arXiv:1804.08596v1 [cond-mat.soft] 23 Apr 2018In symmetric dumbbells and snowman particles...

7
Geometric pinning and antimixing in scaffolded lipid vesicles Melissa Rinaldin, 1, 2 Piermarco Fonda, 1 Luca Giomi, 1, * and Daniela J. Kraft 2, 1 Instituut-Lorentz, Universiteit Leiden, P.O. Box 9506, 2300 RA Leiden, The Netherlands 2 Huygens-Kamerlingh Onnes Lab, Universiteit Leiden, P. O. Box 9504, 2300 RA Leiden, The Netherlands Unravelling the physical mechanisms behind the organisation of lipid domains is a central goal in cell biology and membrane biophysics. Previous studies on cells and model lipid bilayers featuring phase-separated domains found an intricate interplay between the membrane geometry and its chemical composition. However, the lack of a model system with simultaneous control over the membrane shape and conservation of its composition precluded a fundamental understanding of curvature-induced effects. Here, we present a new class of multicomponent vesicles supported by colloidal scaffolds of designed shape. We find that the domain composition adapts to the geometry, giving rise to a novel “antimixed” state. Theoretical modelling allowed us to link the pinning of domains by regions of high curvature to the material parameters of the membrane. Our results provide key insights into the phase separation of cellular membranes and on curved surfaces in general. Cells adapt the curvature of their membrane to create a plethora of functionalities from cargo-trafficking vesicles to cell division and motility. Dynamic remodelling of the membrane shape can be achieved by local changes in the lipid composition, the shape of the cytoskeleton, and the presence of curvature-generating proteins [1]. In turn, areas of high membrane curvature can recruit curvature- sensing proteins and lipids, which may further stabilise or increase local curvature differences [2]. Such a correlation between membrane geometry and chemical composition has also been extensively observed in model lipid membranes consisting of ternary mixtures of lipids and cholesterol. While being in a homogeneously mixed state at high temperatures, below a critical tem- perature, they separate into two different lipid phases. These phases, known as liquid-ordered (LO) and liquid- disordered (LD), differ in their lipid composition and pos- sess different mesoscopic material properties such as layer thickness and resistance to bending. In cells, phase sepa- ration is ubiquitous and exploited to concentrate certain molecules [3]. Similar to cellular membranes [4], exper- iments on giant unilamellar vesicles (GUVs) [58] and supported lipid bilayers (SLBs) [9, 10] reported a twofold correlation between these lipid domains and the mem- brane geometry: on the one hand, the local membrane curvature can favour the segregation of lipids and the nu- cleation and localization of domains. On the other hand, the presence of lipid domains can drive the formation of curved regions, such as buds, necks and protrusions. While these experiments have greatly contributed to our understanding of the physics and chemistry of lipid domains, they could not disentangle the correlation be- tween membrane geometry and chemical composition. In GUVs the geometry cannot be constrained while the overall composition is conserved, whereas SLBs posses a fixed membrane geometry but also allow ex- change of lipids with the environment. Furthermore, the change of the membrane curvature upon phase separation lead to repulsions between domains and non-equilibrium states. This prevented a fundamental understanding of curvature-related effects in phase-separating lipid mem- branes. Here we overcome these limitations by fabri- cating multicomponent scaffolded lipid vesicles (SLVs) which feature both prescribed shape and controllable composition. Our set-up, summarised in Fig. 1A, consists of a ternary mixture of the lipids sphingomyelin (SM), 1- palmitoyl-2-oleoyl-sn -glycero-3-phosphocholine (POPC) and cholesterol (chol) in a 2:1:1 mole ratio, deposited on micron-sized particles of four different shapes: spheres (Fig. 1B), cubes (Fig. 1C), symmetric dumbbells (Fig. 1D) and asymmetric dumbbells (Fig. 1E), also called snowman particles in the following. To preserve the fluidity of the lipid bilayer, we use col- loidal particles with a silica surface and 5% mole of 1,2-distearoyl-sn -glycero-3-phosphoethanolamine-N- [methoxy(polyethylene glycol)-2000] (DOPE-PEG 2000) (see inset of Fig. 1A). After lipid deposition, we lower the temperature to induce phase separation and image the SLVs using confocal microscopy. We identify the LO and LD phases through fluorescent labelling with N-[11-(dipyrrometheneboron difluoride)undecanoyl]- D-erythro -sphingosylphosphorylcholine (C11 TopFluor SM), represented in green, and 1,2-dioleoyl-sn-glycero-3- phosphoethanolamine-N-lissamine rhodamine B sulfonyl 18:1 (Liss Rhod PE), represented in magenta [11, 12]. Remarkably, both the likelihood of phase separation and the spatial arrangement of the phase domains are deeply affected by the geometry of the substrate. By comparing samples of 200 SLVs of each type we found that only 22% of the spheres showed phase separation. In contrast, for symmetric and asymmetric dumbbells, this fraction increased to 68% and 91%, respectively (SI: D.3). Since all experimental conditions were kept constant, we propose that the highly curved neck of the dumbbells promotes phase separation. This phenomenon is further arXiv:1804.08596v1 [cond-mat.soft] 23 Apr 2018

