probing interfacial dynamics and mechanics using …dennin/preprints/bdse14-submerged...probing...

16
Probing interfacial dynamics and mechanics using submerged particle microrheology. II. Experiment Thomas Boatwright, Michael Dennin, Roie Shlomovitz, Arthur A. Evans, and Alex J. Levine Citation: Physics of Fluids 26, 071904 (2014); doi: 10.1063/1.4887084 View online: http://dx.doi.org/10.1063/1.4887084 View Table of Contents: http://scitation.aip.org/content/aip/journal/pof2/26/7?ver=pdfcov Published by the AIP Publishing Articles you may be interested in Probing interfacial dynamics and mechanics using submerged particle microrheology. I. Theory Phys. Fluids 26, 071903 (2014); 10.1063/1.4886996 Interface shear microrheometer with an optically driven oscillating probe particle Rev. Sci. Instrum. 82, 094702 (2011); 10.1063/1.3627410 Multidepth, multiparticle tracking for active microrheology using a smart camera Rev. Sci. Instrum. 82, 033712 (2011); 10.1063/1.3567801 Laser tweezer microrheology of a colloidal suspension J. Rheol. 50, 77 (2006); 10.1122/1.2139098 Molecular dynamics simulation of optically trapped colloidal particles at an oil-water interface J. Chem. Phys. 121, 4292 (2004); 10.1063/1.1779569 This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP: 128.200.29.82 On: Thu, 17 Sep 2015 01:30:10

Upload: others

Post on 14-Jul-2020

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Probing interfacial dynamics and mechanics using …dennin/preprints/BDSE14-submerged...Probing interfacial dynamics and mechanics using submerged particle microrheology. II. Experiment

Probing interfacial dynamics and mechanics using submerged particle microrheology.II. ExperimentThomas Boatwright, Michael Dennin, Roie Shlomovitz, Arthur A. Evans, and Alex J. Levine Citation: Physics of Fluids 26, 071904 (2014); doi: 10.1063/1.4887084 View online: http://dx.doi.org/10.1063/1.4887084 View Table of Contents: http://scitation.aip.org/content/aip/journal/pof2/26/7?ver=pdfcov Published by the AIP Publishing Articles you may be interested in Probing interfacial dynamics and mechanics using submerged particle microrheology. I. Theory Phys. Fluids 26, 071903 (2014); 10.1063/1.4886996 Interface shear microrheometer with an optically driven oscillating probe particle Rev. Sci. Instrum. 82, 094702 (2011); 10.1063/1.3627410 Multidepth, multiparticle tracking for active microrheology using a smart camera Rev. Sci. Instrum. 82, 033712 (2011); 10.1063/1.3567801 Laser tweezer microrheology of a colloidal suspension J. Rheol. 50, 77 (2006); 10.1122/1.2139098 Molecular dynamics simulation of optically trapped colloidal particles at an oil-water interface J. Chem. Phys. 121, 4292 (2004); 10.1063/1.1779569

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

128.200.29.82 On: Thu, 17 Sep 2015 01:30:10

Page 2: Probing interfacial dynamics and mechanics using …dennin/preprints/BDSE14-submerged...Probing interfacial dynamics and mechanics using submerged particle microrheology. II. Experiment

PHYSICS OF FLUIDS 26, 071904 (2014)

Probing interfacial dynamics and mechanics usingsubmerged particle microrheology. II. Experiment

Thomas Boatwright,1 Michael Dennin,1,a) Roie Shlomovitz,2

Arthur A. Evans,2,3 and Alex J. Levine2,4

1Department of Physics and Astronomy, University of California, Irvine,California 92697, USA2Department of Chemistry and Biochemistry, University of California, Los Angeles,California 90095, USA3Department of Physics, University of Massachusetts, Amherst, Massachusetts 01003, USA4Department of Physics and Astronomy, University of California, Los Angeles, California90095, USA and California Nanosystems Institute, University of California, Los Angeles,California 90095, USA

(Received 24 January 2012; accepted 21 April 2014; published online 17 July 2014)

A non-contact microrheological technique to probe the mechanics of the air/waterinterface is explored. Polystyrene spheres dissolved in water are trapped with anoptical tweezer near the free surface of water, allowing the response functions of theparticles to be measured as a function of the distance from the air/water interface.These measurements show that at the surface, the imaginary part of the responsefunction increases by approximately 30% from the Stokes value measured in thebulk. As the particle is moved away from the surface via an optical trap, the responsefunction returns to the bulk value. The method is tested by comparing the responsefunction of particles near a rigid wall to the theory developed by Faxen. A newlydeveloped hydrodynamic theory is used to explain the results at the free interfacethrough a calculation of the linear response function as a function of depth. Theseresults show a range of sensitivity that can be utilized to study the microrheology ofa Langmuir monolayer without distorting its structure. C© 2014 AIP Publishing LLC.[http://dx.doi.org/10.1063/1.4887084]

I. INTRODUCTION

The characterization of interfacial mechanical properties remains an important experimentalchallenge with a wide range of applications. There are two distinct experimental environments inwhich interfacial measurements are of interest: air/water interfaces and fluid/fluid interfaces. Theformer is typically found in the study of Langmuir monolayers, two-dimensional layers of moleculesat the air/water interface,1, 2 while the latter is of interest in biological systems in which lipid bilayersare common as part of the cell membrane and intracellular structures. Langmuir monolayers arerelevant in a number of technological applications and as model systems for a range of biologicalproblems and general studies of two-dimensional phase behavior. While there is a long history ofmacroscopic measurements of their mechanical properties,3 these techniques are difficult to adaptto in situ measurements in biological systems and often do not provide critical information aboutlocal properties in highly heterogeneous systems. Consequently, there has been a strong interest indeveloping interfacial microrheological measurements.

Microrheology4–10 is a method that uses the observed displacement fluctuations (Brownianmotion) of microscopic probes to extract that medium’s rheological properties, by applying thefluctuation dissipation theorem.11 From that well-known result, the observed fluctuation spectrumreports on the frequency-dependent response function of that particle to an applied force. The

a)Electronic mail: [email protected]

1070-6631/2014/26(7)/071904/15/$30.00 C©2014 AIP Publishing LLC26, 071904-1

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

128.200.29.82 On: Thu, 17 Sep 2015 01:30:10

Page 3: Probing interfacial dynamics and mechanics using …dennin/preprints/BDSE14-submerged...Probing interfacial dynamics and mechanics using submerged particle microrheology. II. Experiment

071904-2 Boatwright et al. Phys. Fluids 26, 071904 (2014)

remaining step in the analysis requires one to determine that response function in terms of thehydrodynamic modes of the system (in the generalized sense11) and their associated moduli. Theprincipal advantages of using this indirect rheological measurement stem from the fact that thetechnique does not require the active deformation of large bulk samples of the material and, byusing only thermal forces, it can be applied to the most fragile of structures. These propertiesmake microrheology useful in the study of biopolymer networks in vivo12 and in situ in livingcells.13–15 It also allows one to probe the rheology of fragile structures that have no three-dimensionalrealization—Langmuir monolayers16, 17 and lipid membranes—since one does not have to couplethe system to a macroscopic rheometer. Interfaces have also been studied with related techniquesat liquid-solid interfaces, measuring the effect of surface properties on diffusion.18, 19 The centralchallenge for these microrheological studies of lower dimensional, i.e., interfacial systems, is thedevelopment of the necessary theoretical framework20, 21 to compute the response function of aparticle attached to the monolayer. This analysis is complicated by the role of a three-phase contactline at the particle and the lack of understanding of local structural perturbation of the monolayercaused by the presence of the particle.

