aËœaËœcontinuous Shear Thickening and Colloid Surfacesaa Phys Re

The popular interest in cornstarch and water mixtures known as "oobleck" after the

complex fluid

in one of Dr. Seuss's classic children's books arises from their transition from fluid-like to

solid-like

behavior when stressed. The

viscous liquid

that emerges from a roughly 2-to-1 (by volume) combination of starch to water can be poured into one's hand. When squeezed, the

liquid

morphs into a doughy

paste

that can be formed into shapes, only to "melt" into a puddle when the applied stress is relieved.

Internet

videos show people running across a large pool of the stuff, only to sink once they stop in place, and "monsters" that grow out of the mixture when it's acoustically vibrated (for an example, see the online version of this article).

Shear-thickening

fluids certainly entertain and

spark

our curiosity, but their

effect

can also vex

industrial processes

by fouling pipes and spraying equipment, for instance. And yet, when engineered into

composite materials,

STFs can be controlled and harnessed for such exotic applications as shock-absorptive skis and the soft body armor discussed in box 1

Engineers

and

colloid

scientists have wrestled with the scientific and practical problems of

shear-thickening colloidal

dispersions—typically composed of condensed

polymers, metals,

or oxides suspended in a

liquid—for

more than a century. More recently, the physics community has explored the highly nonlinear

materials

in the context of jamming 1 1. C. B. Holmes et al. , J. Rheol. 49, 237 (2005). https://doi.org/10.1122/1.1814114 (see the article by Anita Mehta, Gary Barker, and Jean-Marc Luck in PHYSICS TODAY, May 2009, page 40) and the more general study of

colloids

as

model systems

for understanding soft condensed matter.

Hard-sphere

colloids

are the "hydrogen atom" of

colloidal dispersions.

Because of their greater size and

interaction

times compared with atomic and molecular

systems, colloidal dispersions

are often well suited for

optical microscopy

and

scattering

experiments using light,

x rays,

and

neutrons.

That makes the

dispersions,

beyond their own intrinsic technological importance, ideal

models

for exploring equilibrium and near-equilibrium phenomena of interest in atomic and molecular physics—for example, phase behavior and "dynamical arrest," in which particles stop moving collectively at the

glass transition.

The relevance of

colloids

to atomic and molecular

systems

breaks down, though, for highly nonequilibrium phenomena. Indeed,

shear thickening

in strongly flowing

colloidal dispersions

may be among the most spectacular, and elucidating, examples of the differences between the

systems.

Figure 1 . illustrates the

effect.

The addition of

colloidal

particles to a

liquid

such as water results in an increase in the

liquid's viscosity

and, with further addition, the onset of non-Newtonian behavior—the dependence of its

viscosity

on an applied shear stress or

shear rate.

At high particle concentrations, the fluid behaves as if it has an apparent

yield stress.

That is, it must be squeezed, like ketchup, before it can actually

flow.

At such concentrations, the

colloidal dispersions

fit into the general paradigm for jamming in soft matter: 2 2. A. J. Liu S. R. Nagel, Nature 396, 21 (1998). https://doi.org/10.1038/23819 At high particle densities and low stresses (and low temperatures, usually), the

system

dynamically arrests, just as atomic, molecular,

polymeric,

and

granular systems

do. But once the

yield stress

is exceeded, the fluid's

viscosity drops,

a response known as

shear thinning.

That rheology is engineered into a range of consumer products, from shampoos and paints to

liquid

detergents, to make them gel-like at rest but still able to

flow

easily under a weak stress. Again, the

colloid model

fits the general paradigm for how matter behaves: It

flows

when sheared strongly enough.

At higher stresses,

shear thickening

occurs:

Viscosity

rises abruptly, sometimes discontinuously, once a critical shear stress is reached. The rise is counterintuitive and inconsistent with our usual experience. Experiments and simulations on atomic and small-molecule

liquids

predict only

shear thinning,

at least until the eventual onset of

turbulence

at

flow

rates that vastly exceed those of interest here.

The ubiquity of the phenomenon in the

flow

of suspended

solids

is a serious limitation for

materials processing,

especially when it involves high

shear-rate

operations. In a 1989 review, Howard Barnes writes,

Concentrated

suspensions

of nonaggregating

solid

particles, if

measured

in the appropriate

shear rate

range, will always show (reversible)

shear thickening.