Transcript of arXiv:1804.08596v1 [cond-mat.soft] 23 Apr 2018In symmetric dumbbells and snowman particles...

Geometric pinning and antimixing in scaffolded lipid vesicles

Melissa Rinaldin,1, 2 Piermarco Fonda,1 Luca Giomi,1, ∗ and Daniela J. Kraft2, †

1Instituut-Lorentz, Universiteit Leiden, P.O. Box 9506, 2300 RA Leiden, The Netherlands2Huygens-Kamerlingh Onnes Lab, Universiteit Leiden,P. O. Box 9504, 2300 RA Leiden, The Netherlands

Unravelling the physical mechanisms behind the organisation of lipid domains is a central goal incell biology and membrane biophysics. Previous studies on cells and model lipid bilayers featuringphase-separated domains found an intricate interplay between the membrane geometry and itschemical composition. However, the lack of a model system with simultaneous control over themembrane shape and conservation of its composition precluded a fundamental understanding ofcurvature-induced effects. Here, we present a new class of multicomponent vesicles supported bycolloidal scaffolds of designed shape. We find that the domain composition adapts to the geometry,giving rise to a novel “antimixed” state. Theoretical modelling allowed us to link the pinning ofdomains by regions of high curvature to the material parameters of the membrane. Our resultsprovide key insights into the phase separation of cellular membranes and on curved surfaces ingeneral.

Cells adapt the curvature of their membrane to create aplethora of functionalities from cargo-trafficking vesiclesto cell division and motility. Dynamic remodelling of themembrane shape can be achieved by local changes in thelipid composition, the shape of the cytoskeleton, and thepresence of curvature-generating proteins [1]. In turn,areas of high membrane curvature can recruit curvature-sensing proteins and lipids, which may further stabiliseor increase local curvature differences [2].

Such a correlation between membrane geometry andchemical composition has also been extensively observedin model lipid membranes consisting of ternary mixturesof lipids and cholesterol. While being in a homogeneouslymixed state at high temperatures, below a critical tem-perature, they separate into two different lipid phases.These phases, known as liquid-ordered (LO) and liquid-disordered (LD), differ in their lipid composition and pos-sess different mesoscopic material properties such as layerthickness and resistance to bending. In cells, phase sepa-ration is ubiquitous and exploited to concentrate certainmolecules [3]. Similar to cellular membranes [4], exper-iments on giant unilamellar vesicles (GUVs) [5–8] andsupported lipid bilayers (SLBs) [9, 10] reported a twofoldcorrelation between these lipid domains and the mem-brane geometry: on the one hand, the local membranecurvature can favour the segregation of lipids and the nu-cleation and localization of domains. On the other hand,the presence of lipid domains can drive the formation ofcurved regions, such as buds, necks and protrusions.

While these experiments have greatly contributed toour understanding of the physics and chemistry of lipiddomains, they could not disentangle the correlation be-tween membrane geometry and chemical composition.In GUVs the geometry cannot be constrained whilethe overall composition is conserved, whereas SLBsposses a fixed membrane geometry but also allow ex-change of lipids with the environment. Furthermore, thechange of the membrane curvature upon phase separation

lead to repulsions between domains and non-equilibriumstates. This prevented a fundamental understanding ofcurvature-related effects in phase-separating lipid mem-branes. Here we overcome these limitations by fabri-cating multicomponent scaffolded lipid vesicles (SLVs)which feature both prescribed shape and controllablecomposition.

