M. Laurati and S. U. Egelhaaf
Condensed Matter Physics Laboratory, Heinrich-Heine University, Universitätsstraße 1, 40225 Düsseldorf, Germany
G. Petekidisa)
Department of Materials Science and Technology and IESL-FORTH, University of Crete, Heraklion, Greece
(Received 23 July 2010; final revision received 30 November 2010; published 4 April 2011兲
Synopsis
The depletion attraction, induced upon addition of a nonadsorbing polymer to a colloidal solution, can lead to gel formation at sufficiently high polymer concentrations, which corresponds to strong attractive interactions. We have investigated the nonlinear rheological response, in particular the yielding, of colloidal gels with an intermediate volume fraction and variable interparticle attraction. Two distinct yielding processes are observed in both oscillatory experiments, namely, dynamic strain sweeps and transient experiments, here step rate, creep, and recovery tests. The first yielding process occurs at strains similar to the range of the interparticle potential and is interpreted as the breaking of bonds, which destroys the particle network and leads to individual clusters. The process of bond breaking is successfully modeled as the escape of a particle from the potential well of its nearest neighbor. The second yield point occurs at larger strains and is related to the deformation and fragmentation of clusters, consistent with the observed dependence of the yield strain on attraction. Both yield stresses exhibit a power-law dependence on attraction strength in agreement with observations of other systems and theoretical predictions. Furthermore, the observed two-step yielding reveals similarities, and also differences, to the rheology of attractive colloidal glasses. © 2011 The Society of Rheology. 关DOI: 10.1122/1.3571554兴
I. INTRODUCTION
Viscoelastic materials are frequently subjected to large deformations as evidenced in many industrial processes and in our daily lives when eating food or using personal care products 关Coussot 共2005兲; Larson共1999兲兴. Upon an applied deformation or stress, many viscoelastic materials respond elastically, i.e., solidlike, below a certain strain or stress. Beyond this characteristic yield strain or stress the response is dissipative, i.e., fluidlike. This behavior is known as yielding and occurs in very different materials, ranging from complex fluids to amorphous metals 关Becu et al. 共2006兲; Brader et al. 共2007, 2008兲;Carrier and Petekidis 共2009兲; Crassous et al.共2008兲; Fuchs and Ballauff
a兲Electronic mail: [email protected]
© 2011 by The Society of Rheology, Inc.
673 J. Rheol. 55共3兲, 673-706 May/June 共2011兲 0148-6055/2011/55共3兲/673/34/$30.00
共2005兲;Gopalakrishnan and Zukoski共2007兲;Johnson and Samwer共2005兲;Le Grand and Petekidis 共2008兲; Miyazaki and Reichman 共2002兲; Miyazaki et al. 共2006兲; Pham et al. 共2006, 2008兲; Rao et al. 共2006兲; Rottler and Robbins 共2003兲; Rouyer et al. 共2003兲; Siebenburger et al. 共2009兲; Stokes and Frith 共2008兲; Walls et al. 共2003兲兴. A thorough understanding of yielding, however, is still missing.
The yielding of dynamically arrested states, namely, colloidal glasses and gels, has recently attracted particular interest. In suspensions of colloidal hard spheres with volume fractions ⱖ0.58, an amorphous solid state, a repulsive glass, has been found 关Pusey 共1991兲;Pusey and van Megen共1986,1987兲兴. If large enough interparticle attractions are present, amorphous solid states can be formed at any volume fraction: Attractive glasses at high colloid volume fractions关Dawson et al.共2001兲;Eckert and Bartsch共2002兲;Pham
et al. 共2002, 2004兲 ; Poon et al. 共2003兲兴 are usually distinguished from gels at lower volume fractions关Poon共2002兲;Shah et al.共2003a兲兴. The location of the transition from a fluid to an attractive glass or gel can be predicted by mode-coupling theory共MCT兲 and is consistent with experiments关Bergenholtz et al. 共2003兲;Laurati et al.共2009兲;Shah et
al.共2003b兲兴. It has been suggested that dynamical arrest in these two systems is driven by
different mechanisms: bonding in attractive glasses as described by MCT关Cates共2003兲; Dawson et al. 共2001兲;Pham et al.共2002兲兴 and phase separation in low volume fraction gels关Cardinaux et al.共2007兲;Ilett et al.共1995兲;Lu et al.共2008兲;Manley et al.共2005兲; Poon et al. 共1997兲; Verduin and Dhont 共1995兲; Verhaegh et al. 共1997兲兴. However, the formation of gels has also been described as a glass transition of clusters by a MCT-based approach 关Kroy et al.共2004兲兴. This suggests that the mechanical properties of gels are governed by the cluster length scale, and those of attractive glasses are governed by the particle size 关Zaccone and Del Gado共2009兲兴. The distinction between attractive glasses and gels becomes, however, difficult to describe at intermediate volume fractions.
The rheology of repulsive and attractive glasses has been studied both experimentally 关Conrad et al.共2010兲;Crassous et al.共2008兲;Fuchs and Ballauff共2005兲;Miyazaki et al. 共2006兲; Petekidis et al.共2004兲;Pham et al. 共2006, 2008兲; Shah et al.共2003b兲;Varadan and Solomon 共2003兲兴 and theoretically 关Brader et al.共2007,2008兲; Fuchs and Ballauff 共2005兲;Fuchs and Cates共2002兲;Miyazaki and Reichman共2002兲;Miyazaki et al.共2006兲; Sollich共1998兲;Zaccone and Del Gado共2009兲兴. Repulsive glasses show a single yielding process 关Petekidis et al. 共2002a, 2003兲; Pham et al. 共2008兲兴. This has been associated with the accelerated decay of structural correlations due to shear advection by MCT 关Brader et al. 共2007兲; Crassous et al. 共2008兲; Fuchs and Cates 共2002兲; Miyazaki and Reichman共2002兲;Miyazaki et al.共2006兲兴 and with particle delocalization due to hopping processes by trap models关Dyre共1987兲;Monthus and Bouchaud共1996兲;Sollich共1998兲兴. These two approaches have also been combined关Kobelev and Schweizer共2005b兲兴: The nonergodicity transition of MCT is used to determine the onset of particle localization due to the emergence of finite free-energy barriers in the energy landscape. The impor-tance of nonlinear contributions in achieving a comprehensive theoretical understanding of the stress-strain relation from nonlinear rheology of dense colloidal suspensions under time-dependent shear has recently been recognized关Brader et al.共2007,2008兲;Fuchs and Cates共2002兲兴, while experimental evidence exists both from steady and oscillatory mea-surements in a variety of systems关Carrier and Petekidis共2009兲; Crassous et al.共2008兲; Gadala-Maria and Acrivos共1980兲;Helgeson et al.共2007兲;Hyun et al.共2007兲;Le Grand and Petekidis 共2008兲;Siebenburger et al.共2009兲兴.
Attractive glasses present an even more complex yielding behavior with two distinct yielding processes 关Pham et al. 共2006, 2008兲兴. These have been associated with the breaking of interparticle bonds at small strains and with the limit of elastic cage defor-mations at large strains. To date, there is no theoretical description of these processes.
Some effort has also been devoted to the study of the nonlinear rheology of colloidal gels, which include gels formed due to van der Waals attraction关Buscall et al.共1987兲;Furst and Pantina 共2007兲; Kobelev and Schweizer共2005a兲;Larson 共1999兲;Russel and Grant 共2000兲兴 or depletion interactions 关Kobelev and Schweizer共2005a兲;Koumakis and Peteki-dis 共2008兲兴 and also thermoreversible gels 关Gopalakrishnan and Zukoski 共2007兲; Wout-ersen and de Kruif共1991兲兴. For strongly flocculated gels, a single yielding process with a weak dependence on volume fraction and attraction strength is observed关Buscall et al. 共1987兲; Larson共1999兲; Russel and Grant 共2000兲兴, while thermoreversible gels show a stronger dependence on volume fraction 关Gopalakrishnan and Zukoski 共2007兲兴, which agrees with the combined MCT-barrier hopping approach 关Kobelev and Schweizer 共2005a兲兴. In contrast, two yielding processes, as in attractive glasses, are found in deple-tion gels with intermediate volume fracdeple-tion with a strong interplay as volume fracdeple-tion is decreasing. The presence of structural heterogeneities and clusterlike structures over a broad range of intermediate volume fractions, as observed by scattering and microscopy 关Dibble et al. 共2006兲;Laurati et al. 共2009兲; Shah et al.共2003c兲; Smith共2004兲;Varadan and Solomon共2003兲兴, suggests that an important role is played by structural heterogene-ities in the yielding of these systems.
