shear banding in crystallizing colloidal suspensions · in contrast to equilibrium crystallization,...

8
Korea-Australia Rheology Journal December 2010 Vol. 22, No. 4 309 Korea-Australia Rheology Journal Vol. 22, No. 4, December 2010 pp. 309-316 Shear banding in crystallizing colloidal suspensions Laura T. Shereda, Ronald G. Larson and Michael J. Solomon* University of Michigan Ann Arbor MI 48109-2136 (Received May 25, 2010; accepted June 11, 2010) Abstract We characterize the shear bands generated in simple shear flow of a crystallizing colloidal suspension. 35 volume % suspensions of poly(methyl methacrylate) colloids of diameter 0.68 µm were dispersed in the viscous solvent dioctyl phthalate and subjected to plane Couette flow. The equilibrium structure of this sus- pension was crystalline and flow accelerated its crystallization kinetics significantly. Confocal laser scan- ning microscopy and particle tracking were used to characterize the height-dependent velocity profile in the gap of the shear flow. Near each of the two boundary surfaces, a region of high shear rate flow was observed. A low shear rate region was observed at the center of the gap. The differences in the shear rate within the two banded regions were a function of the both the applied shear rate and strain. The effect of strain indicated that the shear band development was a transient phenomenon. We found that the boundary between the high shear rate and low shear rate regions correlated with the location of crystalline and amor- phous regions in the gap of the shear cell, as visualized by confocal microscopy. Furthermore, the different local shear rates observed in the banded regions were consistent with the different viscosities of the amor- phous and crystalline suspensions. The results demonstrate that shear banded flows accompany shear- induced colloidal crystallization, and that the bands exhibit transient behavior because the crystallization process itself is strain dependent. Keywords : colloidal crystallization, confocal microscopy, self-assembly 1. Introduction Above a critical volume fraction, dense colloidal sus- pensions undergo phase transitions in order to minimize free energy (Anderson and Lekkerkerker, 2002). Such equilibrium crystallization has been studied extensively in model colloid systems. The rate of crystallization by homogeneous mechanisms can be predicted by means of theories such as classical nucleation theory (Gasser, 2009). In contrast to equilibrium crystallization, the kinetics of crystallization under non-equilibrium conditions are not well characterized. The non-equilibrium conditions of flow have a number of effects on colloidal crystallization. First, flow can stabilize metastable structures by shifting phase boundaries (Holmqvist et al., 2005). Second, it can accel- erate the kinetics of crystallization. In the second case, the rate of colloidal crystal formation depends strongly on the strength of the applied flow field relative to the random- izing effects of Brownian motion, as quantified by a dimensionless Peclet number and first identified by Ack- erson and coworkers (Ackerson 1990). Indeed, application of flow has been shown by scattering to lead to a rich sequence of microstructural transitions and metastable phases (Chen et al., 1994). Here we study the relationship between flow-induced crystallization in concentrated col- loidal suspensions and deviations from homogeneous shear flow, such as shear banding, in such suspensions. Quan- tifying the correlation between shear banding and crys- tallization kinetics is necessary for any predictive theory of flow-induced crystallization. Shear banding has been observed in concentrated col- loidal suspensions in a number of cases and a few studies have connected shear banding and shear-induced crystal- lization or melting. Holmqvist et al. studied crystallization kinetics in a charge-stabilized colloidal suspension (Holm- qvist et al., 2005). At a particle concentration of 25.2 wt%, they found no effect of shear-induced phenomena such as shear banding on crystal growth rates. Cohen et al. studied steady-state shear banding in crystallized colloidal sus- pensions in oscillatory flow (Cohen et al., 2006). Shear banding was asymmetric about the mid-plane of the flow and it was generated by a yielding transition due to the strain amplitude of the high rate flow. Wu et al. reported steady-state shear banding in colloidal suspensions in which the location of the bands was due to the coexistence *Corresponding author: [email protected] © 2010 by The Korean Society of Rheology

Upload: others

Post on 20-Jul-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Shear banding in crystallizing colloidal suspensions · In contrast to equilibrium crystallization, the kinetics of crystallization under non-equilibrium conditions are not well characterized

Korea-Australia Rheology Journal December 2010 Vol. 22, No. 4 309

Korea-Australia Rheology JournalVol. 22, No. 4, December 2010 pp. 309-316

Shear banding in crystallizing colloidal suspensions

Laura T. Shereda, Ronald G. Larson and Michael J. Solomon*

University of Michigan Ann Arbor MI 48109-2136

(Received May 25, 2010; accepted June 11, 2010)