Our set-up, summarised in Fig. 1A, consists of aternary mixture of the lipids sphingomyelin (SM), 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine (POPC)and cholesterol (chol) in a 2:1:1 mole ratio, deposited onmicron-sized particles of four different shapes: spheres(Fig. 1B), cubes (Fig. 1C), symmetric dumbbells(Fig. 1D) and asymmetric dumbbells (Fig. 1E),also called snowman particles in the following. Topreserve the fluidity of the lipid bilayer, we use col-loidal particles with a silica surface and 5% mole of1,2-distearoyl-sn-glycero-3-phosphoethanolamine-N-[methoxy(polyethylene glycol)-2000] (DOPE-PEG 2000)(see inset of Fig. 1A). After lipid deposition, we lowerthe temperature to induce phase separation and imagethe SLVs using confocal microscopy. We identify theLO and LD phases through fluorescent labelling withN-[11-(dipyrrometheneboron difluoride)undecanoyl]-D-erythro-sphingosylphosphorylcholine (C11 TopFluorSM), represented in green, and 1,2-dioleoyl-sn-glycero-3-phosphoethanolamine-N-lissamine rhodamine B sulfonyl18:1 (Liss Rhod PE), represented in magenta [11, 12].

Remarkably, both the likelihood of phase separationand the spatial arrangement of the phase domains aredeeply affected by the geometry of the substrate. Bycomparing samples of 200 SLVs of each type we foundthat only 22% of the spheres showed phase separation. Incontrast, for symmetric and asymmetric dumbbells, thisfraction increased to 68% and 91%, respectively (SI: D.3).Since all experimental conditions were kept constant, wepropose that the highly curved neck of the dumbbellspromotes phase separation. This phenomenon is further

arX

iv:1

804.

0859

6v1

[co

nd-m

at.s

oft]

23

Apr

201

8

2

C

B

0 20 40 60 80 10010 30 50 70 90

0 20 40 60 80 10010 30 50 70 90

D

0 20 40 60 80 10010 30 50 70 90

0 20 40 60 80 10010 30 50 70 90

E

SUVs

A

Lowering T

XLD

PEG

Si

SLB

Calculated compositon Mixed state Phase-separated state

Antimixed state

Figure 1. Experimental phase diagram of SLVs. A Schematic representation of the experimental set-up: colloidal particlesof spherical, cubic, dumbbell and snowman shape are coated with a lipid membrane at T=70 ◦C. After the temperature islowered, the membrane undergoes phase separation. Liquid disordered (LD) and liquid ordered (LO) phases are representedin magenta and green, respectively. PEG molecules are drawn in yellow and cholesterol lipids in light blue. B Left: scanningelectron microscopy (SEM) image of a spherical colloid. Right: top views of 3D reconstruction of spherical SLVs, ordered fromleft to right according to increasing percentage of LD area fraction. C Left: SEM image of a cubic colloid. Right: equatorialsections of SLVs on cubic scaffolds. D Left: SEM image of a dumbbell particle. Right: top views of 3D reconstructions ofdumbbell-shaped SLVs. E Left: SEM image of a snowman colloid. Right: top views of 3D reconstructions of snowman-shapedSLVs. See Fig. S11 and Table 2 of the SI for respectively single channel images and area fraction values. Scale bars 2 µm.

enhanced in snowman particles, because of the differentdiameter, and hence curvature, of the two lobes. Al-though the chemical composition of SUVs is fixed priorto deposition, the actual SM:POPC:chol ratio on indi-vidual SLVs varies. This allows us to capture multiplepoints in the concentration phase space at once.

For successfully phase-separated SLVs, we observe astrong correlation between the curvature of the underly-ing scaffold and the area fractions, structure and locationof the phase domains. We denote by xLO and xLD therelative area fractions occupied by the two liquid phases,so that xLO + xLD = 1. We experimentally determinedthese area fractions by binarising the reconstructed pro-files into the LO and LD phase (Methods: Imaging).

Spherical scaffolds are uniformly curved, hence minimiza-tion of interface length between the LO and LD phasesis the principal driving force of domain dynamics, whichrelaxes towards the systematic formation of two domainswith relative area fractions set by the lipid composition(Fig. 1B). In contrast to previous experiments [13], thisobservation implies that our system has reached equilib-rium.