Here, we investigate the nonlinear rheological properties, in particular the yielding, of depletion gels with intermediate volume fractions as a function of attraction strength. The strength of the attraction is varied by changing the concentration of the depletant, a nonadsorbing polymer. Changes in the interparticle interactions induce changes in the structure and dynamics, in particular in the size of heterogeneities, which have been determined by scattering and microscopy 关Dibble et al. 共2006兲; Laurati et al. 共2009兲; Shah et al.共2003c兲;Smith et al.共2007兲;Varadan and Solomon共2003兲兴. The dependence of the size of heterogeneities on polymer concentration, which shows a maximum around the gelation boundary 关Dibble et al.共2006兲; Laurati et al.共2009兲兴, suggests a nonequi-librium route leading to gelation at intermediate volume fractions关Lu et al.共2008兲兴. We investigate the effect of these structural and dynamic changes on shear-induced rear-rangements and yielding by rheology, namely, dynamic strain sweeps 共DSSs兲, step rate tests, flow curves, and creep and recovery experiments. Both the yield strain and stress were measured under different conditions and the data compared to theoretical predic-tions关Kobelev and Schweizer 共2005a兲;Smith et al.共2007兲兴. These results together with investigations of attractive glasses at large volume fractions关Pham et al.共2006,2008兲兴 and gels with a constant attraction strength but with variable volume fractions关Koumakis and Petekidis 共2008,2011兲兴 provide a comprehensive mapping of nonlinear rheology of attractive glasses and gels.
II. MATERIALS AND METHODS A. Samples
The samples contain mixtures of polymethylmethacrylate 共PMMA兲 hard-sphere col-loids and linear polystyrene共PS兲 dispersed in cis-decalin at a temperature T=23 °C. The average hydrodynamic radius of the PMMA particles, R = 137 nm, was determined in very dilute samples by dynamic light scattering. The absence of crystallization in quies-cent and sheared samples indicates a colloid polydispersity of at least 12%. The radius of gyration of PS 共molecular weight Mw= 132 900 g/mol, from Polymer Laboratories兲 in
cis-decalin is rg= 10.8 nm 关following Berry 共1966兲兴 and the polydispersity Mw/Mn = 1.01 as specified by the supplier. This implies a polymer-colloid size ratio = rg/R = 0.079⫾0.013 in dilute solution. The effective polymer-colloid size ratio ⴱ has been calculated at a colloid volume fraction = 0.4 and different polymer concentrations
0.0⬍cp/cpⴱⱕ2.0, where cpⴱ is the overlap concentration, taking into account the concen-tration dependence of the polymer size and the mesh size using the generalized free volume theory共GFVT兲 关Aarts et al.共2002兲;Fleer and Tuinier共2007兲;Lekkerkerker et al. 共1992兲兴 共TableI兲. The calculations are based on polymers in a good solvent, although our conditions approach those of asolvent, and thus might slightly overestimate the values of ⴱ. Furthermore, the Asakura–Oosawa 共AO兲 description of the depletion potential is not strictly valid around and beyond the polymer overlap concentration cpⴱ 关Fuchs and Schweizer共2002兲兴. The values of the depth of the attraction potential U0 for large poly-mer concentrations, reported in TableI, have to be considered therefore as indicative.
The samples were prepared by mixing appropriate amounts of colloid and polymer stock solutions. The colloid stock solution with volume fraction= 0.6 was obtained by redispersing spun-down sediments, whose volume fraction was estimated to be= 0.67 when taking polydispersity into account 关Schaertl and Silescu 共1994兲兴. Polymer stock solutions of different concentrations were prepared by adding dry polymer to cis-decalin with the concentration cp共mass/volume兲 calculated from the weighed masses of the two components. After mixing, samples were vigorously shaken using a vortex shaker and then homogenized for 3 days in a rotating-wheel mixer. The compositions of the samples are summarized in Table I, with the first column listing the nominal polymer concentra-tions, which are used to refer to the samples in the following discussion. Values of the volume fractionand the polymer concentration cprefer to the total volume and cpfreeto the volume not occupied by colloids, as estimated by GFVT关Aarts et al. 共2002兲;Fleer and Tuinier共2007兲;Lekkerkerker et al. 共1992兲兴. Polymer concentrations are normalized by the polymer overlap concentration cⴱp= 3Mw/4NArg3, where NAis Avogadro’s number. B. Rheology
Strain-controlled rheological measurements were performed on an ARES-HR strain-controlled rheometer with a force balance transducer 10FRTN1 mainly using a cone-plate geometry共cone angle of 0.044 rad and cone diameter of 25 mm兲. Homemade mechani-cally roughened cone-plate fixtures共typically with a cone angle of 0.044 and a diameter of 25 mm and roughness of few hundred microns兲, calibrated with viscoelastic standards,
TABLE I. Investigated samples, which are referred to by their nominal polymer concentration.is the colloid
volume fraction; cp/cpⴱand cp
free/c
p
ⴱare the polymer concentrations in the total and free volume, respectively,
normalized by the overlap concentration cpⴱ;ⴱis the effective polymer-colloid size ratio; and U0is the depth of
the attractive potential.
cp/cpⴱ共nominal兲 cp/cpⴱ cfreep /cⴱp ⴱ U0/kBT 0 0.40 0.1 0.40 0.10 0.21 0.079⫾0.013 ⫺5.0 0.2 0.40 0.20 0.40 0.067⫾0.009 ⫺10.2 0.25 0.40 0.25 0.49 0.063⫾0.008 ⫺12.7 0.32 0.39 0.32 0.62 0.057⫾0.007 ⫺16.3 0.4 0.40 0.40 0.76 0.052⫾0.007 ⫺20.5 0.5 0.41 0.48 0.90 0.048⫾0.006 ⫺24.6 0.7 0.40 0.70 1.28 0.040⫾0.006 ⫺36.0 0.8 0.40 0.82 1.48 0.037⫾0.005 ⫺42.1 1 0.40 0.99 1.78 0.033⫾0.005 ⫺51.1 1.5 0.40 1.49 2.63 0.027⫾0.004 ⫺77.5 2 0.40 1.99 3.48 0.023⫾0.004 ⫺104.2
were also used to avoid wall slip 共Sec. III B兲. Stress-controlled measurements were carried out using a DSR-200 rheometer共Rheometric Scientific, currently TA兲 with sand-blasted cone-plate geometry共cone angle of 0.04 rad and cone diameter of 50 mm兲, which was coated with a mixture of small and large PMMA spheres共radii of about 150 nm and 1 m, respectively兲 by sintering to obtain an irregular surface. 共For the stress-controlled rheometer mechanically roughened geometries were not available.兲 This procedure has been shown to prevent wall slip in concentrated hard-sphere dispersions关Ballesta et al. 共2008兲兴. Solvent evaporation was minimized by using a solvent saturation trap, which is designed to isolate the sample from the surrounding atmosphere by a fluid seal at the top and a permanent seal at the bottom. Inside the enclosure, solvent evaporation leads to a saturated atmosphere.
The effects of sample loading and aging were reduced by performing a standard rejuvenation procedure before each test: Directly after loading, we performed a dynamic strain or stress sweep 共for strain or stress-controlled measurements, respectively兲, i.e., we applied oscillatory shear to the samples with a frequency= 1 rad/s and an increas-ing strain amplitude ␥ 共or stress amplitude 兲 until the sample showed a liquidlike response. In strain-controlled experiments, a strain amplitude␥= 8 was large enough for all polymer concentrations, while in stress-controlled measurements stress amplitudes of 15ⱕⱕ60 Pa were used depending on the polymer concentration, 0.4ⱕcp/cpⴱⱕ1.0. Samples were then subjected to a dynamic time sweep with a frequency= 1 rad/s and a strain ␥= 8 共or a stress 15ⱕⱕ60 Pa for stress-controlled measurements兲 until the elastic, G
⬘
, and viscous, G⬙
, moduli reached constant steady-state values. Subsequently, the samples were left at rest for a waiting time tw= 300 s before the test was started. This waiting time was chosen to ensure a reproducible initial state, on the basis of the time dependence共aging兲 of the linear viscoelastic properties after rejuvenation.III. RESULTS AND DISCUSSION A. Tube inversion
The macroscopic behavior of the samples was studied by observing their flow behav-ior upon tube inversion. All samples had colloid volume fractions of⬇0.4, but different polymer concentrations cp/cpⴱ, where cpⴱis the polymer overlap concentration共TableIand Fig.1兲. Gel samples were identified initially by the absence of flow after tube inversion.