The key evidence that these effects (or potentially others) represent a serious problem in thedevelopment of interfacial microrheology is that there is to date poor agreement between microrheo-logical measurements and macroscopic ones associated with, e.g., a surface Couette device. Variousone-22, 23 and two-particle24 microrheology experiments on monolayers and interfaces25, 26 are re-ported in the literature. These results typically differ from those obtained by macroscopic approachesby at least a few orders of magnitude27–29 with the microrheologically obtained results giving con-sistently smaller moduli.30–33 However, both sets of measurements report similar changes in surfacerheology with changes in, e.g., area pressure, suggesting that microscopic and macroscopic mea-surements are observing the same phenomena. In fact, once a single multiplicative correction ismade, the moduli measured via fluctuations appear to change in a quantitatively consistent mannerwith other mechanical measurements as the material undergoes structural phase transitions inducedby area pressure.

We expect that part of the discrepancy between microscopic fluctuation data and those obtainedfrom more traditional, active mechanical measurements is the challenge associated with correctlycharacterizing the response function, in particular with understanding the nature of the couplingbetween the probe particle and the monolayer. Accurately computing the response function involvesaddressing a number of physics issues that have been identified for particles embedded in interfaces,including the role of the subphase,30 the contact angle between the particle and the interface,34 andchanges in the monolayer itself induced by the particles.24 Because it minimizes these effects, themost promising techniques for probes embedded in a monolayer involve active methods that focuson thin-disks.16, 17

In this article, we report on an alternate, non-contact approach to microrheology of interfacesby observing the fluctuations of a probe not in the monolayer itself, but a few particle radii into thesubphase. This method immediately eliminates the issues associated with tracer-induced structuralperturbations of the monolayer and, by avoiding the complexities of the three-phase contact linesimplifies the nature of the coupling between the monolayer and the probe. Of course, one sacrificesthe coupling strength between the probe and monolayer in order to access these advantages, andintroduces a new set of physics that requires understanding. One of the key points of the workreported here is to show that the reduction of coupling strength and the additional new physics doesnot present an insurmountable obstacle. Indeed, we show that there is a significant and quantitativelyunderstandable difference between the fluctuation measurements beneath a free surface that does notsupport shear stresses (free air/water interface) and an infinitely rigid (water/glass interface) one. Allrheologically interesting monolayers fall between these two extremes so these measurements set therange of all possible surface rheology outcomes using submerged particle microrheology. By usingthese extreme cases we also experimentally verify the theoretical analysis of the response functionscarried out in Paper I of this series35 that will underpin all future submerged particle explorations ofmonolayers and membranes.

In brief, we use a weak laser trap combined with the back focal plane displacement detectionscheme described in Refs. 36 and 37 both to hold the particle at a fixed depth below the interface and

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

128.200.29.82 On: Thu, 17 Sep 2015 01:30:10

Page 4: Probing interfacial dynamics and mechanics using …dennin/preprints/BDSE14-submerged...Probing interfacial dynamics and mechanics using submerged particle microrheology. II. Experiment

071904-3 Boatwright et al. Phys. Fluids 26, 071904 (2014)

to measure its small (i.e., few nanometer scale) fluctuations in the plane parallel to the surface. Weextract from these measurements the in-plane response function of the tracers. Using this techniquenear the interface and with optical access from only one side presents a new set of issues. Specifically,the position of the tracer particle is determined by back scattered light from the tracer and the nearbyinterface generates a significant amount of additional back scattering. We explain in detail how onecan account for these effects.

Our principal conclusions are that we can account for the optical complexities of working witha laser trap near the interface in order to recover the Stokes result at depth. We observe the expectedchanges in the tracer’s mobility consistent with hydrodynamic calculations, detailed in Paper I ofthis series.35 This validates the calculated response function to be used in future work along theselines, and clarifies the role of surface tension and particle size. And finally, we show that the abilityto measure the dependence of the tracer particle’s mobility as a function of depth is essential togive sufficient sensitivity to make reproducible rheological measurements of complex viscoelasticmonolayers in spite of the weakened hydrodynamic coupling between the tracer and that monolayer.In this article, “depth” refers to the absolute value of the distance from the center of the particle tothe air/water interface.

The remainder of this article is organized as follows. In Sec. II we describe the experimentalsetup including details of the laser trap and quadrant photodiode detection scheme used to recordthe positional fluctuations of the tracer at a fixed depth. The dielectric mismatch at the air/waterinterface leads to reflections that can impact the detection methods, but do not modify the trappingstrength with depth. These systematic effects and a method of compensating for them are describedin Appendix A. In Sec. III we discuss our experimental data. We compare the theoretical resultsto the data and discuss their broader significance for microrheology of Langmuir monolayers inSec. IV.

II. MATERIALS AND METHODS

A. Apparatus

The apparatus involves two main systems: the optical components that generate the trap andmeasure the particle fluctuations and a small cell for holding the fluid samples. With both of theseelements, we take advantage of an existing Langmuir monolayer trough37 coupled with a 100×water immersion objective (NA 1.0, Olympus America Inc.) from below the trough. The opticaltrap uses a Nd:YVO4 laser (Spectra Physics BL-106C, 1064 nm). The laser light passes throughan optical system consisting of a beam expander, steering lenses and mirrors, leading to the waterimmersion objective, which focuses the beam to form an optical trap. The trough is attached to avertical translation stage so that the fixed objective can trap particles at various distances belowthe surface. Trapped particles scatter laser light back through the objective, onto both a quadrantphotodiode (New Focus 2903) and an intensified CCD camera. The quadrant photodiode (QPD)allows high frequency (66 kHz) 2-dimensional position measurements to be recorded via a dataacquisition board and custom Labview software. A schematic of the apparatus is shown in Fig. 1.

