Top Banner
Solid Character of Membrane Ceramides: A Surface Rheology Study of Their Mixtures with Sphingomyelin Elisa R. Catapano, Laura R. Arriaga, ‡§ Gabriel Espinosa, Francisco Monroy, †‡ * Dominique Langevin, and Iva ´n Lo ´ pez-Montero * Mechanics of Biological Systems, Departamento de Quı ´mica Fı ´sica I, Universidad Complutense, Madrid, Spain; Laboratoire de Physique des Solides, Universite ´ Paris Sud XI, Orsay, France; and § Department of Physics, Harvard University, Cambridge, Massachusetts ABSTRACT The compression and shear viscoelasticities of egg-ceramide and its mixtures with sphingomyelin were investi- gated using oscillatory surface rheology performed on Langmuir monolayers. We found high values for the compression and shear moduli for ceramide, compatible with a solid-state membrane, and extremely high surface viscosities when compared to typical fluid lipids. A fluidlike rheological behavior was found for sphingomyelin. Lateral mobilities, measured from particle tracking experiments, were correlated with the monolayer viscosities through the usual hydrodynamic relationships. In conclu- sion, ceramide increases the solid character of sphingomyelin-based membranes and decreases their fluidity, thus drastically decreasing the lateral mobilities of embedded objects. This mechanical behavior may involve important physiological conse- quences in biological membranes containing ceramides. INTRODUCTION Ceramides are sphingolipids (1) that can form head-to-head hydrogen bonds (2) to promote strong intermolecular inter- actions in biological membranes. As a consequence, ceram- ides in lipid membranes frequently segregate into ordered domains with gel- or solidlike character (3). In vivo, during natural apoptosis or when cell death is induced under stress, infection, or g-irradiation, ceramide appears in the plasma membrane as a product of sphingosine hydrolysis (4,5). Immunofluorescence experiments (Cy3-labeled anti-ceram- ide), performed on the plasma membrane during apoptosis, have revealed accumulation of ceramide into large plat- forms formed de novo from membrane sphingomyelin (4,5). Although, to our knowledge, the precise biological func- tion of the ceramide-enriched lipid domains has not yet been elucidated, an immobilizing role has been suggested to trigger the signaling process during apoptosis (6). In general, ceramides are mechanically different from the rest of the fluid membrane components, thus their biological function could be related to a certain mechanical interplay with other lipids and membrane proteins. According to recent experiments, we have hypothesized about the possible mechanical role of domains in the clustering and further immobilization of membrane signaling receptors (7,8). Due to its unfavorable molecular shape (inverted cone), ceramides do not form stable bilayers alone, only gaining sufficient stability when mixed with compatible bilayer- forming lipids, particularly sphingomyelin. Langmuir mono- layers of single ceramides could provide useful information to help us understand the mechanical properties of mem- brane ceramides in lipid mixtures including sphingomyelin. Studies on pure ceramide monolayers are relatively scarce and mainly focus on the equilibrium properties obtained from the compression isotherms (9–14). From those studies, ceramide monolayers are known to possess very low com- pressibility compared to typical fluid phospholipids. In a recent article, we demonstrated the strong differences in the shear flow properties between ceramide layers and typical fluid membranes (8). This article explores these differences with an extensive rheological (compression and shear) study of the viscoelasticity of ceramide layers contain- ing the metabolic precursor sphingomyelin. The viscoelastic parameters describing the capacity of the membrane for deformation and flow under an external stim- ulus have been measured in a number of biophysical studies (15). Fluidity, understood as inverse viscosity, is one of the principal attributes of biological membranes essential for maintaining a proper lipid distribution and an adequate diffusivity of the membrane proteins in the lipid bilayer— two molecular aspects key for membrane function (7,16). The shear flow through surface canals was exploited early to measure surface fluidity of monomolecular films of fatty acids (17). Since then, the information accumulated on the structure (18–21) and rheological properties of monomolec- ular layers (22–26) has crucially contributed to the creation of the modern picture of biological membranes as complex systems that are not only able to develop viscous flow under protein motion (27–29), but also rigid enough to endow suf- ficient mechanical resilience to support mechanical stability under external stress (15,30,31). The methods of surface rheology can be broadly divided into devices that operate at constant surface area providing shear properties and those that provide compression pro- perties from changes in surface tension caused by lateral packing deformations (see Fuller and Vermant (32) and ref- erences therein). Relevant studies on model systems require a combination of different experimental methods that can Submitted June 3, 2011, and accepted for publication October 31, 2011. *Correspondence: [email protected] or [email protected] Editor: Paulo F. Almeida. Ó 2011 by the Biophysical Society 0006-3495/11/12/2721/10 $2.00 doi: 10.1016/j.bpj.2011.10.049 Biophysical Journal Volume 101 December 2011 2721–2730 2721
10

Solid Character of Membrane Ceramides: A Surface Rheology Study of Their Mixtures with Sphingomyelin

Mar 06, 2023

Download

Documents

Welcome message from author
This document is posted to help you gain knowledge. Please leave a comment to let me know what you think about it! Share it to your friends and learn new things together.
Transcript
Page 1: Solid Character of Membrane Ceramides: A Surface Rheology Study of Their Mixtures with Sphingomyelin

Biophysical Journal Volume 101 December 2011 2721–2730 2721

Solid Character of Membrane Ceramides: A Surface Rheology Studyof Their Mixtures with Sphingomyelin

Elisa R. Catapano,† Laura R. Arriaga,‡§ Gabriel Espinosa,‡ Francisco Monroy,†‡* Dominique Langevin,‡

and Ivan Lopez-Montero†*†Mechanics of Biological Systems, Departamento de Quımica Fısica I, Universidad Complutense, Madrid, Spain; ‡Laboratoire de Physiquedes Solides, Universite Paris Sud XI, Orsay, France; and §Department of Physics, Harvard University, Cambridge, Massachusetts

ABSTRACT The compression and shear viscoelasticities of egg-ceramide and its mixtures with sphingomyelin were investi-gated using oscillatory surface rheology performed on Langmuir monolayers. We found high values for the compression andshear moduli for ceramide, compatible with a solid-state membrane, and extremely high surface viscosities when comparedto typical fluid lipids. A fluidlike rheological behavior was found for sphingomyelin. Lateral mobilities, measured from particletracking experiments, were correlated with the monolayer viscosities through the usual hydrodynamic relationships. In conclu-sion, ceramide increases the solid character of sphingomyelin-based membranes and decreases their fluidity, thus drasticallydecreasing the lateral mobilities of embedded objects. This mechanical behavior may involve important physiological conse-quences in biological membranes containing ceramides.

INTRODUCTION

Ceramides are sphingolipids (1) that can form head-to-headhydrogen bonds (2) to promote strong intermolecular inter-actions in biological membranes. As a consequence, ceram-ides in lipid membranes frequently segregate into ordereddomains with gel- or solidlike character (3). In vivo, duringnatural apoptosis or when cell death is induced under stress,infection, or g-irradiation, ceramide appears in the plasmamembrane as a product of sphingosine hydrolysis (4,5).Immunofluorescence experiments (Cy3-labeled anti-ceram-ide), performed on the plasma membrane during apoptosis,have revealed accumulation of ceramide into large plat-forms formed de novo from membrane sphingomyelin (4,5).