Abstract

We characterize the shear bands generated in simple shear flow of a crystallizing colloidal suspension.35 volume % suspensions of poly(methyl methacrylate) colloids of diameter 0.68 µm were dispersed in theviscous solvent dioctyl phthalate and subjected to plane Couette flow. The equilibrium structure of this sus-pension was crystalline and flow accelerated its crystallization kinetics significantly. Confocal laser scan-ning microscopy and particle tracking were used to characterize the height-dependent velocity profile in thegap of the shear flow. Near each of the two boundary surfaces, a region of high shear rate flow wasobserved. A low shear rate region was observed at the center of the gap. The differences in the shear ratewithin the two banded regions were a function of the both the applied shear rate and strain. The effect ofstrain indicated that the shear band development was a transient phenomenon. We found that the boundarybetween the high shear rate and low shear rate regions correlated with the location of crystalline and amor-phous regions in the gap of the shear cell, as visualized by confocal microscopy. Furthermore, the differentlocal shear rates observed in the banded regions were consistent with the different viscosities of the amor-phous and crystalline suspensions. The results demonstrate that shear banded flows accompany shear-induced colloidal crystallization, and that the bands exhibit transient behavior because the crystallizationprocess itself is strain dependent.

Keywords : colloidal crystallization, confocal microscopy, self-assembly

1. Introduction

Above a critical volume fraction, dense colloidal sus-

pensions undergo phase transitions in order to minimize

free energy (Anderson and Lekkerkerker, 2002). Such

equilibrium crystallization has been studied extensively in

model colloid systems. The rate of crystallization by

homogeneous mechanisms can be predicted by means of

theories such as classical nucleation theory (Gasser, 2009).

In contrast to equilibrium crystallization, the kinetics of

crystallization under non-equilibrium conditions are not

well characterized. The non-equilibrium conditions of flow

have a number of effects on colloidal crystallization. First,

flow can stabilize metastable structures by shifting phase

boundaries (Holmqvist et al., 2005). Second, it can accel-

erate the kinetics of crystallization. In the second case, the

rate of colloidal crystal formation depends strongly on the

strength of the applied flow field relative to the random-

izing effects of Brownian motion, as quantified by a

dimensionless Peclet number and first identified by Ack-

erson and coworkers (Ackerson 1990). Indeed, application

of flow has been shown by scattering to lead to a rich

sequence of microstructural transitions and metastable

phases (Chen et al., 1994). Here we study the relationship

between flow-induced crystallization in concentrated col-

loidal suspensions and deviations from homogeneous shear

flow, such as shear banding, in such suspensions. Quan-

tifying the correlation between shear banding and crys-

tallization kinetics is necessary for any predictive theory of

flow-induced crystallization.

Shear banding has been observed in concentrated col-

loidal suspensions in a number of cases and a few studies

have connected shear banding and shear-induced crystal-

lization or melting. Holmqvist et al. studied crystallization

kinetics in a charge-stabilized colloidal suspension (Holm-

qvist et al., 2005). At a particle concentration of 25.2 wt%,

they found no effect of shear-induced phenomena such as

shear banding on crystal growth rates. Cohen et al. studied

steady-state shear banding in crystallized colloidal sus-

pensions in oscillatory flow (Cohen et al., 2006). Shear

banding was asymmetric about the mid-plane of the flow

and it was generated by a yielding transition due to the

strain amplitude of the high rate flow. Wu et al. reported

steady-state shear banding in colloidal suspensions in

which the location of the bands was due to the coexistence*Corresponding author: [email protected]© 2010 by The Korean Society of Rheology

Page 2: Shear banding in crystallizing colloidal suspensions · In contrast to equilibrium crystallization, the kinetics of crystallization under non-equilibrium conditions are not well characterized

Laura T. Shereda, Ronald G. Larson and Michael J. Solomon

310 Korea-Australia Rheology Journal

of amorphous and crystalline structures in a shear flow

(Wu et al., 2009). The local shear rate in the crystal band

was greater than the shear rate in the amorphous band and

this difference was linked to the different steady-state vis-

cosities of the two structures. Our work here complements

these studies by: (i) quantifying the transient behavior of

the shear bands as they develop from a suspension with ini-

tially amorphous structure; (ii) connecting the transient

development of the shear bands to the flow-induced crys-

tallization kinetics; (iii) connecting the band development

to the different viscosities of the amorphous and crystalline

suspensions.

Our focus in this study is shear banding which occurs at

overall shear rates for which the dimensionless flow Peclet

number, Pe, is greater than one and less than about 103, 1

< Pe < 102. Within this range, for suspensions with volume

fraction in the coexistence or fully crystalline regions of

the phase diagram, steady or oscillatory flow accelerates

the production of colloidal crystals (Vermant and Solomon,

2005). Here where yx is the shear stress, a is the

particle radius and kbT is the thermal energy. In a prior

study, we showed that the 35 volume% suspensions of ster-

ically stabilized poly(methyl methacrylate) colloids under-

went flow-induced crystallization in the flow generated in

a spin coating device (Shereda et al., 2008). Here we study

the behavior of suspensions at the same volume fraction in

plane Couette flow. We find interesting connections

between shear banded structure, transient crystallization

behavior, and rheology.

