You are on page 1of 47

Accepted Manuscript

Experimental synthesis and characterization of rough particles for colloidal and


granular rheology

Lilian C. Hsiao, Shravan Pradeep

PII: S1359-0294(19)30008-1
DOI: https://doi.org/10.1016/j.cocis.2019.04.003
Reference: COCIS 1286

To appear in: Current Opinion in Colloid & Interface Science

Received Date: 12 February 2019


Revised Date: 11 April 2019
Accepted Date: 12 April 2019

Please cite this article as: Hsiao LC, Pradeep S, Experimental synthesis and characterization of rough
particles for colloidal and granular rheology, Current Opinion in Colloid & Interface Science, https://
doi.org/10.1016/j.cocis.2019.04.003.

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to
our customers we are providing this early version of the manuscript. The manuscript will undergo
copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please
note that during the production process errors may be discovered which could affect the content, and all
legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
EP
C
AC
Increasing roughness and friction
Increasing dissipation in sheared suspensions
ACCEPTED MANUSCRIPT

PT
RI
SC
Experimental synthesis and characterization of rough

U
particles for colloidal and granular rheology
AN
M

Lilian C. Hsiao*,a & Shravan Pradeepa


D

a
TE

Department of Chemical and Biomolecular Engineering, North Carolina State University, 911

Partners Way, Raleigh, NC 27695, USA


EP

*
Corresponding author: lilian_hsiao@ncsu.edu
C
AC
ACCEPTED MANUSCRIPT

1 ABSTRACT

2 We review the experimental synthesis of smooth and rough particles, characterization of

3 surface roughness, and quantification of the pairwise and bulk friction coefficient in the rheology

4 of dense particulate flows. Even in the absence of interparticle attraction or cohesion, such types

PT
5 of flows are broadly ubiquitous, spanning enormous length scales ranging from consumer and

RI
6 food products to earth movements. The increasing availability of model frictional particles is

7 useful to advancing new understanding of particulate rheology. Although hard sphere particles

SC
8 remain the most widely studied system due to their simplicity, their rigid and frictionless nature

9 cannot predict many of the complex flow phenomena in colloidal and granular suspensions.

10
U
Besides a myriad of interparticle forces, the presence of tangential interparticle friction arising
AN
11 from either hydrodynamics or solid contacts of asperities is now thought to be responsible for

12 commonalities in shear thickening and jamming phenomena at high volume fractions and shear
M

13 stresses. The overall richness of the suspension mechanics landscape points to the re-unification
D

14 of colloidal and granular physics in the near future: one in which it may become possible to
TE

15 apply a universal set of physical frameworks to understand the flows of model rough particles

16 across multiple spatiotemporal scales. This can only be accomplished by properly distinguishing
EP

17 between microscopic and bulk friction, and by decoupling hydrodynamics and contact

18 contributions within the context of experimental observations.


C

19
AC

20 Keywords: Rough particles; colloidal suspensions; granular suspensions; friction; rheology


ACCEPTED MANUSCRIPT

21 1. Rheological significance of particle roughness

22 The flow of particulate suspensions plays an important role in a broad variety of

23 geophysical phenomena and engineering applications. These suspensions typically consist of

24 geometrically symmetric particles packed in a continuum fluid in the absence of attractive

PT
25 interactions, where the particle type may span colloids, grains, bubbles, and emulsions.

RI
26 Collective mesoscale rearrangements of the particles under applied stresses and often cause

27 enormous rheological changes in the bulk material, ranging from the sudden clogging of pipes

SC
28 [1], liquefaction and landscape evolution [2-4], to creative applications like robotic grippers [5]

29 and liquid body armor [6]. Despite the importance of suspension rheology and its investigation

30
U
since Reynolds and Einstein [7, 8], there is still a persistent gap between the behavior of
AN
31 industrially relevant particulate systems and the results obtained from academic model systems.

Flows are especially challenging to predict for dense suspensions (volume fraction φ ≥ 0.40) of
M

32

33 colloidal (typically with particle diameters 2a ≤ 2 µm) and granular (typically 2a > 2 µm)
D

34 particles. This is because textbook treatments for low Reynolds number suspension flows are
TE

35 traditionally developed through three simplifications [9-11]: (1) particles are perfectly spherical

36 in shape; (2) interparticle collisions are frictionless and overdamped in the case of colloids or
EP

37 inelastic in the case of larger particles; (3) solvent molecules are much smaller than the particle

38 size, such that continuum approximations can be used to model fluid drag between idealized
C

39 spherical particles. These assumptions have made theoretical developments from the Navier-
AC

40 Stokes equations tractable and reduced computational demands, but have also resulted in a major

41 discrepancy between experimental observations and predictions. A notable example is found in

42 many recent investigations of discontinuous thickening and shear jamming suspensions [12-20],

43 in which rough particles generated jumps in energy dissipation at reduced values of φ and shear

1
ACCEPTED MANUSCRIPT

44 stresses σ when compared to smooth, spherical particles [13, 15, 21, 22]. The prevailing thought

45 is that contact mechanics become important as lubrication films break down at large σ [23-25],

46 although there is a severe lack of in situ experimental evidence to directly support this statement.

47 The rheology of particulate suspensions was historically investigated by a combination of

PT
48 fluid mechanicians and granular physicists [26-28]. Although the two fields diverged in the

RI
49 1950s, they are now beginning to reconvene due to the need to consider both solid and fluid

50 mechanics in dense suspension flows. The convergence of these two fields are found in a number

SC
51 of reviews on granular physics and suspension mechanics [29-32]. In addition, we recommend a

52 comprehensive review by Morris on the computer simulations of lubricated-to-frictional shear

53
U
thickening as parallel reading material [33], which will prove useful as we discuss the
AN
54 experimental results here in light of theoretical findings.

55 Our review article summarizes recent experimental methods that are used to break new
M

56 ground in suspension rheology. First, we list a number of academic and industrial particulate
D

57 materials in which the surface roughness can be precisely controlled and quantified. Second, we
TE

58 describe existing experimental parameters used to characterize the frictional properties of various

59 particulates based on their surface morphologies. Finally, we describe the effects of surface
EP

60 anisotropy on macroscopic rheological properties as seen in dense suspensions of rough or

61 frictional particles, with an emphasis on how interparticle friction impacts their microstructure
C

62 and mechanics. The conclusion provides an outlook on the field of dense suspension rheology
AC

63 based on past work, present observations, and future strategies.

64

65 2. Preparation of smooth spherical particles

2
ACCEPTED MANUSCRIPT

66 A hard sphere particle is assumed to be undeformable and impenetrable, and interacts

67 with other particles solely through contact. In experimental systems, particles possess a finite

68 elastic modulus and can become deformed by strong flows [34]. The collisions between particles

69 are inelastic in the case of wet and dry granular materials where inertia dominates due to large

PT
70 particle sizes [35], or overdamped in the case of colloidal suspensions where viscous dissipation

RI
71 by the solvent is significant [36]. Perfectly smooth hard spheres have represented the ideal model

72 system for many years, allowing researchers to validate simulations and theories of suspension

SC
73 phase behavior and rheology [37-42]. They also provide a benchmarking tool for experimental

74 studies involving rough particles of similar sizes made from the same material. It is worth

75
U
remembering that many interparticle forces (electrostatics, solvophilicity, van der Waals,
AN
76 depletion, hydrogen bonding, etc.) are in play during the shear flow of particulate suspensions [9,

77 43], and that variations in synthesis techniques can produce particles with various types of
M

78 pairwise interactions that generate completely different rheological phenomena.


D

79 Currently, two common ways to generate such particles are through microfluidics and
TE

80 wet chemistry synthesis. Reviews of microfluidic and lithographic tools used to synthesize

81 particles are found elsewhere [44]. While these methods are capable of producing particles from
EP

82 ~101−102 µm with intricate surface anisotropy and nearly zero size polydispersity, they are

83 challenging to scale up to the sheer number of particles required for bulk rheology
C

84 measurements. As a point for comparison, it takes ≈1010 hard sphere particles (2a = 2 µm) to
AC

85 completely fill a small parallel plate rheometer geometry (diameter = 20 mm, gap height = 500

86 µm) with a suspension of φ = 0.50. Bulk chemical synthesis is therefore a much more viable

87 method for producing the large numbers of particles used in the investigation of dense

88 suspension rheology. Due to their well-characterized and highly tunable interaction potentials,

3
ACCEPTED MANUSCRIPT

89 sterically-stabilized silica, polystyrene (PS), and poly(methyl methacrylate) (PMMA) colloids

90 remain three of the most popular systems used in the study of suspension rheology. Each system

91 poses unique advantages and disadvantages. All three types of particles can be chemically or

92 physically tagged with conjugated fluorescent dyes for microscopy imaging.

PT
93

RI
94 2.1 Silica spheres

95 Monodisperse silica colloids are synthesized using the Stöber process [45, 46] in which

SC
96 the precursor, typically tetraethyolorthosilicate (TEOS), is hydrolyzed in alcohols and grown into

97 colloidal particles through a one-step sol-gel process. An octadecyl aliphatic chain is then grafted

98
U
to the bare surface of the silica particles through high temperature silanol esterification [47, 48].
AN
99 This method readily produces hard sphere particles with diameters between 20 nm [49] to 1000

100 nm [50]. If larger particles are desired, additional layers of silica or other materials like
M

101 polystyrene can be grown as shells on seed cores, in a method known as seeded growth
D

102 polymerization [51]. Depending on the solvent quality, the octadecyl grafted chains may undergo
TE

103 a lower critical solution temperature (LCST) crystalline transition that leads to thermoreversible

104 flocculation from a hard sphere suspension [49]. This tendency to flocculate at reduced
EP

105 temperatures leads to the term "adhesive hard spheres" for octadecyl-grafted silica colloids,

106 which are used in multiple gelation and self-assembly studies. Small angle neutron scattering
C

107 (SANS) is typically used to obtain the Baxter temperature, which quantifies the attraction
AC

108 strength through a square well potential [52-55]. A key benefit of silica colloids is that they do

109 not swell or plasticize in most solvents, which can impact measurements of φ as well as hard

110 sphere properties. The refractive index mismatch of silica (n = 1.459) with common solvents (n =

111 1.33 for water, n = 1.429 for tetradecane) is not too large, which does not significantly hinder

4
ACCEPTED MANUSCRIPT

112 their imaging resolution in confocal microscopy or introduce significant van der Waals forces.

113 However, since silica colloids have a high density (ρp = 1.7 − 2.0 g/ml) compared to that of most

114 polar and nonpolar solvents (ρf = 1 g/ml for water, ρf = 0.76 g/ml for tetradecane), the density

115 mismatch poses issues due to sedimentation and detachment from rheometer geometries. This

PT
116 issue is somewhat mitigated if the sedimentation velocities are reduced by decreasing the particle

RI
117 size or increasing the solvent viscosity.

118

SC
119

120 2.2 PMMA and PS spheres

121
U
Polymeric hard spheres form another class of model systems in studies of suspension
AN
122 rheology, with benefits and drawbacks that are almost completely opposite from that of silica

123 colloids. PS and PMMA colloids are generally prepared by emulsion polymerization [56], in
M

124 which conjugated polymer brushes such as poly(vinyl pyrrolidone) (PVP) [57], poly(dimethyl
D

125 siloxane) (PDMS) [58, 59], fluorinated copolymer blends [60], or poly(12-hydroxystearic acid)
TE

126 (PHSA) [61-65] are covalently attached to sterically stabilize a particle. Other stabilizers include

127 electrostatic groups that become charged in specific pH conditions, such as poly(acrylic acid)
EP

128 and aliphatic amines on polystyrene microspheres. The brushes can be grafted through a one-pot

129 synthesis as in free radical polymerization [61-65], with or without the addition of reversible
C

130 addition-fragmentation chain transfer (RAFT) agents [66], or they can be grafted post-synthesis
AC

131 through atom radical transfer polymerization (ATRP) [60]. Proper choice of the polymer brush

132 given the suspending fluid is key, as a fully solvated brush provides the largest range of steric

133 repulsion as compared to a collapsed brush. PHSA-grafted PMMA colloids in nonpolar solvents

134 are widely considered to be the model hard sphere system [36] and have been extensively used

5
ACCEPTED MANUSCRIPT

135 since the pioneering studies of Pusey and van Megen [37, 67, 68]. Direct and indirect measures

136 of the hard sphere properties of PHSA-PMMA colloids are widely available in the literature. The

137 standard PHSA brush length on these PMMA colloids is estimated to be ≈10 nm, although larger

138 chain lengths up to 22.4 nm are possible by varying polycondensation time [65]. The brush

PT
139 length was measured using a surface force apparatus (SFA) to obtain the interaction energy as a

RI
140 function of the surface separation for two flat mica surfaces coated with PHSA brushes [69].

