effect of cholesterol on the structure of a phospholipid bilayer · effect of cholesterol on the...

5
Effect of cholesterol on the structure of a phospholipid bilayer Fre ´ de ´ rick de Meyer a,b and Berend Smit b,1 a Centre Europe ´ en de Calcul Atomique et Mole ´ culaire, Ecole Polytechnique Fe ´de ´ rale de Lausanne, CH-1015 Lausanne, Switzerland; and b Department of Chemical Engineering, University of California, Berkeley, CA 94720 Edited by Michael L. Klein, University of Pennsylvania, Philadelphia, PA, and accepted by the Editorial Board January 7, 2009 (received for review October 5, 2008) Cholesterol plays an important role in regulating the properties of phospholipid membranes. To obtain a detailed understanding of the lipid– cholesterol interactions, we have developed a mesoscopic water–lipid– cholesterol model. In this model, we take into account the hydrophobic– hydrophilic interactions and the structure of the molecules. We compute the phase diagram of dimyristoylphosphati- dylcholine– cholesterol by using dissipative particle dynamics and show that our model predicts many of the different phases that have been observed experimentally. In quantitative agreement with ex- perimental data our model also shows the condensation effect; upon the addition of cholesterol, the area per lipid decreases more than one would expect from ideal mixing. Our calculations show that this effect is maximal close to the main-phase transition temperature, the lowest temperature for which the membrane is in the liquid phase, and is directly related to the increase of this main-phase transition temper- ature upon addition of cholesterol. We demonstrate that no conden- sation is observed if we slightly change the structure of the choles- terol molecule by adding an extra hydrophilic head group or if we decrease the size of the hydrophobic part of cholesterol. biomembrane molecular simulation phase behavior dimyristoylphosphatidylcholine mesoscopic model I n this article we address a seemingly simple thermodynamic question: how does the area per molecule of a phospholipid membrane change if we add cholesterol? This question was first posed by Leathes (1) in 1925 and is still being discussed today. The significance of this question relates directly to the importance of cholesterol for the functioning of membranes of higher eukaryotes. For example, cholesterol regulates the fluidity of the membrane and modulates the function of membrane proteins (2). Understand- ing these mechanisms has motivated many researchers to investi- gate the lipid–cholesterol interactions in detail. Because a mem- brane can be seen as a 2D liquid, a first estimate of how the area per molecule would change upon the addition of cholesterol would be to assume ideal mixing, where the area per molecule is simply a weighted average of the pure-components areas. In 1925 Leathes showed that, instead of ideal mixing, one observes a striking nonideal behavior (1). This nonideal behavior is called the con- densing effect (3) because the area per molecule is much lower compared with ideal mixing. Because a membrane behaves as an incompressible f luid, a decrease of the area per molecule will result in a corresponding significant increase of the total thickness of the bilayer. Such an increase of the thickness signals a reorganization of the structure of the membrane. Because changes in the structure of the membrane may have important consequences for the func- tioning of proteins (2), it is important to have a better molecular understanding of the cholesterol-induced changes. Different conceptual models have been proposed to explain the nonideal cholesterol–lipid interactions. Examples include the condensed-complexes model (4, 5), the superlattice model (6), and the umbrella model (7). The condensed-complexes model explains the condensation effect by assuming that cholesterol induces the reversible formation of a stoichiometric cholesterol–lipid complex. In such a complex the membrane is condensed as the lipid acyl chains are more ordered. At a given cholesterol concentration, an equilibrium composition exists between these condensed choles- terol–lipid complexes and ordinary lipids. The superlattice model assumes that at critical concentrations the cholesterol molecules exhibit a specific long-ranged order. The umbrella model takes the point of view that the hydrophilic part of cholesterol is too small and therefore the lipids need to contribute to the screening of the cholesterol molecules from hydrophobic interactions with water. The phospholipids can only create this umbrella if these molecules straighten to make space for cholesterol. In these models the underlying mechanisms leading to condensing are very different. Interestingly, a recent experimental study concluded that their data supported the condensed-complexes model (8), whereas another set of experiments did not find any indication of the condensed complexes and supported the umbrella model (9). These differ- ences in insights motivated us to develop a mesoscopic model of a lipid–cholesterol–water system. We used molecular simulations to shed some light on the lateral organization of cholesterol in lipid membranes and the underlying lipid–cholesterol interactions that induce the condensation effect. Several molecular simulations of all-atom and coarse-grained models of cholesterol in lipid bilayers have been reported in the literature (for some recent examples, see refs. 10–13 and references therein). Ideally, one would like to use all-atom simulations to study the condensing effect over a large range of temperatures and compositions. At present, however, these simulations are too time- consuming. Therefore, we use a coarse-grained model, in which efficiency is gained by integrating out some of the details of an all-atom model. Our model is based on the model of Kranenburg and coworkers (14, 15). The model uses explicit water molecules. Lipids and cholesterol consist of hydrophilic and hydrophobic particles (see Fig. 1). This model lumps groups of atoms into one mesoscopic pseudoatom. The intramolecular interactions include bond vibrations and bond bending of which the parameters have been optimized to mimic the structure of a single all-atom phos- phatidylcholine molecule in water. The hydrophilic and hydropho- bic interactions are described with soft-repulsive interactions, and the parameters of these interactions are related to the solubility parameters by using the method described by Groot and Rabone (16). The idea of this model is that the main driving forces of cholesterol–phospholipid mixing are the hydrophobic and the hy- drophilic interactions, which is the conclusion of many experimen- tal studies (7, 9, 17). In our model, the unit of length is directly related to the effective size of a pseudoatom, i.e., one pseudoatom occupies a volume of 90 Å 3 . The unit of energy follows from the matching of the soft-repulsion parameters of the mesoscopic water particles to the compressibility of water at ambient conditions. The simplicity of the models requires a reparameterization of these soft repulsions if the temperature is changed. However, in this work, we Author contributions: F.d.M. and B.S. designed research; F.d.M. performed research; F.d.M. analyzed data; and F.d.M. and B.S. wrote the paper. The authors declare no conflict of interest. This article is a PNAS Direct Submission. M.L.K. is a guest editor invited by the Editorial Board. 1 To whom correspondence should be addressed. E-mail: [email protected]. 3654 –3658 PNAS March 10, 2009 vol. 106 no. 10 www.pnas.orgcgidoi10.1073pnas.0809959106 Downloaded by guest on January 15, 2021