The actual nature of the

shear thickening

will depend on the parameters of the suspended phase: phase volume, particle size (distribution), particle shape, as well as those of the suspending phase

(viscosity

and the details of the

deformation,

i.e., shear or

extensional flow,

steady or transient, time and rate of

deformation).

3 3. H. A. Barnes, J. Rheol. 33, 329 (1989). https://doi.org/10.1122/1.550017

Inks,

polymeric

binders for paints,

pastes,

alumina casting

slurries,

blood, and clays are all known to shear thicken. But the earliest searches for the root cause came from industrial laboratories that coated paper at high speeds

(shear rates

typically up to 106 Hz), a process in which the coating's increasing

viscosity

would either tear the paper or ruin the equipment. Industrial labs remain intensely interested in the science. Hundreds of millions of metric tons of cement are used globally each year, for example, and production

engineers

are careful to formulate modern high-strength cements and concretes that don't suffer from the effect—at least in a range of

shear rates

important for processing and construction. 4 4. F. Toussaint, C. Roy, P.-H. Jézéquel, Rheol. Acta, https://doi.org/10.1007/s00397-009-0362-z (2009).

In pioneering work in the 1970s, Monsanto's Richard Hoffman developed novel light-scattering experiments to probe the underlying microstructural transitions that accompanied

shear thickening

in concentrated latex

dispersions.

5 5. R. L. Hoffman, J. Rheol. 42, 111 (1998). https://doi.org/10.1122/1.550884 The transition was observed to correlate with a loss of Bragg peaks in the

scattering measurement.

On that basis, Hoffman developed a micromechanical

model

of

shear thickening

as a flow-induced

order-disorder transition.

In the 1980s and early 1990s BASF's Martin Laun and others interested in products such as paper coatings and emulsion-polymerized

materials

used then emerging small-angle neutron-scattering techniques to demonstrate that an

order-disorder transition

was neither necessary nor alone sufficient to induce significant

shear thickening.

6 6. H. M. Laun et al. , J. Rheol. 36, 743 (1992). https://doi.org/10.1122/1.550314 Because

shear thickening

is a highly nonequilibrium, dissipative state, though, a full understanding had to await the development of new

theoretical

and experimental tools.

Box 1. Soft armor and other applications

The unique

material properties

of increased energy dissipation combined with increased

elastic modulus

make

shear-thickening

fluids (STFs) ideal for damping and shock-absorption applications. For example, so-called EFiRST fluids can be switched between shear-thickened and flowing states using an applied

electric field,

which controls the damping.

Researchers

have also explored the STF response in sporting equipment 14 14. C. Fischer et al. , Smart Mater. Struct. 15, 1467 (2006). https://doi.org/10.1088/0964-1726/15/5/036 and automotive applications, 15 15. H. M. Laun, R. Bung, F. Schmidt, J. Rheol. 35, 999 (1991). https://doi.org/10.1122/1.550257 such as skis and tennis rackets that efficiently dissipate vibrations without losing stiffness or STFs embedded in a passenger compartment liner designed to protect passengers in a car accident.

One commercial application of STF composites is expected to be protective clothing. 16 16. Y. S. Lee, E. D. Wetzel, N. J. Wagner, J. Mater. Sci. 38, 2825 (2003). https://doi.org/10.1023/A:1024424200221 The fabric imaged in these scanning electron micrographs has STFs intercalated into its woven yarns. Initial applications are anticipated in flexible vests for correctional officers. Longer-term research is being performed in one of our laboratories (Wagner's), in conjunction with the

US

Army Research Laboratory, to use STF fabrics for

ballistic,

puncture, and blast protection for the military, police, and first responders.

Tests of the

materials

demonstrate a marked enhancement in performance. Consider this comparison between two STF-based fabrics: The velocity at which a quarter-inch steel ball is likely to penetrate a single layer of Kevlar is

measured

at about 100 m/s. The velocity required to penetrate Kevlar formulated with

polymeric colloids

(polymethyl-methacrylate) is about 150 m/s, and that for Kevlar formulated with

silica colloids

is 250 m/s, 2.5 times that for the Kevlar alone. High-speed video demonstrations and further test details are available at http://www.ccm.udel.edu/STF. Many other composites are now under investigation for armor applications. (Images courtesy of Eric Wetzel,

US

Army Research Laboratory.)

