arXiv:1604.06836v1 [cond-mat.soft] 23 Apr 2016

8
Colloidal particle adsorption at liquid interfaces: Capillary driven dynamics and thermally activated kinetics Amir M. Rahmani, 1 Anna Wang, 2 Vinothan N. Manoharan, 2,3, * and Carlos E. Colosqui 1, 1 Department of Mechanical Engineering, Stony Brook University, Stony Brook, NY 11794, USA. 2 Harvard John A. Paulson School of Engineering and Applied Sciences, Harvard University, Cambridge, MA 02138, USA. 3 Department of Physics, Harvard University, Cambridge, MA 02138, USA. The adsorption of single colloidal microparticles (0.5–1 μm radius) at a water-oil interface has been recently studied experimentally using digital holographic microscopy [Kaz et al., Nat. Mater., 2012, 11, 138–142]. An initially fast adsorption dynamics driven by capillary forces is followed by an unexpectedly slow relax- ation to equilibrium that is logarithmic in time and can span hours or days. The slow relaxation kinetics has been attributed to the presence of surface “defects” with nanoscale dimensions (1–5 nm) that induce multiple metastable configurations of the contact line perimeter. A kinetic model considering thermally activated transi- tions between such metastable configurations has been proposed [Colosqui et al., Phys. Rev. Lett., 2013, 111, 028302] to predict both the relaxation rate and the crossover point to the slow logarithmic regime. However, the adsorption dynamics observed experimentally before the crossover point has remained unstudied. In this work, we propose a Langevin model that is able to describe the entire adsorption process of single colloidal particles by considering metastable states produced by surface defects and thermal motion of the particle and liquid interface. Invoking the fluctuation dissipation theorem, we introduce a drag term that considers significant dissipative forces induced by thermal fluctuations of the liquid interface. Langevin dynamics simulations based on the proposed adsorption model yield close agreement with experimental observations for different micropar- ticles, capturing the crossover from (fast) capillary driven dynamics to (slow) thermally activated kinetics. I. INTRODUCTION The adsorption and binding of colloidal particles at liquid interfaces is a fundamental phenomenon in colloid science that is relevant to applications in biomedical, environmental, food, and materials engineering. 1,2 The “irreversible” bind- ing of particles at liquid interfaces has been extensively ex- ploited for emulsion stabilization, 3 encapsulation, 4 wastewa- ter treatment,, 5 and self-assembly of novel materials. 6–8 De- spite the success of diverse technological applications, many fundamental aspects of the adsorption of colloidal particles at liquid interfaces remain poorly understood. Continuum ther- modynamic descriptions are conventionally employed to pre- dict the conditions under which colloidal particles bind to a liquid-fluid interface. Stable equilibrium conditions, where the system energy is minimized, are prescribed by the (Young) equilibrium contact angle determined by Young’s law. For colloidal particles larger than 10 nm, the reduction of sur- face energy at equilibrium can be several orders of magnitude larger than the thermal energy, and thus binding to the inter- face can become “irreversible”. 9,10 While conventional ther- modynamic models seem to describe certain fundamental as- pects of equilibrium behavior, an increasing number of exper- imental observations highlight the necessity of developing ac- curate models for the (non-equilibrium) dynamics of adsorp- tion. A spontaneous and rapid relaxation to thermodynamic equilibrium is expected for perfectly spherical particles of nano- and microscale dimensions when driving forces associ- ated with wetting and capillarity are much larger than damp- ing forces due to hydrodynamic effects. 11–15 Nevertheless, ex- perimental studies show that colloidal particle adsorption is neither spontaneous nor fast, owing to the presence of signif- icant energy barriers caused by different physicochemical ef- fects. The formation of electric double layers at liquid-liquid and liquid-solid interfaces is a common effect that prevents adsorption and/or hinders the relaxation to equilibrium. Elec- trostatic energy barriers can be suppressed through the addi- tion of electrolyte salts (e.g., NaCl) in sufficiently large con- centration so that surface charges are screened. 16,17 However, even for weak electrostatic effects, external work must be of- ten applied through hydrodynamic shear or sonication in or- der to observe significant adsorption rates. 18–21 Recent exper- imental work by Kaz et al. 22 has studied the adsorption of sin- gle polystyrene microspheres with different surface function- alities at an interface between a glycerol/water mixture and decane. Optical measurements by 3D digital holographic mi- croscopy revealed a “rapid” initial adsorption of the particles following the spontaneous breach of the water-oil interface, which occurs when the salt concentration is sufficiently high (>100 mM NaCl). 22 The initial rapid adsorption is followed by a surprisingly slow relaxation to equilibrium that is loga- rithmic in time and has been attributed to thermally activated pinning/depinning of the contact line at nanoscopic surface defects with areas ranging from 1 to 30 nm 2 . 22 Notably, the observed logarithmic relaxation can take place over several hours or even days and is reminiscent of physical aging in glassy systems with a free energy landscape that is densely populated by multiple local minima. 23,24 The logarithmic relaxation reported by Kaz et al. 22,25 sug- gests that the interfacial energy of a colloidal particle strad- dling a liquid-liquid interface has multiple local minima in- duced by nanoscale surface defects. The interplay between nanoscale surface defects and thermal fluctuations can induce the thermally activated displacement of the three-phase con- tact line, which results in nontrivial wetting processes. 26–28 arXiv:1604.06836v1 [cond-mat.soft] 23 Apr 2016

Transcript of arXiv:1604.06836v1 [cond-mat.soft] 23 Apr 2016

Page 1: arXiv:1604.06836v1 [cond-mat.soft] 23 Apr 2016

Colloidal particle adsorption at liquid interfaces: Capillary driven dynamicsand thermally activated kinetics

Amir M. Rahmani,1 Anna Wang,2 Vinothan N. Manoharan,2, 3, ∗ and Carlos E. Colosqui1, †

1Department of Mechanical Engineering, Stony Brook University, Stony Brook, NY 11794, USA.2Harvard John A. Paulson School of Engineering and Applied Sciences, Harvard University, Cambridge, MA 02138, USA.

3Department of Physics, Harvard University, Cambridge, MA 02138, USA.

