interactions.

    1.4. Particle volume fraction

    In relation to particle elasticity, a consequence of softness is the abil-

    ity of the particles to deform/compress in contrast to hard spheres.

    Emulsion droplets adapt their shape to steric constraints by forming

    facets at contact at constant volume. Hairy particles (like star polymers,

    block copolymer micelles or grafted particles) and microgels can adjust

    both their shape and volume in response to external stimuli like pres-

    sure, ow, pH, and temperature, which makes their properties tunable

    [4]. Moreover, the hairy structure of star-like particles allows for inter-

    digitation in addition to deformability. The above features profoundly

    affect the denition of the volume fraction of soft particles and their

    behavior at high concentrations.The volume fraction of a hard sphere colloidal suspension is usually

    determined in reference to the onset of crystallization at c= 0.494 or

    the maximum packing fraction atHCP = 0.74. Hard particles subjected

    to a centrifugal forceeld will eventually reach the latter limit so that

    desired volume fractions can be subsequently prepared by dilution

    [10,51,52].This procedure does not apply to soft deformable particles

    Fig. 1. Cartoons of different nanoparticles with the arrow pointing to the direction of enhanced elasticity and transition from softto hard repulsive pair interaction potential: polymeric

    coil, star polymer,microgel,emulsion, andhard sphere.Dashedand dottedlines illustrate thevariability of therepulsive potentials uponchanges in molecular characteristics, which affect

    particle elasticity. Particles with grafted or adsorbed polymers (not shown here) are typically positioned between stars and emulsions (see text).

    562 D. Vlassopoulos, M. Cloitre / Current Opinion in Colloid & Interface Science 19 (2014) 561574

  • 7/25/2019 Tunable Rheology of Dense Soft Deformable Colloids

    3/14

    which de-swell osmotically at large volume fraction [5356]. The

    volume fraction of concentrated soft particle suspensions thus remains

    an ambiguous quantity[52]. Generally, the effective volume fraction

    based on the single particle hydrodynamic size is agreed upon as a uni-

    versal parameter, albeit different from the actual volume fraction[51,

    52,57,58]. The determination of the latter is extremely challenging

    and, apart from a few studies for microgels and star polymers, few re-

    sults are available[53,56,59]. The effective volume fraction has been

    shown to work well in terms of scaling the properties (such as zero-shear viscosity or self-diffusion) for different soft particles over a

    range of sizes and concentrations[57,58,60,61]. Hereafter we shall call

    itvolume fractionunless otherwise specied.

    2. State diagram: crystallization, vitrication and jamming

    Particle interactions are reected in the form of the potential and

    eventually in the phase behavior. Since some phases are metastable,

    we prefer to call the respective morphological diagrams of athermal

    (i.e., possessing only entropic interactions) soft colloids as statedia-

    grams. Before studying the state diagrams of soft colloids, it is useful

    to revisit the reference case of hard spheres.Fig. 2a shows the behavior

    of monodisperse and non-interacting colloidal spheres suspended in

    Newtonianuids[62,63].With increasing volume fraction there is a

    transition from the disordereduid phase to a crystal phase (at volume

    fraction c) through an intermediateuidcrystal coexistence regime

    where crystals are in equilibrium with the uid phase. Above volume

    fractiong, the crystalline order is lost and a disordered amorphous

    glass, where the positions of the particles are uncorrelated, appears in-

    stead. The upper limit of the glassy region is the maximum fraction of

    randomly packed spheres,J, where theparticlesare forcedinto contact

    ina jammedstate [64]. AboveJ, samples must have domains of crystal-

    line structure or be entirely crystallized. Polydispersity inuences the

    state diagram through the values of the glass and jamming transition,

    which are largerfor polydisperse samples, and more notably suppresses

    crystallization[65].

    The state diagram of non-interacting soft microgels is shown in

    Fig. 2b for comparison. It exhibits apparent similarities with that of

    hard-sphere suspensions: microgel suspensions undergo the same

    sequence of transitions fromuid to crystal to amorphous solid with

    increasing volume fraction. However several major quantitative differ-

    ences signal the role of softness. First, particles can be compressed

    well above the jamming transition where they come into contact,

    because they can accommodate the increased osmotic pressure by

    adjusting their volume and shape as discussed in Section 1.4[4]. Sec-ondly, crystallization and melting seem to occur at higher volume frac-

    tions and exhibit a narrower phase coexistence region compared to

    hard sphere suspensions[66]. Sometimes, it may happen that crystalli-

    zation is suppressed, whereas the glass and jamming transition are

    shifted to higher volume fractions[58].Finally, the presence of soft in-

    teractions combined with the development of appropriate annealing

    protocols (rapid heating and subsequent cooling after certain rest

    time) [67] and variations in synthesis (for instance changing the degree

    of crosslinking)[13]offers unique possibilities to manipulate the mor-

    phology of soft microgel suspensions to get ordered crystalline phases

    over a wide range of volume fractions. Note that the state diagram in

    Fig. 2b is based on the behavior of microgel suspensions which are not

    purely athermal: poly(methylmethacrylate) (PMMA) spheres swollen

    in benzyl alcohol [68] and thermosensitive poly(N-isopropylacrylamide)

    (PNIPAm) particles in water[39,50,66]. The state diagrams of ionic

    microgels with varying softness qualitatively also conform to that

    inFig. 2b[69], which supports the generality of the trends we have

    identied so far.

    The softer star-like systems are characterized by a large uid region,

    practically up to the overlap concentration (c*), a very small crystal re-

    gion which may be completely masked by the glass, and a relatively

    large glass-jamming region (Fig. 2c). The suppression of crystallization

    in hairy suspensions is related to the fact that the dangling chains inter-

    penetrate above c* and uctuate beyond the limit allowed by the

    Lindeman criterion[5,35,70,71]. At sufciently long times, star glasses

    eventually crystallize (in fact, this has been reported for hard spheres

    as well)[43,72,73]. It appears that a large core contributes to damping

    Fig. 2. Schematic one-dimensional state diagrams of nearly-monodisperse athermal soft colloidal suspensions, as function of their effective volume fraction (see Section 2). The numbers

    arevaluesfor volume fractions of hardspheres;values above 0.74are relevantonly for softsystems. Frombottomto top thediagrams of a):hard spheres,b): microgels, c): star polymers

    andd): polymericcoils areshown.Letters refer todifferentregions:uid(F, theblue-shaded regions), crystal(C, atc = 0.545 forhardspheres),coexistence (F + C, above volume fraction

    0.494 forhard spheres), glass (G,at g = 0.58 forhard spheres), andjammed (J). RCPand HCPare therandomclosepacking andhexagonal close packing volume fractions, which forhard

    spheres are determined to be 0.64 and 0.74, respectively. The solid black vertical lines represent established transitions (even if occurring at different volume fractions for different

    systems) whereas red dashed lines represent transitions whose universality is still debatable. The values of the volume fraction in the horizontal scale are indicative and do not respect

    the actual scale.

    563D. Vlassopoulos, M. Cloitre / Current Opinion in Colloid & Interface Science 19 (2014) 561574

    http://localhost/var/www/apps/conversion/tmp/scratch_4/image%20of%20Fig.%E0%B2%80
  • 7/25/2019 Tunable Rheology of Dense Soft Deformable Colloids

    4/14

    theuctuations, hence promoting crystallization. This explain in part

    why in general block copolymer micelles and densely grafted hard

    nanoparticles can crystallize easily whereas star polymers cannot[5,

    30,74,75]. When the dangling chains are short, the dynamical nature

    of the micelles allows for appropriate adjustment of the number of

    arms to promote crystallization[76,77].

    From the above discussion we can draw some general conclusions

    about the impact of softness on the state diagram of colloidal suspen-

    sions. A general and important trend is that the

    uid phase tends to ex-pand when the softness increases. This observation can be connected to

    the behavior of polymer coil solutions which remainuid throughout

    the whole range of volume fractions (Fig. 2d). For monodisperse sys-

    tems, the same sequence of phases and states leading from uid phase

    to glassy states through crystalline phases is qualitatively observed.