As originally designed, the objective extends through a well in the bottom of the Langmuirtrough. The optical trap would be focused near the air/water interface and used to trap particles in ornear a Langmuir monolayer. However, for the experiments reported here, two smaller sample cellswere used to eliminate flows in the subphase and allow for ease of switching between free and rigidboundary conditions for the interface. The Langmuir trough was used as a sample stage to supportthe cells, and allowed for immersion of the objective in water so that it functioned properly. Forexperiments with a free boundary, the cell consists of a thin block of polytetrafluoroethylene (PTFE)with a circular hole, 19 mm in diameter. A glass cover slip is attached to the bottom of the teflonchamber with double sided tape to form an open, thin cylindrical cell about 2 mm in depth. Thesolution fills this volume to a height of approximately 1 mm, which is shallow enough to allow theobjective access to the air/water interface. A cover is used to reduce air currents across the interface.

For studies of a rigid interface, we measure the fluctuations of trapped particles in a ∼50 μLchamber made from two layered strips of double sided tape sandwiched between a glass slide anda glass coverslip.38 The chamber is approximately 190 μm thick and is filled with the same diluted

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

128.200.29.82 On: Thu, 17 Sep 2015 01:30:10

Page 5: Probing interfacial dynamics and mechanics using …dennin/preprints/BDSE14-submerged...Probing interfacial dynamics and mechanics using submerged particle microrheology. II. Experiment

071904-4 Boatwright et al. Phys. Fluids 26, 071904 (2014)

Mercury lamp

1064 nm laser

beam-spli�er

mirror

dichroic mirror

lenses beam-spli�er

CCD camera

QPD

water-immersionobjec�ve

PTFE open troughglass cover slip

trapped bead

trapped bead

air-water interface

incident lasersca�ered laser

d

FIG. 1. Schematic of the position detection and imaging system. The inset shows a particle (with radius R) trapped in adistance d below the air/water interface. This distance is measured from the center of the particle to the interface.

particle solution as used in the air/water interface experiments. It is rested on the Langmuir trough,just like the air/water interface cell, with the coverslip side of the chamber facing downward towardthe objective. The trapped particle’s depth is defined as the distance below the glass slide, which isthe glass/water interface furthest from the objective. A similar chamber is used to obtain the datashown in Fig. 8, but with polycarbonate in place of a glass slide.

The cells are cleaned with water and ethanol prior to each experiment. For the open cell, beforea sample is deposited, the cell is again filled with water, then aspirated to remove dust and otherparticles that may have entered the cell while waiting for the ethanol to completely evaporate.Carboxylate modified, red fluorescent particles (Invitrogen) are diluted in ultrapure water by a factorof 104 and 0.3 mL of this solution is placed in the circular cell. Two sizes of particles are trapped(0.5 and 5 μm radii as specified by the manufacturer), but the results from 0.5 μm particles areonly used to highlight experimental difficulties with small particles and are not used in our finalresults. Only Figures 2, 7, and 8 contain results from these smaller particles. A mercury lamp excitesfluorescent tags on the particles, allowing them to be imaged by the CCD camera after passingthrough appropriate filters. With the particles in view, the trapping laser is turned on and a singleparticle is trapped. Prior to measuring the particle fluctuations with the QPD, a mirror is used tocenter the scattered light from the particle on the QPD’s chip while the particle is 200 μm beneaththe air/water interface.

A particle’s position relative to the focus can be determined by comparing the images of trappedparticles.39 For example, if a 0.5 μm particle is below the focus, the center of the particle appearsdark; if above, it appears light. When a particle is trapped, it is located very close to the focus,has clear boundaries and no visible interference fringes. These phenomena can be used to find thesurface in the following way. When a particle is trapped beneath the surface, a translation stage isused to lower the cell around the fixed objective lens, moving the optical trap carrying the particlecloser to the water’s surface. Video of the particle shows that it maintains its position relative to thefocus throughout the vertical translation until the particle is pushed below the focus by the interface.At this point the particle appears with a black center and surrounded by interference fringes. Thelocation of the surface can be confirmed by finding where the total intensity of scattered laserlight is at a maximum. Once the location of the interface is determined, the objective is translateddownward with an uncertainty in depth of approximately 5 μm in the air/water interface experiments.

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

128.200.29.82 On: Thu, 17 Sep 2015 01:30:10

Page 6: Probing interfacial dynamics and mechanics using …dennin/preprints/BDSE14-submerged...Probing interfacial dynamics and mechanics using submerged particle microrheology. II. Experiment

071904-5 Boatwright et al. Phys. Fluids 26, 071904 (2014)

1 10 100 10001E-5

1E-4

1E-3

0.01

0.1

1

PS

D (

arb.

uni

ts)

f (Hz)

FIG. 2. Power spectral densities (PSDs) of 0.5 μm (thick dotted line/green online) and 5 μm (thin dotted line/orange online)radius particles trapped in bulk solution. A Lorentzian fit to the spectrum for the 0.5 μm radius particle—see Eq. (1)—isshown in the solid line/red online. This fit shows that the data are consistent with a simple model of overdamped motion ina linearly elastic trap. On the other hand, the PSD of the 5 μm radius particle poorly fits a Lorentzian. The larger particlerequires higher laser power (100 mW instead of 20 mW) for trapping and this introduces more low frequency noise. Inaddition, the high-frequency Brownian motion of larger particles is of a smaller amplitude than for the smaller ones andconsequently harder to detect. For this reason, the observed PSD (above) for that particle beyond ∼2000 Hz decays moresharply than the expected 1/f2 (dashed line/black online).

Between each fluctuation measurement at the air/water interface, the surface position is reacquiredbecause of evaporation. The rate of evaporation was found to be on the order of 0.4 μm/min, whichwould produce a significant error over the series of depth measurements if the surface were notreacquired.

B. Measurement details

We briefly outline the experimental procedure. Using the method discussed above and inRef. 37, we obtain tracer position data at a rate of 66 kHz. These data for the lateral (i.e., inthe plane whose perpendicular is the optical axis) displacements of the tracer particle form a timeseries xt of two-dimensional vectors. We then compute the fast Fourier transform and obtain thepower spectrum of these position fluctuations 〈|xf|2〉 where f is the frequency variable conjugateto time t. Discounting the role of reactive stresses associated with surface deformation, the powerspectrum takes the form

〈|x f |2〉 = D

2π2( f 2c + f 2)

, (1)

where D = kBTμ is the diffusion constant of the particle with mobility μ and fc = ωc/2π = kμis the corner frequency, which arises from the stiffness k of the (harmonic) optical trap (seeFig. 2). All reported results are derived from measured time series with a length of 2 s. The spectraof the fluctuations are blocked and averaged across many of such time series. The fluctuation data of5.0 μm particles submerged below the air/water interface represent ∼40 time series measurements.For the case of 0.5 μm particles at the glass and polycarbonate interfaces we averaged 10 and 5 timeseries, respectively. Low frequency data showing drift were omitted from further analysis. We show

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

128.200.29.82 On: Thu, 17 Sep 2015 01:30:10

Page 7: Probing interfacial dynamics and mechanics using …dennin/preprints/BDSE14-submerged...Probing interfacial dynamics and mechanics using submerged particle microrheology. II. Experiment

071904-6 Boatwright et al. Phys. Fluids 26, 071904 (2014)

power spectral densities (PSDs) over a frequency range of 1 Hz to 2000 Hz. The lower limit is setby fluctuations in laser power on longer times scales while the upper limit is set by the ability toresolve small amplitude Brownian fluctuations of the relatively large sphere.

