Stress-controlled shear cell A stress-controlled shear cell for small-angle light scattering and microscopy S. Aime,1, a) L. Ramos,1 J. M . Fromental,1 G. Pr´evot,1 R. Jelinek,1 and L. Cipelletti1, b) Laboratoire Charles Coulomb (L2C), UMR 5221 CNRS-Universit´ede Montpellier, Montpellier, F-France (Dated: 27 August 2018) We develop and test a stress-controlled, parallel plates shear cell that can be coupled to an optical microscope or a small angle light scattering setup, for simultaneous investigation of the rheological response and the microscopic structure of soft materials under an imposed shear stress. In order to minimize friction, the cell is based on an air bearing linear stage, the stress is applied through a contactless magnetic actuator, and the strain is measured through optical sensors. We discuss the contributions of inertia and of the small residual friction to the measured signal and demonstrate the performance of our device in both oscillating and step stress experiments on a variety of viscoelastic materials. Keywords: rheology; light scattering; microscopy; custom shear cell; stress-controlled

I. INTRODUCTION and creep and yielding in amorphous, dense emulsions and , and surfactant phases3,53–65. Complex fluids such as solutions, surfactant Previous works may be classified according to the phases, foams, emulsions and colloidal suspensions are probe method, the deformation geometry and the rhe- ubiquitous in everyday life and industrial applications. ological quantities that are measured. In a first group Their rheological properties are often of paramount im- of experiments, mostly performed in the last 30 years portance both during the production process and for the of the past century, small angle static light scattering final user1. They are also a topic of intense fundamen- (SALS) was used to monitor the change of the sam- tal research in fields as divers as the physics of poly- ple structure10–12,15–17,20–22,25. Small angle X-ray scat- mers2, the glass transition3,4, foams dynamics5, and bi- tering (SAXS13,14,24,27,29,31,33,39,42,56) and neutron scat- ological fluids and tissues3,6. Conventional rheology is tering (SANS,23,32,34,35,41,44,46,47,49 and Ref.66 and refer- widely used as a powerful characterization tool1, provid- ences therein) have also been used62,67, although they ing valuable information on the macroscopic mechanical require large scale facilities and are therefore less acces- properties of a system. As early as in the 1934 study of sible than microscopy and light scattering. Quite gen- shear-induced emulsification by G. I. Taylor7, however, it erally, these experiments were performed under strain- was recognized that coupling rheology to measurements imposed conditions, often using custom-designed shear of the sample structure and dynamics at the microscopic cells. Stress measurements (or stress-imposed tests) were scale tremendously increases our insight in the material available in just a few cases11,16,21, where a commercial behavior. Most experiments rely on optical and scatter- rheometer was coupled to a scattering apparatus. ing probes of the microstructure, although other meth- With the development of advanced microscopy ods have also been introduced, e.g. acoustic velocimetry8 methods, in particularly scanning con- and nuclear magnetic resonance9. Indeed, in the last 40 focal microscopy, many groups have devel- years a large number of apparatuses have been developed, oped apparatuses that couple rheology and mi- which use microscopy, static and dynamic light scatter- croscopy20,28,30,43,45,48,51,57,58,60,61,64,65,68–73. Appara- ing, neutron and X-ray scattering to investigate the mi- tuses based on a commercial rheometer usually give crostructure of driven samples. The wide spectrum of access to both the shear stress and the strain65,69. topics that have benefitted from such simultaneous mea- With confocal microscopy, both a plate-plate and a surements demonstrates the importance and success of cone and plate geometry are possible, since in confocal this approach: a non-exhaustive list includes the inves- microscopy the sample is illuminated and the image tigation of the orientation dynamics and deformation of is collected from the same side. This allows one to individual objects such as emulsion drops, , liq- avoid any complications in the optical layout due to arXiv:1603.06703v2 [cond-mat.soft] 13 Jul 2016 uid crystals and protein clusters7,10–14, the influence of the wedge-shaped sample volume of the cone and plate shear on demixing and critical phenomena15,16, shear- geometry. Note that for rotational motion the cone induced structure distortion and non-equilibrium phase and plate geometry is preferable to the plate-plate one transitions17–36, the dynamics of foams37,38 and that of when a uniform stress is required, e.g. for yield stress defects in colloidal crystals39–43, shear banding44–50 and fluids or in the non-linear regime. Custom shear cells non-affine deformation in polymer gels and glasses51,52, have also been used in conjunction to microscopy. As for scattering-based setups, custom cells are in general strain-controlled, with no measurement of the stress (see however Ref.72 for a notable exception). In spite a)Electronic mail: [email protected] of this limitation, custom cells may be an interesting b)Electronic mail: [email protected] option in terms of cost and because they allow valuable Stress-controlled shear cell 2 features to be implemented: a fine control and a great up to several tens of µm) and the characteristic structural flexibility on the choice of the surfaces in contact with length scales of most complex fluids, and the possibility the sample, the creation of a stagnation plane through to extend it to the X-ray domain (X-photon correlation counter-moving surfaces28,45,71, which greatly simplifies spectroscopy, XPCS83). the observation under a large applied shear, and the In this paper, we introduce a novel, custom-made implementation of non-conventional shear geometries, shear cell that can be coupled both to a microscope e.g. small channels for investigating confinement and to a static and dynamic small angle light scat- effects48,68,69 or large-amplitude shear with linearly tering apparatus, such as that described in Ref.82. translating parallel plates30. The shear cell is composed by two parallel plates, Real-space microscopy data are unsurpassed in that one of which can undergo translational motion. In they provide full knowledge of both the structure and the contrast to other custom shear cells previously re- dynamics of the sample at the particle level. However, ported in the literature, for which a shear deformation microscopy suffers from several limitations: it is quite is imposed and no stress measurement is avail- 7,10,12,17,22–28,34,37–44,46,47,49,51,53–55,57,58,60,61,64,70,71 sensitive to multiple scattering, making turbid samples able , difficult to study; it requires specifically designed, fluo- our cell is stress-controlled and the strain is accurately rescently labelled particles if confocal microscopy is to be measured. Unlike the stress-controlled custom cell 72 used; only quite small sample volumes can be imaged; a of Ref. that is optimized for confocal microscopy high resolution, a large field of view and a fast acquisi- in an inverted microscope, our cell can be used in tion rate are mutually exclusive, so that a compromise both an inverted or upright conventional or confocal has to be found between these conflicting requirements. microscope, as well as in light scattering experiments in Scattering methods, while not accessing the single parti- the transmission or backscattering geometry; further- cle level, do not suffer from these limitations. Further- more, it can be placed either in a vertical or horizontal more, advanced scattering techniques such as the Time position. This versatile design opens the way to the Resolved Correlation74 (TRC) and the Photon Correla- full characterization of the rheological behavior of a tion Imaging75,76 (PCI) methods can fully capture tem- sample, simultaneously to its structural and dynamical porally and spatially varying dynamics, yielding instan- evolution at the microscopic level. In particular, its taneous coarse-grained maps of the dynamical activity. compact design allows it to be coupled to state-of-the-art Thus, scattering techniques are a valuable alternative to static and dynamic light scattering setups, a significant real-space methods not only for measuring the sample improvement over commercially available light scattering structure, as in the early works mentioned above, but adds-on for rheometers, which are limited to static light also to probe its dynamics. scattering. The rest of the paper is organized as follows: in Sec. II Dynamic light scattering in the highly multiple scat- we briefly present the shear cell, before discussing in tering limit (diffusing wave spectroscopy, DWS77) has Sec. III its main features independently of the inves- been used since the Nineties of the past century to mea- tigated sample (stress calibration, strain measurement, sure the microscopic dynamics associated with the affine plate parallelism, effects of inertia and residual friction). deformation in the shear flow of a simple fluid78 and Section IV demonstrates the cell performances through a the impact of shear on foam dynamics37. In a subse- series of tests on model systems, representative of a New- quent series of works, the so-called ‘echo-DWS’ method tonian fluid, a perfectly elastic solid, and a viscoelastic has been introduced: here, DWS is used to measure the Maxwell fluid. In Sec. V, we illustrate how the cell can irreversible rearrangements occurring in amorphous vis- be coupled to a dynamic light scattering apparatus, by coelastic solids such as emulsions53, foams38, or colloidal measuring the microscopic dynamics of a Brownian sus- gels and glasses54,55,59 subject to an oscillating shear de- pension at rest and under shear. Finally, an overview of formation. More recently, space-resolved DWS has been the characteristics of the shear cell, with an emphasis on applied to the investigation of the microscopic rearrange- its strengths and limits concludes the paper, in Sec. VI. ments in polymeric solids under compression or elonga- tion52,79. Dynamic light scattering in the single scat- tering regime80 (DLS), by contrast, has been much less used as a probe of the microscopic dynamics of a driven II. SETUP system43,81, perhaps because single scattering conditions require greater care than multiple scattering ones in de- A scheme of the setup is shown in fig. 1. The sample signing a scattering apparatus, due to the sensitivity of is confined between two crossed microscope slides, whose DLS to the scattering angle at which light is collected and surfaces are separated by a gap e, typically of the order its vulnerability to stray light scattered by any imperfec- of a few hundreds of µm. The gap can be adjusted us- tions in the optics82. In spite of its greater complexity, ing spacers and a set of precision screws to ensure the DLS has several appealing features, such as its ability parallelism of the plates, as described in more detail in to probe the dynamics on a large range of length scales Sec. III C. One of the two slides is fixed to an optical ta- (by varying the scattering angle), the excellent overlap ble, while the second one is mounted on a mobile frame between the probed length scales (from a few tens of nm that slides on a linear air bearing. The air bearing avoids Stress-controlled shear cell 3 any mechanical contact between the sliding frame and allelism will be given in Sec. III C. The sliding part of the the optical table, therefore minimizing friction. In or- setup has a relatively small mass, M ' 335 g, including der to avoid spurious contributions due to gravity, the the mass of the sliding bar itself, 288.0 g. This allows horizontal alignment of the air bearing rail can be ad- inertial effects to be kept small, as shown in Sec. IV. justed to within a few 10−5 rad using a finely threaded differential screw. To impose the desired shear stress, a controlled force is applied to the mobile frame using III. SETUP CHARACTERIZATION a magnetic actuator (see fig. 1), powered by a custom- designed, computer-controlled current generator. A con- In this Section, we describe the characterization of tactless optical sensor measures the displacement of the the empty shear cell, discussing in particular the force- moving frame and thus the sample strain. The imple- current calibration of the stress actuator, the optical mentation of the strain sensor is discussed in Sec. III B. measurement of the strain, the control and measurement of the plate parallelism, and the effects of inertia and residual friction. I B J L A C A. Calibration of the applied force