The adsorption of single colloidal microparticles (0.5–1 µm radius) at a water-oil interface has been recentlystudied experimentally using digital holographic microscopy [Kaz et al., Nat. Mater., 2012, 11, 138–142].An initially fast adsorption dynamics driven by capillary forces is followed by an unexpectedly slow relax-ation to equilibrium that is logarithmic in time and can span hours or days. The slow relaxation kinetics hasbeen attributed to the presence of surface “defects” with nanoscale dimensions (1–5 nm) that induce multiplemetastable configurations of the contact line perimeter. A kinetic model considering thermally activated transi-tions between such metastable configurations has been proposed [Colosqui et al., Phys. Rev. Lett., 2013, 111,028302] to predict both the relaxation rate and the crossover point to the slow logarithmic regime. However,the adsorption dynamics observed experimentally before the crossover point has remained unstudied. In thiswork, we propose a Langevin model that is able to describe the entire adsorption process of single colloidalparticles by considering metastable states produced by surface defects and thermal motion of the particle andliquid interface. Invoking the fluctuation dissipation theorem, we introduce a drag term that considers significantdissipative forces induced by thermal fluctuations of the liquid interface. Langevin dynamics simulations basedon the proposed adsorption model yield close agreement with experimental observations for different micropar-ticles, capturing the crossover from (fast) capillary driven dynamics to (slow) thermally activated kinetics.

I. INTRODUCTION

The adsorption and binding of colloidal particles at liquidinterfaces is a fundamental phenomenon in colloid sciencethat is relevant to applications in biomedical, environmental,food, and materials engineering.1,2 The “irreversible” bind-ing of particles at liquid interfaces has been extensively ex-ploited for emulsion stabilization,3 encapsulation,4 wastewa-ter treatment,,5 and self-assembly of novel materials.6–8 De-spite the success of diverse technological applications, manyfundamental aspects of the adsorption of colloidal particles atliquid interfaces remain poorly understood. Continuum ther-modynamic descriptions are conventionally employed to pre-dict the conditions under which colloidal particles bind to aliquid-fluid interface. Stable equilibrium conditions, wherethe system energy is minimized, are prescribed by the (Young)equilibrium contact angle determined by Young’s law. Forcolloidal particles larger than 10 nm, the reduction of sur-face energy at equilibrium can be several orders of magnitudelarger than the thermal energy, and thus binding to the inter-face can become “irreversible”.9,10 While conventional ther-modynamic models seem to describe certain fundamental as-pects of equilibrium behavior, an increasing number of exper-imental observations highlight the necessity of developing ac-curate models for the (non-equilibrium) dynamics of adsorp-tion.

A spontaneous and rapid relaxation to thermodynamicequilibrium is expected for perfectly spherical particles ofnano- and microscale dimensions when driving forces associ-ated with wetting and capillarity are much larger than damp-ing forces due to hydrodynamic effects.11–15 Nevertheless, ex-perimental studies show that colloidal particle adsorption isneither spontaneous nor fast, owing to the presence of signif-

icant energy barriers caused by different physicochemical ef-fects. The formation of electric double layers at liquid-liquidand liquid-solid interfaces is a common effect that preventsadsorption and/or hinders the relaxation to equilibrium. Elec-trostatic energy barriers can be suppressed through the addi-tion of electrolyte salts (e.g., NaCl) in sufficiently large con-centration so that surface charges are screened.16,17 However,even for weak electrostatic effects, external work must be of-ten applied through hydrodynamic shear or sonication in or-der to observe significant adsorption rates.18–21 Recent exper-imental work by Kaz et al.22 has studied the adsorption of sin-gle polystyrene microspheres with different surface function-alities at an interface between a glycerol/water mixture anddecane. Optical measurements by 3D digital holographic mi-croscopy revealed a “rapid” initial adsorption of the particlesfollowing the spontaneous breach of the water-oil interface,which occurs when the salt concentration is sufficiently high(>100 mM NaCl).22 The initial rapid adsorption is followedby a surprisingly slow relaxation to equilibrium that is loga-rithmic in time and has been attributed to thermally activatedpinning/depinning of the contact line at nanoscopic surfacedefects with areas ranging from 1 to 30 nm2.22 Notably, theobserved logarithmic relaxation can take place over severalhours or even days and is reminiscent of physical aging inglassy systems with a free energy landscape that is denselypopulated by multiple local minima.23,24

The logarithmic relaxation reported by Kaz et al.22,25 sug-gests that the interfacial energy of a colloidal particle strad-dling a liquid-liquid interface has multiple local minima in-duced by nanoscale surface defects. The interplay betweennanoscale surface defects and thermal fluctuations can inducethe thermally activated displacement of the three-phase con-tact line, which results in nontrivial wetting processes.26–28

arX

iv:1

604.

0683

6v1

[co

nd-m

at.s

oft]

23

Apr

201

6

Page 2: arXiv:1604.06836v1 [cond-mat.soft] 23 Apr 2016

2

Previous work by Colosqui et al.29,30 proposed a kinetic modelemploying Kramers’ theory of thermally activated transitionson a one-dimensional energy profile that has multiple minimainduced by periodic energy barriers with magnitude ∆F ∝Ad and period λ = Ad/2πR; here, Ad is the characteristicdefect area and R is the particle radius. The proposed analyt-ical model predicts that the defect area Ad and particle radiusR prescribe both the logarithmic relaxation rate and the crit-ical distance from equilibrium where the adsorption processbecomes logarithmic in time.

While previous work22,29 has focused on rationalizing theslow logarithmic relaxation observed near equilibrium, no an-alytical models have been proposed to describe the observedadsorption dynamics preceding the slow logarithmic regime.Recent experimental measurements of the (in-plane) diffusiv-ity of microparticles at interfaces by Boniello et al.31 revealedthat hydrodynamic drag models cannot explain the dissipa-tive forces experienced by a particle straddling a liquid-fluidinterface at different contact angles. Boniello et al. pro-posed that thermal motion of the liquid interface can inducea dominant contribution to (in-plane) dissipative forces thatcan be estimated via the fluctuation-dissipation theorem.32

Building on prior work by Colosqui et al.,29 this work in-troduces a Langevin model that is able to describe the en-tire adsorption dynamics, comprising (fast) capillary-drivenand (slow) thermally activated regimes, by considering (1) asurface energy landscape with metastable states induced bynanoscale defects, (2) random fluctuations of surface forcesinduced by thermal motion of the colloidal particle and liquid-liquid interface, and (3) dissipative forces required to satisfythe fluctuation-dissipation relation.

