30.08.2013 Views

Mesoscopic models of lipid bilayers and bilayers with embedded ...

Mesoscopic models of lipid bilayers and bilayers with embedded ...

Mesoscopic models of lipid bilayers and bilayers with embedded ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>Mesoscopic</strong> <strong>models</strong> <strong>of</strong> <strong>lipid</strong><br />

<strong>bilayers</strong> <strong>and</strong> <strong>bilayers</strong> <strong>with</strong><br />

<strong>embedded</strong> proteins<br />

A Dissipative Particle Dynamics study


<strong>Mesoscopic</strong> <strong>models</strong> <strong>of</strong> <strong>lipid</strong><br />

<strong>bilayers</strong> <strong>and</strong> <strong>bilayers</strong> <strong>with</strong><br />

<strong>embedded</strong> proteins<br />

A Dissipative Particle Dynamics study<br />

ACADEMISCH PROEFSCHRIFT<br />

ter verkrijging van de graad van doctor<br />

aan de Universiteit van Amsterdam,<br />

op gezag van de Rector Magnificus<br />

pr<strong>of</strong>. mr. P.F. van der Heijden<br />

ten overstaan van een door het college voor promoties ingestelde commissie,<br />

in het openbaar te verdedigen in de Aula der Universiteit<br />

op vrijdag 10 december 2004, te 10.00 uur<br />

door<br />

Maddalena Venturoli<br />

geboren te Roma, Italie


Promotiecommissie:<br />

Promotor:<br />

• pr<strong>of</strong>. dr. ir. B. Smit<br />

Overige leden:<br />

• pr<strong>of</strong>. dr. G. Ciccotti<br />

• pr<strong>of</strong>. dr. D. Frenkel<br />

• pr<strong>of</strong>. dr. K.J. Hellingwerf<br />

• pr<strong>of</strong>. dr. J.A. Killian<br />

• pr<strong>of</strong>. dr. B. de Kruijff<br />

• dr. E.J. Meijer<br />

• dr. M.M. Sperotto<br />

Faculteit der Natuurwetenschappen, Wiskunde en Informatica<br />

The research reported in this thesis was carried out at the Department <strong>of</strong> Chemical Engineering,<br />

Faculty <strong>of</strong> Science, University <strong>of</strong> Amsterdam (Nieuwe Achtergracht 166, 1018 WV Amsterdam,<br />

The Netherl<strong>and</strong>s). Financial support was provided by FOM (Stichting voor Fundamenteel<br />

Onderzoek der Materie).


Contents<br />

1 Introduction 1<br />

1.1 The cell membrane . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2<br />

1.2 Study <strong>of</strong> <strong>lipid</strong> <strong>bilayers</strong> <strong>with</strong> computer simulations . . . . . . . . . . . . . 3<br />

1.3 This thesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4<br />

2 Simulation method for coarse-grained <strong>lipid</strong>s 7<br />

2.1 Dissipative Particle Dynamics . . . . . . . . . . . . . . . . . . . . . . . . 8<br />

2.2 Coarse-grained model <strong>of</strong> <strong>lipid</strong>s . . . . . . . . . . . . . . . . . . . . . . . 12<br />

2.3 Self-assembly . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17<br />

3 Surface tension in <strong>lipid</strong> <strong>bilayers</strong> 21<br />

3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22<br />

3.2 Method <strong>of</strong> calculation <strong>of</strong> surface tension . . . . . . . . . . . . . . . . . . 23<br />

3.3 Constant surface tension ensemble . . . . . . . . . . . . . . . . . . . . . 27<br />

3.4 Surface tension in <strong>lipid</strong> <strong>bilayers</strong> . . . . . . . . . . . . . . . . . . . . . . . 30<br />

3.5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36<br />

4 Structural characterization <strong>of</strong> <strong>lipid</strong> <strong>bilayers</strong> 37<br />

4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38<br />

4.2 Structural quantities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39<br />

4.3 Computational details . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41<br />

4.4 Results <strong>and</strong> discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42<br />

4.4.1 Density pr<strong>of</strong>iles . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42<br />

4.4.2 Effect <strong>of</strong> chain length . . . . . . . . . . . . . . . . . . . . . . . . . 45<br />

4.4.3 Lateral pressure pr<strong>of</strong>iles in tensionless <strong>bilayers</strong> . . . . . . . . . . 45<br />

4.4.4 Double-tail <strong>lipid</strong>s . . . . . . . . . . . . . . . . . . . . . . . . . . . 53<br />

5 Phase behavior <strong>of</strong> coarse-grained <strong>lipid</strong> <strong>bilayers</strong> 57<br />

5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58<br />

5.2 Single-tail <strong>lipid</strong> <strong>bilayers</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59<br />

5.2.1 Computational details . . . . . . . . . . . . . . . . . . . . . . . . 59<br />

5.2.2 Results <strong>and</strong> Discussion . . . . . . . . . . . . . . . . . . . . . . . . 60


ii CONTENTS<br />

5.3 Double-tail <strong>lipid</strong> <strong>bilayers</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . 73<br />

5.3.1 Computational details . . . . . . . . . . . . . . . . . . . . . . . . 73<br />

5.3.2 Results <strong>and</strong> Discussion . . . . . . . . . . . . . . . . . . . . . . . . 73<br />

5.4 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76<br />

6 Interaction <strong>of</strong> small molecules <strong>with</strong> <strong>bilayers</strong> 79<br />

6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80<br />

6.2 Computational details . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80<br />

6.3 Results <strong>and</strong> Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82<br />

6.3.1 Partitioning <strong>of</strong> the solutes in the bilayer . . . . . . . . . . . . . . 82<br />

6.3.2 Effect on the bilayer properties . . . . . . . . . . . . . . . . . . . 83<br />

6.3.3 Effect on the pressure distribution . . . . . . . . . . . . . . . . . 90<br />

7 <strong>Mesoscopic</strong> model for <strong>lipid</strong> <strong>bilayers</strong> <strong>with</strong> <strong>embedded</strong> proteins 95<br />

7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96<br />

7.2 Computational details . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98<br />

7.2.1 Lipid <strong>and</strong> protein <strong>models</strong> . . . . . . . . . . . . . . . . . . . . . . 98<br />

7.2.2 Method <strong>of</strong> calculation <strong>of</strong> statistical quantities . . . . . . . . . . . 101<br />

7.3 Results <strong>and</strong> discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104<br />

7.3.1 Protein-induced bilayer perturbation . . . . . . . . . . . . . . . 105<br />

7.3.2 Lipid-induced protein tilting . . . . . . . . . . . . . . . . . . . . 109<br />

7.3.3 Thermotropic behavior . . . . . . . . . . . . . . . . . . . . . . . . 111<br />

7.4 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117<br />

Summary 125<br />

Samenvatting 129<br />

Curriculum Vitae 135<br />

Publications 137<br />

Acknowledgments 139


I<br />

Introduction


2 Introduction<br />

1.1 The cell membrane<br />

Cells are the building blocks <strong>of</strong> every living organism, <strong>and</strong> the smallest units <strong>of</strong> life<br />

capable <strong>of</strong> independently sustaining <strong>and</strong> reproducing themselves. Every cell is a<br />

biochemical reaction center, where energy <strong>and</strong> matter are produced <strong>and</strong> converted,<br />

<strong>and</strong> where the genetic information is stored. A common structure found in all cells<br />

<strong>and</strong> their inner organelles is the membrane. Biological membranes act as semipermeable<br />

barriers, allowing a selected passage <strong>of</strong> small molecules <strong>and</strong> ions. Biomembranes<br />

are constituted <strong>of</strong> a <strong>lipid</strong> matrix in which molecules, such as proteins or cholesterol,<br />

are <strong>embedded</strong> or attached. The <strong>lipid</strong> matrix is formed by the non-covalent<br />

self-assembly <strong>of</strong> two <strong>lipid</strong> monolayers made <strong>of</strong> a variety <strong>of</strong> <strong>lipid</strong> types. Lipids are amphiphilic<br />

molecules, i.e. molecules constituted <strong>of</strong> an hydrophilic polar headgroup,<br />

which is water soluble, <strong>and</strong> hydrophobic tails, which are water insoluble.<br />

The combination <strong>of</strong> hydrophobic <strong>and</strong> hydrophilic groups in the same molecule<br />

is a key factor for the assembly <strong>of</strong> <strong>lipid</strong>s into supra-molecular aggregates, such as<br />

micelles or vesicles, the latter being the templates for the cell membranes. Due to<br />

the hydrophobic effect [1, 2], membrane <strong>lipid</strong>s assemble in such a way that their hydrophobic<br />

part is excluded from a direct contact <strong>with</strong> the water environment, while<br />

the hydrophilic or polar parts are in direct contact <strong>with</strong> the water. The resulting<br />

pseudo two-dimensional system (depicted in figure 1.1) is a fluid structure where<br />

the <strong>lipid</strong>s can diffuse in the membrane plane, can flip-flop from one monolayer to<br />

another, or may even move out <strong>of</strong> the system.<br />

aqueous environment<br />

polar heads<br />

hydrocarbon tails<br />

Figure 1.1: Schematic representation <strong>of</strong> a <strong>lipid</strong> bilayer.<br />

Biological membranes are not inert walls, but complex, organized, dynamic, <strong>and</strong><br />

highly cooperative structures, whose physical properties are important regulators <strong>of</strong><br />

vital biological functions ranging from cytosis <strong>and</strong> nerve processes, to transport <strong>of</strong><br />

energy <strong>and</strong> matter [3].


1.2 Study <strong>of</strong> <strong>lipid</strong> <strong>bilayers</strong> <strong>with</strong> computer simulations 3<br />

1.2 Study <strong>of</strong> <strong>lipid</strong> <strong>bilayers</strong> <strong>with</strong> computer simulations<br />

To relate the structure <strong>and</strong> dynamics <strong>of</strong> biomembranes to their biological function—<br />

the ultimate goal <strong>of</strong> biomembrane science—it is <strong>of</strong>ten necessary to consider simpler<br />

systems. Lipid <strong>bilayers</strong> composed <strong>of</strong> one or two <strong>lipid</strong> species, <strong>and</strong> <strong>with</strong> <strong>embedded</strong><br />

proteins or natural or artificial peptides, are <strong>of</strong>ten used as model systems. Computer<br />

simulations can be used as an approach complementary to experiments for the study<br />

<strong>of</strong> such simplified, s<strong>of</strong>t-condensed matter, systems.<br />

Because <strong>of</strong> the many degrees <strong>of</strong> freedom involved, the processes that take place<br />

even in model biomembranes occur over a wide range <strong>of</strong> time <strong>and</strong> length scales<br />

[4]. The typical time <strong>and</strong> length scales <strong>of</strong> the processes under investigation do pose<br />

limitations on the level <strong>of</strong> chemical <strong>and</strong> molecular details chosen to represent the<br />

model. Often, this necessity follows the fact that some theoretical methods are limited<br />

in their applicability by the long computational time needed to calculate statistical<br />

quantities. To model membranes, it is thus necessary to decide a priori the level<br />

<strong>of</strong> description <strong>of</strong> the system, i.e. to neglect those details which are not important to<br />

the process one wants to study.<br />

Molecular Dynamics (MD) simulation methods on atomistic detailed <strong>models</strong> have<br />

been used to study the structural <strong>and</strong> dynamic properties <strong>of</strong> membranes [5], the selfassembly<br />

<strong>of</strong> phospho<strong>lipid</strong>s into <strong>bilayers</strong> [6], as well as the interaction <strong>of</strong> membrane<br />

proteins or other molecules <strong>with</strong> the <strong>lipid</strong> bilayer [7–12]. MD simulations can provide<br />

detailed information about the phenomena that occur in biomembrane systems at<br />

the nanoscopic level <strong>and</strong> on a nanosecond time-scale. However, many membrane<br />

processes happen at the mesoscopic length <strong>and</strong> time scale, i.e. >1-1000 nm, ns, respectively,<br />

<strong>and</strong> involve the collective nature <strong>of</strong> the system. This is the case for phase<br />

separation, the gel-fluid phase transition, the formation <strong>of</strong> domains, or the transition<br />

from a bilayer to a non-bilayer phase. Even though the speed <strong>of</strong> numerical computation<br />

is increasing very rapidly, it will be some time before it will be possible, by<br />

MD on realistic all-atom <strong>models</strong>, to predict the cooperative behavior <strong>of</strong> biosystems<br />

at mesoscopic time <strong>and</strong> length scales.<br />

An alternative modeling approach consists in neglecting most <strong>of</strong> the molecular<br />

details <strong>of</strong> the system. The resulting lattice [13, 14], interfacial [15], or phenomenological<br />

<strong>models</strong> [16–18], are computationally very efficient, <strong>and</strong> can give insight into<br />

the physical properties <strong>of</strong> reconstituted membranes [19, 20]. However, using these<br />

<strong>models</strong> it is difficult to study the structural <strong>and</strong> conformational properties <strong>of</strong> the system<br />

that derive from some molecular details. To overcome this difficulty, we have<br />

developed a model for <strong>lipid</strong> systems which can be seen as an intermediate between<br />

the all-atom <strong>models</strong> <strong>and</strong> the <strong>models</strong> briefly just mentioned. This mesoscopic model<br />

considers a system <strong>of</strong> ‘particles’, or ‘beads’, in which each particle represents a complex<br />

molecular component <strong>of</strong> the system whose details are not important to the process<br />

under investigation. These types <strong>of</strong> <strong>models</strong>, which use simplified interactions


4 Introduction<br />

between the beads, are called coarse-grain (CG) <strong>models</strong>. With the coarse-grain approach<br />

one can explore longer time <strong>and</strong> length scales than it is possible by the more<br />

traditional atomistic approach. Also, compared to thermodynamic or lattice <strong>models</strong><br />

for <strong>lipid</strong> <strong>bilayers</strong>, CG <strong>models</strong> retain a number <strong>of</strong> structural details <strong>of</strong> the molecular<br />

components <strong>of</strong> the system. In the recent years, CG <strong>models</strong> have been developed to<br />

study biomembrane-like systems at the mesoscopic level, <strong>and</strong> both MD <strong>and</strong> Monte<br />

Carlo (MC) simulation methods were used on such <strong>models</strong> [21–25].<br />

However, despite the advantages which arise by minimal modeling in connection<br />

<strong>with</strong> simulation methods like MD <strong>and</strong> MC, the possibility to study processes<br />

that involve the collective behavior <strong>of</strong> the system is still limited. To try to overcome<br />

this limitation, we have used a faster simulation technique, Dissipative Particle Dynamics<br />

(DPD) [26–28], on CG <strong>models</strong>. This approach can be seen as a middle way<br />

between the approach based on pseudo three-dimensional <strong>models</strong>—such as lattice<br />

<strong>and</strong> interfacial <strong>models</strong>—<strong>and</strong> the MD on all-atom model approach.<br />

1.3 This thesis<br />

The scope <strong>of</strong> the work presented in this thesis is to develop a CG model for bilayer<br />