    However, the actual values of the effective volume fraction marking

    the transitions depend on softness. External stimuli, such as shear[78]

    or thermal annealing[43,72,73], can promote local rearrangement and

    even lead to crystallization.

    Polydispersity systematically inhibits crystallization so that in poly-

    disperse soft colloids we are left with the glassy and jammed regimes

    only, which are often, but not always, the states of principal interest

    for rheology. The glass transition can be described to a rst approach

    by the cage model originally proposed for hard spheres where colloidal

    particles are topologically constrained by their neighbors in virtual

    cages[9,62,79,80]. In the jammed state, soft particles are still arranged

    in a disordered amorphous way but they form polyhedral facets at con-

    tact[4]. Interpenetration of star-like particles introduces an additional

    complexity, which has not been investigated in detail yet [81,82].

    Today there is growing evidence suggesting that solidication of poly-

    disperse soft particle suspensions does not obey a universal description,

    anddifferent scenarioshave been proposed as particle softness is varied.

    Emulsions, which are relatively rigid particles ( 106), exhibit a glass

    transition at g = 0.58 and a jamming transition around J= 0.64

    [8385]. Star polymer solutions ( 103) exhibit a liquid to solid tran-

    sition at a relatively well-dened concentration, but the glass or

    jammed nature of the solid phase is unclear[86]. These experimental

    observations qualitatively conform to the predictions of a recent

    model which combines the contributions of Brownian motion and har-monic interactions[87,88]. Observations for microgels are somewhat

    contradictory. The approach of the glass transition from the liquid

    phase seems to becomemore gradual for softer microgels [89]. Aqueous

    suspensions of submicron thermosensitive PNIPAm microgel or core

    shell particles have been shown to exhibit a well-dened glass regime.

    They represent an exquisite model system for studying the rheology

    of entropic glasses in relation with Mode Coupling Theory (MCT) [90].

    However, it is interesting to note that the rheology of very similar

    PNIPAm suspensions near the liquid-solid transition has been

    interpreted in terms of jamming[91]of a mixed glass-jamming state

    [92]. Understanding the details of particle microstructure is important

    in order to further advance the eld in this direction.

    3. Mixtures and osmotic interactions

    3.1. Colloidpolymer mixtures

    Athermal mixtures of colloidal hard spheres with non-adsorbing

    linear polymers constitute an archetypical situation where osmotic

    forces induce new morphologies and original rheological behavior

    (Fig. 3a)[10,93104].For example, in the colloidal gas domain (i.e.

    uidphase, F),the addition of linearpolymers to hard spheres is respon-

    sible for attractive depletion interactions which induce the aggregation

    of particles[93104]. At low volume fraction phase separation (PS) is

    observed whereas at larger volume fractions colloidal aggregates ulti-

    mately ll all the available space leading to gel formation (Gel). The in-

    terplay between strength and range of particle interactions (both

    repulsive and attractive), and gravity remains a subject of debate

    [105108].There are reports pointing to arrested phase separation as

    the ultimate equilibrium state in the system (with short-range attrac-

    tions) and others suggesting that percolation prevails[106112]. The

    most intriguing phenomena occur in the high-volume fraction glass re-

    gime. When linear polymers are added at increasing concentrations to

    hard spheres at a size ratio of about 1/10, the initial repulsive glass

    (RG) melts due to depletion interactions and eventually a re-entrant

    attractive glass (AG) forms[99,100,103,113].

    A natural question concerns the role of softness. Long-range soft po-tentials are expected to impart not only quantitative but also qualitative

    differences in the behavior of soft colloidlinear polymer mixtures[71,

    114]. Let us rst discuss the case of three-component mixtures made

    of solvent (background), soft colloids and linear chains. Earlier work

    with highly crosslinked microgel systems, which behave very much

    like hard spheres, conrmed the repulsive glass to liquid to re-entrant

    attractive glass or depletion occulation scenario[115118].Slightly

    crosslinked microgels, which respond to external elds by both volume

    and shape adjustments (Section 1.4), were found to de-swell in the

    presence of small linear polymers[118121]. More systematic studies

    have dealt with hairy particles and in particular star polymers

    (Fig. 3b)[55,122129]. Due to the osmotic pressure of the linear poly-

    mers, star polymers are expected to shrink (de-swell), causing a change

    of volume fraction, and eventually experience depletion interactions.

    This has been studied in reasonable detail both theoretically and exper-

    imentally[55,125,128,129]. Whereas the main physics associated with

    osmotic pressure effects is the same as for polymer/hard sphere mix-

    tures, the state diagrams are different (Fig. 3b). More importantly, the

    re-entrant solid state is not an attractive glass (or at least not necessar-

    ily) but a gel instead (Gel)[128,129]. For given attraction strength,

    phase separation is observed[55]. Due to the range of the potential,

    the liquid pocket (F) which exists at intermediate volume fractions of

    linear chains and stars is larger than in the hard sphere case. Moreover,

    polydispersity of the hard spheres may reduce or even eliminatethe er-

    godic phase, but not the glass-to-glass transition[113]. These conclu-

    sions are based on systematic rheological measurements supported by

    Small Angle Neutron and Dynamic Light Scattering (SANS, DLS), simula-

    tions and theoretical models in terms of effective interactions and Mode

    Coupling Theory[55,122129]. Nonlinear rheology and in particularLarge Amplitude Oscillatory Shear (LAOS) is an exquisite tool to detect

    the softening and eventual melting of the repulsive glass (RG) upon ad-

    dition of linear chains through the reduction of the yield stress, yield

    strain and cage modulus[128130]. It also supports the conjecture

    that yielding in these metastable colloidal systems is a gradual process

    involving multiple cooperative breakage events[131135]. For large

    enough fractions of linear polymers, the ergodic pocket (F) disappears

    and the colloidal mixture undergoes a transition from a repulsive glass

    (RG) to a gel (Gel). This is unambiguously probed by linear and nonlin-

    ear rheology since the LissajousBowditch stressstrain signatures of

    gels and glasses measured in LAOS experiments are distinct, the former

    being a nearly square curve characteristic of viscoplastic behavior and

    the latter a tilted ellipse as is the case for viscoelastic materials

    [132135].The star polymer paradigm has motivated or put into perspective

    studies involving hairy particles, and in particular the structure and

    dynamics of mixtures of polymer grafted nanoparticles and linear

    polymers in good solvents [136]or block copolymers and linear

    homopolymers in selective solvents[26,31,137]. In the latter case

    in particular, the homopolymers not only melt the crystalline phases

    formed by the block copolymer micelles but can also inuence their

    aggregation number, hence providing a control parameter for self-

    assembly[31,137].

    Recently, the role of linear polymers in inducing liquid-like behavior

    in concentrated microgel suspensions via depletion was exploited [33,

    138,139]. Adding polymers to microgel glasses induced melting and

    eventual re-entrance attractive glass formation, with the extension of

    the ergodic region depending on the microstructure of the particles

    564 D. Vlassopoulos, M. Cloitre / Current Opinion in Colloid & Interface Science 19 (2014) 561574

  • 7/25/2019 Tunable Rheology of Dense Soft Deformable Colloids

    5/14

    (e.g., density of crosslinks), the size ratio and the potentially changing

    solvent quality for the polymer. Furthermore, the role of depletion in

    the aggregation and phase separation of food protein systems, whichare also dominated by softness and exhibit qualitatively similar features

    to those reported here, was recently reviewed[140].