II. THEORETICAL MODEL

To model the adsorption dynamics we will assume that theparticle undergoes uncorrelated Brownian motion. Thereforewe employ a classical Langevin equation29,32,33 for the evo-lution of the (center-of-mass) vertical position z(t). For con-venience, z(t) is measured relative to the liquid-fluid inter-face located at z = 0 (see Fig. 1(a)). Neglecting electrostaticand gravitational effects, the resulting force on the particleF = Fs + Fd + Ft is the sum of surface forces Fs, dissipa-tive forces Fd, and forces due to thermal motion Ft. We willconsider that the surface force Fs = 〈Fs〉 + F ′s comprisesa mean deterministic component 〈Fs〉 = −∂F/∂z deter-mined by the surface free energy F , and a zero-mean randomforce F ′s due to thermal fluctuations of the liquid interface.The fluctuation-dissipation theorem32,33 states that dissipativeforces Fd = −ξz and random thermal forces Ft =

√2kBTξη

are related through the effective friction coefficient ξ; here,η(t) is a (zero-mean unit-variance) white noise, kB is theBoltzmann constant and T is the temperature of the fluids.The proposed Langevin equation for the particle adsorption

dynamics thus reads29

mz = −∂F∂z− ξz +

√2kBTξη(t), (1)

where m is the particle mass. The Langevin model in eqn (1)is valid when the particle is in thermal equilibrium withthe surrounding fluids and its position is subject to uncorre-lated Brownian motion. Despite the simplifying assumptions,eqn (1) has been able to produce close agreement with fullyatomistic molecular dynamics simulations in previous work29

where the fluctuations of the interface and particle may be cor-related.

A. Surface forces

Within the framework of continuum thermodynamics,changes in the Helmholtz free energy of a colloidal particleat a liquid-fluid interface give the reversible work performedby the system at constant volume and temperature. If we as-sume symmetry about the polar and azimuth axes, the surfacefree energyF = F(z) can be parametrized with respect to theparticle center of mass position z, measured normal to the flatliquid-fluid interface (see Fig. 1a). The surface free energyof a spherical particle of radius R straddling a perfectly flatliquid interface located at z = 0 can be expressed as10,29

FS(z) = γπ(z − zE)2 + τ2π√R2 − z2 + C (2)

for |z| ≤ R, where zE = −R cos θE is the equilibrium posi-tion determined by the (Young) equilibrium contact angle θE ,γ is the liquid-fluid surface tension, τ is the line tension, andC is an additive constant.

While the surface tension γ is always positive and has val-ues on the order of 10 mN/m for water-oil interfaces, the linetension τ , associated with the excess free energy at the three-phase contact line, can be either positive or negative with amagnitude on the order of 10 pN, according to theoreticalestimates.34 It can be challenging to determine the line ten-sion empirically. Different experimental studies with sim-ple liquids on smooth substrates report values of τ rangingbetween 1 and 1000 pN.35–38 Defining a dimensionless linetension τ = τ/γR, we can identify two situations in whichforces induced by line tension cannot be neglected: (1) whenthe particle barely straddles the liquid interface, such that|z−R| < τ2/2(1+cos θE)2, and (2) when the particle is veryclose to equilibrium, such that |z − zE | < τ/ tan θE . For themicroparticles and water-oil interface studied in this work22

we haveR ' 1 µm, γ ' 4×10−2 N/m and θE = 110◦, whichgives τ . 10−4–10−1. Thus, the line tension contributions ineqn (2) can be neglected for the experimentally observed tra-jectories z(t).

For perfectly spherical particles and negligible line tensioncontributions, eqn (2) predicts a single energy minimum atz = zE and a relaxation to equilibrium driven by mono-tonic reduction of surface energy. According to eqn (2),

Page 3: arXiv:1604.06836v1 [cond-mat.soft] 23 Apr 2016

3

𝜃

Δ𝜃𝑖

𝑧′

𝑟

Δ𝐹′𝑆

𝜃𝑅

𝑧′

𝑟𝑧

𝑅

(a) (b)

𝛾

𝛾𝑠2

𝛾𝑠1

Phase 2

(decane oil)

Phase 1

(water/glycerol+NaCl)

Phase 2

Phase 1𝜇1 = 𝜇

𝜇2 = 𝜇/𝜅

FIG. 1. Colloidal particle at an interface. (a) Spherical particleof radius R straddling a sharp flat interface between two immisci-ble phases at a contact angle θ = arccos(−z/R). The surfaceenergies γ, γs1, γs2 determine the expected equilibrium positionzE = −R cos θE , where cos θE = (γs2 − γs1)/γ. For nanoscalesurface defects (

√Ad � R), the free energy profile F(z) is given

by eqn (3). (b) Fluctuating surface forces F ′s =∑

i ∆F ′i along thez-direction are induced by thermal fluctuations of the local contactangle θi = θ + ∆θi. The total surface force Fs = −∂F/∂z + F ′scomprises average (deterministic) and fluctuating (random) compo-nents.

when colloidal particles with nano- or microscale dimensions(R > 10 nm) approach equilibrium z → zE , they become “ir-reversibly” adsorbed at the interface because the depth of theenergy reduction (i.e., the binding energy) is several orders ofmagnitude larger than the thermal energy.