<strong>lipid</strong>s <strong>and</strong> molecules interacting <strong>with</strong> <strong>lipid</strong> <strong>bilayers</strong>, <strong>and</strong> to study the cooperative behavior<br />

<strong>of</strong> these systems. The model is studied by the DPD simulation technique. We<br />

start by investigating the structural <strong>and</strong> thermodynamic properties <strong>of</strong> the pure <strong>lipid</strong><br />

systems. We validate the model by comparing the results for model-<strong>lipid</strong>s <strong>with</strong> those<br />

<strong>of</strong> reconstituted <strong>lipid</strong>-bilayer systems, which are studied experimentally. Once we<br />

have established that the CG model that we have developed is a reliable one to reproduce<br />

some key features <strong>of</strong> a pure <strong>lipid</strong> bilayer, we proceed by extending the model to<br />

the study <strong>of</strong> <strong>bilayers</strong> interacting <strong>with</strong> small solute molecules (which can be considered<br />

as a model for anesthetics), <strong>and</strong> proteins. This thesis is structured as follows.<br />

In Chapter 2 we describe the mesoscopic model for <strong>lipid</strong> <strong>bilayers</strong>. According to<br />

this approach, the <strong>lipid</strong> molecules are coarse-grained <strong>with</strong> sets <strong>of</strong> beads, each <strong>of</strong><br />

which represents a portion <strong>of</strong> the molecule. The interactions between the beads are<br />

simplified respect to atomistic MD representations, but the molecular nature <strong>of</strong> the<br />

<strong>lipid</strong> is retained. As an example, in figure 1.2 the chemical structure <strong>of</strong> the phospho<strong>lipid</strong><br />

dimyristoylphosphatidylcholine (DMPC) <strong>and</strong> its CG representation are shown.<br />

In Chapter 2 we also describe the DPD simulation technique <strong>and</strong> its application<br />

to the CG model. It will be shown that CG <strong>lipid</strong>s spontaneously self-assemble into<br />

supramolecular aggregates, like micelles <strong>and</strong> <strong>bilayers</strong>.<br />

In Chapter 3 we address the debated issue <strong>of</strong> which is the correct value <strong>of</strong> the<br />

surface tension <strong>of</strong> simulated <strong>lipid</strong> <strong>bilayers</strong> in order to reproduce the structural characteristics<br />

<strong>of</strong> self-assembled, unconstrained membranes, which are known to be in a<br />

tensionless state. To this purpose, we introduce a fast <strong>and</strong> efficient simulation technique,<br />

based on the Monte Carlo method, which allows to impose on the bilayer a


1.3 This thesis 5<br />

13<br />

12<br />

11<br />

9<br />

H 10<br />

2 H2<br />

C<br />

H<br />

C<br />

2<br />

H C<br />

H<br />

3 C C 2 H<br />

C<br />

2<br />

H2 C H O<br />

C 2<br />

H2 C C<br />

H2 C<br />

C<br />

H2<br />

C<br />

H2 C O H2 H2<br />

C<br />

CH<br />

O<br />

H2 H3 C C<br />

O CH2 C C<br />

H2 H2 3<br />

7<br />

8<br />

5<br />

6<br />

4<br />

O<br />

P O<br />

O<br />

2<br />

C<br />

C<br />

1<br />

−<br />

H2 N<br />

H2 +<br />

H2 H2 H2 H2 H2 C<br />

C C C C<br />

C C C C C<br />

H2 H2 H2 H2 O<br />

CH3 CH3 CH3 Figure 1.2: The atomistic representation <strong>of</strong> DMPC <strong>and</strong> its corresponding coarse-grained<br />

model.<br />

chosen value <strong>of</strong> the surface tension, <strong>of</strong> which the zero value is a particular case.<br />

In Chapter 4 the structural properties <strong>of</strong> the model bilayer, like the bilayer thickness<br />

or the area per <strong>lipid</strong>, are studied in relation to the characteristics <strong>of</strong> the model<br />

<strong>lipid</strong>s. We address the important question <strong>of</strong> how much chemical detail should—<strong>and</strong><br />

can—be included in CG <strong>models</strong> <strong>of</strong> bilayer <strong>lipid</strong>s. To this purpose, we compare singletail<br />

<strong>and</strong> double-tail model <strong>lipid</strong>s, the latter more closely resembling a real phospho<strong>lipid</strong>,<br />

as illustrated in figure 1.2. We show that, depending on the <strong>lipid</strong> architecture,<br />

like stiffness, acyl-chain length, <strong>and</strong> size <strong>of</strong> the headgroup, <strong>and</strong> on the choice <strong>of</strong> interaction<br />

parameters, the model <strong>bilayers</strong> have different structural characteristics, which<br />

we compare <strong>with</strong> the ones <strong>of</strong> real phospho<strong>lipid</strong> <strong>bilayers</strong>.<br />

In Chapter 5 we determine the phase diagram <strong>of</strong> CG <strong>lipid</strong> <strong>bilayers</strong>, <strong>of</strong> both single<strong>and</strong><br />

double-tail <strong>lipid</strong>s. By performing simulations in the constant surface tension<br />

ensemble, we are able to study structural rearrangements <strong>of</strong> the bilayer in which the<br />

area per <strong>lipid</strong> changes, <strong>and</strong> thus directly observe phase transitions. We show that CG<br />

single-tail <strong>lipid</strong>s can spontaneously form interdigitated <strong>bilayers</strong>. In an interdigitated<br />

bilayer the two monolayers are not separated, instead the terminal tail groups <strong>of</strong> the<br />

<strong>lipid</strong>s in one monolayer extend further into the bilayer, <strong>and</strong> face the headgroups <strong>of</strong><br />

the <strong>lipid</strong>s in the opposing monolayer. Conversely, double-tail <strong>lipid</strong>s, <strong>with</strong> the appropriate<br />

degree <strong>of</strong> stiffness in the tails, do not spontaneously interdigitate, while they<br />

correctly reproduce the gel <strong>and</strong> fluid phases <strong>of</strong> <strong>lipid</strong> <strong>bilayers</strong>.<br />

In Chapter 6 we address the important, <strong>and</strong> not yet fully understood, problem <strong>of</strong><br />

the mechanism <strong>of</strong> action <strong>of</strong> anesthetics. Whether general anesthetics directly bind<br />

to membrane proteins, or act indirectly through changes in the packing properties <strong>of</strong><br />

the <strong>lipid</strong> <strong>bilayers</strong>, is still a matter <strong>of</strong> debate. An attractive hypothesis was recently proposed<br />

by Robert Cantor [29, 30], who described a possible non-specific mechanism<br />

to explain the action <strong>of</strong> anesthetic molecules. Cantor postulated that the activity <strong>of</strong><br />

anesthetics is <strong>lipid</strong>-mediated; conformational changes in transmembrane proteins<br />

whose functionality is related to depth dependent changes in their cross sectional<br />

area might be related to shifts <strong>of</strong> the lateral pressure pr<strong>of</strong>ile across the bilayer induced<br />

by anesthetic molecules. Experimental measurements <strong>of</strong> pressure pr<strong>of</strong>iles in<br />

<strong>lipid</strong> <strong>bilayers</strong> are not yet available; on the other h<strong>and</strong>, pressure pr<strong>of</strong>iles can be directly<br />

computed in molecular simulations. By means <strong>of</strong> simulations, we investigate


6 Introduction<br />

whether the preferred absorption sites <strong>of</strong> small molecules follow the distribution <strong>of</strong><br />

pressure across the bilayer, <strong>and</strong> if—<strong>and</strong> to which extent—these molecules in turn<br />

modify the pressure distribution.<br />

Finally, in Chapter 7 <strong>of</strong> this thesis we extend the mesoscopic model to the case<br />

<strong>of</strong> <strong>lipid</strong> <strong>bilayers</strong> <strong>with</strong> <strong>embedded</strong> proteins spanning both monolayers <strong>of</strong> the bilayer.<br />

We assume that the matching between the hydrophobic length <strong>of</strong> a membrane protein<br />

<strong>and</strong> the hydrophobic thickness <strong>of</strong> the <strong>lipid</strong> bilayer plays an important role in the<br />

<strong>lipid</strong>-protein interaction. If the hydrophobic section <strong>of</strong> a transmembrane protein<br />

does not match the <strong>lipid</strong> bilayer thickness, then a mismatch occurs; the energetic<br />

cost <strong>of</strong> exposing the hydrophobic section <strong>of</strong> either the protein or the <strong>lipid</strong>s to the water<br />

environment is so high that compensating mechanisms take place. Among these,<br />

the <strong>lipid</strong> bilayer thickness can locally increase (or decrease) around the transmembrane<br />

protein, or the protein can tilt [20]. In Chapter 7 we show the results <strong>of</strong> a systematic<br />

study <strong>of</strong> the <strong>lipid</strong>-protein model. The aim <strong>of</strong> the study is to underst<strong>and</strong> under<br />

which conditions—via the cooperative behavior <strong>of</strong> the system, <strong>and</strong> due to hydrophobic<br />

mismatch—the protein may perturb the surrounding <strong>lipid</strong> bilayer (as illustrated<br />

in figure 1.3), <strong>and</strong>, in turn, the <strong>lipid</strong> bilayer may affect the orientation <strong>of</strong> the protein<br />

<strong>with</strong> respect to the bilayer normal.<br />

Figure 1.3: Model bilayer <strong>with</strong> <strong>embedded</strong> a protein.


II<br />

Simulation method for coarse-grained <strong>lipid</strong>s


8 Simulation method for coarse-grained <strong>lipid</strong>s<br />

In this Chapter we describe the Dissipative Particle Dynamics (DPD) simulation<br />

technique that we have applied to the study <strong>of</strong> biological membranes. In addition,<br />

we introduce the mesoscopic model that we use to represent coarse-grained (CG)<br />

<strong>lipid</strong> molecules. In this CG model that we propose, the <strong>lipid</strong>s are represented by s<strong>of</strong>t<br />

spheres interacting by purely repulsive forces, <strong>and</strong> are constructed by connecting<br />

beads via harmonic springs. We also demonstrate that DPD in combination <strong>with</strong><br />

these mesoscopic <strong>models</strong> gives us a sufficiently large gain in CPU-time such that we<br />

can form the bilayer by self-assembly.<br />

2.1 Dissipative Particle Dynamics<br />

Dissipative Particle Dynamics was introduced by Hoogerbrugge <strong>and</strong> Koelman [26] as<br />

a particle based method to simulate complex fluids <strong>with</strong> the correct hydrodynamics<br />

on mesoscopic time <strong>and</strong> length scales. DPD describes the motion <strong>of</strong> particles, or<br />

beads, which represent the center <strong>of</strong> mass <strong>of</strong> a fluid ’droplet’, i.e. each DPD particle is<br />

a momentum carrying group <strong>of</strong> molecules. All the degrees <strong>of</strong> freedom smaller than a<br />

bead radius are assumed to have been integrated out, resulting in a coarse-grain representation<br />

<strong>of</strong> fluid elements <strong>and</strong> in the possibility to adopt s<strong>of</strong>t interactions between<br />

the beads.<br />

DPD resembles Molecular Dynamics (MD) in that the particles move in continuous<br />

space <strong>and</strong> discrete time step, according to Newton’s laws, but, due to the s<strong>of</strong>t<br />

interactions, larger time <strong>and</strong> length scales are accessible, compared to st<strong>and</strong>ard MD.<br />

This allows for the study <strong>of</strong> physical behavior on time scales many orders <strong>of</strong> magnitude<br />

greater than possible <strong>with</strong> MD.<br />

In addition to the s<strong>of</strong>t-repulsive (conservative) interaction, DPD <strong>models</strong> include<br />

two other forces: a dissipative force that slows the particles down <strong>and</strong> removes energy,<br />

which can be seen as a viscous resistance, <strong>and</strong> a stochastic force, which, on<br />

average, adds energy to the system <strong>and</strong> accounts for the degrees <strong>of</strong> freedom which<br />

have been removed by the coarse-graining process. These two forces act together<br />

as a thermostat for the system. Each <strong>of</strong> the three DPD forces is pairwise additive,<br />

conserves momentum, <strong>and</strong> acts along the line joining two particles.<br />

One <strong>of</strong> the attractive features <strong>of</strong> DPD is the easy way in which complex fluids can<br />

be constructed by introducing additional conservative interactions between the particles,<br />

like harmonic potentials to connect particles to build a “bead-<strong>and</strong>-spring” representation<br />

<strong>of</strong> polymer chains.<br />

DPD has been used to study phase separation <strong>and</strong> domain growth [31–33] <strong>of</strong> binary<br />

immiscible fluids by assuming two types <strong>of</strong> particles which repel each other<br />

<strong>with</strong> a strength larger than the repulsion between equal particles. The kinetics <strong>of</strong><br />

microphase separation <strong>of</strong> diblock copolymer has also been investigated <strong>with</strong> DPD<br />

[34,35] showing the critical role <strong>of</strong> hydrodynamic forces. Another successful application<br />

<strong>of</strong> DPD is the modeling <strong>of</strong> dilute polymer solutions [36, 37] <strong>and</strong> the effect <strong>of</strong> sol-


2.1 Dissipative Particle Dynamics 9<br />

vent quality on the static <strong>and</strong> rheological properties <strong>of</strong> polymer solutions has been<br />

investigated [38]. The DPD method applied to the study <strong>of</strong> polymer melts [39] found<br />

results consistent <strong>with</strong> the Rouse-Zimm model. Other applications <strong>of</strong> DPD are to<br />

simulation <strong>of</strong> suspensions <strong>of</strong> colloidal particles [40–42], polymer mixtures [28,36,43],<br />

tethered polymers in shear flow [44], amphiphilic mesophases [45], <strong>and</strong> polymersurfactant<br />

aggregation [46]. For an exhaustive review <strong>of</strong> the DPD technique <strong>and</strong> its<br />

applications we refer the reader to refs. [47] <strong>and</strong> [48].<br />

The simulation technique<br />

The total force acting on a DPD particle i is expressed as a summation over all the<br />

other particles, j, <strong>of</strong> three forces [26, 28] <strong>of</strong> the pairwise-additive type:<br />

fi = <br />

(F C ij + F D ij + F R ij). (2.1)<br />

j=i<br />

The first term in the above equation refers to a conservative force, the second to a<br />

dissipative force <strong>and</strong> the third to a r<strong>and</strong>om force.<br />

The conservative force has two contributions, one describing non bonded interactions<br />

<strong>and</strong> the other describing the interactions used to tie the beads together in the<br />

building <strong>of</strong> a polymer like molecule (see equation 2.13). Most DPD studies express<br />

the former contribution as a s<strong>of</strong>t repulsive force in the form<br />