    Interesting phenomena also occur in the melt, i.e. in the absence of

    solvent background, where grafted nanoparticles or star polymers can

    be thought as solvent-free colloids[141144], with signicance for

    nanocomposite systems. Although we do not review the eld of nano-

    composites here, it is worth mentioning that recent work suggested

    that adding linear homopolymers to polymer-grafted nanoparticles

    originally in theuid state can result in the formation of anisotropic do-

    mains comprising trapped particles[145147]. This brings analogies to

    the asymmetric caging of binary star mixtures [148]which will be

    discussed below. A further worthy example is the systematic investiga-

    tions of grafted particles in polymer melts (say polystyrene-grafted

    polyorganosiloxane in linear polystyrene matrices) whereby increasing

    the size ratio between the radius of gyration of the polymer and the

    particle radius it is possible to reduce the strength of depletion and pro-

    mote dispersion of the particles, with important consequences on theirdynamic response[144,149153]. More recently, blends of crosslinked

    PMMA microgels with linear PMMA of much smaller size were investi-

    gated[154]. In the case of slightly crosslinked microgels the linear

    chains penetrate and swellthe microgels and homogeneous dispersions

    are obtained. However, highly crosslinked microgelsbehave likeimpen-

    etrable hard spheres and aggregate due to the depletion effects induced

    by linear chains.

    3.2. Binary colloidal mixtures

    The consequences of strong entropic effects discussed inSection 3.1

    can be extended to more complex, yet interesting from the standpoint

    of applications, systems, by replacing the linear polymer by another col-

    loidal particle. For instance the morphology of binary asymmetric

    Fig. 3.Schematic state diagrams of (from top to bottom): a) linear polymerhard sphere mixture atRpolymer/RHS b 0.2; b) linear polymercolloidal star mixture at Rpolymer/Rstar b 0.5;

    c) binary hard sphere mixture at Rsmall/Rlarge b 0.2; d) binary star mixture atRsmall/Rlarge= 0.3; and e) hard spherestar mixture atRHS/Rstar= 0.25. Notation: F = uid, PS = phase

    separation, RG = repulsive glass, AG = attractive glass, SG = single glass, DG = double glass, C = crystal, F + C = uidcrystal coexistence, APS = arrested phase separation,

    AsG = asymmetric glass.

    565D. Vlassopoulos, M. Cloitre / Current Opinion in Colloid & Interface Science 19 (2014) 561574

    http://localhost/var/www/apps/conversion/tmp/scratch_4/image%20of%20Fig.%E0%B3%80
  • 7/25/2019 Tunable Rheology of Dense Soft Deformable Colloids

    6/14

    mixtures of hard spheres was explored with emphasis given to the ex-

    istence of freezing transitions and the formation of superlattice struc-

    tures, driven by the maximization of entropy as the system adopts its

    most efcient packing arrangement.Fig. 3c shows that, for a give size

    ratio and depending on the relative volume fraction, hard sphere mix-

    tures exhibit both crystalline (C) and glassy regions (SG, DG). Single

    glasses (SG) and double glass (DG) form when either the small (here)

    particles or both small and large particles become kinetically trapped

    in amorphous structures[155,156].Mixtures of small star polymers oflow-functionality (up to 32) and large hard colloids were found to

    phase separate for size ratios not exceeding 0.5 and colloid volume frac-

    tions below 0.5[157]. Microgel particles were used as depletants (in-

    stead of polymers) to hard spheres and it was found that gels were

    formed more easily and eventually demixed[158].

    A much richer behavior is observed and actually predicted in star

    mixtures (with functionalities exceeding 64), placing them in the soft

    colloidal domain[71,159], where the softness leads to the formation

    of differenttypes of arrested states: not only those with one or bothpar-

    ticle populations trapped as in hard sphere mixtures, but also a novel

    asymmetric glass (AsG) where the effective large particle cages become

    anisotropic due to the osmotic pressure of the small particles (Fig. 3d)

    [148,160162]. Recently, mixtures involving block copolymer micelles

    with frozen and livingcores and short arms were studied, and it was

    found that a balance between crystallization and vitrication results

    from an optimum adjustment of the number of arms[163].

    Combining soft and hard interactions opens another route for ex-

    ploring and tuning macroscopic properties. A mixture of small hard

    spheres and large star polymers represents such a situation, where re-

    pulsive pair potentials which have different ranges are coupled

    [164166]. As depicted schematically in Fig. 3e, at low polymer star vol-

    ume fractions the colloidal liquid phase separates upon adding hard

    spheres. At higher volume fractions in the star glass regime, there is a

    small domain where the glass melts due to depletion and further addi-

    tion of hard spheres results in APS. Rheology was again used in order to

    identify the different transitions and the strength of the moduli for dif-

    ferent mixture compositions, whereas thegradual natureof the yielding

    process under shear was conrmed[166].

    This plethora of morphologies has important consequences on themacroscopic properties of mixtures. The various stable and metastable

    states discussed above in the context of colloidal mixtures possess dif-

    ferent rheological properties so that both linearand nonlinear viscoelas-

    tic characterizations can very sensitively diagnose their differences. At

    the same time this richness of behavior allows tailoring the ow of

    soft colloidal composites at wish[4,5,167].

    4. Linear viscoelasticity, diffusion and zero-shear viscosity

    4.1. Viscoelastic relaxation spectrum and plateau modulus

    The viscoelastic properties of soft particle suspensionsare exquisite-

    ly reected in the dependence of the storage and loss moduli on fre-

    quency. As the volume fraction increases, the rheological responsealters from viscoelastic liquid behavior where terminal regime is ob-

    served over the experimental frequency window, to solid-like behavior

    where the storage modulus dominates. The evolution of the viscoelastic

    spectra leading to solidication hasbeen studied in detail forsoft particles

    in the colloidal regime like emulsions and nanoemulsions[83,168,169],

    coreshell particles[3739,90,170172], and microgel particles[40,92,

    173175], which have softness parametersN104. The approach of the

    glass transition is signaled by the superpositionofG and G at higher fre-

    quencies beforea plateauin G anda shallowminimum in G appears. The

    onset of terminal relaxation shifts to lower frequencies as the volume

    fraction increases up to the point where it falls outside the experimental

    frequency window. These viscoelastic properties can be qualitatively

    understood in the framework of the cage model where a colloidal

    particle is topologically constrained by its neighbors in a hypothetical

    cage[9,79,80]. It moves locally within the cage (beta relaxation) but

    will escape only if the environment of the cage can renew at long

    times (alpha relaxation). The minimum ofGexpresses the transition

    from in-cage motion to out-of-cage motion, and it can be also probed

    in creep recovery experiments[176,177].

    The shape of the viscoelastic spectra remains qualitatively the same

    as the suspensions cross over the jamming transition althoughthere are

    some important quantitative differences that will be discussed later on

    in this section. The solidi

    cation scenario is not well-known yet forsofter particles with10104 although it is denitely more gradual

    [86]. Inside the solid state, the viscoelastic spectra of star-like particle

    suspensions also exhibit a plateau in Gand a minimum inGbut new

    features are observed at low frequencies where the relaxation of the

    interdigitated arms apparently contributes to an additional relaxation

    process[81]. Before alpha relaxation is activated, the arms can disen-

    gage from their neighbors and deform the cage. As a result, star-like

    colloidal glasses with many dangling arms often have an experimentally

    accessible alpha relaxation, marking a clear departure for other systems

    [178,179]. This may not necessarily be the case for colloidal block

    copolymer micelles which have a larger core and smaller dangling

    ends than stars and usually exhibit slightly larger particle elasticity

    [180182].

    In view of the huge diversity of soft particles in terms of effective

    elasticity and architecture, a real challenge is to draw a one to one

    comparison of their viscoelastic properties. This is a formidable task,

    which in general seems inaccessible, because of the difculty to deter-

    mine the glass and jamming transitions (Section 2) and the actual vol-

    ume fraction of the suspensions (Section 1.4). Even for a given class of

    particles such as star polymers, manifestations of softness like osmotic

    shrinking and interdigitation depend on functionality, and there are

    no systematic experimental studies[81,183]. An additional difculty

    comes from the fact that, unlike for polymers, mean-eld theories of

    colloids are in general not possible since the coordination number

    does not exceed a value of 12. The rationalization of experimental re-

    sults often relies on semi-phenomenological approaches or simulations.