1. Nanoscale surface defects.

Experimental observations for colloidal microparticles22,25

have revealed an unexpectedly slow approach to equilibriumdue to the presence of nanoscale surface defects not consid-ered in eqn (2). As in previous work by Colosqui et al.29

we aim to develop an analytical model that incorporates lo-cal minima in the surface energy profile F(z) that are in-duced by multiple surface defects with nanoscale dimensions.For this purpose, we assume that the surface of the particle isdensely populated by defects with a characteristic nanoscalearea Ad ∼ O(1 nm2) that produce surface energy perturba-tions of magnitude ∆F ' γAd > kBT . The displacement ofthe contact line over a single defect of characteristic area Ad

changes the surface energy by an amount ∆F ∼ γAd and theaverage contact line position by an amount λ = Ad/2πR �R. To model this process we introduce a periodic spatial per-turbation of magnitude ∆F and period λ in the particle freeenergy so that

F(z) = FS(z) +1

2∆F sin

(2π

λ(z − zE)− ϕ

). (3)

The phase ϕ of the spatial perturbation produces a negligibleshift ∆zE ≤ λ � R of the equilibrium position and thuscan be arbitrarily chosen; we adopt ϕ = −π/2 so that theglobal energy minimum in eqn (3) remains at the equilibrium

position z = zE determined by the equilibrium contact angleθE .

2. Fluctuating surface forces.

For the case of a liquid-fluid interface fluctuating due tothermal motion, the surface force can be decomposed intomean and fluctuating components Fs = 〈Fs〉 + F ′s, where〈Fs〉 = −∂F/∂z and 〈F ′s〉 = 0; hereafter, angle brackets 〈 〉indicate ensemble averages. The effects of spatial fluctuationsof the contact line induced by localized defects on the parti-cle surface are considered in the free energy profile F(z) ineqn (3), and thus in the deterministic mean surface force 〈Fs〉.We will consider the case when the fluctuating force F ′s is dueto thermal motion of the liquid-fluid interface, which causesrandom fluctuations of the local contact angle θi = 〈θi〉+∆θi,where the average contact angle is 〈θi〉 = θ = arccos(−z/R)(see Fig. 1b) and 〈∆θi〉 = 0.

The fluctuating surface force normal to the interfacecan be estimated by summing small contributions ∆F ′s 'γsi(cos θi − cos θ)/ sin θ (see Fig. 1b) from contact line seg-ments having nanoscale dimensions si '

√Ad. In similar

fashion, Boniello et al. have estimated fluctuating surfaceforces parallel to the interface.31 Assuming small random fluc-tuations |∆θi| < 1 of the local contact angle, we obtain thefluctuating surface force

F ′s ' γ√Ad

N∑i=1

∆θi, (4)

where N = 2πR sin θ/√Ad. The fluctuating surface force

F ′s given by eqn (4) has zero mean 〈F ′s〉 = 0 but finite vari-ance 〈F ′s

2〉 > 0, and thus produces a finite energy input to thetranslational motion of the particle.

B. Dissipative forces

A key aspect of the Langevin dynamics modeled by eqn (1)is that deterministic dissipative forces and random forcesdue to thermal fluctuations are related through the effec-tive friction coefficient ξ, in accordance with the fluctuation-dissipation theorem. For systems in thermal equilibrium,the fluctuation-dissipation relation ensures that the energy in-put due to random thermal forces is strictly balanced, in the“long-time” limit, by dissipation of energy to the surround-ing fluids.32,33 The friction coefficient ξ can be estimated fromdifferent theoretical models for predicting dissipative forces.Alternatively, it is possible to determine the effective fric-tion coefficient by employing the Green–Kubo relation32,39

ξ = (1/kBT )∫∞0〈Ft(0)Ft(t)〉dt, which involves the time au-

tocorrelation of the random thermal forces Ft acting on theparticle. We will assume that the thermal force Ft = F ′z + F ′snormal to the liquid interface is the sum of the random trans-lational force F ′z due to thermal fluctuations of the fluid bulk

Page 4: arXiv:1604.06836v1 [cond-mat.soft] 23 Apr 2016

4

and the random surface force F ′s due to thermal fluctuationsof the liquid interface. Further assuming that these randomforces are uncorrelated (〈F ′zF ′s〉 = 0), the friction coeffi-cient ξ = ξz + ξs is the sum of ξz , which will be deter-mined by dynamic arguments, and ξs, which will be estimatedby the fluctuation-dissipation relation as proposed recently byBoniello et al.31

1. Translational drag: hydrodynamics and contact line dynamics.

Under creeping flow conditions the friction coefficient as-sociated with translational motion can be generally expressedas ξz(z) = (µ1K1 + µ2K2)R = (K1 + K2/κ)µR, whereµ1 = µ and µ2 = µ/κ are the viscosities of each fluid phase,κ = µ1/µ2 > 1 is the viscosity ratio, and K1,2 are position-dependent drag coefficients that can be estimated using avail-able hydrodynamic models.11–15 Given the relatively large vis-cosity ratio κ = 12.8 for the experimental conditions studiedin this work,22 µ = µ1 is adopted hereafter as the character-istic viscosity. As in previous studies,22,31 we find that trans-lational friction coefficients ξz predicted by different hydro-dynamic models for creeping flow conditions11–15 are two tothree orders of magnitude smaller than those required to agreewith experimental observations [see Sec. III C for details].

The translational motion of the contact line perimetercan be considered to produce dissipation of energy. Ac-cording to the molecular kinetic theory (MKT) developedby Blake and co-workers,40–42 the energy “consumed” inthe adsorption and desorption of molecules at the movingcontact line can be modeled as a virtual dissipative forceFd = −ξz z determined by the friction coefficient ξz =2πR sin θµ(ν/λ3M ) exp(Wa/kBT ), where ν is the molecu-lar volume of the liquid in phase 1, Wa = γλ2M (1 + cos θE)is the work of adhesion, and λM is the characteristic size ofthe adsorption sites. The characteristic size λM of the adsorp-tion sites is treated as a free parameter that must, however,remain within molecular dimensions (i.e., λM < 1 nm) forconsistency with the premises of MKT.43 As discussed in de-tail later in Sec. III C, maximum friction coefficients ξs(z =0) = O(0.1–1µN s/m) predicted by MKT for λM = 0.5–1 nmproduce dissipative forces that are two to three orders of mag-nitude smaller than those required to describe the experimen-tally observed adsorption dynamics. As reported in previousexperimental studies,22,31 we find that damping forces due tohydrodynamics and molecular kinetic effects considered byMKT cannot fully explain the dynamic behavior of particlesat interfaces.