Although, to our knowledge, the precise biological func-tion of the ceramide-enriched lipid domains has not yetbeen elucidated, an immobilizing role has been suggestedto trigger the signaling process during apoptosis (6). Ingeneral, ceramides are mechanically different from the restof the fluid membrane components, thus their biologicalfunction could be related to a certain mechanical interplaywith other lipids and membrane proteins. According torecent experiments, we have hypothesized about the possiblemechanical role of domains in the clustering and furtherimmobilization of membrane signaling receptors (7,8).

Due to its unfavorable molecular shape (inverted cone),ceramides do not form stable bilayers alone, only gainingsufficient stability when mixed with compatible bilayer-forming lipids, particularly sphingomyelin. Langmuir mono-layers of single ceramides could provide useful informationto help us understand the mechanical properties of mem-brane ceramides in lipid mixtures including sphingomyelin.Studies on pure ceramide monolayers are relatively scarce

Submitted June 3, 2011, and accepted for publication October 31, 2011.

*Correspondence: [email protected] or [email protected]

Editor: Paulo F. Almeida.

� 2011 by the Biophysical Society

0006-3495/11/12/2721/10 $2.00

and mainly focus on the equilibrium properties obtainedfrom the compression isotherms (9–14). From those studies,ceramide monolayers are known to possess very low com-pressibility compared to typical fluid phospholipids. In arecent article, we demonstrated the strong differences inthe shear flow properties between ceramide layers andtypical fluid membranes (8). This article explores thesedifferences with an extensive rheological (compression andshear) study of the viscoelasticity of ceramide layers contain-ing the metabolic precursor sphingomyelin.

The viscoelastic parameters describing the capacity of themembrane for deformation and flow under an external stim-ulus have been measured in a number of biophysical studies(15). Fluidity, understood as inverse viscosity, is one of theprincipal attributes of biological membranes essential formaintaining a proper lipid distribution and an adequatediffusivity of the membrane proteins in the lipid bilayer—two molecular aspects key for membrane function (7,16).The shear flow through surface canals was exploited earlyto measure surface fluidity of monomolecular films of fattyacids (17). Since then, the information accumulated on thestructure (18–21) and rheological properties of monomolec-ular layers (22–26) has crucially contributed to the creationof the modern picture of biological membranes as complexsystems that are not only able to develop viscous flow underprotein motion (27–29), but also rigid enough to endow suf-ficient mechanical resilience to support mechanical stabilityunder external stress (15,30,31).

The methods of surface rheology can be broadly dividedinto devices that operate at constant surface area providingshear properties and those that provide compression pro-perties from changes in surface tension caused by lateralpacking deformations (see Fuller and Vermant (32) and ref-erences therein). Relevant studies on model systems requirea combination of different experimental methods that can

doi: 10.1016/j.bpj.2011.10.049

Page 2: Solid Character of Membrane Ceramides: A Surface Rheology Study of Their Mixtures with Sphingomyelin

2722 Catapano et al.

properly isolate the different rheological features into a bio-logical context. This refers to as physicochemical conditionsresembling the structure of the real lipid bilayer (18,33).Indeed, the rheological tool is so useful for getting newinsights into the mechanical function of biological mem-branes that such methods are currently being used to revealdynamical aspects related to the complex functional inter-play between the different membrane components and theirstructure (8,34).

This article addresses the fundamental question ofmembrane viscoelasticity of mixed ceramide/sphingomyelinlayers measured both in compression and in shear, witha special focus on the solidlike character of the membraneemplacements enriched in ceramide. We compared the me-chanical behavior of Langmuir monolayers of egg-ceramide(eggCer) and their mixtures with egg-sphingomyelin(eggSM), which we discussed in terms of the lateral diffusiv-ities of motion probes embedded in the monolayer (35).For the sake of relevance, the work is restricted to a nearphysiological state, usually assumed to correspond to amonolayer packing compatible with a surface pressure ofpbilz 30 mN/m (33).

MATERIALS AND METHODS

Materials

Lipids and chemicals

Egg ceramide (eggCer) and egg sphingomyelin (eggSM) were supplied by

Avanti Polar Lipids (Alabaster, AL) as high-purity powder. Just before use,

eggCer/eggSM mixtures (1:0 mol, 9:1 mol, 2:1 mol, 1:1 mol, 1:2 mol, and

0:1 mol) were dissolved in chloroform (Sigma-Aldrich, St. Louis, MO) to

achieve a final concentration of 0.5 mg/mL. The stock solutions were kept

at –20�C. Experiments on Langmuir monolayers were performed on a water

subphase. High-purity water was produced by a Milli-Q source system (con-

ductivity<18 MU cm; g ¼ 72.6 mN/m at 20�C; Millipore, Billerica, MA).

Methods

The equilibrium characterization of the Langmuir monolayers was per-

formed by a combination of surface pressure (p-A) isotherms with Brewster

angle microscopy (BAM). The rheological study was performed upon

oscillatory compression (36) and shear (37–39). Surface states were

characterized by a surface pressure p0 and a packing area AB0B. All exper-

iments were performed at room temperature (255 1�C). For the oscillatorybarrier experiments, we used a conventional Langmuir balance (NIMA,

Coventry, UK). The shear rheometer (Physica MCR301; Anton Paar, Ost-

fildern, Germany) was equipped with a biconical bob tool working in the

oscillatory mode (see the Supporting Material for details). Both techniques

take advantage of the general concept of mechanical causality relating the

stress response (s) and strain (u),

sðtÞ ¼ AuðtÞ; (1)

where A ¼ A0 þ iA00 is the complex viscoelastic function. In the linear

regime, for oscillatory strain (u ¼ u0 eiut), one measures the force stressed

by the monolayer s¼ s0 ei(ut�f) after an applied strain. Normally, a delay f

appears as a consequence of viscous friction. The elastic modulus is repre-

sented by the real part of the viscoelastic function, which is calculated as

Biophysical Journal 101(11) 2721–2730

A0 ¼ s0

u0cosf: (2)

Viscous losses are represented by A00, the imaginary part of the visco-

elastic modulus that, for oscillatory motion, is related to the respective

(compression and shear) viscosities as

h ¼ A00

u¼ s0

uu0sinf: (3)

Given the uncertainty in reproducing identical surface states, the final

viscoelastic parameters are affected by a typical experimental error ~dA0,dA00 z 51 mN/m, larger than the instrumental resolution of the surface

stresses (ds z 50.01 mN/m) (a detailed discussion is provided in the

Supporting Material).