Geometric effects start to become visible on cubicSLVs, where the curvature is predominantly localised atthe corners and along the edges (see Fig. 1C). At low xLD

(magenta), we predominantly observed the LD domainsto be located at the edges, whereas at higher xLD wecould not identify a correlation with curvature. The re-

3

gions of higher curvature localize the softer LD domainsdue to the lower cost of bending and we refer to this ef-fect as geometric pinning. This leads to the growth ofmultiple LD domains and hinders coalescence, althoughline tension still plays an important role in merging sepa-rate domains, similarly to the case of spherical SLVs. Asa result, in a sample of 569 phase-separated cubes, 29%showed three or more domains (see Fig. SI S13).

In symmetric dumbbells and snowman particles geo-metric effects become more dramatic because of the pres-ence of regions of negative (i.e. saddle-like) Gaussian cur-vature and high mean curvature. Fig. 1D shows how therelative area fraction of the lipids determines the loca-tion of the interface in dumbbells. For low xLD (below∼ 23%), the softer disordered phase is pinned along thehighly curved neck. For higher xLD we observe only twodomains: in the range 24− 50%, the interface lies alwaysalong the neck, while for higher values it is positionedon one lobe. We find that 87% of the 200 dumbbellsexhibited an interface along the neck. This observationindicates that the energetic advantage resulting from po-sitioning the interface along the neck is large enough tomake any other configuration significantly unfavourable.Remarkably, the fact that the interface remains localisedalong the neck, for a whole range of concentrations, im-plies that the lipid composition of at least one of the twodomains must vary. We name this unprecedented phe-nomenon antimixing, to emphasise that the membranecomposition is sharply inhomogeneous and yet featurestwo coexisting mixed domains.

The additional curvature asymmetry (radii ratio 0.57)in snowman particles increases this striking effect (Fig.1E). For xLD equal to 13%, three domains are present;above this value, we observed only two. In the lattercase, the interface always lies along the neck and thecompositions of each lobe vary continuously. Out of 200snowman SLVs, 91% showed antimixing. We conjecturethat these unexpected results are a combination of threeeffects: the bending-rigidities-induced preference of oneof the two phases for curved regions, the high stabilityof the interface provided by the neck region, and, cru-cially, the influence of curvature on the thermodynamicsof phase separation.

To shed light on the experimental results, we havemodelled the membrane as a closed surface Σ endowedwith a continuous order parameter φ representing the lo-cal concentration of unsaturated lipids, so that

∫Σ

dAφcorresponds to the area occupied by the LD phase. Theequilibrium configurations of φ are determined by min-ima of the following free-energy inspired by Julicher-Lipowsky theory of multicomponent vesicles [14, 15]:

F =

∫Σ

dA

2|∇φ|2 +

1

εV (φ) + k(φ)H2 + k(φ)K

],

(1)where H and K are the mean and Gaussian curvatures

25%

50%

75%

- 0.5 - 0.4 - 0.3 - 0.2 - 0.1 0 0.1 0.2 0.3 0.4 0.50

0.1

0.2

0.3

0.4

0.5

0.09 0.18 0.27 0.36 0.45 0.55 0.64 0.73 0.82 0.91 1.00

A

B

XLD

C

D

Figure 2. Numerical simulation of phase domains. AAllowed regions in the difference of bending rigidities param-eter space where experimentally observed domain configura-tions can be reproduced through numerical simulations for adumbbell-shaped membrane. The unit of length is such thatthe area of the dumbbell is equal to one. Shaded regions referto the approximate portion of the phase space where obser-vations can be reproduced, with each colour representing adifferent xLD value, as indicated in the top-right inset. Theyellow highlighted overlap region is the allowed parameterrange for the bending rigidities, for fixed lateral line tensionσ. The black dot in the highlighted regions corresponds tothe point (∆k,∆k) = (−0.3σ, 0.3σ). B-D Equilibrium con-figurations for cubes, dumbbells and snowman particles forvarying area fractions, using as coupling values the black dotin the top panel. We are unable to put upper bounds on ∆k,nor to ∆k, and estimate that ∆k ' −∆k in agreement with[6].

of Σ. The gradient term in the energy favours gently-varying field configurations, while the second term,V (φ) = V0φ

2(φ − 1)2, is a double-well potential thatfavours phase-separated configurations, where φ is uni-form over portions of Σ which we identify as the LOand LD phases. The parameter ε controls the thick-ness of the one-dimensional interfaces separating the twophases. In the limit ε → 0 the first two terms of Eq.(1) converge to the interfacial energy of Ref. [14] withline tension σ = 2