2. Materials and Methods

2.1. Colloidal suspensionsColloidal particle synthesis was carried out using meth-

ods described in the literature (Antl et al., 1986; Path-

mamanoharan et al., 1989; Campbell and Bartlett, 2002).

The resulting poly(methyl methacrylate) (PMMA) parti-

cles contained Nile Red dye, which allowed for imaging by

confocal microscopy using a 488 nm laser excitation. The

particles were negatively charged and interacted repul-

sively.

The colloids were dispersed in the viscous, non-volatile

solvent dioctyl phthalate at a concentration of 35 volume

percent. (This viscous solvent was chosen to shift the con-

ditions necessary for shear-induced crystallization, Pe > 1,

into a range accessible to confocal microscopy visualiza-

tion during shear flow (Solomon and Solomon 2006;

Kogan et al., 2008).) From scanning electron microscopy,

the diameter of the particles was 0.68 µm and their stan-

dard deviation in diameter was 2.3%. Addition compo-

nents added to the solvent were 7% photopolymer (CD501,

Sartomer), 1% photoinitiator (Irgacure 2100, Ciba Spe-

cialty) and 10 µM tetrabutylammoniumchloride (TBAC).

The photopolymer and initiator were added to allow immo-

bilization of the structure developed under flow so as to

characterize the crystallinity of the structure by 3D con-

focal laser scanning microscopy (Shereda et al., 2008). The

disassociating salt TBAC was added to screen charge in

the system (Royall et al., 2003) such that the 35 volume

percent concentration studied here is within the fully crys-

talline region of the equilibrium phase diagram. In addition

to the principal measurements at 35 volume%, additional

measurements were conducted under dilute conditions

(2.5 volume%), to assess the performance of the shear cell

device, as discussed in the next section.

The steady-state rheology of the colloidal suspension

prepared was measured by cone and plate rheometry

(ARG2, TA Instruments) with a cone of diameter 6 cm and

cone angle 2o, as reported in Fig. 1. For these measure-

ments, the steady-state viscosity at each shear rate was

measured for 600 s after an initial delay of 30 s so as to

allow the system to reach steady-state. Fig. 1 shows the

significant shear thinning and modest high rate shear thick-

ening of this concentrated colloidal suspension. Measure-

ments reported here are within the shear thinning region of

the Fig. 1 flow curve.

Peτyxa

3

kbT-----------=

Fig. 1. Shear-rate dependent viscosity, η, as a function of shear

rate in 35 volume% suspensions of colloidal PMMA of

average diameter 0.68 m. The solvent, as described in the

text, is comprised of 7% photopolymer and 1% initiator

dissolved in the viscous solvent dioctyl phthalate. A 6 cm

cone geometry with an angle of 2o was used for mea-

surements. The preshear protocol used is described in the

text. Error bars, which are standard error of the mean of

three replications, are plotted, but are not much larger than

the size of a datum point.

Page 3: Shear banding in crystallizing colloidal suspensions · In contrast to equilibrium crystallization, the kinetics of crystallization under non-equilibrium conditions are not well characterized

Shear banding in crystallizing colloidal suspensions

Korea-Australia Rheology Journal December 2010 Vol. 22, No. 4 311

2.2. Plane couette shear cellFig. 2 shows a schematic of the shear cell constructed to

allow confocal microscopy imaging of the gap region of

plane Couette flow. Briefly, the flow was generated by dis-

placement of a top plate comprised of glass of thickness

3.18 mm. The length of the plate was 9.62 cm, its width

was 2.39 cm, and its maximum distance of travel was

~1 cm. The top plate was displaced with a 5-phase stepper

motor (PMM3B2, Oriental motor company ltd) mounted

on a slider (537233, Coherent). LabMotion v1.0 software

was used to control the movement. The bottom plate was

a glass coverslip (Fisherbrand #1, 40 mm×22 mm) care-

fully glued with a thin film of GB341-UV (Dymax) glue to

an aluminum surface with a 1.91 cm square hole in the

center. The angle of the bottom plate could be adjusted to

ensure the plates were parallel to each other. The paral-

lelism of the plates was determined to be ±0.2 µm/mm by

direct visualization with the confocal microscope. The gap

was adjusted by changing the height of the top plate using

a micrometer. It was set to 150 m for these experiments.

For this study, plate velocities of 12.5, 25 and 40 µm/s

were used. At a gap of 150 µm, these conditions corre-

sponded to shear rates of 0.083, 0.166, and 0.266 s-1. In

terms of the microstructural response, the relevant dimen-

sionless group is the Peclet number, as defined earlier.

From the particle size reported above and the steady-state

shear viscosity presented in Fig. 1, the Peclet numbers of

the conditions studied (for the case of the 35 vol% sus-

pensions) were 12.9, 17.4 and 23.3. Prior to experiments at

these rates, specimens were shear melted by applying a

deformation rate of 100 s-1 for 120 s.