2. Interfacial fluctuations: Green–Kubo relations.

The thermal motion of the liquid interface causes fluctuat-ing surface forces on the particle (cf. Fig. 1b) that must bebalanced by dissipative forces Fd = −ξsz in order to satisfy

the fluctuation-dissipation relation. Hence, the friction coeffi-cient ξs can be estimated by the Green–Kubo relation

ξs =1

kBT

∫ ∞0

〈F ′s(0)F ′s(t)〉dt '1

kBT〈F ′s

2〉 × tc, (5)

where tc is the correlation time, assuming an Ornstein–Uhlenbeck process for the relaxation of the interface. In orderto estimate the friction coefficient in eqn (5) we will begin byconsidering that contact angle fluctuations ∆θi along the con-tact line perimeter relax through capillary waves that travelat a speed vc = γ/µ. We can then estimate the correlationtime tc = χ(2πR sin θ/vc) where χ = O(1) is a scaling fac-tor accounting for the number of times capillary waves travelaround the contact line perimeter before correlation betweenlocal fluctuations is completely lost.

Invoking eqn (4) the variance of the fluctuating force is〈F ′s

2〉 = γ2√Ad(2πR sin θ)〈∆θ2i 〉, where the mean square

fluctuation of the local contact angle 〈∆θ2i 〉 ' kBT/γAd isdetermined by a balance between thermal energy and elasticsurface deformation. Incorporating the expressions above intoeqn (5) we arrive at the simple expression

ξs = χµ4π2R2

√Ad

(1− (z/R)2

)(6)

for the friction coefficient determining dissipative forces in-duced by thermal fluctuations of the interface. For the caseof microparticles (R = O(1 µm)) with nanoscale defect ar-eas

(Ad = O(1 nm2)

), eqn (6) predicts a maximum friction

coefficient ξs(z = 0) = O(100 µN s/m), which has the mag-nitude required for agreement between the Langevin model ineqn (1) and experimental observations.22,25

C. Crossover to thermally activated adsorption

As the particle approaches equilibrium, where 2γ|z−zE | <∆F/λ, the free energy F in eqn (3) has local energy minimawhere ∂F/∂z = 0 and the mean surface force vanishes. Asa result, sufficiently close to equilibrium the adsorption dy-namics described by eqn (1) is no longer driven by a mono-tonic reduction of surface energy and becomes dominated bythermally activated transitions between multiple metastablestates.29 The crossover to thermally activated adsorption is agradual process that takes place over a finite position interval|z − zC | > 0 that is approximately centered at the crossoverpoint z(t) = zC . According to previous work by Colosqui etal.29 the crossover point zC can be estimated by the equation

|zC − zE | = α∆F2γλ

(7)

for 0.5 . α . 1. The crossover position zC is prescribed bythe free energy profile and is independent of frictional and dis-sipative effects. Provided that both the energy barrier ∆F andperiod λ employed in eqns (1)–(3) are properly determined,the crossover criterion in eqn (7) can estimate the critical dis-tance from equilibrium below which the adsorption becomeslogarithmic in time.

Page 5: arXiv:1604.06836v1 [cond-mat.soft] 23 Apr 2016

5

III. RESULTS

The proposed adsorption model consisting of eqns (1)–(3)and (6) has formally three adjustable parameters: (i) the en-ergy barrier magnitude ∆F , and (ii) the period λ prescrib-ing the free energy profile in eqn (3), along with (iii) the di-mensionless correlation time, or correlation factor, χ ∼ O(1)determining the friction coefficient in eqn (6). However, theenergy barrier magnitude and period must be determined bythe same nanoscale defect area Ad for consistency with phys-ical arguments. Employing a correlation factor χ = 3 for thefriction coefficient in eqn (6) produces the dynamics reportedin Figs. 2–4. While small variations in the energy barrier ∆Fand period λ produce large differences in the modeled dynam-ics, the correlation factor χ can be varied by about 20% with-out observing significant effects.

For the functionalized polystyrene microspheres(R ' 1 µm) studied in this work, previous experiments22,25

have found that the defect area Ad ranges from 5 to 30 nm2

and is uncorrelated with the particle radius. The experimentalsystem is at room temperature (T = 21 ± 3◦C) and consistsof two immiscible phases: phase 1 is a NaCl solution (>100mM NaCl) in a water/glycerol (45:55% w/w) mixture withviscosity 11.5 mPa s, and phase 2 is a anhydrous decanewith viscosity 0.9 mPa s. The viscosity µ = 11.5 mPa s ofthe aqueous phase (phase 1) is adopted as the characteristicviscosity given the large viscosity ratio (κ > 10). The surfacetension at the water-oil interface is γ = 0.037 N/m andthe equilibrium contact angle for all the studied particles isθE = 110◦.22 The experimental measurements have a timeresolution of 0.2–10 ms and spatial resolution of 1 nm.

The energy barrier induced by a (three-dimensional) sur-face defect on the particle surface involves a change of (1)the liquid-liquid interface surface area ∆A12 and (2) the solidwetted area ∆As1, both comparable but not necessarily equalto characteristic defect area Ad.29,30 Hence, the magnitudeof the energy barrier ∆F in eqn (3) is given by ∆F '|γ(∆A12 + ∆As1 cos θd)| = βγAd where β ∼ O(1) is ashape factor prescribed by the specific geometry of the defectand the Young equilibrium contact angle θd determined bylocal surface energies. Naively assuming homogeneous sur-face energy so that θd = θE and a spherical cap defect withprojected area Ad and protruding height hd = 0–0.5

√Ad/π

leads to β = 0.35–0.2, which is close to the values employedfor fitting the experimental results [see Table I].