3

√2V0 (see SI §C.2). The functions

k(φ) and k(φ) represent the elastic moduli of the LO andLD phases and can be parametrised as k(φ) = ∆k f(φ)and k(φ) = ∆k f(φ), with ∆k and ∆k the differencein bending rigidity and saddle-splay modulus betweenthe LO and LD phase and f(φ) a dimensionless func-tion interpolating between 0 and 1. Here we choose

4

f(φ) = φ2(3−2φ), but the precise form of this function isuninfluential. Finally, mass conservation is enforced viaa Lagrange multiplier λ, so that the total energy is givenby F + λ

∫Σ

dAφ.Free-energy minimisation in the sharp interface limit

(i.e. ε→ 0) yields:

σκg = ∆kH2 + ∆k K + λ , (2)

where κg is the geodesic curvature of the interface (seeSI §C.1). Eq. (2) is the two-dimensional analogue ofthe Young-Laplace equation for liquid-liquid interfaces:on a flat substrate, where both H and K vanish, so-lutions describe a circular droplet of radius σ/λ, withthe Lagrange multiplier effectively working as a Laplacepressure across the interface. For spherical particles ofradius R, for which H2 = K = 1/R2, the equilibriuminterface lies along constant geodesic curvature lines, i.enon-maximal circles. The solid angle underlying suchcircles is either fixed by mass conservation (λ 6= 0) or bythe elastic moduli differences of the two phases. On moregeneral curved substrates, the effective Laplace pressureis affected by the geometric properties of the environ-ment, modulated by the difference in rigidity betweenthe two lipid phases.

Although intuitive, Eq. (2) is extremely difficult tosolve for more complex shapes, such as those consideredin our experiments. However, one can construct approxi-mated solutions of Eq. (2) by numerical minimisation ofthe free energy Eq. (1) for finite ε (see SI §C.3). Fig. 2A(inset) shows three examples for symmetric dumbbellshaving xLD equal to 25%, 50% and 75% correspondingto the 4th, 6th and 9th configuration of Fig. 1D. Bycomparison with the experimental results we identifiedthe region of appropriate values for the material param-eters ∆k/σ and ∆k/σ, indicated by the yellow hue inFig. 2. We found good agreement for ∆k > 0, ∆k < 0and |∆k| ' |∆k| ' 10−1σ (for units of length normalisedwith the area of the particle), consistent with previousestimates for free-standing bilayers [6]. Choosing the val-ues ∆k = 0.3σ and ∆k = −0.3σ from the region of ap-propriate parameters identified for dumbbells, we thencalculated the equilibrium configurations for cubic andsnowman shaped particles (Fig. 2B-D). By comparingFig. 1C and Fig. 2B, we see that the model qualita-tively reproduces the observed phase-separation patternson cubic scaffolds. Similarly, Fig. 1D and Fig. 2C showgood agreement for most area fractions on dumbbell par-ticles. On the other hand, comparing Figs. 1E and 2D,we find that antimixing cannot be captured by the sharpinterface limit of Eq. (1), as expected.

To shed light on the origin of antimixing in snowmanparticles (Fig. 3A,B), we have developed a simplifiedvariant of the model, where the effect of curvature on thethermodynamics of phase separation is explicitly takeninto account. Snowman particles are approximated bytwo disjoint spheres which are allowed to exchange lipids.

Axial position

I norm

Figure 3. The antimixed state. A From left to right:snowman shaped SLVs in phase-separated state, antimixedstates and mixed state. Scale bar 1 µm. B Normalised in-tensity profiles along the axial direction of the TOP FluorSM and DOPE Rhodamine in green and magenta respec-tively. Phase-separated SLVs show different intensity peaksin both lobes, while mixed configurations exhibit overlap ofthe whole intensity profile. Antimixed SLVs display overlapof peaks in one lobe and different peaks in the other. C Mini-mal model of antimixing on snowman particles. The particlesare approximated by two disjoint spheres allowed to exchangelipids. The diagram shows the local concentration φ of un-saturated lipids on each sphere, with the diagonal dot-dashedlines indicating the configurations of equal average concen-tration Φ0 = 1

A

∫dAφ. The vertical and horizontal dashed

(solid) lines mark the spinodal (binodal) point on each sphere.The solid curve crossing the diagram diagonally represents theline of equilibria obtained by analytically minimising the freeenergy. When such a line is in between the spinodal lines, thelipids are phase-separated. The black dots are computed fromnumerical simulations and match the analytical values outsidethe spinodal region. (Inset) 3D reconstruction of representa-tive configurations. The colouring of mixed and antimixedconfigurations is pure green for φ = 0 and pure magenta forφ = 1. Lipids are antimixed for 34% ≤ Φ0 ≤ 46% and phase-separated for 46% ≤ Φ0 ≤ 63%.