From the power spectrum we can directly compute the imaginary part of the response function,which we report in the data presented. This response function describes the linear relationshipbetween the position of the tracer and the force applied to it,

x f = χ ( f )Ff , (2)

in a frequency-resolved (i.e., f-dependent) manner. It is generally applicable to viscoelastic systems.Writing this complex function in its real and imaginary parts we define

χ ( f ) = χ ′( f ) + iχ ′′( f ). (3)

The fluctuation-dissipation theorem11 immediately relates the imaginary part of the response functionto the PSD of the fluctuations via

χ ′′( f ) = π f 〈|x f |2〉kBT

. (4)

The more common hydrodynamic quantity is the mobility of the particle, describing the linearrelationship between the particle’s velocity and the force applied to it. Of course, this quantity andthe position response function defined above are directly proportional in the frequency domain. Theadvantage of discussing χ is that it relates thermodynamically conjugate variables and thus entersthe expression in Eq. (4). Framing our results in this form also allows a more direct connectionto other microrheological analyses. A more complete discussion of the response function for thisexperimental system can be found in the first of these two papers.35

The real part of the response function can then be found from a Kramers-Kronig integral.40

In practice, it is necessary to logarithmically block and bin the experimentally obtained PSD forsmoothing purposes and the remaining integrals are performed using a discrete sine and cosinetransform.41

While it is true that the surface tension of the air/water interface introduces a reactive part tothe particle’s response function due to the elastic stresses associated with normal displacementsof that interface, our calculations suggest that these effects are vanishingly small for high surfacetension interfaces such as that of air and water.35 Our experiments confirm that, within uncertainty,the real part of the response function is zero. Thus, in the two cases of current experimental interest(the air/water and water/glass interfaces), the particle’s response function can be considered to bethe combination of a purely dissipative and depth-dependent part due to the hydrodynamics and asimple elastic part due to the laser trap.

Hereafter we discuss our results in terms of the imaginary part of the response function definedin Eqs. (2) and (3). It is convenient when discussing these results to nondimensionalize them by theresult expected for a spherical particle of radius a in bulk water of viscosity η:

χStokes = i

6πηa(2π f ). (5)

For all experiments reported here, we expect to observe that the imaginary part of the responsefunction nondimensionalized in this way approaches unity as the depth of the particle increases.

To connect the QPD signals and the particle fluctuations, there are two main corrections dueto two related optical effects: changes in the background intensity of reflected light as the depthis changed and particle lensing effects acting on that reflected light. These are discussed in theAppendices. There we show that the systematic errors are minimized for 5 μm size particles,justifying the focus on these particles in the data section.

III. RESULTS

We plot in Fig. 3 the power spectra of the position fluctuations of 5 μm radius spheres atthe air/water interface, geometrically averaging five sets of measurements at constant depth, each

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

128.200.29.82 On: Thu, 17 Sep 2015 01:30:10

Page 8: Probing interfacial dynamics and mechanics using …dennin/preprints/BDSE14-submerged...Probing interfacial dynamics and mechanics using submerged particle microrheology. II. Experiment

071904-7 Boatwright et al. Phys. Fluids 26, 071904 (2014)

1 10 100 10001E-6

1E-5

1E-4

1E-3

0.01

0.1

1

PSD

(arb

. uni

ts)

frequency (Hz)

FIG. 3. The power spectra of 5 μm radius particles in water at depths ranging from 10 − 205 μm beneath the air/waterinterface. The PSDs with highest magnitudes (thin dots/black online and thick dots/red online), and therefore the highestmobility, correspond to particles nearest to the surface at a depth of 2 particle radii. A black arrow denotes these two spectra.As the particle’s depth is increased, the magnitudes of the power spectra decrease monotonically converging to the curveconsistent with a particle trapped in bulk water (see cluster of solid lines/multicolored online). The dashed black line indicatesa slope of −2 expected for the high frequency region of the PSD.

with a two second duration. In an ideal system, the power spectrum would be Lorentzian, but lowfrequency noise from the laser increases the magnitude of the power spectrum at low frequencies.Additionally, the corner frequency for the 5 μm radius particles is close to the laser noise frequencyband, smearing out what would be a plateau in the power spectrum. Despite the low frequencynoise, a clear trend is visible in the spectra: the spectral magnitude for the particles at the surfaceis significantly higher than the other particles. The arrow in the figure indicates two power spectrataken for particles positioned about two particle radii away from the surface. As the particles moveaway from the surface, the power spectrum decreases, converging to a depth-independent curve verysimilar to that measured in the bulk fluid. The increase in the power spectrum near the surface,reflecting an increase in the position fluctuations of the particle, suggests a greater particle mobility.A more quantitative analysis of the particle mobility follows. We exclude data corresponding toparticles at depths of ∼1 particle radius or less. At such short distances to the air/water interface, thepower spectrum becomes somewhat unreproducible. We attribute this to the fact that the air/waterinterface is, in fact, moving slowly relative to the optically trapped particle due to evaporation andperhaps minor leakage of the cell. Over the course of the experiment, water level drops a distancecomparable to a particle radius. We estimate the uncertainty in the depth as ±5 μm as a result ofthese effects. Experiments done inside a closed chamber have a much smaller uncertainty in depth,about ±2 μm, since there is no detectable change in height of the surface.

Applying the methods described in Sec. II B, we first test the measurement of particles near a rigidinterface. We used 5 μm radius particles and calculate their mobility from sets of 40 measurementsat each depth, using data obtained on 5 different days. We plot the Stokes-normalized imaginaryresponse function (which is directly related to the mobility) as a function of depth from the glass/waterinterface in Fig. 4. At depths of approximately 7 particle radii, the observed mobility is experimentallyindistinguishable from that of bulk water.