A. Langevin simulations

In Table I we report the model parameters employed ineqns (1)–(3) for the Langevin Dynamics (LD) simulations inthis work. The LD simulations must employ extremely smalltime steps ∆t = 0.1 ps to accurately resolve forces pro-duced by the free energy in eqn (3), which has spatial oscil-lations with very small periods λ = Ad/2πR = 1.6–3.6 pm[see Table I]. Simulated particle trajectories z(t) reported in

77

70

63

56

48

38

26

0

10-6

10-4

10-2

100

0.5

0.6

0.7

0.8

0.9

1

1.1

1.2

79

73

66

59

52

44

34

20

10-6

10-4

10-2

100

0.6

0.8

1

1.2

𝑡 [𝑠]

|𝑧−𝑧 𝐸|[𝜇m]

𝜃[deg]

Experimental data

LD [eqns (1)-(3)]

eqn (7) (𝛼 = 0.9)

(b)

𝑡 [𝑠]

|𝑧−𝑧 𝐸|[𝜇m]

𝜃[deg]Experimental data

LD [eqns (1)-(3)]

(c)

from oil phase

from water phase

110

97

84

70

56

38

0

10-6

10-4

10-2

100

0

0.2

0.4

0.6

0.8

1

1.2

𝑡 [𝑠]

|𝑧−𝑧 𝐸|[𝜇m]

𝜃[deg]

Experimental data

LD [eqns (1)-(3)]

eqn (7) (𝛼 = 0.8)

(a)

from oil phase

from water phase

from oil phase

from water phase

eqn (7) (𝛼 = 0.9)

FIG. 2. Adsorption dynamics: Langevin dynamics versus experi-ments. Separation from equilibrium |z(t) − zE | versus time forthree polystyrene particles of radius R = 0.9µm with differ-ent surface functional groups. (a) Carboxyl, (b) Sulphate, and (c)Carboxylate-modified.22 The contact angle on the right axis corre-sponds to θ = arccos(−z/R). Markers: Experimental observa-tions from Ref. 22. Solid lines: LD simulation (eqns (1)–(3)) us-ing model parameters reported in Table I; the adsorption is initiatedfrom the water and oil phase. Dashed lines: Logarithmic-in-time fits|z(t)− zE | = c1 + c2 log t (c1,2 = const.). Horizontal dotted lines:Crossover distance |zC−zE | determined by eqn (7) for α = 0.8–0.9.

Fig. 2 for three different functionalized particles (carboxyl,sulphate, and carboxylate-modified surface groups) of radiusR = 0.9 µm are in close agreement with experimental obser-vations for adsorption initiated from the aqueous phase.22

Notably, LD simulations describe both the initial adsorp-tion dynamics, driven by capillary forces induced by reduc-tion of surface energy, and the transition to a thermally acti-vated regime where the particle follows a logarithmic-in-timetrajectory as predicted in Ref. 29. The crossover positionzC in all studied cases can be estimated by the criterion ineqn (7) using α = 0.8–0.9 [see Fig. 2]. Simulations for threestudied particles show that the adsorption dynamics can bedramatically different when it is initiated from the oil phase,

Page 6: arXiv:1604.06836v1 [cond-mat.soft] 23 Apr 2016

6

TABLE I. Model parameters employed for Langevin simulations (eqns (1)–(3)) of colloidal particles experimentally studied in Ref. 22.Particle class Ad (nm2) ∆F = βγAd λ = Ad/2πR (pm) βCarboxyl 9.2 19.0 kBT 1.63 0.23Sulphate 12.9 48.9 kBT 2.34 0.41Carboxylate-modified 85.7 360 kBT 15.5 0.45

where the initial position z(t = 0) is much closer to equilib-rium. Extremely low relaxation rates and kinetically trappedstates are observed for large energy barriers ∆F > 20 kBT[cf. Fig. 2b–c] when the adsorption is initiated in the region|z(t = 0) − zE | < |zC − zE |, and the entire process is dom-inated by thermally activated transitions between metastablestates.

B. Particle size effects

Experimental observations by Kaz et al.22 have reportedthat the defect area Ad is uncorrelated with the particle sizeand that it is predominantly determined by the surface func-tionalization. A series of LD simulations for particles witha fixed defect area (Ad = 9.2 nm2 and 12.9 nm2) and threedifferent radii R = 0.5, 1, and 1.5 µm are reported in Fig. 3.Despite the fact that damping forces scale with R2, the ad-sorption dynamics of smaller particles is not significantlyfaster. When the radius of the particle is reduced the ad-sorption dynamics enters the slow logarithmic regime muchearlier, which is consistent with the crossover criterion ineqn (7). Simulations for the studied range of microparticleradii R = 0.5–1.5 µm with energy barriers much larger thanthe thermal energy ∆F > 20 kBT (cf. sulphate functional-ized particles in Fig. 3(b)) reveal that the logarithmic-in-timerelaxation to equilibrium cannot be observed within the ex-perimental observation window (0 to 2 s) if the adsorption isinitiated from the oil phase; these findings are consistent withrecent experimental observations.25

C. Dissipative effects: Drag models comparison

To determine the dissipative force Fd = −ξz employedin eqn (1) we considered different contributions predicted byhydrodynamic models, Blake’s MKT theory for contact linedynamics, and eqn (6) derived by invoking the fluctuation-dissipation theorem. For the case of hydrodynamic drag con-tributions, we found that models based on Stokes flow pre-dict friction coefficients that are smaller than or comparableto those predicted by Scriven’s corner flow model in the limitof physically meaningful values of the molecular cutoff lengthε.11 In Fig. 4a we report the value of the position-dependentfriction coefficient ξ(z) predicted for the studied experimen-tal conditions by three different models: (1) Scriven’s hy-drodynamic model for corner flow adopting a molecular cut-off length ε = 1 nm,11 (2) MKT predictions41 for two ad-sorption site sizes λM = 1.3 nm and λM = 1.5 nm, and

10-6

10-4

10-2

100

0

0.5

1

1.5

2

10-6

10-4

10-2

100

0.5

1

1.5

2

𝑡 [𝑠]

𝑡 [𝑠]

𝑧−𝑧 𝐸

[𝜇m]

|𝑧−𝑧 𝐸|[𝜇m]𝑅 = 0.5 𝜇m𝑅 = 1 𝜇m𝑅 = 1.5 𝜇m

𝑅 = 0.5 𝜇m𝑅 = 1 𝜇m𝑅 = 1.5 𝜇m

(a)

(b)

FIG. 3. Particle size variation. Separation from equilibrium |z(t) −zE | versus time for particles of radius R = 0.5, 1, and 1.5 µm: (a)Carboxyl-functionalized particles, (b) Sulphate-functionalized par-ticles. Solid lines: LD simulations of eqns (1)–(3) using parame-ters reported in Table I for adsorption initiated from the water phase(thicker lines) and from the oil phase (thinner lines). Horizontal dot-ted lines: Crossover position predicted by eqn (7) for α = 0.8 (a) andα = 0.9 (b).