F The shear stress imposed to the sample is controlled G D E M using a magnetic actuator (Linear Voice Coil DC Motor LVCM-013-032-02, by Moticont), consisting of a perma- nent magnet and a coil winded on an empty cylinder. N The magnet is fixed to the sliding part of the appara- H I tus; it moves with no contact within the empty cylinder, J which is mounted on the fixed part of the apparatus. The L device can be easily controlled via a PC, which allows one to synchronously apply a given stress and acquire strain M and optical data. Due to the small coil resistance (5.9 E Ω) and inductance (1 mH), the magnetic actuator re- D F sponse is very fast and its characteristic time (less than K 1 ms) will be neglected in the following. The force ex- N erted by the voice coil is proportional to the current fed to the device. We use a custom-designed current gener- FIG. 1. Scheme of the experimental setup. The moving parts ator, which outputs a controlled current I that can be are shown in blue. A) Sample; B) fixed glass plate; C) mobile adjusted over a very large range by selecting the appro- glass plate; D) air bearing - fixed body; E) air bearing - sliding priate full scale value, between 10−5 A and 1 A, in steps of rail; F) air supply; G) gap setting spacer; H) plate alignment one decade. The current generator noise is smaller than screws; I) coil (fixed); J) magnet (mobile); K) cable to PC; 10−4 of the full scale. The current generator is in turn L) frosted glass; M) tilt adjustment; N) mechanical stops. controlled through a voltage input: by feeding different voltage waveforms to the current generator, it is there- The linear air bearing (model RAB1, from Nelson Air fore possible to impose a stress with an arbitrary time Corp.) is guided by air films of thickness ∼ 10 µm, which dependence. In our implementation, the voltage signal is confine its motion in the longitudinal x direction. The generated by a D/A card (USB-6002 by National Instru- viscous drag of the air films can typically be neglected, ments) controlled by a PC. We calibrate the voice coil whereas a small residual friction, most likely due to dust by measuring the force, F , resulting from the imposed particles, is seen when a force smaller than 1 mN is ap- current. The force is measured using a balance, with a plied, as it will be shown in Sec. III D. The maximum precision of 1 mg. The F vs. I curve is shown in fig. 2. linear travel of the system, set by the length of the mov- F is remarkably proportional to I over 4.5 decades, with ing plate, is 75 mm. However, we typically restrain the a proportionality constant k = (0.916 ± 0.003) NA−1. travel distance to 5-10 mm, over which the imposed force An important issue in designing the shear cell appara- is virtually independent of position (see Sec. III A). This tus is the requirement that the applied stress be constant typically corresponds to a strain γ ≤ 10. The nominal regardless of the resulting strain. To check this point, straightness of the sliding bar (provided by the manufac- we have measured several calibration curves similar to turer) is better than 1 µm over the whole travel length, that shown in fig. 2, each time slightly changing d, the whereas the nominal thickness of the plates is constant relative axial position between the magnet and the coil, to within 10 µm throughout their entire surface. There- with d = 0 the position of the magnet when it is fully fore, the gap e can be considered to be uniform over the inserted in the coil. The d dependence of the propor- typical working distance. More details on the plates par- tionality constant k is shown in the inset of fig. 2. In Stress-controlled shear cell 4 the range 7 mm < d < 15 mm, we find that k changes required, we use a custom optical setup, inspired by76 by less than 0.5%. In our experiments we typically work and schematized in fig. 3a. A frosted glass is fixed to in this range, thus ensuring that the applied force is es- the mobile frame of the shear cell. A laser beam illu- sentially independent of sample deformation. In order to minates the frosted glass, which is imaged on a CMOS convert the applied force to a stress value, we measure camera (BU406M by Toshiba Teli Corp). The camera the surface S of the sample by taking a picture of the has a 2048 × 2048 pixel2 detector, with a pixel size of shear cell after loading it. The surface is calculated from 5.5 µm; its maximum acquisition rate at full frame is 90 the image using standard image processing tools. Typ- Hz. The image is formed by a plano-convex lens with ical values of S and its uncertainty are 250 mm2 and a a focal length f = 19 mm, placed in order to achieve a few mm2, respectively. When taking into account both high magnification, m = 20.5, such that the pixel size the uncertainty of the F − I calibration (0.3%) and that corresponds to 268 nm on the frosted glass. Due to the on S, the stress σ = F/S is known to within about 1%. of the laser light, the image consists of a highly contrasted (see fig. 3a), which translates 1 0 0 0 as the mobile part of the cell, and hence the frosted glass, is displaced. Using image cross-correlation techniques84 the speckle drift and thus the sample shear can be mea- 1 0 0 sured. In practice, a speckle image acquired at time t is spatially cross-correlated with an image taken at a later 0 d [ m m ] time t . The position of the peak of cross-correlation as ] 1 0 a function of the spatial shift yields the desired displace-

N 0 5 1 0 1 5 0 ment between times t and t . An example is shown in m [

fig. 3b, which shows a cut of the cross-correlation func- 1 0 . 9 0 F tion along the x direction. The line is a Gaussian fit to ] A

the peak: when taking into account the magnification, 0 . 1 / 0 . 8 5 N the fit yields a speckle size σsp = 16.5 µm in the plane of [

k the sensor. The speckle size is controlled by the diameter 0 . 8 0 D of the lens and the system magnification; it has been 0 . 0 1 0 . 0 1 0 . 1 1 1 0 1 0 0 1 0 0 0 optimized in order to correspond to a few pixels, which minimizes the noise on the cross-correlation85. I [ m A ] To estimate the typical noise on the measurement of the displacement, we take a series of full-frame images FIG. 2. Calibration of the magnetic actuator: force (mea- of the speckle pattern over a period of 60 s, while keep- sured by a precision balance) as a function of the applied ing the frosted glass immobile. The displacement with current. The symbols are the experimental data and the line is the best fit of a straight line through the origin, yielding respect to the first image (t = 0), as measured from a proportionality constant k = (0.916 ± 0.003) NA−1. The the position of the cross-correlation peak, is shown in upper limit of the imposed current is set by the coil specifica- fig. 3c. The displacement fluctuates around a zero av- tions, whereas the smallest current chosen here is limited by erage value; the standard deviation of the signal over the balance precision. Inset: variation of k with the relative the full acquisition time, σn, may be taken as an esti- axial position between the magnet and the coil, d. The main mate of the typical noise on the measured position. We plot has been measured for d = 7 mm. find σn = 23 nm, more than 40 times smaller than the nominal precision of the commercial laser sensor. For a typical gap e = 200 µm it is therefore possible to reliably measure strains as small as 10−4. The main limitations B. Strain measurement of this technique are its time resolution and the largest displacement speed that can be directly measured. The In order to measure the displacement of the mobile former is limited by the camera acquisition rate and the frame and hence the sample strain, we use two differ- image processing time: in practice, the position can be ent optical methods, based on a commercial sensor and sampled at a maximum rate of about 50 Hz. The latter on a scattering technique, respectively. The former is is limited by the acquisition rate and size of the field of a commercial laser position sensor (model IL-S025, by view, lv ≈ 0.5 mm: if the frosted glass translates by more Keyence) that can acquire and send to a PC up to 1000 than lv between two consecutive images, the speckle pat- points per second with a nominal precision of 1 µm. tern is completely changed and no cross-correlation peak This sensor is very convenient to follow large, fast de- is observed. In practice, displacement speeds as high as formations. However, its precision sets a lower bound 1mm s−1 can be measured, corresponding to strain rates on the error on the strain measurements of at least 0.5% up toγ ˙ = 2 s−1 orγ ˙ = 5 s−1 for gaps of 500 µm and for a typical gap e = 200 µm; moreover, we find that 200 µm, respectively. Since the commercial sensor and the sensor output tends to artifactually drift over time. our custom device have complementary strengths and When more precise, more stable strain measurements are limitations, we typically use both of them simultaneously. Stress-controlled shear cell 5

A final point concerns the conversion of the displacement screws. In addition to the finely spaced fringes, due to the to strain units, for which the gap thickness is required. tilt, some larger-scale, irregular fringes are also observed, We determine it after adjusting the plates parallelism (see due to slight deviations of the plates from perfectly flat Sec. III C below), by placing the empty shear cell under surfaces. When optimizing the alignment (fig. 4a), only an optical microscope and by measuring the vertical dis- the irregular, large-scale fringes are visible. To gauge the placement required to focus the inner surface of the fixed precision with which the plates can be aligned, we impose and moving plates, respectively. a series of increasingly large tilt angles αimposed, acting on the two differential screws. αimposed is estimated from IMAGED a) the length of the plate, the nominal thread of the screws, SPECKLE FROSTED and the imposed screw rotation. For each α , we PATTERN DETECTOR LENS GLASS imposed measure the tilt angle αmeasured = Nλ/2L, where N is LASER the number of fine fringes observed over a distance L 퐷 BEAM and λ = 633 nm is the laser wavelength. Figure 4c shows the measured tilt angle as a function of the imposed one: 푧′ 푧 the data are very well fitted by a straight line with unit 100 slope and a small offset, σα ∼ 0.3 mrad, most likely due 1.0 Data Gauss fit b) c) to the difficulty of finding the optimum alignment since 50 the plate surfaces are not perfectly flat. We conclude that the plates can be tuned to be parallel to within a