2.0 1.5 1.0 0.5 0.0
c
p/
c
p*
0.50 0.45 0.40 0.35 0.30 gelation boundary appearance of wall slipFIG. 1. Results from tube inversion of samples with similar volume fraction⬇0.4 but different polymer
concentrations cp/cpⴱ. Samples exhibiting no flow共gels兲 共쎲兲, high viscosity fluids 共䊏兲, and low viscosity fluids 共䉭兲.
This was observed for samples with cp/cⴱpⱖ0.4. Samples with 0.2⬍cp/cpⴱ⬍0.4 still flowed upon inversion but exhibited high viscosities, while samples with cp/cⴱpⱕ0.2 showed low viscosities.
B. Wall slip
The response using smooth and mechanically roughened geometries to dynamic strain sweeps共DSS兲 is shown in Fig.2. A comparison of the results for samples with different polymer concentrations cp/cⴱp at a fixed frequency= 1 rad/s Fig. 2共a兲 shows that for
cp/cpⴱ= 0.4, comparable results are obtained with different degrees of surface roughness, while markedly different results are obtained for cp/cⴱp= 1.0 in the nonlinear regime. Indeed, for cp/cⴱp= 1.0 the storage modulus G
⬘
is smaller at all strains␥ when smooth surfaces are used. In addition, the strain dependences of both the elastic, G⬘
, and viscous,G
⬙
, moduli are qualitatively different. One particular difference when using smooth sur-faces is in G⬘
; a minimum is evident at␥⬇30% followed by a maximum at␥⬇100%. These features disappear when roughened geometries are used. For cp/cpⴱ= 1.0, DSS measurements performed at different frequencies关Fig.2共b兲兴 indicate quantitative dif-ferences for different surface roughness with the difdif-ferences in the qualitative trends becoming more pronounced at lower frequencies. Such discrepancies between the re-sponse using smooth and roughened geometries have been attributed to slip at the smooth10-1 100 101 102 103 G ',G '' [Pa] 100 101 102 G ',G '' [Pa] 10-1 100 101 102 103
γ [%]
(a) (b)FIG. 2. Comparison of DSS measurements performed with roughened共symbols兲 and smooth 共lines兲 cone-plate
geometries. Full symbols and solid lines represent the elastic modulus G⬘, while open symbols and dotted lines represent the viscous modulus G⬙.共a兲 Dependence on polymer concentration 关cp/cpⴱ= 0.4共쎲兲 and 1.0 共䉱兲兴 for a fixed frequency= 1 rad/s. 共b兲 Dependence on frequency 关= 1 rad/s 共䉱兲 and 100 rad/s 共⽧兲兴 for a fixed polymer concentration cp/cpⴱ= 1.0.
surfaces关Walls et al.共2003兲兴, while the frequency dependence is in qualitative agreement with findings in colloidal glasses under steady shear where slip is observed at low rates 关Ballesta et al.共2008兲兴. Responses similar to those shown for cp/cⴱp= 1.0 were measured for samples with cp/cpⴱⱖ0.7. Slip during oscillatory measurements was also detectable either by the existence of even harmonics in the stress signal 共typically around 2%兲 or by microscopic dynamics measured with the light-scattering echo 共LS-echo兲 technique 关Petekidis et al.共2002a,2002b兲兴, as will be reported in detail elsewhere.
Wall slip thus strongly influences the measurements at large strains in the nonlinear regime for samples with polymer concentrations cp/cpⴱⱖ0.7 共as indicated in Fig.1兲, but has little effect on the linear rheological response at small strains. All measurements reported here共and previously performed to investigate the linear response 关Laurati et al. 共2009兲兴兲 have been done with roughened surfaces to minimize wall slip. As our experi-ments indicate, the results with roughened and smooth surfaces agree where wall slip can be excluded.
Besides wall slip, shear banding has been observed in colloidal glasses under steady shear关Besseling et al.共2010兲兴 and colloidal crystals under oscillatory shear 关Cohen et al. 共2006兲兴. A direct detection of such phenomena requires monitoring the velocity profile during shear关Ravindranath and Wang共2008兲;Tapadia et al.共2006兲兴. Preliminary confo-cal microscopy experiments under steady shear on gels similar to the ones investigated here have revealed shear banding only at quite low shear rates and high attraction strengths关Ballesta et al. 共in preparation兲兴. However, no oscillatory shear experiments are presently available. An indication for the absence of nonperiodic shear banding effects during large amplitude oscillatory shear共LAOS兲 was obtained by Fourier transform 共FT兲 analysis of the stress response, where even harmonics were absent 关Laurati et al. 共in preparation兲兴. Hence, although shear banding cannot be entirely excluded, indirect evi-dence suggests that it is absent in the current LAOS measurements.
C. Linear rheological response
Linear viscoelastic measurements presented before in the same system关Laurati et al. 共2009兲兴 are briefly summarized here to serve as a reference for the discussion of the nonlinear viscoelastic properties. Dynamic frequency sweep 共DFS兲 tests have been per-formed with samples below and above the macroscopic gelation boundary.
Below the macroscopic gelation boundary关Fig.3共a兲兴, the crossing points between the elastic共storage兲, G
⬘
, and viscous共loss兲, G⬙
, moduli shift toward smaller frequencies upon increasing the polymer concentration cp/cⴱp. This indicates a transition from a liquidlike to a solidlike response which occurs at increasingly smaller frequencies共corresponding to larger characteristic times兲 as the polymer concentration, i.e., interparticle attraction, is increased. The characteristic time is related to the “bond” lifetime estimated by the time needed to escape from the attractive potential well. Bond breaking was found to be the dominant process close to but still below the gelation boundary, where the samples already have a network structure with a finite lifetime共transient network兲, while particle or cluster diffusion seems to control structural relaxation at the lowest polymer concen-trations 关Laurati et al. 共2009兲兴. In this regime, i.e., below the gelation boundary, MCT successfully describes the dependence of the elastic modulus G⬘
on polymer concentra-tion 关Fig. 3共c兲, blue line兴, while there are discrepancies in the frequency dependence 关Laurati et al. 共2009兲兴.In the gel region关Fig.3共b兲兴, solidlike behavior is observed, with G
⬘
being larger thanG
⬙
for all investigated frequencies and weakly frequency-dependent. The frequency de-pendence of G⬙
displays a minimum, indicating structural relaxation of the gels at long10-2 10-1 100 101 102 G '( ), G '' ( ) [Pa] 101 102 103 G '( ), G '' ( ) [Pa] 10-2 10-1 100 101 102 ω [rad/s] 10-2 10-1 100 101 102 103 G '[Pa] 2.0 1.8 1.6 1.4 1.2 1.0 0.8 0.6 0.4 0.2 0.0 cp/cp*
(c)
I
II III
(a)
(b)
FIG. 3. DFSs with elastic共storage兲 modulus G⬘共filled symbols兲 and viscous 共loss兲 modulus G⬙共open symbols兲
as a function of frequency. Samples共a兲 below and 共b兲 above the macroscopic gelation boundary 共cp/cpⴱ = 0.4兲 are shown with polymer concentrations cp/cⴱp= 0共⽧兲, 0.1 共䉲兲, 0.2 共⽧兲, 0.25 共䉱兲, 0.32 共䊏兲, 0.4 共쎲兲, 0.5 共⽧兲, 0.7 共䉲兲, 0.8 共⽧兲, 1.0 共䉱兲, 1.5 共䊏兲, and 2.0 共쎲兲. 共c兲 Elastic modulus G⬘extracted from DFS as a function of polymer concentration cp/cpⴱat frequency= 10 rad/s 共䊏兲. Dashed lines separate the different states: fluid 共I兲, transient network 共II兲, and gel 共III兲. Solid lines represent theoretical predictions for the elastic modulus of transient networks共blue line,Laurati et al., 2009兲 and gels 共red line,Zaccone and Del Gado, 2009兲.
times. The almost linear dependence of G
⬘
on polymer concentration in this region关Fig. 3共c兲, region III兴 was attributed to the dominant contribution of intercluster links; this was theoretically confirmed recently关Zaccone and Del Gado共2009兲兴 关Fig.3共c兲, red line兴. The intermediate volume fraction gel was modeled as a dispersion of clusters. The intracluster bonds were assumed to be permanent, and the elastic modulus of the gel thus determined by the intercluster elastic constant was calculated on the basis of the interaction potential as well as the coordination number, effective volume fraction, and characteristic size of the clusters关Zaccone and Del Gado 共2009兲兴.D. Nonlinear rheological response
1. DSSs
DSSs at a frequency = 1 rad/s were performed with samples below and above the macroscopic gelation boundary关Figs.4共a兲and4共b兲, respectively兴. The moduli measured in DSS tests represent the average response of the samples. The detailed nonlinear re-sponse can be accessed by analysis of the higher order harmonics and allows for a more detailed understanding of the intracycle yielding mechanisms 关Ewoldt et al. 共2008兲;
100 101 102 103
G
',G
''
[Pa]
10-1 100 101 102 103γ [%]
1.8 0.9 10-2 10-1 100 101 102G
',G
''
[Pa]
1.6 0.8(a)
(b)
10-2 10-1 101 102 103FIG. 4. DSSs with elastic共storage兲 modulus G⬘共filled symbols兲 and viscous 共loss兲 modulus G⬙共open symbols兲
as a function of strain␥at= 1 rad/s. Samples 共a兲 below and 共b兲 above the macroscopic gelation boundary 共cp/cpⴱ= 0.4兲 are shown with polymer concentrations cp/cpⴱ= 0.1共䉲, inset兲, 0.2 共⽧, inset兲, 0.25 共䉱兲, 0.32 共䊏兲, 0.4共쎲兲, 0.5 共⽧兲, 0.7 共䉲兲, 0.8 共⽧兲, 1.0 共䉱兲, 1.5 共䊏兲, and 2.0 共쎲兲.