Upload: others

Post on 22-Sep-2020

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Effect of cholesterol on the structure of a phospholipid bilayer · Effect of cholesterol on the structure of a phospholipid bilayer Fre ´derick de Meyera,b and Berend Smitb,1 aCentre

Effect of cholesterol on the structureof a phospholipid bilayerFrederick de Meyera,b and Berend Smitb,1

aCentre Europeen de Calcul Atomique et Moleculaire, Ecole Polytechnique Federale de Lausanne, CH-1015 Lausanne, Switzerland; and bDepartment ofChemical Engineering, University of California, Berkeley, CA 94720

Edited by Michael L. Klein, University of Pennsylvania, Philadelphia, PA, and accepted by the Editorial Board January 7, 2009 (received for reviewOctober 5, 2008)

Cholesterol plays an important role in regulating the properties ofphospholipid membranes. To obtain a detailed understanding ofthe lipid–cholesterol interactions, we have developed a mesoscopicwater–lipid–cholesterol model. In this model, we take into accountthe hydrophobic–hydrophilic interactions and the structure of themolecules. We compute the phase diagram of dimyristoylphosphati-dylcholine–cholesterol by using dissipative particle dynamics andshow that our model predicts many of the different phases that havebeen observed experimentally. In quantitative agreement with ex-perimental data our model also shows the condensation effect; uponthe addition of cholesterol, the area per lipid decreases more than onewould expect from ideal mixing. Our calculations show that this effectis maximal close to the main-phase transition temperature, the lowesttemperature for which the membrane is in the liquid phase, and isdirectly related to the increase of this main-phase transition temper-ature upon addition of cholesterol. We demonstrate that no conden-sation is observed if we slightly change the structure of the choles-terol molecule by adding an extra hydrophilic head group or if wedecrease the size of the hydrophobic part of cholesterol.

biomembrane � molecular simulation � phase behavior �dimyristoylphosphatidylcholine � mesoscopic model

In this article we address a seemingly simple thermodynamicquestion: how does the area per molecule of a phospholipid

membrane change if we add cholesterol? This question was firstposed by Leathes (1) in 1925 and is still being discussed today. Thesignificance of this question relates directly to the importance ofcholesterol for the functioning of membranes of higher eukaryotes.For example, cholesterol regulates the fluidity of the membraneand modulates the function of membrane proteins (2). Understand-ing these mechanisms has motivated many researchers to investi-gate the lipid–cholesterol interactions in detail. Because a mem-brane can be seen as a 2D liquid, a first estimate of how the areaper molecule would change upon the addition of cholesterol wouldbe to assume ideal mixing, where the area per molecule is simply aweighted average of the pure-components areas. In 1925 Leathesshowed that, instead of ideal mixing, one observes a strikingnonideal behavior (1). This nonideal behavior is called the con-densing effect (3) because the area per molecule is much lowercompared with ideal mixing. Because a membrane behaves as anincompressible fluid, a decrease of the area per molecule will resultin a corresponding significant increase of the total thickness of thebilayer. Such an increase of the thickness signals a reorganizationof the structure of the membrane. Because changes in the structureof the membrane may have important consequences for the func-tioning of proteins (2), it is important to have a better molecularunderstanding of the cholesterol-induced changes.