0

0.5 tilt angle of about 0.3 mrad, a value comparable to or x [nm] -50 even better than for commercial rheometers86. Once the 0.0 plates have been aligned, they can be removed and placed -100 -6 -4 -2 0 2 4 6 Correlation value 0 20 40 60 again in the setup (e.g. for cleaning them) with no need x [m] t [s] to realign the setup. In this case, we checked that the parallelism is preserved to within 1 mrad. FIG. 3. (a) Optical scheme of the speckle imaging system for measuring the strain: D = 12.7 mm, z = 20 mm and z0 = 410 0 . 0 0 3 mm. (b) Cut along the x direction of the crosscorrelation a ) c ) between two speckle images taken at time t and t0. The peak position, ∆x = 1.34 µm, corresponds to the translation of the mobile part of the cell between t and t0. The line is a Gaussian fit to the peak, yielding a speckle size (in the sensor 0 . 0 0 2 plane) σ = 16.5 µm. (c) Time dependence of the measured ]

sp d a

displacement while the cell is at rest. The standard deviation r [

σn = 23 nm of the measured position yields an estimate of d b ) e r

the measurement error. u s a

e 0 . 0 0 1 m ؁

C. Adjusting the parallelism between the moving and fixed plates 0 . 0 0 0 0 . 0 0 0 0 . 0 0 1 0 . 0 0 2 ؁ The parallelism between the two plates is tuned using two differential screws (model DAS110, by ThorLabs), i m p o s e d [ r a d ] which control the position of the upper side of the fixed plate (see the pictures in fig. 1). The quality of the align- FIG. 4. (a) and (b): images of the interference fringes ob- ment is checked by visualizing the interference fringes served when the plates are parallel or tilted by an angle −4 formed by an auxiliary laser beam reflected by the two αimposed = 8 × 10 rad, respectively. The size of the field of inner surfaces of the plates. A change in the local separa- view is 15 × 22 mm2. (c) Measured tilt angle as a function of tion between the two plates modifies the relative phase of the manually imposed tilt angle. The symbols are the data the two reflected beams, thus changing the fringe pattern. and the line is a linear fit with slope 1. When the plates are parallel, no fringes should be visi- ble. The fringes are imaged on a CMOS camera (DMM 22BUC03-ML, by The Imaging Source, GmbH) with a 744 × 480 pixel2 detector, the pixel size being 6 µm, cor- D. Effects of inertia and friction responding to 31 µm in the sample plane (magnification factor m = 0.19). Figure 4b shows the fringe pattern ob- Inertia and friction effects may lead to spurious results served when a tilt angle of 8 × 10−4 rad between the two if they are neglected in the modelling of the setup re- plates is imposed on purpose, using the two differential sponse. Both effects are best characterized by measuring Stress-controlled shear cell 6 the cell motion under an applied force in the absence cell, varying the driving frequency from 0.75 rad s−1 to −1 of any sample. Under these conditions, the equation of 62.8 rad s . For each ω, we adjust Fω so as to keep motion for the moving part of the cell reads: the amplitude of the oscillations roughly constant and measure both the modulus and the phase ofx ˜ω from the (ext) Mx¨(t) = F (t) + Ffr [x(t), x˙(t)] , (1) time dependence of the displacement in the stationary regime. Figure 5c shows that the magnitude of the oscil- where M is the mass of the moving part, x its position lations has the expected value up to 4 rad s−1, beyond (with x(0) = 0 the position before applying any force), which it drops, due to the frequency dependence of the (ext) F the (time-dependent) force applied by the elec- response of the current generator and magnetic actuator. tromagnetic actuator, and Ffr a friction term that may The phase ofx ˜ω is close to zero, as expected, except for depend on both the position (in the case of solid friction) the lowest frequencies, for which the applied force is too and the velocity (for viscous-like friction). To character- small for friction effects to be neglected. ize the friction term, we measure x(t) after imposing a (ext) step force, F = F0Θ(t), with Θ(t) the Heaviside func- 1 0 0 0 tion. Figure 5a shows that x(t) follows remarkably well a a ) 1 0 b ) quadratic law, suggesting that the friction contribution, ] if any, is independent of both x andx ˙. We perform the 1 0 0 m 5 m [ same experiment for a variety of applied forces and ex- x - 1 [ tract, for each F0, the acceleration a from a quadratic 0 ؗ [ r a d s

fit to x(t). The results are shown in fig. 5b: while at ] 1 0

2 1 1 0 1 0 0 0 . 5 1 . 0 1 . 5 s 0 . 4 large F0 one finds a ∝ F0, as expected if friction is neg- / t [ s ] c ) ligible, for F < 1 mN the data clearly depart from a 0 . 3 0 m 1 1 | ؗ

proportionality relationship, suggesting that friction re- m

[ F 0 . 2

/ ؐ duces the effective force acting on the moving part of the 2 a / ؗ

؄ 1 . 0 cell. The line in fig. 5b is a fit to the data of the affine 0 . 1 M law a = (F0 − Ffr)/M, with M fixed to 335 g , yielding ؗ 0 . 0 x

Ffr = 0.12 mN. The fit captures very well the behavior | 0 . 1 - 0 . 1 of a over 4 decades, showing that a static friction term 0 . 0 1 accounts satisfactorily for the experimental data. Slight 0 . 1 1 1 0 1 0 0 deviations are observed for large forces, presumably due to the large uncertainty in the measurement of the dis- F 0 [ m N ] placement, a consequence of the very fast motion in this regime. Some small deviations are also observed at low FIG. 5. (a) time-dependent translation x(t) following the ap- forces, possibly because the friction term may slightly de- plication of a step force F0 = 0.9 mN to the empty cell. The viate from the simple position- and velocity-independent line is a quadratic fit. (b) Acceleration of the empty cell expression that we have assumed. Since the sample sur- vs. the amplitude of the step force applied. Symbols are face is typically A ∼ 2 cm2, the friction term sets a lower the data point and the line is the best fit of an affine law, F0−Ffr bound of the order of 1 Pa on the stress that may be a = M , with M = 335 g the mass of the sliding frame, applied in our apparatus. yielding a residual friction force Ffr = 0.12 mN as the only To probe the frequency dependence of the setup re- fitting parameter. (c) Normalized amplitude (black, left axis) sponse, we apply an oscillatory force to the empty cell, and phase (red, right axis) of the oscillating part of the re-   sponse to a cosinusoidal external force. The symbols are the ˜ iωt F (t) = Re Fωe , where the tilded variables are com- data points and the lines are the theoretical behavior. plex quantities. Assuming x(0) =x ˙(0) = 0 and neglect- ing friction, the expected response is

iωt  x˜(t) =x ˜ω e − 1 − iωt , (2) IV. TESTS WITH MODEL SAMPLES

˜ 2 withx ˜ω = −Fω/(Mω ). Note that, because the physical Having characterized the response of the empty cell, we equation of motion is given by the real part of Eq. 2, in- test its performances on standard rheological samples in ˜ 2 ertial effects yield either a nonzero offset Fω/(Mω ) (for oscillatory and step stress experiments. We use a purely a cosine-like applied force), or a nonzero drift velocity viscous fluid, a purely elastic solid, and a viscoelastic ˜ iFω/(Mω) (for a sinusoidal force), or both of them if the material well described by a Maxwell model. For each real and imaginary parts of F˜ω are nonzero. Thus, in os- sample, we briefly discuss the expected behavior in the cillatory shear experiments inertia effects add a constant presence of inertia and compare it to the outcome of ex- or linearly growing strain to the sample, whose contribu- periments, in order to assess the sensitivity and limits tion may or may not be negligible depending on the sam- of our setup. For the sake of simplicity, we will not in- ple rheological properties and the excitation frequency. troduce the friction term in the equations, but its effects To test Eq. 2, we apply a cosinusoidal force to the empty will be highlighted in discussing the experiments. Stress-controlled shear cell 7

When a sample is loaded in the shear cell, Eq. 1 has The frequency-dependent complex modulus extracted ∗ to be modified in order to take into account the force from the oscillating part of γ(t), G =σ ˜ω/γ˜ω, has mag- exerted by the sample on the cell, F (s)(x, x˙): nitude and phase given by