Our main results are presented in Fig. 5. Here we directly compare the data from Fig. 4 for arigid boundary (open triangles) with the data for a free surface (solid blue circles). As predicted,

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

128.200.29.82 On: Thu, 17 Sep 2015 01:30:10

Page 9: Probing interfacial dynamics and mechanics using …dennin/preprints/BDSE14-submerged...Probing interfacial dynamics and mechanics using submerged particle microrheology. II. Experiment

071904-8 Boatwright et al. Phys. Fluids 26, 071904 (2014)

0 5 10 15 20

0.1

1

χ"/|χ

Sto

kes|

depth (d/a)

FIG. 4. The measured imaginary response function of a particle in water as a function of distance from a glass/water interface.Theoretical values at a rigid wall are represented by a solid line (red online). Five sets of data were averaged together at eachdistance. In this experiment, the uncertainty in the depth is about ±0.4 radii.

the free surface exhibits an enhanced response function, corresponding to an enhanced mobility. Wefind that effect of the free surface (in agreement with theory) leads to a maximum increase of theparticles’ mobility by about 30% at a depth of about two particle radii. The mobility monotonicallydecreases to the bulk value as a function of distance from the surface. In addition, though currently

0 5 10 15 200

1

2

χ"/|χ

Sto

kes|

depth (d/a)

FIG. 5. The imaginary response function of a 5 μm radius particle in water as a function of distance from an air/waterinterface (solid circles/blue online) and a glass/water interface (open triangles/black online). The glass/water interface datafollow the expected curve derived by Faxen (dashed line/red online). Our prediction of the imaginary response function (solidline/red online) fits the air/water interface data well. See paper I35 for theoretical details.

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

128.200.29.82 On: Thu, 17 Sep 2015 01:30:10

Page 10: Probing interfacial dynamics and mechanics using …dennin/preprints/BDSE14-submerged...Probing interfacial dynamics and mechanics using submerged particle microrheology. II. Experiment

071904-9 Boatwright et al. Phys. Fluids 26, 071904 (2014)

at the limit of our error bars, the two sets of data are measurably different, especially when oneconsiders the entire curve and not just a single data point. This emphasizes the trade-off betweencoupling with the interface (decreasing measurement sensitivity) and measuring the behavior asa function of depth (increasing measurement sensitivity). Finally, we also plot the hydrodynamictheory for the free interface theory (solid curve/red online) presented in paper I35 and Faxen’s theory(dashed curve/red online).

The agreement of both measurements with their corresponding theories demonstrates the fea-sibility of probing the rheology of an interface without contacting it. The effect of a deformableand viscoelastic interface on the mobility lies between the two extremes of a free air/water surfaceand rigid wall. The experiment can probe the particle mobilities near viscoelastic interfaces in afrequency-resolved manner with sufficient precision to allow for tractable non-contact surface orinterfacial rheometry for a variety of more complex and viscoelastic monolayers.

Finally, we have tested this technique with a standard set of monolayers, with the full resultsreported in Ref. 42. For the purposes of this paper, we provide the details of how one determinesthe dynamic range of the instrument for a given set of parameters. The key step in determiningtwo-dimensional surface viscosities (η2D) is the measurement of the imaginary part of χ as afunction of frequency. This is compared to the values of χ computed from theory as a functionof surface viscosity, and by matching theory and experiment, the monolayer surface viscosity isextracted. The dynamic range is determined by the theoretical values for χ for a free surface anda rigid boundary, as discussed in detail in this paper. These provide an upper and lower bound,respectively, as shown in Fig. 6. For comparison, Fig. 6 also plots the value of χ for a monolayer ofDPPC (dipalmitoylphosphatidylcholine) at two relatively low values of pressure, � = 1 mN/m and� = 2 mN/m. One can see that for � = 1 mN/m, the behavior of χ is essentially that of a free surface;however, careful comparison with theory gives a surface viscosity of η2D = 1.4 × 10−9 Ns/m forthis monolayer. For the case of � = 2 mN/m, one finds η2D = 7.2 × 10−8 Ns/m.

1000 10000106

107

108

ω χ

(Pa

s m)-1

frequency (Hz)

FIG. 6. The results for a free surface (dotted line/light blue online) and a rigid surface (dashed line/red online) bound thedynamic range of the instrument. These bounds depend the ratio of the depth of the particle below the surface and its radiusd/a; we show data here for d/a = 2. Typical data are shown for a DPPC surfactant monolayer at area pressures of � = 1 mN/m(thin solid line/black online) and � = 2 mN/m (thick solid line/green online). Surface viscosities for these monolayers aremeasured to be η2D = 1.4 × 10−9 Ns/m and η2D = 7.2 × 10−8 Ns/m, respectively.

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

128.200.29.82 On: Thu, 17 Sep 2015 01:30:10

Page 11: Probing interfacial dynamics and mechanics using …dennin/preprints/BDSE14-submerged...Probing interfacial dynamics and mechanics using submerged particle microrheology. II. Experiment

071904-10 Boatwright et al. Phys. Fluids 26, 071904 (2014)

IV. SUMMARY

The effect of a nearby rigid wall on the drag coefficient of a particle is a well-understoodproblem in low Reynolds number hydrodynamics, but one of continuing interests with regard tomicrofluidics.43–45 The general observation regarding this is that the presence of a wall reduces themobility of the particles near to it. Here we find that the free surface enhances that mobility, in linewith expectations from micro hydrodynamic theory.46 Because the distinction between the effect ofa rigid wall and a zero shear stress wall is so dramatic, there is a significant range available to observethe effects of complex surface rheology as a function of frequency. Additionally, we illustrate thedetermination of the surface viscosity of a simple monolayer (DPPC at low surface pressure) andthe determination of the dynamic range of the apparatus. Full details of its application to a numberof standard monolayers is given in Ref. 42.

We find the depth dependent changes in the imaginary response function, and thus the mobility,of a submerged 5 μm radius particle. Comparing the experimental data shown in Fig. 5 with thetheoretical prediction35 for the free interface of water, we see that the increase in the imaginaryresponse near the surface is well explained. This increase is due to the fact that the motion near thefree surface leads to less viscous dissipation. The agreement confirms the correct identification inRef. 35 of the role of particle size and surface tension. The calculation establishes that particle sizeis a higher order effect, allowing the use of relative large 5 μm particles. Surface tension introducesa reactive part of the response function near the free surface since the elastic deformation of theinterface does recoverable work. As shown in Paper I,35 however, we expect that the effect of surfacetension is vanishing small for high surface energy interfaces because the large surface tension forcesthe surface to remain nearly flat. The relevant measure of effect of surface tension τ on the responsefunction of a tracer of radius a is controlled by the capillary number Ca = ηa(2π f)/τ . For theair/water interface at the frequencies of experimental interest, Ca � 1 so the interface remains flatand surface tension should produce no measurable change in the response function. Indeed, we donot observe a surface tension effect in our data.