    However, recent progress in theeld allows us to identify some impor-

    tant trends.

    Mode Coupling Theory (MCT)[184]has proven highly successful inpredicting the linear rheology of hard colloids on approaching the glass

    transition by accounting for density correlations from the static struc-

    ture factor and then calculating the viscoelastic moduli[185]. It also

    works well for soft glassy colloids with the same t parameters[84,

    186188]. Recent extensions of MCT[189191]have been developed

    to predict the frequency-dependent moduli of coreshell particles and

    star polymer suspensions well-within the glassy regime, which is be-

    yond the expected range of validity (up to and at the glass transition)

    of the initial theory[39,192]. However, MCT is unable to capture the

    low frequency relaxation which occurs in addition to alpha relaxation

    as already mentioned. For the sake of comparison with predictions,

    the experimental viscoelastic moduli have to be scaled with the entro-

    pic elasticity of individual particles, kT/R3, while the angular frequency

    is converted into the Peclet number:Pe = (6R3)/kT. These scalingforms are expected to be valid in the thermal glass regimeg b b Jdue to the entropic nature of the suspension dynamics [134]. Note

    that a successful alternative, albeit similar in spirit, statistical mechani-

    cal theory exists, which describes glassy dynamics based on a nonequi-

    librium free energy that incorporates local cage correlations and

    activated barrier hopping processes[193196].

    The properties of jammed suspensions exhibit several quantitative

    differences from those of entropic glasses because of the presence of

    repulsive elastic interactions which exceed thermal forces. The plateau

    modulus scales with the contact modulusE* of the particles (E* = E/

    (1 2); Eand are the Young modulus and the Poisson ratio, respec-

    tively) and exhibits a rapid linear increase with the distance to the jam-

    ming transition: G0/E* ( J). This has been observed for

    concentrated emulsions, microgel suspensions and even star polymer

    566 D. Vlassopoulos, M. Cloitre / Current Opinion in Colloid & Interface Science 19 (2014) 561574

  • 7/25/2019 Tunable Rheology of Dense Soft Deformable Colloids

    7/14

    solutions [45,86,197]. At high angular frequency, the loss modulus of

    many soft systems like emulsions and microgels increases likeG~ 0.5,

    which is the sign of anomalous dissipation[83,174,178,198,199].

    A recent comprehensive study combines particle scale simulations

    and a microscopic mean-eld theory[200]. The particles are treated

    like elastic non-Brownian spheres dispersed in a solvent (viscosity: s)

    at high volume fraction, which interact through Hertz-like potential

    and elastohydrodynamic lubrication forces[48]. The theory gives pre-

    dictions for the pair correlation function of the suspensions, the highfrequency moduli, and the osmotic modulus [200]. Simulations also

    allow computing the frequency dependence of the viscoelastic moduli

    including the high frequency regime and the variations of the low-

    frequency plateau modulus with volume fraction[45,135]. The predic-

    tions compare well with experiments without adjustable parameters

    when the moduli are scaled by the contact modulusE* of the particles

    and the frequency by the characteristic time = s/E*. Other models

    have been proposed to take into account the complex interfacial struc-

    ture of emulsions and the non-homogeneity of coreshell particles

    [201,202]. Similar investigations for star-like particles are still missing

    but are highly desirable.

    4.2. Zero-shear viscosity and self-diffusion

    The most striking manifestation of the difference in softness (hence

    interaction potential) among colloidal particles appears in the terminal

    viscoelastic regime, and a relevant macroscopic quantity is the zero-

    shear viscosity of the suspensions. It turns out that it is possible to

    build a generic plot where the zero-shear viscosity of various systems

    scaled with the solvent viscosity is represented versus the effective hy-

    drodynamic volume fraction [57,203]. Thisis depicted in Fig. 4 whichfor

    clarity only includes one data set from each particle type: hard spheres

    [204], high functionality stars[86], block copolymer micelles[32]and

    microgels[53]. Interestingly, the limited long-time self-diffusion data

    available for hard spheres and stars (not shown here) are in excellent

    agreement with the viscosity data when plotted accordingly, hence

    conrming the validity of the StokesEinsteinSutherland relation

    even at large volume fractions [81,205207]. Data for colloidal particlesof any softness (e.g., stars, grafted particles, block copolymer micelles,

    microgels) could be included in such a plot, which reveals the role of

    softness in reducing the relative viscosity and increasing the effective

    maximum packing fraction[27,61,170,208].However, it is important

    to emphasize that the transition to a weaker viscosity at higher frac-

    tions, and eventually to divergence, also signals osmotic de-swelling

    or shrinking. This phenomenon, which can also be triggered externally

    by temperature of ionic strength variation,fullyreects particle softness

    and reduces the actual volume fraction and stiffness as discussed above

    [54,61,120,209,210]. Forpolyelectrolyte particlesit hasbeen shown that

    the divergence of the viscosity exhibits little departure from the hard

    sphere case once the volume fraction is appropriately corrected for

    particle shrinkage[53].The type of representation shown inFig. 4emphasizes the fact that

    softness affects only the high volume fraction regime, where particle in-

    teractions come into play. At low volume fractions reduced viscosities

    for all particles collapse to the EinsteinSutherland and Batchelor curves

    [57]. But on increasing volume fraction the data depart according to

    their softness. Recent extensive mesoscale simulation studies, accounting

    for the effects of hydrodynamics as well, successfully predict the volume

    fraction dependent viscosity of linear and star polymers shown inFig. 4

    [211213]. Several empirical models have been proposed to describe

    the viscosity departure and divergence at the glass transition. The

    KriegerDougherty and Quemada models are very successful for a wide

    range of systems[11,203]. The approach to glass transition is a subtle

    issue because in reality there do not exist true hard sphere experimental

    systems due to stabilization needs [214]. Inthis case even models used in

    molecular glasses were invoked to capture the steep viscosity increase

    [215]. To further characterize the variations of the viscosity before solidi-

    cation, the concept of fragility was used [47,89]. It was proposed to build

    Angell-type fragility plots by dening a glass transition volume fraction

    and plotting the normalized viscosity of the suspension against its dis-

    tance from this glass transition[89]. The outcome of this analysis is that

    softer colloids make stronger glasses and indeed, a closer look at the

    trend ofthe datain Fig. 4 conrms this conclusionand suggeststhat poly-

    mer coils are potentially the strongest glasses.

    4.3. Aging

    A comment on aging is in order to close this section. Disordered

    crowded materials, such as glassy and jammed colloids exhibit a slow

    time evolution of their dynamic and structural properties, also termedaging. There are several features of aging which are shared by many col-

    loidal glasses, such as a certain time regime characterized by a logarith-

    micincrease of storage modulus and slowing-downof dynamics. On the

    other hand, the dependence or not of macroscopic properties on the age

    and the exact form of the respective scaling, as well as the presence of

    different aging regimes for different systems, represent distinct signa-

    tures of softness[134,179,199,216,217].We shall not discuss aging fur-

    ther, as this has been the subject of other reviews[218,219]. However

    we emphasize that irrespectively of its study, aging must be properly

    accounted for when analyzing properties of glassy and jammed mate-

    rials. In this respect, an appropriate preparation protocol is necessary

    to place the material in a reproducible state. It consists of shear-

    melting the glass or jammed state in analogy to thermal annealing of

    molecular or polymeric glasses, and following the dynamics over timeuntil a quasi-steady state is reached. This is not necessary a trivial task

    since internal stresses may affect the dynamics, hence proper denition

    of a protocol and careful implementation are important for obtaining

    consistent data[179,220222]. The properties discussed above such as

    the volume fraction dependence of the moduli or the frequency spec-

    trum of viscoelastic moduli, all refer to aged systems.