Renou et al.共2010兲兴. Such an analysis, performed on a selection of the samples and well inside the nonlinear regime, revealed amplitudes of the third harmonic as high as 20% and will be the subject of a forthcoming publication. Samples with polymer concentra-tions cp/cpⴱ= 0.1 and 0.2 show a fluidlike response in agreement with the linear viscoelas-tic data 关Fig.3共a兲兴, with shear thinning behavior at high strains. For cp/cpⴱ= 0.25, G
⬘
is smaller than G⬙
at large strains, but becomes very similar below␥⬇2%, i.e., in the linear regime. The response is thus intermediate between a highly viscous fluid and a low yield-stress solid and may also be used to determine the liquid-gel transition. This is attributed to a transient network structure, which is also consistent with its linear response 关Laurati et al.共2009兲兴. At frequencies⬎1 rad/s 共not shown兲, G⬘
becomes larger thanG
⬙
in the linear regime, in agreement with DFS关Fig.3共a兲, cp/cⴱp= 0.25兴.For larger polymer concentrations 共cp/cpⴱⱖ0.32兲, G
⬘
is larger than G⬙
in the linear regime, and thus samples behave solidlike at low frequencies. Entering the nonlinear regime, the shear moduli decrease with increasing strain, with G⬘
showing a stronger decrease than G⬙
, which leads to a crossover between G⬘
and G⬙
at strains␥ⲏ10%. Such a crossover in LAOS is often interpreted as a yielding point since at strains beyond the crossover many materials, for example, hard-sphere glasses关Derec et al.共2003兲;Mason and Weitz共1995兲;Miyazaki et al.共2006兲;Pham et al.共2006,2008兲兴 and other soft glassy materials关Carrier and Petekidis 共2009兲;Cloitre et al.共2000兲; Daniel et al.共2001兲; Hel-geson et al.共2007兲;Hyun et al. 共2002兲;Le Grand and Petekidis共2008兲;Sollich共1998兲; Wyss et al. 共2007兲兴, show a fluidlike response with G⬘
⬃␥0 and G⬙
⬃␥1 where 0 ⬇21. In the present case, however, such a fluidlike response is not observed immedi-ately after the crossover. Instead, over a broad range of strains the decay of the two moduli is weaker, indicating that some residual structure remains under flow. Only for ␥⬎100% is a strain regime with 0⬇21 observed with −1.8ⱕ0ⱕ−1.4. Such a re-sponse indicates the existence of two different yielding mechanisms.
DSSs show a similar frequency dependence for all polymer concentrations investi-gated; Fig. 5 shows data for cp/cpⴱ= 0.7. Both G
⬘
and G⬙
increase with increasing fre-quencyin the linear and nonlinear regimes 共except the 1 and 10 rad/s values of G⬙
in the linear regime, which show the opposite trends but are very similar兲. The strain dependence of G⬘
and G⬙
does not change significantly with frequency, except for strains100 101 102 G ',G '' [Pa] 10-1 100 101 102 103 γ [%]
FIG. 5. DSSs with elastic共storage兲 modulus G⬘共filled symbols兲 and viscous 共loss兲 modulus G⬙共open symbols兲
as a function of strain␥for a polymer concentration cp/cpⴱ= 0.7 and frequencies= 1 rad/s 共䉲兲, 10 rad/s 共쎲兲, 69 rad/s共䊏兲, and 100 rad/s 共⽧兲.
␥⬎100% where, with increasing frequency, both moduli show a weaker decay prior to the final shear thinning regime 共Fig.5兲. The two yielding processes thus remain for all frequencies.
An alternative representation of the DSS data is to plot the strain dependence of the total stress and its elastic component G
⬘
␥ 共Fig. 6兲. In this representation the two yielding mechanisms inferred from the trends of the shear moduli in Fig. 4 can be quantitatively determined as the inflection points in the strain dependence of the total stress共␥兲, at which the stress-strain relation becomes increasingly sublinear 关Fig.6共a兲兴. The maxima or shoulders of the elastic component of the stress G⬘
␥关Fig.6共b兲, arrows兴 provide an alternative way of locating the yield points, as has been done for other soft matter systems 关Kobelev and Schweizer 共2005a兲; Pai and Khan 共2002兲; Yang et al. 共1986兲兴. Both 共␥兲 and G⬘
␥共␥兲 support the presence of two yielding processes for samples with polymer concentrations cp/cpⴱⱖ0.32.Two yielding processes have also been observed for attractive glasses 关Pham et al. 共2006,2008兲兴. In this case, the yielding mechanisms have been identified from distinct maxima of G
⬙
, i.e., maxima of energy dissipation, and are associated with bond breaking 共at low strains of about 5%兲 and the limit of elastic cage distortions 共at large strains of about 50%兲. For the gels studied here, the first yield point can also be associated with10-2 10-1 100 101 102
σ
[Pa]
10-1 100 101 102G
'γ
[Pa]
10-1 100 101 102 103γ [%]
(a)
(b)
FIG. 6. DSSs represented as 共a兲 total stress and 共b兲 its elastic component G⬘␥ as a function of strain
␥, which were performed at a frequency = 1 rad/s for different polymer concentrations cp/cpⴱ = 0.25, 0.32, 0.4, 0.5, 0.7, 0.8, 1.0, 1.5, 2.0共from bottom to top, same data and symbols as in Fig.4兲. Yield points
共indicated by arrows兲 are identified from 共a兲 the inflection points in the stress response, located through the crossing points of the corresponding tangents, and共b兲 the maxima or shoulders in the elastic stress-strain curve.
bond breaking, where the absence of a maximum at low polymer concentrations is con-sistent with the low strength of the bonds. The second weaker yield point, in contrast to glasses, relates to a more gradual restructuring resulting in a fluidized state. In the fol-lowing, we compare this suggestion with our experimental data to test for consistency and to extract further information.
We now consider the dependence of the two yield points, as well as the crossing point between G
⬘
and G⬙
, on the polymer concentration cp/cpⴱ 共or depletion attraction兲 and frequencyand discuss the corresponding yield strains and yield stresses共Figs.7and8兲. The first yield point reported 共␥1,1兲 corresponds to the onset of the sublinear stress-strain relation in the stress-stress-strain plot关Fig.6共a兲兴. We chose these values as they are better determined than the values obtained from the maximum or the onset of monotonous120 100 80 60 40 20 0
γ
X [%] 5 4 3 2 1 0γ
1 [%] 400 300 200 100 0γ
2 [% ] 2.0 1.5 1.0 0.5 0.0c
p/ c
p*
(a) (b) (c)FIG. 7. Yield strains at共a兲 the first yield point,␥1;共b兲 the crossing point,␥X; and共c兲 the second yield point,␥2,
as a function of polymer concentration cp/cpⴱfor different frequencies= 1 rad/s 共䊊兲, 10 rad/s 共䊐兲, 40 rad/s 共䉮兲, 69 rad/s 共䉭兲, and 100 rad/s 共〫兲. The yield strains have been extracted from the DSSs with 共a兲 the first yield point estimated as the onset of sublinear stress-strain relation,共b兲 the crossing point extracted from the strain dependence of G⬘and G⬙, and共c兲 the second yield point identified with the second inflection point of the stress-strain curve.