Different conceptual models have been proposed to explain thenonideal cholesterol–lipid interactions. Examples include thecondensed-complexes model (4, 5), the superlattice model (6), andthe umbrella model (7). The condensed-complexes model explainsthe condensation effect by assuming that cholesterol induces thereversible formation of a stoichiometric cholesterol–lipid complex.In such a complex the membrane is condensed as the lipid acyl

chains are more ordered. At a given cholesterol concentration, anequilibrium composition exists between these condensed choles-terol–lipid complexes and ordinary lipids. The superlattice modelassumes that at critical concentrations the cholesterol moleculesexhibit a specific long-ranged order. The umbrella model takes thepoint of view that the hydrophilic part of cholesterol is too small andtherefore the lipids need to contribute to the screening of thecholesterol molecules from hydrophobic interactions with water.The phospholipids can only create this umbrella if these moleculesstraighten to make space for cholesterol. In these models theunderlying mechanisms leading to condensing are very different.Interestingly, a recent experimental study concluded that their datasupported the condensed-complexes model (8), whereas anotherset of experiments did not find any indication of the condensedcomplexes and supported the umbrella model (9). These differ-ences in insights motivated us to develop a mesoscopic model of alipid–cholesterol–water system. We used molecular simulations toshed some light on the lateral organization of cholesterol in lipidmembranes and the underlying lipid–cholesterol interactions thatinduce the condensation effect.

Several molecular simulations of all-atom and coarse-grainedmodels of cholesterol in lipid bilayers have been reported in theliterature (for some recent examples, see refs. 10–13 and referencestherein). Ideally, one would like to use all-atom simulations to studythe condensing effect over a large range of temperatures andcompositions. At present, however, these simulations are too time-consuming. Therefore, we use a coarse-grained model, in whichefficiency is gained by integrating out some of the details of anall-atom model. Our model is based on the model of Kranenburgand coworkers (14, 15). The model uses explicit water molecules.Lipids and cholesterol consist of hydrophilic and hydrophobicparticles (see Fig. 1). This model lumps groups of atoms into onemesoscopic pseudoatom. The intramolecular interactions includebond vibrations and bond bending of which the parameters havebeen optimized to mimic the structure of a single all-atom phos-phatidylcholine molecule in water. The hydrophilic and hydropho-bic interactions are described with soft-repulsive interactions, andthe parameters of these interactions are related to the solubilityparameters by using the method described by Groot and Rabone(16). The idea of this model is that the main driving forces ofcholesterol–phospholipid mixing are the hydrophobic and the hy-drophilic interactions, which is the conclusion of many experimen-tal studies (7, 9, 17). In our model, the unit of length is directlyrelated to the effective size of a pseudoatom, i.e., one pseudoatomoccupies a volume of 90 Å3. The unit of energy follows from thematching of the soft-repulsion parameters of the mesoscopic waterparticles to the compressibility of water at ambient conditions. Thesimplicity of the models requires a reparameterization of these softrepulsions if the temperature is changed. However, in this work, we

Author contributions: F.d.M. and B.S. designed research; F.d.M. performed research; F.d.M.analyzed data; and F.d.M. and B.S. wrote the paper.

The authors declare no conflict of interest.

This article is a PNAS Direct Submission. M.L.K. is a guest editor invited by the Editorial Board.

1To whom correspondence should be addressed. E-mail: [email protected].

3654–3658 � PNAS � March 10, 2009 � vol. 106 � no. 10 www.pnas.org�cgi�doi�10.1073�pnas.0809959106

Dow

nloa

ded

by g

uest

on

Janu

ary

15, 2

021

Page 2: Effect of cholesterol on the structure of a phospholipid bilayer · Effect of cholesterol on the structure of a phospholipid bilayer Fre ´derick de Meyera,b and Berend Smitb,1 aCentre

assume the parameters to be independent of temperature. Thetemperature scale is set by fitting to the experimental-phase tran-sition temperatures. Further details and applications of this modelcan be found in ref. 18.

Our model of cholesterol, shown in Fig. 1B, is based on the sameassumptions about the effective size and interactions as the lipidmodel. Following McMullen et al. (19), we made the hydrophobicpart of the cholesterol model slightly longer than the hydrophobicpart of the lipid model, which aims to represent dimyristoylphos-phatidylcholine (DMPC). For simplicity, we have assumed that thehydrophobic and hydrophilic interactions of a cholesterol moleculeare similar to those in a lipid. To gain insight into the molecularmechanism of the condensing effect of cholesterol, we have intro-

duced three cholesterol-like molecules in which we perturb thehydrophobic–hydrophilic balance of the molecule: one in which wedecrease the hydrophobic tail length (see Fig. 1C), one in which weadd an additional hydrophobic group (Fig. 1D), and one in whichwe replace the ring by a simple chain (Fig. 1E).