The

dynamics

of

colloidal dispersions

is inherently a many-body, multiphase fluid-mechanics problem. But first consider the case of a single particle. Fluid drag on the particle leads to the Stokes-Einstein- Sutherland

fluctuation-dissipation

relationship:

The diffusivity D 0 scales with the thermal energy kT divided by the suspending medium's

viscosity

µ and the particle's

hydrodynamic

radius a. That diffusivity sets the characteristic time scale for the particles'

Brownian motion;

it takes the particle a 2/D 0 seconds to diffuse a distance equal to its radius. The time scale defines high and low

shear rates

γ .

A dimensionless number known as the Péclet number, Pe, relates the

shear rate

of a

flow

to the particle's

diffusion

rate; alternatively, the Péclet number can be defined in terms of the applied shear stress τ:

The number is useful because

dispersion

rheology is often

measured

by applied

shear rates

or shear stresses. Low Pe is close enough to equilibrium that

Brownian motion

can largely restore the equilibrium microstructure on the time scale of slow

shear flow.

At sufficiently high

shear rates

or stresses, though,

deformation

of the

colloidal

microstructure by the

flow

occurs faster than

Brownian motion

can restore it.

Shear thinning

is already evident around Pe ≈ 1. And higher

shear rates

or stresses (higher Pe) trigger the onset of

shear thickening.

The presence of two or more particles in the

suspension

fundamentally alters the

Brownian motion

due to the inherent coupling, or

hydrodynamic interaction,

between the

motion

of the particles and the displacement of the suspending fluid. In a series of seminal articles in the 1970s, Cambridge University's George Batchelor laid a firm foundation for understanding the

colloidal dynamics.

7 7. W. B. Russel, D. A. Saville, W. R. Schowalter, Colloidal Dispersions, Cambridge U. Press, New York (1989). https://doi.org/10.1017/CBO9780511608810 In essence, because any particle

motion

must displace incompressible fluid, a long-ranged—and inherently many-body—force is transmitted from one particle through the intervening fluid to neighboring particles; the result is that all particles collectively disturb the local

flow

field through

hydrodynamic interactions.

Such

interactions

are absent in atomic and molecular fluids, where the intervening medium is vacuum.

Batchelor's calculation of the trajectories of non-Brownian particles under

shear flow

identified the critical importance of what's known as

lubrication hydrodynamics,

which describes the behavior of particles

interacting

via the suspending medium at very close range. Those

hydrodynamics

were already well known in the

fluid mechanics

of journal bearings, which Osborne Reynolds investigated in the late 1800s and which remain of great importance to the workings of modern machines. As box 2 explains, the force required to push two particles together in a fluid diverges inversely with their separation distance. Of particular significance is that at close range, the trajectories that describe their relative

motion

become correlated. That is, the particles effectively orbit each other—indefinitely if they are undisturbed.

Batchelor's work also led to a formal understanding of how

hydrodynamic

coupling alters the

fluctuation-dissipation

relationship, which, in turn, enabled him to calculate the

diffusion

coefficient and

viscosity

of dilute

dispersions

of Brownian

colloids

at equilibrium. 7 7. W. B. Russel, D. A. Saville, W. R. Schowalter, Colloidal Dispersions, Cambridge U. Press, New York (1989). https://doi.org/10.1017/CBO9780511608810 Although it was not fully appreciated at the time, the

effect

of

hydrodynamic interactions

on

particle trajectories

is the basis for understanding the

shear-thickening effect.
Hydrodynamic interactions

in real

colloidal suspensions

require numerical methods to solve. The method of Stokesian

dynamics

outlined in box 3 calculates the

properties

of ensembles of

colloidal

and noncolloidal spheres under

flow.

A great advantage of the simulations is their ability to resolve which forces contribute to the

viscosity.

Moreover, they demonstrate that the ubiquitous

shear thinning

in hard-sphere

colloidal dispersions

is a direct consequence of particle rearrangement due to the applied shear.