To address the possibility of wall slip, two different sur-

faces were tested. The first were smooth, unmodified glass

surfaces. In the second, both the bottom and top plates

were dip coated in a dilute solution of polydisperse PMMA

particles (mean diameter ~2 µm) dispersed in hexane. The

plates were heated to a temperature of 170oC and held for

3 hours. The melted PMMA coated the surfaces with a

layer of roughness of order ~2 µm. No significant dif-

ference in measured velocity profiles was observed for

either 2.5 vol% or 35 vol% suspensions of colloidal

PMMA (data now shown).

2.3. Confocal laser scanning microscopyTo image the microstructure of the concentrated colloidal

suspensions, we used confocal laser scanning microscopy

(CLSM, Leica SP2, 100x NA=1.4 oil immersion objec-

tive) to visualize the colloidal particles. Two kinds of

experiments were conducted. The first was to determine

the velocity profile by particle tracking. The second was to

characterize the crystalline and amorphous microstructure

developed during the flow.

In the first type of experiment, a time series of two-

dimensional (2D) images at a fixed height in the shear flow

were collected. Typically 25 to 400 images at 512×512 res-

olution (0.0456 µm/pixel) were acquired with a delay

(exposure) time of 0.823 s. These image series were used to

determine the mean suspension velocity at the gap height of

the measurement, as discussed in the next section.

In the second type of experiment, a three-dimensional

(3D) image volume was acquired after the specimen had

been sheared and then its structure immobilized (fixed) by

UV-induced polymerization. This polymerization occurred

due to the monomer and initiator suspended in the solvent,

as discussed in the materials section, and was initiated by

illumination of a UV source (propagated through the

microscope optics) onto the specimen. The structure, as

discussed in Shereda et al., 2008, was rapidly polymerized

in less than < 1 s with no significant distortion of the exist-

ing microstructure. These 3D images volumes were

acquired across the lower half of the shear flow gap by a

collecting a set of 2D images (0.0456 µm/pixel) spaced at

a slice distance of 0.0814 µm. In a typical experiment

about 1,000 slices were acquired parallel to the flow direc-

tion from the bottom plate to the mid-plane of the flow.

These 3D image volumes were used to determine crys-

talline and amorphous regions in the suspensions induced

by the shear flow, as discussed following.

2.4. Particle tracking method for determination of

height-dependent velocity profileThe following describes the method to characterize the

gap height dependence of the velocity profile in the shear

cell. The method is illustrated in Fig. 3 for crystallizing

suspensions comprised of a 35 volume percent solution of

colloidal PMMA. In Fig. 3, the particles that appear red

(dark grey) are those of the crystallizing colloidal sus-

pension which contain Nile Red dye. Those that appear

Fig. 2. Schematic of the shear cell used for the measurement of

the height-dependent velocity profile by confocal micros-

copy in crystallizing suspensions of colloidal PMMA of

diameter 0.68 µm. (a) depicts the device which is

mounted above the confocal microscope; (b) shows a

view from top down of the shearing surfaces. Both sur-

faces are transparent to allow confocal imaging.

Page 4: Shear banding in crystallizing colloidal suspensions · In contrast to equilibrium crystallization, the kinetics of crystallization under non-equilibrium conditions are not well characterized

Laura T. Shereda, Ronald G. Larson and Michael J. Solomon

312 Korea-Australia Rheology Journal

green (light grey) are the probe particles used to charac-

terize the velocity profile. Probe particles, used for particle

tracking, are of diameter 0.71 µm. They are dispersed at a

concentration of 0.5 volume %. They contain the dye Cou-

marin 30, which fluoresces in an emission band that is dis-

tinct from the Nile Red dye used for the primary particles.

A time series of images during flow was captured at a par-

ticular height in the sample at a particular shear rate. The

image plane was located at a gap height of interest for

velocity measurements.

Two consecutive images, as shown in Fig. 3, and sep-

arated by a delay time of 0.823 s, were extracted from the

time series for further analysis. The probe particles, which

because of their dilute concentration, were widely sepa-

rated, were easily located in the two samples and their

coordinates in the objective plane determined by selecting

two consecutive frames near the end of each time series

and locating four (green) tracer particles in the nth frame

and locating the same four particles in the n+1th frame.

The coordinates of each of the particles were determined

using the public domain software package ImageJ and the

distance the particles traveled in the flow direction was

divided by the delay time (0.823 s) to yield the particle

velocity. The velocities of the four particles tracked were

averaged to determine velocity. This was repeated for three

sets of frames in each time series.

2.5. Quantitative image processing to determinecrystalline and amorphous regions of the sus-

pensionsThe centroids of the colloids in the suspension volumes

acquired by 3D confocal microscopy were located by

image processing tools that have been previously described

(Crocker and Grier, 1996). Centroids were determined to a

resolution of ±35 nm in the objective plane and to ±45 nm

along an axis perpendicular to the objective plane (Dibble

et al., 2006). The relative positions of the centroidal coor-

dinates were analyzed to determine if a given particle was

part of a locally crystalline structure. The method used is

adapted from the simulations of ten Wolde et al. (ten

Wolde et al., 1996), and previously used for experiments

by Gasser et al. (Gasser et al., 2001), Solomon and

Solomon (Solomon and Solomon, 2006) and Shereda et al.