F C ij =<br />

aij(1 − rij/Rc)^rij (rij < Rc)<br />

0 (rij ≥ Rc)<br />

(2.2)<br />

where the coefficient aij > 0 represents the maximum repulsion strength, rij = ri−rj<br />

is the distance between particles i <strong>and</strong> j (rij = |rij|), <strong>and</strong> Rc is a cut<strong>of</strong>f radius which<br />

gives the range <strong>of</strong> the interactions.<br />

The dissipative <strong>and</strong> r<strong>and</strong>om forces are expressed as:<br />

F D ij = −ηw D (rij)(^rij · vij)^rij<br />

F R ij = σw R (rij)ζij^rij (2.3)<br />

where vij = vi − vj is the velocity difference between particles i <strong>and</strong> j, η is a friction<br />

coefficient <strong>and</strong> σ is the noise amplitude. ζij is a Gaussian (or uniform) distributed<br />

r<strong>and</strong>om number, independent for each pair <strong>of</strong> particles, <strong>with</strong> zero mean <strong>and</strong> unit<br />

variance. The requirement ζij = ζji enforces momentum conservation.<br />

Español <strong>and</strong> Warren [27] demonstrated that, in the limit <strong>of</strong> infinitesimal timestep,<br />

the system satisfies detailed balance <strong>and</strong> achieves a well-defined equilibrium<br />

state, the Gibbs Canonical NVT ensemble, <strong>with</strong> a well defined temperature, derived<br />

from a fluctuation dissipation theorem, if the weight functions <strong>and</strong> coefficients <strong>of</strong> the


10 Simulation method for coarse-grained <strong>lipid</strong>s<br />

dissipative <strong>and</strong> the r<strong>and</strong>om force satisfy the relations:<br />

w D (r) = [w R (r)] 2<br />

σ 2 = 2ηkBT.<br />

(2.4)<br />

This is because the time evolution <strong>of</strong> the distribution function <strong>of</strong> the system f(r, p, t),<br />

which specifies the probability <strong>of</strong> finding the system at time t in a state where the<br />

particles have positions r={ri} <strong>and</strong> momenta p={pi}, is governed by the Fokker-Planck<br />

equation<br />

∂f<br />

∂t = LC f + L D f, (2.5)<br />

where L C <strong>and</strong> L D are the evolution operators. L C is the Liouville operator <strong>of</strong> the<br />

Hamiltonian system interacting <strong>with</strong> conservative forces, <strong>and</strong> L D contains the dissipative<br />

<strong>and</strong> the noise terms. If only L C is present, the system is Hamiltonian <strong>with</strong><br />

equilibrium distribution, feq, the Gibbs distribution. In order to have the Gibbs distribution<br />

also when the r<strong>and</strong>om <strong>and</strong> the dissipative forces are present, the corresponding<br />

evolution operator should satisfy L D feq = 0. The relations 2.4 specify this<br />

condition. The existence <strong>of</strong> a Hamiltonian implies that only the conservative part <strong>of</strong><br />

the force (or better, the potential UC related to it: F C = −∇UC), determines the equilibrium<br />

averages <strong>of</strong> the system observables. Furthermore, as shown by Willemsen et<br />

al. in [49], it implies that DPD can be combined <strong>with</strong> Monte-Carlo (MC) methods.<br />

We will use this property in section 3.3 <strong>of</strong> Chapter 3, where we describe an hybrid<br />

DPD-MC scheme to ensure that the simulations are performed at constant surface<br />

tension.<br />

All the forces assume the same functional dependence on the interparticle distance<br />

rij, as the conservative force F C ij does, if the weight function wR (r) has the following<br />

functional form<br />

w R (r) =<br />

(1 − r/Rc) (r < Rc)<br />

0 (r ≥ Rc)<br />

(2.6)<br />

<strong>and</strong> w D (r) follows from equation 2.4.<br />

The Newton’s equations <strong>of</strong> motion are integrated using a modified version <strong>of</strong> the<br />

velocity-Verlet algorithm [28]:<br />

ri(t + ∆t) = ri(t) + ∆tvi(t) + 1<br />

2 (∆t)2 fi(t)<br />

˜vi(t + ∆t) = vi(t) + λ∆tfi(t) (2.7)<br />

fi(t + ∆t) = fi[r(t + ∆t), ˜v(t + ∆t)]<br />

vi(t + ∆t) = vi(t) + 1<br />

2 ∆t [fi(t) + fi(t + ∆t)].<br />

This integration algorithm becomes the original velocity-Verlet algorithm when the


2.1 Dissipative Particle Dynamics 11<br />

force is independent <strong>of</strong> velocity <strong>and</strong> λ = 0.5. Since in DPD the force depends on<br />

the velocity, the iterative scheme <strong>of</strong> equation 2.7 is used, where ˜v is a prediction for<br />

the velocity. In practice λ is an empirical parameter which takes into account the<br />

fact that the underlying equations <strong>of</strong> DPD are stochastic differential equations [28].<br />

Following Groot <strong>and</strong> Warren in [28] we choose the values λ = 0.65 <strong>and</strong> σ = 3 which<br />

are a compromise between fast simulations (large timestep ∆t) <strong>and</strong> stability <strong>of</strong> the<br />

system <strong>with</strong>out large deviations from the imposed temperature.<br />

Reduced <strong>and</strong> physical units<br />

Usually <strong>with</strong>in the DPD approach one makes use <strong>of</strong> reduced units for the mass, length<br />

<strong>and</strong> energy. [24, 28]. The unit <strong>of</strong> length is defined by the cut-<strong>of</strong>f radius Rc, the unit <strong>of</strong><br />

mass by the mass m <strong>of</strong> a DPD bead (where all the beads in the system have equal<br />

mass), <strong>and</strong> the unit <strong>of</strong> energy by kBT. From these, the unit <strong>of</strong> time τ follows as<br />

<br />

τ = Rc m/kBT. (2.8)<br />

In the following, unless otherwise stated, all the reported quantities will be expressed<br />

in these reduced units.<br />

An important—<strong>and</strong> non trivial—aspect in mesoscopic simulations is the mapping<br />

<strong>of</strong> the above mentioned units onto physical units. The level <strong>of</strong> coarse-graining,<br />

i.e. the number Nm <strong>of</strong> molecules represented by a DPD bead, can be seen as the<br />

renormalization factor for this mapping [24]. As an example we describe below the<br />

procedure to derive the physical unit <strong>of</strong> length. If a DPD bead corresponds to Nm<br />

water molecules, then a cube <strong>of</strong> volume R 3 c represents ρNm water molecules, where ρ<br />

is the number density, i.e. the number <strong>of</strong> DPD beads per cubic Rc. Considering that<br />

a water molecule has approximately a volume <strong>of</strong> 30 ˚A 3 , we then have<br />

R 3 c = 30ρNm [˚A 3 ]. (2.9)<br />

Taking a bead density <strong>of</strong> ρ = 3 [28], from equation 2.9 we have<br />

Rc = 4.4814(Nm) 1/3 [˚A]. (2.10)<br />

For instance, if Nm = 1 then Rc = 4.4814 ˚A, <strong>and</strong> if Nm = 3 then Rc = 6.4633 ˚A.<br />

Different mapping criteria have been used to derive the physical unit <strong>of</strong> time, all<br />

based on the mapping <strong>of</strong> diffusion constants for the system components. For example,<br />

Groot <strong>and</strong> Rabone [24] used a mapping based on the comparison <strong>of</strong> the experimental<br />

value <strong>of</strong> the self-diffusion constant <strong>of</strong> water <strong>and</strong> the corresponding value<br />

computed in DPD simulations, while Groot [46] used the diffusion constant <strong>of</strong> a surfactant<br />

micelle. In both references a value <strong>of</strong> the integration timestep (see equation<br />

2.7) <strong>of</strong> ∆t = 0.06τ was taken, which gives ∆t ≈5ps or ∆t ≈25ps depending on which<br />

diffusion constant is considered. Both values show that DPD simulations allow for


12 Simulation method for coarse-grained <strong>lipid</strong>s<br />

a timestep at least three orders <strong>of</strong> magnitude larger than in atomistic molecular dynamics<br />

simulations, where a timestep <strong>of</strong> the order <strong>of</strong> few fs is typically used.<br />

In the next section we describe the coarse-grained model we use to represent <strong>lipid</strong><br />

molecules <strong>and</strong> we show that these model <strong>lipid</strong>s, simulated <strong>with</strong> DPD, spontaneously<br />

self-assemble into micelles <strong>and</strong> <strong>bilayers</strong>.<br />

2.2 Coarse-grained model <strong>of</strong> <strong>lipid</strong>s<br />

Within the mesoscopic approach, each molecule <strong>of</strong> the system (or groups <strong>of</strong> molecules)<br />

is coarse-grained by a set <strong>of</strong> beads. We consider three types <strong>of</strong> beads: a water-like<br />

bead (denoted as ’w’), which <strong>models</strong> the solvent, an hydrophilic bead (denoted as ’h’)<br />

which <strong>models</strong> a section <strong>of</strong> the <strong>lipid</strong> headgroup, <strong>and</strong> an hydrophobic bead (denoted<br />

as ’t’) which <strong>models</strong> a segment <strong>of</strong> the <strong>lipid</strong> hydrocarbon tail. A <strong>lipid</strong> is constructed<br />

by connecting head- <strong>and</strong> tail-beads <strong>with</strong> springs. The simplest <strong>lipid</strong> consists <strong>of</strong> a linear<br />

chain <strong>of</strong> one hydrophilic head-bead <strong>and</strong> one tail <strong>of</strong> hydrophobic beads. A more<br />

realistic model <strong>of</strong> a phospho<strong>lipid</strong> can be constructed by connecting two hydrophobic<br />

tails to an headgroup consisting <strong>of</strong> one or more head-beads. In both single- <strong>and</strong><br />

double-tail <strong>lipid</strong>s the tail(s) can have different length. We denote a single-tail <strong>lipid</strong><br />

<strong>with</strong> one head-bead <strong>and</strong> n tail beads as htn, <strong>and</strong> a double-tail <strong>lipid</strong> <strong>with</strong> m headbeads<br />

<strong>and</strong> n tail-beads as hm(tn)2 (see figure 2.1).<br />

Figure 2.1: Schematic drawing <strong>of</strong> the single- <strong>and</strong> double-tail model <strong>lipid</strong>s described in the text<br />

<strong>and</strong> their nomenclature. The black particles represent the hydrophilic head-beads <strong>and</strong> the<br />

white particles the hydrophobic tail-beads.<br />

A mapping <strong>of</strong> coarse-grained <strong>lipid</strong>s onto real phospho<strong>lipid</strong>s can be established<br />

through the factor Nm. We have chosen a mapping factor <strong>of</strong> Nm = 3, corresponding<br />

to three water molecules represented by one DPD-bead <strong>of</strong> volume <strong>of</strong> 90 ˚A 3 . In terms


2.2 Coarse-grained model <strong>of</strong> <strong>lipid</strong>s 13<br />

<strong>of</strong> methyl groups <strong>of</strong> an actual <strong>lipid</strong> molecule, this volume corresponds to three CH2<br />

groups (or one CH2 plus one CH3 group). By also mapping the choline, phosphate,<br />

<strong>and</strong> glycerol groups <strong>of</strong> the phospho<strong>lipid</strong> hydrophilic head on one DPD bead each, a<br />

real phospho<strong>lipid</strong>, like, for example, dimyristoylphosphatidylcholine (DMPC), can<br />

be represented by a double-tail coarse-grained <strong>lipid</strong> <strong>with</strong> three hydrophilic headbeads<br />

<strong>and</strong> five hydrophobic beads in each tail (<strong>lipid</strong> h3(t5)2), as shown in figure 2.2.<br />

13<br />

12<br />

11<br />

H3 C<br />

H3 C<br />

9<br />

H 10<br />

2 H<br />

C<br />

2 H<br />

C<br />

2 H<br />

C C 2 H<br />

C<br />

2<br />

H2 C H<br />

C 2<br />

H2 C C<br />

H2 C<br />

C<br />

H2<br />

H2 C<br />

H2<br />

H2 H2 H2 H2 H2 H2 C C C C C C<br />

C C C C C C C<br />

H2 H2 H2 H2 H2 H2<br />

O<br />

O<br />

C<br />

O<br />

O H2 C<br />

CH<br />

CH2 3<br />

O<br />

8<br />

7<br />

6<br />

5<br />

4<br />

O −<br />

P<br />

O<br />

H<br />

O 2<br />

C<br />

C N<br />

H2 +<br />

CH3 CH3 CH 3<br />

2 1<br />

Figure 2.2: The atomistic representation <strong>of</strong> DMPC <strong>and</strong> the corresponding coarse-grained<br />

model used in this work. Hydrophilic head-beads are indicated in gray <strong>and</strong> hydrophobic tailbeads<br />

in white.<br />

Model parameters<br />

Non-bonded interactions<br />

The non-bonded interactions between the beads are described by the conservative<br />

force <strong>of</strong> equation 2.2. The relative strength <strong>of</strong> the force between different bead types<br />

is represented by setting the repulsion parameter between two beads–either both hydrophilic<br />

or hydrophobic– to a smaller value than the one between two beads where<br />

one is hydrophilic <strong>and</strong> the other is hydrophobic.<br />

In particular, the interaction parameter between water-beads (aww) can be derived<br />

by fitting the calculated value <strong>of</strong> the compressibility <strong>of</strong> water at room temperature<br />

to the experimental one [24, 28], according to<br />

1<br />

kBT<br />

<br />

∂p<br />

=<br />

∂ρ sim<br />

Nm<br />

kBT<br />

<br />

∂p<br />

∂n exp<br />

(2.11)<br />

where Nm is the number <strong>of</strong> water molecules that are represented by one DPD bead, p<br />

is the pressure, <strong>and</strong> n <strong>and</strong> ρ are the water <strong>and</strong> DPD water-like bead densities, respectively.<br />

Groot <strong>and</strong> Warren [28] have shown that for densities ρ > 2, the equation <strong>of</strong><br />

state <strong>of</strong> a DPD single-component system for different densities <strong>and</strong> different repulsion<br />

parameters follows a simple scaling relation. Since the higher the system density<br />

the larger the number <strong>of</strong> interactions for each particle, it is convenient to choose the


14 Simulation method for coarse-grained <strong>lipid</strong>s<br />

minimum density for which the scaling relation holds, i.e. ρ = 3. For this value <strong>of</strong> the<br />

density, the correct compressibility <strong>of</strong> water is matched for aww = 25NmkBTo. In this<br />

fitting procedure a temperature <strong>of</strong> T ∗ = 1.0 (where the star indicates the temperature<br />

in reduced units) corresponds to room temperature To. In principle, the same procedure<br />

could be used to match the compressibility <strong>of</strong> water at different temperatures.<br />

This would result, however, in temperature dependent aij parameters. Such temperature<br />

dependent parameters would make the interpretation <strong>of</strong> the results more<br />

complex <strong>and</strong> therefore we assume that these parameters are temperature independent.<br />