    5. Nonlinear rheological phenomena

    5.1. Yielding behavior

    Concentrated suspensions of soft particles exhibit solid-like proper-

    ties at rest but ow at large stresses. When the applied stress is de-

    creased, the material is progressively slowed-down and eventually

    Fig. 4.Dependence of the zero-shear viscosity (normalized to the solvent viscosity) of

    various colloidal systems on the effective volume fraction. Selected data are shown for

    suspensions of hard spheres[204], microgels[53], block copolymer micelles [32], and

    high functionality star polymers[86].The dashed line is the bestt of hard sphere data

    to the Quemada model[53]. The low-shear viscosity data of star polymer data are well

    represented by a double exponential function (dotted line)[86].

    567D. Vlassopoulos, M. Cloitre / Current Opinion in Colloid & Interface Science 19 (2014) 561574

    http://localhost/var/www/apps/conversion/tmp/scratch_4/image%20of%20Fig.%E0%B4%80
  • 7/25/2019 Tunable Rheology of Dense Soft Deformable Colloids

    8/14

    immobilized at a nite value of the stress, called dynamic yield stress,

    which for the present discussion will be simply called yield stress,y.

    Note that the static yield stress represents the stress needed to apply

    in order to induce macroscopicow from rest. In general, the yield

    stresscan be viewed as a viscosity bifurcation: the application of a stress

    is followed by ow stoppage (b y) or steady deformation ( Ny)

    depending on the stress value[216,223225]. There are many other

    ways of determining the yield stress[226], the most straightforward

    and frequently used being the strain amplitude sweep tests[131,174,188,199,227,228]. Recently, the application and analysis of Large

    Amplitude Oscillatory Shear (LAOS) has emerged as a powerful tool

    for probing the yielding andow properties of glasses and jammed sus-

    pensions[133,135,227,229232].

    Referring to the old dilemma whether glasses ow or not, there is

    consensus that yielding exists, i.e. an external eld can drastically alter

    the structure of a material, hence its macroscopic properties[233,234].

    However, there is stilla debate as to whether yield stress is a true mate-

    rial parameter or not. A useful distinction has been made between true

    yield stress materials and thixotropic materials exhibiting apparent

    yield stress[11,235238].

    Densely packed systems such as glassy and jammed particle suspen-

    sions (sometimes termed pastes) made of soft repulsive colloids, which

    are the focus in this review, are true yield stress materials. Yielding oc-

    curs through a continuous break and reformation of the cages which

    keep the particles motionless at rest. In this context, LAOS is a unique

    tool due to the possibility to reproduce within a cycle the sequence of

    events that characterize ow and yielding [133,229,239,240]. One inter-

    esting outcome is the effective cage modulus, which is dened as the

    slope of the stress vs. strain amplitude plot at zero stress and appears

    to match the value of the bulk (plateau) modulus G0 and remains insen-

    sitive to the strain or stress amplitude[130,133]. A natural question at

    this point is how do yield stress and strain compare for different soft

    jammed colloids and depend on volume fraction[131,134]. Due to the

    prominence of elastic effects, yis often proportional toG0with the co-

    efcient of proportionality being the yield strain y: y = G 0y. As

    discussed inSection 4.1,the storage modulus of soft particle suspen-

    sionsscales with kT/R3 in the entropic glass regime, andthe particle con-

    tact modulus E* in the jammed regime. In the literature not muchattention has been paid to the yield strain [199,227,228].Simulations

    and theoretical modeling have predicted that the yield strain of soft

    elastic spheres in the jammed regime should increase exponentially

    with the distance to the jamming transition (it varies from low values

    of about 0.02 nearJto about 0.08 for = 0.9)[49]. This prediction is

    relatively well obeyed by emulsions[48,49,85,168]and microgel parti-

    cles [48,134]. Note that entropic hard-sphere glasses have a non-

    monotonicyield strain with a maximum value of about0.15at a volume

    fraction between the glassand close-packing volumefractions[176,227,

    241]. Combining the scaling expected forG0andyprovides the varia-

    tions of the yield stress. Due to their long-hairy conformation, star-like

    colloids exhibit distinct yielding properties[86,130,134,199,228]. Since

    terminal alpha relaxation is accessible in these systems even at very

    high concentrations, as explained inSection 4.1, yielding exists only ina limited range of frequencies or time scales. Moreover, the yield strain

    depends very weakly on concentration. These ndings call for more

    detailed investigations.

    Colloidal gels ofocculated dispersions with attractive interactions

    mostly belong to the class of thixotropic materials[11,242,243]. The

    microstructure responsible for solid-like properties at b y is

    destroyed by the ow, which induces a sudden drop of viscosity and

    substantial structural changes. The disruption of the structure under

    ow is reversible but typically characterized by a large hysteresis as

    the structural timescale is usually longer than the inverse shear rate

    [243,244]. Due to their microstructure, stars in the weak glassy regime

    in temperature-sensitive solvents exhibit thixotropic-like response

    [245]. Although there is some confusionin the terms,the slow evolution

    of thixotropic materials is distinct from the aging and rejuvenation

    phenomena which are known to occur in purely repulsive systems

    [246249]. On the other hand, the yield stress reects in part the

    strength of bonds and/or cages. Its determination from the ow curves

    poses serious problems because homogeneous ow ends-up over a

    limiting range of low shear rates[250]. We will not review the eld of

    thixotropy here[11,251]. However, it is worth emphasizing that some

    features of thixotropic materials like non-homogeneous ow at low

    shear rates are also found in repulsive star-like colloids[178,249]. This

    shows that the distinction between true and apparent yield stressmaterials is not as denite as often believed, and that some materials

    share properties of different classes.

    5.2. Flow properties

    An important question concerns the establishment of steady-state

    ow conditions in the form of stressshear rate relationships or ow

    curves. In most experiments, measurements start at high shear rates

    (associated with the so-called rejuvenated state) where steady state is

    reached rapidly, and the shear rate is progressively decreased down to

    the yield point. Applying a stress larger but close to the yield stress is

    followed by extremely long transient phenomena, after which steady

    state is eventually reached[249,252254]. In practice, the shape of the

    steady state ow curves constitutes a remarkable ngerprint of the

    system under investigation. In many soft colloidal suspensions like

    microgels, emulsions, polystyrenePNIPAm coreshell particles, and

    star polymers, the onset of the glass transitionis signaled byow curves

    exhibiting two branches at low and large shear rates, separated by a

    slowly increasing plateau at intermediate rates (in double logarithmic

    coordinates) [86,90,168]. Thelow-shear rate branchcorresponds to ter-

    minal behavior, and the high shear-rate branch reects shear-thinning.

    As the volume fraction increases, terminal behavior shifts to very low

    shear rates due to the huge increase of the -relaxation time, up to

    the point where it falls outside the experimental window; meanwhile

    the intermediate plateau widens and attens. In the entropic glass re-

    gime, the ow curves are conveniently represented in dimensionless

    coordinates as: kT=R3 f

    R2=D0

    . Here, kT/R3 represents the char-

    acteristic scale of the yield stress while R2=D0 is a characteristic time

    scale associated with the self-diffusion of individual particles (D0isthe self-diffusion coefcient at innite dilution). In this regime, MCT

    provides remarkable quantitative predictions of theow curves, sug-

    gesting that the effects of softness and deformability can be effectively

    lumped into the structure factor at least at low shear-rates[90,189,

    190,255]. Other nonlinear features like creep behavior [256], stress

    overshoots during shear ow start-up[257,258]and the built up of

    residual stress[222,259]are also predicted by MCT[260].