The equilibrium microstructure is set by the balance of stochastic and interparticle forces at play—including

electrostatic

and van der Waals forces—but is not affected by hydro-dynamic

interactions.

The low-shear (Pe ≪ 1)

viscosity

has two components, one due to direct interparticle forces, which dominate, and one due to

hydrodynamic interactions.

7 7. W. B. Russel, D. A. Saville, W. R. Schowalter, Colloidal Dispersions, Cambridge U. Press, New York (1989). https://doi.org/10.1017/CBO9780511608810 Under weak but increasing

shear flow

(Pe ~ 1), the fluid structure becomes

anisotropic

as particles rearrange to reduce their

interactions

so as to

flow

with less resistance. Figure 2 illustrates the evolution schematically. Near equilibrium, the resistance to

flow

is naturally high because shearing the random distribution of particles causes them to frequently collide, like cars would if careening haphazardly along a road. With increasing

shear rates,

though, particles behave as if merging into highway traffic: The

flow

becomes streamlined and the increasingly efficient

transport

of

colloidal

particles reduces the system's

viscosity.

Simulations that ignore

hydrodynamic

coupling between particles show that the ordered, low-viscosity state persists even as the Péclet number approaches infinity. Think of particles sliding by in layers orthogonal to the shear-gradient direction. Stokesian

dynamics

simulations, however, demonstrate that

hydrodynamic

forces become larger at high

shear rates

(Pe ≫ 1) than do interparticle forces that drive

Brownian motion.

So when the particles are driven close together by applied shear stresses,

lubrication hydrodynamics

strongly couple the particles' relative

motion.

The result is a

colloidal dispersion

that has a microstructure significantly different from the one near equilibrium, and hence, the energy dissipation increases. In hindsight, that should not be surprising given Batchelor's calculation of closed trajectories.

In both semidilute and concentrated

dispersions,

the strong

hydrodynamic

coupling between particles leads to the formation of hydroclusters—transient concentration

fluctuations

that are driven and sustained by the applied shear field. Here again, the analogy to traffic collisions disrupting organized, low-dissipation

flow

may be helpful. Unlike the seemingly random microstructure observed close to equilibrium, however, this microstructure is highly organized and

anisotropic.

The transient hydroclusters are the defining feature of the

shear-thickening

state.

Referring back to figure 1 , one can see that a

colloidal

volume fraction φ = 0.50 produces a latex

dispersion

whose

viscosity

is 1 Pa·s at a low shear stress and again at one more than four orders of magnitude higher. The same

viscosity

emerges for very different reasons, though. Changes in the particles' size, shape, surface chemistry, and ionic strength and in

properties

of the suspending medium all affect the interparticle forces, which dominate the

viscosity

at low shear stress.

Hydrodynamic

forces, in contrast, dominate at high shear stress. Understanding the difference is critical to formulating a

dispersion

that behaves as needed for specific processes or applications.

As shown in figure 3 ,

rheo-optical measurements

on

model dispersions

experimentally confirm the predictions of simulations that the shear-thickened state is driven by dissipative

hydrodynamic interactions.

The

flow

generates strong

anisotropy

in the nearest-neighbor distributions (see box 3 ). The

anisotropies

give rise to clusters of particles and concomitant large stress

fluctuations

8 8. J. R. Melrose R. C. Ball, J. Rheol. 48, 961 (2004). https://doi.org/10.1122/1.1784784 that, in turn, lead to high dissipation rates and thus a high shear

viscosity.

The formation of hydroclusters is generally reversible, though, so reducing the

shear rate

returns the

suspension

to a stable, flowing

suspension

with lower

viscosity.

Moreover, even very dilute

dispersions

will shear thicken, although the

effect

is hard to observe. 9 9. J. Bergenholtz, J. F. Brady, M. Vicic, J. Fluid Mech. 456, 239 (2002). https://doi.org/10.1017/S0022112001007583

Controlling

shear thickening

requires different strategies from those typically employed to control the low-shear

viscosity.

The addition, for example, of a

polymer

"brush" grafted or adsorbed onto the particles' surface can prevent particles from getting close together. With the right selection of graft density, molecular weight, and

solvent,

the onset of

shear thickening

moves out of the desired processing regime. 10 10. L.-N. Krishnamurthy, N. J. Wagner, J. Mewis, J. Rheol. 49, 1347 (2005). https://doi.org/10.1122/1.2039867 The strategy is often used to reduce the

viscosity

at high processing rates but could increase the suspension's low-shear

viscosity.