Fig. 2 shows the computed temperature composition phasediagram of the water–phospholipid–cholesterol system. The phaseboundaries were obtained from a visual inspection of the snapshotsand, more quantitatively, from the inflection points of the curvesthat give the area per lipid, the average hydrophobic thickness of themembrane, and the tail order and tilt parameters. These propertieswere computed as a function of temperature and cholesterolcontent.

Let us first focus on the pure-lipid phases, and the effect ofcholesterol will be discussed next. For the pure-lipid bilayer, thephase diagram has been calculated by Kranenburg and Smit (14) fora system that is four times smaller. We used the same methodologyto locate the phase boundaries as Kranenburg and Smit did (14).Our results are in excellent agreement with this study, whichindicates that the finite-size effects are small. For the pure phos-pholipid, we observe at high temperatures a liquid phase (L�) inwhich the tails are disordered. At low temperatures, the tails areordered and tilted, which defines the gel phase (L�c). These twophases are separated by the rippled phase (P��), in which we observea microphase separation of domains in which the bilayer is thick andthe lipids are ordered and domains in which the bilayer is thin andthe lipids disordered. The presence of these three phases indicatesthat the resulting phase diagram is in very good agreement with theexperimental diagram of the pure lipid (20). The temperature scaleis set by matching the temperatures of the phase transitions of thegel phase to the ripple phase (Tp) and the ripple phase to the liquidphase (Tm) with the corresponding experimental-phase transitiontemperatures of pure DMPC. A further comparison with theexperimental data is made for the average area per molecule of thebilayer (Fig. 3A), for the bilayer thickness (Fig. 4A), and for the lipidtail order (Fig. 4B). For the area per lipid we obtain 56 Å2 permolecule compared with the experimental (21) 60 Å2 per molecule.For the bilayer thickness we computed a value of 38.7 Å, whichcompares well with the experimental value of 36 Å (21), and asimilar agreement is obtained for the tail order (see Fig. 4B). Tocompute the area per molecule for pure cholesterol, we simulateda bilayer of pure cholesterol. We obtained a value of 40.3 Å, whichcompares very well with the most recent experimental value of 41Å (22) for a monolayer of cholesterol. Considering the approxi-mations made in our model, the agreement between the experi-mental and the simulated values is surprisingly good.

A B C D E

Fig. 1. Schematic drawing of the mesoscopic models that are studied in thiswork. (A and B) Figure represents DMPC (A) and cholesterol (B). The modelcontains hydrophobic (white) and hydrophilic (black) beads that are connectedwith springs and bond-bending potentials. The model contains explicit watermolecules that are modeled as hydrophilic beads. To study the effect of changein the chemical structure of cholesterol we introduce three ‘‘new’’ molecules inwhich we change the hydrophobic–hydrophilic balance of cholesterol. (C) Cho-lesterol with a shorter tail length. (D) Cholesterol that is more hydrophilic. (E)Cholesterol that is less hydrophobic.

0 10 20 30 40 50mol% cholesterol

0

40

60

80

Tem

pera

ture

LII

LO

Lβ’L

c’

Pβ’

0 1 2 3 4 5 6 7

LII

Lc’

Lβ’

Pβ’L

O

Tm

Tp

20

Condensation effect [Å2]

Fig. 2. Phase diagram and the structure of the variousphases. (Left) Computed-phase diagram as a function oftemperature (in degrees centigrade) and cholesterol con-centration. The black lines give the phase boundaries. Thecolor coding gives the condensation effect at a given statepoint, where blue indicates very little condensation andorange a large condensation effect. (Right) Schematicdrawing of the various phases. L�, lipids in the liquid phase;P��, ripple phase; L��, gel phase with tilted lipid chains; L�c, gelphase with lipid chains not tilted; LII, gel phase, similar to L�c,containing small cholesterol clusters; Lo, liquid-orderedphase.Thecondensationeffect isdefinedasthedifference,in Å2, between AM, sim and AM, ideal.

de Meyer and Smit PNAS � March 10, 2009 � vol. 106 � no. 10 � 3655

BIO

PHYS

ICS

AN

DCO

MPU

TATI

ON

AL

BIO

LOG

YA

PPLI

EDPH

YSIC

AL

SCIE

NCE

S

Dow

nloa

ded

by g

uest

on

Janu

ary

15, 2

021

Page 3: Effect of cholesterol on the structure of a phospholipid bilayer · Effect of cholesterol on the structure of a phospholipid bilayer Fre ´derick de Meyera,b and Berend Smitb,1 aCentre

Let us now turn to the effect of cholesterol on the properties ofthe bilayer. The first question we will address is whether our modelcan reproduce the condensation effect. Fig. 3A shows the effect ofcholesterol on the area per molecule as a function of the cholesterolconcentration. Comparison with the experimental data once againshows very good agreement. In this figure we also show the area permolecule assuming ideal mixing. This figure convincingly illustratesthe condensation effect; the area per molecule decreases muchmore than one would expect on the basis of ideal mixing. Otherexperimental data include the effect of cholesterol on the bilayerswelling (Fig. 4A) and the tail order parameter (Fig. 4B). Both theexperimental data and the simulation show that cholesterol in-creases the thickness of the bilayer and its order. Also for these twoproperties, our simulation results are in very good agreement withthe experimental data. The simulation and experimental data (Figs.3A and 4 A and B) show that the addition of cholesterol stronglymodifies the properties of the lipid bilayer up to �30 mol%cholesterol. After this, a region is reached where further addition ofcholesterol has only a slight effect. At 30 mol% cholesterol, both thearea per molecule and the lipid tail order and tilt parameters havevalues that are typical for the gel phase.