    At higher volume fractions corresponding to the regime of jammed

    suspensions, the ow curves (again in double-logarithmic representa-

    tion) of a wide range of viscoplasticuids with yielding properties are

    characterized by a stress plateau at low-shear rates and a shear-

    thinning behavior at high shear rates[217,246,261,262]. Empirically,

    they arevery well represented by theHerschelBulkley phenomenolog-

    ical equation: y kn

    withan exponent n which for different ma-terials has beenfound to be close to 0.5. The valueof the yield stress isin

    good agreement with independent measurements based on other tech-

    niques. The consistency parameterkdepends on material properties. It

    has been shown[262]that normalizing the stress and shear rate with

    yield stressyand time scale S/G0(Sbeing the solvent viscosity), re-

    spectively, collapses the data obtained for different jammed materials

    onto a master curve (Fig. 5). A recent micromechanical model quantita-

    tively accounts for these results and provides a physical understanding

    of yielding and ow in soft colloidal particles[49]. The suspensions are

    modeled as three-dimensional packings of non-Brownian elastic

    spheres which are subject to repulsive central forces and

    elastohydrodynamic (EHD) drag forces analogous to those at the origin

    of slip near solid surfaces [263,264]. This EHD model provides quantita-

    tive predictions for the ow curves[49], the microscopic dynamics and

    568 D. Vlassopoulos, M. Cloitre / Current Opinion in Colloid & Interface Science 19 (2014) 561574

  • 7/25/2019 Tunable Rheology of Dense Soft Deformable Colloids

    9/14

    macroscopic rheology in LAOS[135], as well as the internal stress that

    persists upon ow cessation[221].

    Inan attemptto proposea unied description, computer simulations

    of concentrated suspensions of soft repulsive particles involving ther-

    maluctuations and viscous dissipation were performed[87,88]. They

    predicted a transition in ow curves from power-law to plateau stress

    vs. rate at jamming, and an empirical model of the form Glass

    Jammed s

    was proposed to describe the entire rheology. Scaling

    relations extracted from simulations were used to model the glassy

    and jamming contributions to the stress.

    As discussed above, the distinction between the entropic glass re-gime and jammedregime is not alwaysclear. Recent work with PNIPAm

    microgel particles at volume fractions close to the jamming transition

    suggests that the stress plateau occurs only above the jamming transi-

    tion whereas a non-Newtonian behavior is observed below [91,92,

    265]. These results conrm that the stress plateau is a robust signature

    of jamming. However, here the system jumps directly from the liquid

    state to the jammed state without crossing the entropic glass regime.

    The ow curves measured above and below the yield point collapse

    onto two branches exhibiting power-law scalings of shear stress and

    shear rates with the distance to the jamming transition. The same

    type of critical scaling collapse was reported using numerical simula-

    tions and simple models of soft particles interacting through repulsive

    interactions, but the exact values of the exponents are largely model-

    dependent[266269]. These results are qualitatively consistent withthe EHD micromechanical model, albeit again with different exponents,

    leaving the issue partially unsettled. Two parameters seem to be impor-

    tant: the distance to the jamming point (EHD model results are for

    suspensions well inside the jammed regime) and the exact form of the

    viscous drag force between particles[270].

    Despite the link between the existence of a stress plateau at low

    shear rates and jamming that has been mentioned in the previous

    paragraph, there are situations where the plateau is not indicative of

    true yielding behavior or does not even exist. A deviation of the low-

    rate stressfroma true plateaumay also reect theparticle's internal mi-

    crostructure, even when aging is carefully accounted for. A powerful

    manifestation of these effects for systems where a yield stress is dened

    from dynamic strain sweep experiments was reported for star polymers

    [86]. An interpretation of this remains elusive, but with slip being

    excluded, effects like the cooperative local particle rearrangement,

    which for stars and long hairy particles is manifested via interpenetra-

    tion[5,81,82,205]can contribute to slow relaxations visible at low

    shearrates. A phenomenological descriptionof such behavior is possible

    using a modied HerschelBulkley model where Am

    B n

    [86].

    In some cases a stressplateauexists but it is associated instead to in-

    homogeneousow structure where the ow splits into two bands of

    different viscosities, one that is essentially uid (Ny), whereas the

    other remains solid (b

    y), separated along the

    ow gradient directionandcoexisting at thesame shear stress. Such a situation hasbeen mainly

    reported in thixotropic suspensions [225,236,237]but it also exists in

    some star polymer suspensions [178,249,271] where depending on vol-

    ume fraction and softness (which can be tuned via functionality as

    discussedabove) a plateauor weak power-law in stresscan be observed

    (Fig. 5). Interested readers are referred to different reviews on the topic

    [272276]. A numberof colloidal systems have been shown experimen-

    tally to exhibit gradient banding in addition to starpolymers: surfactant

    solutions[277], hard sphere suspensions[278], block copolymer mi-

    celles[275,279281], and carbopol microgels[282]. Presently there is

    no consensus on the origin of shear-banding in soft colloidal material

    and many different explanations have been proposed: coupling be-

    tweenow and concentration eld[278],mechanical instability and

    mobileimmobile phase coexistence in the low-shear region [178,

    283], slow relaxation dueto transient bonding [44,284], competitionbe-

    tween aging and rejuvenation[271,285], competition between micro-

    structural time and the characteristicow time scale[286]. Returning

    toFig. 5,it is remarkable that one can tune the ow properties of soft

    colloids (such as stars) at a molecular level, and in particular induce or

    avoid shear banding. Note that for stars with short arms, mesoscopic

    simulations did not produce any shear banding for functionalitiesf 50[212,213]. The fundamental question, which is still under debate

    and needs further investigations, is whether there are generic features

    of banding in colloids and what is the exact role of softness. For the

    star case in particular, the exact role of molecular parameters (number

    and size of arms) remains elusive[212,213,271,284]. In general, the in-

    terplay of shear banding, yielding, slip and aging in dense soft colloids is

    subtle. Clearly, there are ample opportunities for further exploring the

    nonlinear shear rheology of colloidal materials including mixtures andunderstanding the possible connection of banding to shear-induced

    transitions in soft matter.

    A nal aspect of signicance in this topic is the role of particle

    surface interactions in the macroscopic rheological response of jammed

    materials. In a recent investigation of the ow and velocity proles of

    microgel suspensions with purely repulsive interactions it was found

    that surface roughness and surface chemistry have a strong effect on

    ow properties[287]. Below theyield stress, wall slip is of hydrodynam-

    ic origin when the surfaces are smooth and the particlesurface interac-

    tions repulsive. Slip can be suppressed or reduced by attractive surfaces

    [288]. When attractiveinteractionsare weak, slip canoccur as a resultof

    EHD lubrication. Moreover, yielding depends on the surfaces again: it is

    uniform near rough or smooth repulsive surfaces and, hence, the bulk

    ow properties are recovered. It is non-uniform near smooth attractivesurfaces, suggesting that different dynamical states coexist, forming a

    macroscopic boundary layer (and not an immobile band). This marks

    a clear difference from bulk properties and from the behavior of non-

    thixotropic materials discussed above. This important development,

    which was recently extended to account for slip heterogeneities in

    microchannel ows with different particlesurface interactions[289],

    demonstrates another possibility to tailor the ow properties of soft

    colloids[290].

    5.3. Mesoscopic modeling of soft colloidal rheology

    Besides MCT and micromechanical models discussed above, the rich

    behavior of soft colloids at high volume fractions has stimulated a

    wealth of phenomenological modeling activities. While these models

    Fig. 5.Master ow curves representing normalized shear stress vs. normalized shear rate

    relationship for microgels (, ) and star solutions (,;) at high effective volume

    fractions eff, withf: star functionality and Ma: arm molar mass. The dotted line is at of

    the data for microgels to the HerschelBulkley model[48].The dashed line is a t of the

    data forstarpolymers withf= 390 andMa = 24 kg/mol tothemodied HerschelBulkley

    model (see text)[86].The data for star polymers withf= 122 and Ma= 72 kg/mol col-

    lapse onto the data for the other star polymers and the microgels at high shear rates;

    the plateau observedat low-shear ratesis indicative of shear-banding(continuousstraightline)[178].