The neck region acts as a barrier for lipid domains tomigrate from one lobe to the other, but does not stopsingle lipids to diffuse freely. Then, in Eq. (1) we takeV (φ) = Jφ(1 − φ) + φ log φ + (1 − φ) log(1 − φ), withJ a constant. Such a thermodynamic potential, corre-sponding to the mean-field free-energy for a lattice-gassystem, can account for mixed and phase-separated con-figurations alike and reduces to the usual double-well po-

5

tential in the presence of phase-separation. Finally, weset k(φ) + k(φ) = J/ε f(φ), so that the curvatures canaffect directly the thermodynamic potential arbitrarilyclose to the thin interface limit (i.e. ε→ 0, SI §C.4). Min-imising the free energy for fixed average concentration,Φ0 = 1

A

∫dAφ, yields the phase diagram of Fig. 3C,

inclusive of the mixed, antimixed and phase-separatedstate. We stress here that Φ0 equates the area fractionxLD only in the case of phase-separated configurations,whereas for mixed and antimixed configurations it cor-responds to the relative amount of unsaturated lipids.The mixed and antimixed configurations form a line ofequilibria (solid line) for which the concentration φ is uni-form across each lobe, but differs between the two lobesin case of antimixed configurations. The latter scenariodoes not occur if lipid mixing is unaffected by the cur-vature of the substrate (in which case φ = Φ0 and theline coincides with the bisectrix of the diagram). Wetherefore conclude that the observed antimixed state ofFig. 1E and Fig. 3A,B originates mainly from the factthat the geometry of the substrate affects not only thelateral organisation of lipid domains but also the free-energy landscape of the order parameter. Though thisphenomenon shares some similarity with the curvature-driven lipid sorting observed by Sorre et al. [16], it doesnot require our system to be on the verge of demixingand is a system-wide effect.

In this article we have introduced a new experimentalmodel system to decipher the complex physics of phaseseparation on curved surfaces and cells in particular:scaffolded lipid vesicles (SLVs). Like the cytoskeletonand proteins shape the cellular membrane, SLVs imposetheir geometry onto the supported lipid membrane allow-ing an independent study of geometry-induced effects.With this unique setup, we discovered a novel state oflipid membranes, the antimixed state, in which the com-position of the lipid domains adapts to the geometry ofthe scaffold. The antimixed state occurs when the under-lying geometry is highly inhomogeneous, and leads to co-existing mixed phases, whose concentrations of saturatedand unsaturated lipids adapts to the overall composition.

While the interplay between geometry and localizationof the domains had been previously observed, setting thescaffold geometry allowed us to fully understand the del-icate balance between the curvature of the surface, thematerial parameters and the composition of the mem-brane that leads to geometric pinning. Surprisingly, thereis an upper and a lower bound for the area fraction of thesofter phase for geometric pinning to occur, whose precisevalues are determined by the scaffold shape. Our resultson geometric pinning and antimixing could be extendedto other physical systems where phase separation occurs[17]. They provide key insights into how cells might regu-late local concentrations of lipids and proteins by adapt-ing the membrane shape. The here developed membrane-coated colloids can furthermore switch their surface prop-

erties and, upon further chemical functionalization, theirinteractions [18], from uniform to site-specific, paving theway towards smart materials and biomedical applicationssuch as sensing, imaging and drug delivery.

Acknowledgements This work was supported bythe Netherlands Organisation for Scientific Research(NWO/OCW), as part of the Frontiers of Nanoscienceprogram (MR), VENI grant 680-47-431 (DJK) and theVIDI scheme (LG,PF). We thank Vera Meester for helpwith particle synthesis and electron microscopy imaging.