Hence in our simulations we have chosen to keep the parameters fixed <strong>and</strong><br />

only change the temperature whenever we use a temperature different than the room<br />

temperature. Since the translation to experimental temperatures is not straightforward,<br />

in Chapter 5 we will convert the reduced temperatures into physical units by a<br />

mapping based on the temperatures <strong>of</strong> the phase transitions in <strong>lipid</strong> <strong>bilayers</strong>.<br />

To obtain the repulsion parameters for a multicomponent system, where beads<br />

<strong>of</strong> different type are present, mutual solubilities can be matched by relating the DPD<br />

repulsion parameters to the Flory-Huggins χ-parameters that represent the excess<br />

free energy <strong>of</strong> mixing [24, 28, 43].<br />

This route to derive repulsion parameters for mixed DPD systems has been applied<br />

by various authors. However, the set <strong>of</strong> mesoscopic parameters used in the<br />

literature to model amphiphilic surfactants, <strong>and</strong> in particular phospho<strong>lipid</strong>s, is not<br />

unique. Different levels <strong>of</strong> coarse-graining result in different parameter sets [24, 28,<br />

46]. The ionic nature <strong>of</strong> the molecules [24, 46] can produce different values for the<br />

interactions. Also, the interactions parameters have been tuned to reproduce the<br />

self-assembly <strong>of</strong> amphiphilic surfactants into spherical micelles [46], or into a bilayer<br />

<strong>with</strong> thickness <strong>and</strong> <strong>lipid</strong> end-to-end distance consistent <strong>with</strong> experimental values<br />

[50]. Despite these differences, it can be shown that if the repulsion between hydrophobic<br />

<strong>and</strong> hydrophilic particles is sufficiently larger than the repulsion between<br />

particles <strong>of</strong> the same type, phase separation occurs, <strong>and</strong> that, under the appropriate<br />

conditions <strong>of</strong> concentration <strong>and</strong> amphiphiles architecture, the self-assembled<br />

supra-molecular structure is a bilayer. We use the parameter set derived by Groot<br />

for amphiphilic surfactants [46] <strong>and</strong> reported in table 2.2, <strong>with</strong> the exception <strong>of</strong> att<br />

(tail-tail), which we have increased from 15 to 25 to avoid unrealistic high densities<br />

in the bilayer hydrophobic core. The low repulsion between water-beads <strong>and</strong> head-<br />

aij w h t<br />

w 25 15 80<br />

h 15 35 80<br />

t 80 80 25<br />

Table 2.1: Repulsion parameters aij (see equation 2.2) used in our simulations. Water beads<br />

are indicated as w, hydrophilic head-beads as h, <strong>and</strong> hydrophobic tail-beads as t. The parameters<br />

are in units <strong>of</strong> kBT.


2.2 Coarse-grained model <strong>of</strong> <strong>lipid</strong>s 15<br />

beads a wh represents the hydration <strong>of</strong> the headgroup by water molecules. The higher<br />

value <strong>of</strong> the head-head repulsion parameter a hh, respect to the water-water aww <strong>and</strong><br />

tail-tail att repulsion parameters, takes into account the bulkier volume <strong>of</strong> the headgroup<br />

<strong>and</strong> the repulsion due to the charged nature <strong>of</strong> the <strong>lipid</strong> headgroup.<br />

The value aww = 25 for the interaction parameter between water beads, reported<br />

in table 2.2, gives the correct compressibility <strong>of</strong> water at room temperature if a mapping<br />

factor <strong>of</strong> Nm = 1 is used. However, it is worth to point out that the interaction<br />

parameters have an effective nature, <strong>and</strong> that the relative strength <strong>of</strong> the interactions<br />

between different beads is the factor that mainly determines the properties <strong>of</strong><br />

the mesoscopic bilayer. In the following <strong>of</strong> this thesis we will show that a mapping<br />

factor <strong>of</strong> Nm = 3 can also be used in combination <strong>with</strong> the parameter set <strong>of</strong> table<br />

2.2, <strong>and</strong> that the properties <strong>of</strong> the resulting coarse-grained mesoscopic <strong>bilayers</strong> can<br />

be consistently compared <strong>with</strong> experimental data. In particular, in Chapter 3, we will<br />

show that the bilayer area compressibility has a value comparable <strong>with</strong> the area compressibility<br />

<strong>of</strong> real phospho<strong>lipid</strong> <strong>bilayers</strong>; <strong>and</strong>, in Chapter 5, that the bilayer structural<br />

quantities, like the area per <strong>lipid</strong> <strong>and</strong> bilayer thickness, for the coarse-grained <strong>bilayers</strong>,<br />

are in remarkably good quantitative agreement <strong>with</strong> experimental data.<br />

Bonded interactions<br />

The beads that form a <strong>lipid</strong> molecule are connected via harmonic springs. Groot <strong>and</strong><br />

Warren [28] used a harmonic force in the form<br />

fspring = krij<br />

(2.12)<br />

<strong>with</strong> a spring constant k = 2. This spring force, however, does not prevent the beads<br />

to be located far more than a cut-<strong>of</strong>f radius, Rc, apart. In this way it would be easy<br />

for the <strong>lipid</strong> chains to cross each other <strong>with</strong>out experiencing any mutual interaction.<br />

This means that, due to the s<strong>of</strong>t interactions between DPD beads, the model can not<br />

simulate entanglement if the spring in equation 2.12 is used. This problem can be<br />

controlled by adjusting the length <strong>and</strong> intensity <strong>of</strong> the spring [51]. To avoid bond<br />

crossing we use a spring force in the form <strong>of</strong> a Hookean spring<br />

Fspring = Kr(rij − req)^rij. (2.13)<br />

The equilibrium distance, giving the chosen number density <strong>of</strong> 3, is req = 0.7. The<br />

force constant Kr is chosen <strong>with</strong> a value 100, which guarantees that 98% <strong>of</strong> the cumulative<br />

bond distance distribution lies <strong>with</strong>in one Rc. To compare the two spring<br />

<strong>models</strong>, in figure 2.3 we plot the total energy given by<br />

Utot = UC + Uspring, (2.14)


16 Simulation method for coarse-grained <strong>lipid</strong>s<br />

where UC is the energy due to the force F C in equation 2.2:<br />

<br />

r<br />

UC = a r − 1 +<br />

2Rc<br />

Rc<br />

<br />

2<br />

(2.15)<br />

<strong>and</strong> the value <strong>of</strong> the repulsion parameter is chosen as a = 25kBT. Uspring in equation<br />

2.14 is either the potential from the spring force in equation 2.12 or in equation 2.13.<br />

It can be seen that using the potential from equation 2.12 <strong>and</strong> k = 2 (dashed line in<br />

figure 2.3) the minimum <strong>of</strong> the total potential energy is very broad <strong>and</strong> located at a<br />

value <strong>of</strong> the interparticle distance ro ≈ 0.9, which is almost the value <strong>of</strong> the cut-<strong>of</strong>f<br />

radius (Rc = 1). The same potential <strong>with</strong> k = 100 (long-dashed line in figure 2.3)<br />

presents a deeper minimum but results in a value <strong>of</strong> the interparticle equilibrium<br />

distance ro ≈ 0.2 which is too small compared to the average distance between the<br />

beads in a system at density ρ = 3, while the Hookean spring <strong>of</strong> equation 2.13 (solid<br />

line) <strong>with</strong> the chosen parameterization, gives the correct bond length.<br />

U tot (r)=U C (r)+U spring (r)<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

0.0 0.2 0.4 0.6 0.8 1.0<br />

r/Ro Figure 2.3: Comparison <strong>of</strong> the total potential energy Utot (see equation 2.14 <strong>and</strong> 2.15) for the<br />

two spring <strong>models</strong> in equation 2.12 (dashed line k=2, long-dashed line k=100) <strong>and</strong> 2.13 (solid<br />

line Kr=100).<br />

The flexibility <strong>of</strong> the <strong>lipid</strong> can also be controlled by adding an extra bond-bending<br />

potential between consecutive bonds in the form:<br />

U bend = Kθ<br />

2<br />

= Kθ<br />

2<br />

(θ − θo) 2<br />

<br />

ri−1,i · ri+1,i<br />

− θo<br />

|ri−1,i||ri+1,i|<br />

2<br />

(2.16)<br />

where Kθ is the strength <strong>of</strong> the potential, θ is the angle between two consecutive<br />

bonds (ri−1,i <strong>and</strong> ri+1,i), <strong>and</strong> θo is the equilibrium angle. If no angle is set in the<br />

hydrophobic tail, the <strong>lipid</strong> is fully flexible. A resistance to bending can be modeled by


2.3 Self-assembly 17<br />

setting the equilibrium angle <strong>of</strong> the bond-bending potential to the value θo = 180 o .<br />

This topology can represent saturated carbon chains or unsaturated all-trans chains.<br />

To model a cis-unsaturation due to a double bond, an equilibrium angle θo = 135 o<br />

can be chosen since cis-unsaturation creates a kink in the <strong>lipid</strong> tail [52].<br />

2.3 Self-assembly<br />

The scope <strong>of</strong> this section is to show that DPD coarse-grained <strong>lipid</strong>s, modeled <strong>with</strong><br />

the interaction parameters <strong>of</strong> table 2.2, can demix <strong>and</strong> form self-assembled structures.<br />

To study the self-assembly process we consider mixtures <strong>of</strong> water <strong>and</strong> singletail<br />

<strong>lipid</strong>s <strong>with</strong> five tail-beads (ht5). The <strong>lipid</strong> tail is fully flexible, i.e. no bond-bending<br />

potential is applied.<br />

We consider mixtures <strong>of</strong> water <strong>and</strong> <strong>lipid</strong>s at different mole fraction, cs, defined as<br />

NL<br />

cs =<br />

NL + Nw<br />

, (2.17)<br />

where NL <strong>and</strong> Nw are the total number <strong>of</strong> <strong>lipid</strong>s <strong>and</strong> water particles in the system,<br />

respectively. For each <strong>of</strong> these systems the number density — defined as the total<br />

number <strong>of</strong> beads in the system divided by the simulation box volume — was ρ = 3,<br />

<strong>and</strong> the box lengths were fixed as Lx = Ly = 10 <strong>and</strong> Lz = 12, <strong>with</strong> periodic boundary<br />

conditions in all three Cartesian directions. For each concentration we started <strong>with</strong><br />

a r<strong>and</strong>om initial configuration <strong>of</strong> the <strong>lipid</strong>s in water <strong>and</strong> we let the system evolve<br />

until the equilibrium state was reached, i.e. until the supramolecular aggregate <strong>of</strong><br />

the <strong>lipid</strong>s was stable in time.<br />

For the simulations presented in this section we have chosen a temperature <strong>of</strong><br />

kBT = 1. We will show in Chapter 5 that the parameters in table 2.2 lead to stable<br />

<strong>bilayers</strong> in a wide range <strong>of</strong> temperatures.<br />

Results <strong>and</strong> discussion<br />

At low concentrations one small spherical micelle is formed. By increasing the concentration<br />

the size <strong>of</strong> the spherical micelle increases until a concentration <strong>of</strong> 0.044<br />

is reached, when a cylindrical micelle is formed. This cylindrical micelle is stable up<br />

to cs ≈ 0.057, then, at cs = 0.062, it deforms into a flat cylindrical micelle. When<br />

the number <strong>of</strong> <strong>lipid</strong>s is further increased by just 10 more molecules (corresponding<br />

to a concentration <strong>of</strong> cs = 0.067), a bilayer <strong>with</strong> a pore is formed. To obtain a complete<br />

bilayer a concentration <strong>of</strong> cs = 0.077 must be reached, corresponding to 200<br />

<strong>lipid</strong>s. In the first stage <strong>of</strong> the bilayer formation the <strong>lipid</strong>s aggregate in a cluster. This<br />

cluster then takes the shape <strong>of</strong> a cylinder. The cylindrical micelle is not stable <strong>and</strong><br />

shows large fluctuations in shape. Some <strong>of</strong> these fluctuations result in a percolation<br />

<strong>of</strong> the micelle across the periodic boundary conditions <strong>and</strong> a bilayer is formed. It


18 Simulation method for coarse-grained <strong>lipid</strong>s<br />

takes between 10000 <strong>and</strong> 20000 time steps for a bilayer to self-assembly, <strong>and</strong> the bilayer,<br />

once formed, is stable <strong>and</strong> symmetric, i.e. it contains approximately the same<br />

number <strong>of</strong> <strong>lipid</strong>s in each monolayer. A bilayer is the stable structure up to 210 <strong>lipid</strong>s.<br />

By further increase in the concentration, the bilayer starts to branch <strong>and</strong> deforms in<br />

a shapeless structure tough still interconnected. For 300 <strong>lipid</strong>s (cs = 0.14) a second,<br />

separate, bilayer starts to form, parallel to the first one. In figure 2.4 we show some <strong>of</strong><br />

the described structures as snapshots taken from the simulations.<br />

Our results are in remarkably good agreement <strong>with</strong> the results by Goetz <strong>and</strong> Lipowsky<br />

[21]. These authors studied the self-assembly <strong>of</strong> linear CG Lennard-Jones<br />

amphiphilic surfactants — <strong>with</strong> the same topology as our DPD single-tail <strong>lipid</strong>s — by<br />

molecular dynamics <strong>and</strong> Monte Carlo simulations. We find the same self-assembled<br />

structures that these authors find, <strong>and</strong> the same type <strong>of</strong> structure is formed in the<br />

same range <strong>of</strong> concentrations in the two studies. A small difference between our<br />

study <strong>and</strong> the one in ref. [21] is that in our simulations the transition from one type <strong>of</strong><br />

structure to the other (like, for example, from spherical micelle to cylindrical micelle)<br />

occurs at slightly lower <strong>lipid</strong>/water ratios than observed by Goetz <strong>and</strong> Lipowsky. This<br />

is likely due to the fact that the amphiphiles in [21] have four beads in the tail <strong>and</strong><br />

the ones we use have five, while both <strong>models</strong> have the same headgroup size (one<br />

bead). As a consequence, closed spherical structures (like micelles <strong>and</strong> cylindrical<br />

micelles) that contain the same number <strong>of</strong> molecules become unstable at lower concentrations<br />

for amphiphiles <strong>with</strong> a longer tail. The hydrophilic surface is the same<br />

in the two cases, but the hydrophobic volume is larger for amphiphiles <strong>with</strong> a longer<br />

tail. To verify this hypothesis we studied the self-assembly <strong>of</strong> <strong>lipid</strong>s <strong>with</strong> four tailbeads<br />

(ht4) in the same range <strong>of</strong> concentrations used for the ones <strong>with</strong> five tail-beads.<br />

We indeed found that, for these shorter chains, the transitions between the different<br />

supramolecular aggregates occur at higher concentrations compared to the ones <strong>of</strong><br />

longer <strong>lipid</strong>s. These concentrations are still slightly lower than the corresponding<br />

ones reported in [21].<br />

It is also interesting to compare out results <strong>with</strong> the recent simulation <strong>of</strong> Marrink<br />

et al. [6] whose work is the first simulation <strong>of</strong> aggregation <strong>of</strong> phospho<strong>lipid</strong>s into <strong>bilayers</strong><br />

<strong>with</strong> atomic details <strong>of</strong> the structure <strong>and</strong> interactions. These authors find that a bilayer<br />

<strong>with</strong> an hydrophilic pore, like the one observed in our <strong>and</strong> Goetz <strong>and</strong> Lipowsky’s<br />

CG simulations, is the intermediate configuration in the formation <strong>of</strong> the bilayer. The<br />

breakdown <strong>of</strong> the pore is the rate limiting step in the overall process. Although in our<br />

simulations the bilayer <strong>with</strong> a pore is stable <strong>and</strong> it is observed at concentrations too<br />

low for a full bilayer to be formed, the fact that this structure has been produced using<br />

such different <strong>models</strong> <strong>and</strong> simulation techniques suggests that it could indeed be<br />

a fundamental step in the kinetics <strong>of</strong> bilayer self-assembly.