The color coding in Fig. 2 shows the difference between thesimulated area per lipid and the value estimated by assuming idealmixing. We observe that at high and low temperatures, the con-densation effect is relatively small. The condensation effect ismaximal in a well-defined region in phase space that is located justabove the main-phase transition of the pure phospholipid. Tounderstand better the nature of the condensation effect, it isimportant to understand the effect of adding cholesterol on thephase behavior of the membrane.

Fig. 2 shows the most important features of the phase diagram.The different phases we observed in our simulations have also

been observed experimentally, although not always for thespecific mixture of DMPC and cholesterol (20, 23, 24). However,the different experimental studies show qualitatively very dif-ferent phase diagrams, which limits our possibilities to make adetailed comparison. Snapshots of the various phases can befound in Fig. 5.

At very high temperatures, the addition of up to 50 mol%cholesterol has little effect on the structure of the membrane, andwe observe the L� phase for all concentrations (24). At tempera-tures below Tp, we observe that cholesterol changes the structure ofthe gel phase by inhibiting the tilt of the lipid tails, causing theformation of the L�c phase (20) (compare Fig. 5 D with E). At highercholesterol concentrations (�20%), we observe the formation ofsmall, cholesterol-rich clusters. We denote this phase by LII, and thisphase is shown in Fig. 5E. At temperatures between Tp and Tm, thepure bilayer is in the ripple phase, and cholesterol transforms thisripple phase (see Fig. 5C) into a liquid-ordered phase (23) (Fig. 5B).The term liquid-ordered phase was introduced by Ipsen et al. (25).The thickness of the bilayer is intermediate between the thicknessof the liquid and the gel phase. The lipid tail-order parameters havevalues that are close to the gel phase; however, contrary to the gelphase, the lipids are more disordered and do not tilt. Cholesterolgradually increases the temperature at which the L� to Lo phasetransition occurs. At very high cholesterol concentrations, theliquid-ordered phase transforms into a gel phase (LII) when thetemperature is decreased below Tm. This phase has been observedexperimentally for dipalmitoylphosphatidylcholine (26), but wehave not found experimental data for DMPC at these conditions.Let us now return to the condensation effect. Fig. 2 shows that thecondensation effect is maximal at a temperature just above the

0 1010 2020 3030 4040 50504040

4545

5050

5555

6060 Ideal mixingExp. Hung et al.Simulations

Are

a p

er m

ole

cule

mol% Cholesterol

0 1010 2020 3030 4040 50503535

4040

4545

5050

5555

6060

Ideal mixingCholes. more hydrophilicCholes. less hydrophilicCholes. shorter tail

Are

a p

er m

ole

cule

mol% Cholesterol

A

B

Fig. 3. Area per molecule as a function of the cholesterol concentration for themolecules shown in Fig. 1. Data for cholesterol (A) and for the modified choles-terol (B) molecules shown in Fig. 1 C–E. (A) We compare experimental data ofHung et al. (21) with our simulation results and the ideal mixing estimates. Thisestimate is given by AM,mix � (1 � xc)AL � xcAC, with xc as the mole fraction ofcholesterol. AL and AC are the pure-component area per lipid and the area percholesterol, respectively.TheexperimentaldataandsimulationswerebothatT�30 °C. (B) Effect of changes of the cholesterol hydrophobic–hydrophilic balance;the circles are for cholesterol in which the hydrophilic part is increased, thesquares are for cholesterol with a decreased hydrophobic part, and the trianglesare for cholesterol with a shorter tail length (see Fig. 1).