(Shereda et al., 2008). The output of this analysis was the

3D position of colloids in the suspension and the assign-

ment of their microstructure as either amorphous or crys-

talline.

The local structure assessment method uses a structural

measure that is sensitive to the overall degree of crystal-

linity but insensitive to both the type of structure (i.e. FCC

vs. HCP) and its orientation in space. The method com-

putes a vector q6 for each particle that is characteristic of

the local order. As computed from spherical harmonics, the

mth component of q6 is (ten Wolde et al. 1996):

(1)

where

(2)

Here i was the reference particle, j was a neighboring

particle, the coordination number Nb was the number of

neighboring particles, and were spherical harmon-

ics evaluated along the axis connecting the ith and jth

particles. For this calculation, a neighboring particle was

defined as a particle located within a distance of the first

minimum of g(r) from the centeroid of the reference par-

ticle (i).

To determine the local structure of a sample, each indi-

vidual particle in the sample was determined to be crys-

talline or amorphous by calculating the bond coherence

between the reference particle and each of its neighbors.

Bond coherence was measured as the dot product of q6(i)

and q6(j) as shown below.

(3)

A bond was considered crystalline if q6(i) ·q6(j) > 0.5.

q6m i( )q6m i( )

q6m i( )2

m 6–=

m 6=

∑0.5

--------------------------------------=

q6m i( )1

Nb i( )------------ Y6m rij( )

j 1=

Nbi( )

∑=

Y6m rij( )

rij

q6 i( ) q6 j( )⋅ q6m i( )q6m j( )∗

m 6–=

6

∑=

Fig. 3. Images demonstrating the particle tracking method used to

measure the height dependent velocity profile in the crys-

tallizing colloidal suspensions. See text for description of

the method from which the velocity at a particular height

was determined from a sequence of two images (a) and

(b), acquired at the gap height of interest, and separated

by a delay time of 0.832 s. Samples here shown are com-

prised of a 35 volume percent solution of colloidal

PMMA of diameter 0.68 µm. These images are from an

experiment at a height of 50 µm with a gap of 150 µm.

The applied shear rate was 0.266 s-1 and the strain was 40.

All scale bars are 5 µm.

Page 5: Shear banding in crystallizing colloidal suspensions · In contrast to equilibrium crystallization, the kinetics of crystallization under non-equilibrium conditions are not well characterized

Shear banding in crystallizing colloidal suspensions

Korea-Australia Rheology Journal December 2010 Vol. 22, No. 4 313

Particles with eight or more crystalline bonds were con-

sidered crystalline (ten Wolde et al., 1996).

To determine the accuracy of the algorithm two samples

that were known to be amorphous and crystalline were

studied. The coordination number distribution and the num-

ber of crystalline bonds distribution were compared for the

two samples. It was found that the coordination number dis-

tributions for the two samples were very similar in shape

and peaked at twelve (data not shown). For the number of

crystalline bonds distribution, the peak in the amorphous

sample was in the range of one to two while for the crys-

talline sample the peak occurred at twelve (data not shown).

3. Results and Discussion

3.1. Homogeneous shear flow of dilute colloidalsuspensions

To establish if any deviations from simple shear flow in

our shear cell device were a consequence of material prop-

erties of the crystallizing suspensions, rather than being

due to possible non-idealities in the flow cell geometry

itself or its alignment, we measured the height-dependent

velocity profile of a dilute colloidal suspensions at shear

rates that spanned the region of interest for colloidal crys-

tallization. The results are plotted in Fig. 4 for measure-

ments of a 2.5 vol% suspension of 0.68 µm PMMA

colloids dispersed in dioctyl phthalate. Pe equals 4, 8 and

12.7 for these measurements. Thus, the experiments are in

a regime in which the displacement of the colloids is dom-

inated by the convective effects of the flow, rather than by

the effects of Brownian motion. The error bars plotted are

the standard error of the mean of four independent mea-

surements. The lines, which are the predicted velocity pro-

file in the gap of the shear flow given the velocity imposed

on the top plate, agree well with the measurements. Thus,

we conclude that the shear cell geometry, its alignment,

and the method to characterize the height dependent veloc-

ity profile are all sufficient to conclude that the base flow is

homogeneous simple shear, absent any effects of the col-

loids as high concentration, as addressed in the next section.