Indeed, because the separation between hydroclustered particles is predicted to be on the order of nanometers for typical

colloidal dispersions, shear-thickening

behavior directly reflects the particles'

surface structure

and any short-range interparticle forces at play. Fluid slip, adsorbed ions,

surfactants, polymers,

and

surface roughness

all significantly influence the onset of

shear thickening.

Simple

models

based on the hydrocluster mechanism have proven valuable in predicting the onset of

shear thickening

and its dependence on those stabilizing forces. 11 11. B. J. Maranzano N. J. Wagner, J. Chem. Phys. 114, 10514 (2001). https://doi.org/10.1063/1.1373687

Figure 4 shows a toy-model calculation in which

shear thickening

is suppressed by imposing a purely repulsive force field—akin to the

effect

of a

polymer

brush—around each particle that prevents the particles from getting too close to each other. 9 9. J. Bergenholtz, J. F. Brady, M. Vicic, J. Fluid Mech. 456, 239 (2002). https://doi.org/10.1017/S0022112001007583 When the range of the repulsive force approaches 10% of the particle radius, the

shear thickening

is effectively eliminated and the

suspension flows

with low

viscosity.

Manipulating those nanoscale forces, the particles' composition and shape, and

properties

of the suspending fluid so as to control the sheer thickening, however, remains a challenge for the

suspension

formulator.

Box 2. Lubrication hydrodynamics and hydroclusters

When two

colloidal

particles approach each other, rising

hydrodynamic

pressure between them squeezes fluid from the gap. At close range, the

hydrodynamic

force increases inversely with the distance between the particles' surfaces and diverges to a singularity. The graph at right plots the normalized force required to drive two particles together (along a line through their centers) at constant velocity. The

Navier-Stokes equations

that govern the

flow

behavior between particles are time reversible, so the force is the same one required to separate two particles.

In simple

shear flow, particle trajectories

are strongly coupled by the hydro-dynamic

interactions

if the particles are close together. The inset of the plot shows a test particle's trajectories, sketched as paths as it moves in a

shear flow

relative to a reference particle (gray sphere). The trajectories are reversible and can be divided into two classes: those that come and go to infinity and those that lead to correlated orbits—so-called closed trajectories—around the reference particle.

Simulations and

theory

of concentrated

dispersions

that account for those short-range

hydrodynamics

show that at high

shear rates,

particles that are driven into close proximity remain strongly correlated and are reminiscent of the closed trajectories observed in dilute

suspensions.

The flow-induced density

fluctuations

are known as hydroclusters. Because the particle concentration is higher in the clusters, the fluid is under greater stress, which leads to an increase in energy dissipation and thus a higher

viscosity.

The illustration at right during a stage of the Stokesian

dynamics

simulation shows

colloidal

particles in hydroclusters. 8 8. J. R. Melrose R. C. Ball, J. Rheol. 48, 961 (2004). https://doi.org/10.1122/1.1784784

Box 3. Stokesian dynamics

The

flow

of particles suspended in an incompressible Newtonian fluid is a challenging fluid-mechanics problem that can be handled analytically for a single sphere and semianalytically for two spheres. For three or more particles, though, it requires a

numerical solution

to the Stokes equation—the

Navier-Stokes equation

without inertia. Solution strategies range from brute-force

finite-element

calculations, to more elegant

boundary integral methods,

to coarse-grained methods, such as smoothedparticle

hydrodynamics

or

lattice Boltzmann

techniques, for representing the fluid. The method of Stokesian

dynamics

17 17. For a review, see J. F. Brady G. Bossis, Annu. Rev. Fluid Mech. 20, 111 (1988). https://doi.org/10.1146/annurev.fl.20.010188.000551 starts with the

Langevin equation

for N-particle

dynamics,

in which the

tensor

M is a generalized mass, a 6N × 6N mass and moment-of-inertia matrix; U is the 6N-dimensional particle translational and rotational velocity vector; and the 6N-dimensional force and

torque

vectors represent the interparticle and external forces F P (such as gravity),

hydrodynamic

forces F H acting on the particles due to their

motion

relative to the fluid, and stochastic forces F B that give rise to

Brownian motion.