0 1010 2020 3030 4040 50501.001.00

1.051.05

1.101.10

1.151.15

1.201.20

SimulationsExp. Hung et al.Exp. Pan et al.

d/d

0

mol% Cholesterol

0 1010 2020 3030 4040 50500.00.0

0.20.2

0.40.4

0.60.6

0.80.8

SNMR

Simulations S

X-ray Simulations

SNMR

Exp. Mills et al. (DPPC)S

X-ray Exp. Pan et al.

ord

er p

aram

eter

mol% Cholesterol

A

B

Fig. 4. The relative bilayer thickness (A) and order parameter (B) of theDMPC–cholesterol system as a function of cholesterol concentration. (A) Wecompare the experimental data of Pan et al. (30) and Hung et al. (21) with theresults of our simulation. The relative bilayer thickness swelling is defined as d/d0,withd thephosphorus-to-phosphorusdistance intheelectrondensityprofileandd0 the thickness of the pure bilayer. (B) Experimental data are from Pan et al. (30)and Mills et al. (31). The orientational lipid tail order parameter, SNMR, is definedas SNMR � 0.5 �3 cos �2 � 1�, where � is defined as the angle between theorientation of the vector along two beads in the chain and the normal to thebilayer plane, and the average is taken of the ensemble average over all beads.SX-ray quantifies the average tilt of the chain of the lipids by using the sameformulawhere theangle � isbetweentheorientationof thevectoralongthefirstand the last tail beads and the normal to the bilayer plane. The experimental dataand simulations were both at T � 30 °C.

3656 � www.pnas.org�cgi�doi�10.1073�pnas.0809959106 de Meyer and Smit

Dow

nloa

ded

by g

uest

on

Janu

ary

15, 2

021

Page 4: Effect of cholesterol on the structure of a phospholipid bilayer · Effect of cholesterol on the structure of a phospholipid bilayer Fre ´derick de Meyera,b and Berend Smitb,1 aCentre

main transition Tm. The reason is that at these conditions the purebilayer is in a liquid-disordered state, whereas the addition ofcholesterol to the bilayer transforms it into a liquid-ordered phase,which has an area per lipid that is much smaller compared with theliquid-disordered state. This large difference causes a large con-densation effect. At higher temperatures, the liquid phase remainsstable for all cholesterol concentrations, giving a much smallercondensation effect. At lower temperatures, the pure lipid bilayerhas an area per lipid that is much closer to the area per lipid of theliquid-ordered phase, and as a result the condensation effect is farless.

The above results indicate that the condensation effect is a directconsequence of particular changes in the phase behavior thatcholesterol is inducing. In the literature there are various specula-tions about those aspects of the cholesterol structure that arespecifically responsible for its condensing effect. For example, the

umbrella model is based on the notion that compared with phos-pholipids, the hydrophilic part of cholesterol is much smaller andneeds the phospholipid, as an umbrella, for additional screeningfrom interactions with water. This suggests that an additionalhydrophilic group would change the properties completely. An-other important factor is the bulky ring structure; if we replace thering by a tail we obtain a molecule that resembles more an alcoholmolecule (27). However, shortening the hydrophobic tail wouldhave little effect. Fig. 1 shows the modified cholesterol moleculesthat mimic these changes. Indeed, the results in Fig. 3B show thatshortening the tail of cholesterol shows the same condensationeffect. However, Fig. 3B shows that for both other modifications ofthe cholesterol molecule, adding an additional hydrophilic groupand replacing the ring by a linear chain, no condensing effect isobserved. We observe the opposite effect: adding these moleculescauses the bilayer to become more expanded compared with ideal

Fig. 5. Snapshots of a side view of the bilayer. (A) L�

phase for 10% cholesterol at T � 37 °C. (B) L0 phase for40% cholesterol at T � 37 °C. (C) Ripple (P��) phase for5% cholesterol at T � 20 °C. (D) L�� phase for 5% cho-lesterol at T � 5 °C. (E) L�c phase for 15% cholesterol atT � 5 °C. (F) LII phase for 40% cholesterol at T � 5 °C.The hydrophilic and the hydrophobic beads of thephospholipids are depicted in dark blue and in lightblue, respectively. The end beads of the lipid tails aredepicted in gray. The cholesterol headgroup is de-picted in yellow, the cholesterol tetrameric ring andtail beads are depicted in red. For clarity, water beadsare not shown. The difference in the width of thebilayers illustrates the condensation effect nicely.

de Meyer and Smit PNAS � March 10, 2009 � vol. 106 � no. 10 � 3657

BIO

PHYS

ICS

AN

DCO

MPU

TATI

ON

AL

BIO

LOG

YA

PPLI

EDPH

YSIC

AL

SCIE

NCE

S

Dow

nloa

ded

by g

uest

on

Janu

ary

15, 2

021

Page 5: Effect of cholesterol on the structure of a phospholipid bilayer · Effect of cholesterol on the structure of a phospholipid bilayer Fre ´derick de Meyera,b and Berend Smitb,1 aCentre

mixing. The effect of (smaller) alcohols on the area per moleculehas been measured experimentally, and the experimental data alsoshow an increase (28). Closely connected to this, we observed thatfor both cases in the phase diagram the liquid phase was stable overthe entire concentration range. In fact, we observe that adding thesemolecules decreases the main transition temperature, and hencethere is no region in the phase diagram where there is a largecondensation effect.