3.2. Inhomogeneous shear flow (shear banding) ofa concentrated, crystallizing colloidal suspension

Fig. 5 reports results of an experiment similar to that of

Fig. 4, except the volume fraction of the colloidal sus-

pension was 35%. This volume fraction was in a regime

that was both concentrated (ie. significant shear thinning

was observed) and above the crystalline phase boundary

(i.e. given many days under quiescent conditions, this sam-

ple underwent the slow process of nucleation and growth

to yield a polycrystalline colloidal crystal). In Fig. 5,

results for Pe = 13, 17 and 23 are reported. Because Pe >

1 for these experiments, they are in a regime in which

Fig. 4. Measured velocity profile for 2.5 volume% colloidal sus-

pensions of PMMA particles. For these experiments, the

top shearing surface, which is displaced at the velocity

shown in the legend is located at height of 150 µm. The

lines are the prediction for simple shear flow based on the

known velocity of the top shearing surface.

Fig. 5. Measured velocity profile for 35 volume% colloidal sus-

pensions of PMMA particles. For these experiments, the

top shearing surface, which is displaced at the velocity

shown in the ledged is located at height of 150 µm. The

lines are the results expected for simple shear flow, as

measured in the case of the dilute colloidal suspensions of

Fig. 4. These measurements were acquired after an

applied strain, γ=10.

Page 6: Shear banding in crystallizing colloidal suspensions · In contrast to equilibrium crystallization, the kinetics of crystallization under non-equilibrium conditions are not well characterized

Laura T. Shereda, Ronald G. Larson and Michael J. Solomon

314 Korea-Australia Rheology Journal

shear flow accelerates crystallization, according both to the

predictions of Ackerson (Ackerson, 1990), as well as our

own measurements, as discussed below. The lines in Fig. 5

are the expected behavior for simple shear flow based on

the velocity imposed on the top plate. For the 35 volume%

suspensions, deviations from simple shear flow were

observed as the plate velocity (and therefore the Peclet

number) was increased. These deviations were localized

near the top and bottom plates of the deformation, and

were roughly symmetric, in agreement with the report of

Wu et al (Wu et al., 2009). The deviations increased with

Pe and were highly significant at the largest Pe number

studied. For example, at the largest plate velocity of

40 µm/s, the velocity at a gap height ~20 µm was about

12 µm/s. This measurement was equivalent to a near wall

shear rate ~0.6 s-1, more than a factor of two greater than

the imposed rate of 0.27 s-1. Fig. 5 thus demonstrates that the

35 vol% suspension exhibited an inhomogeneous, height-

dependent strain rate that is evidence of shear banding.

The Fig. 5 measurements were collected at an applied

strain, ~10. For a colloidal suspension with equilibrium

microstructure, this large amount of strain would be

expected to have yielded a steady-state result. However, in

this case the initially disordered, amorphous microstructure

of the suspension was metastable and the Peclet number of

the imposed deformation was sufficiently large that shear-

induced crystallization was observed. Thus, at the largest

Peclet number (Pe = 23) of the Fig. 5 conditions, we inves-

tigated the effect of the strain deformation on the height-

dependent velocity profile. Measurements were collected

at applied strain deformations of γ = 10, 40 and 160 and

results are plotted in Fig. 6. As in Fig. 5, deviations from

homogeneous simple shear were roughly symmetric about

the mid-plane of the flow (the mid-plane is at 75 µm). The

results show that in all cases the flow could be divided into

three regions – a central region of low strain rate and two

regions of high strain rate adjacent to the top and bottom

surfaces of the flow. Although the precision of the mea-

surements precluded quantitative analysis, the deviations

of the flow from simple shear increased with the applied

strain deformation. The magnitude of the applied strain

deformation for these measurements was too large to be

explained by local distortion of the particle level micro-

structure, as observed in recent simulations (Foss and

Brady, 2000) and experiments (Gao et al., 2010) for dis-

ordered suspensions. However, the question of the micro-

structural origin of this shear banding could be addressed

by confocal microscopy direct visualization, as discussed

in the next section.

3.3. Correlation between shear banding and shear-induced crystallization

At the conditions of the Fig. 6 experiments (γ = 10, 40

and 160; Pe = 23), we illuminated the suspension with UV

light, induced photopolymerization of monomers dissolved

in the solvent, and therefore rapidly immobilized the col-

loidal structure. Previous work has indicated that the time

for immobilization by this method is sufficiently fast that

there is no significant change in the particle-level micro-

structure before and after photopolymerization (Shereda et

al., 2008). After immobilization, the specimens were

imaged in 3D (up to the half-gap of the flow) by confocal

microscopy. Image processing yielded the coordinates of

all the particles as well as an indentification of each par-

ticle as either crystalline or amorphous, based on the bond

order parameters of ten Wolde et al (ten Wolde et al.,

1996). Renderings of the colloidal microstructure are

reported in Fig. 7. Fig. 7 shows that this suspension under-

went shear-induced crystallization and that the progress of

the crystallization was dependent on the magnitude of the

applied strain.