(ext) (s) |G∗| = ηωp1 + ω2τ 2 Mx¨(t) = F (t) + Ffr [x(t), x˙(t)] + F [x(t), x˙(t)] . v (9) arg G∗ = δ = arctan 1 . (3) ωτv In principle, F (s)(x, x˙) can be divided into an elastic (x- dependent) and a viscous (x ˙-dependent) parts. To recast At short times t  τv, viscous dissipation is always iωt  Eq. 3 in a form suitable to describe the sample rheological negligible and γ(t) ≈ γ˜ω e − 1 − iωt , as in Eq. 2, properties, we divide both sides by the sample surface A which was derived for an empty cell. At larger times,  iωt  and express position and velocity as strain γ = x/e and Eq. 7 simplifies to γ(t) ≈ iσ˜ω/(ηω) 1 − e / (1 + iωt) : strain rateγ ˙ , respectively: in the high frequency regime ωτv  1 inertial effects are still relevant, but in the opposite regime ωτv  1 the Iγ¨(t) + σ [γ(t), γ˙ (t)] = σ(ext)(t) (4) usual behavior for a viscous fluid is recovered: the cell oscillates around a negligible equilibrium positionγ ˜ωωτv, (ext) Here, I = eM/A is the inertia term, σ the applied with an amplitudeγ ˜ω ≈ −iσ˜ω/(ηω), and the strain lags stress and σ = −F (s)/A the sample stress, which has the stress by an angle δ = π/2. opposite sign with respect to F (s) since we adopt the usual notation where σ is the stress exerted by the setup on the sample. This is the general equation that has to 2. Experimental tests be solved given a constitutive equation for σ [γ(t), γ˙ (t)], an experimental protocol (i.e. the temporal profile of

σ(ext)), and the initial conditions. In the following, we 2 . 0 1 0 0 will systematically assume γ(0) =γ ˙ (0) = 0, and will 1 . 5 (ext) only consider either creep tests, σ (t) = σ0Θ(t), or (ext) iωt 1 . 0 oscillatory stresses, σ (t) =σ ˜ωe . ؇ 0 . 5

1 0 0 . 0 A. Newtonian fluid ] 0 . 0 0 . 5 1 . 0 1 . 5 2 . 0 1 -

s [

t [ s ]

1. Theoretical background.

. 8 ؇ 1 For a Newtonian fluid of viscosity η, σ = ηγ˙ . Equa- tion 4 therefore reads:

Iγ¨ + ηγ˙ = σ(ext) , (5) 0 . 1 0 . 1 1 1 0 1 0 0 which, for a step stress test with stress amplitude σ0 , ؓ has as solution 0 [ P a ] h  − t i γ(t) =γ ˙ ∞ t + τv e τv − 1 , (6) FIG. 6. Main plot: steady shear rateγ ˙ ∞ as a function of the applied stress in a step stress test, for a silicon oil with where τv = I/η is a characteristic time arising from the nominal viscosity η = 1.02 Pa s. Empty and filled symbols interplay between inertial and viscous effects andγ ˙ ∞ = refer to two independent measurements. The line is a fit of 2 σ0/η. For t  τv, inertia dominates and γ ≈ σ0t /(2I). γ˙ ∞ = (σ − σfr)/η where the offset σfr ∼ 0.25 Pa is due to In the opposite limit t  τv, the usual simple viscous friction. Inset: time evolution of the strain following a step flow γ =γ ˙ ∞t is recovered. Typical values of the crossover stress of amplitude σ0 = 1.58 Pa. The symbols are the data −1 time may be estimated from τ [s] ≈ (2η[Pa s])−1, where and the line is a fit of Eq. 6, with τv = 0.7 s andγ ˙ ∞ = 1.18 s , v −1 we have assumed M = 0.3 kg, e = 300 µm and A = in good agreement with (σ0 − σfr)/η = 1.27 s as obtained 2 cm2. Thus, the inertial regime lasts less than 1 s if the from the nominal viscosity. viscosity exceeds 0.5 Pa s−1. For an oscillatory stress, the solution to Eq. 5 is Measurements for step-stresses have been performed with 3 different silicon oils, with viscosity 1.02, 11.98 and h iωt  − t i 91.68 Pa s respectively (nominal values at 25◦C). As an γ(t) =γ ˜ω e − 1 + iωτv e τv − 1 , (7) example, we show in fig. 6 results for the less viscous with oil, for which the applied stress is varied between 0.3 and 100 Pa. The inset shows the time evolution of the σ˜ω i + ωτv strain following the application of a step stress with σ0 = γ˜ω = − 2 2 . (8) ηω 1 + ω τv 4.5 Pa. At large times, a pure viscous flow is found, Stress-controlled shear cell 8

γ(t) =γ ˙ ∞t. Deviations at short times from this linear plitude (20% γ) –i.e. for the less viscous samples and behavior are due to inertia. The data are in excellent the lowest frequencies–, deviations are observed due to agreement with the theoretical expression, Eq. 6 (line), the residual friction discussed above. Some deviations where τv = eM/(ηA) andγ ˙ ∞ are fitting parameters. We are also observed at the highest ω, due to the response plot in the main figure the steady-state shear rateγ ˙ ∞ vs of the current generator and voice coil, as mentioned in σ0, obtained by fitting γ(t) for all applied stresses. At Sec. III D. For the more viscous sample (η = 100 Pa s), large stresses, data are in excellent agreement withγ ˙ = no inertial regime is observed: |G∗| scales as ω, as ex- σ0/η using the nominal value of the viscosity, whereas pected for a Newtonian fluid. This is consistent with the friction causes deviations from this linear dependence for fact that for this sample inertial effects should become −1 the smallest stresses. Note that friction becomes relevant significant only for ω & 1/τv = 100 rad s , which is for σ0 . 1 Pa, in agreement with the lower bound on the beyond the range of probed frequencies. For the sample stress estimated in Sec. III D. Data over the whole range with intermediate viscosity (η = 10 Pa s), a crossover be- of applied stresses are very well accounted for by Eq. 6 tween |G∗| ∼ ω (viscous regime) and |G∗| ∼ ω2 (inertia- −1 (line in fig. 6), with τv as the only fitting parameter and dominated regime) is observed at ω ≈ 10 − 15 rad s , −1 using the nominal viscosity. The fit yields τv = 0.7 s, in consistent with 1/τv = 10 − 20 s (depending on e). good agrement with the expected value eM/(ηA) = 0.8 s. Finally, for the less viscous sample we estimate 1/τv = 2 s−1: for this sample, all data for which friction is neg- ligible lay in the high frequency regime ωτv > 1 where -M o d e l D a t a ؈ [ P a . s ] inertia dominates, leading to the observed |G∗| ∼ ω2 scal 3 1 0 1 ing. The inset of fig. 7 shows the argument of the complex

1 0 modulus as a function of the normalized frequency ωτ . v 1 0 The data nicely exhibit the transition between the low 2 1 0 1 0 0 and high frequency regimes predicted by Eq. 9, charac-

] terized by δ = π/2 (viscous regime) and δ = 0 (inertial ؔ a ؗ P v regime), respectively. The discrepancy seen for the fluid [

1 - 4 - 3 - 2 - 1 0

| 1 0 1 0 1 0 1 0 1 0 1 0 with η = 1 Pa s is again due to residual friction. * G

| 0 . 5 0

B. Purely elastic solid ؐ 1 0 / ؄ 0 . 0 1. Theoretical background 1 0 - 1 0 . 1 1 1 0 The constitutive law for a purely elastic sample is σ = ؗ Gγ, with G the elastic shear modulus. Equation 4 then [ r a d / s ] becomes

FIG. 7. Main graph: magnitude of the complex modulus of Iγ¨ + Gγ = σ(ext) , (10) Newtonian viscous fluids measured by oscillatory rheology, as a function of angular frequency. Solid (open) symbols are ex- whose solution involves the characteristic frequency Ω = periments performed with a gap e = 300 µm (e = 600 µm). pG/I, related to an inertial time scale. Inertia is ex- The data are labelled by the nominal viscosity. The dashed pected to dominate on time scales smaller than Ω−1, lines are the theoretical predictions, Eq. 9, using the nominal whose typical value for our setup, using M = 0.3 kg, viscosity and the inertia term issued from measurements on e = 300 µm, and A = 2 cm2, is given by Ω−1 [s] ≈ the empty cell. Inset: phase δ, as a function of angular fre- 0.7/pG [Pa]. quency normalized by the visco-inertial time τv. The line is (ext) the theoretical prediction, Eq. 9. In a step stress experiment with σ = σ0Θ(t), the solution to Eq. 10 is

Measurements for oscillating stresses have been per- γ(t) = γ∞ [1 − cos (Ωt)] , (11) formed with the same 3 samples and two different values of the gap: e = 300 µm and e = 600 µm (solid and with γ∞ = σ0/G. In the inertial regime Ωt  1 one open symbols of fig. 7, respectively). The stress ampli- finds the characteristic quadratic time dependence of the 1 σ0 2 tude was chosen in order to keep the strain amplitude strain, γ(t) ≈ 2 I t , while at later times the cos (Ωt) |γ˜ω| fixed at 20% for all ω. In fig. 7 the magnitude of term leads to a distinctive “ringing” behavior. Note how- the complex modulus |G∗| (main plot) and its phase δ ever that when friction is included, the strain oscillations (inset) are plotted as a function of angular frequency, to- eventually are damped. gether with the theoretical expectations given by Eq. 9. For an oscillatory imposed stress, the solution to Eq. 10 The data are in very good agreement with the theory for reads applied stresses with |σ˜ω| > 1 Pa. By contrast, when h ω i γ(t) =γ ˜ eiωt − cos (Ωt) − i sin (Ωt) , (12) lower stresses are applied to obtain the same strain am- ω Ω Stress-controlled shear cell 9