Simulations with these structural variations of cholesterol indi-cate how surprisingly subtle the mechanism is. The main transitionin a pure bilayer is very sensitive to the hydrophobic interactions.The headgroups of the lipids screen the hydrophobic tails from thewater. At high temperatures, the area per lipid is high, and thisscreening is far from optimal; but at these conditions the chainentropy dominates. Decreasing the temperature makes it increas-ingly important to screen the hydrophobic interactions and at themain transition eventually induces an ordering of the chains. A keyaspect is to understand how cholesterol destabilizes the liquidphase. Cholesterol has a smaller hydrophilic head and is thereforeless efficient in shielding the hydrophobic interactions. At hightemperatures, the lipid bilayer can accommodate this, but at lowertemperatures the lipids can only contribute to the screening of thecholesterol by decreasing its area per lipid. This causes the observedordering and explains why the main transition increases. The twochanges we introduced to the cholesterol structure influence itshydrophobic screening; in both variants the intrinsic undershieldingof cholesterol disappears. If these molecules are added to thebilayer, there is no need for additional screening of the hydrophobicinteractions, and these molecules prevent the formation of anordered phase.

Let us compare our observations with the previous models thathave been introduced to explain the condensation effect. First, ourmodel does not give any indication of long-range ordering as isassumed in the superlattice model. Implicit in both the umbrellamodel and the condensed complexes is the assumption of some kindof local organization. For example, in the umbrella model it isassumed that one lipid molecule could screen one or two neigh-boring cholesterol molecules (see e.g., ref. 2). Our simulations showa much more disordered structure in which we cannot identify theseordered structures. At this point it is important to recall that ourmodel contains many assumptions, and this raises the question of

whether the conclusions we draw from our simulations are relevantfor the experimental systems. We were very surprised to see that wewere able to obtain such a rich phase behavior by using a coarse-grained model that uses purely repulsive forces. Our model gives avery reasonable quantitative description of recent experimentaldata on the structure of the bilayer. The other interesting aspect isthat our calculations predict that the condensation effect is maximalin a narrow temperature range above the main transition. It mightbe possible to verify this experimentally. A very stringent test ofour model would have been a detailed comparison with theexperimental-phase diagram. In this context, it is encouraging thatthe phases we have found have been observed experimentally,although not always for exactly the system simulated. By carefullyselecting those experimental data that agree with our simulationswe could even claim very good agreement. A possible reason for thedisagreement between the various experiments is that differenttechniques are used, and not all techniques are equally sensitive tothe differences in the structure of the various phases. We hope thatthe combination of a phase diagram and detailed informationon the structure of the various phases gives some guidelines as towhether a particular experimental technique can identify a partic-ular phase transition.

Materials and MethodsOur mesoscopic model was studied by using dissipative particle dynamics(DPD) (29). The equations of motion were integrated by using a modifiedversion of the velocity Verlet algorithm with a reduced time step 0.03. Themain modification of the standard DPD algorithm is a method we haveimplemented to ensure that the membrane is simulated in a tensionless state.After on average 15 time steps, a Monte Carlo step was made that involved anattempt to change the area of the lipid in such a way that the total volumeremained constant. The acceptance rule for this move involves the imposedinterfacial tension (15), which was set to zero for our simulations. Furtherdetails on the simulation techniques can be found in ref. 15. To ensuresufficient hydration, we used a system of 100,000 water molecules for a totalof 4,000 cholesterol and lipid molecules. Cholesterol molecules were added tothe system by randomly replacing a lipid molecule by a cholesterol moleculein such a way that the concentration of cholesterol molecules remained thesame on the two sides of the membrane.

ACKNOWLEDGMENTS. We thank Jay Groves for stimulating discussions andDavid Chandler, George Oster, and Jocelyn Rodgers for a critical reading of ourmanuscript.

1. Leathes JB (1925) Croonian lectures on the role of fats in vital phenomena. The Lancet205:853–856.

2. Ikonen E (2008) Cellular cholesterol trafficking and compartmentalization. Nat Rev MolCell Biol 9:125–138.

3. McConnell HM, Radhakrishnan A (2003) Condensed complexes of cholesterol and phos-pholipids. Biochim Biophys Acta Biomembr 1610:159–173.

4. Radhakrishnan A, McConnell HM (1999) Condensed complexes of cholesterol and phos-pholipids. Biophys J 77:1507–1517.

5. Radhakrishnan A, Anderson TG, McConnell HM (2000) Condensed complexes, rafts, andthechemicalactivityof cholesterol inmembranes. ProcNatlAcadSciUSA 97:12422–12427.

6. Chong PLG (1994) Evidence for regular distribution of sterols in liquid-crystalline phos-phatidylcholine bilayers. Proc Natl Acad Sci USA 91:10069–10073.

7. Huang JY (2002) Exploration of molecular interactions in cholesterol superlattices: Effectof multibody interactions. Biophys J 83:1014–1025.