The stochastic forces are related to the

hydrodynamic interactions

through the

fluctuation-dissipation theorem.

In

Stokes flow

the

hydrodynamic

forces and torques are linearly related to the particle translational and rotational velocities as F H = −R · U, where R is the configuration-dependent

hydrodynamic

resistance matrix. In the Stokesian

dynamics

method, the necessary matrices are computed by taking advantage of the linearity of the Stokes

equations

and their integral solutions. Long-range many-body hydro-dynamic

effects

are accurately computed by a force-multipole expansion and combined with the exact, analytic

lubrication hydrodynamics

to calculate the forces.

Armed with that method, one can predict the

colloidal

microstructure associated with a particular shear

viscosity.

Take, for instance, a concentrated

colloidal dispersion

whose particles occupy nearly half the volume. If the positions of those particles are represented as dots, the figure illustrates how the hydro-dynamic forces affect their probable locations around some arbitrary test particle (black). The three panels differ only in the

shear rate,

represented by the Péclet number Pe, the ratio of the shear and

diffusion

rates.

At low Péclet number (0.1), the distribution of neighboring particles is isotropic, which is typical of a concentrated

liquid.

Red indicates the most probable particle positions as nearest neighbors and green the least probable. At Pe = 1, significant shear distortion appears in neighbor distributions, such that particles are convected together along the compression axes (135° and −45°) relative to the

shear flow.

At high Péclet numbers, the

shear-thickening

regime, particles aggregate into closely connected clusters, which is manifest as yet greater

anisotropy

in the microstructure. Particles are more closely packed and thus occupy a narrower region (red) around the test particle than at lower Péclet numbers, indicative of being trapped by the

lubrication

forces. 18 18. D. R. Foss J. F. Brady, J. Fluid Mech. 407, 167 (2000). https://doi.org/10.1017/S0022112099007557

Although the basic

micromechanics

of shear behavior in

colloidal suspensions

are understood, many aspects of the fascinating and

complex fluids

remain active research problems. At very high particle densities,

dispersions

can undergo discontinuous

shear thickening

whereby the

suspension

will not shear at any higher rate. Rather, increasing the power to a rheometer, for example, leads to such dramatic increases in

viscosity

and large

fluctuations

in stress that the

suspension

either refuses to

flow

any faster or

solidifies.

Samples that exhibit strong

shear thickening

are particularly interesting as candidates for soft body armor (see box 1 ), and that application has prompted investigations of transient

shear thickening

at microsecond time scales and at stresses that approach the ideal strength of the particles.

Another active research topic concerns jamming transitions under

flow.

As figure 1 suggests, concentrated

suspensions

could be jammed at low and high shear stresses but

flow

in between. Evidence also exists, as the figure more subtly suggests, that

dispersions

may exhibit a second regime of

shear thinning

at the highest stresses rather than continuing to resist the increasing

shear rate.

The

effect

can be understood as a manifestation of the finite

elasticity

of the particles—relatively soft plastic in this case. At very high stresses, particles stop behaving like billiard balls and elastically

deform,

which alters their rheology. The same forces that drive the hydrocluster formation, which is reversible as the

flow

is reduced, can also lead to irreversible

aggregation.

That is, particles forced into contact remain in contact even as the

flow

weakens. Such shear-sensitive

dispersions

irreversibly thicken and are often undesirable in practice.

Conversely, in

dispersions

composed of particle aggregates or fillers such as fumed

silica

or

carbon

black, the extreme forces can lead to particle breakage and

thixotropy

(time-dependent

viscosity).

Indeed, propagating those forces into the

colloids

may be key to splitting the

colloids

into

nanoparticles.

It's thought, for instance, that the extreme mechanical stress required to grind up and pulverize particles is more effectively transferred to the particles when they are in a shear-thickened state in the

slurry

of a mill.

Interesting questions arise in the role of

shear thickening

in chemical mechanical planarization, a critical step in

semiconductor

processing. Concentrated

dispersions

are useful for other

polishing

operations as well, and the control of their

shear thickening

can be critical to performance.