141 The basic principle of emulsion polymerization reactions in the formation of polymer

SC
142 lattices is as follows: a monomer, such as styrene or methyl methacrylate, is dissolved in a

143 solvent mixture in which it is barely soluble. Heat-activated initiators such as potassium

144
U
persulfate (KPS) and azobisisobutyronitrile (AIBN), or ultraviolet light-activated photoinitiators
AN
145 such as hydroxymethylpropiophenones (Darocur), are triggered to release free radicals which

146 initiate and propagate the polymerization reaction. When the oligomers grow to a certain size,
M

147 they become insoluble in the solvent and phase separate out of the solution as nuclei for further
D

148 colloidal growth. A thorough review of the mechanisms involved in emulsion polymerization is
TE

149 given by Thickett and Gilbert [56]. The major benefits of using most polymeric colloids are that

150 benign solvents can be used for complete density and refractive index matching, and that the
EP

151 polymer brushes could potentially be functionalized to introduce stimuli-responsiveness into

152 particles [70]. Some disadvantages include problematic charge screening in nonpolar solvents
C

153 [71, 72], particle plasticization and swelling in organic solvents [73], and the added complexity
AC

154 that comes with the synthesis of polymer brushes such as PHSA.

155

156 2.3 Other polymeric materials

6
ACCEPTED MANUSCRIPT

157 A few other materials are used in the formation of spherical particles. Poly(N-

158 isopropylacrylamide) (PNIPAM) microgels have highly tunable Young's moduli depending on

159 the concentration of the added crosslinker (102 Pa ≤ E ≤ 104 Pa) and can be synthesized by one-

160 pot emulsion polymerization. They are used to understand the flow and self-assembly physics of

PT
161 soft, deformable particles above the random closed packing volume fraction of hard spheres (φrcp

RI
162 = 0.64) because they expand at temperatures below the LCST [74-76]. The applications of

163 PNIPAM are especially promising in biomedical engineering. Due to their softness, microgels

SC
164 with fibrin-protofibril binding motifs have been used as platelet-like particles to induce blood

165 clotting more rapidly under physiological flow conditions [77], and composite nanostructures

166 have been


U
added to PNIPAM to generate stimuli-responsive hydrogels that are highly
AN
167 stretchable [78]. Another highly versatile colloidal system is trimethoxysilyl propyl methacrylate
M

168 (TPM), which has been used to create polyhedral clusters [79], light-activated colloidal surfers

169 [80], colloidal alloys [81], and lock-and-key particles [82]. Other materials, such as fatty acid-
D

170 coated superparamagnetic iron oxide colloids, are prepared by alkali-mediated precipitation [83]
TE

171 and used in magnetic field-driven assembly studies [84, 85].

172 Because of the bottoms-up nature of these synthesis methods, in reality, even so-called
EP

173 smooth particles are never completely smooth at length scales close to that of the homopolymer

174 constituents. Smith et al. found that PHSA-PMMA colloids are slightly porous [86], with the
C

175 density of the PMMA cores having a slightly reduced density from the homopolymers. Silica and
AC

176 PS colloids are both subject to fluctuations in the particle porosity. This may shift the phase

177 behavior of hard sphere suspensions, which is a function of the osmotic pressure of the solvent. It

178 could also affect the attraction potential in colloidal gel investigations, because the interactions

179 are generated by excluded volumes of small depletant molecules.

7
ACCEPTED MANUSCRIPT

180

181 3. Preparation of rough particles

182 Rough particles were traditionally considered to be unsuitable as model systems.

183 Nevertheless, they are widely found in industrial formulations due to the use of milling as a

PT
184 common technique to grind up solids, for example in foods, paints, and coatings [87-89]. Recent

RI
185 developments in chemical and physical methods to synthesize bulk quantities of rough or bumpy

186 particles have made it possible to investigate the effects of roughness on suspension rheology

SC
187 (Fig. 1). We provide an overview of industrial and academic methods used to create

188 geometrically symmetric, yet surface anisotropic particles spanning the colloidal to granular

189
U
length scales. An in-depth review on the synthesis of porous polymeric particle is given by
AN
190 Gokmen and Du Prez [90]. Although there is a non-trivial relationship between roughness and

191 friction [91-93], in general, surface roughness increases the interparticle friction coefficient. In
M

192 this section, we discuss various physical and chemical routes to the formation of rough particles
D

193 in quantities large enough for bulk rheological characterization. The physical methods include
TE

194 milling and grinding processes, self-assembly of smaller particles on larger ones through

195 interparticle forces, in situ templating methods, and the external application of mechanical
EP

196 stresses. The chemical methods include the seeded growth of small particles on larger cores, acid

197 or base etching, linker chemistry and charge compensation, and the addition of crosslinkers
C

198 during emulsion polymerization. The advent of 3D printing has also made it possible to create
AC

199 granular particles with highly complex geometries.

8
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
C EP
AC

200

201 Figure 1. Surface morphologies accessible for geometrically symmetric particles and their

202 synthesis methods. Most of these micron-sized colloidal particles are synthesized using wet

203 chemistry methods [13, 28, 94-119], although 3D printing is also able to fabricate large

204 quantities of granular particles with arbitrary shapes [120].

9
ACCEPTED MANUSCRIPT

205 3.1. Grinding and milling

206 A mill applies kinetic energy to solid materials in order to break them up via friction and

207 attrition. Grains formed by milling are typically very polydisperse in their size distributions (p.d.

208 >> 10%) and may contain sharp and irregular facets. In fact, it is not unusual to obtain

PT
209 polydispersity values ranging from 100% to 300%. A few papers on the effect of breakage

RI
210 mechanisms on particulate sizes are available [121, 122], but ultimately it is an engineering

211 process in which the large number of process parameters makes predictive capabilities difficult.

SC
212 Nevertheless, because a mill is easy to use and can handle large quantities of wet or dry material,

213 milling remains one of the most common manufacturing techniques to grind materials like

214
U
organic crystals in pharmaceuticals [123], calcite powders [124], nanocrystalline metals [89],
AN
215 pigments [88], and various other particulates down to the desired size range. Cornstarch, one of

216 the most common particulates used to study the physics of shear thickening [17, 125, 126], is
M

217 formed by the wet milling of corn kernels and is therefore highly subject to size polydispersity
D

218 and shape irregularities.


TE

219

220 3.2. Surface heterocoagulation and seeded growth polymerization


EP

221 Electrostatic forces are commonly leveraged to decorate large core particles with smaller,

222 oppositely charged particles, forming composite raspberry-like particles with a bumpy exterior.
C

223 This so-called heterocoagulation mechanism was first used by Ottewill et al. [94] in which
AC

224 negatively charged PS particles are coated with smaller, positively charged poly(butyl

225 methacrylate) (PBMA) particles at reaction temperatures above the glass transition temperature

226 of the PBMA. The authors proposed a simple theory to obtain bumpy particles by considering

227 the interfacial energy of the two polymers. The most important parameter is the ratio of

10
ACCEPTED MANUSCRIPT

228 interfacial energies, as found in the Young-Dupré equation, which should be kept at an

229 intermediate value to avoid complete wetting or dewetting. A mass balance can be used to

230 deduce the proper ratio of particle radii and particle numbers for hexagonal close packing of the

231 PS particles on the PBMA cores. Various electrostatic stabilization and energy minimization

PT
232 methods were used successfully by a number of research groups to fabricate a dizzying array of

RI
233 Pickering emulsions and colloidosomes [99], surface-modified polystyrene particles [102, 127],

234 and raspberry-like silica particles [128, 129]. Other types of particle interactions, such as

SC
235 hydrogen bonding and π-π bonding, can also result in the same type of raspberry-like

236 morphology [98, 104, 130]. Removal of the bumps by chemical etching is also possible if golf

237 ball-like morphologies are desired [113].


U
AN
238 A variant of heterocoagulation coupled with seeded growth polymerization can be used to

239 synthesize larger raspberry-like particles, in which a core polymeric particle is coated with a
M

240 solid shell [95, 100, 105, 116, 117]. A secondary coating step is used to grow the shell covalently
D

241 on top of the composite bumpy particle. This step is thought to provide improved mechanical
TE

242 stability to the asperities during shear, such that they do not detach as easily [118]. Although the

243 particle morphology obtained with heterocoagulation methods is desirable due to their ease of
EP

244 reproducibility in simulation studies, a major issue is that each chemical synthesis and cleaning

245 step reduces the yield of particles.


C

246
AC

247 3.3. Confinement templating

248 Raspberry-like particles with multiple bump functionalities are also obtainable through a

249 microstructural confinement templating method, in which a small number of large particles are

250 dispersed in a concentrated bath of smaller particles. This method was used by Gaulding et al. to

11
ACCEPTED MANUSCRIPT

251 deposit functionalized PNIPAM microparticles onto PS spheres, where the surface bumpiness

252 could be precisely tuned [106]. However, this method suffers from two drawbacks, in that the

253 yield of the composite particles is limited and that the unused PNIPAM microparticles would go

254 to waste if not recycled. A similar method was used to confine TPM droplets by co-

PT
255 sedimentation with a dense suspension of PS spheres. The deformed TPM droplets are then

RI
256 polymerized by heat into polyhedral shapes [107]. The yield of particles obtained through

257 templating is likely to be much lower than other types of bulk synthesis methods.

SC
258

259 3.4. In situ mechanical stresses

260
U
Other researchers leveraged methods to generate internal mechanical stresses within
AN
261 particles to create macroporous microstructures. Peterson et al. describe an internal templating

262 method in which the precipitation of low melting point salts together with silica nanoparticles
M

263 using high temperature aerosol spraying generates various types of macroporous silica colloids
D

264 [109]. Degassing of PS and PMMA spheres in a polyelectrolyte solvent was also used to create
TE

265 nanometer scale roughness on their surfaces due to changes in the charge density on the polymer

266 surface [110]. A similar gas producing mechanism was used to create raspberry-like protrusions
EP

267 on silica particles as they pass through a flow-focusing microfluidic channel [111]. The surface

268 morphology is thought to result from the addition of hydrochloric acid and sodium bicarbonate,
C

269 which participate in the Stöber sol-gel process and produce carbon dioxide gas at the silica-water
AC

270 interfacial subphase. Finally, PS dimers and triangles were made by swelling of PS seed particles

271 in the presence of a crosslinker [108]. The concentration gradient of the crosslinker is proposed

272 to mechanically control the directionality of the phase separations during the seeded growth

273 polymerization step. These methods are more likely able to produce the quantities of particles

12
ACCEPTED MANUSCRIPT

274 needed for rheological testing, although the maximum roughness achievable may be limited to

275 the nanometer or submicron range.

276

277 3.5. Crosslinker-aided polymerization

PT
278 The addition of crosslinking monomers during the early nucleation step of emulsion

RI
279 polymerization can be used to generate surface morphologies that range from slight dimpling to

280 that reminiscent of crumpled paper and golf balls [113, 131]. By increasing the concentration of

SC
281 the crosslinker (ethylene glycol methacrylate) up to 2 wt% of the monomer, our group has

282 fabricated sterically stabilized PHSA-PMMA colloids that have root-mean-squared (RMS)

283
U
roughness values up to 20% of the mean particle radius [13]. Increasing the crosslinker
AN
284 concentration further tends to result in gelation of the entire polymer network, causing a failed

285 synthesis. This wet chemistry method is likely to be applicable for any polymeric materials, such
M

286 as PS and PNIPAM particles, that are formed by one-pot free radical polymerization reactions.
D

287 The size and polydispersity of the particles can be tuned somewhat independently of the
TE

288 roughness, although care should be taken such that the elastic modulus of the particles does not

289 increase significantly due to the addition of the crosslinker [132]. The thermodynamic
EP

290 mechanism for the formation of the rough features is currently unknown, primarily due to the

291 innate complexity of emulsion polymerization. We speculate that changes to the oligomer
C

292 solubility and nuclei shape during microphase separation could be possible reasons for the
AC

293 formation of rough particles. One of the biggest benefits of this method is that the entire particle

294 is made out of the same material, which simplifies the linking of particle interactions and

295 elasticity to bulk suspension mechanics.

296

13
ACCEPTED MANUSCRIPT

297 3.6. Other wet chemistries

298 If metallic oxide particles with high surface areas are desired, one-pot and template-free

299 hydrothermal synthesis can be used to prepare copper oxide colloids coated with a layer of

300 cerium oxide. The particle morphology resembles the type created by internal mechanical

PT
301 stresses, and the available surface area for catalysis can be dramatically increased by up to five

RI
302 times using this technique [119]. Hydrolytic polycondensation of poly(methyl silsesquioxane)

303 with inorganic calcium carbonate particles produced composite and roughened colloids with

SC
304 increased hydrophobicity [114]. Photoresponsive raspberry-like colloids were synthesized by

305 covalently attaching iron oxide nanoparticles to silica cores using curcubit[8]uril and azobenzene

306
U
linkers, which generated reversible changes in particle morphology from rough to smooth upon
AN
307 the illumination of ultraviolet light [115]. The authors show that the particle shape can be

308 controlled to switch between shear thinning and shear thickening behavior. These examples
M

309 illustrate the large variety of wet chemistries that can be used to tailor particle shapes in an
D

310 equally large variety of particulate materials.