2.3 Self-assembly 19<br />

(a) (b)<br />

(c) (d)<br />

(e)<br />

Figure 2.4: Snapshots <strong>of</strong> self-assembled structure <strong>of</strong> ht5 <strong>lipid</strong>s. Following increasing concentration:<br />

large spherical micelle (a), cylindrical micelle (b), flat cylindrical micelle (c), bilayer<br />

<strong>with</strong> a pore (d), <strong>and</strong> bilayer (side <strong>and</strong> top view) (e,f). The black beads represent the headgroups,<br />

the gray beads the tails, <strong>and</strong> the dots the water particles.


III<br />

Surface tension in <strong>lipid</strong> <strong>bilayers</strong>


22 Surface tension in <strong>lipid</strong> <strong>bilayers</strong><br />

3.1 Introduction<br />

Lipid <strong>bilayers</strong> are self-assembled structures, which are not constrained by the total<br />

area, <strong>and</strong> hence will adopt a conformation that will have the lowest free energy. Since<br />

the thermodynamic definition <strong>of</strong> surface tension is the derivative <strong>of</strong> the free energy<br />

<strong>with</strong> respect to the area <strong>of</strong> the interface [53], for a unconstrained bilayer the free energy<br />

minimum will be a tensionless state [54]. Experimental results on unilamellar<br />

vesicles have also indicated that <strong>bilayers</strong> are in a stress free state [55]. In molecular<br />

simulation, for both self-assembled <strong>and</strong> pre-assembled membranes, a fixed number<br />

<strong>of</strong> <strong>lipid</strong> molecules <strong>and</strong> a fixed area are usually combined <strong>with</strong> periodic boundary<br />

conditions. The periodic boundary conditions correspond to an infinitely large<br />

membrane, but the fixed size <strong>of</strong> the simulation box, <strong>and</strong> the fixed number <strong>of</strong> <strong>lipid</strong>s<br />

at the interface, impose a constraint on the bilayer area which results in a finite surface<br />

tension. Although the constraint on the fixed area can be released by performing<br />

simulations <strong>of</strong> membranes at constant pressure or constant surface tension [56, 57],<br />

it is an important — <strong>and</strong> still open — question which value <strong>of</strong> the surface tension<br />

should be used in simulations to reproduce the state <strong>and</strong> the area per <strong>lipid</strong> <strong>of</strong> a real<br />

membrane.<br />

In their molecular dynamics simulations Feller <strong>and</strong> Pastor [58,59] observed that a<br />

tensionless state did not reproduce the experimental value <strong>of</strong> the area per <strong>lipid</strong>. They<br />

explained this result by considering that, since the typical undulations <strong>and</strong> <strong>and</strong> out<strong>of</strong>-plane<br />

fluctuations <strong>of</strong> a macroscopic membrane do not develop in a small patch <strong>of</strong><br />

a membrane (theirs was composed <strong>of</strong> 36 <strong>lipid</strong>s for monolayer), a positive surface tension<br />

(stretching) must be imposed in order to compensate for the suppressed undulations,<br />

<strong>and</strong> hence to recover the experimental values <strong>of</strong> the area per <strong>lipid</strong>. Recently,<br />

Marrink <strong>and</strong> Mark [60] investigated the system size dependence <strong>of</strong> the surface tension<br />

in membrane patches ranging from 200 to 1800 <strong>lipid</strong>s, simulated for times up<br />

to 40ns. Their calculations show that, in a stressed membrane, the surface tension<br />

is size dependent, i.e. it drops if the system size is increased (at fixed area per <strong>lipid</strong>),<br />

which is in agreement <strong>with</strong> the results by Feller <strong>and</strong> Pastor. On the other h<strong>and</strong>, their<br />

results show that at zero stress simulation conditions the equilibrium does not depend<br />

on the system size. Marrink <strong>and</strong> Mark then concluded that simulations at zero<br />

surface tension correctly reproduce the experimental surface areas for a stress free<br />

membrane.<br />

Simulations at constant surface tension have been introduced by Chiu et al. in<br />

[56]. A constant surface tension ensemble (NVγ) has been considered in literature<br />

<strong>and</strong> the corresponding equations <strong>of</strong> motion for Molecular Dynamics simulations<br />

have been derived [57], <strong>and</strong> applied to the simulation <strong>of</strong> phospho<strong>lipid</strong> <strong>bilayers</strong> [59,<br />

61–65]. Alternatively, to ensure a tensionless state, Goetz <strong>and</strong> Lipowsky [21] performed<br />

several simulations to determine the area per <strong>lipid</strong> that gives a state <strong>of</strong> zero<br />

tension.


3.2 Method <strong>of</strong> calculation <strong>of</strong> surface tension 23<br />

In our simulations we use a different approach, based on a Monte Carlo (MC)<br />

scheme, to simulate a membrane at a given state <strong>of</strong> tension, <strong>of</strong> which the tensionless<br />

state is a particular case. Similar to constant pressure simulation, we impose a given<br />

value for the surface tension <strong>and</strong> from the simulation we obtain the average area per<br />

molecule. Such method has the advantage that we do not have to perform several<br />

– relatively expensive – simulations to locate the area <strong>of</strong> zero tension. Moreover, by<br />

releasing the constraint <strong>of</strong> a priori chosen area <strong>and</strong> allowing dynamic fluctuations<br />

<strong>of</strong> the bilayer area, the system is able to explore the phase space <strong>and</strong> assume the<br />

configurational structure that corresponds to the free energy minimum at the given<br />

thermodynamic conditions. One <strong>of</strong> the implications <strong>of</strong> this extra degree <strong>of</strong> freedom is<br />

that it allows to observe directly phase transitions in which the area per <strong>lipid</strong> changes.<br />

We will exploit this advantage in Chapter 5 where we study the phase behavior <strong>of</strong> <strong>lipid</strong><br />

<strong>bilayers</strong>.<br />

In this Chapter we first introduce the definition <strong>of</strong> surface tension <strong>and</strong> its calculation<br />

method in computer simulations. We then describe the scheme we use to<br />

impose a given value <strong>of</strong> the surface tension <strong>and</strong> we validate the method by applying<br />

it to a monolayer <strong>of</strong> amphiphilic dumb-bells at oil-water interface. We then utilize<br />

the constant surface tension scheme in simulations <strong>of</strong> <strong>lipid</strong> <strong>bilayers</strong> to study the dependence<br />

<strong>of</strong> the area per <strong>lipid</strong> on the system size, <strong>and</strong> on the value <strong>of</strong> the applied<br />

surface tension.<br />

3.2 Method <strong>of</strong> calculation <strong>of</strong> surface tension<br />

In homogeneous systems at equilibrium the pressure is constant <strong>and</strong> equal at each<br />

point in space, while for an inhomogeneous system the pressure is a tensor P(r) that<br />

depends on the spatial direction <strong>and</strong> on the position r where it is calculated. Here<br />

we follow the discussion presented in [53] <strong>and</strong> [66] to derive the properties <strong>of</strong> the<br />

pressure tensor.<br />

Consider a system <strong>of</strong> two immiscible liquids forming a planar interface normal to<br />

the z direction. In equilibrium, mechanical stability requires that the gradient <strong>of</strong> the<br />

pressure tensor is zero everywhere<br />

∇ · P = 0. (3.1)<br />

Shear forces are also zero <strong>and</strong> the non diagonal components <strong>of</strong> P vanish; also, because<br />

<strong>of</strong> planar symmetry, the components <strong>of</strong> the pressure tensor parallel to the interface<br />

should be identical. The pressure tensor is then diagonal<br />

⎛<br />

P = ⎝<br />

Pxx 0 0<br />

0 Pyy 0<br />

0 0 Pzz<br />

⎞<br />

⎠ , (3.2)


24 Surface tension in <strong>lipid</strong> <strong>bilayers</strong><br />

<strong>with</strong><br />

From equations 3.1 <strong>and</strong> 3.2 we have<br />

Pxx(r) = Pyy(r). (3.3)<br />

∂Pxx<br />

∂x ex + ∂Pyy<br />

∂y ey + ∂Pzz<br />

∂z ez = 0 (3.4)<br />

where eα (α = x, y, z) are the orthogonal basis vectors in the Cartesian space. From<br />

equations 3.4 <strong>and</strong> 3.3 results that the lateral (also called tangential) components <strong>of</strong><br />

the pressure tensor are function <strong>of</strong> z only: PL(z) = Pxx(z) = Pyy(z). The normal<br />

component is constant throughout the system <strong>and</strong> equal to the external pressure:<br />

PN(z) = Pzz(z) = Pext. The lateral components are also equal to the external pressure<br />

in the bulk phases.<br />

The surface tension is defined as the integral over the interface <strong>of</strong> the difference<br />

between the normal <strong>and</strong> lateral components <strong>of</strong> the pressure tensor [67, 68]<br />

γ =<br />

z2<br />

z1<br />

dz [PN(z) − PL(z)] =<br />

z2<br />

z1<br />

dγ(z) (3.5)<br />

where z1 <strong>and</strong> z2 are positions in the bulk phases <strong>and</strong> γ(z) is the local surface tension<br />

at position z.<br />

In molecular simulations, statistical mechanics is used to relate thermodynamic<br />

quantities to ensemble averages over microscopic degrees <strong>of</strong> freedom. For an homogeneous<br />

system the pressure is a scalar <strong>and</strong> it can be expressed by the virial equa-<br />

tion [69, 70]<br />

P = ρkBT + 1<br />

<br />

N<br />

<br />

ri · Fi<br />

3V<br />

i=1<br />

(3.6)<br />

where Fi is the total internal force on particle i <strong>and</strong> the brackets indicate an ensemble<br />

average. If the intermolecular forces are pairwise additive, the above may be written<br />

as<br />

P = ρkBT + 1<br />

3V<br />

<br />

<br />

i<br />

j>i<br />

where Fij is the force on particle i due to particle j.<br />

rij · Fij<br />

<br />

(3.7)<br />

For an inhomogeneous system the pressure tensor at position r can still be expressed<br />

in a tensorial form <strong>of</strong> the virial equation <strong>and</strong> it can be split in a kinetic P K<br />

<strong>and</strong> a potential part P U [66],<br />

P(r) = P K (r) + P U (r) (3.8)


3.2 Method <strong>of</strong> calculation <strong>of</strong> surface tension 25<br />

The kinetic part can be expressed as a generalization <strong>of</strong> the ideal gas contribution<br />

P K (r) = kBT ρ(r)^1 (3.9)<br />

where ρ(r) is density at position r <strong>and</strong> ^1 is the 3x3 unit matrix. This kinetic part is a<br />

single particle property <strong>and</strong> it is well localized in space. Conversely, there is no unambiguous<br />

way <strong>of</strong> expressing the potential part <strong>of</strong> the pressure tensor. P U (r) can be<br />

defined as the force acting across a microscopic element <strong>of</strong> area located at r. Because<br />

the force depends on the position <strong>of</strong> two particles (for pair additive potentials), there<br />

is no unique way to determine which pairs <strong>of</strong> particles should contribute to the pressure<br />

across the microscopic element <strong>of</strong> area at r [71], <strong>and</strong> to reduce the non local<br />

two-particles force to a local force at r. Irving <strong>and</strong> Kirkwood [72] derived the equations<br />

<strong>of</strong> hydrodynamics by means <strong>of</strong> classical statistical mechanics <strong>and</strong> obtained the<br />

expression <strong>of</strong> the pressure tensor in terms <strong>of</strong> molecular variables. They required that,<br />

in any definition <strong>of</strong> the pressure tensor, the local virial should be located near the line<br />

connecting the two interacting particles.<br />

Different methods have been proposed to compute the potential part <strong>of</strong> the pressure<br />

tensor, like the Irving <strong>and</strong> Kirkwood method [72] or the Harashima method [73].<br />

The various definitions correspond to different choices <strong>of</strong> the contour which connects<br />

the position in which the microscopic pressure tensor is calculated <strong>with</strong> the<br />

particles position. It is important to underline that all methods give the same expression<br />

for the total pressure <strong>and</strong> interfacial tension when integrated over the whole<br />

system, while the expression <strong>of</strong> the local pressure depends on the applied method.<br />

For a detailed description <strong>of</strong> these different methods see [66]. In our simulations we<br />

use the Kirkwood-Buff convention [67, 74], which takes as a contour a straight line.<br />

The simulation box <strong>of</strong> sizes Lx, Ly <strong>and</strong> Lz, is divided into Ns slabs parallel to the interface<br />

(xy-plane) <strong>and</strong> the contribution <strong>of</strong> each pair <strong>of</strong> interacting particles to the<br />

local pressure tensor is evenly split through all the slabs which intersect the line the<br />

connects the two particles (line <strong>of</strong> centers).<br />

The normal <strong>and</strong> lateral components <strong>of</strong> the local pressure tensor in slab k, including<br />

the kinetic contribution, are then given by<br />

PL(k) = kBT 〈ρ(k)〉 − 1<br />

<br />

<br />

2Vs<br />

(i,j)<br />

PN(k) = kBT 〈ρ(k)〉 − 1<br />

<br />

<br />

Vs<br />

(k) x2 ij + y2 ij<br />

u<br />

rij<br />

′ (rij)<br />

(k) z<br />

(i,j)<br />

2 ij<br />

u<br />

rij<br />

′ (rij)<br />

<br />

<br />

(3.10)<br />

(3.11)<br />

where ρ(k) is the average density in slab k, Vs = LxLyLz/Ns is the volume <strong>of</strong> a slab,<br />

u ′ (r) is the derivative <strong>of</strong> the intramolecular potential, <strong>and</strong> the brackets denote an ensemble<br />

average. (k)<br />

(i,j) means that the summation runs over all pairs <strong>of</strong> particles (i, j)<br />

<strong>of</strong> which the slab k (partially) contains the line <strong>of</strong> centers. A slab k gets a contribu-


26 Surface tension in <strong>lipid</strong> <strong>bilayers</strong><br />

tion 1/No from a given pair (i, j), where No is the total number <strong>of</strong> slabs that intersect<br />

the line <strong>of</strong> centers between particle i <strong>and</strong> j. Periodic boundary conditions are always<br />

taken into account in this calculation. The integral in equation 3.5 now becomes<br />