decrease of G
⬙
共Fig. 4兲, and the maximum or shoulder of the elastic stress G⬘
␥ 关Fig. 6共b兲兴, which show a comparable dependence on cp/cpⴱwith only small differences in the absolute values. In comparison, the second yield point 共␥2,2兲 is identified with the second inflection of the stress-strain curve 关Fig. 6共a兲兴. The crossing point 共␥X,X兲 is extracted from the strain dependence of G⬘
and G⬙
, at G⬘
= G⬙
共Fig.4兲.The first yield strain ␥1 increases in value with increasing polymer concentration 关Fig. 7共a兲兴. At a low polymer concentration of cp/cpⴱ= 0.25, it ranges, depending on frequency , from 2.3%ⱗ␥1ⱗ3.7% and then increases to 4.2%ⱗ␥1ⱗ5.2%, with the saturation value reached around cp/cpⴱ⬇1. With increasing frequency, the dependence of␥1on polymer concentration is qualitatively unchanged, but the value of␥1increases 关Fig.7共a兲兴. 10-1 100 101 σ1 [Pa] 10-1 100 101 102 σ2 [Pa] 2.0 1.5 1.0 0.5 0.0 cp/ cp* 10-1 100 101 102 σX [Pa] 100 101 102 3 4 5 6 7 8 100 2 100 101 102 3 4 5 6 7 8 100 2 100 101 102 3 4 5 6 7 8 100 2 (a) (b) (c)
FIG. 8. Yield stresses at共a兲 the first yield point,1;共b兲 the crossing point,X; and共c兲 the second yield point,
2, as a function of polymer concentration cp/cpⴱfor different frequencies= 1 rad/s 共쎲兲, 10 rad/s 共䊏兲, 40 rad/s共䉲兲, 69 rad/s 共䉱兲, and 100 rad/s 共⽧兲. The yield stresses have been extracted from the DSSs with 共a兲 the first yield point estimated as the onset of sublinear stress-strain relation,共b兲 the crossing point extracted from the strain dependence of G⬘and G⬙, and共c兲 the second yield point identified with the second inflection point of the stress-strain curve. Lines are fits to a power-law dependence=ˆ关共cp/cgelp 兲− 1兴 共Kobelev and Schweizer, 2005a兲. Insets: Log-log plots of the same data.
The values of the first yield strain ␥1 are comparable to the range of the attractive interparticle interactions;␥1⬇ⴱ共TableI兲. This suggests, by comparison to the behavior of attractive glasses关Pham et al.共2006,2008兲兴, that the first yield strain␥1reflects the breaking of interparticle bonds. This dissipates energy as indicated by the initial increase of G
⬙
and leads to breaking of the space-spanning gel network into disconnected clusters. An underlying mechanism of such a process could be the breakage of bonds between particles that connect two clusters. Confocal microscopy studies of similar samples 关Smith et al.共2007兲;Smith共2004兲兴 suggest that a significant number of such particles are present. In such a picture, after intercluster bonds are broken, the system would represent a “cluster fluid.” Cluster release would be responsible for the observed decrease of G⬘
and G⬙
共Fig.4兲 as well as the sublinear stress-strain relation 关Fig.6共a兲兴. Furthermore, the residual elasticity of the remaining clusters and the possibility of dynamically reforming bonds could cause a viscoelastic liquidlike response distinct from a simple Newtonian liquid. The presence of clusters or structural heterogeneities is consistent with scattering and microscopy results 关Laurati et al.共2009兲兴.Although the values of the first yield strain␥1and the interaction rangeⴱ are com-parable, their dependencies on polymer concentration display competing behavior; the interaction range decreases with increasing polymer concentration due to the transition from a dilute to a semidilute polymer solution. That the first yield strain␥1nevertheless increases with polymer concentration is attributed to the additional effect of the increas-ing interaction strength with increasincreas-ing polymer concentration, which tends to suppress bond breaking. Below, we combine both effects, the range and strength of the attractive interactions, in a simple model that considers the escape from the attractive potential well of the nearest neighbors and leads to semiquantitative agreement with our data共Sec. IV兲. The frequency dependence of the first yield strain ␥1 is attributed to the changing balance between shear-induced motion and Brownian motion. This is quantified by the Peclet number Pe, which describes the balance between Brownian motion, characterized in our case by the time required by the particle to diffuse over a distance corresponding to the interaction range of the potential 共B= 3.5⫻10−4 s兲, and shear-induced motion, represented by the period of the oscillatory shear共␥兲−1, with the strain amplitude␥. In our experiments, the Peclet number Pe=␥Bis always smaller than 1, and thus Brown-ian motion dominates. This suggests that a particle bond, when it is considerably stretched by shear, has a high probability to break because Brownian motion allows the particle to escape from the attractive potential well. Within one period, the residence time of a particle in the stretched bond condition is determined by the frequency , which implies that a higher frequency reduces the residence time and thus the probability of escape. Thus, a larger strain is required for higher frequencies. This qualitatively agrees with the observed increase in ␥1 with increasing frequency and is in semiquantitative agreement with the model proposed in Sec. IV.
In contrast to the first yield strain ␥1, the second yield strain ␥2 decreases with in-creasing polymer concentration 关Fig.7共c兲兴. At low polymer concentrations, ␥2 depends strongly on frequencywith values ranging from about 200% to 400%, while at higher polymer concentrations共cp/cⴱpⲏ1兲 a value of␥2⬇170% is attained, almost independent of frequency . These strains correspond to displacements of about two to four particle diameters, depending on polymer concentration cp/cpⴱ. This length scale and its
cp/cpⴱ-dependence are consistent with the typical cluster sizes observed by differential interference contrast 共DIC兲 and confocal microscopy 关Laurati et al. 共2009兲; Smith 共2004兲兴. In the same study, the shape of clusters composing the gel network was found to become more elongated with increasing polymer concentration. Similar trends for the dependence of the number of bonds and cluster shape on cp/cpⴱ were observed by
con-focal microscopy for PMMA-PS gels at a lower volume fraction, namely,= 0.2关Dibble
et al.共2006兲;Lee and Furst共2008兲兴. A response resembling that of a Newtonian liquid is thus only expected when not only the intercluster bonds are broken but also the clusters themselves are broken down or restructured. In order to quantitatively model this process, detailed structural information is required, which is not available at the moment. Quali-tatively, cluster rearrangements occur on much larger length scales共about 150%–400%兲 than typical cage distortions of about 10% and 50% as implicated in the yielding of repulsive and attractive glasses, respectively 关Pham et al.共2006,2008兲兴. This is consis-tent with the heterogeneous共cluster兲 structure of gels, whose length scale decreases with increasing volume fraction, i.e., by moving from gels to attractive glasses关Zaccone and Del Gado共2009兲兴. This connection is more precisely demonstrated in a recent investiga-tion of the yielding of similar depleinvestiga-tion gels along a line of constant attracinvestiga-tion and the range of volume fractions covering attractive glasses as well as low volume fraction gels 关Koumakis and Petekidis 共2011兲兴. As in our experiments, this probes the effect of a decrease of the characteristic size of heterogeneities, which occurs with increasing attrac-tion strength or polymer concentraattrac-tion 关Dibble et al. 共2006兲;Laurati et al. 共2009兲; Lee and Furst 共2008兲;Smith共2004兲兴.
With increasing frequency, the second yield strain␥2decreases, which indicates a decreasing size of the “free” clusters which are left after the first network disruption. This indicates an enhanced efficiency of shear in disintegrating the gel structure at larger frequencies or, in other words, a shear-rate dependence of such a cluster size.