TE

311

312 3.7. Etching and 3D printing


EP

313 Chemical etching is a simple method to generate rough surfaces on almost all types of

314 materials – whether they are inorganic or organic in nature. By immersing particles in a strong
C

315 acid or base for a sufficiently long period of time, the solvent will etch away parts of the surface,
AC

316 leaving behind a slightly roughened exterior. Silica and soda-lime glass beads are etched using

317 sodium hydroxide [21] and salt derivatives of hydrofluoric acid [133]. The RMS roughness can

318 be tuned by controlling the immersion time of the particles [133], although the roughness

14
ACCEPTED MANUSCRIPT

319 achievable may be limited to the nanometer range. On the other hand, it is possible for metallic

320 particles with high purity to become smoother after etching.

321 Finally, 3D printing has emerged as a powerful way to fabricate reusable granular

322 building blocks with highly complex geometries. Engineering software such as AutoCAD are

PT
323 typically used to design the particle shape. Many types of 3D shapes, such as polyhedra and

RI
324 stars, can be generated with polymer resins that have different elastic moduli values [120, 134].

325 Acrylic, nylon, ceramics, and hydrogel particles can all be printed this way. The particle sizes

SC
326 and shape details are limited by the resolution of current 3D printers (≈ 100 µm). A disadvantage

327 with 3D printing is that particles with very thin, interlocking features may break easily during

328
U
packing or shear due to the low fracture strength of commercial resins.
AN
329

330 4. Quantifying roughness and friction


M

331 The relationships between surface roughness, interparticle friction, and macroscopic
D

332 suspension mechanics are nonlinear. In the field of tribology, it is understood that surface
TE

333 roughness can completely change the frictional dissipation between two surfaces in near contact.

334 The friction between two surfaces is dictated by the contact area, material elasticity, shearing
EP

335 velocities, as well as the presence of lubricant fluid between the surfaces [91, 135]. Friction may

336 arise from either solid-solid contact or from hydrodynamic drag of the lubricant at higher sliding
C

337 speeds. The generic friction coefficient, µ, is an adjustable parameter used in many particulate
AC

338 simulations and theories. Because of its importance, a significant amount of effort has been

339 dedicated to measuring µ for various types of particles. No matter how smooth a particle might

340 appear, a non-zero static friction coefficient is always present in experimental observations [32].

341 Here, we review the methods used to obtain surface roughness parameters, as well as the

15
ACCEPTED MANUSCRIPT

342 interparticle friction coefficient (µp) and bulk stress ratio (µb) in particulate systems. The two

343 parameters represent physics from distinctly different length scales and we suggest exercising

344 great caution in distinguishing them.

345 Table 1 is a list of µp a measured using different methods for various types of particulate

PT
346 materials in the literature, although the list is by no means exhaustive. Roughness parameters and

RI
347 particle sizes are provided where available. This table is meant to provide a general reference to

348 the values of µ, which can differ greatly between static and sliding conditions and when additives

SC
349 are present in wet systems.

350

351
U
Table 1. List of particulate materials with various roughness and friction coefficients
AN
Material Method Reference 2aeff µp Rq
Acrylic Angle of repose [136] ≈ 0.3 cm 0.88 0.7 ± 0.3 µm
M

Angle of repose [136] ≈ 0.3 cm 0.96 2.6 ± 0.1 µm


Alumina LFM [137] 6 and 60 µm 0.03−0.07 189 nm
Aluminum Angle of repose [136] ≈ 0.3 cm 0.62 0.32 ± 0.14 µm
D

Cellulose LFM [138] 16−32 µm 0.22−0.64 13 nm



LFM [139] 15 µm 0.05
TE


LFM [139] 25 µm 0.35
Cornstarch LFM [140] 8 µm 0.02 14 nm
Glass LFM [141] 9.78 µm 0.9 (Bare) 12.4 nm
EP

LFM [141] 9.78 µm 0.02 (Brushes) 12.4 nm


LFM [137] 6 and 39 µm 0.18−0.60 12−19 nm

Sliding [142] 70−110 µm 0.25−0.65

Angle of Repose [133] 140−240 µm 0.42−0.52
C


Angle of repose [143] 3.5 mm 0.20

AC

Angle of Repose [144] 0.5−10 mm 0−1



Sliding [145] 6 mm 0.13
Limestone LFM [137] 4 µm 0.67 102 nm

Ottawa sand Cyclic shear [146] 0.40 mm 0.40

Pea gravel Cyclic shear [146] 9.8 mm 0.42
PMMA Simulations [13] 1.6−2.3 µm 0 20 nm
Simulations [13] 1.9−2.8 µm 0.30−1.00 50−110 nm

PS Angle of repose [147] 540 nm 0−0.11
PTFE Angle of repose [136] ≈ 0.3 cm 0.54 1.1 ± 0.6 µm

16
ACCEPTED MANUSCRIPT

PVC LFM [140] 1 µm 0.45 2.2 nm



Salmeterol Angle of repose [148] 15−36 µm 0.25−0.87

Silica LFM [15] 550−700 nm 0.02−0.12

LFM [149] 1.7−2.2 µm 0.34−1.04

LFM [150] 2.5 µm 0.08 (Silanated)

LFM [150] 2.5 µm 0.39 (Bare)

PT
Steel Angle of repose [136] ≈ 0.3 cm 0.66 0.1 ± 0.02 µm
Titania LFM [137] 0.2 µm 0.04−1.5 131−147 nm
Zeolite LFM [137] 2 µm 0.69 148 nm

RI
352 Data unavailable from reference.

353

SC
354 4.1. Atomic force microscopy and roughness parameters

355 Atomic force microscopy (AFM) is the leading experimental technique to measure

U
356 surface roughness at nanometer resolutions. It involves the use of a cantilever and tip with
AN
357 known spring constants to directly probe the topography of a surface in wet or dry conditions. As

358 an extension to the surface force apparatus [69, 151], AFM is also be used to obtain the pairwise
M

359 interactions of surfaces and particles through approach-retraction measurements [152, 153]. In
D

360 this section, we focus on the use of AFM for topographical characterization of surface roughness.
TE

361 Contact mode involves moving the AFM tip up and down as the instrument scans the surface,

362 but is typically avoided because the presence of liquids and adhesive interactions between the tip
EP

363 and the surface may erroneously influence topography measurements. A more convenient way to

364 avoid the pitfalls encountered in contact mode is to use tapping mode, which involves oscillating
C

365 the tip up and down at its resonance frequency using piezoelectric elements in the cantilever
AC

366 holder [154]. To obtain 3D surface topography measurements for particles, the particles must be

367 fixed such that the AFM tip does not move them around during imaging. This is accomplished

368 by partially embedding particles in special epoxies that soften when heated [129], or by spin

369 coating dilute suspensions onto a layer of porous polymer filter media adhered to a flat substrate

370 [13]. The surface topography measurements are limited to a relatively small field of view at the

17
ACCEPTED MANUSCRIPT

371 top most part of a particle, because imaging artifacts arise from dragging the AFM tip close to

372 the vertical sides of micron-sized or larger particles. Ensemble averaging of the topography

373 across multiple particles is required to generate enough statistics for quantification, especially in

374 the case of highly anisotropic particles. In addition to AFM, other optical methods such as white

PT
375 light scanning interferometry [133] are reported to generate 3D topography data of particles.

RI
376 Confocal laser scanning microscopy is also a possible method, but the difference in resolution in

377 the horizontal and vertical planes should be accounted for [155].

SC
378 Before roughness parameters can be extracted, the curved surface profiles of spherical

379 particles must be flattened and compared to a reference surface. This is best done by first fitting

380
U
an ideal sphere to the 3D topography data, centering the sphere location based on available
AN
381 height information, and minimizing the deviation between the two surfaces (Fig. 2) [156]. This

382 procedure effectively flattens the curved surface for further analysis. The diameter of the fitted
M

383 sphere, 2aeff, should be close to the values obtained from independent measurements of the
D

384 particle size, such as from scanning electron microscopy. Experimental measures of the
TE

385 roughness can then be obtained through the discretized form of the two-dimensional surface

386 roughness autocorrelation function [93, 157-159]:


EP

( ) ( )
Ly Lx

R ( x, y ) ≈
1
387
Nx N y
∑ ∑ z xi , y j ⋅ z xi + ∆x, y j + ∆y (1)
j=(− Ly ) i=(− Lx )
C

388 where Nx and Ny are the number of AFM data points in the x and y directions, Lx and Ly are the
AC

389 lengths of the x and y directions, ∆x and ∆y are the pixel sizes, and z(xi, yi) is the deviatory

390 surface height at a position (xi, yi). It is also possible to compute a 1D form of the function in

391 each direction and simply average them if Nx = Ny. From this function, a variety of surface

392 statistical parameters such as the spatial distribution of heights, the mean surface gradient, and

393 the RMS roughness can be obtained. All of these parameters are known to affect µp, with the

18
ACCEPTED MANUSCRIPT

394 RMS roughness Rq = (R2)1/2 being a popular parameter used in the engineering literature [160].

395 To account for different particle sizes, the value of Rq can be normalized by the particle radius.

396 We emphasize that even nominally smooth particles will have some level of roughness at the

397 nanoscale; for example, Rq was found to be 0.01% for smooth silica colloids [21] and 2.6% for

PT
398 smooth PMMA colloids [13].

RI
A

SC
2 µm 1 µm

U
B
AN
M

500 nm 500 nm

C
2 2
z (µm)

z (µm)
D

aeff
TE

0 0
-2 -1 0 1 2 -2 -1 0 1 2
x (µm) x (µm)
399
EP

400 Figure 2. Measurements and quantification of surface roughness for particles. Atomic force

401 microscopy (AFM) is one of the most viable methods of probing surface roughness, as long as
C

402 the particles are fixed on a flat substrate. The surface morphology of (A) sterically-stabilized
AC

403 PMMA colloids [13] and (B) raspberry-like silica colloids [15] have been successfully quantified

404 using AFM. (C) The root-mean-squared roughness values for smooth (left) and rough (right)

405 PMMA colloids are calculated by fitting an effective sphere of radius aeff to the raw data, then

406 minimizing the deviation (red arrows) between the measured profiles and the fitted sphere.

19
ACCEPTED MANUSCRIPT

407 Because the AFM tip is likely to produce imaging artifacts near the edge of the particles, regions

408 below a certain height z (indicated by grey areas) should not be considered for analysis.

409

410 4.2. Lateral force microscopy

PT
411 Lateral force microscopy (LFM) is a specialized operating mode of AFM, in which the

RI
412 AFM tip is dragged horizontally across a substrate at a fixed normal force (FN) and at finite

413 sliding speeds [139, 141, 149, 150, 161, 162]. The horizontal deflection of the cantilever is used

SC
414 to generate the frictional shear force (FS) according to Hooke's Law. The attachment of a

415 colloidal particle onto the AFM tip is termed colloidal probe microscopy. Many researchers have

416
U
used LFM with colloidal particles to either measure the value of µp for a particle sliding on a
AN
417 surface (which may or may not be coated with particles), or to obtain µp for two non-rotational

418 particles. Again, we emphasize that even though LFM provides a measure of the frictional
M

419 dissipation between two particles, it is not representative of the multi-body and multiscale
D

420 physics found in flowing particulate suspensions.


TE

421

422 4.3. Angle of repose


EP

423 Consider building a sand pile on a beach: continue to pour dry sand in one spot and a pile

424 of sand with a constant angle of repose, θ, forms. Adding water to the interstitial spaces reduces
C

425 θ, while the use of coarser grains might increase θ. The angle of repose for a particulate material
AC

426 is defined as the steepest angle to which the granular pile can be built without failure. Two types

427 of experimental setups are used to obtain the angle of repose: the first involves a quasi-2D

428 rotating drum, which spins at a fixed speed and provides the dynamic angle of repose for a

429 particulate material [32, 163, 164]; and the second involves a funnel setup where grains are

20
ACCEPTED MANUSCRIPT

430 continuously built into a pile [136, 147, 148]. The stress ratio can be estimated from the angle of

431 repose through Coulomb's criterion, µb = FS FN = tan θ , for a solid block sliding down an

432 inclined plane. This ratio is commonly understood as the macroscopic friction coefficient of a

PT
433 material [144]. More sophisticated theoretical treatments of the angle of repose involve using a

434 granular temperature to explain the local rearrangement events of grains [165, 166]. This so-

RI
435 called shear transformation zone (STZ) theory applies statistical mechanics principles [167],

436 reminiscent of the type found in soft glassy rheology [168, 169], was able to explain

SC
437 experimental measurements and computational simulations of granular flows down an inclined

U
438 plane [170, 171]. AN
439

440 4.4. Rheometry and cyclic shear cells


M

441 A rotational rheometer is a standard experimental tool used to quantify the relationship

442 between the bulk deformation and bulk stresses of particulate suspensions. When the normal
D

443 force is fixed and the shear force is measured, a measure of the bulk friction coefficient µb can be
TE

444 obtained with various sliding speeds and lubricating solvents. Comprehensive reviews of

445 rheometric methods are found in many textbooks [9, 11, 172]. The constant volume cyclic shear
EP

446 cells used in granular and soil mechanics testing is simply a larger version of the rheometer,

447 capable of plane shearing a packed bed of grains that have diameters on the order of millimeters
C

448 [145, 146, 173]. These instruments belong to a general class of rheometry techniques that
AC

449 provide bulk material properties that are useful in the modeling industrial and geophysical

450 rheological phenomena. Simple shear and oscillatory shear are two operating modes that are

451 widely used in the understanding of suspension rheology. They simulate the type of flows

452 encountered in realistic scenarios, and provide useful dynamical information about the bulk

21
ACCEPTED MANUSCRIPT

453 material through macroscopic parameters like the yield stress, shear strength, viscoelastic

454 moduli, and relaxation spectra. Most instruments are either strain rate-controlled or shear stress-

455 controlled, with the cone-and-plate, parallel plate, and annular Couette cells being some of the

456 most common geometries. Recently, Boyer et al. designed a specialized porous annular cell to

PT
457 maintain a constant particle pressure Pp for dense granular suspensions [174].