Ns<br />

Ns<br />

<br />

<br />

γ = [PN(k) − PL(k)]∆zs = γ(k) (3.12)<br />

k=1<br />

where ∆zs = Lz/Ns is the (uniform) width <strong>of</strong> the slabs <strong>and</strong> γ(k) is the local surface<br />

tension in slab k.<br />

To illustrate the application <strong>of</strong> this method <strong>with</strong> a very simple example, we plot in<br />

figure 3.1 the distribution <strong>of</strong> the local surface tension γ(z) across an oil/water interface<br />

(the interface is perpendicular to the z-axis), calculated from a DPD simulation<br />

<strong>of</strong> oil- <strong>and</strong> water-like phase separated particles. In the same figure we also plot the<br />

oil, water <strong>and</strong> total densities pr<strong>of</strong>iles to better characterize the shape <strong>of</strong> the pressure<br />

pr<strong>of</strong>ile. Note that, because <strong>of</strong> periodic boundary conditions, the interfaces in<br />

the simulation box are actually two. The maximum <strong>of</strong> the surface tension is at the<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

−0.2<br />

γ(z)<br />

oil water<br />

k=1<br />

0 2 4 6 8 10<br />

Z<br />

Figure 3.1: Surface tension pr<strong>of</strong>ile γ(z) (solid black), <strong>and</strong> oil (dashed black), water (dot-dashed<br />

black), <strong>and</strong> total (solid gray) density pr<strong>of</strong>iles across the interfaces (perpendicular to the z-axis)<br />

between phase separated oil <strong>and</strong> water. Because <strong>of</strong> periodic boundary conditions, there are<br />

two interfaces. Note that the densities have been rescaled <strong>and</strong> shifted for better comparison<br />

<strong>with</strong> the surface tension (hence the scale in the graph ordinate refers to the surface tension<br />

only).<br />

oil/water interfaces where there are large repulsive forces between oil <strong>and</strong> water, <strong>and</strong><br />

the density is lower than in the bulk phases. Moving away from the interface (in<br />

both directions since the system is fully symmetric), as a consequence <strong>of</strong> the repulsive<br />

forces at the interface, there is a thin region <strong>of</strong> compressed fluid where the surface<br />

tension becomes (slightly) negative. The surface tension becomes again positive<br />

oil


3.3 Constant surface tension ensemble 27<br />

where a second density minimum occurs <strong>and</strong> then goes to zero in the bulk phases.<br />

In the literature, when describing the distribution <strong>of</strong> pressures in a <strong>lipid</strong> bilayer,<br />

the quantity that is usually reported is the difference between the lateral <strong>and</strong> the normal<br />

pressure. Hence, we find it convenient to define here the pressure pr<strong>of</strong>ile π(z)<br />

as<br />

π(z) = [PL(z) − PN(z)]. (3.13)<br />

3.3 Constant surface tension ensemble<br />

In this section we introduce a Monte-Carlo (MC) scheme to impose a constant surface<br />

tension in the presence <strong>of</strong> a planar interface [75].<br />

Consider a system <strong>with</strong> constant number <strong>of</strong> particles N, constant temperature<br />

T, <strong>and</strong> constant volume V, in which an interface <strong>of</strong> area A is present. The interface<br />

gives an additional term in the energy <strong>of</strong> the system, i.e. the energy associated <strong>with</strong><br />

the creation <strong>of</strong> the interface, which is expressed by the surface tension γ times the<br />

area <strong>of</strong> the interface. The work done on the system by compressing or stretching the<br />

interface by dA, is given by [53] dW = γdA. The partition function for such a system<br />

can be written as<br />

Q =<br />

1<br />

Λ3N <br />

dr<br />

N! V<br />

N exp −β(U(r N ) − γA) . (3.14)<br />

where U denotes the potential energy, γ the surface tension, A the area <strong>of</strong> the interface,<br />

<strong>and</strong> β = 1/kBT.<br />

L⊥<br />

z<br />

¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡<br />

¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡<br />

x<br />

L||<br />

A<br />

y<br />

L||<br />

Figure 3.2: Schematic representation <strong>of</strong> a simulation box for a system <strong>with</strong> a flat interface parallel<br />

to the xy-plane. The area <strong>of</strong> the interface is A = L 2<br />

<strong>and</strong> the box dimension perpendicular<br />

to the interface (z axis) is L⊥.<br />

Consider a simulation box (see figure 3.2) <strong>with</strong> edges Lx = Ly = L parallel to


28 Surface tension in <strong>lipid</strong> <strong>bilayers</strong><br />

the interface (xy plane), <strong>and</strong> Lz = L⊥ perpendicular to the interface (z axis), so that<br />

the system volume is V = L⊥L2 <strong>and</strong> the area <strong>of</strong> the interface A = L2 . We define a<br />

transformation <strong>of</strong> the box sizes which changes the area <strong>and</strong> the height but keeps the<br />

volume constant. Such a transformation can be written in the form<br />

L ′ = λ L (3.15)<br />

L ′ ⊥ = 1<br />

L⊥<br />

λ2 where λ is the parameter <strong>of</strong> the transformation. By changing λ, the above expression<br />

generates a transformation <strong>of</strong> coordinates which preserves the total volume <strong>of</strong> the<br />

system, hence no work against the external pressure is performed. The coordinate<br />

phase space has now an extra degree <strong>of</strong> freedom represented by the parameter λ.<br />

To write the partition function corresponding to this ensemble it is convenient to<br />

introduce a set <strong>of</strong> scaled coordinates s ∈ [0, 1], defined as<br />

r = (L sx, L sy, L⊥sz). (3.16)<br />

By a transformation <strong>of</strong> the box sizes <strong>with</strong> λ (equation 3.15) the coordinates <strong>of</strong> the<br />

particles rescale as<br />

r ′ <br />

= λLsx, λLsy, 1<br />

<br />

L⊥sz .<br />

λ2 (3.17)<br />

In terms <strong>of</strong> these scaled coordinates the partition function <strong>of</strong> the system takes the<br />

expression<br />

Q = VN<br />

Λ3N <br />

dλ ds<br />

N!<br />

N exp −β U(s N ; λ) − γA(λ) . (3.18)<br />

The probability <strong>of</strong> finding a configuration <strong>with</strong> scaled positions s N <strong>and</strong> parameter λ<br />

is then given by [76]<br />

N(s N , λ) ∝ exp −β U(s N ; λ) − γA(λ) . (3.19)<br />

In a MC move an attempt <strong>of</strong> changing the parameter λ is then accepted <strong>with</strong> a probability<br />

Pacc(λ → λ ′ <br />

exp −β U(s<br />

) =<br />

′N<br />

<br />

′ ′ ; λ ) − γA(λ )<br />

exp {−β [U(sN ; λ) − γA(λ)]}<br />

(3.20)<br />

where γ is the imposed surface tension. If we choose the particular value γ = 0,<br />

then the explicit term depending on the area in equation 3.20 drops. The described<br />

scheme can be applied to impose any value <strong>of</strong> the surface tension. It is important to<br />

remark that this scheme assumes that the stress tensor is diagonal, which is true for<br />

fluid systems. Since, as we have discussed in Chapter 2, an Hamiltonian for the conservative<br />

part <strong>of</strong> the interaction energy in DPD can be defined, it is possible to imple-


3.3 Constant surface tension ensemble 29<br />

ment the described MC scheme in combination <strong>with</strong> DPD simulations. The method<br />

combines DPD to evolve the positions <strong>of</strong> the particles <strong>and</strong> MC moves to change the<br />

shape <strong>of</strong> the simulation box [75]. A simulation done <strong>with</strong> this hybrid method consists<br />

<strong>of</strong> cycles. In each cycle we choose at r<strong>and</strong>om whether to perform a number<br />

<strong>of</strong> DPD steps or an attempt to change the parameter λ (i.e. the box aspect ratio) by<br />

δλ. The number <strong>of</strong> DPD steps to perform when a DPD-type move is chosen is also<br />

selected at r<strong>and</strong>om between one <strong>and</strong> a maximum number <strong>of</strong> steps Nmax. We have<br />

chosen Nmax = 50. The fraction <strong>of</strong> accepted box shape moves was set at 30%, <strong>and</strong><br />

δλ is automatically adjusted at regular intervals during the simulation to match this<br />

percentage. Unless otherwise stated, these are the values used in all the simulations<br />

presented in this thesis.<br />

Validation <strong>of</strong> the constant surface tension scheme<br />

To validate the constant surface tension scheme, we performed simulations <strong>of</strong> a monolayer<br />

<strong>of</strong> amphiphilic dumb-bell surfactants at water-oil interface. The system consist<br />

<strong>of</strong> 1400 oil-like beads (o), 1400 water-like beads (w), <strong>and</strong> 200 dumb-bell surfactants<br />

built by connecting via a spring one water-like bead to one oil-like bead. The DPD<br />

repulsion parameter between beads <strong>of</strong> the same type was chosen as aoo = aww = 25,<br />

<strong>and</strong> the one between different beads as aow = 80. The overall number density <strong>of</strong> the<br />

system was ρ = 3. We considered the monolayer <strong>with</strong> two different initial values <strong>of</strong><br />

the area, corresponding to a surface tensions <strong>of</strong> γ = 0 <strong>and</strong> γ = 4, respectively. In<br />

both systems we imposed a surface tension <strong>of</strong> γ = 2 <strong>and</strong> calculated the area <strong>and</strong> the<br />

surface tension as function <strong>of</strong> MC cycles. The results are plotted in figure 3.3. In both<br />

cases the imposed value <strong>of</strong> the surface tension was already achieved after only 1000<br />

MC cycles, <strong>and</strong> the area <strong>of</strong> both systems converged to the same value.<br />

100<br />

85<br />

A<br />

70<br />

55<br />

3<br />

γ 2<br />

1<br />

0<br />

0 1000 2000 3000 4000 5000<br />

MC cycles<br />

Figure 3.3: Area, A, <strong>and</strong> surface tension, γ, as function <strong>of</strong> MC cycles. The two sets <strong>of</strong> lines<br />

correspond to different initial conditions: area A = 53.8 <strong>and</strong> surface tension γ = 0 (black<br />

lines), area A = 101.4 <strong>and</strong> surface tension γ = 4 (gray lines). The plotted values <strong>of</strong> the surface<br />

tension are running averages <strong>of</strong> length 100.


30 Surface tension in <strong>lipid</strong> <strong>bilayers</strong><br />

3.4 Surface tension in <strong>lipid</strong> <strong>bilayers</strong><br />

To investigate the system size dependence <strong>of</strong> the area per <strong>lipid</strong> on the surface tension,<br />

we consider <strong>bilayers</strong> formed by 100, 200, 400, 900, <strong>and</strong> 1600 <strong>lipid</strong>s to which we<br />

impose different values <strong>of</strong> the surface tension using the above described hybrid DPD-<br />

MC scheme. The area per <strong>lipid</strong>, AL, is calculated in our simulations as the bilayer<br />

projected area divided by half the number <strong>of</strong> <strong>lipid</strong>s that form the bilayer, considered<br />

that, on average, there is an equal number <strong>of</strong> <strong>lipid</strong>s in each monolayer.<br />

Computational details<br />

The <strong>lipid</strong> model used for these simulations is a double-tail <strong>lipid</strong> <strong>with</strong> three headbeads<br />

<strong>and</strong> five hydrophobic beads in each <strong>of</strong> the tails, denoted as h3(t5)2, <strong>and</strong> shown<br />

in figure 3.4.<br />

Figure 3.4: Schematic representation <strong>of</strong> the model <strong>lipid</strong> h3(t5)2, used in the simulations presented<br />

in this section. The black particles represent the hydrophilic head-beads <strong>and</strong> the gray<br />

particles the hydrophobic tail-beads.<br />

The interaction parameters between the beads are the ones described in section<br />

2.2 <strong>of</strong> Chapter 2. Two consecutive beads in the <strong>lipid</strong> are connected by harmonic<br />

springs (equation 2.13) <strong>with</strong> spring constant Kr = 100 <strong>and</strong> equilibrium distance<br />

ro = 0.7. To control the tail flexibility, a bond-bending potential (equation<br />

2.16) between two consecutive bonds in the <strong>lipid</strong> tails was added <strong>with</strong> bending constant<br />

Kθ = 6 <strong>and</strong> equilibrium angle θo = 180 o . An additional bond-bending potential<br />

was applied between the vectors connecting the tails to the headgroup, <strong>with</strong> Kθ = 3<br />

<strong>and</strong> θo = 90 o .<br />

The <strong>lipid</strong> h3(t5)2, <strong>with</strong> the chosen interaction parameters, self-assemble into a<br />

stable bilayer. First a self-assembled bilayer <strong>of</strong> 100 <strong>lipid</strong>s <strong>and</strong> 2500 water particles,<br />

corresponding to a fully hydrated bilayer, was formed by using DPD steps only, then<br />

the bilayer was replicated in the x <strong>and</strong> y directions (plane <strong>of</strong> the bilayer) to form <strong>bilayers</strong><br />

<strong>of</strong> different size: i.e. <strong>of</strong> NL=100, 200, 400, 900, <strong>and</strong> 1600 <strong>lipid</strong>s. In each case, 25<br />

water particles per <strong>lipid</strong> were considered, resulting in a maximum total number <strong>of</strong><br />

beads <strong>of</strong> 60800, in the case <strong>of</strong> the bilayer formed by 1600 <strong>lipid</strong>s. For each considered<br />

system the overall density was ρ = 3 <strong>and</strong> the reduced temperature T ∗ =1. To each


3.4 Surface tension in <strong>lipid</strong> <strong>bilayers</strong> 31<br />

<strong>of</strong> these <strong>bilayers</strong> different values <strong>of</strong> the surface tension were then imposed by applying<br />

the hybrid MC-DPD scheme, <strong>with</strong> a probability <strong>of</strong> 70% <strong>of</strong> performing DPD steps<br />

(chosen at r<strong>and</strong>om <strong>with</strong> a maximum <strong>of</strong> 50), or to attempt a change <strong>of</strong> the box aspect<br />

ratio. The total number <strong>of</strong> MC-DPD cycles was 100,000 for each simulation <strong>and</strong> the<br />

instantaneous values <strong>of</strong> the area <strong>and</strong> surface tension were sampled every 50 cycles.<br />