Although it's impossible to completely survey the science surrounding

shear thickening

in

colloidal dispersions

and its applications, we hope the highly counterintuitive rheology has piqued your interest. A wealth of fascinating challenges and applications awaits.

  1. 1. C. B. Holmes et al. , J. Rheol. 49, 237 (2005). https://doi.org/10.1122/1.1814114 , Google Scholar Crossref , ISI , CAS
  2. 2. A. J. Liu S. R. Nagel, Nature 396, 21 (1998). https://doi.org/10.1038/23819 , Google Scholar Crossref , ISI , CAS
  3. 3. H. A. Barnes, J. Rheol. 33, 329 (1989). https://doi.org/10.1122/1.550017 , Google Scholar Crossref , ISI , CAS
  4. 4. F. Toussaint, C. Roy, P.-H. Jézéquel, Rheol. Acta, https://doi.org/10.1007/s00397-009-0362-z (2009). Google Scholar Crossref
  5. 5. R. L. Hoffman, J. Rheol. 42, 111 (1998). https://doi.org/10.1122/1.550884 , Google Scholar Crossref , CAS
  6. 6. H. M. Laun et al. , J. Rheol. 36, 743 (1992). https://doi.org/10.1122/1.550314 , Google Scholar Crossref , ISI , CAS
  7. 7. W. B. Russel, D. A. Saville, W. R. Schowalter, Colloidal Dispersions, Cambridge U. Press, New York (1989). https://doi.org/10.1017/CBO9780511608810 , Google Scholar Crossref
  8. 8. J. R. Melrose R. C. Ball, J. Rheol. 48, 961 (2004). https://doi.org/10.1122/1.1784784 , Google Scholar Crossref , CAS
  9. 9. J. Bergenholtz, J. F. Brady, M. Vicic, J. Fluid Mech. 456, 239 (2002). https://doi.org/10.1017/S0022112001007583 , Google Scholar Crossref , CAS
  10. 10. L.-N. Krishnamurthy, N. J. Wagner, J. Mewis, J. Rheol. 49, 1347 (2005). https://doi.org/10.1122/1.2039867 , Google Scholar Crossref , CAS
  11. 11. B. J. Maranzano N. J. Wagner, J. Chem. Phys. 114, 10514 (2001). https://doi.org/10.1063/1.1373687 , Google Scholar Crossref , ISI , CAS
  12. 12. H. M. Laun, Angew. Makromol. Chem. 123, 335 (1984). https://doi.org/10.1002/apmc.1984.051230115 , Google Scholar Crossref
  13. 13. J. Bender N. J. Wagner, J. Rheol. 40, 899 (1996). https://doi.org/10.1122/1.550767 , Google Scholar Crossref , ISI , CAS
  14. 14. C. Fischer et al. , Smart Mater. Struct. 15, 1467 (2006). https://doi.org/10.1088/0964-1726/15/5/036 , Google Scholar Crossref
  15. 15. H. M. Laun, R. Bung, F. Schmidt, J. Rheol. 35, 999 (1991). https://doi.org/10.1122/1.550257 , Google Scholar Crossref , ISI , CAS
  16. 16. Y. S. Lee, E. D. Wetzel, N. J. Wagner, J. Mater. Sci. 38, 2825 (2003). https://doi.org/10.1023/A:1024424200221 , Google Scholar Crossref , CAS
  17. 17. For a review, see J. F. Brady G. Bossis, Annu. Rev. Fluid Mech. 20, 111 (1988). https://doi.org/10.1146/annurev.fl.20.010188.000551 , Google Scholar Crossref , ISI
  18. 18. D. R. Foss J. F. Brady, J. Fluid Mech. 407, 167 (2000). https://doi.org/10.1017/S0022112099007557 , Google Scholar Crossref , ISI , CAS
  1. © 2009 American Institute of Physics.

cambelldilgin.blogspot.com

Source: https://physicstoday.scitation.org/doi/10.1063/1.3248476?feed=most-cited

0 Response to "aËœaËœcontinuous Shear Thickening and Colloid Surfacesaa Phys Re"

Post a Comment

Iklan Atas Artikel

Iklan Tengah Artikel 1

Iklan Tengah Artikel 2

Iklan Bawah Artikel