RI
458

459 4.5. In situ force visualization

SC
460 The methods described in Sections 4.1 to 4.5 provide measures of µb and µp, but these two

461 parameters may not be representative of the multi-body interparticle friction coefficient when a

462
U
dense suspension is undergoing deformation. Currently, the best known method to directly
AN
463 visualize grain-scale forces in situ is through the use of photoelastic disks made out of

464 commercially available birefringent polymers. A detailed review of photoelastic force


M

465 measurements in granular packings is provided by Daniels et al. [175] Briefly, this technique
D

466 was pioneered by the Behringer group to display contact force networks created by a dense
TE

467 packing of 2D disks in a biaxial shear cell [176, 177]. Non-circular shapes are also possible.

468 When each photoelastic disk is subjected to external normal stresses from its neighbors, local
EP

469 regions of the material rotates the polarization of light in accordance with the stress-optic

470 coefficient of the polymer. This change in the polarization angle is observable as bright and dark
C

471 fringe patterns when viewed with a circular polarizer. Photoelastic force measurements were
AC

472 used to characterize both the spatial and temporal distributions of the microscopic force network

473 for various 2D granular packings under linear and nonlinear impact [178]. Nevertheless, the

474 photoelastic quantification of force chains in 3D packings has not been accomplished due to

22
ACCEPTED MANUSCRIPT

475 challenges in analyzing the polarization angle and in obtaining the solution to the photoelasticity

476 problem.

477

478 5. Rheological phenomena

PT
479 A reason for the recent surge of interest in synthesizing rough particles is the attribution

RI
480 of shear thickening and jamming to interparticle friction. These flow scenarios are associated

481 with large increases in the dissipative energy of a dense suspension under applied shear stresses

SC
482 [12, 179-184]. In dense suspensions of colloidal and non-Brownian hard spheres, when φ and σ

483 are both sufficiently large, there is a gradual transition from Newtonian flow to mild continuous

484
U
shear thickening (CST), to discontinuous shear thickening (DST) where σ jumps discontinuously
AN
485 as a function of [185-187]. Both experiments and simulations found that the suspension
M

(φ − φ )
−α
486 viscosity for hard spheres diverges as ηs max
, where φmax ≈ 0.58−0.60 and α ~ 2.2−2.6

for colloidal glasses [188-190] and φmax ≈ 0.59−0.64 and α ~ 2.0 for athermal suspensions [174,
D

487

488 191]. Other particle shapes, such as fibrous rods and ellipsoids with high aspect ratios, are known
TE

489 to decrease the value of φmax even further [192, 193].


EP

490 Colloidal suspensions in fluids behave in an overdamped fashion due to viscous

491 dissipation from the solvent, while inertia becomes important for granular flows involving very
C

492 large particles. A few dimensionless numbers are useful in conceptualizing the relative
AC

493 importance of inertial and viscous forces, with subtle differences between each number. The

494 Stokes number, provides a measure of the inertial and viscous forces

495 experienced by a particle, while the Reynolds number describes the

496 competing forces experienced by the fluid around the particle. These two dimensionless numbers

23
ACCEPTED MANUSCRIPT

497 are different from the conventional Reynolds number for a continuum fluid in say, a parallel

498 plate rheometer geometry, , where H is the gap height and R is the radius

499 of the geometry. It is well known in the fluid mechanics community that pure fluids and dilute

PT
500 suspensions undergo inertial or turbulent flows when Ref >> 1. These types of inertial flows

501 exhibit a shear stress to shear rate scaling of , and should be differentiated from the

RI
502 Bagnoldian scaling ( ) in dense granular flows, which is observed even at low Rep values

SC
503 depending on φ [194].

504

U
505 5.1. The onset of shear thickening and dilatancy
AN
506 At extremely large values of φ and σ, both colloidal and granular suspensions may

507 experience an expansion in volume and dilate against its confining boundaries [194], or even
M

508 stop flowing altogether [18, 24, 195, 196]. Although dilatancy is thought to be related to shear
D

509 thickening [197-199], recent data on silica and PMMA colloids show that their onset stresses and
TE

510 volume fractions do not always coincide [13, 21, 200]. Further investigation is necessary to

511 comprehensively probe the role of surface roughness on the onset conditions of dilatancy and
EP

512 shear thickening.

513 The critical stress of thickening, σc, scales inversely as the particle radius squared (σcaeff2
C

514 ~ 1) for spherical PHSA-PMMA and silica colloids with or without electrostatic repulsion [68,
AC

515 201-203]. This scaling can be interpreted as a force balance between two particles at "contact",

516 where the hydrodynamic forces acting on a particle pair is equivalent to the derivative of the

517 interparticle potential that prevents them from overlapping. If other types of interparticle forces

518 are present, then the onset stress may scale in a different way [202, 204], or even become

519 obscured by a yield stress in the case of attractive interactions [193]. Stokesian Dynamics (SD)

24
ACCEPTED MANUSCRIPT

520 [205] and other simulations [34], in tandem with rheo-visualization [206-208] experiments,

521 showed that short-range lubrication forces are important in generating compact microstructures

522 in flows [209, 210]. Brownian motion introduces a time scale into the onset of thickening [179,

523 211-213]. The conundrum was that squeeze flow for perfectly spherical particles could not

PT
524 explain the large viscosity increases in DST, nor the observations that dense suspensions

RI
525 generate a positive macroscopic normal force due to dilatancy [214]. We note that

526 experimentalists should exercise care in reading out normal stress differences directly from a

SC
527 standard rheometer, because the instrument assumes that surface tension effects are negligible on

528 the boundaries of the fluid-air interface. Dilation cause particles to protrude at the interface [194,

529
U
215], producing in a large decrease in the first normal stress difference, N1. The axial force, Fz,
AN
530 exerted on the rheometer geometry by the suspension is a more appropriate way of understanding

531 dilation going forward. If surface tension terms are negligible, then [172]
M

 σ − σ φφ   1 2
532 Fz = −π R 2  θθ  + σ + Patm 
= π R N1 (2)
D

rr
 2   2
TE

533 where R is the radius of the geometry, σθθ, σφφ, and σrr are the normal stresses of the fluid in the θ,

534 φ, and r directions, and Patm is the atmospheric pressure that holds the fluid boundaries in place.
EP

535 For the conventional N1 relation in equation (2) to apply, the instrument software assumes that

536 σrr is balanced by Patm - Psurf, where Psurf ~ γ/Rcurv is negligible due to the large radius of
C

537 curvature (Rcurv) of the interface. Here, γ is the surface tension of the liquid. When particles
AC

538 protrude from the interface due to dilatancy, then Psurf ~ γ/aeff generates a significant contribution

539 to σrr, making it difficult to define the first normal stress difference N1 as read out by the

540 rheometer. Particle imaging experiments have shown that it takes a finite amount of time for the

541 particles to protrude from the surface of the suspension [14] and to generate a large negative N1

25
ACCEPTED MANUSCRIPT

542 value. These considerations suggest that it is more appropriate to use Fz instead of N1 to

543 characterize a dilating suspension, and that confinement from boundaries play an important role

544 in measuring dilation.

545

PT
546 5.2. Hydrodynamics and granular friction

RI
547 The term "hydrodynamics" in suspension rheology refers to the stresses borne by the

548 fluid phase in which particles are suspended, which are strong functions of the particle

SC
549 concentration φ. Einstein famously derived the relation between the viscosity of a dilute

550 suspension and its volume fraction, η/ηf = 1 + 2.5φ, by considering the mechanical work done on

551
U
a continuum fluid by the presence of a rigid sphere. This was accomplished through a surface
AN
552 integral for the expanding sphere [8]. The Stokes-Einstein-Sutherland relations for dilute

553 suspensions (translational diffusivity DT = kBT/6πηfa, rotational diffusivity DR = kBT/8πηfa3) were


M

554 obtained by considering the Stokes drag around a sphere, the osmotic pressure exerted by a
D

555 particle, and Fick's first law of diffusion. As the volume fraction increases beyond φ ≥ 0.05, the
TE

556 Einstein relation for viscosity no longer holds because multi-body interactions cause the fluid

557 stresses to become a function of the sheared microstructure of the particles [209]. It took nearly
EP

558 70 years before an analytical form of the relative viscosity that includes higher order terms of φ2

559 was obtained [216, 217], but the Batchelor-Green relation is still unable to predict the viscosity
C

560 when φ > 0.20. In fact, most experimental studies still rely on empirical correlations and
AC

561 benchmarking against previous studies to verify the validity of their measured suspension

562 viscosities at φ ≥ 0.50 [13, 29, 68].

563 Stokesian Dynamics simulations developed by Brady and Bossis in the late 1980s created

564 a major inroad into suspension rheology by incorporating multibody short-range and long-range

26
ACCEPTED MANUSCRIPT

565 lubrication hydrodynamic forces (represented by the grand resistance and mobility tensors) to

566 predict the structure and dynamics of a sheared suspension up to φ ≈ 0.50 [205, 212, 213, 218-

567 220]. Lubrication hydrodynamics lead to negative N1 values in dense suspensions of hard spheres

568 [180] (up to φ ≈ 0.60) due to anisotropies in the shear-induced microstructure and the formation

PT
569 of hydroclusters [12, 221]. Although surface roughness was explored using lubrication

RI
570 hydrodynamics models in the early 2000s, at that time simulation results did not match that of

571 experiments in terms of the large increase in suspension stresses that characterize DST and

SC
572 dilatancy [222-226]. Promisingly, the inclusion of surface roughness or the adaptation of a

573 roughness-corrected tangential lubrication force below a cutoff distance in Stokesian Dynamics

574
U
simulations appear to generate results that agree well with experimental observations even at
AN
575 high ϕ [227].

576 An alternative mechanism is proposed by researchers working in granular physics, who


M

577 have considered the concept that solid friction is responsible for the flow behavior of dry
D

578 granular matter [228]. Specifically, it is known that dense granular flows exhibit an intermediate
TE

579 fluid regime where particles interact by inelastic collisions and contact friction [229-231], which

580 are quantified by the restitution coefficient and the friction coefficient. They also undergo flow-
EP

581 arrest (jamming) transitions, as illustrated in a series of state diagrams by Liu, Nagel, and

582 coworkers [232, 233]. Although this problem may appear simple at a glance, difficulties arise
C

583 from the lack of constitutive relations for granular flows. The reasons for these challenges are
AC

584 twofold: (1) the continuum approximation for small solvent molecules, as in complex fluids and

585 colloidal suspensions, cannot be applied to discretized granular particles [234]; (2) there must be

586 a macroscopic friction coefficient imposed on the suspension [32]. Because there are no internal

587 stress scales in perfectly frictionless and rigid spheres, the lack of a macroscopic frictional

27
ACCEPTED MANUSCRIPT

588 criterion means that they would not be able to form any solid-like structures such as a granular

589 pile.

590 The introduction of this macroscopic friction coefficient (µb) is the origin of the growing

591 interest in bridging suspension and granular mechanics. Cates and co-workers first proposed

PT
592 shearing concentrated colloids as a specific example of inducing fragility in soft matter, in which

RI
593 force chains are formed to support compressive load without plastic rearrangement [235]. This

594 concept was further developed to explain S-shaped flow curves [24, 236] and to derive local

SC
595 constitutive relations, in which granular ideas such as jamming and friction were introduced into

596 shear thickening [125, 237]. When incorporated into simulations, a frictional criterion generally

597
U
does well in predicting the long-range velocity correlations and force networks displayed by
AN
598 granular flows [143, 238-243]. In particular, the relationship between the bulk stress ratio and the

599 viscous number above the material yield stress, or so-called µ(I) rheology, is used to generate
M

600 predictive models for the nonlocal rheology of granular flows [244-246].
D

601 Currently, most researchers agree that both hydrodynamic and contact forces are
TE

602 important in sheared suspensions, with a joint effort directed at decoupling the fluid and

603 frictional contributions to the overall dissipation [34, 247-249].