Comparing Figs. 6 and 7 addresses the microstructural

origin of the shear-banded flow discussed in the previous

section. First, the results show that the 35 vol% colloidal

suspension underwent shear-induced crystallization (Fig. 7)

in the same strain deformation range for which shear

banded flow was observed (Fig. 6). Second, just as a near

wall region of high strain rate was observed in the mea-

surements of shear banding, so was a crystalline region

observed in that same near well region. Thus, we infer cor-

relation between the shear banded flow observed by par-

Fig. 6. Strain dependence of the velocity profile measured in the

gap between the shearing surfaces. The data show that the

deviations from simple shear increase with the applied

strain. For all experiments gap=150 µm, shear rate =

0.266 s-1 and Pe = 23.3. The line represents the results for

simple shear flow, as measured in the case of the dilute

colloidal suspensions of Fig. 4.

Page 7: Shear banding in crystallizing colloidal suspensions · In contrast to equilibrium crystallization, the kinetics of crystallization under non-equilibrium conditions are not well characterized

Shear banding in crystallizing colloidal suspensions

Korea-Australia Rheology Journal December 2010 Vol. 22, No. 4 315

ticle tracking and the progress of shear-induced

crystallization in the suspension observed by analysis of

3D image volumes acquired by confocal microscopy.

Based on the comparison of Figs. 5 and 6 with Fig. 7 we

hypothesize that: (i) the viscosity of the crystalline sus-

pension is less than the viscosity of the amorphous, dis-

ordered suspensions, because the shear rate in the

crystalline region of the flow is greater than the shear rate

in the amorphous region; (ii) that on start-up of steady-

shear, the viscosity evolves from a high value characteristic

of the disordered suspension to a low value characteristic

of the crystalline suspension, because the strain require-

ments for crystallization at this gap dimension require

approximately 160 units of strain. These hypotheses were

addressed by cone and plate rheology, as discussed in the

next section.

3.4. Effect of crystallization on suspension rheologyFig. 8 reports the results of cone and plate rheology of

the 35 volume% colloidal suspension. Samples were pre-

sheared at a high rate prior to the tests to insure an initial

state of a metastable, disordered microstructure. The exper-

iments were conducted at an applied shear rate that brack-

ets the region of interest of the direct visualization

experiments reported in Figs. 6 and 7. The rheological

responses, which are each the average of three replications,

were strain (or time) dependent. This transient response

persisted over a much longer strain than could be explained

by structural distortion of the amorphous microstructure.

Such microstructural changes, because they involve struc-

tural distortion (anisotropy) on the order of a particle diam-

eter, are anticipated to result in transient responses of ~ one

strain unit or less. Instead, the transient rheology here

observed required approximately 100 strain units or more

to generate a steady-state response. This strain requirement

was very similar to that found for shear banding (observed

in Fig. 6) and for shear-induced crystallization (observed in

Fig. 7). Moreover, the decrease in viscosity from the initial

state to the steady-state was consistent with the Fig. 6 and 7

measurements. In particular, Fig. 6 showed that the shear-

banded region close to the shearing surfaces was at a

higher shear rate then the region in the center of the flow.

Furthermore, the direct visualization of Fig. 7 showed that

the near wall region was crystalline, and that the steady-

state microstructure of the suspension was also crystalline.

Thus, the different shear rates of the shear banded regions

are consistent with the rheology of the different micro-

structures observed for them.

4. Conclusions

This study has shown that the transient process of shear-

induced crystallization in plane Couette flow is accom-

Fig. 7. Renderings of sheared suspensions showing the crystalline

and amorphous regions after the 35 volume% colloidal

suspension is subjected to different amounts of strain. For

panel (a) strain = 10, panel (b) strain = 40, and panel (c)

strain = 160. The rendering shows the sheared region from

bottom plate to the midplane of the shear flow. The shear

rate of this experiment is 0.266 s-1 (Pe = 23). Green (light

grey) particles are amorphous particles and purple (dark

grey) particles are crystalline particles, as analyzed from

3D confocal microscopy volumes and quantitative image

processing as described in the methods.

Fig. 8. Transient rheology of crystallizing suspensions at colloid

volume fraction of 35%. A cone with a 2o cone angle and

6 cm diameter was used for all experiments. Samples

were shear melted at a shear rate of 100 s-1 for two min-

utes before each experiment. Error bars are standard error

of the mean of three replications.

Page 8: Shear banding in crystallizing colloidal suspensions · In contrast to equilibrium crystallization, the kinetics of crystallization under non-equilibrium conditions are not well characterized

Laura T. Shereda, Ronald G. Larson and Michael J. Solomon

316 Korea-Australia Rheology Journal

panied by shear banded flow. The equilibrium structure of

the 35 vol% suspension studied were in the fully crystal-

line region of the phase diagram. Direct visualization with

confocal microscopy showed that the system progressed

from an initially amorphous, metastable microstructure to a

fully crystalline microstructure. At the gap height studied,

150 µm, this crystallization process required about 160

strain units at Pe = 23. At these same conditions, particle-

tracking measurements of the height-dependent velocity

profile showed that the flow deviated significantly from

simple shear. The local shear rate in the near wall regions

of the flow was as large as a factor of two greater than the

flow in the center of the gap. These differences in the

velocity profile were consistent with the transient rheology

of the crystallizing suspensions, which showed about a fac-

tor of two decrease in the viscosity over a strain interval of

approximately 100 units. The results demonstrate the com-

plex coupling between flow and crystallization in con-

centrated colloidal suspensions. In addition to the effect of

flow accelerating the kinetics of crystallization, the crys-

tallization process itself feedbacks on the local deforma-

tion, thereby generating shear banded flow that is

sensitively linked to the progress of the crystallization.