The yield strain ␥X, which corresponds to the crossing point of G
⬘
and G⬙
, first increases and then decreases as a function of polymer concentration, with a maximum around cp/cpⴱ⬇1 关Fig.7共b兲兴. The yield strain␥Xis intermediate between the first,␥1, and second,␥2, yield strains, and thus the behavior is also expected to be intermediate, with the intercluster bonds broken and thus the gel network destroyed. However, not all intra-cluster bonds would be broken, and hence the intra-clusters would not yet be completely restructured and a viscoelastic liquid of clusters would be present. The observed trend indicates a gradual transition from bond breaking to cluster rearrangements共around the gelation boundary, cp/cpⴱ= 0.4兲 with the size of clusters decreasing upon increasing poly-mer concentration共within the gel region兲. The two processes take place at characteristic strains that become increasingly close with increasing cp/cpⴱ. Therefore, in close proxim-ity to the gelation boundary, ␥X follows the trend of the bond breaking process, while well inside the gel 共cp/cⴱp⬎1兲 the processes begin to superimpose and ␥X follows the smooth decrease of␥2, with␥1being saturated. Similar trends have been predicted for the yield strain by a model for the nonlinear viscoelasticity of depletion gels 关Kobelev and Schweizer共2005a兲兴. Note also that the trend of␥Xreflects the ratio between G⬘
and G⬙
in the linear regime due to the fact that the strain dependences of the two moduli are comparable at different polymer concentrations. This suggests that␥X follows the solid/ fluid balance as a function of polymer concentration in the linear regime. The frequency dependence of ␥X also shows intermediate behavior, with the initial increase 共 ⱕ40 rad/s兲 following the frequency dependence of␥1 and the subsequent decrease共 ⬎40 rad/s兲 following the frequency dependence of␥2.We now consider the yield stresses at the three yield points:1,2, and X 共Fig.8兲. All three yield stresses show similar behavior, increasing with polymer concentration and frequency. The dependences on polymer concentration, i.e., strength of attraction, follow a power-law increase共log-log plots in the insets兲. Such a form has already been found for colloidal gels formed due to van der Waals attractions 关Buscall et al. 共1987兲; Larson 共1999兲;Russel and Grant 共2000兲兴 and thermoreversible gels 关Gopalakrishnan and Zuko-ski 共2007兲兴 for both increasing polymer and colloid concentrations. Furthermore, for
depletion gels similar to ours, a power-law dependence has been predicted by an acti-vated barrier hopping approach关Kobelev and Schweizer共2005a兲兴. This model is based on a shear-induced deformation and reduction of the free-energy barrier, which favors the escape of bonded particles. Adapting this approach to our system with its intercluster and intracluster bonds, our particles can be considered as trapped in a free-energy landscape involving two barriers. At the first yield point, intercluster bonds are broken, thus dis-rupting the gel network structure. The first yield stress 1 is thus expected to increase with attraction strength controlled by the polymer concentration. A fit of a power law =ˆ关共cp/cp
gel兲− 1兴 关Kobelev and Schweizer共2005a兲兴 to the data results in good agree-ment关Fig.8共a兲兴. The obtained exponent= 1.6 is lower than experimentally determined values for similar systems 关Buscall et al. 共1987兲; Gopalakrishnan and Zukoski 共2007兲; Larson共1999兲; Russel and Grant共2003兲兴 and theoretical predictions 关Kobelev and Sch-weizer共2005a兲兴, which suggest values of 2.0ⱕⱕ2.5. Note that in our system experi-mental evidence exists关Dibble et al.共2006兲;Smith共2004兲兴 that the number of bonds per particle decreases as a function of cp/cpⴱ, in contrast to the assumption used in theoretical calculations关Kobelev and Schweizer共2005a兲兴. This can be attributed to the fact that the theory is based on homogeneous gels while our gels, possibly as a result of a nonequi-librium route to gelation, present a large degree of heterogeneity. Such an additional dependence could be a reason for the smaller exponents found in our case. The position of the gel boundary derived from the power-law fit, cgelp , is found to be in the range 0.25cⴱp⬍cp
gel⬍0.32c p
ⴱ. It has been fixed to c
p/cpⴱ= 0.25, which is consistent with our previous work 关Laurati et al. 共2009兲兴 and leads to good fits at all frequencies. Fitted values of the prefactor ˆ increase with increasing frequency over the range 2ⱕˆ
ⱕ5 Pa. The agreement between the experimental data and the power-law scaling pro-posed by Kobelev and Schweizer 共2005a兲supports the interpretation of the first yield stress 1 as the one necessary to overcome the free-energy barrier localizing each par-ticle, which is also consistent with the concept of bond breaking.
The incomplete yielding of samples at the first yield point suggests that particles are still trapped within clusters, which—however—can flow past each other, representing a cluster liquid. We propose that the second yield stress corresponds to a further disruption of these clusters under larger deformations. While the theory ofKobelev and Schweizer 共2005a,2005b兲 in its present form does not describe such a process, we speculate that in analogy to the first yield point, it might be described as a shear-induced reduction of entropic intracluster barriers where clusters are broken down to individual particles or smaller clusters allowing for a free flow. Again, this barrier increases with increasing polymer concentration, i.e., attraction strength. The second yield stress also follows a power-law dependence on cp/cpⴱwith a slightly smaller exponent 1.2ⱕⱕ1.4. However, the agreement is not as good as for the first yield point. It is interesting to note that in this case the prefactor ˆ ranges from 10 Pa at= 1 rad/s to 30 Pa at= 100 rad/s. These considerably different levels of stress characterizing the first and second yielding pro-cesses indicate a clear distinction between intracluster bond breaking and intercluster rearrangements. In addition, the average number of bonds in similar samples has been observed by confocal microscopy to decrease from 6 at cp/cⴱp= 0.4 to 4 at cp/cⴱp= 0.8 and further to 2 at cp/cpⴱ= 1.5关Smith共2004兲兴. The ratio 2/1 shows a comparable depen-dence, thus reflecting the fact that, while the increase of the strength of bonds with increasing cp/cpⴱ contributes to the increase of 2 共as for 1兲, less compact elongated clusters formed at larger polymer concentrations would need to break less bonds to rearrange, hence contributing to a relative decrease of2. Finally, the stress at the cross-ing point,X, also follows a power law with= 1.9 and 3ⱕˆⱕ6 Pa, as a result of being intermediate between the first and second yield points.
All three yield stresses increase in value with increasing frequency共Fig.8兲, which is reflected in the increasing ˆ , as mentioned above. The same frequency dependence is
observed in the linear regime关Laurati et al.共2009兲兴, which suggests that in the case of the first yielding point the elastic strength of the samples in the quiescent state governs the stress needed to break the structure and fluidize the sample. In terms of the free-energy picture described above, the increase as a function of frequency of the prefactor used to fit the power-law dependence of stress on polymer concentration indicates an increase of the barriers with increasing frequency.
2. Step rate tests
In step rate tests, a constant rate of deformation is applied, the shear rate␥˙ , and the
stressmeasured as a function of time t共Fig.9兲. With strain␥=␥˙ t, the stressinitially increases linearly and then overshoots before it seems to tend toward a constant steady-state stress共within the measurement time of 1000 s a well-developed steady-state could only be observed for the largest shear rates兲. Two distinct stress overshoots are observed for polymer concentrations cp/cⴱpⱕ1, whereas for larger polymer concentrations, cp/cpⴱ = 2.0关Fig.9共d兲兴, only the second stress overshoot is clearly visible, while the first stress overshoot is reduced to a shoulder. With increasing polymer concentration, the strain at the first overshoot or shoulder, ␥1step, is initially constant with␥1step⬇12% for cp/cpⴱ⬍1 and then increases to about 20% for cp/cpⴱ= 1.5关Fig.10共a兲兴. The corresponding stress at the first overshoot, 1step, increases with polymer concentration 关Fig. 10共a兲, inset兴. The second stress overshoot moves to smaller strains upon increasing the polymer concentra-tion;␥2stepdecays from about 300% to 150%关Fig.10共b兲兴. The corresponding stress at the second overshoot, 2step, increases with polymer concentration 关Fig. 10共b兲, inset兴. The shear rate␥˙ has only a weak effect on the overshoots; increasing␥˙ slightly shifts the first
overshoot to higher strains and the second overshoot to lower strains共arrows in Fig.9兲. Two stress overshoots are consistent with the presence of two yielding processes, as found in DSS 共Sec. III D 1兲. A stress overshoot in step rate tests has also been reported for other systems such as polymer melts 关Osaki et al.共2000兲兴, nanocomposite materials 关Letwimolnun et al.共2007兲兴, metallic glasses 关van Aken et al.共2000兲兴, and concentrated colloidal suspensions with both repulsive sphere and attractive interactions关Carrier and Petekidis共2009兲;Zausch et al.共2008兲兴. MCT relates the stress overshoot seen in suspen-sions of hard spheres to superdiffusive motion of the particles, which has been observed in the transient dynamics of this system 关Zausch et al. 共2008兲兴. Similarly, simulations relate the overshoot to a distortion of the structure, characterized by an anisotropic com-ponent of the pair correlation function 关Zausch and Horbach 共2009兲兴. While only one stress overshoot has been found in hard-sphere systems, two have been observed in attractive glasses关Pham et al. 共2008兲兴.
The strain at which the first stress overshoot appears,␥1step, is slightly larger than, but still comparable to, the first yield strain observed in DSS, ␥1, and furthermore shows a similar dependence on polymer concentration 关Fig. 7共a兲兴. A power-law fit to the stress 共1
step兲 关Kobelev and Schweizer 共2005a兲兴 yields good agreement with the data and an exponent= 1.7, consistent with the findings of the DSS experiments共Sec. III D 1兲 and other studies 关Buscall et al. 共1987兲; Gopalakrishnan and Zukoski 共2007兲; Kobelev and Schweizer共2005a兲;Larson共1999兲;Russel and Grant共2000兲兴. The prefactor of the power-law fit, 1ⱗˆⱗ2 Pa, is also in agreement with values obtained for 1 from DSS mea-surements. This suggests that the first stress overshoot is also related to the breaking of intercluster bonds. Shear induces a stretching of the bonds, which represents an elastic deformation with the buildup of stress until the bonds break at the first yield strain共␥1step兲.