h 2i−1 σ˜ω ω  on several factors (cleanness of the air bearing, alignment withγ ˜ω = 1 − . This corresponds to a com- G Ω with respect to the vertical direction...) and that 1 Pa plex modulus whose magnitude and phase are should be regarded as a higher bound on its magnitude. h 2i |G∗| = G 1 − ω  Ω (13) δ = 0 . A second series of experiments is performed by im- posing oscillating stresses. The sample is prepared at a At small times t  Ω−1 we have again γ(t) ≈ higher crosslinker density (1:40 v/v), with e = 590 µm iωt  2 γ˜ω e − 1 − iωt , whereas at large t the strain exhibits and A = 2.5 cm . After curing the PDMS as described oscillations at the frequency Ω associated to inertia, su- above, an oscillating stress is applied and the amplitude perimposed to the elastic response at the forcing fre- of the strain oscillations is extracted from a sinusoidal fit quency ω: their amplitude is |σ˜ω|/G and they are in to γ(t). Figure 8c shows the frequency dependent elastic phase with the driving force, a distinctive feature of elas- modulus obtained from G(ω) = |σω/γω|, as a function of ticity. angular frequency. Experiments performed at different stress amplitudes spanning two decades within the lin- ear regime nicely overlap in the whole frequency range. 2. Experimental tests Additionally, data taken in the custom shear cell over- lap reasonably well with data acquired in a commercial We use polydimethylsiloxane (PDMS; Sylgard 184 by rheometer for a similar sample, thus validating the new Dow Corning), a popular elastomer whose elastic mod- setup. The slight differences that are seen are most likely ulus can be conveniently tuned by varying the crosslink due to the different protocols for preparing and loading density87, as a model elastic solid. Using a commercial the sample in the two devices. rheometer, we check that the linear regime extends up to at least γ = 40%, larger than the largest strain probed in our tests. A sample with a crosslinker/base volume ratio of 1:60 is used for the step stress measurements. In order to maximize the adhesion to the glass plates, the sample is prepared in situ: after adding the crosslinker, the fluid 1 0 - 1 a ) 0 . 2 b ) solution is placed between the two plates and cured in an oven at 90◦C for 90 minutes. The two plates with the 0 . 1

؇ 2 - 1 0 cured sample are then mounted on the shear cell. Dur- L a s e r m e t e r ing this operation, care is taken not to stress nor damage S p e c k l e d e t e c t o r 0 . 0 [ ؗ [ r a d / s the sample. Control measurements are run in a com- - 3 0 1 0 0 2 0 0 mercial rheometer, with the sample cured in situ under 1 0 0 . 1 1 1 0 1 0 0 8 t [ s ] 1 0 0 similar conditions. We perform a series of step stress ؇ 1 8 P a 1 8 0 P a c ) experiments, with a gap e = 570 µm (as measured be- - 4 ] 1 8 0 0 P a a 1 0 2 0 0 P a ( r h ) fore the experiment) and a sample surface A = 3.3 cm2. P k

[ Stresses of various amplitude σ are applied, the result-

0 | ing strain being recorded over time. Figure 8a shows a 1 0 - 5 * 1 0 G typical time series of strain data acquired simultaneously | with the commercial laser sensor (symbols) and with our 0 . 0 1 0 . 1 1 1 0 1 0 0 custom made optical setup (line), for σ0 = 275 Pa: note ؓ the excellent agreement between the two devices. After [ P a ] א an initial transient, lasting a few seconds, γ reaches a constant value γ∞ dictated by the elastic modulus. Note that the oscillations due to inertia predicted by Eq. 11 FIG. 8. (a) Time evolution of the strain following a step are not seen here, indicating that the (small) dissipation stress applied at t = 0, measured with both the commercial laser meter (black dots) and the home made optical setup of PDMS is sufficient to overdamp them. In fig. 8b, γ∞ estimated by averaging γ(t) over 5 minutes is plotted as (red line). (b) Response to a step stress. The asymptotic a function of the imposed stress. Each data point is an strain γ∞ is plotted against the amplitude of the applied step stress. Data are averaged over 8 experiments, the error bars average over 8 experiments, the error bars representing are the standard deviation over the experiment repetitions. data dispersion. The experimental data are very well fit- The solid line is an affine fit, yielding an elastic modulus G = ted by an affine law (red line): γ∞ = (σ0 − σoff )/G, 1620 Pa. (c) Frequency dependence of the elastic modulus with G = 1620 Pa the elastic modulus of the sample and in an oscillatory test, as measured in the custom shear cell σoff = 13 mPa an offset stress due to friction. We note at different stress amplitudes (symbols) and in a commercial that σoff is smaller than the typical friction stress esti- rheometer (line). mated from the experiments with an empty cell, which is of the order of 1 Pa. This discrepancy suggests that the exact value of the friction term depends sensitively Stress-controlled shear cell 10

C. Maxwell fluid 2. Experimental tests

1. Theoretical background As a model Maxwell fluid, we use a self-assembled transient network, comprising surfactant-stabilized mi- croemulsions reversibly linked by triblock copolymers, As an example of a viscoelastic fluid, we consider a 88 Maxwell fluid, for whichγ ˙ =σ/G ˙ + σ/η. Here G is the described in detail in . After loading the sample in the shear cell, we wait 5 minutes before applying a step stress, plateau modulus and η = GτM the viscosity, with τM the Maxwell relaxation time. Using this constitutive law, the in order to let the sample fully relax. Similarly, after each equation of motion, Eq. 4, yields: creep experiment a waiting time of 2 minutes is applied before starting the next one. Creep experiments are per- formed for different stress amplitudes, ranging from 1.5 d2γ˙ I dγ˙ σ(ext) (ext) to 150 Pa. The time evolution of the strain amplitude is I 2 + + Gγ˙ =σ ˙ + , (14) dt τM dt τM shown in fig. 9 for several applied stresses σ0. The data are very well fitted by the theoretical expression (Eq. 16) whose solutions again involve the characteristic frequency using the viscoelastic parameters of the Mawxell fluid, Ω = pG/I. τM and G, as fitting parameters. As shown in the inset The general solution for a step stress is of fig. 9, we find over the whole range of applied stresses a nearly constant value of the elastic plateau (G = 165±20 σ  t  0 −λ+t −λ−t γ(t) = c0 + + c+e + c−e , (15) Pa), in excellent agreement with the one measured in a G τM frequency sweep test in a conventional rheometer (180 Pa). A somehow poorer agreement between the shear cell   data and conventional rheometry is seen for the charac- where c = 1 − 1 , c = − 1 c ∓ √ c0+2 , and 0 τ 2 Ω2 ± 2 0 2 2 M 1−4τM Ω teristic relaxation time, which is found to decrease by a   1 p 2 2 factor of almost 2 as the applied stress increases. This λ± = 1 ± 1 − 4Ω τ . The regime of slowly re- 2τM M is most likely due to the fact that, for this sample, the laxing Maxwell fluids (as compared to the inertial time) relaxation time τM is close to the characteristic ringing corresponds to ΩτM  1. In this limit, one recovers 2 time due to inertia, 2π/Ω ≈ 0.3 s, which makes it diffi- γ ≈ σt /(2I), as for a purely viscous fluid (see Eq. 6). In cult to independently and reliably retrieve the two times the opposite limit ΩτM  1, the solution is: when fitting the data to Eq. 16. σ  t  t   γ(t) = 0 1 + − exp − cos(Ωt) , (16) G τM 2τM V. DYNAMIC LIGHT SCATTERING UNDER SHEAR for which three regimes may be distinguished. For We couple the shear cell to a custom made small an- −1 2 t  Ω , inertia dominates and γ = σ0t /(2I). For, gle light scattering setup82 and measure the microscopic −1 Ω  t  τM , the typical oscillations due to the elas- dynamics of a colloidal suspension of Brownian particles, tic part of the sample response are observed: γ(t) ≈ σ both at rest and under shear. The setup uses a CMOS G [1 − cos(Ωt)]. Finally, at long times t  τM the sample camera as a detector; the microscopic dynamics is quan- flows as a purely viscous fluid: γ(t) ≈ σ0t/η. 80,89 tified by the intensity correlation function g2 − 1: For the sake of completeness, we report here the equa- * + tions for an applied oscillating stress, although in the hIp(t + τ)Ip(t)iq following we test the Maxwell fluid only in step stress ex- g2(q, τ) − 1 = − 1 . (19) hIp(t)i hIp(t + τ)i periments. Focussing on the complex modulus calculated q q t from the oscillating part of the cell response, one finds: Here, Ip is the time-dependent intensity measured by the

 2 2 2 p − th pixel, < ··· >t denotes an average over the ex- 1− ω + ω Ω2 Ω4τ2 periment duration and < ··· > is an average over a set |G∗| = G M q s 2  2  of pixels corresponding to a small solid angle centered 1− ω − 1 + 1 Ω2 τ2 Ω2 ω2τ2 (17) M M around the direction associated to a scattering vector q, 1 Ω2 −1 tan δ = 1 . whose magnitude is q = 4πnλ sin(θ/2), with n the sol- ωτM Ω2−ω2− τ2 M vent , λ = 532 nm the laser wavelength and θ the scattering angle. −1 In the Ω  τM and Ω  ω limit where inertia is negli- The intensity correlation function is directly related gible Eq. 17 simplifies to yield the usual expressions for to the microscopic dynamics: for particles undergoing 2 a Maxwell fluid: , g2(q, τ) − 1 = exp(−2Dq τ), where D is the particle diffusion coefficient80. Under shear, − 1 ∗  1  2 the relative motion due to the affine displacement also |G | = G 1 + ω2τ 2 M (18) contributes to the decay of g − 1. Neglecting hy- tan δ = 1 . 2 ωτM drodynamic interactions and assuming an homogeneous Stress-controlled shear cell 11

1 1 . 5 P a 1 5 P a representative correlation functions; for a given magni- 1 0 3 3 P a 3 0 P a

] tude of the scattering vector, the decay time is shorter ] a 4 . 5 P a 7 5 P a s P 1 0 [ [ when q is oriented parallel to the shear velocity (open 7 . 5 P a 1 5 0 P a ׭ G -symbols), reflecting faster dynamics due to the contribu ؔ 2 0 . 1 1 0 tion of the affine shear field. The lines are fits to the data 1 1 0 1 0 0 -using Eq. 20: for both q orientations, an excellent agree ؓ 1 [ P a ] א