The transport properties were obtained from the lateral mobilities

measured by particle tracking (40,41). We used micron-sized PMMA

spherical particles as an interfacial probe of motion (5.75-mm diameter;

Microparticles, Berlin, Germany). The particles were spread on the lipid

monolayers previously formed in a homebuilt trough installed on the micro-

scope stage (see the Supporting Material for preparation details). Brownian

trajectories were tracked with an inverted microscope (Eclipse TE2000,

bright field, 20�; Nikon, Melville, NY) and a cooled charge-coupled device

camera (DS-1QM, 1Mpx; Nikon) (a detailed description of the tracking

method is supplied as the Supporting Material). The trajectories must

follow the Einstein diffusion equation; in this method, the mean-square

displacements were (41)

MSDðtÞ ¼ 8Dt; (4)

where D is a bare diffusion coefficient, which is calculated as the slope of

the short-time diffusive part of the mean-square displacement (MSD)

trajectory (t < 1 s, typically). Finally, D was calculated as a statistical

average with its corresponding standard deviation.

RESULTS

Equilibrium

Fig. 1A shows the experimental pressure-area (p-A) isothermof eggCer measured at room temperature (25 5 1�C). Formolecular areas above 75 A2, the p-A isotherm remained ata near-zero pressure that is compatible with a diluted state.Upon compression below 75 A2, the lateral pressure in-creased up to a pseudoplateau state at a pressure ~5 mN/m.Then, at an area close to 50 A2, the isotherm suddenly rose,eventually entering a collapse regime at Az 40 A2, charac-terized by a constant pressure (~49 mN/m). Generally, theisotherm displays a condensedlike shape characterized bya high slope, typical of solid phases (14). This isotherm isqualitatively similar to that recently obtained by Busto et al.(13) for synthetic C16-ceramide.

To obtain more information about the structure of eggCermonolayers, BAM images were simultaneously capturedalong a compression cycle. Fig. 1, in images a–h, showsthe evolution of monolayer textures for progressive pressurestates. In the diluted state (p < 5 mN/m), the system wasfound to be heterogeneous: gaseouslike regions occurredas apparently uncovered areas (dark) between ceramideplatforms (bright). At this temperature, lipid patches showed

Page 3: Solid Character of Membrane Ceramides: A Surface Rheology Study of Their Mixtures with Sphingomyelin

FIGURE 1 (A) Surface pressure p-A isotherm of eggCer monolayer at

compression rate of 5 cm2/min. (Inset) Elasticity modulus E of eggCer rep-

resented in function of the mean molecular area. (B) Elasticity modulus E of

eggCer represented in function of the surface pressure. E is obtained from

the experimental compression isotherm as the direct derivative of p(A),

using Eq. 5. (Images a–h) BAM micrographs of eggCer monolayers (scale

bar corresponds to 100 mm). Images represent monolayer states raised along

a continuous compression isotherm at 20�C. (a) 0.3 mN/m, (b) 0.6 mN/m,

(c) 2 mN/m, (d) 5 mN/m, and (e–h) image sequence showing the surface

topography of eggCer monolayers at the collapse upon further compression.

Brighter regions correspond to thicker surface ruffles emerging from the

ceramide monolayer.

Mechanics of Ceramide-Based Monolayers 2723

no flat edges, suggesting a molecular arrangement typical ofamorphous solids (glassylike). Upon further compression,the gaseous regions became progressively covered by theceramide monolayer in such a way that the irregular-shapedholes evolved into smaller circular areas and the solid frag-ments merged into larger patches (Fig. 1, image c). Above5 mN/m, the monolayer became homogenous, entering acontinuous solid phase, characteristic of the bilayer packingstate (pbil; see Fig. 1 d). At collapse (p R 45mN/m), themonolayer thickened into a ruffled structure (Fig. 1, imagese–h). Upon further compression, thicker regions appeared assurface ruffles progressively thickening into multilayerfringes separated by monolayer furrows (Fig. 1, imagese–h). These textures, typical of fracturing solids upon lateralcompression, are compatible with buckling instability (42).

On the other side, sphingomyelin is well known to havea fluidlike phase behavior in Langmuir monolayers (43).Different to ceramide, SM monolayers displayed a coexis-tence plateau (pcoex z 15–25 mN/m, depending on the

chain length) between the expanded low-pressure phase(LE) and the condensed liquid phase (LC), typical of thebilayer packing state (14). At higher pressures, SM mole-cules undergo a packing transition into a disordered solidstate (44). The monolayer of eggSM showed a broad LE/LC coexistence plateau at pcoex

(SM) z 20 mN/m (seeFig. S3 in the Supporting Material). Above 20 mN/m upto collapse, BAM images revealed a homogenous LC statefor the single SM monolayer. For the Cer/SM binary mix-tures, the p-A isotherms were observed to expand to higherareas and smaller slopes with increasing SM contents(Fig. S3), similar to the Cer/SM system studied by Bustoet al. (13). At the relevant packing (pbil z 30 mN/m), themonolayers were invariably observed at the homogeneousLC-state, although irregularly edged platforms were seenat low pressures for the ceramide-rich states (Fig. S3). Wethus confirm uniform LC-monophasic behavior over thewhole compositional range at the biologically relevant state.

Compression viscoelasticity

The equilibrium compression modulus, E0, is defined as thechange in monolayer pressure caused by an infinitesimalchange in the molecular area. It can be easily calculatedfrom the numerical derivative of the experimental p-Aisotherms as

E0 ¼ �Avp

vA: (5)

In the diluted regime (A> 50 A2, p< 5 mN/m), the egg-Cermonolayer was highly compressible (E0< 50 mN/m). Whencompressing below 50 A2, the equilibrium modulusincreased up to a maximum value (E0z 450mN/m) reachedat a surface area of ~47 A2 (Fig. 1 A, inset). A detailed anal-ysis showed that the maximal rigidity appeared at pressuresranging from 25 to 35 mN/m, corresponding to the biologi-cally relevant surface packing (33). As expected, furthercompression reduced E0 down to smaller values compatiblewith the collapsed arrangement (Fig. 1 B). Fig. 2 shows thecompression modulus for EggCer/EggSM mixtures calcu-lated as Eq. 5 from the compression p-A isotherms inFig. S3. As expected, the compression modulus drasticallyincreased with increasing eggCer (see Fig. 2, inset), a factintrinsically related to the compaction effect induced byhydrogen-bonding with ceramide molecules (13).

The dynamic response against oscillatory compressionwas measured at a constant frequency (u0 ¼ 2p/T ¼0.29 s�1). We explored strain amplitudes ranging from0.3% up to 6% of the initial equilibrium area. Fig. 3 showsthe strain-stress plots measured for eggCer and eggSM at thesurface state pbil ¼ 30 mN/m, well within the homogeneousLC phase of both lipids (see Fig. S1 for a typical time trace).For eggSM, the experimental time traces were fitted tothe linear Eqs. S1 and S2 in the Supporting Material,

Biophysical Journal 101(11) 2721–2730

Page 4: Solid Character of Membrane Ceramides: A Surface Rheology Study of Their Mixtures with Sphingomyelin

FIGURE 2 p-Dependence of the elasticity modulus, E0, for lipid mono-

layers made of EggCer/EggSM mixtures: (solid line) 1:0 mol; (dashed

line) 2:1 mol; (dash-dotted line) 1:1 mol; (dotted line) 1:2 mol; and (dash-

dot-dot-dashed line) 0:1 mol; and that obtained from the p-A isotherms

(Eq. 5). (Inset) % SM-dependence of the maximal elasticity modulus,

E0max, measured at the biologically relevant lateral pressurepbilz 30mN/m.