The balance between the decreasing range and the increasing strength of the bonds is again consistent with the dependence of the yield strain on polymer concentration 共Sec. IV兲. The increase of␥1
step
with increasing shear rate is also supportive of the findings from the DSS experiments.
The second stress overshoot at larger strains,␥2step关Fig.10共b兲兴, supports the existence of a second yielding process which involves the breaking of intracluster bonds along with the deformation of clusters. Again, the value of␥2stepas well as its dependence on polymer
20 15 10 5 0
σ
[Pa
]
14 12 10 8 6 4 2 0 6 5 4 3 2 1 0 100 80 60 40 20 0 10-1 100 101 102 103 104 γ [%] (a) (b) (c) (d)FIG. 9. Step rate tests with stressas a function of strain␥=␥˙ t for samples with polymer concentrations
cp/cpⴱ= 0.5共a兲, 0.7 共b兲, 1.0 共c兲, and 2.0 共d兲 and shear rates 共increasing from bottom to top at high strains兲␥˙ = 0.01, 0.05, 0.10 s−1 关共a兲,共b兲,共d兲兴 and 0.005,0.03,0.10 s−1 共c兲. The duration of the experiments was t
concentration and shear rate resembles that of the second yield strain␥2observed in DSS experiments关Fig.7共c兲兴, where the decreases of␥2with polymer concentration and shear rate have been attributed to the decreasing cluster size共Sec. III D 1兲. The corresponding stress again shows a power-law dependence on cp/cⴱp with an exponent = 2.0. The values of the stresses at the second yield point are very similar to those of the first yield point with a similar prefactorˆ⬇2 Pa, which is an order of magnitude lower than the
corresponding values from DSS measurements.
To further investigate the relation between the overshoots in the step rate tests and the yield points in the DSS experiments, the stress-strain data from the step rate tests共Fig.9兲 are regrouped; the results for different polymer concentrations cp/cⴱpbut a fixed shear rate ␥˙ = 0.10 s−1are collected and shown in Fig.11. This representation reveals the similarity to the strain dependence of the elastic stress G
⬘
␥ as determined in a DSS experiment 关Fig.6共b兲兴. In both cases, the strain dependence shows two maxima or shoulders before a steady-state value is attained. With increasing polymer concentration, the first maxi-mum evolves into a shoulder and both maxima共or shoulders兲 move closer, i.e., the strains of the first maxima,␥1and␥1step, increase and the strains of the second maxima,␥2 and␥2 step
, decrease. The steady-state values of the stress⬁stepincrease with polymer concen-tration and rate␥˙ 关Fig.12共b兲, full symbols兴.
600 400 200 0 γ2 step [%] 2.0 1.5 1.0 0.5 0.0 cp/cp* 20 15 10 5 0 γ1 step [%]
(a)
(b)
10-1 100 101 102 σ1 step [Pa] 2.0 1.5 1.0 0.5 0.0 10-1 100 101 102 σ2 step [Pa] 2.0 1.5 1.0 0.5 0.0FIG. 10. Strains at共a兲 the first overshoot or shoulder,␥1step, and共b兲 the second overshoot,␥
2
step, as a function of
polymer concentration cp/cpⴱfor different shear rates ␥˙ = 0.01 s−1共䊏兲, 0.05 s−1共䉱兲, and 0.10 s−1共쎲兲. The insets show the corresponding stresses,1stepand2step, for the same shear rates with corresponding power-law fits. The strains and stresses have been extracted from step rate tests.
These steady-state values of the stress⬁stepare compared to flow curves, i.e., the stress as a function of shear rate␥˙ , as determined in a steady rate sweep experiment
per-formed upon decreasing rate关Fig.12共a兲兴; the data essentially agree 共the stress is, particu-larly at low rates, slightly underestimated in the standard steady rate sweeps, which we attribute to the long relaxation time needed to reach the steady-state兲. The flow curves do not exhibit a clear yield-stress plateau. This could be due to the fact that measurements were conducted upon decreasing shear rate starting from high rates where the network structure is broken. Thus, at low rates a long time would be needed for network forma-tion. Furthermore, the “plateau” has a finite slope in step rate experiments, especially for
cp/cpⴱ= 1.0, where a very small shear rate was measured. This is consistent with activated hopping processes leading to bond breaking, and thus a plateau is not expected to develop 关Kobelev and Schweizer共2005b兲兴. The experimental data in the lowest 2 decades exhibit a smooth power-law dependence on shear rate with an exponent of about 0.23, in agree-ment with such theoretical predictions. We can extract an indicative yield stress from the value at the lowest shear rate measured, as a function of polymer concentration. Again, the yield stress 0 follows a power-law dependence 关Kobelev and Schweizer 共2005a兲兴 with an exponent= 1.9, consistent with the findings of other tests共Sec. III D 1兲 as well as other studies关Buscall et al.共1987兲;Gopalakrishnan and Zukoski共2007兲;Kobelev and Schweizer 共2005a兲;Larson共1999兲;Russel and Grant共2000兲兴.
3. Step stress and recovery tests
In a step stress共creep兲 test, a step of stress is applied and the induced strain␥共t兲 is measured as a function of time t共Fig.13兲. After the initial oscillations due to coupling of the tool inertia with the viscoelastic properties of the sample, the strain ␥共t兲 increases sublinearly as a function of time provided the applied stressis well below the first yield stress 1 关1= 3 Pa for cp/cpⴱ= 0.5 and 0= 10 Pa for cp/cpⴱ= 1.0 at = 1 rad/s, from DSS experiments; Fig.8共a兲兴. The strain␥共t兲 does not attain a constant value as would be expected for an ideal elastic solid 关Figs.13共a兲 and 13共c兲兴, i.e., we observe a so-called
FIG. 11. Step rate tests with stressas a function of strain␥=␥˙ t at a shear rate␥˙ = 0.10 s−1for samples with
solid creep response. Correspondingly, the viscosity共t兲 increases with time 关Fig.13共b兲, = 1 Pa, and Fig.13共d兲,= 2 and 8 Pa兴. If the applied stressbecomes comparable or larger than the first yield stress1, a complex response is observed as a function of time, with regions of decreasing viscosity 共shear thinning,␥⬃tv, withv⬎1兲 alternating with regions of increasing viscosity共shear thickening or creeping, ␥⬃tv, withv⬍1兲, before the system reaches a constant viscosity共steady-state flow,␥⬃t兲. For values of the stress significantly larger than 1but still lower than 2, the viscosity shows a maximum as a result of the transition from shear thickening to shear thinning 关Figs.13共b兲 and13共d兲兴. Such a maximum, or equivalently the onset of shear thinning, shifts toward shorter times with increasing stress until, for ⬎2, shear thinning is observed immediately after application of stress.
The complex response upon applying a stresswith1⬍⬍2reflects the presence of two yielding processes and is reminiscent of the behavior of attractive glasses关Pham
et al. 共2008兲兴. As the strain increases with time, the samples pass through regimes with
intercluster and intracluster bond breaking as well as bond reformation.
To investigate the transition from creep to flow behavior upon increasing the applied
100 2 4 6 101 2 4 6 102 2 σ [P a ] 10-3 10-2 10-1 100 101 dγ/dt [s-1] 10-2 10-1 100 101 102 σO [Pa] 2.0 1.5 1.0 0.5 0.0 cp/ cp*
(a)
(b)
FIG. 12.共a兲 Stressas a function of shear rate␥˙共flow curve兲 as determined in a standard measurement 共open
symbols兲 and the steady-state value⬁stepfrom step rate tests共full symbols兲 for different polymer concentrations
cp/cpⴱ= 0.5, 0.7, 1.0, and 2.0共from bottom to top兲. 共b兲 Stress0, corresponding to the stress measured at the
lowest rate in steady rate sweeps, as a function of polymer concentration. The line is a fit to the power-law dependence=ˆ关共cp/cp
stress , a power-law dependence ␥共t兲⬃twas fitted to the data at long times, 150⬍t ⬍200 s, although this is only a limited region. The strain dependence of the creep exponent provides the transition from creep 共⬍1兲 to flow 共⬇1兲, which occurs at the yield stress共Fig.13, arrows in insets兲. The values of共Fig.14兲 and their dependence on polymer concentration are consistent with the values of the first yield stress obtained from DSS,1关Fig.8共a兲兴, and the first and second yield stresses obtained from step rate tests, 1step and2step共Fig. 10, insets兲. The dependence of on polymer concentration
10-1 100 101 102 103 104 η [Pa·s] 10-3 10-1 101 103 105 107 γ [%] 10-2 100 102 104 106 γ [%] 100 101 102 103 104 105 η [Pa·s] 10-2 10-1 100 101 102 t [s] 1.0 0.8 0.6 0.4 0.2 0.0 λ 25 20 15 10 5 0 σ [Pa] 1.0 0.8 0.6 0.4 0.2 0.0 λ 60 50 40 30 20 10 0 σ [Pa] (a) (b) (c) (d)
FIG. 13.共a兲,共c兲 Strain␥and共b兲,共d兲 viscosityas functions of time t as determined in step stress共creep兲 tests
for samples with polymer concentrations cp/cⴱp= 0.5关共a兲,共b兲兴 and 1.0 关共c兲,共d兲兴 for different applied stresses = 1.0, 2.0, 2.4, 3.0, 4.0, 5.0, 7.0, 10.0, and 15.0 Pa关共a兲 from bottom to top; 共b兲 from top to bottom兴 and 2.0, 8.0, 14.0, 18.0, 26.0, 35.0, 45.0, 55.0, and 62.0 Pa关共c兲 from bottom to top; 共d兲 from top to bottom兴. According to the DSS experiments, these stresses are below the first yield stress,⬍1共open symbols兲; between first and second
yield stresses,1⬍⬍2共full symbols兲; and beyond the second yield stress,⬎2共crosses and plus兲. Insets:
Creep exponent as a function of stressfor the same samples. Arrows indicate when the plateaux⬇1 are reached.