    569D. Vlassopoulos, M. Cloitre / Current Opinion in Colloid & Interface Science 19 (2014) 561574

    http://localhost/var/www/apps/conversion/tmp/scratch_4/image%20of%20Fig.%E0%B5%80
  • 7/25/2019 Tunable Rheology of Dense Soft Deformable Colloids

    10/14

    involve many parameters not directly accessible in experiments, they

    enjoyreasonable success in reproducingseveral features of the rheology

    of dense suspensions. For example, the Soft Rheology Model (SGR) uses

    trap dynamics for local mesoscopic non-ergodic events, with the effec-

    tive noise temperature parameterx characterizing the jump rate out

    of local traps[291,292]. Appropriate variation ofx allows to reproduce

    macroscopic yield and ow dynamics observed in steady, oscillatory

    and transient ows as well as viscosity bifurcation and thixotropy

    [285]. Historically the SGR model has opened the route for many phe-nomenological models, which are impossible to review in depth here.

    In the following we mention briey a few of them and we refer the in-

    terested reader to more specialized articles [293,294]. The starting

    point of all these models is a coarse-grained representation of the mate-

    rial divided into mesoscopic cells, where the stress evolves according to

    a prescribed law. In the uidity model, the local state of the system is

    expressed by a Maxwell equation which involves the local relaxation

    frequencyof thestress, calledthe uidity [172,217,295]. The timeevo-

    lution of the uidity results from the competition between aging, which

    decreases theuidity as time passes, andow-induced rearrangements,

    which increases uidity. Further developments of the uidity model

    take into account the spatial variation of the uidity, which can be

    used to reproduce the occurrence of various spatial heterogeneities

    like shear banding[249,271].An approach bearing similarities with

    theuidity model is the non-local cooperative model which accounts

    for the cooperative nature of rearrangements by relating the spatial

    stress response in each cell to the global stress uctuations occurring

    over the entire system[294,296298]. A characteristic cooperative

    length was proposed to exist and be intrinsic material property, some-

    thing that is not totally supported by existing experiments. This coopera-

    tive model has been used to explain deviations observed in concentrated

    emulsions from bulk rheology in conned geometries[299301].

    6. Conclusions and perspectives

    This partial presentation of the views of the community on the

    rheology of soft colloids demonstrates the unprecedented richness

    imparted by softness on the properties of colloidal suspensions. It is ev-

    ident that the eld of soft colloids is picking-up. We believe that themajor advantage of these systems is their tunability which allows tailor-

    ing theirow properties at molecular level. This indicates that under-

    standing the ne details of internal microstructure of the particles is

    often of crucial importance[4,5,128,302]. The two most important dis-

    tinct features are their change of shape/size (e.g., microgels) and their

    interdigitation (e.g., stars) in the dense state. It is evident from the

    above discussion that several outstanding challenges remain. We list

    them below.

    The existence and the nature of the glass and jamming transitions in

    soft dispersions pose several important questions. Does a glass to jam-

    ming transition exist for all types of soft colloids (say microgels and

    stars)? Is there universality in scaling the viscoelastic properties with

    the actual volume fraction? In relation to this, aging and the related

    role of internal stresses need further elaboration for the different softcolloidal archetypes.

    Osmotic effects due to additives are shown to yield new dynamically

    arrested states. New types of mixtures and other combinations of size

    ratio and volume fractions must be investigated to deeply understand

    the state diagrams. The rheology of these states and the possible ow-

    induced transitions remain largely unexplored.

    In the eld of nonlinear rheology, linking the scaling ofow curves

    in the glassy and jammed regions represents a grand challenge. Along

    the same lines, the high-shear and extensional rheology of soft colloids

    in these states are entirely unexplored territories and their possible

    combinations with rheo-physical observations of particle/structure de-

    formation will advance the eld and also contribute to our understand-

    ing of the relations among various types of shear-banding, yielding, and

    slip with respect to particle architecture and thenature of the conning

    surfaces. Moreover, the effects of softness and volume fraction on the

    remarkable shear thickening phenomenon associated with colloidal

    dispersions have received very little attention so far [303,304]. Recently,

    discontinuous shear thickening was discussed in friction-dominated

    large hard sphere suspensions (somehow equivalent to jammed parti-

    cles) and the principal role of contact friction was identied[305,306].

    To go beyond entropic spherical systems, variations of temperature

    or pressure can change the quality of the background solvent and

    hence introduce attractions. Likewise, the presence of functional end-groups can lead to clustering and hierarchical order with the possibility

    of obtaining well-characterized patchy supramolecular structures[307]

    whose rheology is unknown. Finally, departure from spherical shape

    combined with softness is expected to change the packing of colloids

    and hence their structure and rheology. These are promising new direc-

    tions. The eld remains as exciting as ever!

    Acknowledgments

    This review was written when D.V. was visiting ESPCI ParisTech

    with the support of the Michelin Chair. We thank Jan Mewis for helpful

    comments on the manuscript ESPCI ParisTech is member ofPSL Re-

    search University.

    References

    [1] Caruso F, editor. Colloids and colloid assemblies: synthesis, modication, organiza-tion and utilization of colloid particles. Weinheim: Wiley-VCH; 2004.

    [2] Stokes JR, Frith WJ. Rheology of gelling and yielding soft matter systems. Soft Mat-ter 2008;4:113340.

    [3] Chen DTN, Wen Q, Janmey PA, Crocker JC, Yodh AG. Rheology of soft materials.Annu Rev Condens Matter Phys 2010;1:30122.

    [4] Bonnecaze R, Cloitre M. Micromechanics of soft particle glasses. Adv Polym Sci2010;236:11762.

    [5] Vlassopoulos D, Fytas G. Frompolymers to colloids: engineering the dynamic prop-erties of hairy particles. Adv Polym Sci 2010;236:154.

    [6] Tadros TF, editor. Colloids and interface science series, volumes 16. Weinheim:Wiley-VCH; 2010.

    [7] Dupont J, LiuG. ABC triblock copolymer hamburger-like micelles, segmented cylin-ders, and Janus particles. Soft Matter 2010;6:365461.

    [8] Bosko JT, Prakash RJ. Universal behaviour of dendrimer solutions. Macromolecules2011;44:660.

    [9] Pusey PN. Colloidal suspensions. In: Hansen JP, Levesque D, Zinn-Justin J, editors.Liquids, freezing and glass transition. Amsterdam: North Holland; 1991.

    [10] Russel WB, Saville DA, Schowalter WR. Colloidal dispersions. New-York:Cambridge University Press; 1989.

    [11] Mewis J, Wagner NJ. Colloidal suspension rheology. New York: Cambridge Univer-sity Press; 2012.

    [12] Daoud M, Cotton JP. Star shaped polymers: a model for the conformation and itsconcentration dependence. J Phys (Paris) 1982;43:5318.

    [13] Senff H, Richtering W. Inuence of cross-link density on rheological properties oftemperature-sensitive microgel suspensions. Colloid Polym Sci 2000;278:83040.

    [14] Fernandez-Nieves A, Wyss H, Mattsson J, Weitz DA, editors. Microgel suspensions:fundamentals and applications. Weinheim: Wiley-VCH; 2011.

    [15] Princen HM. Rheology of foams and highly concentrated emulsions: I. Elastic prop-erties and yield stress of a cylindrical model system. J Colloid Interface Sci 1983;91:16075.

    [16] Abreu D, Levant M, Steinberg V, Seifert U. Fluid vesicles in ow. Adv Colloid Inter-face Sci 2014;208:129.

    [17] Eichenbaum GM,Kiser PF,Dobrynin AV, Simon SA,NeedhamD. Investigation of theswelling response and loading of ionic microgels with drugs and proteins: the de-pendence on cross-link density. Macromolecules 1999;32:486778.

    [18] Hashmi SM, Dufresne ER. Mechanical properties of individual microgel particlesthrough the deswelling transition. Soft Matter 2009;5:36828.