2724 Catapano et al.

independent of the applied strain. As expected, only thefundamental peak at u0, corresponding to the linear re-sponse, was found in the power spectrum obtained byFourier transforming the experimental p(t) curves (Fig. 3,inset). In this fluid case, a near-linear response typical ofa fluid system was measured over the considered deforma-tion span. Nonlinear effects were clearly visible for the solid

FIGURE 3 Experimental stress-strain curves obtained for (,) eggCer;

(6) eggCer/egg/SM 1:1 mol; and (>) eggSM monolayers. (Solid lines)

Asymptotic linear behavior s0 ¼ E0u0, well characterized by the linear

compression modulus E0 reported in Fig. 2. (Inset) Fast Fourier transform

stress functions of eggCer (black solid line) and eggSM (gray dashed

line) measured at ~4% strain amplitude. The spectrum for eggCer is char-

acterized by the presence of a series of harmonics (ku0) of the fundamental

mode u0 (excitation frequency). By contrast, only the fundamental mode is

found for eggSM monolayers.

Biophysical Journal 101(11) 2721–2730

cases. Particularly, for eggCer monolayers, the experimentaltime-traces could be fitted to the linear Eqs. S1 and S2 inthe Supporting Material only for strains smaller than 1%.However, a number of harmonics are necessary to fit stressresponses at higher strains (see Eq. S5 in the SupportingMaterial). Furthermore, the Fourier transform shows inthis case a typical nonlinear pattern where the smaller peakscorrespond to higher harmonics of the fundamental excita-tion mode u0 (uk ¼ ku0, k ¼ 2, 3, .) (Fig. 3; inset)(45,46). In this case, above a critical deformation thresholduC, the stress-strain curve showed a softening deviationcharacterized by smaller stresses than expected for linearHookean behavior s ¼ E0u0 (E0 taken from Fig. 2).

Oscillatory compression in the linear regime allows for anadequate evaluation of time-effects. The strain amplitudewas set in all cases at 0.5% of the initial equilibrium area,well within the linear regime. Fig. 4 A shows the values ofthe dynamic compression elasticity measured as a functionof the frequency of the deformation u0 for the different sys-tems at pbil ¼ 30 mN/m. For eggSM, the dynamic modulusE(u) was constant, independent of the frequency, takingvalues compatible with the equilibrium compression mod-ulus E0. Similar qualitative behavior was observed for theSM-rich monolayers, up to an eggSM content ~50%. Quitedifferent were the frequency-dependences found for ceram-ide-rich monolayers, where the compression elasticity dis-played a significant decrease with increasing frequency.

FIGURE 4 (A) Frequency dependence of the compression elasticity

modulus E (solid symbols). Equilibrium value E0 obtained from the p-A

isotherms (dashed lines) (Fig. 2). (B) Frequency dependence of the

compression viscosity hE (open symbols) as obtained from oscillatory

barrier experiments for different eggCer/eggSM monolayers: (-) 1:0 mol;

(C) 2:1 mol; (:) 1:1 mol; (;) 1:2 mol; and (A) 0:1 mol.

Page 5: Solid Character of Membrane Ceramides: A Surface Rheology Study of Their Mixtures with Sphingomyelin

FIGURE 5 (A) Experimental strain-stress plots under shear deformation

measured at fixed frequency of 1 Hz for different eggCer/eggSM mono-

layers: (-, solid line) 1:0 mol; (:, dashed line) 1:1 mol; and (A, dots)

0:1 mol. The yield stress sY defines a plastic plateau at gY. (Dashed line)

Boundary of the linear regime from the nonlinear regime. (B) Strain depen-

dence of the shear modulus G0 (solid symbols) and (C) the frictional shear

losses G00 (open symbols) as obtained from oscillatory barrier experiments

Mechanics of Ceramide-Based Monolayers 2725

This behavior demonstrated softening plasticity at highfrequencies, suggesting the presence of yielding effects athigh compression rates (7).

The compression viscosity hE was evaluated using Eq. S4in the Supporting Material (the particular expression ofEq. 3). Fig. 4 B plots the values of hE as a function of u0

for the different systems. As expected in compression,ceramide-rich monolayers (i.e., solidlike) exhibited muchhigher viscosities than the SM-ones (i.e., fluidlike). Withthe exception of the 1Cer:2SM mixture, the binary mono-layers were characterized by intermediate viscosities. Thisapparent anomaly (high viscosity) could be related to the1:2 maximal compaction due to the molecular complexationpreviously hypothesized by Busto et al. (13) at this particularmolar composition. Regarding the frequency dependence,compression viscosities showed a linear power-law depen-dence (hE ~ u�1), typical of a Voigt-like element withconstant viscous losses (Eloss ¼ uhE z constant). This ischaracteristic of solid bodies deforming with a significanthigher elastic storage than the associated viscous losses(E0[Eloss¼uhE). Like the ceramide monolayer, the mixedones exhibited a similar dynamic behavior, indicating thedense character of these compacted systems with a solidlikebehavior under compression (hE ~ u�1), regardless of theirchemical nature.

for the same eggCer/eggSM mixtures shown in panel A.

Shear viscoelasticity

The solid character of ceramide-rich monolayers is betterobserved in mechanical experiments probing a finite shearresponse (G0 > 0). Adequate account requires previous iden-tification of the linear regime. Fig. 5A shows the stress-straincurves determined in shearing the monolayers at the biolog-ically relevant state (pbil¼ 30mN/m) at a constant frequency(1 Hz). For eggSM monolayers, viscous flow was character-ized by a near-linear stress response curve, s ¼ sV ¼ G00g(Newtonian flow) and sEz 0, corresponding to a zero shearmodulus, G0 z 0 (no shear resistance) (see Fig. 5 B).Conversely, the stress response of eggCer contained bothelastic and flow components (s ¼ sE þ sV), with higherelastic storage than viscous losses (G0 > G00; see Fig. 5, Band C). In this case, the response curve exhibited a plasticplateau characterized by a relatively low yield stress (sY z0.8 mN/m; see Fig. 5 A). This defines the onset of plasticityat shear strains >1% for eggCer, which is thus identified asa soft solid. The frequency dependence of the shear responsewas probed at low strains (g¼ 0.5%), which were chosen tobe lower than the yielding limit (gY z 1%; see Fig. 5 A),therefore well within the linear regime.