؇ ment is seen between the data and the model. We extract

from the fits the characteristic relaxation time τ ∗, defined ∗ ∗ by g2(q, τ ) − 1 = exp(−2), and plot τ vs q in the inset 0 . 1 ∗ of Fig. 10. For q = qyuˆx, τ is in excellent agreement with the theoretical prediction, dotted line; its q depen- dence is very close to a q−1 scaling, indicative of ballis- 0 . 0 1 tic motion and suggesting that over the range of probed q vectors and time scales the displacement due to the 0 . 0 1 0 . 1 1 1 0 affine deformation is significantly larger than that due to Brownian motion. This is confirmed by inspecting τ ∗(q) t [ s ] for a quiescent suspension (solid squares and solid line): for all q, the relaxation time due to Brownian motion is FIG. 9. Main plot: strain for a Maxwell sample as a func- larger than that under shear. The solid circles are the tion of time following a step stress with amplitude σ0 as given relaxation time of the dynamics in the vorticity direction by the labels. Open and filled symbols represent two consec- ∗ while shearing the sample. Overall, τ (qy) closely follows utive measurements on the same sample; lines are best fits the data for the quiescent suspension, demonstrating that to the data using Eq. 16, where the fitting parameters are the affine and non-affine components of the particle dy- the characteristic relaxation time, τM , and the plateau mod- ulus, G. Inset: Maxwell parameters extracted from the fits of namics can be resolved by dynamic light scattering. Only the experimental creep curves shown in the main plot, plot- at the smallest q vector do the data exhibit some devi- ted against the applied stress. Squares and left axis: plateau ations with respect to purely Brownian dynamics: most modulus; circles and right axis: relaxation time. The symbols likely, this discrepancy stems from slight deviations of are experimental data obtained in the custom shear cell, the the velocity field from the ideal one, due to the finite size lines are τM and G as measured in a commercial rheometer, of the sample. in a frequency sweep test at 10% strain amplitude.

VI. CONCLUSION shear flow in the xz plane (ˆux andu ˆz being the ve- locity and the velocity gradient direction, respectively), the particle dynamics are described by the displacement We have presented a custom-made stress-controlled field ∆r(r, γ,˙ τ) = ∆r(diff)(τ) + ∆r(aff)(r, γ,˙ τ), where shear cell in the linear translation plate-plate geometry, (aff) ∆r (r, γ,˙ τ) =γτ ˙ (r · uˆz)ˆux is the affine shear field, which can be coupled to both a microscope and a static whereas ∆r(diff)(τ) describes Brownian diffusion. As- and dynamic light scattering apparatus. The setup has suming that the two contributions are uncorrelated, the been successfully tested on a variety of samples represen- intensity correlation function can be expressed as90: tative of simple fluids, ideal solids and viscoelastic fluids; its coupling to dynamic light scattering measurements 2  e  2  g2(q, γ,˙ τ) − 1 = sinc qxγτ˙ exp −2Dq τ , (20) has been illustrated by measuring the dynamics of Brow- 2 nian particles in a suspension at rest or under shear. The where qx = q·uˆx is the projection of the scattering vector main features of the cell include a gap that can be ad- along the velocity direction. Note that no contribution justed down to 100 µm keeping the parallelism to within of the applied shear on the dynamics is expected when 0.3 mrad, the acquisition of strain data at up to 1000 Hz selecting a scattering vector q = qyuˆy parallel to the with very good precision (typically better than 0.01%) vorticity direction. and for strains as large as 104%, and the possibility of im- We measure the dynamics of a suspension of posing a user-defined, time-varying stress, ranging from polystyrene particles with diameter 2a = 1.2µm (Invitro- 0.1 Pa (limited by residual friction) to 10 kPa, with a gen Molecular Probes), suspended at a weight fraction of frequency bandpass 0 ≤ ω ≤ 4 rads−1. Most of these 0.01% in a 1:1 v/v mixture of H2O and D2O that matches specifications meet or even exceed those of a commercial the density of polystyrene, to avoid sedimentation. A rheometer. Among the limitations intrinsic to the setup, time series of images of the scattered light is acquired the most stringent is probably inertia, which restricts vis- 91 using the scheme described in , both at rest and while cosity measurements to η & 1 Pa s. Other limitations in imposing a constant shear rateγ ˙ = 0.03s−1. The inten- the current implementation could be improved: for ex- sity correlation function is calculated using Eq. 19, for ample, the upper stress limit could be pushed to 100 kPa several q vectors both along the velocity and the vortic- by changing the magnetic actuator and the cutoff fre- −1 ity direction (qx and qy, respectively). Figure 10 shows quency ω ≈ 4 rads could be increased by modifying Stress-controlled shear cell 12

is gratefully acknowledged. 1 . 0 Q u i e s c e n t S h e a r - q ] y Bibliography s

[ S h e a r - q 1 x *

׀ 8 . 0 ؔ 1R. G. Larson, The Structure and Rheology of Complex Fluids, - 1 1st ed. (Oxford University Press, New York, 1998).

1 0 . 6 0 . 1 2 S. M. Rubinstein, Colby, and Ralph H, , 1st ed. -

1 2 3 4

) (Oxford University Press, Oxford ; New York, 2003).

3

D. T. N. Chen, Q. Wen, P. A. Janmey, J. C. Crocker, and 1 - ؔ

0 . 4 µ

, q [ m ] q x q y A. G. Yodh, “Rheology of Soft Materials,” Annual Review of q

( Condensed Matter Physics 1, 301–322 (2010). 2 0 . 9 g 0 . 2 4K. Binder and W. Kob, Glassy Materials and Disordered Solids: 1 . 4 An Introduction to Their Statistical Mechanics (World Scientific, 2 . 0 2011). 0 . 0 3 . 1 5I. Cantat, S. Cohen-Addad, F. Elias, F. Graner, R. H¨ohler, R. Flatman, O. Pitois, F. Rouyer, and A. Saint-Jalmes, Foams: 0 . 0 1 0 . 1 1 1 0 1 0 0 Structure and Dynamics (OUP Oxford, 2013). 6P. A. Janmey, P. C. Georges, and S. Hvidt, “Basic Rheology for .Biologists,” (Academic Press, 2007) pp. 1–27 ؔ [ s ] 7G. I. Taylor, “The Formation of Emulsions in Definable Fields of Flow,” Proceedings of the Royal Society of London. Series A, FIG. 10. Main plot: representative intensity correlation func- Containing Papers of a Mathematical and Physical Character tions for a sheared suspension of Brownian particles, for scat- 146, 501–523 (1934). 8T. Gallot, C. Perge, V. Grenard, M.-A. Fardin, N. Taberlet, and tering vectors parallel (qx, open symbols) and perpendicular S. Manneville, “Ultrafast ultrasonic imaging coupled to rheome- (q , solid symbols) to the shear direction. Data are fitted with y try: Principle and illustration,” Review of Scientific Instruments Eq. 20 (dashed and solid lines for qx and qy, respectively) with ∗ 84, 045107 (2013). D andγ ˙ as fitting parameters. Inset: characteristic time τ 9P. T. Callaghan, “Rheo-NMR: nuclear magnetic resonance and as defined in the text for the two orientations of q and for a the rheology of complex fluids,” Rep. Prog. Phys. 62, 599 (1999). quiescent sample, as a function of the scattering vector. The 10V. B. Tolstoguzov, A. I. Mzhel’sky, and V. Y. Gulov, “Deforma- lines are the predictions of Eq. 20, using the nominal values tion of emulsion droplets in flow,” and Polymer Science of the diffusion coefficient, D = 0.36 µm2s−1, and the shear 252, 124–132 (1974). rate,γ ˙ = 0.03 s−1. 11T. Hashimoto, T. Takebe, and S. Suehiro, “Apparatus to Mea- sure Small-Angle Light Scattering Profiles of Polymers under Shear Flow,” Polym J 18, 123–130 (1986). 12J. W. van Egmond, D. E. Werner, and G. G. Fuller, “Time- the current generator design. Finally, we remind that dependent small-angle light scattering of shear-induced concen- care must be taken in aligning both the plate parallelism tration fluctuations in polymer solutions,” The Journal of Chem- ical Physics 96, 7742 (1992). and the horizontality of the setup, which makes its use 13 less straightforward than that of a commercial rheome- E. Di Cola, C. Fleury, P. Panine, and M. Cloitre, “Steady Shear Flow Alignment and Rheology of Lamellae-Forming ABC Tri- ter. These drawbacks are more than offset by the advan- block Copolymer Solutions: Orientation, Defects, and Disorder,” tages afforded by our shear cell: the possibility of easily Macromolecules 41, 3627–3635 (2008). changing the plates, e.g. to optimize them against slip- 14D. C. F. Wieland, V. M. Garamus, T. Zander, C. Krywka, page or to replace them when the surface quality doesn’t M. Wang, A. Dedinaite, P. M. Claesson, and R. Willumeit- R¨omer,“Studying solutions at high shear rates: a dedicated mi- meet anymore the stringent requirements for microscopy crofluidics setup,” Journal of Synchrotron Radiation 23, 480–486 or light scattering; the open design that leaves full access (2016). on both sides, thus making it suitable for both inverted 15D. Beysens, M. Gbadamassi, and L. Boyer, “Light-scattering and upright microscopes and for small angle light scatter- study of a critical mixture with shear flow,” Physical Review ing; the linear translation geometry that insures uniform Letters 43, 1253 (1979). 16J. L¨augerand W. Gronski, “A melt rheometer with integrated stress in the (optically optimal) parallel plate configura- small angle light scattering,” Rheologica acta 34, 70–79 (1995). tion; the flexibility provided by the choice of the orien- 17B. J. Ackerson, J. van der Werff, and C. G. de Kruif, “Hard- tation in the vertical or horizontal plane; and, last but sphere dispersions: Small-wave-vector structure-factor measure- not least, the reduced cost as compared to a commercial ments in a linear shear flow,” Physical Review A 37, 4819 (1988). 18 rheometer (about 4000 Euros including a PC and an air Y. D. Yan and J. K. G. Dhont, “Shear-induced structure distor- tion in nonaqueous dispersions of charged colloidal spheres via filter for the air bearing rail). light scattering,” Physica A: Statistical Mechanics and its Appli- cations 198, 78–107 (1993). 19R. Linemann, J. L¨auger,G. Schmidt, K. Kratzat, and W. Rich- ACKNOWLEDGMENTS tering, “Linear and nonlinear rheology of micellar solutions in the isotropic, cubic and hexagonal phase probed by rheo-small-angle light scattering,” Rheologica acta 34, 440–449 (1995). We thank J. Barbat and E. Alibert for help in in- 20T. Kume, K. Asakawa, E. Moses, K. Matsuzaka, and strumentation and S. Arora for providing us with the T. Hashimoto, “A new apparatus for simultaneous observation of optical microscopy and small-angle light scattering measure- bridged microemulsion sample. Funding from the French ments of polymers under shear flow,” Acta polymerica 46, 79–85 ANR (project FAPRES, ANR-14-CE32-0005-01) and the (1995). E.U. (Marie Sklodowska-Curie ITN Supolen, no 607937) 21J. Vermant, L. Raynaud, J. Mewis, B. Ernst, and G. G. Fuller, Stress-controlled shear cell 13