(3) the expression derived in eqn (6) with a correlation fac-tor χ = 3. For a carboxyl functionalized particle withR = 0.9 µm and nanoscale defect area Ad = 12 nm2 [seeTable 1], eqn (6) predicts a maximum friction coefficientξs(z = 0) ' 6.2×10−4 N s/m, which is within the magnituderange required for agreement between the adsorption dynam-ics modeled by eqn (1) and experimental observations.22,25 Asshowed in Figs. 4a–b, large dissipative forces are required todescribe the experimentally observed trajectories. Such dis-sipative forces are predicted by the friction coefficient ξs ineqn (6), which can be several orders of magnitude larger thanfriction coefficients due to hydrodynamic and molecular ki-netic effects. Employing MKT predictions with an adsorptionsite of size λM = 1.5 nm results in a friction coefficient ξthat is sufficiently large to produce a reasonable (but not op-timal) fit to the experimental data [cf Fig. 4b]. However, theenergy barrier ∆F = 19.2 kBT employed to describe the

Page 7: arXiv:1604.06836v1 [cond-mat.soft] 23 Apr 2016

7

|𝑧 − 𝑧𝐸|

𝜉(𝑧)[𝑚

N/m

]𝜃[deg]

eqn (6) (𝜒 = 3)

MKT (𝑙𝑀 = 1.5 nm)

MKT (𝑙𝑀 = 1. 3 nm)

Scriven (𝜀 = 1 nm)

𝑡 [𝑠]

|𝑧−𝑧 𝐸|[𝜇m]

𝜃[deg]

110

97

84

70

56

38

0

10-8

10-6

10-4

10-2

100

0

0.2

0.4

0.6

0.8

1

1.2

(a)

(b)

FIG. 4. Dissipative forces and adsorption dynamics. (a) Friction co-efficient ξ(z) versus particle separation from equilibrium |z − zE |.(b) Separation from equilibrium |z − zE | versus time adopting fric-tion coefficients reported in (a). Results correspond to the carboxylfunctionalized particle reported in Table I. Solid lines: Friction co-efficient derived via Green–Kubo relations (eqn (6)) with χ = 3.Dashed-dotted lines: MKT friction coefficient for λM = 1.5 nm.Dashed lines: MKT friction coefficient for λM = 1.3 nm. Dottedlines: Scriven’s hydrodynamic models with molecular cutoff param-eter ε = 1 nm.

logarithmic relaxation rate and estimate the crossover pointvia eqn (7) is significantly larger than the work of adsorptionWa = 13.9 kBT corresponding to λM = 1.5 nm. More-over, the use of molecular adsorption sites with dimensionsλM ≥ 1 nm is beyond the original premises of Blake’s MKT.As indicated by Blake and co-workers,44 modeling the effectsof nano- and mesoscale surface defects is likely to require theuse of alternative models based on Kramers theory.29,30 No-tably, for the range of defect sizes Ad = 9.2–85.7 nm2 em-ployed to fit the logarithmic relaxation rates of different par-ticles, the friction coefficient ξs (eqn (6)) derived from thefluctuation-dissipation theorem accounts quantitatively for theobserved adsorption dynamics.

IV. CONCLUSIONS

We have proposed a Langevin dynamics approach for theadsorption of single colloidal particles that considers deter-ministic spatial fluctuations of the free energy induced bynanoscale surface defects and random thermal fluctuations ofthe particle position and local contact angle with the liquid

interface. The proposed Langevin model is able to describethe entire adsorption dynamics of single microparticles withdifferent surface functional groups that has been reported inprevious experimental work. The adsorption dynamics ex-hibits two remarkably different regimes: (i) capillary-drivenadsorption dominated by reduction of surface energy, and (ii)thermally activated adsorption governed by random transi-tions between multiple metastable states. The capillary-drivenregimes are observed during the initial stages of the adsorp-tion process and are damped by friction forces that are muchlarger than predicted by previously available drag models. Inour model, the thermal motion of the interface produces off-plane fluctuating surface forces that have zero mean and finitevariance and are uncorrelated with the off-plane translationalmotion of the particle. The energy injected into the particletranslation by fluctuating surfaces forces must be dissipatedby a large damping force in order to satisfy the fluctuation-dissipation relation for a system in thermal equilibrium. Bymeans of the Green-Kubo formalism we have derived a newanalytical expression that predicts dissipative forces with themagnitude and functional dependence on the particle positionthat are required to describe the experimentally observed dy-namics of adsorption. The crossover to slow thermally ac-tivated regimes observed in experiments and simulations canbe predicted by a simple analytical criterion (eqn (7)) based onthe emergence of local energy minima when thermodynamicequilibrium is approached.

The proposed Langevin model allows us to elucidate novelaspects of the dynamics of adsorption of single colloidal mi-crospheres at liquid-fluid interfaces and make a few usefulpredictions for further experimental validation. For exam-ple, invoking the crossover criterion in eqn (7) the charac-teristic defect area Ad = ∆F/(γπα|zC/R − cosE |), whereα = 0.8–0.9, could be estimated from the experimentally ob-served crossover position zC , once the energy barrier ∆F isdetermined by fitting the logarithmic regime. In the absence ofnanoscale characterization techniques to determine the shapeand characteristic areaAd of surface defects, the proposed an-alytical model along with experimental observations can serveas a probe of such nanoscale features. A better understand-ing of the adsorption dynamics, as provided by the modelsin this work, will help in determining fundamental propertiesof particles at interfaces, such as their equilibrium and non-equilibrium binding energy and contact angle.