Fig. 6 A plots the frequency dependence of the experi-mental values of the shear parameters. For eggCer, the valueof the shear modulus is found high (G0 z 75 mN/m) andsystematically larger than the loss modulus (G0 >> G00),as expected for a solid material able to resist lateral shear(21). On the other hand, no measurable shear rigidity was

detected for eggSM monolayers (G0 z 0), again an irrefut-able proof of fluidity at the considered packing state (pbil ¼30 mN/m). The shear viscosity was found in this case to beseveral orders-of-magnitude lower than that for the eggCermonolayer (see Fig. 6 B), indicating the pasty character ofceramide with respect to its fluid precursor sphingomyelin.Indeed, eggSM was so fluid with respect to eggCer that, atlow frequencies, the shear viscosity fell below the experi-mental resolution of the bicone rheometer (see eggSMat u < 1 s�1 in Fig. 6 B). Likewise, in compression, thefrequency dependence of the shear viscosity of eggCer andtheir mixtures with SM also displayed a linear powerlaw, ~u�1, again a signature of the constant losses presentin these dense monolayers. However, a distinct behaviorwas shown by the fluid eggSM monolayer, with a very low,nearly frequency-independent viscosity (hS ~ u

0), character-istic of Newtonian flow. The binary monolayers exhibitedshear parameters intermediate between the single-lipidmonolayers (Fig. 6 B) with an inversion of the absolutevalues at a composition ~2:1 (Cer/SM) (Fig. 6 C). Thus,the Cer-rich states (xCer > 66%) can be classified as ‘‘pasty’’because they display a high shear rigidity and a high, butfinite, viscosity (G0 > G00).

However, because solids are able to resist shear, theseceramide-rich monolayers are also able to develop flowunder lateral stress. Obviously, fluid sphingomyelin causesdisorder at the monolayers, quantified as a decrease in the

Biophysical Journal 101(11) 2721–2730

Page 6: Solid Character of Membrane Ceramides: A Surface Rheology Study of Their Mixtures with Sphingomyelin

FIGURE 6 (A) Frequency dependence of the shear modulus G0 and (B)

the shear viscosity hS obtained from oscillatory shear experiments at a

fixed strain amplitude of 0.5% for different eggCer/eggSM monolayers:

(-) 1:0 mol; (=) 9:1 mol; (C) 2:1 mol; (:) 1:1 mol; (;) 1:2 mol;

and (A) 0:1 mol. (C) Shear modulus G0 (-) and the frictional shear losses

G00(C) measured at a fixed frequency of 10 Hz and represented as a function

of the SM content present in the same eggCer/eggSM mixtures shown in

panels A and B. (D) Solid character (defined as a percentage of G0/G00 ratio)of the same eggCer/eggSM mixtures shown in panels A and B as a function

of the % SM content.

FIGURE 7 (A) Typical image showing several PMMA microparticles

(at low surface covering) spread on a pure eggSM monolayer. (B) Mean-

square displacements (MSDs) for PMMA particles of 5.75-mm diameter

on different eggCer/eggSM monolayers: (,) 1:0 mol; (B) 2:1 mol; (6)

1:2 mol; and (7) 0:1 mol. (Inset) Diffusion coefficient obtained from the

linear fit of MSDs as a function of the SM content present in the eggCer/

eggSM mixtures.

2726 Catapano et al.

shear modulus (i.e., softening) followed by a discrete de-crease in viscosity (i.e., thinning). The increase of disorderis such that at sphingomyelin contents >33%, the mono-layer undergoes a mechanical inversion into higher viscouslosses than elastic storage. In this compositional regime(xCer < 66%), the monolayers become a viscoelastic bodywith flow properties dominant over elasticity (G0 < G00).At optimal binary compaction (1:2 Cer/SM) (13), the mixedsystem still preserves certain mechanically rigidity (G0 z3 mN/m). However, at higher SM contents (xCer < 33%),the systems might be certainly classified as viscous fluids(G00 > G0 z 0), with flow properties similar to fluid SM.Using the experimental values of G0 and G00, we constructeda relative scale of hardness defined as a percentG0/G00 ratio. Inthis scale, eggCer was assigned 100% hardness in contrast to0% for pure eggSM (G0/G00 ¼ 0). The way the monolayerhardness varies with SM content is plotted in Fig. 6 D. Ineffect, introducing eggSM made the ceramide monolayerprogressively softer in such a way that only 10% SM pro-duces ~50% softening with respect to pure ceramide. HigherSM contents induced progressive softening on ceramide

Biophysical Journal 101(11) 2721–2730

monolayers, which became practically fluid at the physio-logic composition 1:2 Cer/SM and beyond (G0/G00 % 20%).

Lateral mobility by particle tracking

Embedded membrane objects (such as transmembrane pro-teins or hydrophobic particles) undergo Brownian motionunder the action of thermal energy. Obviously, a high fric-tion imposed by a viscous membrane introduces high dragrestrictions to particle motion. Consequently, lateral mobil-ities might be correlated with the flow behavior describedabove in such a way that lateral motion in membranescontaining ceramides might be largely slowed down withrespect to motion in a fluid SM-based environment. Toquantitatively address this conjecture, we tracked the lateraldiffusion of microparticles embedded within the same lipidmonolayers explored above. We used noncharged PMMAmicroparticles, a hydrophobic probe of motion interactingwith the lipid layer under sticking conditions. Particledensity was chosen quite low, well below the packing tran-sition, so that interparticle interactions could be reasonablyneglected.

Fig. 7 A shows a typical microscopy field containing afew PMMA microparticles adsorbed into a lipid monolayer.

Page 7: Solid Character of Membrane Ceramides: A Surface Rheology Study of Their Mixtures with Sphingomyelin

Mechanics of Ceramide-Based Monolayers 2727

Microparticle diffusion coefficients were measured indifferent lipid monolayers based on egg ceramide and itsmixtures with sphingomyelin. Under these conditions, theparticles typically undergo Brownian motion in two dimen-sions, the mean-square displacements (MSDs) closelyfollowing Eq. 4 (see the Supporting Material and Eqs. S8and S9 therein). Fig. 7 B shows the time evolution ofMSDs. Each MSD trajectory represents an average over~200 particles at different places in a sample and in differentsamples (see Methods for details). At a given time, MSDsincreased with the content of eggSM, as expected forincreasing fluidity.

The analysis of MSDs in terms of Eq. 4 provides effectivevalues for the particle diffusion coefficient. From the linearfits to experimental MSD curves, absolute values of D(SM) ¼0.68 5 0.04 mm2/s and D(cer) ¼ 0.15 5 0.05 mm2/s wereobtained for pure eggSM and eggCer, respectively (Fig. 7,inset). These values are in agreement with previous litera-ture data obtained in similar systems (40). Intermediatemobilities were observed for the binary mixtures (Fig. 7,inset). Remarkably, lateral diffusivity is reduced in solideggCer monolayers by ~78% with respect to fluid eggSM.In general, the data in Fig. 7 show particle diffusivitiesdecreasing with increasing ceramide content—a trend com-patible with the viscosity increase reported above.

DISCUSSION

From our surface rheological study, we show unequivocalpieces of evidence about the solid character of egg-ceramidemonolayers, a behavior radically different from most lipidsforming biological membranes. Important differences areparticularly found in the following properties.