“Large-scale bundle ordering in sterically stabilized latices,” 1944–1948 (2004). Journal of colloid and interface science 211, 221–229 (1999). 41S. Reinicke, M. Karg, A. Lapp, L. Heymann, T. Hellweg, and 22P. Varadan and M. J. Solomon, “Shear-Induced Microstructural H. Schmalz, “Flow-Induced Ordering in Cubic Gels Formed by Evolution of a Thermoreversible Colloidal Gel,” Langmuir 17, P2VP- b -PEO- b -P(GME- co -EGE) Triblock Terpolymer Mi- 2918–2929 (2001). celles: A Rheo-SANS Study,” Macromolecules 43, 10045–10054 23L. Porcar, W. A. Hamilton, P. D. Butler, and G. G. (2010). Warr, “Scaling of Shear-Induced Transformations in Membrane 42M. A. Torija, S.-H. Choi, T. P. Lodge, and F. S. Bates, “Large Phases,” Physical Review Letters 89 (2002), 10.1103/Phys- Amplitude Oscillatory Shear of Block Copolymer Spheres on a RevLett.89.168301. Body-Centered Cubic Lattice: Are Micelles Like Metals?” The 24L. Ramos and F. Molino, “Shear Melting of a Hexagonal Colum- Journal of Physical Chemistry B 115, 5840–5848 (2011). nar Crystal by Proliferation of Dislocations,” Physical Review 43E. Tamborini, L. Cipelletti, and L. Ramos, “Plasticity of a Col- Letters 92 (2004), 10.1103/PhysRevLett.92.018301. loidal Polycrystal under Cyclic Shear,” Phys. Rev. Lett. 113, 25R. Scirocco, J. Vermant, and J. Mewis, “Effect of the viscoelas- 078301 (2014). ticity of the suspending fluid on structure formation in suspen- 44M. E. Helgeson, P. A. Vasquez, E. W. Kaler, and N. J. Wagner, sions,” Journal of Non-Newtonian Fluid Mechanics 117, 183–192 “Rheology and spatially resolved structure of cetyltrimethylam- (2004). monium bromide wormlike micelles through the shear banding 26I. Cohen, T. G. Mason, and D. A. Weitz, “Shear-Induced Con- transition,” Journal of Rheology 53, 727 (2009). figurations of Confined Colloidal Suspensions,” Physical Review 45D. Derks, H. Wisman, A. van Blaaderen, and A. Imhof, “Con- Letters 93 (2004), 10.1103/PhysRevLett.93.046001. focal microscopy of colloidal dispersions in shear flow using a 27L. Yang, R. H. Somani, I. Sics, B. S. Hsiao, R. Kolb, H. Fruitwala, counter-rotating cone–plate shear cell,” Journal of Physics: Con- and C. Ong, “Shear-Induced Crystallization Precursor Studies densed Matter 16, S3917–S3927 (2004). in Model Polyethylene Blends by in-Situ Rheo-SAXS and Rheo- 46R. Angelico, C. O. Rossi, L. Ambrosone, G. Palazzo, WAXD,” Macromolecules 37, 4845–4859 (2004). K. Mortensen, and U. Olsson, “Ordering fluctuations in a shear- 28Y. L. Wu, J. H. J. Brand, J. L. A. van Gemert, J. Verkerk, banding wormlike micellar system,” Physical Chemistry Chemi- H. Wisman, A. van Blaaderen, and A. Imhof, “A new parallel cal Physics 12, 8856 (2010). plate shear cell for in situ real-space measurements of complex 47M. E. Helgeson, L. Porcar, C. Lopez-Barron, and N. J. Wag- fluids under shear flow,” Review of Scientific Instruments 78, ner, “Direct Observation of Flow-Concentration Coupling in 103902 (2007). a Shear-Banding Fluid,” Physical Review Letters 105 (2010), 29Y. Kosaka, M. Ito, Y. Kawabata, and T. Kato, “Lamellar-to- 10.1103/PhysRevLett.105.084501. Onion Transition with Increasing Temperature under Shear Flow 48T. J. Ober, J. Soulages, and G. H. McKinley, “Spatially resolved in a Nonionic Surfactant/Water System,” Langmuir 26, 3835– quantitative rheo-optics of complex fluids in a microfluidic de- 3842 (2010). vice,” Journal of Rheology (1978-present) 55, 1127–1159 (2011). 30V. Grenard, N. Taberlet, and S. Manneville, “Shear-induced 49A. K. Gurnon, C. R. Lopez-Barron, A. P. R. Eberle, L. Porcar, structuration of confined carbon black gels: steady-state features and N. J. Wagner, “Spatiotemporal stress and structure evolu- of vorticity-aligned flocs,” Soft Matter 7, 3920 (2011). tion in dynamically sheared polymer-like micellar solutions,” Soft 31L. Ramos, A. Laperrousaz, P. Dieudonn´e, and C. Ligoure, Matter 10, 2889–2898 (2014). “Structural Signature of a Brittle-to-Ductile Transition in Self- 50V. Chikkadi, D. Miedema, M. Dang, B. Nienhuis, and P. Schall, Assembled Networks,” Physical Review Letters 107 (2011), “Shear Banding of Colloidal Glasses: Observation of a Dynamic 10.1103/PhysRevLett.107.148302. First-Order Transition,” Physical Review Letters 113 (2014), 32C. R. L´opez-Barr´on,L. Porcar, A. P. R. Eberle, and N. J. Wag- 10.1103/PhysRevLett.113.208301. ner, “Dynamics of Melting and Recrystallization in a Polymeric 51A. Basu, Q. Wen, X. Mao, T. C. Lubensky, P. A. Janmey, and Micellar Crystal Subjected to Large Amplitude Oscillatory Shear A. G. Yodh, “Nonaffine Displacements in Flexible Polymer Net- Flow,” Phys. Rev. Lett. 108, 258301 (2012). works,” Macromolecules 44, 1671–1679 (2011). 33R. Akkal, N. Cohaut, M. Khodja, T. Ahmed-Zaid, and 52M.-Y. Nagazi, G. Brambilla, G. Meunier, P. Marguer`es,J.-N. F. Bergaya, “Rheo-SAXS investigation of organoclay water in oil P´eri´e, and L. Cipelletti, “Space-resolved diffusing wave spec- emulsions,” Colloids and Surfaces A: Physicochemical and Engi- troscopy measurements of the macroscopic deformation and the neering Aspects 436, 751–762 (2013). microscopic dynamics in tensile strain tests,” ArXiv e-prints 34L. Gentile, M. A. Behrens, L. Porcar, P. Butler, N. J. Wag- (2016), arXiv:1603.06384 [cond-mat.soft]. ner, and U. Olsson, “Multilamellar Vesicle Formation from a 53P. Hebraud, F. Lequeux, J. P. Munch, and D. J. Pine, “Yielding Planar Lamellar Phase under Shear Flow,” Langmuir 30, 8316– and Rearrangements in Disordered Emulsions,” Phys. Rev. Lett. 8325 (2014). 78, 4657–4660 (1997). 35J. M. Kim, A. P. Eberle, A. K. Gurnon, L. Porcar, and 54G. Petekidis, A. Moussaid, and P. N. Pusey, “Rearrangements N. J. Wagner, “The microstructure and rheology of a model, in hard-sphere glasses under oscillatory shear strain,” Phys. Rev. thixotropic nanoparticle gel under steady shear and large am- E 66, 051402 (2002). plitude oscillatory shear (LAOS),” Journal of Rheology (1978- 55G. Petekidis, P. N. Pusey, A. Moussaıd,¨ S. Egelhaaf, and present) 58, 1301–1328 (2014). W. C. K. Poon, “Shear-induced yielding and ordering in con- 36N. Y. C. Lin, X. Cheng, and I. Cohen, “Biaxial shear of con- centrated particle suspensions,” Physica A: Statistical Mechan- fined colloidal hard spheres: the structure and rheology of the ics and its Applications Invited Papers from the 21th IUPAP vorticity-aligned string phase,” Soft Matter 10, 1969 (2014). International Conference on St atistical Physics, 306, 334–342 37A. D. Gopal and D. J. Durian, “Nonlinear bubble dynamics in a (2002). slowly driven foam,” Physical Review Letters 75, 2610 (1995). 56T. Bauer, J. Oberdisse, and L. Ramos, “Collective Rearrange- 38R. H¨ohler,S. Cohen-Addad, and H. Hoballah, “Periodic non- ment at the Onset of Flow of a Polycrystalline Hexagonal Colum- linear bubble motion in aqueous foam under oscillating shear nar Phase,” Physical Review Letters 97 (2006), 10.1103/Phys- strain,” Physical review letters 79, 1154 (1997). RevLett.97.258303. 39F. R. Molino, J.-F. Berret, G. Porte, O. Diat, and P. Lind- 57R. Besseling, E. Weeks, A. B. Schofield, and W. C. K. Poon, ner, “Identification of flow mechanisms for a soft crystal,” The “Three-Dimensional Imaging of Colloidal Glasses under Steady European Physical Journal B-Condensed Matter and Complex Shear,” Phys. Rev. Lett. 99, 028301–028301 (2007). Systems 3, 59–72 (1998). 58P. Schall, D. A. Weitz, and F. Spaepen, “Structural Rearrange- 40P. Schall, I. Cohen, D. A. Weitz, and F. Spaepen, “Visualiza- ments That Govern Flow in Colloidal Glasses,” Science 318, tion of Dislocation Dynamics in Colloidal Crystals,” Science 305, 1895–1899 (2007). Stress-controlled shear cell 14