We thank D.M. Kaz, J. Koplik, R. McGorty, and J.M. Mor-ris for useful discussions. AW and VNM acknowledge sup-port from the National Science Foundation under grant no.DMR-1306410 and from the Harvard MRSEC, supported byNSF grant no. DMR-1420570. AMR and CEC acknowl-edge support from the SEED Grant Program by The Officeof Brookhaven National Laboratory (BNL) Affairs at StonyBrook University (SBU). This work employed computationalresources from the Institute for Advanced Computational Sci-ence at SBU and the BNL Center for Functional Nanoma-terials supported by the U.S. DOE under Contract No. DE-SC0012704.

Page 8: arXiv:1604.06836v1 [cond-mat.soft] 23 Apr 2016

8

[email protected][email protected] B. P. Binks and T. S. Horozov, Colloidal particles at liquid inter-

faces, Cambridge University Press, 2006.2 P. Kralchevsky and K. Nagayama, Particles at fluid interfaces and

membranes, Elsevier Science New York, NY, USA, 2001.3 R. Aveyard, B. P. Binks and J. H. Clint, Adv. Colloid Interface

Sci., 2003, 100, 503–546.4 A. Dinsmore, M. F. Hsu, M. Nikolaides, M. Marquez, A. Bausch

and D. Weitz, Science, 2002, 298, 1006–1009.5 J. Rubio, M. Souza and R. Smith, Miner. Eng., 2002, 15, 139–155.6 Y. Lin, H. Skaff, T. Emrick, A. Dinsmore and T. Russell, Science,

2003, 299, 226–229.7 S. Biswas and L. T. Drzal, Nano Lett., 2008, 9, 167–172.8 R. McGorty, J. Fung, D. Kaz and V. N. Manoharan, Mater. Today,

2010, 13, 34–42.9 P. Pieranski, Phys. Rev. Lett., 1980, 45, 569.

10 R. Aveyard and J. H. Clint, J. Chem. Soc., Faraday Trans., 1996,92, 85–89.

11 C. Huh and L. Scriven, J. Colloid Interface Sci., 1971, 35, 85–101.12 J. T. Petkov, N. D. Denkov, K. D. Danov, O. D. Velev, R. Aust and

F. Durst, J. Colloid Interface Sci., 1995, 172, 147–154.13 K. D. Danov, R. Dimova and B. Pouligny, Phys. Fluids, 2000, 12,

2711–2722.14 P. Singh and D. Joseph, J. Fluid Mech., 2005, 530, 31–80.15 T. M. Fischer, P. Dhar and P. Heinig, J. Fluid Mech., 2006, 558,

451–475.16 H. Ohshima, T. W. Healy and L. R. White, J. Colloid Interface

Sci., 1982, 90, 17–26.17 E. Mbamala et al., J. Phys.: Condens. Matter, 2002, 14, 4881.18 A. Stocco, E. Rio, B. P. Binks and D. Langevin, Soft Matter, 2011,

7, 1260–1267.19 K. Du, E. Glogowski, T. Emrick, T. P. Russell and A. D. Dins-

more, Langmuir, 2010, 26, 12518–12522.20 V. Garbin, J. C. Crocker and K. J. Stebe, J. Colloid Interface Sci.,

2012, 387, 1–11.21 H. Wang, V. Singh and S. H. Behrens, J. Phys. Chem. Lett., 2012,

3, 2986–2990.22 D. M. Kaz, R. McGorty, M. Mani, M. P. Brenner and V. N.

Manoharan, Nat. Mater., 2012, 11, 138–142.

23 L. Struik, Polym. Eng. Sci., 1977, 17, 165–173.24 R. G. Palmer, D. L. Stein, E. Abrahams and P. W. Anderson, Phys.

Rev. Lett., 1984, 53, 958.25 A. Wang, D. M. Kaz, R. McGorty and V. N. Manoharan, AIP

Conference Proceedings, 2013, 1518, 336–343.26 A. Prevost, E. Rolley and C. Guthmann, Phys. Rev. Lett., 1999,

83, 348.27 K. Davitt, M. S. Pettersen and E. Rolley, Langmuir, 2013, 29,

6884–6894.28 C. E. Colosqui, T. Teng and A. M. Rahmani, Phys. Rev. Lett.,

2015, 115, 154504.29 C. E. Colosqui, J. F. Morris and J. Koplik, Phys. Rev. Lett., 2013,

111, 028302.30 S. Razavi, I. Kretzschmar, J. Koplik and C. E. Colosqui, J. Chem.

Phys., 2014, 140, 014904.31 G. Boniello, C. Blanc, D. Fedorenko, M. Medfai, N. B. Mbarek,

M. In, M. Gross, A. Stocco and M. Nobili, Nat. Mater., 2015,908–911.

32 R. Kubo, Rep. Prog. Phys., 1966, 29, 255.33 E. Hinch, J. Fluid Mech., 1975, 72, 499–511.34 A. Amirfazli and A. Neumann, Adv. Colloid Interface Sci., 2004,

110, 121–141.35 B. Widom, J. Phys. Chem., 1995, 99, 2803–2806.36 J. Drelich, Colloids Surf., A, 1996, 116, 43–54.37 A. Checco, P. Guenoun and J. Daillant, Phys. Rev. Lett., 2003, 91,

186101.38 J. K. Berg, C. M. Weber and H. Riegler, Phys. Rev. Lett., 2010,

105, 076103.39 D. J. Searles and D. J. Evans, J. Chem. Phys., 2000, 112, 9727–

9735.40 T. Blake and J. Haynes, J. Colloid Interface Sci., 1969, 30, 421–

423.41 T. D. Blake, J. Colloid Interface Sci., 2006, 299, 1–13.42 M. J. De Ruijter, T. Blake and J. De Coninck, Langmuir, 1999,

15, 7836–7847.43 M. Ramiasa, J. Ralston, R. Fetzer and R. Sedev, Adv. Colloid In-

terface Sci., 2014, 206, 275–293.44 T. Blake and J. De Coninck, Eur. Phys. J. Spec. Top., 2011, 197,

249–264.