Compression elasticity

For egg ceramide at the biologically relevant state (pbil z30 mN/m), the equilibrium modulus was abnormally high(E0 z 450 mN/m), compared to typical fluid lipids(z100 mN/m), e.g., POPC and Escherichia coli lipids,representative of eukaryote and prokaryote membranes,respectively (46). Such a high compression rigidity is evenmuch bigger than that found for saturated sphingolipids(E0 z 100 mN/m for C16-SM) (13) and phospholipids(E0 z 300 mN/m for DPPC) (47) in the gel phase, andliquid-ordered monolayers compacted by high cholesterolcontents (E0 z 400 mN/m for 1SM/1Chol mixtures) (48).This fact underlies the idea of ceramide monolayers asvery rigid solid assemblies, stiffer than other solidlike lipidsystems. Regarding dynamics, ceramide monolayers alsodiffer significantly with respect to fluid systems upon lateralcompression (see Fig. 4 A).

The dynamic softening observed for ceramide-rich mono-layers, E(u) R E0, indicates smaller instantaneous rigiditythan under quasistatic compression conditions, a pseudo-

plastic behavior typical of soft solids that flow better athigh rates. Similar to Bingham plastics unable to flow atlow rates, such solid materials increasingly resist compres-sion at low frequencies, but undergo yielding flow whendeformed at higher rates (49). Such a yield plasticity wasindeed revealed in ceramide monolayers as a stress soft-ening observed upon faster compression rates at increasingdeformation (see Fig. 3). For SM-rich monolayers, however,the dynamic modulus was found essentially constant inde-pendent of the compression frequency—a mechanical char-acteristic typical of Newtonian fluids.

Shear elasticity

Solid membranes exhibit structural rigidity and so resistshear deformations, which are characterized by a finite shearmodulus. Such a solidlike behavior was observed for mono-layers containing eggCer (see Fig. 6), a higher relativecontent of ceramide defining a stiffer solid character ofthe membrane (see Fig. 6, C and D). Particularly, at the bio-logically relevant state (pbil), pure ceramide monolayerswere characterized by a shear modulus as high as G0 z80–100 mN/m, a value three-orders-of-magnitude higherthan that measured for DPPC monolayers at gel state (8).Although increasing SM contents decreased the solid char-acter, G0 remained finite up to a relatively high SM content(G0 > 0 for xCer < 66%), indicating the high capacity ofceramide to condense SM molecules and impart solid orderto SM-based membranes. On the other hand, single SMmonolayers were characterized by a zero shear modulus,an unequivocal evidence for fluidity.

Surface viscosity

Flow properties were remarkably different for ceramidemonolayers compared to typical fluid systems. At the biolog-ically relevant state (pbil), the surface viscosities of solid ce-ramidemonolayers had values as high as hEz hSz 200mNs/m (at a reference frequency,uz 0.1 s�1; Figs. 4 B and 6 B,respectively). These values were comparatively higher thanthe moderate viscosities measured for gel-like DPPC(hE z 100 mN s/m (47) and hS z 1 mN s/m (8)) and therelatively low ones measured for fluid monolayers (hE z50 mN s/m (50), hS z 0.3 mN s/m (8) for POPC, andhE z 20 mN s/m (51), hS <<0.1 mN s/m (8) for E. colilipids; at u z 0.1 s�1). The monolayers of pure SM indeedshowed a high fluidity, characterized by surface viscosities aslow as hE z 30 mN s/m (Fig. 4 B) and hS z 0.1 mN s/m(Fig. 6 B) (at u z 0.1 s�1). Mixing ceramide with SMelicited a progressive increase in fluidity due to SM, whichoccurred in parallel to the decrease in solid character im-posed by ceramide. The inversion between predominantlysolid- and a fluidlike behavior occurred close to a 2:1Cer/SM compositional ratio. Higher SM contents were char-acterized by a relatively low solid character (<<50%, as

Biophysical Journal 101(11) 2721–2730

Page 8: Solid Character of Membrane Ceramides: A Surface Rheology Study of Their Mixtures with Sphingomyelin

2728 Catapano et al.

defined in Fig. 6 C) and a high fluidity, typical of SM. At thephysiological composition 1:2 Cer/SM, the system behavedas essentially fluid (G00 >G0), although with a shear viscosity2–3 times higher than pure SM and a very low, but still finite,shear rigidity (see Fig. 6). Consequently, the ceramide plat-forms existing in vivo should be expected to be much moresolid and viscous than typical fluid membrane phases.

Lateral mobility versus solid character

On the basis of the above results, ceramide-rich membraneenvironments should mechanically behave as weakly rigid(soft) solids and exhibit less fluidity than typical fluidmembrane sites formed by sphingomyelin, phospholipids,and their mixtures with cholesterol. This hypothesis hasbeen previously stated in the literature (8,52), but the pre-sent results confirm the picture. Therefore, if the enzymaticproduction of ceramide is triggered at a given membranesite, a decrease in fluidity followed by a partial solidificationis expected to localize at that emplacement. Consequently,protein diffusion would be extremely slowed down insidethese ceramide-rich domains. In an attempt to understandthe constitutive relationships between mechanical and trans-port properties, how ceramide affects lateral mobilities andhow it could eventually shape their physiological function,we will discuss the correlation between the diffusionalmobilities with the solid character imposed by ceramide.The conclusions will be later put into a biological context.In Fig. 8, we correlated the relative lateral diffusivitiesobtained from the particle tracking experiments with theprimitive transport property expressed as a shear viscosity.As expected, there was an inverse correlation betweenlateral diffusivity and viscosity. In other words, membranefluidity and lateral diffusivity were mutually proportional,D ~ hS

�1, as predicted by the Stokes-Einstein (SE) relation-

FIGURE 8 Diffusion coefficient D relative to the diffusion coefficient of

SM DSM as a function of surface shear viscosity hS relative to the shear

surface viscosity of SM hSSM as obtained from the shear oscillatory ex-

periments for different eggCer/eggSM mixtures. (Vertical dashed line)

Two different viscoelastic behaviors characterized by G0 < G00 (left) andG0 > G00 (right). (Solid line) Stokes-Einstein dependence. (Dashed line)

Saffmann-Delbruck prediction (see text for details).

Biophysical Journal 101(11) 2721–2730

ship for the diffusional motion of a spherical object withradius R in a bulk fluid of shear viscosity h:

DSE ¼ kBT

6phR: (6)

This SE expression has been successfully exploited todetermine apparent microviscosities from experimentalmobilities. However, using Eq. 6 to analyze membranediffusivities requires turning the bulk viscosity h into anequivalent surface value. This correspondence is usuallyperformed by transforming the surface shear viscosity (hS)of a membrane of thickness h into an equivalent bulkvalue using the approximate dimensional correspondence,h ~ h hS. However, no exact quantitative correspondenceis provided by this relation, thus no absolute values ofsurface viscosity can be calculated using Eq. 6 from particletracking experiments. Nevertheless, the surface-modifiedSE relationship can be successfully exploited to calculaterelative viscosities from relative diffusivities as

D

Dð0Þ ¼

hS

hð0ÞS

!�1

: (7)

In this work, Eq. 7 was used to test the inverse correlation

between lateral diffusivities and shear viscosities, indepen-dently measured in tracking and rheology experiments,respectively. For the fluid monolayers (xCer < 33%), theagreement between theory and experiment was quantitative.However, slower mobilities were predicted by Eq. 7 for thesolidlike systems (xCer > 33%). Indeed, Saffman andDelbruck showed that a particle trapped in a viscous mem-brane also feels drag from the adjacent bulk phases, itsmobility enhanced by a logarithmic factor depending onthe ratio of the surface/bulk viscosities,