It is essential to understand the physics and confirm the hydrodynamic calculations in the twosimpler limits presented here: a perfectly rigid surface and a free and deformable surface. Havingdone so, we provide proof of principle that one can extract the complex frequency dependent rheologyof a viscoelastic interface, such as a Langmuir monolayer, from observing the in-plane fluctuationsof tracers submerged a short distance below it. The out-of-plane fluctuations provide an excellentmeasure of the bending mechanics of the interface as well. The non-contact microrheological probeis highly desirable since the presence of the submerged probe particle should in no way perturbthe monolayer eliminating the need for more difficult two-particle microrheology.47, 48 Moreover,the (purely hydrodynamic) coupling of the probe to the monolayer is well understood and nowverified eliminating the need to understand the complex physics of the three-phase contact linein order to interpret the fluctuation data rheologically. Using other methods, the interaction ofthe probe and monolayer domains may govern the measured rheological response, rather than therheology of the monolayer. Kurnaz and Schwartz49 illustrate this by showing that, depending on themeasurement technique, shear thickening or shear thinning may be observed in the same system.In addition, it appears that for the case of highly elastic monolayers, probe particles are typicallyexpelled from the monolayer presumably due to the elastic stress they produce at the interface.We believe that this effect accounts for the large and consistent discrepancies between fluctuation-based and active mechanical measurements of such monolayer systems. In a following article wewill address the cases of principal interest in which the interface that is not completely rigid, butrather has a finite and frequency-dependent complex shear modulus. Furthermore, we will use anactive microrheological technique in a viscoelastic monolayer to more precisely probe the responsefunction.

ACKNOWLEDGMENTS

We acknowledge the support of NSF-DMR-1309402. M.D. acknowledges the Research Corpo-ration for support.

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

128.200.29.82 On: Thu, 17 Sep 2015 01:30:10

Page 12: Probing interfacial dynamics and mechanics using …dennin/preprints/BDSE14-submerged...Probing interfacial dynamics and mechanics using submerged particle microrheology. II. Experiment

071904-11 Boatwright et al. Phys. Fluids 26, 071904 (2014)

APPENDIX A: CORRECTION FOR SURFACE REFLECTIONS

The four quadrant photodiode detection of tracer position relies on variations of light intensityreflected back through the optical system. Unfortunately, in the submerged particle problem thereare reflections both from the particle, which measure its displacement, and from the interface, whichmake spurious contributions to that signal. One needs to account for those spurious signals in a newway in order to make the fluctuation measurements on the submerged particles. Reflected laser lightis collected by the objective and reflected onto the QPD, which has three outputs: X, Y, and SUM.The X and Y outputs are difference measurements between halves of the QPD used to establishthe particle position, and the SUM output is a measure of the total intensity on the photodiode. Acommon issue with a laser/QPD position measurement is drift in the overall laser intensity, whichcan introduce systematic errors in the X and Y output. There is a standard correction which is todivide the X and Y outputs by the SUM. As a result of this normalization, changes in laser powerin a standard particle trapping experiment do not significantly change the magnitude of a powerspectrum generated by a times series of X or Y values.38

In our experiments, due to the index of refraction mismatch at the air/water interface, there arealso strong surface reflections in addition to the reflected light from the tracer. The diffuse componentof the surface reflection produces a significant signal in the SUM channel of the QPD at ≈80 μmfrom the surface. As shown in Fig. 7, the signal is present with and without particles in the laser trapand has a strong dependence on distance from the interface. Given an incident laser power of A weexpect that the contribution to the SUM signal from surface reflections should vary with distanced as A

(d/B+1)2 + C . Here B is the focal distance and C corrects for the background illumination. InFig. 7, we confirm this where we show that both the SUM signal with (dotted) and without (dashed)a particle fits to this function (red online). Because this additional signal is independent of the lightreflected from the particle, it increases the value of the SUM channel but does not impact the X andY channel signals. Therefore, if we were to follow the standard normalization procedures, we wouldintroduce new systematic errors due to the surface reflection enhancement of the SUM signal.

To avoid these systematic errors, we normalize using the average SUM value from the bulkdepths (taking “bulk” to refer to measurement depths of 120 to 200 μm), rather than the SUMsignal measured concurrently with X and Y. Without this procedure, the data show an unphysicalminimum with depth, as shown by the open symbols in Fig. 8. However, when when one usesour new procedure, one obtains the results shown by the closed symbols in Fig. 8; the unphysical

0 50 100 150 2000.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

5.0

SU

M (

arb.

uni

ts)

FIG. 7. The SUM signal, a measure of the total intensity on the quadrant photodiode (QPD), is plotted as a function of depth.The following are represented on the plot: a trapped 0.5 μm radius particle (solid triangles/black online), a trapped 5 μmradius particle (solid circles/blue online), and the result without a particle (open triangles/black online). In the data with aparticle trapped, a single particle is trapped throughout the measurement. The lines (red online) running through the datapoints are fits to the data of the form A

(d/B+1)2 + C .

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

128.200.29.82 On: Thu, 17 Sep 2015 01:30:10

Page 13: Probing interfacial dynamics and mechanics using …dennin/preprints/BDSE14-submerged...Probing interfacial dynamics and mechanics using submerged particle microrheology. II. Experiment

071904-12 Boatwright et al. Phys. Fluids 26, 071904 (2014)

0 50 100 150 200 250

1

10

χ"/|χ

Sto

kes|

depth (d/a)

FIG. 8. The imaginary response function of 0.5 μm radius particles naıvely near a wall with stick boundary conditionscalculated with normalization by the bulk SUM signal (solid symbols) and with normalization by the SUM signal at eachindividual particle depth (open symbols). To better understand the role of surface reflections, we studied data from particlesnear walls made of two optically different materials: polycarbonate (n ≈ 1.592) and sodalime glass (n ≈ 1.520), representedby square (green online) and triangle (blue online) symbols, respectively. The dashed (red online) line is the expected resultdue to Faxen’s theory. Using the bulk SUM signal for normalization, the unphysical minimum near d ≈ 30a disappears.Despite this correction, the reflected light has an additional impact due to particle lensing that systematically increases thefluctuations near the surface. Since the index of refraction of polycarbonate differs from water more than does glass, theformer materials generate more intense reflections, causing a larger variation in the position signal and resulting in a largerapparent mobility. The particle lensing effect is remedied by using sufficiently large tracer particles.

minimum has been eliminated. These corrected data, however, still deviate significantly from simplehydrodynamic theories particularly at small depths. The residual error is due to a particle lensingeffect discussed in Appendix B.

APPENDIX B: CORRECTION FOR PARTICLE LENSING

In addition to the diffusive component of surface reflections, we hypothesize that, very nearthe surface, there is a second order reflection component that enhances the variance of the X andY signals. The surface reflected light diffracts back around the trapped particle (or refracts throughit) and passes back to the objective. This particle lensing forms a time varying image on the QPD,causing an additional signal in the X and Y channels. This stronger signal variance in the positionchannels would naıvely be interpreted as a greater mobility of the particle. This is apparent inFig. 8, where the measured values near the fixed walls disagree with the expected Faxen’s result.