    [19] Wyss HM, Franke T, Mele E, Weitz DA. Capillary micromechanics: measuring theelasticity of microscopic soft objects. Soft Matter 2010;6:45505.

    [20] Nordstrom KN, Verneuil E, Ellenbroek WG, Lubensky TC, Gollub JP, Durian DJ. Cen-trifugal compression of soft particle packings: theory and experiments. Phys Rev E2010;82:041403.

    [21] Sierra-Martin B, Frederick JA, Laporte Y, Markou G, Lietor-Santos JL, Fernandez-Nieves A. Determination of the bulk modulus of microgel particles. Colloid PolymSci 2011;289:7218.

    [22] Abate AR, Han L, Jin L, Suob Z, Weitz DA. Measuring the elastic modulus ofmicrogels using microdrops. Soft Matter 2012;8:100325.

    [23] Gonzalez-Rodriguez D, Guevorkian K, Douezan S, Brochard-Wyart F. Soft mattermodels of developing tissues and tumors. Science 2012;338:9107.

    [24] Ballauff M, Likos CN. Dendrimers in solution: insight from theory and simulation.Angew Chem Int Ed 2004;43:2998.

    [25] Nikoubashman A, Likos CN. Branched polymers under shear. Macromolecules

    2010;43:1610.

    570 D. Vlassopoulos, M. Cloitre / Current Opinion in Colloid & Interface Science 19 (2014) 561574

    http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0005http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0005http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0005http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0005http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0010http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0010http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0010http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0010http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0015http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0015http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0015http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0015http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0020http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0020http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0020http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0020http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0025http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0025http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0025http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0025http://refhub.elsevier.com/S1359-0294(14)00094-6/rf1495http://refhub.elsevier.com/S1359-0294(14)00094-6/rf1495http://refhub.elsevier.com/S1359-0294(14)00094-6/rf1495http://refhub.elsevier.com/S1359-0294(14)00094-6/rf1495http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0030http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0030http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0030http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0030http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0035http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0035http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0040http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0040http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0045http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0045http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0050http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0050http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0055http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0055http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0055http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0055http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0060http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0060http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0060http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0060http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0060http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0060http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0065http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0065http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0070http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0070http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0070http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0070http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0070http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0075http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0075http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0075http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0075http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0080http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0080http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0080http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0080http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0080http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0085http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0085http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0085http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0085http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0090http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0090http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0090http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0090http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0095http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0095http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0095http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0100http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0100http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0100http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0100http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0100http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0105http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0105http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0105http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0105http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0110http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0110http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0110http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0110http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0115http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0115http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0120http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0120http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0120http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0120http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0115http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0115http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0110http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0110http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0105http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0105http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0100http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0100http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0100http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0095http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0095http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0095http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0090http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0090http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0085http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0085http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0080http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0080http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0080http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0075http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0075http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0070http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0070http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0070http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0065http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0065http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0060http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0060http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0055http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0055http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0050http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0050http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0045http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0045http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0040http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0040http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0035http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0035http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0030http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0030http://refhub.elsevier.com/S1359-0294(14)00094-6/rf1495http://refhub.elsevier.com/S1359-0294(14)00094-6/rf1495http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0025http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0025http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0020http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0020http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0015http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0015http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0010http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0010http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0005http://refhub.elsevier.com/S1359-0294(14)00094-6/rf0005
  • 7/25/2019 Tunable Rheology of Dense Soft Deformable Colloids

    11/14

    [26] Watanabe H, Kotaka T. Rheology of ternary mixtures of styrenebutadiene diblockcopolymer, homopolybutadiene, and n-tetradecane. J Rheol 1983;27:22340.

    [27] BuitenhuisJ, ForsterS. Block copolymer micelles: viscoelasticity andinteraction po-tential of soft spheres. J Chem Phys 1997;107:26272.

    [28] Pham QT, Russel WB, Thibeault JC, Lau W. Micellar solutions of associative triblockcopolymers: entropic attraction and gasliquid transition. Macromolecules 1999;32:29963005.

    [29] Li Z, Kesselman E, Talmon Y, Hillmyer MA, Lodge TP. Multicompartment micellesfrom ABC miktoarm stars in water. Science 2004;306:98101.

    [30] Laurati M, Stellbrink J, Lund R, Willner L, Richter D, Zaccarelli E. Starlike micelleswith starlike interactions: a quantitative evaluation of structure factors and phase

    diagram. Phys Rev Lett 2005;94:195504.[31] Sayeed A, Lodge TP. Depletion interactions: a new control parameter for the self-assembly of diblock copolymer micelles. Phys Rev Lett 2007;99:137802.

    [32] Merlet-Lacroix N, Di Cola E, Cloitre M. Swelling and rheology of thermoresponsivegradient copolymer micelles. Soft Matter 2010;6:98493.

    [33] Willenbacher N, Vesaratchanon JS, Thorwarth O, Bartsch E. An alternative route tohighly concentrated, freely owing colloidal dispersions. Soft Matter 2011;7:577788.

    [34] Mewis J, Frith WJ, Strivens TA, Russel WB. The rheology of suspensions containingpolymerically stabilized particles. AIChE J 1989;35:41522.

    [35] Witten TA, Pincus PA. Colloid stabilization by long grafted polymers. Macromole-cules 1986;19:250913.

    [36] ZackrissonM, StradnerA, Schurtenberger P, Bergenholtz J. Structure, dynamics andrheology of concentrated dispersions of poly(ethylene glycol)-grafted colloids.Phys Rev E 2006;73:011408.

    [37] Senff H, Richtering W, Norhausen C, Weiss A, Ballauff M. Rheologyof a temperaturesensitive coreshell latex. Langmuir 1999;15:1026.

    [38] Deike I, Ballauff M, Willenbacher N, Weiss A. Rheology of thermosensitive latexparticles including the high-frequency limit. J Rheol 2001;45:70920.

    [39] Crassous JJ, Siebenbrger M, Ballauff M, Drechsler M, Henrich O, Fuchs M.Thermosensitive coreshell particles as model systems for studying the ow be-havior of concentrated colloidal dispersions. J Chem Phys 2006;125:204906.

    [40] Ballauff M. Spherical polyelectrolyte brushes. Prog Polym Sci 2007;32:113551.[41] Likos CN, Lwen H, Watzlawek M, Abbas B, Jucknischke O, AllgaierJ, et al.Star poly-

    mers viewed as ultrasoft colloidal particles. Phys Rev Lett 1998;80:44504.[42] Fof G, Sciortino F, Tartaglia P, Zaccarelli E, Verso FL, Reatto L, et al. Structuralarrest

    in dense star-polymer solutions. Phys Rev Lett 2003;90:238301.[43] Stiakakis E, Wilk A, Kohlbrecher J, Vlassopoulos D, Petekidis G. Slow dynamics,

    aging, and crystallization of multiarm star glasses. Phys Rev E 2010;81:020402(R).

    [44] Padding JT,van Ruymbeke E, Vlassopoulos D, Briels WJ.Computersimulation of therheology of concentrated star polymer suspensions. Rheol Acta 2010;49:47384.

    [45] Seth JR, Cloitre M, Bonnecaze RT. Elastic properties of soft particle pastes. J Rheol2006;50:35376.

    [46] Zhang Z, Xu N, Chen DTN, Yunker P, Alsayed AN, Aptowicz KB, et al. Thermal ves-tige of the zero-temperature jamming transition. Nature 2009;459:2303.

    [47] Yang J, Schweizer KS. Glassy dynamics and mechanical response in denseuids ofsoft repulsive spheres. I. Activated relaxation, kinetic vitrication, and fragility. JChem Phys 2011;134:204908.

    [48] Seth JR, Mohan L, Locatelli-Champagne C, Cloitre M, Bonnecaze RT. Amicromechanical model to predict the ow of soft particle glasses. Nat Mater2011;10:8