DSD ¼ kBT

6phShlog

�hSh

2h� 0:577

�: (8)

Predictions from the Saffman and Delbruck model are also

plotted in Fig. 8.As expected, Eq. 8 predicts highermobilitiesthan effective SE-friction in Eqs. 6 and 7. At high SMcontents (xCer < 33%), the monolayers are so fluid that thelogarithmic correction is not quite relevant. However, athigher viscosities, Saffman and Delbruck’s model predictshigher mobilities, although still slower than those experi-mentally measured in tracking experiments. Lateral mobil-ities might be just a matter of viscous flow in SM-richfluid membranes; however, the viscoelastic character intro-duced by ceramide might introduce an additional elasticcomponent to particle impulse at xCer> 33%. Consequently,dominant membrane elasticity in ceramide-rich membranesmight impart higher mobility than expected from the usualSaffman-Delbruck approach (see Fig. 8; G0 < G00), whichjust accounts for frictional control in a fluid environment
Page 9: Solid Character of Membrane Ceramides: A Surface Rheology Study of Their Mixtures with Sphingomyelin

Mechanics of Ceramide-Based Monolayers 2729

(G00 [ G0). Overall, despite these details, data in Fig. 8depict ceramide as a pasty agent able to impart profoundchanges on membrane dynamics, particularly a largedecrease in the lateral mobilities of the objects embeddedwithin.

Biological implications

From this study, the pasty nature assigned to ceramidemembranes may involve strong physiological consequences.The large increase in viscosity and the solid character,together with the decrease in molecular mobility imposedby ceramide, could be relevant in apoptosis. Ceramides aregenerated in the cell in response to different stresses thatcan harm the cell (even to activate a programmed cell deathcascade, i.e., apoptosis). Recently, Cremesti et al. (6) sug-gested an immobilization role for ceramide-rich domainsformed in the plasma membrane during apoptosis. Such aprocess is conceived as an accretion of immobilized proteinsthat triggers the signaling events before apoptosis.

From our results, the immobilization capacity of ceramideplatforms can be easily conceived, with the ceramide solid-like domains working as sequestrating elements for sig-naling proteins. Protein diffusion within a ceramide-richdomain would be extremely slowed down and the signalingproteins would be frozen (capping) within the domains. Apossible role for ceramidewould be related to the entrapmentof these proteins to amplify the signal response. The partitionof some membrane proteins into ceramide domains hasrecently been reported (53). In that study, although none ofthe studied proteins is involved in signaling processes, theresults suggested a functional role of ceramide platforms inconnection with their singular mechanical properties (53).Therefore, we conclude that ceramidesmight have amechan-ical role in membrane physiology.

SUPPORTING MATERIAL

Additional methods and supporting equations, three figures, and one

scheme are available at http://www.biophysj.org/biophysj/supplemental/

S0006-3495(11)01315-4.

We thank the ‘‘CAI de Infrarrojo-Raman-Correlador’’ facility at the Univer-

sidad Complutense for Brewster-angle microscopy time.

This work was supported under grants FIS2009-14650-C02-01, the Consol-

ider-Ingenio en ‘‘Nanociencia Molecular’’ CSD2007-0010, and

S2009MAT-1507 (NOBIMAT) from the Comunidad autonoma de Madrid.

E.R.C. was supported by the FPU program at the Ministerio de Educacion.

I.L.M. thanks the ‘‘Juan de la Cierva’’ program (Ministerio de Ciencia e In-

novacion) F.M. thanks Universidad Complutense and Triangle de la

Physique for financial support during a sabbatical stay at Laboratoire de

Physique des Solides. G.E. was supported by CONACYT.

REFERENCES

1. Goni, F. M., and A. Alonso. 2006. Biophysics of sphingolipids. I.Membrane properties of sphingosine, ceramides and other simplesphingolipids. Biochim. Biophys. Acta. 1758:1902–1921.

2. Pascher, I. 1976. Molecular arrangements in sphingolipids. Confor-mation and hydrogen bonding of ceramide and their implicationon membrane stability and permeability. Biochim. Biophys. Acta.455:433–451.

3. Staneva, G., A. Momchilova, ., K. Koumanov. 2009. Membranemicrodomains: role of ceramides in the maintenance of their structureand functions. Biochim. Biophys. Acta. 1788:666–675.

4. Grassme, H., A. Jekle,., E. Gulbins. 2001. CD95 signaling via ceram-ide-rich membrane rafts. J. Biol. Chem. 276:20589–20596.

5. Grassme, H., V. Jendrossek,., E. Gulbins. 2003. Host defense againstPseudomonas aeruginosa requires ceramide-rich membrane rafts. Nat.Med. 9:322–330.

6. Cremesti, A. E., F. M. Goni, and R. Kolesnick. 2002. Role of sphingo-myelinase and ceramide in modulating rafts: do biophysical propertiesdetermine biologic outcome? FEBS Lett. 531:47–53.

7. Lopez-Montero, I., F. Monroy, ., P. F. Devaux. 2010. Ceramide:from lateral segregation to mechanical stress. Biochim. Biophys.Acta. 1798:1348–1356.

8. Espinosa, G., I. Lopez-Montero, ., D. Langevin. 2011. Shearrheology of lipid monolayers and insights on membrane fluidity.Proc. Natl. Acad. Sci. USA. 108:6008–6013.

9. Carrer, D. C., and B. Maggio. 1999. Phase behavior and molecularinteractions in mixtures of ceramide with dipalmitoylphosphatidylcho-line. J. Lipid Res. 40:1978–1989.

10. Carrer, D. C., and B. Maggio. 2001. Transduction to self-assembly ofmolecular geometry and local interactions in mixtures of ceramidesand ganglioside GM1. Biochim. Biophys. Acta. 1514:87–99.

11. Maggio, B. 2004. Favorable and unfavorable lateral interactions ofceramide, neutral glycosphingolipids and gangliosides in mixed mono-layers. Chem. Phys. Lipids. 132:209–224.

12. Holopainen, J. M., H. L. Brockman,., P. K. Kinnunen. 2001. Interfa-cial interactions of ceramide with dimyristoylphosphatidylcholine:impact of the N-acyl chain. Biophys. J. 80:765–775.

13. Busto, J. V., M. L. Fanani,., A. Alonso. 2009. Coexistence of immis-cible mixtures of palmitoylsphingomyelin and palmitoylceramide inmonolayers and bilayers. Biophys. J. 97:2717–2726.

14. Fanani, M. L., and B. Maggio. 2010. Phase state and surfacetopography of palmitoyl-ceramide monolayers. Chem. Phys. Lipids.163:594–600.

15. Boal, D. H. 2002. Mechanics of the Cell. Cambridge University Press,New York.

16. Singer, S. J., and G. L. Nicolson. 1972. The fluid mosaic model of thestructure of cell membranes. Science. 175:720–731.