Author contributions MR and DJK designed the ex-periment. MR performed the experiments and the dataanalysis. PF and LG developed the theoretical model.PF performed the analytical and numerical calculations.LG and DJK conceived the study and supervised the re-search. All authors wrote the article.

[email protected][email protected]

[1] McMahon, H. T. & Gallop, J. L. Membrane curvatureand mechanisms of dynamic cell membrane remodelling.Nature 438, 590 (2005).

[2] Peter, B. J. et al. Bar domains as sensors of membranecurvature: the amphiphysin bar structure. Science 303,495–499 (2004).

[3] Dolgin, E. What lava lamps and vinaigrette can teach usabout cell biology. Nature 555, 300 (2018).

[4] Rayermann, S. P., Rayermann, G. E., Cornell, C. E.,Merz, A. J. & Keller, S. L. Hallmarks of reversible sepa-ration of living, unperturbed cell membranes into two liq-uid phases. Biophysical Journal 113, 2425–2432 (2017).

[5] Baumgart, T., Hess, S. T. & Webb, W. W. Imaging co-existing fluid domains in biomembrane models couplingcurvature and line tension. Nature 425, 821 (2003).

[6] Baumgart, T., Das, S., Webb, W. & Jenkins, J. Mem-brane elasticity in giant vesicles with fluid phase coexis-tence. Biophys. J. 89, 1067–1080 (2005).

[7] Hess, S. T., Gudheti, M. V., Mlodzianoski, M. & Baum-gart, T. Shape analysis of giant vesicles with fluid phasecoexistence by laser scanning microscopy to determinecurvature, bending elasticity, and line tension. In Meth-ods in membrane lipids, 367–387 (Springer, 2007).

[8] Semrau, S. & Schmidt, T. Membrane heterogeneity–fromlipid domains to curvature effects. Soft Matter 5, 3174–3186 (2009).

[9] Parthasarathy, R., Yu, C.-h. & Groves, J. T. Curvature-modulated phase separation in lipid bilayer membranes.Langmuir 22, 5095–5099 (2006).

[10] Subramaniam, A. B., Lecuyer, S., Ramamurthi, K. S.,Losick, R. & Stone, H. A. Particle/fluid interface repli-cation as a means of producing topographically patternedpolydimethylsiloxane surfaces for deposition of lipid bi-layers. Adv. Mater. 22, 2142–2147 (2010).

[11] De Almeida, R. F., Fedorov, A. & Prieto, M. Sph-ingomyelin/phosphatidylcholine/cholesterol phase dia-gram: boundaries and composition of lipid rafts. Bio-phys. J. 85, 2406–2416 (2003).

6

[12] Sezgin, E. et al. Partitioning, diffusion, and ligand bind-ing of raft lipid analogs in model and cellular plasmamembranes. Biochim. Biophys. Acta 1818, 1777–1784(2012).

[13] Madwar, C., Gopalakrishnan, G. & Lennox, R. B. Inter-facing living cells and spherically supported bilayer lipidmembranes. Langmuir 31, 4704–4712 (2015).

[14] Julicher, F. & Lipowsky, R. Domain-induced budding ofvesicles. Phys. Rev. Lett. 70, 2964–2967 (1993).

[15] Elliott, C. M. & Stinner, B. A surface phase field modelfor two-phase biological membranes. SIAM J. Appl.Math. 70, 2904–2928 (2010).

[16] Sorre, B. et al. Curvature-driven lipid sorting needs prox-imity to a demixing point and is aided by proteins. Proc.Natl. Acad. Sci. U.S.A. 106, 5622–5626 (2009).

[17] Franzmann, T. M. et al. Phase separation of a yeast prionprotein promotes cellular fitness. Science 359, eaao5654(2018).

[18] Chakraborty, I., Meester, V., van der Wel, C. & Kraft,D. J. Colloidal joints with designed motion range andtunable joint flexibility. Nanoscale 9, 7814–7821 (2017).

[19] Kim, J.-W., Larsen, R. J. & Weitz, D. A. Synthesis ofnonspherical colloidal particles with anisotropic proper-ties. J. Am. Chem. Soc. 128, 14374–14377 (2006).

[20] Sugimoto, T., Khan, M. M., Muramatsu, A. & Itoh, H.Formation mechanism of monodisperse peanut-type α-fe2o3 particles from condensed ferric hydroxide gel. Col-loids Surf. A 79, 233–247 (1993).