In order to confirm that the systematic effects are in fact due to surface reflections, we usedtwo materials with differing indices of refraction for the fixed wall. In this way, we keep the same“stick” boundary conditions for the hydrodynamics, but we change the amount of reflected light. InFig. 8 we plot the apparent imaginary part of response function of a half micron radius sphere inwater near either a glass wall (blue symbols) or a polycarbonate wall (green symbols). It is clearfrom these data that the spurious increase in particle mobility depends on the material making up thewall, and that the more reflective polycarbonate boundary (n ≈ 1.592) generates a larger error thanthat of the less reflective glass (n ≈ 1.520). The dependence of the effect on the wall’s refractiveindex demonstrates that it cannot be hydrodynamic in origin, but instead is due to the ≈55% greaterreflectance of the polycarbonate wall over that of the glass one.

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

128.200.29.82 On: Thu, 17 Sep 2015 01:30:10

Page 14: Probing interfacial dynamics and mechanics using …dennin/preprints/BDSE14-submerged...Probing interfacial dynamics and mechanics using submerged particle microrheology. II. Experiment

071904-13 Boatwright et al. Phys. Fluids 26, 071904 (2014)

0 5 10 15 20

0.1

1

χ"/|χ

Sto

kes|

depth (d/a)

FIG. 9. The imaginary response function from Figure 4 showing how results normalized by the bulk SUM signal (solidcircles/blue online) match theory better than the results normalized by the SUM signal at each depth (open squares/blackonline). The latter values of the response function are shown to be significantly below the theoretical curve at shallow depths.Values normalized by the bulk SUM signal closely follow the theoretical curve.

A further test of this idea is to use larger spheres to modify the lensing effect. This is illustrated inFig. 9, where data for 5 μm particles are presented. In contrast to the half micron sphere data shownin Fig. 8, we find that correctly normalizing the data by the bulk SUM signal provides excellentagreement with theory, effectively eliminating the particle lensing effect. This is presumably dueto the fact that the 5 μm sphere is larger than the wavelength of the laser light and thus producesless diffraction effects. The trade-off for using a larger size particle is that the amplitude of theBrownian motion becomes smaller. For the 5 μm radius tracers, we cannot detect their Brownianmotion at frequencies above 2 kHz. The power spectra in Fig. 3 show f−2 behavior between 200 −2000 Hz as expected for a viscous fluid, but beyond this frequency, the power spectrum shows asteeper slope, which we attribute to having reached our detection floor. As both of the systematicerrors inherent in technique stem from our use of reflected light for tracer position detection, weexpect that an alternative geometry using QPD to detect the transmitted light may reduce or eliminateboth problems. This involves a redesign of our basic instrument, but will be implemented in futureexperiments.

1 H. Mohwald, “Phospholipid and phospholipid-protein monolayers at the air/water interface,” Annu. Rev. Phys. Chem.41(1), 441–476 (1990).

2 V. M. Kaganer, H. Mohwald, and P. Dutta, “Structure and phase transitions in Langmuir monolayers,” Rev. Mod. Phys.71(3), 779–819 (1999).

3 J. S. Olafsen, Experimental and Computational Techniques in Soft Condensed Matter Physics (Cambridge UniversityPress, 2010).

4 T. G. Mason and D. A. Weitz, “Optical measurements of frequency-dependent linear viscoelastic moduli of complexfluids,” Phys. Rev. Lett. 74(7), 1250–1253 (1995).

5 T. G. Mason, K. Ganesan, J. H. van Zanten, D. Wirtz, and S. C. Kuo, “Particle tracking microrheology of complex fluids,”Phys. Rev. Lett. 79(17), 3282–3285 (1997).

6 F. C. MacKintosh and C. F. Schmidt, “Microrheology,” Curr. Opin. Colloid Interface Sci. 4(4), 300–307 (1999).7 T. A. Waigh, “Microrheology of complex fluids,” Rep. Prog. Phys. 68(3), 685–742 (2005).8 T. M. Squires and J. F. Brady, “A simple paradigm for active and nonlinear microrheology,” Phys. Fluids 17(7), 073101

(2005).9 P. Cicuta and A. M. Donald, “Microrheology: A review of the method and applications,” Soft Matter 3(12), 1449–1455

(2007).

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

128.200.29.82 On: Thu, 17 Sep 2015 01:30:10

Page 15: Probing interfacial dynamics and mechanics using …dennin/preprints/BDSE14-submerged...Probing interfacial dynamics and mechanics using submerged particle microrheology. II. Experiment

071904-14 Boatwright et al. Phys. Fluids 26, 071904 (2014)

10 T. M. Squires and T. G. Mason, “Fluid mechanics of microrheology,” Annu. Rev. Fluid Mech. 42(1), 413–438 (2010).11 P. Chaikin and T. Lubensky, Principles of Condensed Matter Physics (Cambridge University Press, 1995).12 D. Mizuno, D. A. Head, F. C. MacKintosh, and C. F. Schmidt, “Active and passive microrheology in equilibrium and

nonequilibrium systems,” Macromolecules 41(19), 7194–7202 (2008).13 S. Yamada, “Mechanics of living cells measured by laser tracking microrheology,” Biophys. J. 78(4), 1736–1747

(2000).14 J. C. Crocker and B. D. Hoffman, Multiple-Particle Tracking and Two-Point Microrheology in Cells (Elsevier, Waltham,

Massachusetts, 2007), Vol. 83, pp. 141–178.15 D. Wirtz, “Particle-tracking microrheology of living cells: Principles and applications,” Annu. Rev. Biophys. 38(1), 301–

326 (2009).16 S. Q. Choi, S. Steltenkamp, J. A. Zasadzinski, and T. M. Squires, “Active microrheology and simultaneous visualization

of sheared phospholipid monolayers,” Nat. Commun. 2, 312 (2011).17 K. Kim, S. Q. Choi, J. A. Zasadzinski, and T. M. Squires, “Interfacial microrheology of DPPC monolayers at the air-water

interface,” Soft Matter 7, 7782–7789 (2011).18 Y. Almeras, J. L. Barrat, and L. Bocquet, “Influence of wetting properties on diffusion in a confined fluid,” preprint

arXiv:cond-mat/9912067 (1999).19 L. Joly, C. Ybert, and L. Bocquet, “Probing the nanohydrodynamics at liquid-solid interfaces using thermal motion,” Phys.