17. Harkins, W. D., and R. J. Myers. 1937. Viscosity of monomolecularfilms. Nature. 140:465.

18. MacDonald, R. C., and S. A. Simon. 1987. Lipid monolayer states andtheir relationships to bilayers. Proc. Natl. Acad. Sci. USA. 84:4089–4093.

19. Mohwald, H. 1990. Phospholipid and phospholipid-protein monolayersat the air/water interface. Annu. Rev. Phys. Chem. 41:441–476.

20. McConnell, H. M. 1991. Structures and transitions in lipid monolayersat the air- water interface. Annu. Rev. Phys. Chem. 42:171–195.

21. Kaganer, V. M., H. Mohwald, and P. Dutta. 1999. Structure and phasetransitions in Langmuir monolayers. Rev. Mod. Phys. 71:779–819.

22. Gaines, H. E. 1966. Insoluble Monolayers at the Liquid-Gas Interface.Interscience, New York.

23. Kragel, J., G. Kretzschmar,., H. Mohwald. 1996. Surface rheology ofmonolayers. Thin Solid Films. 284–285:361–364.

24. Ding, J., H. E. Warriner, and J. A. Zasadzinski. 2002. Viscosity of two-dimensional suspensions. Phys. Rev. Lett. 88:168102.

25. Brooks, C. F., G. G. Fuller, ., C. R. Robertson. 1999. An interfacialstress rheometer to study rheological transitions in monolayers at theair-water interface. Langmuir. 15:2450–2459.

Biophysical Journal 101(11) 2721–2730

Page 10: Solid Character of Membrane Ceramides: A Surface Rheology Study of Their Mixtures with Sphingomyelin

2730 Catapano et al.

26. Wilke, N., F. Vega Mercado, and B. Maggio. 2010. Rheological prop-erties of a two phase lipid monolayer at the air/water interface: effect ofthe composition of the mixture. Langmuir. 26:11050–11059.

27. Sprong, H., P. van der Sluijs, and G. van Meer. 2001. How proteinsmove lipids and lipids move proteins. Nat. Rev. Mol. Cell Biol.2:504–513.

28. Thompson, T. E., M. B. Sankaram, ., W. L. Vaz. 1995. Effects ofdomain structure on in-plane reactions and interactions. Mol. Membr.Biol. 12:157–162.

29. Bhalla, U. S., and R. Iyengar. 1999. Emergent properties of networks ofbiological signaling pathways. Science. 283:381–387.

30. Rawicz, W., K. Olbrich,., E. Evans. 2000. Effect of chain length andunsaturation on lipid bilayer elasticity. Biophys. J. 79:328–339.

31. Marguet, D., P. F. Lenne, ., H. T. He. 2006. Dynamics in the plasmamembrane: how to combine fluidity and order. EMBO J. 25:3446–3457.

32. Fuller, G. G., and J. Vermant. 2011. Editorial. Dynamics and rheologyof complex fluid-fluid interfaces. Soft Matter. 7:7583–7585.

33. Marsh, D. 1996. Lateral pressure in membranes. Biochim. Biophys.Acta. 1286:183–223.

34. Choi, S. Q., S. Steltenkamp, ., T. M. Squires. 2011. Active micro-rheology and simultaneous visualization of sheared phospholipidmonolayers. Nat. Commun. 2:312.

35. Saxton, M. J., and K. Jacobson. 1997. Single-particle tracking: applica-tions to membrane dynamics. Annu. Rev. Biophys. Biomol. Struct.26:373–399.

36. Monroy, F., F. Ortega, and R. G. Rubio. 1998. Dilatational rheology ofinsoluble polymer monolayers: Poly(vinylacetate). Phys. Rev. E.58:7629–7641.

37. Erni, P., P. Fischer, ., J. Lauger. 2003. Stress- and strain-controlledmeasurements of interfacial shear viscosity and viscoelasticity atliquid/liquid and gas/liquid interfaces. Rev. Sci. Instrum. 74:4916–4924.

38. Kragel, J., S. R. Derkatch, and R. Miller. 2008. Interfacial shearrheology of protein-surfactant layers. Adv. Colloid Interface Sci.144:38–53.

39. Vandebril, S., A. Franck, ., J. Vermant. 2010. A double wall-ringgeometry for interfacial shear rheometry. Rheol. Acta. 49:131–144.

40. Sickert, M., and F. Rondelez. 2003. Shear viscosity of Langmuir mono-layers in the low-density limit. Phys. Rev. Lett. 90:126104.

Biophysical Journal 101(11) 2721–2730

41. Ortega, F., H. Ritacco, and R. G. Rubio. 2010. Interfacial microrheol-ogy: particle tracking and related techniques. Curr. Opin. Colloid Inter-face Sci. 15:237–245.

42. Stine, K. J., C. M. Knobler, and R. C. Desai. 1990. Buckling instabilityin monolayer network structures. Phys. Rev. Lett. 65:1004.

43. Li, X. M., J. M. Smaby, ., R. E. Brown. 2000. Sphingomyelin inter-facial behavior: the impact of changing acyl chain composition.Biophys. J. 78:1921–1931.

44. Vaknin, D., M. S. Kelley, and B. M. Ocko. 2001. Sphingomyelin at theair-water interface. J. Chem. Phys. 115:7697–7704.

45. Hilles, H., F. Monroy, ., R. G. Rubio. 2006. Fourier-transformrheology of polymer Langmuir monolayers: analysis of the non-linearand plastic behaviors. Adv. Colloid Interface Sci. 122:67–77.

46. Lopez-Montero, I., L. R. Arriaga, ., F. Monroy. 2010. Lipid domainsand mechanical plasticity of Escherichia coli lipid monolayers. Chem.Phys. Lipids. 163:56–63.

47. Arriaga, L. R., I. Lopez-Montero,., F. Monroy. 2010. Domain-growthkinetic origin of nonhorizontal phase coexistence plateaux in Langmuirmonolayers: compression rigidity of a raft-like lipid distribution.J. Phys. Chem. B. 114:4509–4520.

48. Hac-Wydro, K., and P. Dynarowicz-qatka. 2008. The impact of sterolstructure on the interactions with sphingomyelin in mixed Langmuirmonolayers. J. Phys. Chem. B. 112:11324–11332.

49. Bingham, E. C. 1922. Fluidity and Plasticity. McGraw-Hill, NewYork. 219.

50. Isanta, S., G. Espinosa, ., F. Monroy. 2011. Active membranes withbound F-actin: sliding vs. sticking conditions. Soft Matter. 7:3100–3107.

51. Lopez-Montero, I., L. R. Arriaga,., M. Velez. 2008. High fluidity andsoft elasticity of the inner membrane of Escherichia coli revealed bythe surface rheology of model Langmuir monolayers. Langmuir.24:4065–4076.

52. Wilke, N., and B. Maggio. 2009. The influence of domain crowding onthe lateral diffusion of ceramide-enriched domains in a sphingomyelinmonolayer. J. Phys. Chem. B. 113:12844–12851.

53. Chiantia, S., J. Ries, ., P. Schwille. 2008. Role of ceramide inmembrane protein organization investigated by combined AFM andFCS. Biochim. Biophys. Acta. 1778:1356–1364.