8. Ege C, Ratajczak MK, Majewski J, Kjaer K, Lee KYC (2006) Evidence for lipid/cholesterolordering in model lipid membranes. Biophys J 91:L1–L3.

9. Ali MR, Cheng KH, Huang JY (2007) Assess the nature of cholesterol–lipid interactionsthrough the chemical potential of cholesterol in phosphatidylcholine bilayers. Proc NatlAcad Sci USA 104:5372–5377.

10. Smondyrev AM, Berkowitz ML (2001) Molecular dynamics simulation of the structure ofdimyristoylphosphatidylcholine bilayers with cholesterol, ergosterol, and lanosterol. Bio-phys J 80:1649–1658.

11. Izvekov S, Voth GA (2006) Multiscale coarse-graining of mixed phospholipid/cholesterolbilayers. J Chem Theory Comput 2:637–648.

12. Marrink SJ, Risselada HJ, Yefimov S, Tieleman DP, de Vries AH (2007) The martini forcefield: Coarse grained model for biomolecular simulations. J Phys Chem B 111:7812–7824.

13. Murtola T, Falck E, Patra M, Karttunen M, Vattulainen I (2004) Coarse-grained model forphospholipid/cholesterol bilayer. J Chem Phys 121:9156–9165.

14. Kranenburg M, Smit B (2005) Phase behavior of model lipid bilayers. J Phys Chem B109:6553–6563.

15. Venturoli M, Sperotto MM, Kranenburg M, Smit B (2006) Mesoscopic models of biologicalmembranes. Phys Rep 437:1–54.

16. Groot RD, Rabone KL (2001) Mesoscopic simulation of cell membrane damage, morphol-ogy change and rupture by nonionic surfactants. Biophys J 81:725–736.

17. McMullen TPW, Lewis R, McElhaney RN (2000) Differential scanning calorimetric andFourier transform infrared spectroscopic studies of the effects of cholesterol on thethermotropic phase behavior and organization of a homologous series of linear saturatedphosphatidylserine bilayer membranes. Biophys J 79:2056–2065.

18. de Meyer FJM, Venturoli M, Smit B (2008) Molecular simulations of lipid-mediated pro-tein–protein interactions. Biophys J 95:1851–1865.

19. McMullen TPW, Lewis R, McElhaney RN (1993) Differential scanning calorimetric study ofthe effect of cholesterol on the thermotropic phase-behavior of a homologous series oflinear saturated phosphatidylcholines. Biochemistry 32:516–522.

20. Karmakar S, Raghunathan VA (2005) Structure of phospholipid–cholesterol membranes:An X-ray diffraction study. Phys Rev E 71:061924.

21. Hung WC, Lee MT, Chen FY, Huang HW (2007) The condensing effect of cholesterol in lipidbilayers. Biophys J 92:3960–3967.

22. Brzozowska I, Figaszewski ZA (2002) The equilibrium of phosphatidylcholine–cholesterolin monolayers at the air/water interface. Colloid Surf B 23:51–58.

23. Mortensen K, Pfeiffer W, Sackmann E, Knoll W (1988) Structural properties of a phos-phatidylcholine–cholesterol system as studied by small-angle neutron-scattering: Ripplestructure and phase-diagram. Biochim Biophys Acta 945:221–245.

24. Sankaram MB, Thompson TE (1991) Cholesterol-induced fluid-phase immiscibility in mem-branes. Proc Natl Acad Sci USA 88:868–8690.

25. IpsenJH,KarlstromG,MouritsenOG,WennerstromH,ZuckermannMJ(1987)Phase-equilibriain the phosphatidylcholine–cholesterol system. Biochim Biophys Acta 905:162–172.

26. Wilson-Ashworth HA, et al. (2006) Differential detection of phospholipid fluidity, order,and spacing by fluorescence spectroscopy of bis-pyrene, prodan, nystatin, and merocya-nine 540. Biophys J 91:4091–4101.

27. Kranenburg M, Smit B (2004) Simulating the effect of alcohol on the structure of amembrane. FEBS Lett 568:15–18.

28. Westh P, Trandum C (2000) Partitioning of small alcohols into dimyristoylphosphatidyl-choline (DMPC) membranes: Volumetric properties. J Phys Chem B 104:11334–11341.

29. Frenkel D, Smit B (2002) Understanding Mol Simuls: from Algorithms to Applications(Academic, San Diego), 2nd Ed.

30. Pan JJ, Mills TT, Tristram-Nagle S, Nagle JF (2008) Cholesterol perturbs lipid bilayersnonuniversally. Phys Rev Lett 100:198103.

31. Mills TT, et al. (2008) Order parameters and areas in fluid-phase oriented lipid membranesusing wide angle X-ray scattering. Biophys J 95:669–681.

3658 � www.pnas.org�cgi�doi�10.1073�pnas.0809959106 de Meyer and Smit

Dow

nloa

ded

by g

uest

on

Janu

ary

15, 2

021