[21] Rossi, L. et al. Cubic crystals from cubic colloids. SoftMatter 7, 4139–4142 (2011).

7

MATERIALS AND METHODS

Reagents

1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine(POPC), porcine brain sphingomyelin (BrainSM), ovine wool cholesterol, 1,2-dioleoyl-sn-glycero-3-phosphoethanolamine-N-lissamine rho-damine B sulfonyl 18:1 (Liss Rhod PE), N-[11-(dipyrrometheneboron difluoride)undecanoyl]-D-erythro-sphingosylphosphorylcholine (C11 TopFluorSM), 1,2-dioleoyl-sn-glycero-3-phosphoethanolamine-N-[methoxy(polyethylene glycol)-2000] (DOPE PEG 2000)were purchased from Avanti Polar Lipids.

Silica spheres were purchased from MicroparticlesGmbH (2.06 ± 0.05 µm and 7±0.29 µm). Polystyrene-3-(Trimethoxysilyl)propyl methacrylate (PS-TPM) dumb-bell and snowman particles were synthesised by mak-ing a protrusion from swollen polystyrene particles [19].Hematite cubic particles were made following the methodof [20] and coated with silica [21]. Details about particlesyntheses in SI §A.1.

Hepes buffer was made with 115 mM NaCl, 1.2 mMCaCl2, 1.2 mM MgCl2, 2.4 mM K2HPO4 and 20 mMHepes.

Lipid bilayer coating

A mixture of 500 µg of SM, POPC and cholesterolin 2:1:1 mole ratio was prepared in chloroform. 0.2%mole fraction of Liss Rhod PE and C11 TopFluor SMwere added to fluorescently label the liquid-disorderedand liquid-ordered phases respectively. 5% mole frac-tion of DOPE PEG 2000 was added to improve themobility of the bilayer. The lipids were dried in twohours by vacuum desiccation for two hours and thenre-suspended to a 2 mg/mL solution with Hepes buffer.The solution was vortexed for 15 minutes and heated to70 ◦C. The lipid solution was extruded 21 times witha mini extruder (Avanti Polar Lipids) equipped withtwo 250 µg gas-tight syringes (Hamilton), four draindiscs and nucleopore track-etch membranes (Whatman)and placed on a heating plate set at 70◦C. Then, in a

rotavapor (Buchi) 50 µL of SUVs were added to 1 mLof 0.05 %w/v of particles dispersed in Hepes. Particleswere left undergoing gentle rotation at 70◦C for onehour. The temperature and the time were chosen tokeep the SUVs in the mixed state during the spreadingand to give them enough time to spread on the surfaceof the particles. The solution was then centrifuged at3000 rpm for 10 minutes and the supernatant replacedwith Hepes to remove any SUVs in excess.

Imaging and analysis

The SLVs were imaged at room temperature withan inverted confocal microscope (Nikon Eclipse Ti-E)equipped with a Nikon A1R confocal scanhead withgalvano and resonant scanning mirrors. A 100x oilimmersion objective (NA=1.4) was used. 488 and 561nm lasers were used to excite excited respectively TopFluor and Lissamine Rhodamine dyes. Lasers passedthrough a quarter wave plate to avoid the polarisationof the dyes and the emitted light was separated into twochannels using a dichroic mirror equipped with 500-550nm and 565-625 nm filters. Mobility of the bilayerwas check by fluorescence recovery after photobleaching(FRAP) experiments in which the recovery of the signalwas fitted with an exponential curve (SI §A.2). Three-dimensional image stacks were acquired by scanning thesample in the z direction with a MCL Nano-drive stageand reconstructed with Nikon AR software.The composition of the SLVs was analysed by usingPython scripts. For spherical, dumbbell and snowmanparticles every slice of the z-stack was fitted witha sphere, a nephroid and a two circles. For cubicparticles only the equatorial plane was fitted with asquircle. The intensity along the fitted shapes wasnormalized by the maximum intensity and binarised,to obtain the ratio between the liquid ordered anddisordered phases. For z-stacks this ratio was thenaveraged along the z-direction to obtain the totalratio of area fraction between the two phases, that isreported in Figure 1. The scripts can be found on Github(https://github.com/RinaldinMelissa/SLVsAreaFraction)and a more detailed analysis description of the analysismethod is reported in SI §A.4.