EP

604

605 5.3. Bridging rheology at the macroscale


C

606 Recent efforts at bridging colloidal suspensions with granular matter have suggested bulk
AC

607 µ(I) rheology as a possible unifying framework. At the macroscopic level, dimensional analysis

608 on frictional granular flows generated the Bagnoldian relations in which the shear stress scales as

609 and the normal stress scales as , where f1 and f2 are functions

610 of packing density [250]. The stresses for a flowing colloidal suspension without frictional

28
ACCEPTED MANUSCRIPT

611 interactions scale as , with thermal motion generating a dynamic yield stress at low

612 [9, 43]. Boyer et al. used a constant pressure shear cell to develop a general constitutive

613 framework for millimeter-sized PS and PMMA granules, in which the suspension viscosity (ηs)

PT
614 and µb is measured as a function of the viscous number, , where is the applied

615 shear rate and η f is the solvent viscosity (Fig. 3A). The measured rheology of the granular

RI
616 suspensions obey the frictional framework of granular matter and the viscous framework of

SC
617 colloidal suspensions [174]. Confined pressure Brownian Dynamics simulations of hard sphere

618 colloidal suspensions by Wang and Brady showed a similar µ(I) scaling, although the yield point

619
U
is slightly lower [251] (Fig. 3B). Because hydrodynamic and frictional interactions were not
AN
620 imposed in the simulations, the authors showed that excluded volume alone in colloidal

621 suspensions can generate similar types of bulk rheological behavior as the kind seen with
M

622 frictional granular flows.


D

A B
TE
C EP
AC

623

624 Figure 3. Macroscopic µ(I) rheology show commonalities in dense granular and colloidal

625 suspensions. (A) Experiments by Boyer et al. [174] in a pressure-controlled Couette shear cell

626 showed that the bulk stress ratio plotted against the viscous number for a dense suspension of

627 millimeter-sized beads (φ = 0.565) collapse on a master curve. The solvent is a Newtonian fluid.

29
ACCEPTED MANUSCRIPT

628 (B) Brownian Dynamics simulations by Wang and Brady [251] using a pressure-controlled

629 simulation box, enabled by a compressible solvent, show that hard sphere colloidal suspensions

630 without friction or hydrodynamics exhibit qualitatively similar rheology.

631

PT
632 5.3. Bridging rheology at the microscale

RI
633 The pairwise friction coefficient µp offers a more intuitive way of thinking about

634 frictional effects on dense suspension rheology. Simulation studies explicitly impose Coulomb's

SC
635 friction criterion for the particle tangential and normal contact forces, by applying the relation

636 Fc,tan = µ p Fc,nor when neighboring particles make contact or overlap with one another [23, 185,

U
AN
637 252]. This contact load model (CLM) was developed as an attempt to overcome the breakdown

638 in continuum approximations as particles approach within nanometers of one another in strongly
M

639 sheared flows. Its success in capturing the viscosity increases in DST and a positive first normal

640 stress difference representative of dilatancy led to a subsequent influx of granular concepts into
D

641 suspension rheology, including theoretical predictions of S-shaped flow curves based on a
TE

642 phenomenological relation between the jamming volume fraction and the fraction of broken

643 ( )
lubrication films [24]: φ J = φµ f + φrcp 1− f . Constitutive models are beginning to become
EP

644 available [253]. Here, the jamming volume fraction refers to the value of φ at which ηs diverges,
C

645 in which φ = φrcp = 0.64 is the isostatic criterion of frictionless hard spheres (µp = 0) and φµ ≈
AC

646 0.54 is the random close packing of frictional particles (µp ≈ 1) [254-257]. This relation spawned

647 a flurry of other investigations, including a shear reversal study in which the authors claimed to

648 have measured contact force contributions to the suspension viscosity in CST [248] in

649 contradiction with previous studies. The idea is that contact forces go to zero immediately upon

650 shear reversal, while hydrodynamic forces stay constant. While the experimental data were in

30
ACCEPTED MANUSCRIPT

651 good agreement with CLM simulations, it is important not to over-analyze raw rheometric data

652 below an initial time of 0.5 seconds even with modern stress-controlled rheometers. Our lab's

653 calibration data showed that inertia of the motor combined with the suspension may cause

654 artifacts in stress measurements across all shear rates tested [172]. To truly validate the presence

PT
655 of solid contact friction in dense suspensions, in situ measurements of particle-level forces like

RI
656 that of the photoelastic disks [176], but in 3D, are necessary.

657 Unfortunately, there is currently no experimental way to measure the value of µp as a 3D

SC
658 suspension is undergoing flow, as discussed in Section 4.5. Furthermore, the friction between

659 two surfaces should not be quantified using an ensemble-averaged value. This is because of

660
U
evidence that show the tendency for DST and dense granular flows to be locally and dynamically
AN
661 inhomogeneous, punctuated by shear bands with jammed and shear thinning regions that change

662 over time [126, 142, 167, 258, 259]. Because the friction coefficient varies nonlinearly as a
M

663 function of the roughness, sliding speed, and separation gap between two surfaces [260-262],
D

664 simply substituting a singular value of µp into models will not capture the relevant
TE

665 micromechanics. An excellent simulation study by Fernandez et al. correctly accounted for the

666 Stribeck behavior between two particles, which captured the transition from Newtonian to shear
EP

667 thickening flows [16]. Experimental measurements of polymer-adsorbed quartz microparticles

668 supported the simulation data, although nanotribology friction tests were performed using flat
C

669 surfaces in this study.


AC

670 Recently, rheological studies quantified the shear thickening and dilatant properties of

671 colloids with different roughness parameters or frictional interactions [15, 22, 140, 200, 236,

672 263-266] (Fig. 4). Although the Isa group used LFM to obtain µp for bumpy silica colloids [15],

673 and our group used the match between simulations and experimental rheology to back-calculate

31
ACCEPTED MANUSCRIPT

674 µp [13], the take-home message from both studies are similar: frictional dissipation is enhanced

675 by the presence of submicron-sized surface roughness on particles, possibly due to particle

676 interlocking mechanisms that slow down the stress relaxation of the hydroclusters or force

677 networks that persist in shear thickening [267-269].

PT
Silica PMMA Simulations

RI
102
ηr/ηr,N

101

SC
100

10-1
400
N1 (Pa, kBT/a3)

U
200

0
AN
-200

-400

10-2 10-1 100 101 102 103


M

σ/σc

678
D

679 Figure 4. Experiments and simulations of shear thickening suspensions agree when
TE

680 interparticle tangential friction is considered. Dense suspensions of silica [200] and PMMA

681 [13] colloids exhibit larger increases in the measured viscosities as φ and σ increase, representing
EP

682 a transition from weak to strong shear thickening. The first normal stress differences transition

683 from negative to positive signs, reminiscent of granular dilatancy in which particles push against
C

684 their confining boundaries in order to maintain flowing states. Surface roughness shifts these
AC

685 transition points to lower values of φ and σ. When interparticle friction is explicitly entered into

686 the equations of motion in dissipative particle dynamics simulations [247], they capture a

687 qualitatively similar trend as the experiments.

688

689 6. Outlook

32
ACCEPTED MANUSCRIPT

690 This review summarizes the techniques used to synthesize particles with tunable

691 roughness, methods of quantifying their surface morphology, characterization of the pairwise

692 and bulk friction coefficient, as well as the implication of hydrodynamic and contact friction in

693 flows of dense particulate suspensions. Many of the sections are presented from an experimental

PT
694 point of view, with the exception of rheological phenomena in which connections between

RI
695 interparticle friction and bulk rheology are primarily drawn from theory and simulations.

696 Outstanding questions remain as to how surface roughness affects the interparticle potential

SC
697 [270-272], how softness of the potential affects shear properties [62, 273], how force networks

698 propagate suspension stresses in space and time [265] and how to measure them experimentally,

699
U
and how hydrodynamic and contact contributions to the flow forces can be separated [247, 248].
AN
700 Although many predictive models already exist, validation is only made possible if new ways of

701 measuring interparticle forces within flowing suspensions are developed. Parallel advances in
M

702 optical imaging and mechanochemistry may turn this pipe dream into a real possibility.
D

703 The richness of the suspension mechanics landscape points to the re-unification of
TE

704 colloidal and granular physics in the near future [58, 174, 274]. In this version of the future, we

705 envision that it would be acceptable to apply a universal set of frictional frameworks to
EP

706 understand the flows of colloids and soils across multiple spatiotemporal scales, and in which

707 rough particles will be considered model systems for frictional flows. However, care must be
C

708 taken to ensure that proposed models are truly representative of the physics of flowing
AC

709 suspensions in different regimes. A look at introductory tribology textbooks immediately

710 illustrates that a friction coefficient may arise from either hydrodynamic or contact origins, and

711 that the friction between two sliding surfaces depend non-trivially on lubricant properties and

712 sliding speeds [91]. Furthermore, a force chain of fixed length can bear the same amount of load

33
ACCEPTED MANUSCRIPT

713 regardless of whether the particles are held together by lubrication forces, electrostatic repulsion,

714 or attractive interactions [227]. The current pool of evidence suggests that particle

715 micromechanics contribute in a seemingly non-unique way to the bulk rheological phenomena.

716 Therefore, physicists and engineers must continue to work together within the broader

PT
717 framework of experimental observations if we are to truly understand and design the

RI
718 contributions of interparticle forces and friction in the complex flows of dense suspensions.

719

SC
720 Acknowledgements

721 This work is supported in part by North Carolina State University startup funds, the

722
U
National Science Foundation (CBET-1804462), and the American Chemical Society Petroleum
AN
723 Research Fund (grant # 59208-DNI9).
M
D
TE
C EP
AC

34
ACCEPTED MANUSCRIPT

724 References
725 [**1] C. Ness, R. Mari, and M. E. Cates, Science Advances 4, eaar3296 (2018).
726 Simulations of superimposed cross oscillatory flows on steady shear show that the frequency of
727 the cross flows can be adjusted to decrease frictional particle contacts and thus suspension
728 viscosity.
729 [2] M. Houssais, C. P. Ortiz, D. J. Durian, and D. J. Jerolmack, Nature Communications 6, 6527
730 (2015).

PT
731 [3] B. Ferdowsi, C. P. Ortiz, and D. J. Jerolmack, Proceedings of the National Academy of
732 Sciences 115, 4827 (2018).
733 [4] A. Athanasopoulos-Zekkos and R. Seed, Journal of Geotechnical and Geoenvironmental
734 Engineering 139, 1911 (2013).

RI
735 [5] E. Brown, N. Rodenberg, J. Amend, A. Mozeika, E. Steltz, M. R. Zakin, H. Lipson, and H. M.
736 Jaeger, Proceedings of the National Academy of Sciences 107, 18809 (2010).
737 [6] Y. S. Lee, E. D. Wetzel, and N. J. Wagner, Journal of Materials Science 38, 2825 (2003).

SC
738 [7] O. Reynolds, The London, Edinburgh, and Dublin Philosophical Magazine and Journal of
739 Science 20, 469 (1885).
740 [8] A. Einstein, Annalen der Physik 324, 371 (1906).
741 [9] J. Mewis and N. J. Wagner, Colloidal suspension rheology (Cambridge University Press,

U
742 2012).
743 [10] J. Happel and H. Brenner, Low Reynolds number hydrodynamics: with special applications to
744 particulate media (Springer Science & Business Media, 2012), Vol. 1.
AN
745 [11] R. G. Larson, The structure and rheology of complex fluids (Oxford University Press New
746 York, 1999).
747 [12] N. J. Wagner and J. F. Brady, Physics Today 62, 27 (2009).
748 [**13] L. C. Hsiao, S. Jamali, E. Glynos, P. F. Green, R. G. Larson, and M. J. Solomon, Physical
M

749 Review Letters 119, 158001 (2017).


750 Experiments of PHSA-PMMA colloids ranging from smooth to a variety of surface roughness
751 show transition from weak to strong thickening as well as a sign change in N1. The experiments
D

752 are in agreement with DPD simulations in which the particle friction coefficient is explicitly
753 adjusted.
TE

754 [14] E. Brown and H. M. Jaeger, Reports on Progress in Physics 77, 046602 (2014).
755 [**15] C.-P. Hsu, S. N. Ramakrishna, M. Zanini, N. D. Spencer, and L. Isa, Proceedings of the
756 National Academy of Sciences, 201801066 (2018).
757 Raspberry-like silica core-shell particles are used to demonstrate DST and a sign change in N1
EP

758 in dense suspensions. No CST was observed. The authors attached the colloids onto an LFM
759 and measured the pairwise friction between particles.
760 [**16] N. Fernandez et al., Physical Review Letters 111, 108301 (2013).
761 A Stribeck friction law, in which the pairwise friction coefficient changes nonlinearly as a
C

762 function of the sliding speed, solvent viscosity, and normal load, is imposed within simulations
763 to obtain shear thickening curves. Experiments of quartz particle with different polymer
AC

764 coatings, representing different boundary friction as measured with a nanotribometer,


765 qualitatively support the simulation results.
766 [17] Y. Madraki, S. Hormozi, G. Ovarlez, E. Guazzelli, and O. Pouliquen, Physical Review Fluids 2,
767 033301 (2017).
768 [18] D. Bi, J. Zhang, B. Chakraborty, and R. P. Behringer, Nature 480, 355 (2011).
769 [*19] N. M. James, E. Han, R. A. L. de la Cruz, J. Jureller, and H. M. Jaeger, Nature Materials 17,
770 965 (2018).
771 Interparticle hydrogen bonding, when adjusted experimentally using charge-stabilized polymer
772 colloids and cornstarch suspensions, elicit shear jamming in dense suspensions.
773 [20] D. Lootens, H. Van Damme, and P. Hébraud, Physical Review Letters 90, 178301 (2003).