Acknowledgments

This study was partially supported by the National Sci-

ence Foundation (NSF CBET 0707383). LTS acknowl-

edges support of a GAANN fellowship. The shear cell

schematics were prepared by Eric Viges and Lilian Hsiao.

References

Ackerson, B.J., 1990, Shear induced order and shear processing

of model hard sphere suspensions, J. Rheol. 34, 553.

Anderson, V.J. and H.N.W. Lekkerkerker, 2002, Insights into

phase transition kinetics from colloid science, Nature 416, 811.

Antl, L., J.W. Goodwin, R.D. Hill, R.H. Ottewill, S.M. Owens

and S. Papworth, 1986, The preparation of poly(methyl meth-

acrylate) latices in non-aqueous media, Colloids and Surfaces

17, 67.

Campbell, A.I., and P. Bartlett, 2002, Fluorescent hard-sphere

polymer colloids for confocal microscopy, Journal of Colloid

and Interface Science 256, 325.

Chen, L.B., B.J. Ackerson and C.F. Zukoski, 1994, Rheological

consequences of microstructural transitions in colloidal crys-

tals, J. Rheology 38, 193.

Cohen, I., B. Davidovitch, A.B. Schofield, M.P. Brenner and

D.A. Weitz, 2006, Slip, yield, and bands in colloidal crystals

under oscillatory shear, Phys. Rev. Lett. 97, 215502.

Crocker, J.C. and D.G. Grier, 1996, Methods of digital video

microscopy for colloidal studies, Journal Colloid Interface Sci-

ence 179, 298.

Dibble, C.J., M. Kogan and M.J. Solomon, 2006, Structure and

dynamics of colloidal depletion gels: Coincidence of transi-

tions and heterogeneity, Physical Review E 74, 041403.

Foss, D.R. and J.F. Brady, 2000, Structure, diffusion and rhe-

ology of Brownian suspensions by Stokesian Dynamics sim-

ulation, Journal of Fluid Mechanics 407, 167.

Gao, C., S.D. Kulkarni, J.F. Morris and J.F. Gilchrist, 2010,

Direct investigation of anisotropic suspension structure in pres-

sure-driven flow, Physical Review E 81, 041403.

Gasser, U., 2009, Crystallization in three- and two-dimensional

colloidal suspensions, Journal of Physics - Condensed Matter

21, 203101.

Gasser, U., E.R. Weeks, A. Schofield, P.N. Pusey and D.A.

Weitz, 2001, Real-space imaging of nucleation and growth in

colloidal crystallization, Science 292, 258.

Holmqvist, P., M.P. Lettinga, J. Buitenhuis and J.K.G. Dhont,

2005, Crystallization kinetics of colloidal spheres under sta-

tionary shear flow, Langmuir 21, 10976.

Kogan, M., C.J. Dibble, R.E. Rogers and M.J. Solomon, 2008,

Viscous solvent colloidal system with well-characterized inter-

actions for direct visualization of suspension structure, Journal

Colloid and Interface Science 318, 252.

Pathmamanoharan, C., C. Slob and H.N.W. Lekkerkerker, 1989,

Preparation of poly(methyl methacrylate) latices in non-polar

media, Colloid Polym Sci 267, 448.

Royall, C.P., M.E. Leunissen and A. van Blaaderen, 2003, A new

colloidal model system to study long-range interactions quan-

titatively in real space, Journal of Physics: Condensed Matter

15, S3581.

Shereda, L.T., R.G. Larson and M.J. Solomon, 2008, Local stress

control of spatiotemporal ordering of colloidal crystals in com-

plex flows, Phys. Rev. Lett. 101, 038301.

Solomon, T. and M.J. Solomon, 2006, Stacking fault structure in

shear-induced colloidal crystallization, J. Chem. Phys. 124, art.

no. 134905.

ten Wolde, P.R., M.J. Ruiz-Montero, and D. Frenkel, 1996,

Numerical calculation of the rate of crystal nucleation in a Len-

nard-Jones system as moderate undercooling, J. Chem. Phys.

104, 9932.

Vermant, J. and M.J. Solomon, 2005, Flow-induced structure in

colloidal suspensions, J. Phys.-Condes. Matter 17, R1.

Wu, Y.L., D. Derks, A. van Blaaderen, and A. Imhof, 2009, Melt-

ing and crystallization of colloidal hard-sphere suspensions

under shear, Proceedings of the National Academy of Sciences

of the United States of America 106, 10564.