Negative values <strong>of</strong> applied surface tension, γ, correspond to a compression <strong>of</strong> the<br />

bilayer; γ = 0, corresponds to a tensionless, unconstrained bilayer; <strong>and</strong> positive values<br />

to a stretched bilayer. For each <strong>of</strong> these values <strong>of</strong> surface tension <strong>and</strong> for each <strong>of</strong><br />

the bilayer sizes considered, the bilayer configuration was the stable state throughout<br />

the simulation.<br />

Results <strong>and</strong> discussion<br />

The instantaneous value <strong>of</strong> the area per <strong>lipid</strong> for a bilayer <strong>of</strong> 400 <strong>lipid</strong>s is shown in<br />

figure 3.5(a), as function <strong>of</strong> MC cycles <strong>and</strong> at three different values <strong>of</strong> imposed surface<br />

tension (one positive, one negative <strong>and</strong> one zero). All the systems start from the<br />

same value <strong>of</strong> the area per <strong>lipid</strong> <strong>and</strong>, after about 20,000 MC cycles, have converged<br />

to the average values: 〈AL〉 = 1.50 ± 0.03 for imposed γ = −2, 〈AL〉 = 1.66 ± 0.02<br />

for imposed γ = 0, <strong>and</strong> 〈AL〉 = 1.83 ± 0.02 for imposed γ = 2. The fluctuations <strong>of</strong><br />

the computed surface tension are very large, though decreasing in magnitude <strong>with</strong><br />

increasing value <strong>of</strong> the imposed surface tension. For example, at imposed value <strong>of</strong><br />

γ = −2, the calculated value <strong>of</strong> the surface tension was -2.0 ± 1.4, at imposed γ = 0,<br />

the calculated surface tension was 0.0 ± 1.3, <strong>and</strong> at imposed γ = 2, the calculated<br />

surface tension was 2.0 ± 1.1. Despite these large fluctuations, the average value <strong>of</strong><br />

the surface tension is equal to the imposed one, as can be seen from the running<br />

averages plotted in figure 3.5(b).<br />

A further verification that the equilibrium area per <strong>lipid</strong> does correspond to the<br />

imposed surface tension, was obtained by making the histogram <strong>of</strong> the sampled values<br />

<strong>of</strong> the area at equilibrium, <strong>and</strong> by calculating the corresponding average surface<br />

tension in each bin. In figure 3.6 three <strong>of</strong> these histograms for the bilayer <strong>of</strong> 400 <strong>lipid</strong>s<br />

are shown, relative to negative, zero <strong>and</strong> positive values <strong>of</strong> the imposed surface tension,<br />

respectively. Each <strong>of</strong> the histograms is computed after the area has reached<br />

its equilibrium value, hence the width <strong>of</strong> the distribution <strong>of</strong> the area per <strong>lipid</strong> corresponds<br />

to the spontaneous fluctuations at equilibrium.<br />

The average values <strong>of</strong> γ show a linear increase <strong>with</strong> increasing area. The larger the<br />

area, the more stretched is the membrane, corresponding to higher values <strong>of</strong> the surface<br />

tension, <strong>and</strong> in correspondence <strong>with</strong> the maximum in the area distribution the<br />

surface tension is equal to the imposed value. These trends are more pronounced for<br />

positive <strong>and</strong> zero imposed surface tension (graphs (b) <strong>and</strong> (c) in figure 3.6), while in<br />

the case <strong>of</strong> negative imposed surface tension (graph (a)) there are deviations from a<br />

linear dependence <strong>of</strong> the surface tension on the area. The dependence <strong>of</strong> the area per<br />

<strong>lipid</strong> on system size, NL, <strong>and</strong> on imposed surface tension, γ, is shown in figure 3.8.


32 Surface tension in <strong>lipid</strong> <strong>bilayers</strong><br />

A L<br />

1.9<br />

1.8<br />

1.7<br />

1.6<br />

1.5<br />

1.4<br />

0e+00 2e+04 4e+04 6e+04 8e+04 1e+05<br />

MC cycles<br />

(a)<br />

γ= 2<br />

γ= 0<br />

γ=−2<br />

γ<br />

2.0<br />

0.0<br />

−2.0<br />

0e+00 2e+04 4e+04 6e+04 8e+04 1e+05<br />

MC cycles<br />

Figure 3.5: (a) Instantaneous value <strong>of</strong> the area per <strong>lipid</strong>, AL, as function <strong>of</strong> MC cycles for three<br />

values <strong>of</strong> imposed surface tension γ. In figure (b) the corresponding running averages (<strong>of</strong><br />

length 100) <strong>of</strong> the instantaneous value <strong>of</strong> the surface tension are plotted. The straight lines<br />

parallel to the graph abscissa correspond to the imposed values <strong>of</strong> the surface tension. The<br />

data refer to a bilayer <strong>of</strong> 400 h3(t5)2 <strong>lipid</strong>s.<br />

The area per <strong>lipid</strong> increases <strong>with</strong> increasing value <strong>of</strong> the imposed surface tension,<br />

for all the system sizes. Furthermore, the area per <strong>lipid</strong> at positive or zero surface<br />

tension does not depend on the system size, while if the bilayer is compressed (negative<br />

values <strong>of</strong> γ), a size dependence <strong>of</strong> the area is found, <strong>and</strong> the (projected) area per<br />

<strong>lipid</strong> decreases <strong>with</strong> increasing system size. For clarity reasons, the projected area per<br />

<strong>lipid</strong> for NL = 900 <strong>and</strong> NL = 1600 at the lowest surface tension considered, γ = −2,<br />

are not shown in figure 3.8, but they were found to be about two times smaller than<br />

the value for NL = 400. The reason <strong>of</strong> these deviations becomes clear by looking at<br />

instantaneous snapshots <strong>of</strong> the two largest <strong>bilayers</strong> (900 <strong>and</strong> 1600 <strong>lipid</strong>s) at negative<br />

values <strong>of</strong> the surface tension, as shown in figure 3.7. In large bilayer patches, like<br />

the ones shown in the figure, the lateral compression activates bending modes in<br />

the bilayer: the bilayer is not flat anymore, but shows out-<strong>of</strong>-plane undulations. The<br />

amplitude <strong>of</strong> these undulations increases <strong>with</strong> increasing system size <strong>and</strong> <strong>with</strong> decreasing<br />

surface tension. These bending modes are completely suppressed in small<br />

bilayer patches <strong>of</strong> 100 <strong>and</strong> 200 <strong>lipid</strong>s, <strong>and</strong> partially suppressed in bilayer patches <strong>of</strong><br />

400 <strong>lipid</strong>s. Considering that, in most atomistic-detailed molecular dynamics simulations<br />

<strong>of</strong> bilayer membranes, the typical number <strong>of</strong> <strong>lipid</strong>s considered is usually<br />

smaller than 200, care must be used when choosing boundary conditions that compress<br />

the bilayer. On the other h<strong>and</strong>, our results suggest that, for stretched or tensionless<br />

bilayer, small bilayer patches can be considered <strong>with</strong> no finite-size effects.<br />

This result partly contradicts the findings <strong>of</strong> Marrink <strong>and</strong> Mark [60] discussed in the<br />

Introduction to this Chapter. These authors found a system size dependence not only<br />

(b)


3.4 Surface tension in <strong>lipid</strong> <strong>bilayers</strong> 33<br />

γ<br />

1.0<br />

0.0<br />

−1.0<br />

−2.0<br />

−3.0<br />

P(A L )<br />

−4.0<br />

=1.50<br />

−5.0<br />

1.42 1.44 1.46 1.48 1.50<br />

AL 1.52 1.54 1.56<br />

(a) γ = −2<br />

γ<br />

3.0<br />

2.0<br />

1.0<br />

0.0<br />

−1.0<br />

P(A L )<br />

−2.0<br />

=1.66<br />

−3.0<br />

1.60 1.62 1.64 1.66<br />

AL 1.68 1.70 1.72<br />

(b) γ = 0<br />

γ<br />

5.0<br />

4.0<br />

3.0<br />

2.0<br />

1.0<br />

P(A L )<br />

0.0<br />

=1.83<br />

−1.0<br />

1.76 1.78 1.80 1.82 1.84<br />

AL 1.86 1.88 1.90<br />

(c) γ = 2<br />

Figure 3.6: Histograms <strong>of</strong> the area per <strong>lipid</strong> AL <strong>and</strong> corresponding average value <strong>of</strong> the calculated<br />

surface tension γ (filled circles) in each bin. The three plots refer to a bilayer <strong>of</strong> 400<br />

h3(t5)2 <strong>lipid</strong>s <strong>with</strong> different values <strong>of</strong> the imposed surface tension: (a) γ = −2, (b) γ = 0, <strong>and</strong><br />

(c) γ = 2. The values on the graph ordinate refer to γ only, while the height <strong>of</strong> the bins in the<br />

histogram is proportional to P(AL), i.e. to the probability distribution <strong>of</strong> the area per <strong>lipid</strong>. The<br />

basis line <strong>of</strong> each <strong>of</strong> the histograms corresponds to the imposed value <strong>of</strong> γ. The average values<br />

<strong>of</strong> the area per <strong>lipid</strong>, 〈AL〉, are also reported in the graphs. The error on these values is ±0.02.<br />

at negative values <strong>of</strong> the surface tension, but also for stretched <strong>bilayers</strong>, <strong>and</strong> only tensionless<br />

<strong>bilayers</strong> were found to display no finite-size effects. It must be pointed out<br />

that the more complex <strong>and</strong> long range (electrostatic) interactions used in atomistic<br />

MD, <strong>and</strong> not represented in our mesoscopic model, might have an influence on the<br />

calculation <strong>of</strong> the surface tension <strong>and</strong> on its dependence on system size.<br />

A L<br />

2.1<br />

1.9<br />

1.7<br />

1.5<br />

N L =100<br />

N L =200<br />

N L =400<br />

N L =900<br />

N L =1600<br />

1.3<br />

−3 −2 −1 0 1 2 3 4<br />

γ<br />

Figure 3.8: Average values <strong>of</strong> the area per <strong>lipid</strong> at equilibrium, AL, as function <strong>of</strong> applied surface<br />

tension γ, for bilayer patches <strong>of</strong> different sizes, i.e. number <strong>of</strong> <strong>lipid</strong>s NL. The lines connecting<br />

the symbols are only a guide to the eye.


34 Surface tension in <strong>lipid</strong> <strong>bilayers</strong><br />

(a) NL = 900, γ = −1<br />

(b) NL = 1600, γ = −1<br />

(c) NL = 900, γ = −2 (d) NL = 1600, γ = −2<br />

Figure 3.7: Snapshots <strong>of</strong> the equilibrium configurations <strong>of</strong> compressed <strong>bilayers</strong>, i.e. <strong>with</strong> imposed<br />

negative surface tension, γ, for two, large, system sizes (number <strong>of</strong> <strong>lipid</strong>s in the bilayer,<br />

NL). (a) NL = 900, <strong>and</strong> γ = −1. (b) NL = 1600, <strong>and</strong> γ = −1. (c) NL = 900, <strong>and</strong> γ = −2. (d)<br />

NL = 1600, <strong>and</strong> γ = −2. Two periodic images (in the x direction) <strong>of</strong> each bilayer are shown<br />

to better display the undulation modes. Note that in (b) the bilayer undulations are in both x<br />

<strong>and</strong> y directions, while in (c) <strong>and</strong> (d) they are only in the x direction. The <strong>lipid</strong> headgroups are<br />

depicted in black, <strong>and</strong> the <strong>lipid</strong> tails in gray, <strong>with</strong> the terminal tail beads (located in the bilayer<br />

center) in darker gray. The water is not shown.


3.4 Surface tension in <strong>lipid</strong> <strong>bilayers</strong> 35<br />

Area compressibility<br />

From the dependence <strong>of</strong> the area per <strong>lipid</strong> on the surface tension, the bilayer area<br />

compressibility can be calculated. The bilayer area compressibility KA is defined as<br />

[77]<br />

Integrating equation 3.21 we have<br />

KA = A<br />

γ = KA ln A<br />

<br />

∂γ<br />

. (3.21)<br />

∂A<br />

Ao<br />

(3.22)<br />

where Ao is the area at the free energy minimum, i.e. at zero surface tension. Exp<strong>and</strong>ing<br />

equation 3.22 around Ao we obtain a linear dependence <strong>of</strong> γ on A<br />

γ ≈ KA<br />

(A − Ao). (3.23)<br />

Ao<br />

In simulation, the area compressibility can be calculated from the above equation,<br />

either by fixing the area <strong>of</strong> the bilayer <strong>and</strong> computing the corresponding surface tension,<br />

or by imposing the surface tension <strong>and</strong> calculating the area. In this study we<br />

use the latter method. If the area is taken as the dependent variable, then, from 3.23,<br />

we have<br />

A = Ao<br />

γ + Ao<br />

(3.24)<br />

KA<br />

To fit the area as function <strong>of</strong> γ using equation 3.24, we used two approaches. The<br />

first one consists in computing the area at different values <strong>of</strong> imposed surface tension,<br />

in the regime were the area dependence on surface tension is linear. We found<br />

the behavior <strong>of</strong> the average projected area per <strong>lipid</strong> to be linear in the applied surface<br />

tension for values <strong>of</strong> the tension in the range γ=[-0.5,1.5]. The second method<br />

consists in considering the spontaneous fluctuations at equilibrium, <strong>of</strong> both the area<br />

<strong>and</strong> the measured surface tension, in a tensionless bilayer, as computed in the histogram<br />

shown in figure 3.6(b). The results <strong>of</strong> both fitting methods are in very good<br />

agreement: in both cases we find KAR 2 c/kBT=20.8 from the slope <strong>of</strong> the fitted lines.<br />

Also, a value <strong>of</strong> 1.66 R 2 c is found as the intercept, which is the same value <strong>of</strong> the area<br />

at zero surface tension, Ao, as directly measured from the simulation at γ = 0.<br />

Using Rc=6.4633 ˚A <strong>and</strong> taking kBT as room temperature, i.e. kBT = 4.14 10 −21 J, we<br />

can convert the area compressibility into physical units, resulting in KA ≈ 210 mN/m.<br />

This value is comparable to the value <strong>of</strong> the compressibility equal to 300 mN/m found<br />

by Lindahl <strong>and</strong> Edholm [77] in atomistic MD simulations <strong>of</strong> <strong>lipid</strong> <strong>bilayers</strong>, <strong>and</strong> to<br />

the values measured by micropipette aspiration experiments which, for saturated<br />

phosphatidylcholine <strong>bilayers</strong> in the fluid phase, are also in the range 230-240 mN/m<br />

[78].