35
ACCEPTED MANUSCRIPT

774 [21] D. Lootens, H. Van Damme, Y. Hémar, and P. Hébraud, Physical Review Letters 95, 268302
775 (2005).
776 [22] S. Gallier, E. Lemaire, F. Peters, and L. Lobry, Journal of Fluid Mechanics 757, 514 (2014).
777 [**23] R. Seto, R. Mari, J. F. Morris, and M. M. Denn, Physical Review Letters 111, 218301 (2013).
778 One of the first simulations that combine hydrodynamics and contact friction to recover DST
779 behavior in dense suspensions. Force chains are observed in sheared suspensions.
780 [**24] M. Wyart and M. Cates, Physical Review Letters 112, 098302 (2014).

PT
781 A phenomenological theory for the jamming fraction as a function of the fraction of lubrication
782 films is used to generate S-shaped flow curves and shear jamming state diagrams.
783 [25] R. Ball and J. Melrose, Advances in Colloid And Interface Science 59, 19 (1995).
784 [26] G. I. Taylor, Proceedings of the Royal Society A 138, 41 (1932).

RI
785 [27] M. Von Smoluchowski, Annalen der Physik 326, 756 (1906).
786 [28] P. W. Rowe, Proceedings of the Royal Society A 269, 500 (1962).
787 [29] J. J. Stickel and R. L. Powell, Annual Review of Fluid Mechanics 37, 129 (2005).

SC
788 [30] J. Mewis and N. J. Wagner, Journal of Non-Newtonian Fluid Mechanics 157, 147 (2009).
789 [31] J. F. Morris, Rheologica Acta 48, 909 (2009).
790 [32] Y. Forterre and O. Pouliquen, Annual Review of Fluid Mechanics 40, 1 (2008).
791 [**33] J. F. Morris, Physical Review Fluids 3, 110508 (2018).

U
792 A review paper summarizing recent simulation and theoretical work investigating the
793 lubrication-to-frictional transition in sheared dense suspensions.
794 [34] S. Jamali, A. Boromand, N. Wagner, and J. Maia, Journal of Rheology 59, 1377 (2015).
AN
795 [35] J. T. Jenkins and S. B. Savage, Journal of Fluid Mechanics 130, 187 (1983).
796 [36] C. P. Royall, W. C. Poon, and E. R. Weeks, Soft Matter 9, 17 (2013).
797 [37] P. N. Pusey and W. Van Megen, Nature 320, 340 (1986).
798 [38] G. Batchelor, Journal of Fluid Mechanics 83, 97 (1977).
M

799 [39] R. Lionberger and W. Russel, Journal of Rheology 38, 1885 (1994).
800 [40] S.-E. Phan, W. B. Russel, Z. Cheng, J. Zhu, P. M. Chaikin, J. H. Dunsmuir, and R. H. Ottewill,
801 Physical Review E 54, 6633 (1996).
D

802 [41] L. Silbert, J. Melrose, and R. Ball, Journal of Rheology 43, 673 (1999).
803 [42] W. Schaertl and H. Sillescu, Journal of Statistical Physics 77, 1007 (1994).
TE

804 [43] W. B. Russel, W. Russel, D. A. Saville, and W. R. Schowalter, Colloidal dispersions


805 (Cambridge University Press, 1991).
806 [44] D. Dendukuri and P. S. Doyle, Advanced Materials 21, 4071 (2009).
807 [45] W. Stöber, A. Fink, and E. Bohn, Journal of Colloid and Interface Science 26, 62 (1968).
EP

808 [46] Y. J. Wong, L. Zhu, W. S. Teo, Y. W. Tan, Y. Yang, C. Wang, and H. Chen, Journal of the
809 American Chemical Society 133, 11422 (2011).
810 [47] A. Van Helden, J. W. Jansen, and A. Vrij, Journal of Colloid and Interface Science 81, 354
811 (1981).
C

812 [48] A. Woutersen, R. May, and C. d. de Kruif, Journal of Colloid and Interface Science 151, 410
813 (1992).
AC

814 [49] A. P. Eberle, R. n. Castañeda-Priego, J. M. Kim, and N. J. Wagner, Langmuir 28, 1866 (2012).
815 [50] P. Varadan and M. J. Solomon, Journal of Rheology 47, 943 (2003).
816 [51] K. A. Whitaker and E. M. Furst, Langmuir 30, 584 (2013).
817 [52] A. P. Eberle, N. J. Wagner, and R. Castañeda-Priego, Physical Review Letters 106, 105704
818 (2011).
819 [53] J. M. Kim, J. Fang, A. P. Eberle, R. Castañeda-Priego, and N. J. Wagner, Physical Review
820 Letters 110, 208302 (2013).
821 [54] M. Chen and W. Russel, Journal of Colloid and Interface Science 141, 564 (1991).
822 [55] M. A. Miller and D. Frenkel, The Journal of Chemical Physics 121, 535 (2004).
823 [56] S. C. Thickett and R. G. Gilbert, Polymer 48, 6965 (2007).
824 [57] H. Pertoft, T. C. Laurent, T. Låås, and L. Kågedal, Analytical Biochemistry 88, 271 (1978).

36
ACCEPTED MANUSCRIPT

825 [58] L. C. Hsiao, R. S. Newman, S. C. Glotzer, and M. J. Solomon, Proceedings of the National
826 Academy of Sciences (2012).
827 [59] H. Tsurusawa, J. Russo, M. Leocmach, and H. Tanaka, Nature Materials 16, 1022 (2017).
828 [*60] T. E. Kodger, R. E. Guerra, and J. Sprakel, Scientific Reports 5, 14635 (2015).
829 The synthesis of fluorinated PMMA colloids with hard sphere behavior is provided in this
830 paper.
831 [61] L. Antl, J. Goodwin, R. Hill, R. H. Ottewill, S. Owens, S. Papworth, and J. Waters, Colloids

PT
832 and Surfaces 17, 67 (1986).
833 [62] G. Bryant, S. R. Williams, L. Qian, I. Snook, E. Perez, and F. Pincet, Physical Review E 66,
834 060501 (2002).
835 [63] M. T. Elsesser and A. D. Hollingsworth, Langmuir 26, 17989 (2010).

RI
836 [64] G. Bosma, C. Pathmamanoharan, E. H. de Hoog, W. K. Kegel, A. van Blaaderen, and H. N.
837 Lekkerkerker, Journal of Colloid and Interface Science 245, 292 (2002).
838 [65] L. Palangetic, K. Feldman, R. Schaller, R. Kalt, W. R. Caseri, and J. Vermant, Faraday

SC
839 Discussions 191, 325 (2016).
840 [66] U. C. Palmiero, A. Agostini, E. Lattuada, S. Gatti, J. Singh, C. T. Canova, S. Buzzaccaro, and
841 D. Moscatelli, Soft Matter 13, 6439 (2017).
842 [67] S. Auer, W. Poon, and D. Frenkel, Physical Review E 67, 020401 (2003).

U
843 [**68] B. Guy, M. Hermes, and W. C. Poon, Physical Review Letters 115, 088304 (2015).
844 Experimental data of shear thickening for PHSA-PMMA colloids of various sizes show a
845 transition in the critical onset stress, and an explanation is offered for this scaling in terms of
AN
846 the particle contact area during flow.
847 [69] B. d. L. Costello, P. Luckham, and T. F. Tadros, Langmuir 8, 464 (1992).
848 [70] M. J. Solomon, Langmuir 34, 11205 (2018).
849 [71] Y. Kim, A. A. Shah, and M. J. Solomon, Nature Communications 5, 3676 (2014).
M

850 [72] N. Park and J. C. Conrad, Soft Matter 13, 2781 (2017).
851 [73] W. C. K. Poon, E. R. Weeks, and C. P. Royall, Soft Matter 8, 21 (2012).
852 [74] G. M. Conley, P. Aebischer, S. Nöjd, P. Schurtenberger, and F. Scheffold, Science Advances 3,
D

853 e1700969 (2017).


854 [75] L. Mohan, R. T. Bonnecaze, and M. Cloitre, Physical Review Letters 111, 268301 (2013).
TE

855 [76] Y. Peng, F. Wang, Z. Wang, A. M. Alsayed, Z. Zhang, A. G. Yodh, and Y. Han, Nature
856 Materials 14, 101 (2015).
857 [77] A. C. Brown et al., Nature Materials 13, 1108 (2014).
858 [78] L.-W. Xia, R. Xie, X.-J. Ju, W. Wang, Q. Chen, and L.-Y. Chu, Nature Communications 4,
EP

859 2226 (2013).


860 [79] K. V. Edmond, M. T. Elsesser, G. L. Hunter, D. J. Pine, and E. R. Weeks, Proceedings of the
861 National Academy of Sciences, 201203328 (2012).
862 [80] J. Palacci, S. Sacanna, A. P. Steinberg, D. J. Pine, and P. M. Chaikin, Science 339, 936 (2013).
C

863 [81] É. Ducrot, M. He, G.-R. Yi, and D. J. Pine, Nature Materials 16, 652 (2017).
864 [82] S. Sacanna, W. T. Irvine, P. M. Chaikin, and D. J. Pine, Nature 464, 575 (2010).
AC

865 [83] D. Kim, Y. Zhang, W. Voit, K. Rao, and M. Muhammed, Journal of Magnetism and Magnetic
866 Materials 225, 30 (2001).
867 [84] B. Bharti, A.-L. Fameau, M. Rubinstein, and O. D. Velev, Nature Materials 14, 1104 (2015).
868 [85] J. W. Swan, J. L. Bauer, Y. Liu, and E. M. Furst, Soft Matter 10, 1102 (2014).
869 [86] G. N. Smith et al., Journal of Colloid and Interface Science 479, 234 (2016).
870 [87] J. E. Fannon, R. J. Hauber, and J. N. BeMiller, Cereal Chemstry 69, 284 (1992).
871 [88] E. Bilgili, R. Hamey, and B. Scarlett, Chemical Engineering Science 61, 149 (2006).
872 [89] H. Fecht, E. Hellstern, Z. Fu, and W. Johnson, Metallurgical Transactions A 21, 2333 (1990).
873 [90] M. T. Gokmen and F. E. Du Prez, Progress in Polymer Science 37, 365 (2012).
874 [91] G. Stachowiak and A. W. Batchelor, Engineering tribology (Butterworth-Heinemann, 2013).
875 [92] G. Yakubov, T. Branfield, J. Bongaerts, and J. Stokes, Biotribology 3, 1 (2015).

37
ACCEPTED MANUSCRIPT

876 [93] B. Persson, O. Albohr, U. Tartaglino, A. Volokitin, and E. Tosatti, Journal of Physics:
877 Condensed Matter 17, R1 (2004).
878 [94] R. Ottewill, A. Schofield, J. Waters, and N. S. J. Williams, Colloid and Polymer Science 275,
879 274 (1997).
880 [95] M. Yu, Q. Wang, M. Zhang, Q. Deng, and D. Chen, RSC Advances 7, 39471 (2017).
881 [96] S.-H. Kim, C.-J. Heo, S. Y. Lee, G.-R. Yi, and S.-M. Yang, Chemistry of Materials 19, 4751
882 (2007).