59P. A. Smith, G. Petekidis, S. U. Egelhaaf, and W. C. K. 085702–4 (2009). Poon, “Yielding and crystallization of colloidal gels under os- 76L. Cipelletti, G. Brambilla, S. Maccarrone, and S. Caroff, “Si- cillatory shear,” Physical Review E 76 (2007), 10.1103/Phys- multaneous measurement of the microscopic dynamics and the RevE.76.041402. mesoscopic displacement field in soft systems by speckle imag- 60J. Zausch, J. Horbach, M. Laurati, S. U. Egelhaaf, J. M. Brader, ing,” Optics Express 21, 22353–22366 (2013). T. Voigtmann, and M. Fuchs, “From equilibrium to steady state: 77D. A. Weitz and D. J. Pine, “Diffusing-wave spectroscopy,” in the transient dynamics of colloidal liquids under shear,” Journal Dynamic Light scattering, edited by W. Brown (Oxford, 1993) of Physics: Condensed Matter 20, 404210 (2008). pp. 652–720. 61N. Koumakis, M. Laurati, S. U. Egelhaaf, J. F. Brady, and 78X.-L. Wu, D. J. Pine, P. M. Chaikin, J. S. Huang, and D. A. G. Petekidis, “Yielding of Hard-Sphere Glasses during Start- Weitz, “Diffusing-wave spectroscopy in a shear flow,” Journal of Up Shear,” Physical Review Letters 108 (2012), 10.1103/Phys- the Optical Society of America B 7, 15 (1990). RevLett.108.098303. 79M. Erpelding, A. Amon, and J. Crassous, “Diffusive wave spec- 62 D. Denisov, M. T. Dang, B. Struth, G. Wegdam, and P. Schall, troscopy applied to the spatially resolved deformation of a solid,” “Resolving structural modifications of colloidal glasses by com- Phys. Rev. E 78, 046104 (2008). bining x-ray scattering and rheology,” Scientific Reports 3 80B. J. Berne and R. Pecora, Dynamic Light Scattering (Wiley, (2013), 10.1038/srep01631. New York, 1976). 63 N. Y. C. Lin, S. Goyal, X. Cheng, R. N. Zia, F. A. Escobedo, and 81N. Ali, D. Roux, L. Cipelletti, and F. Caton, “RheoSpeckle: I. Cohen, “Far-from-equilibrium sheared colloidal liquids: Dis- a new tool to investigate local flow and microscopic dy- entangling relaxation, advection, and shear-induced diffusion,” namics of soft matter under shear,” ArXiv e-prints (2016), Physical Review E 88 (2013), 10.1103/PhysRevE.88.062309. arXiv:1604.07611 [cond-mat.soft]. 64 E. D. Knowlton, D. J. Pine, and L. Cipelletti, “A microscopic 82E. Tamborini and L. Cipelletti, “Multiangle static and dynamic view of the yielding transition in concentrated emulsions,” Soft light scattering in the intermediate scattering angle range,” Re- Matter 10, 6931–6940 (2014). view of Scientific Instruments 83, 093106 (2012). 65 T. Sentjabrskaja, P. Chaudhuri, M. Hermes, W. C. K. Poon, 83A. Madsen, R. L. Leheny, H. Guo, M. Sprung, and O. Czakkel, J. Horbach, S. U. Egelhaaf, and M. Laurati, “Creep and flow “Beyond simple exponential correlation functions and equilib- of glasses: strain response linked to the spatial distribution of rium dynamics in x-ray photon correlation spectroscopy,” New dynamical heterogeneities,” Scientific Reports 5, 11884 (2015). Journal of Physics 12, 055001 (2010). 66 A. P. Eberle and L. Porcar, “Flow-SANS and Rheo-SANS applied 84P. T. Tokumaru and P. E. Dimotakis, “Image Correlation Ve- to soft matter,” Current Opinion in Colloid & Interface Science locimetry,” Experiments in Fluids 19, 1–15 (1995). 17, 33–43 (2012). 85V. Viasnoff, F. Lequeux, and D. J. Pine, “Multispeckle diffusing- 67 A. Dhinojwala and S. Granick, “Micron-gap rheo-optics with par- wave spectroscopy: A tool to study slow relaxation and time- allel plates,” The Journal of chemical physics 107, 8664–8667 dependent dynamics,” Rev. Sci. Instrum. 73, 2336–2344 (2002). (1997). 86J. Rodr´ıguez-L´opez, L. Elvira, F. Montero de Espinosa Freijo, 68 J. Goyon, A. Colin, G. Ovarlez, A. Ajdari, and L. Bocquet, and J. de Vicente, “Using ultrasounds for the estimation of “Spatial cooperativity in soft glassy flows,” Nature 454, 84–87 the misalignment in plate–plate torsional rheometry,” Journal (2008). of Physics D: Applied Physics 46, 205301 (2013). 69 R. Besseling, L. Isa, E. R. Weeks, and W. C. Poon, “Quantitative 87E. Gutierrez and A. Groisman, “Measurements of Elastic Mod- imaging of colloidal flows,” Advances in Colloid and Interface uli of Silicone Gel Substrates with a Microfluidic Device,” PLoS Science 146, 1–17 (2009). ONE 6, e25534 (2011). 70 D. Chen, D. Semwogerere, J. Sato, V. Breedveld, and E. R. 88E. Michel, M. Filali, R. Aznar, G. Porte, and J. Appell, “Per- Weeks, “Microscopic structural relaxation in a sheared super- colation in a Model Transient Network: Rheology and Dynamic cooled colloidal liquid,” Phys. Rev. E 81, 011403 (2010). Light Scattering,” Langmuir 16, 8702–8711 (2000). 71 J.-B. Boitte, C. Vizca¨ıno,L. Benyahia, J.-M. Herry, C. Michon, 89A. Duri, H. Bissig, V. Trappe, and L. Cipelletti, “Time-resolved- and M. Hayert, “A novel rheo-optical device for studying complex correlation measurements of temporally heterogeneous dynam- fluids in a double shear plate geometry,” Review of Scientific ics,” Phys. Rev. E 72, 051401–17 (2005/11/00/). Instruments 84, 013709 (2013). 90S. Aime and L. Cipelletti, (2016), in preparation. 72 H. K. Chan and A. Mohraz, “A simple shear cell for the direct 91A. Philippe, S. Aime, V. Roger, R. Jelinek, G. Prvot, L. Berthier, visualization of step-stress deformation in soft materials,” Rheo- and L. Cipelletti, “An efficient scheme for sampling fast dynam- logica Acta 52, 383–394 (2013). ics at a low average data acquisition rate,” Journal of Physics: 73 N. Y. C. Lin, J. H. McCoy, X. Cheng, B. Leahy, J. N. Israelachvili, Condensed Matter 28, 075201 (2016). and I. Cohen, “A multi-axis confocal rheoscope for studying shear 92E. Gutierrez and A. Groisman, “Measurements of Elastic Mod- flow of structured fluids,” Review of Scientific Instruments 85, uli of Silicone Gel Substrates with a Microfluidic Device,” PLoS 033905 (2014). ONE 6, e25534 (2011). 74 L. Cipelletti, H. Bissig, V. Trappe, P. Ballesta, and S. Mazoyer, 93C. Baravian and D. Quemada, “Using instrumental inertia in “Time-resolved correlation: a new tool for studying temporally controlled stress rheometry,” Rheologica acta 37, 223–233 (1998). heterogeneous dynamics,” J. Phys.: Condens. Matter 15, S257– 94D. Bonn, J. Paredes, M. M. Denn, L. Berthier, T. Divoux, and S262 (2003). S. Manneville, “Yield Stress Materials in Soft Condensed Mat- 75 A. Duri, D. A. Sessoms, V. Trappe, and L. Cipelletti, “Resolv- ter,” arXiv:1502.05281 [cond-mat] (2015). ing Long-Range Spatial Correlations in Jammed Colloidal Sys- tems Using Photon Correlation Imaging,” Phys. Rev. Lett. 102,