also agrees with a power law 关Kobelev and Schweizer 共2005a兲兴 with an exponent = 1.8共Fig.8兲, which is consistent with the results of our DSS and step rate experiments 共Secs. III D 1 and III D 2兲 and other studies 关Buscall et al.共1987兲;Gopalakrishnan and Zukoski 共2007兲; Kobelev and Schweizer 共2005a兲; Larson 共1999兲; Russel and Grant 共2000兲兴.
When the applied stress is removed after a step stress experiment, the system relaxes. This relaxation was followed and the recovered strain␥r共t兲 was measured over time t, where␥r共t兲=␥max−␥共t兲, with␥maxbeing the maximum strain reached at the end of the application of stress 关Figs. 15共a兲 and 15共b兲兴. A part of the strain recovers almost immediately within the fist second. This is followed by a slow relaxation to a steady-state, the maximum recovered strain ␥rmax, within 100–400 s, depending on the applied stress. The maximum recovered strain␥rmaxfirst increases with increasing applied stress until it reaches a maximum,␥1creepat1creep, and then decreases to a steady-state value ␥2
creep, which is attained at 2
creepand only clearly detected at lower c
p/cpⴱ, probably due to the lower absolute value of the yield stresses 关Fig. 15共c兲兴. With increasing polymer concentration, the maximum␥1creepfirst grows from cp/cpⴱ= 0.4 to 0.5, then decreases at
cp/cpⴱ= 0.7, and successively increases again at larger polymer concentrations 关Fig. 16共a兲兴. At the same time, the position of the maximum,1creep, increases with polymer concentration 关Fig. 16共a兲, inset兴. The plateau ␥2creep decreases with increasing polymer concentration, and its beginning shifts to larger stresses2creep关Fig.16共b兲兴.
In all cases, the maximum recovered strain ␥rmaxis much smaller than the maximum strain reached at the end of the application of stress, ␥max关Fig.15共c兲兴. At small applied stresses, the sample elastically relaxes with the recovered strain ␥r proportional to the applied stress . An almost elastic recovery is observed up to the maximum recovered strain ␥1creepat 1creep, which is in agreement with the transition from creep to flow ob-served upon application of stress,1creep⬇共Fig.14兲. At the same stress1creep⬇, i.e., at the end of creep and the beginning of flow, plastic deformation starts关Fig.15共c兲, inset兴. In the plastic regime, the sample recovers only a very small fraction of its deformation. This is consistent with increasing structural disruption of the particle network and the presence of independent clusters or, at higher stresses, individual particles. Interestingly, the values of␥r
max
at cp/cpⴱ= 0.5 and 1.0 coincide with the onset of structural and dynamic
10-1 100 101 102 σλ [Pa ] 2.0 1.5 1.0 0.5 0.0 cp/ cp* 10-1 100 101 102 3 4 5 6 7 8 9 100 2
FIG. 14. Stressat the transition from creep to flow as a function of polymer concentration cp/cpⴱas extracted
changes observed in confocal microscopy and scattering experiments of very similar samples 关Smith 共2004兲; Smith et al. 共2007兲兴. The stress 1creep and strain ␥1creep at the maximum are expected to be affected by both cluster and bond elasticity. At the end of stress application, the network structure seems to have been broken 共they creep兲, but nevertheless no simple flow is observed, indicating that some structure 共clusters兲 is present. This is similar to what was observed for the crossing point of G
⬘
and G⬙
. The values of ␥1creep and 1creep are indeed similar to those of ␥X and X observed in DSS experiments 关Figs. 7共b兲 and 8共b兲兴. An effect of the cluster size is suggested by the dependence on polymer concentration;␥1creepshows a minimum that coincides with the maximum size of structural heterogeneities 共or clusters兲 at this polymer concentration 关Laurati et al. 共2009兲兴. Furthermore, with increasing polymer concentration cp/cⴱp the8x103 6 4 2 0 γmax [%] 60 50 40 30 20 10 0 σ [Pa] 16 14 12 10 8 6 4 2 0 γr max [%] 100 80 60 40 20 0 σ [Pa] 14 12 10 8 6 4 2 0 γr (t) [%] 0.01 0.1 1 10 100 t [s] 12 10 8 6 4 2 0 γr (t) [%] (a) (b) (c)
FIG. 15.共a兲,共b兲 Recovered strain␥ras a function of time t and共c兲 maximum recovered strain␥rmaxas a function
of applied stress. Both are from the relaxation after step stress共creep兲 experiments for samples with polymer concentrations cp/cpⴱ= 0.5关共a兲; 쎲 in 共c兲兴 and 1.0 关共b兲; 䊏 in 共c兲兴 for different applied stresses= 1.0, 2.0, 2.4, 3.0, 4.0, 5.0, 7.0, 10.0, and 15.0 Pa共a兲 and 2.0, 8.0, 14.0, 18.0, 26.0, 35.0, 45.0, 55.0, and 62.0 Pa 共b兲, with symbols as in Fig.13. Arrows in共c兲 indicate the maxima of␥rmaxand the beginning of the plateaux. Inset: Maximum strain at the end of the step stress as a function of applied stress. Arrows indicate onset of yielding.
attraction strength increases, which is consistent with the observed increase of1creepwith polymer concentration关Fig.16共a兲, inset兴. The predicted power-law dependence 关Kobelev and Schweizer 共2005a兲兴 again shows good agreement and results in an exponent = 1.8, consistent with the results of our DSS, step rate, and step stress experiments共Secs. III D 1 and III D 2 and above兲 and other reports 关Buscall et al.共1987兲;Gopalakrishnan and Zukoski共2007兲; Kobelev and Schweizer共2005a兲;Larson共1999兲;Russel and Grant 共2000兲兴.
At very high stresses, beyond 2creep, the recovered strain saturates at ␥2creep. This residual elastic response suggests that some structures, small clusters, or heterogeneities remain or are reformed, depending on the strength and range of the attractions. The steady-state strain ␥2creep decreases with increasing polymer concentration 关Fig. 16共b兲兴, reflecting the decreasing effective range of the potentialⴱ共TableI兲. Its value in addition is similar to the effective range; ␥2creep⬇ⴱ. The value of 2creep is related to the force required to break down clusters or to overcome a second free-energy barrier localizing particles within clusters. This is consistent with the power-law increase of 2creep with polymer concentration cp/cpⴱ and thus attraction strength with an exponent= 2.0. Fur-thermore, the values of 2creep and 2 extracted from DSS experiments 关Fig. 8共c兲兴 are found to be in agreement. 14 12 10 8 6 4 2 0 γ 1 creep [%] 100 101 102 σ1 creep [Pa] 2.0 1.5 1.0 0.5 0.0 12 10 8 6 4 2 0 γ 2 creep [%] 1.0 0.8 0.6 0.4 0.2 0.0 cp/cp* 100 101 102 σ2 creep [Pa] 2.0 1.5 1.0 0.5 0.0 (a) (b)
FIG. 16. Strains and stresses共insets兲 at 共a兲 the maximum,␥1creepand1creep, and共b兲 the beginning of the plateau,
␥2creepand 2creep, of the maximum recovered strain,␥rmax共兲 关arrows in Fig.15共c兲兴 as functions of polymer