PT
883 [97] X. Liu and J. He, Journal of Colloid and Interface Science 314, 341 (2007).
884 [98] Y. Bao, Q. Li, P. Xue, J. Huang, J. Wang, W. Guo, and C. Wu, Materials Research Bulletin 46,
885 779 (2011).
886 [99] A. Dinsmore, M. F. Hsu, M. Nikolaides, M. Marquez, A. Bausch, and D. Weitz, Science 298,

RI
887 1006 (2002).
888 [100] G. Li, X. Yang, and J. Wang, Colloids and Surfaces A 322, 192 (2008).
889 [101] C. S. Wagner, S. Shehata, K. Henzler, J. Yuan, and A. Wittemann, Journal of Colloid and

SC
890 Interface Science 355, 115 (2011).
891 [102] T. Taniguchi, S. Obi, Y. Kamata, T. Kashiwakura, M. Kasuya, T. Ogawa, M. Kohri, and T.
892 Nakahira, Journal of Colloid and Interface Science 368, 107 (2012).
893 [103] Y. Li, Y. Pan, C. Yang, Y. Gao, Z. Wang, and G. Xue, Colloids and Surfaces A 414, 504

U
894 (2012).
895 [104] H. Minami, Y. Mizuta, and T. Suzuki, Langmuir 29, 554 (2012).
896 [105] Y. Liu, K. V. Edmond, A. Curran, C. Bryant, B. Peng, D. G. Aarts, S. Sacanna, and R. P.
AN
897 Dullens, Advanced Materials 28, 8001 (2016).
898 [106] J. C. Gaulding, S. Saxena, D. E. Montanari, and L. A. Lyon, ACS Macro Letters 2, 337 (2013).
899 [107] Y. Wang, J. T. McGinley, and J. C. Crocker, Langmuir 33, 3080 (2017).
900 [108] J. W. Kim, R. J. Larsen, and D. A. Weitz, Advanced Materials 19, 2005 (2007).
M

901 [109] A. K. Peterson, D. G. Morgan, and S. E. Skrabalak, Langmuir 26, 8804 (2010).
902 [110] B. Ilhan, C. Annink, D. Nguyen, F. Mugele, I. Sîretanu, and M. H. Duits, Colloids and Surfaces
903 A 560, 50 (2019).
D

904 [111] C.-X. Zhao and A. P. Middelberg, RSC Advances 3, 21227 (2013).
905 [112] S. K. Pattanayek and A. K. Ghosh, Colloid and Polymer Science 295, 1117 (2017).
TE

906 [113] D. Kim, K. Lee, and S. Choe, Macromolecular Research 17, 250 (2009).
907 [114] B. Li, X. Gao, J. Zhang, Y. Sun, and Z. Liu, Powder Technology 292, 80 (2016).
908 [115] C. Hu, J. Liu, Y. Wu, K. R. West, and O. A. Scherman, Small 14, 1703352 (2018).
909 [116] C. Graf, D. L. Vossen, A. Imhof, and A. van Blaaderen, Langmuir 19, 6693 (2003).
EP

910 [117] S. Shi, S. Kuroda, K. Hosoi, and H. Kubota, Polymer 46, 3567 (2005).
911 [118] M. D'Acunzi, L. Mammen, M. Singh, X. Deng, M. Roth, G. K. Auernhammer, H.-J. Butt, and
912 D. Vollmer, Faraday Discussions 146, 35 (2010).
913 [119] Z. Jin, J. Li, L. Shi, Y. Ji, Z. Zhong, and F. Su, Applied Surface Science 359, 120 (2015).
C

914 [**120] A. G. Athanassiadis et al., Soft Matter 10, 48 (2014).


915 3D printing was used to generate a range of granular particle shapes, which were placed in a
AC

916 customized triaxial testin machine to draw connections between the material elastic properties
917 and the confining pressure.
918 [121] Y. Altintaş and E. Budak, CIRP Annals-Manufacturing Technology 44, 357 (1995).
919 [122] D. Montgomery and Y. Altintas, Journal of Engineering for Industry 113, 160 (1991).
920 [123] S. P. Chamarthy and R. Pinal, Colloids and Surfaces A: Physicochemical and Engineering
921 Aspects 331, 68 (2008).
922 [124] E. Yoğurtcuoğlu and M. Uçurum, Powder Technology 214, 47 (2011).
923 [125] A. Fall, N. Huang, F. Bertrand, G. Ovarlez, and D. Bonn, Physical Review Letters 100, 018301
924 (2008).
925 [**126] B. Saint-Michel, T. Gibaud, and S. Manneville, arXiv preprint arXiv:1803.03558 (2018).

38
ACCEPTED MANUSCRIPT

926 Magnetic resonance imaging experiments with a Taylor-Couette cell show that shear bands
927 form in DST flows of cornstarch, which are thought to be responsible for stress fluctuations
928 reminiscent of turbulent flows.
929 [127] G. Li, X. Yang, F. Bai, and W. Huang, Journal of Colloid and Interface Science 297, 705
930 (2006).
931 [*128] M. Zanini, C.-P. Hsu, T. Magrini, E. Marini, and L. Isa, Colloids and Surfaces A 532, 116
932 (2017).

PT
933 The synthesis method for raspberry-like silica colloids for model rough particles is provided
934 here.
935 [*129] M. Zanini, C. Marschelke, S. E. Anachkov, E. Marini, A. Synytska, and L. Isa, Nature
936 Communications 8, 15701 (2017).

RI
937 The particles from ref. 126 are found to have interesting wetting and interfacial pinning
938 properties at oil/water and water/oil interfaces due to their surface roughness.
939 [130] D. Crespy, M. Stark, C. Hoffmann-Richter, U. Ziener, and K. Landfester, Macromolecules 40,

SC
940 3122 (2007).
941 [131] R. Rice, R. Roth, and C. P. Royall, Soft Matter 8, 1163 (2012).
942 [132] K. Tamareselvy and F. A. Rueggeberg, Dental Materials 10, 290 (1994).
943 [133] S. Utermann, P. Aurin, M. Benderoth, C. Fischer, and M. Schröter, Physical Review E 84,

U
944 031306 (2011).
945 [134] J. Zhao, M. Jiang, K. Soga, and S. Luding, Granular Matter 18, 59 (2016).
946 [135] A. Tiwari, L. Dorogin, A. Bennett, K. Schulze, W. Sawyer, M. Tahir, G. Heinrich, and B.
AN
947 Persson, Soft Matter 13, 3602 (2017).
948 [136] G. R. Farrell, K. M. Martini, and N. Menon, Soft Matter 6, 2925 (2010).
949 [137] R. Jones, H. M. Pollock, D. Geldart, and A. Verlinden-Luts, Ultramicroscopy 100, 59 (2004).
950 [138] A. A. Feiler, J. Stiernstedt, K. Theander, P. Jenkins, and M. W. Rutland, Langmuir 23, 517
M

951 (2007).
952 [139] S. Zauscher and D. J. Klingenberg, Colloids and Surfaces A 178, 213 (2001).
953 [140] J. Comtet, G. Chatté, A. Niguès, L. Bocquet, A. Siria, and A. Colin, Nature Communications 8,
D

954 15633 (2017).


955 [141] N. Fernandez, J. Cayer-Barrioz, L. Isa, and N. D. Spencer, Langmuir 31, 8809 (2015).
TE

956 [142] S. Nasuno, A. Kudrolli, A. Bak, and J. P. Gollub, Physical Review E 58, 2161 (1998).
957 [143] D. M. Mueth, H. M. Jaeger, and S. R. Nagel, Physical Review E 57, 3164 (1998).
958 [144] Y. Zhou, B. H. Xu, A.-B. Yu, and P. Zulli, Powder Technology 125, 45 (2002).
959 [145] J. Härtl and J. Y. Ooi, Granular Matter 10, 263 (2008).
EP

960 [146] J. F. Hubler, A. Athanasopoulos-Zekkos, and D. Zekkos, Journal of Geotechnical and


961 Geoenvironmental Engineering 143, 04017043 (2017).
962 [147] C. P. Ortiz, R. Riehn, and K. E. Daniels, Soft Matter 9, 543 (2013).
963 [148] F. Podczeck, J. Newton, and M. James, Journal of Materials Science 31, 2213 (1996).
C

964 [149] X. Ling, H.-J. Butt, and M. Kappl, Langmuir 23, 8392 (2007).
965 [150] S. Ecke and H.-J. Butt, Journal of Colloid and Interface Science 244, 432 (2001).
AC

966 [151] J. N. Israelachvili and D. Tabor, Proceedings of the Royal Society A 331, 19 (1972).
967 [152] L.-O. Heim, J. Blum, M. Preuss, and H.-J. Butt, Physical Review Letters 83, 3328 (1999).
968 [153] S. Biggs, Langmuir 11, 156 (1995).
969 [154] S. Morita, F. J. Giessibl, E. Meyer, and R. Wiesendanger, Noncontact atomic force microscopy
970 (Springer, 2015), Vol. 3.
971 [155] J. Pawley, Handbook of biological confocal microscopy (Springer Science & Business Media,
972 2010).
973 [156] J. M. Bennett, Applied Optics 15, 2705 (1976).
974 [157] R. S. Sayles and T. R. Thomas, Nature 271, 431 (1978).
975 [158] J. Greenwood, Proceedings of the Royal Society A 393, 133 (1984).

39
ACCEPTED MANUSCRIPT

976 [159] E. Sidick, in Modeling Aspects in Optical Metrology II (International Society for Optics and
977 Photonics, 2009), p. 73900L.
978 [160] A. J. McGhee et al., Tribology Letters 65, 157 (2017).
979 [161] H.-J. Butt, B. Cappella, and M. Kappl, Surface Science Reports 59, 1 (2005).
980 [162] S. Ecke, R. Raiteri, E. Bonaccurso, C. Reiner, H.-J. Deiseroth, and H.-J. Butt, Review of
981 Scientific Instruments 72, 4164 (2001).
982 [163] N. A. Pohlman, B. L. Severson, J. M. Ottino, and R. M. Lueptow, Physical Review E 73,

PT
983 031304 (2006).
984 [164] S. C. Du Pont, P. Gondret, B. Perrin, and M. Rabaud, Europhysics Letters 61, 492 (2003).
985 [165] A. Lemaître, Physical Review Letters 89, 195503 (2002).
986 [166] O. R. Walton and R. L. Braun, Journal of Rheology 30, 949 (1986).

RI
987 [167] D. Fenistein, J. W. van de Meent, and M. van Hecke, Physical Review Letters 92, 094301
988 (2004).
989 [168] P. Sollich, F. Lequeux, P. Hébraud, and M. E. Cates, Physical Review Letters 78, 2020 (1997).

SC
990 [169] G. Ovarlez, Q. Barral, and P. Coussot, Nature Materials 9, 115 (2010).
991 [170] L. E. Silbert, D. Ertaş, G. S. Grest, T. C. Halsey, D. Levine, and S. J. Plimpton, Physical
992 Review E 64, 051302 (2001).
993 [171] O. Pouliquen, Physics of Fluids 11, 542 (1999).

U
994 [172] C. W. Macosko, Rheology: principles, measurements, and applications (Wiley, New York NY,
995 1994).
996 [173] K. H. Zehtab, X. Fei, J. Hubler, A. Athanasopoulos-Zekkos, D. Zekkos, and W. A. Marr,
AN
997 Geotechnical Testing Journal 41 (2018).
998 [**174] F. Boyer, É. Guazzelli, and O. Pouliquen, Physical Review Letters 107, 188301 (2011).
999 The authors designed a Couette-cone shear cell that with a porous top plate, allowing solvent to
1000 flow freely in and out of a dense suspension of beads. They used this pressure-controlled device
M

1001 to obtain flow curves for the suspension that display both granular and colloidal rheology.
1002 [**175] K. E. Daniels, J. E. Kollmer, and J. G. Puckett, Review of Scientific Instruments 88, 051808
1003 (2017).
D

1004 A review of photoelastic 2D disks for force chain sensing is given here. Both experimental and
1005 analysis details are provided in great detail.
TE

1006
1007 [176] T. S. Majmudar and R. P. Behringer, Nature 435, 1079 (2005).
1008 [177] D. S. Bassett, E. T. Owens, M. A. Porter, M. L. Manning, and K. E. Daniels, Soft Matter 11,
1009 2731 (2015).
EP

1010 [178] L. Papadopoulos, M. A. Porter, K. E. Daniels, and D. S. Bassett, Journal of Complex Networks
1011 6, 485 (2018).
1012 [179] J. Bender and N. J. Wagner, Journal of Rheology 40, 899 (1996).
1013 [180] C. D. Cwalina and N. J. Wagner, Journal of Rheology 58, 949 (2014).
C

1014 [181] W. J. Frith, P. d’Haene, R. Buscall, and J. Mewis, Journal of Rheology 40, 531 (1996).
1015 [182] L.-N. Krishnamurthy, N. J. Wagner, and J. Mewis, Journal of Rheology 49, 1347 (2005).
AC

1016 [183] M. Lee, M. Alcoutlabi, J. Magda, C. Dibble, M. Solomon, X. Shi, and G. McKenna, Journal of
1017 Rheology 50, 293 (2006).
1018 [184] V. T. O'Brien and M. E. Mackay, Langmuir 16, 7931 (2000).
1019 [**185] R. Mari, R. Seto, J. F. Morris, and M. M. Denn, Journal of Rheology 58, 1693 (2014).
1020 A follow up to Mari et al. PRL (2013), in which details of the simulations and contact load
1021 model is described. The rate dependence of the interparticle forces is discussed in the
1022 simulation details.
1023 [186] R. Mari, R. Seto, J. F. Morris, and M. M. Denn, Physical Review E 91, 052302 (2015).
1024 [187] Y. Madraki, G. Ovarlez, and S. Hormozi, Physical Review Letters 121, 108001 (2018).
1025 [188] W. B. Russel, N. J. Wagner, and J. Mewis, Journal of Rheology 57, 1555 (2013).
1026 [189] A. Ikeda, L. Berthier, and P. Sollich, Physical Review Letters 109, 018301 (2012).