Rev. Lett. 96(4), 046101 (2006).20 P. G. Saffman and M. Delbruck, “Brownian motion in biological membranes,” Proc. Natl. Acad. Sci. 72(8), 3111–3113

(1975).21 K. D. Danov, R. Dimova, and B. Pouligny, “Viscous drag of a solid sphere straddling a spherical or flat surface,” Phys.

Fluids 12(11), 2711–2722 (2000).22 M. H. Lee, D. H. Reich, K. J. Stebe, and R. L. Leheny, “Combined passive and active microrheology study of protein-layer

formation at an air-water interface,” Langmuir 26(4), 2650–2658 (2009).23 H. M. Hilles, H. Ritacco, F. Monroy, F. Ortega, and R. G. Rubio, “Temperature and concentration effects on the equilibrium

and dynamic behavior of a Langmuir monolayer: From fluid to gel-like behavior,” Langmuir 25(19), 11 528–11 532 (2009).24 V. Prasad, S. A. Koehler, and E. R. Weeks, “Two-particle microrheology of quasi-2D viscous systems,” Phys. Rev. Lett.

97(17), 176001 (2006).25 V. Prasad and E. R. Weeks, “Two-dimensional to three-dimensional transition in soap films demonstrated by microrheol-

ogy,” Phys. Rev. Lett. 102(17), 178302 (2009).26 V. Prasad and E. R. Weeks, “Flow fields in soap films: Relating viscosity and film thickness,” Phys. Rev. E 80(2), 026309

(2009).27 M. Sickert and F. Rondelez, “Shear viscosity of Langmuir monolayers in the low-density limit,” Phys. Rev. Lett. 90(12),

126104 (2003).28 M. Sickert, F. Rondelez, and H. A. Stone, “Single-particle Brownian dynamics for characterizing the rheology of fluid

Langmuir monolayers,” Europhys. Lett. 79(6), 66005 (2007).29 F. Ortega, H. Ritacco, and R. G. Rubio, “Interfacial microrheology: Particle tracking and related techniques,” Curr. Opin.

Colloid Interface Sci. 15(4), 237–245 (2010).30 D. K. Schwartz, C. M. Knobler, and R. Bruinsma, “Direct observation of Langmuir monolayer flow through a channel,”

Phys. Rev. Lett. 73, 2841–2844 (1994).31 R. S. Ghaskadvi, J. B. Ketterson, and P. Dutta, “Nonlinear shear response and anomalous pressure dependence of viscosity

in a Langmuir monolayer,” Langmuir 13(19), 5137–5140 (1997).32 R. S. Ghaskadvi, S. Carr, and M. Dennin, “Effect of subphase ca++ ions on the viscoelastic properties of Langmuir

monolayers,” J. Chem. Phys. 111(8), 3675–3678 (1999).33 C. F. Brooks, G. G. Fuller, C. W. Frank, and C. R. Robertson, “An interfacial stress rheometer to study rheological

transitions in monolayers at the air/water interface,” Langmuir 15(7), 2450–2459 (1999).34 S. Wurlitzer, C. Lautz, M. Liley, C. Duschl, and T. Fischer, “Micromanipulation of Langmuir monolayers with optical

tweezers,” J. Phys. Chem. B 105(1), 182–187 (2001).35 R. Shlomovitz, A. A. Evans, T. Boatwright, M. Dennin, and A. J. Levine, “Probing interfacial dynamics and mechanics

using submerged particle microrheology. I. Theory,” Phys. Fluids 26, 071903 (2014).36 F. Gittes and C. Schmidt, Chapter 8 Signals and Noise in Micromechanical Measurements, Methods in Cell Biology Vol.

55 (Elsevier, Waltham, Massachusetts, 1997), pp. 129–156.37 R. Walder, C. F. Schmidt, and M. Dennin, “Combined macro- and microrheometer for use with Langmuir monolayers,”

Rev. Sci. Instrum. 79(6), 063905 (2008).38 M. Atakhorrami, J. I. Sulkowska, K. M. Addas, G. H. Koenderink, J. X. Tang, A. J. Levine, F. C. MacKintosh, and

C. F. Schmidt, “Correlated fluctuations of microparticles in viscoelastic solutions: Quantitative measurement of materialproperties by microrheology in the presence of optical traps,” Phys. Rev. E 73(6), 061501 (2006).

39 J. Curtis, “Dynamic holographic optical tweezers,” Opt. Commun. 207(1–6), 169–175 (2002).40 L. D. Landau and E. M. Lifshitz, Statistical Physics (Part 1), 3rd ed. (Pergamon Press, Oxford, 1980).41 B. Schnurr, F. Gittes, F. C. MacKintosh, and C. F. Schmidt, “Determining microscopic viscoelasticity in flexible and

semiflexible polymer networks from thermal fluctuations,” Macromolecules 30(25), 7781–7792 (1997).42 R. Shlomovitz, A. Evans, T. Boatwright, A. J. Levine, and M. Dennin, “Measurement of monolayer viscosity using

noncontact microrheology,” Phys. Rev. Lett. 110(13), 137802 (2013).43 H. A. Lorentz, Abhandlungen uber theoretische Physik (B. G. Teubner, Braunschweig, 1907).44 K. B. Sørensen and H. Flyvbjerg, “Power spectrum analysis for optical tweezers,” Rev. Sci. Instrum. 75(3), 594–612

(2004).45 M. Fischer, A. C. Richardson, S. Nader, L. B. Oddershede, and K. B. Sørensen, “Active-passive calibration of optical

tweezers in viscoelastic media,” Rev. Sci. Instrum. 81(1), 015103 (2010).

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

128.200.29.82 On: Thu, 17 Sep 2015 01:30:10

Page 16: Probing interfacial dynamics and mechanics using …dennin/preprints/BDSE14-submerged...Probing interfacial dynamics and mechanics using submerged particle microrheology. II. Experiment

071904-15 Boatwright et al. Phys. Fluids 26, 071904 (2014)

46 S. Kim and S. J. Karrila, Microhydrodynamics: Principles and Selected Applications (Dover Publications, Mineola, NewYork, 2005).

47 A. J. Levine and T. C. Lubensky, “One- and two-particle microrheology,” Phys. Rev. Lett. 85(8), 1774–1777 (2000).48 J. C. Crocker, M. T. Valentine, E. R. Weeks, T. Gisler, P. D. Kaplan, A. G. Yodh, and D. A. Weitz, “Two-point microrheology

of inhomogeneous soft materials,” Phys. Rev. Lett. 85(4), 888–891 (2000).49 M. L. Kurnaz and D. K. Schwartz, “Channel flow in a Langmuir monolayer: Unusual velocity profiles in a liquid-crystalline

mesophase,” Phys. Rev. E 56, 3378–3384 (1997).

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

128.200.29.82 On: Thu, 17 Sep 2015 01:30:10