40
ACCEPTED MANUSCRIPT

1027 [190] G. Brambilla, D. El Masri, M. Pierno, L. Berthier, L. Cipelletti, G. Petekidis, and A. B.


1028 Schofield, Physical Review Letters 102, 085703 (2009).
1029 [191] A. Ikeda, L. Berthier, and P. Sollich, Soft Matter 9, 7669 (2013).
1030 [192] R. G. Egres, F. Nettesheim, and N. J. Wagner, Journal of Rheology 50, 685 (2006).
1031 [193] E. Brown, N. A. Forman, C. S. Orellana, H. Zhang, B. W. Maynor, D. E. Betts, J. M.
1032 DeSimone, and H. M. Jaeger, Nature materials 9, 220 (2010).
1033 [194] E. Brown and H. M. Jaeger, Journal of Rheology 56, 875 (2012).

PT
1034 [195] I. R. Peters, S. Majumdar, and H. M. Jaeger, Nature 532, 214 (2016).
1035 [196] S. R. Waitukaitis and H. M. Jaeger, Nature 487, 205 (2012).
1036 [197] H. Laun, Journal of Non-Newtonian Fluid Mechanics 54, 87 (1994).
1037 [198] M. I. Smith, R. Besseling, M. E. Cates, and V. Bertola, Nature Communications 1, 114 (2010).

RI
1038 [199] A. Metzner and M. Whitlock, Transactions of the Society of Rheology 2, 239 (1958).
1039 [200] J. R. Royer, D. L. Blair, and S. D. Hudson, Physical Review Letters 116, 188301 (2016).
1040 [201] B. J. Maranzano and N. J. Wagner, The Journal of Chemical Physics 114, 10514 (2001).

SC
1041 [202] R. Mari, R. Seto, J. F. Morris, and M. M. Denn, Proceedings of the National Academy of
1042 Sciences 112, 15326 (2015).
1043 [203] J. Mewis and G. Biebaut, Journal of Rheology 45, 799 (2001).
1044 [204] B. J. Maranzano and N. J. Wagner, Journal of Rheology 45, 1205 (2001).

U
1045 [205] J. F. Brady and G. Bossis, Annual Review of Fluid Mechanics 20, 111 (1988).
1046 [206] D. P. Kalman and N. J. Wagner, Rheologica Acta 48, 897 (2009).
1047 [207] A. K. Gurnon and N. J. Wagner, Journal of Fluid Mechanics 769, 242 (2015).
AN
1048 [208] X. Cheng, J. H. McCoy, J. N. Israelachvili, and I. Cohen, Science 333, 1276 (2011).
1049 [209] F. Gadala‐Maria and A. Acrivos, Journal of Rheology 24, 799 (1980).
1050 [210] J. R. Melrose and R. C. Ball, Journal of Rheology 48, 937 (2004).
M

1051 [211] J. F. Brady, The Journal of Chemical Physics 98, 3335 (1993).
1052 [212] J. F. Brady, R. J. Phillips, J. C. Lester, and G. Bossis, Journal of Fluid Mechanics 195, 257
1053 (1988).
1054 [213] D. R. Foss and J. F. Brady, Journal of Fluid Mechanics 407, 167 (2000).
D

1055 [214] A. Sierou and J. Brady, Journal of Rheology 46, 1031 (2002).
1056 [215] M. Cates, M. Haw, and C. Holmes, Journal of Physics: Condensed Matter 17, S2517 (2005).
TE

1057 [216] G. Batchelor and J. Green, Journal of Fluid Mechanics 56, 401 (1972).
1058 [217] G. Batchelor and J.-T. Green, Journal of Fluid Mechanics 56, 375 (1972).
1059 [218] J. F. Brady and J. F. Morris, Journal of Fluid Mechanics 348, 103 (1997).
1060 [219] R. Phillips, J. Brady, and G. Bossis, The Physics of Fluids 31, 3462 (1988).
EP

1061 [220] J. F. Brady, Journal of Fluid Mechanics 272, 109 (1994).


1062 [221] J. Bergenholtz, J. Brady, and M. Vicic, Journal of Fluid Mechanics 456, 239 (2002).
1063 [222] J. R. Smart and D. T. Leighton Jr, Physics of Fluids A: Fluid Dynamics 1, 52 (1989).
1064 [223] F. Da Cunha and E. Hinch, Journal of Fluid Mechanics 309, 211 (1996).
C

1065 [224] K. Galvin, Y. Zhao, and R. Davis, Physics of Fluids 13, 3108 (2001).
1066 [225] H. J. Wilson and R. H. Davis, Journal of Fluid Mechanics 452, 425 (2002).
AC

1067 [226] R. H. Davis, Y. Zhao, K. P. Galvin, and H. J. Wilson, Philosophical Transactions of the Royal
1068 Society of London A: Mathematical, Physical and Engineering Sciences 361, 871 (2003).
1069 [227] J. Brady, (Kavli Institute of Theoretical Physics, 2018).
1070 [228] R. P. Behringer and B. Chakraborty, Reports on Progress in Physics 82, 012601 (2018).
1071 [229] C. Lun and S. Savage, Journal of Applied Mechanics 54, 47 (1987).
1072 [230] O. Pouliquen and F. F. Chevoir, Comptes rendus. Physique 3, 163 (2002).
1073 [231] G. MiDi, The European Physical Journal E 14, 341 (2004).
1074 [232] A. J. Liu and S. R. Nagel, Nature 396, 21 (1998).
1075 [233] C. S. O’hern, L. E. Silbert, A. J. Liu, and S. R. Nagel, Physical Review E 68, 011306 (2003).
1076 [234] B. Luan and M. O. Robbins, Nature 435, 929 (2005).
1077 [235] M. Cates, J. Wittmer, J.-P. Bouchaud, and P. Claudin, Physical review letters 81, 1841 (1998).

41
ACCEPTED MANUSCRIPT

1078 [236] Z. Pan, H. de Cagny, B. Weber, and D. Bonn, Physical Review E 92, 032202 (2015).
1079 [237] H. De Cagny, A. Fall, M. M. Denn, and D. Bonn, Journal of Rheology 59, 957 (2015).
1080 [238] C. E. Maloney and A. Lemaître, Physical Review E 74, 016118 (2006).
1081 [239] K. Mair, K. M. Frye, and C. Marone, Journal of Geophysical Research: Solid Earth 107, ECV
1082 4 (2002).
1083 [240] P. A. Cundall and O. D. Strack, Geotechnique 29, 47 (1979).
1084 [241] F. Da Cruz, S. Emam, M. Prochnow, J.-N. Roux, and F. Chevoir, Physical Review E 72,

PT
1085 021309 (2005).
1086 [242] H. A. Vinutha and S. Sastry, Nature Physics 12, 678 (2016).
1087 [243] R. P. Jensen, P. J. Bosscher, M. E. Plesha, and T. B. Edil, International Journal for Numerical
1088 and Analytical Methods in Geomechanics 23, 531 (1999).

RI
1089 [244] K. Kamrin and G. Koval, Physical Review Letters 108, 178301 (2012).
1090 [245] M. Bouzid, M. Trulsson, P. Claudin, E. Clément, and B. Andreotti, Physical Review Letters
1091 111, 238301 (2013).

SC
1092 [246] Z. Tang, T. A. Brzinski, M. Shearer, and K. E. Daniels, Soft Matter 14, 3040 (2018).
1093 [247] A. Boromand, S. Jamali, B. Grove, and J. M. Maia, Journal of Rheology 62, 905 (2018).
1094 [**248] N. Y. Lin, B. M. Guy, M. Hermes, C. Ness, J. Sun, W. C. Poon, and I. Cohen, Physical Review
1095 Letters 115, 228304 (2015).

U
1096 The authors perform shear reversal experiments on a strain-controlled rheometer and propose a
1097 decoupling of the hydrodynamic and contact contributions to the shear viscosity. Simulations
1098 using the contact load model are used to support the experimental observations.
AN
1099 [**249] C. Clavaud, A. Bérut, B. Metzger, and Y. Forterre, Proceedings of the National Academy of
1100 Sciences, 201703926 (2017).
1101 Silica suspensions are placed in a pressure-controlled rotating drum and observations of the
1102 angle of repose are reported for transient shear flows, ranging from compaction to avalanches.
M

1103 Analyses of the static and dynamic angles of repose are used to generate a bulk friction
1104 coefficient.
1105 [250] R. A. Bagnold, Proceedings of the Royal Society A 225, 49 (1954).
D

1106 [**251] M. Wang and J. F. Brady, Physical Review Letters 115, 158301 (2015).
1107 Brownian Dynamics simulations without friction and hydrodynamics, but with a compressible
TE

1108 solvent, show that dense colloidal suspensions exhibit the type of bulk stress ratio to viscous
1109 number rheology that is seen with granular suspensions. Qualitative agreement with the data in
1110 ref. 172 is found.
1111 [252] T. Kawasaki, D. Coslovich, A. Ikeda, and L. Berthier, Physical Review E 91, 012203 (2015).
EP

1112 [253] A. Singh, R. Mari, M. M. Denn, and J. F. Morris, Journal of Rheology 62, 457 (2018).
1113 [254] M. van Hecke, Journal of Physics: Condensed Matter 22, 033101 (2009).
1114 [255] S. Papanikolaou, C. S. O’Hern, and M. D. Shattuck, Physical Review Letters 110, 198002
1115 (2013).
C

1116 [256] L. E. Silbert, Soft Matter 6, 2918 (2010).


1117 [257] M. P. Ciamarra, R. Pastore, M. Nicodemi, and A. Coniglio, Physical Review E 84, 041308
AC

1118 (2011).
1119 [258] A. Fall, G. Ovarlez, D. Hautemayou, C. Mézière, J.-N. Roux, and F. Chevoir, Journal of
1120 Rheology 59, 1065 (2015).
1121 [259] R. N. Chacko, R. Mari, M. E. Cates, and S. M. Fielding, arXiv preprint arXiv:1802.10586
1122 (2018).
1123 [260] C. Marone, Annual Review of Earth and Planetary Sciences 26, 643 (1998).
1124 [261] G. He, M. H. Müser, and M. O. Robbins, Science 284, 1650 (1999).
1125 [262] A. Fall, B. Weber, M. Pakpour, N. Lenoir, N. Shahidzadeh, J. Fiscina, C. Wagner, and D.
1126 Bonn, Physical Review Letters 112, 175502 (2014).
1127 [263] J. Castle, A. Farid, and L. Woodcock, in Trends in Colloid and Interface Science X (Springer,
1128 1996), pp. 259.

42
ACCEPTED MANUSCRIPT

1129 [264] Z. Pan, H. de Cagny, M. Habibi, and D. Bonn, Soft Matter 13, 3734 (2017).
1130 [265] J. E. Thomas, K. Ramola, A. Singh, R. Mari, J. Morris, and B. Chakraborty, arXiv preprint
1131 arXiv:1804.03155 (2018).
1132 [266] R. I. Tanner and S. Dai, Journal of Rheology 60, 809 (2016).
1133 [267] L. C. Hsiao, I. Saha-Dalal, R. G. Larson, and M. J. Solomon, Soft Matter 13, 9229 (2017).
1134 [268] F. Ianni, D. Lasne, R. Sarcia, and P. Hébraud, Physical Review E 74, 011401 (2006).
1135 [269] B. Guy, J. Richards, D. Hodgson, E. Blanco, and W. Poon, Physical Review Letters 121,

PT
1136 128001 (2018).
1137 [270] L. Suresh and J. Y. Walz, Journal of Colloid and Interface Science 183, 199 (1996).
1138 [271] G. B. Basim, I. U. Vakarelski, and B. M. Moudgil, Journal of Colloid and Interface Science
1139 263, 506 (2003).

RI
1140 [272] A. A. Feiler, L. Bergström, and M. W. Rutland, Langmuir 24, 2274 (2008).
1141 [273] T. Kawasaki and L. Berthier, arXiv preprint arXiv:1804.06800 (2018).
1142 [274] E. DeGiuli, G. Düring, E. Lerner, and M. Wyart, Physical Review E 91, 062206 (2015).

SC
1143

U
AN
M
D
TE
C EP
AC

43

You might also like