36 Surface tension in <strong>lipid</strong> <strong>bilayers</strong><br />

3.5 Conclusion<br />

In this Chapter we have introduced the definition <strong>of</strong> surface tension, <strong>and</strong> we have<br />

described the basic formalism <strong>and</strong> computational technique to compute pressure<br />

pr<strong>of</strong>iles (i.e. the local values <strong>of</strong> the pressure) in molecular simulations.<br />

We have discussed the necessity <strong>of</strong> performing simulations <strong>of</strong> <strong>lipid</strong> <strong>bilayers</strong> in an<br />

ensemble where the surface tension is one <strong>of</strong> the thermodynamic variables that can<br />

be controlled. To this purpose, we have developed an efficient MC based scheme to<br />

impose a given value <strong>of</strong> the surface tension, <strong>of</strong> which the zero value, that corresponds<br />

to the natural state <strong>of</strong> tension <strong>of</strong> unconstrained <strong>bilayers</strong>, is a particular case.<br />

We have then applied this MC scheme to investigate the dependence <strong>of</strong> the area<br />

per <strong>lipid</strong> on the value <strong>of</strong> the applied surface tension <strong>and</strong> on the system size. We<br />

observed that out-<strong>of</strong>-plane undulations are suppressed in small patches <strong>of</strong> a compressed<br />

bilayer (i.e. negative surface tension), while they are not in larger patches. As<br />

a consequence, finite-size effects on the projected area per <strong>lipid</strong> in compressed <strong>bilayers</strong><br />

are significant. However, for tensionless or stretched (positive surface tension)<br />

<strong>bilayers</strong> we do not observe any finite-size effect.<br />

We have also shown that the area per <strong>lipid</strong> has a linear dependence on the surface<br />

tension for values <strong>of</strong> the latter close to the free energy minimum, i.e. zero surface tension.<br />

From this linear dependence the bilayer area compressibility was derived, <strong>and</strong><br />

compared <strong>with</strong> experimental <strong>and</strong> MD simulation values, as well as <strong>with</strong> the value<br />

<strong>of</strong> area compressibility derived from the spontaneous area fluctuations at equilibrium<br />

in a tensionless bilayer. Good quantitative agreement between these values<br />

was found.


IV<br />

Structural characterization <strong>of</strong> <strong>lipid</strong> <strong>bilayers</strong>


38 Structural characterization <strong>of</strong> <strong>lipid</strong> <strong>bilayers</strong><br />

4.1 Introduction<br />

An important aspect in coarse-grained (CG) <strong>models</strong> for <strong>lipid</strong> <strong>bilayers</strong> is the level <strong>of</strong><br />

chemical <strong>and</strong> molecular detail chosen to represent the components <strong>of</strong> the system.<br />

This choice is determined by the properties one wants to investigate, <strong>and</strong> on how<br />

these properties depend on the details <strong>of</strong> the model. In this Chapter we address these<br />

questions by studying the effect <strong>of</strong> modifications in the topology <strong>of</strong> the bilayer constituents<br />

(i.e. the model <strong>lipid</strong>s) on the structural properties <strong>of</strong> the bilayer. Examples<br />

<strong>of</strong> <strong>lipid</strong> characteristics that can be varied in our CG model are the length <strong>of</strong> the acylchain,<br />

the size <strong>of</strong> the <strong>lipid</strong> headgroup, or the chain stiffness. Also, we will consider<br />

model <strong>lipid</strong>s <strong>with</strong> one or two hydrophobic tails, the latter better resembling a real<br />

phospho<strong>lipid</strong>.<br />

While for some aspects the lack <strong>of</strong> molecular detail implicit in a simplified representation<br />

can be seen as the limitation <strong>of</strong> CG <strong>models</strong>, on the other h<strong>and</strong> it allows<br />

to identify some general features that might be responsible for the structure <strong>and</strong> behavior<br />

<strong>of</strong> phospho<strong>lipid</strong> <strong>bilayers</strong>. Of particular interest is the characterization <strong>of</strong> the<br />

pressure pr<strong>of</strong>ile in <strong>lipid</strong> <strong>bilayers</strong>. It is well known, both from experiments <strong>and</strong> simulations<br />

[12, 79], that <strong>bilayers</strong> are very different than simple bulk hydrocarbon/water<br />

interfaces, in that they exhibit an internal structure. The <strong>lipid</strong>s are oriented, <strong>with</strong> the<br />

headgroups sticking in the water phase, while the tails extend into the bilayer core,<br />

<strong>and</strong> segments <strong>of</strong> the <strong>lipid</strong> chains are located at different depths in the bilayer. This<br />

“ordering” <strong>of</strong> the <strong>lipid</strong>s results in a characteristic, non-uniform, density distribution<br />

in the bilayer. This inhomogeneity in the internal structure is also reflected in the<br />

distribution <strong>of</strong> the lateral pressure across the bilayer.<br />

Structural properties <strong>of</strong> <strong>lipid</strong> <strong>bilayers</strong>, such as order parameters, density pr<strong>of</strong>iles,<br />

area per <strong>lipid</strong>, <strong>and</strong> bilayer thickness, have been extensively studied <strong>and</strong> determined<br />

for a wide range <strong>of</strong> phospho<strong>lipid</strong>s, temperatures <strong>and</strong> bilayer compositions, both experimentally<br />

<strong>and</strong> by simulations. On the other h<strong>and</strong>, an exhaustive characterization<br />

<strong>of</strong> the distribution <strong>of</strong> local pressures in <strong>lipid</strong> <strong>bilayers</strong> is still lacking. Since direct experimental<br />

measurements <strong>of</strong> the pressure pr<strong>of</strong>ile in <strong>lipid</strong> <strong>bilayers</strong> are not yet available,<br />

although some attempts have been reported [80], theoretical <strong>models</strong> or computer<br />

simulation can be a valuable tool in the investigation <strong>of</strong> this quantity.<br />

Mean-field approaches have been applied to the calculation <strong>of</strong> pressure pr<strong>of</strong>iles.<br />

However, those approaches are <strong>of</strong>ten limited by the fact that they are lattice <strong>models</strong>,<br />

in which the headgroups <strong>of</strong> the <strong>lipid</strong>s have to be constrained at the interface in<br />

order to obtain a bilayer structure. Also, in such <strong>models</strong>, it is difficult to incorporate<br />

details <strong>of</strong> the chemical structure <strong>of</strong> the <strong>lipid</strong>s. By atomistic molecular dynamics<br />

(MD), many more details <strong>of</strong> the <strong>lipid</strong> chemistry <strong>and</strong> topology can be implemented,<br />

<strong>and</strong> their effect on the distribution <strong>of</strong> pressure can be studied. However, very long<br />

runs are needed for a sufficient statistical samplings, due to the large fluctuations<br />

related to this quantity. Also, to date, very few simulations on pressure pr<strong>of</strong>iles have


4.2 Structural quantities 39<br />

been reported in literature.<br />

Lindahl <strong>and</strong> Edholm were the first to compute the pressure distribution in an<br />

atomistic MD simulation <strong>of</strong> a fully hydrated dipalmitoylphosphatidylcholine (DPPC)<br />

bilayer [77]. Very recently Gullinsgrud <strong>and</strong> Schulten [81] reported an extensive MD<br />

study on the effect <strong>of</strong> changes in the <strong>lipid</strong> topology on the pressure pr<strong>of</strong>iles in <strong>lipid</strong><br />

<strong>bilayers</strong>, by considering differences in the <strong>lipid</strong> headgroup (choline or ethanolamine)<br />

<strong>and</strong> chain unsaturation. The CG approach has also been used to study pressure<br />

pr<strong>of</strong>iles. Harries <strong>and</strong> Ben-Shaul in [82] presented a study on the comparison between<br />

mean-field calculations <strong>and</strong> Monte Carlo (MC) simulations <strong>of</strong> <strong>bilayers</strong> formed<br />

by flexible linear chains <strong>of</strong> bonded identical spheres interacting <strong>with</strong> 6-12 Lennard-<br />

Jones potentials. They found good agreement in the shape <strong>of</strong> the pressure pr<strong>of</strong>iles<br />

calculated <strong>with</strong> the two approaches. Pressure pr<strong>of</strong>iles in CG model <strong>bilayers</strong> were<br />

computed by Goetz <strong>and</strong> Lipowsky [21], Shillcock <strong>and</strong> Lipowsky [50], <strong>and</strong> Groot <strong>and</strong><br />

Rabone [24]. In all the cited works, <strong>with</strong> the exception <strong>of</strong> [81], the effect <strong>of</strong> changes in<br />

the <strong>lipid</strong> topology on the shape <strong>of</strong> the pressure pr<strong>of</strong>iles has not been considered, <strong>and</strong><br />

a systematic study is still lacking.<br />

Using the DPD-CG model <strong>and</strong> the constant surface tension ensemble introduced<br />

in the previous Chapters, we investigate the effect <strong>of</strong> <strong>lipid</strong> architecture on the bilayer<br />

structure <strong>and</strong> compare our results <strong>with</strong> atomistic MD <strong>and</strong> CG simulations.<br />

First we describe the structural quantities we use to characterize a bilayer. By<br />

systematically changing chain length <strong>and</strong> stiffness <strong>of</strong> the model <strong>lipid</strong>s, we investigate<br />

how these quantities depend on the <strong>lipid</strong> architecture. We then characterize<br />

the shape <strong>of</strong> the pressure pr<strong>of</strong>ile in these different <strong>lipid</strong> <strong>bilayers</strong>, <strong>and</strong> show that the<br />

distribution <strong>of</strong> the pressure across a bilayer can be affected by modifications at specific<br />

sites in the <strong>lipid</strong> architecture. Finally, we show that the lateral pressure pr<strong>of</strong>ile<br />

in <strong>bilayers</strong> <strong>of</strong> CG <strong>lipid</strong>s <strong>with</strong> two tails is very similar in shape to the one computed in<br />

atomistic MD simulations <strong>of</strong> phosphatidylicholine <strong>bilayers</strong>.<br />

4.2 Structural quantities<br />

Orientational order parameter<br />

An important, <strong>and</strong> accurately determined property <strong>of</strong> <strong>lipid</strong> <strong>bilayers</strong>, is the orientational<br />

order parameter. This order parameter can be directly measured by deuterium<br />

substitution NMR spectroscopy [83], <strong>and</strong> is given by<br />

S = 1 <br />

2<br />

3 cos φ − 1<br />

2<br />

(4.1)<br />

where φ is the angle between the orientation <strong>of</strong> the vector along a given C-H bond<br />

<strong>and</strong> the bilayer normal. In our coarse-grained model, however, the hydrogen atoms<br />

are not present, hence we use a different definition. The mathematical expression is


40 Structural characterization <strong>of</strong> <strong>lipid</strong> <strong>bilayers</strong><br />

the same as in equation 4.1, but the angle φ is now defined as the angle between the<br />

orientation <strong>of</strong> the vector along two beads in the chain <strong>and</strong> the normal to the bilayer<br />

plane:<br />

cos φ = rij · ^n<br />

rij<br />

= zij<br />

rij<br />

(4.2)<br />

where ^n is a unit vector normal to the bilayer, <strong>and</strong> rij = ri − rj is the vector between<br />

beads i <strong>and</strong> j (rij = |rij|). The order parameter has value 1 if this vector is on average<br />

parallel to the bilayer normal, 0 if the orientation is r<strong>and</strong>om, <strong>and</strong> −0.5 if the bond is<br />

on average parallel to the bilayer plane. With this definition <strong>of</strong> the angle φ, we can<br />

compute the order parameter for a vector between any two beads in the <strong>lipid</strong>. In particular<br />

we are interested in characterizing the overall order <strong>of</strong> the chains <strong>and</strong> the local<br />

order. For the first quantity we define the indexes <strong>of</strong> the vector rij in equation 4.2 as:<br />

i = tn <strong>and</strong> j = t1, where tn is the last bead in the <strong>lipid</strong> tail <strong>and</strong> t1 is the first one. We<br />

call S tail the corresponding order parameter. For the local order we define i = tm+1<br />

<strong>and</strong> j = tm <strong>with</strong> the index m increasing going toward the tail end, <strong>and</strong> call the corresponding<br />

order parameter Sm. If m is taken progressively from the headgroup to the<br />

tail-end <strong>of</strong> the molecule, a plot <strong>of</strong> the corresponding order parameters, Sm, gives an<br />

indication <strong>of</strong> the persistence <strong>of</strong> order from the interfacial region to the bilayer core.<br />

Area per <strong>lipid</strong> <strong>and</strong> bilayer thickness<br />

The area per <strong>lipid</strong> can be experimentally estimated from X-ray or neutron scattering<br />

[84, 85], or from the just described <strong>lipid</strong> order parameter pr<strong>of</strong>iles [86]. We compute<br />

the area per <strong>lipid</strong>, AL, by dividing the total bilayer projected area by half the number<br />

<strong>of</strong> <strong>lipid</strong>s in the bilayer, since we find that, on average, there is an equal number <strong>of</strong><br />

<strong>lipid</strong>s in each monolayer.<br />

Experimentally the bilayer (total) thickness is computed from the peak-to-peak<br />

headgroup distance measured by X-ray diffraction. In simulations the same approach<br />

can also be used, <strong>and</strong> the thickness can be computed as the distance between the<br />

peaks in the density pr<strong>of</strong>ile.<br />

To compute the thickness <strong>of</strong> the bilayer hydrophobic core, Dc, we consider the average<br />

distance along the bilayer normal (which we assume to be the z-axis) between<br />

the tail bead (or beads in case <strong>of</strong> double-tail <strong>lipid</strong>s) connected to the headgroup <strong>of</strong><br />

the <strong>lipid</strong>s in one monolayer <strong>and</strong> the ones in the opposite monolayer:<br />

where zi t1<br />

(i = 1, 2).<br />

Dc = z 1 t1<br />

− z 2 t1<br />

(4.3)<br />

is the average z position <strong>of</strong> the first tail beads <strong>of</strong> the <strong>lipid</strong>s in monolayer i<br />

It is also useful to consider the <strong>lipid</strong> end-to-end distance, Lee, defined as the distance<br />

between the positions <strong>of</strong> the first bead(s), rt1 , <strong>and</strong> the last bead(s) rtn , <strong>of</strong> the


4.3 Computational details 41<br />

<strong>lipid</strong> tail(s):<br />

Lee = 〈|rtn − rt1 |〉 (4.4)<br />

<strong>and</strong> its projection, L n ee, onto the normal to the bilayer plane:<br />

L n ee = 〈|ztn<br />

− zt1 |〉 (4.5)<br />

where the bilayer is taken parallel to the xy-plane.<br />

If the bilayer consists <strong>of</strong> two opposing monolayers which are in contact by the<br />

terminal carbons in the tails, the area per <strong>lipid</strong> <strong>and</strong> bilayer thickness are related by<br />

VL = ALDc/2, where VL is the volume <strong>of</strong> one <strong>lipid</strong>; <strong>and</strong> the thickness <strong>and</strong> the (projected)<br />

end-to-end distance are related by Dc = 2L n ee. However, if the two monolayers<br />

are interdigitated, the above relations do not hold [86]. For example, for partially interdigitated<br />

<strong>bilayers</strong> it will be Dc < 2L n ee <strong>and</strong> for fully interdigitated <strong>bilayers</strong> Dc = L n ee.<br />

To investigate the presence <strong>of</strong> an interdigitated phase we define a measure for the<br />

extent <strong>of</strong> interpenetration <strong>of</strong> the hydrophobic cores (tails) <strong>of</strong> the <strong>lipid</strong>s on opposite<br />

sides <strong>of</strong> the bilayer by defining the chain overlap D overlap, as<br />

Doverlap = 2Lnee − Dc<br />

Ln . (4.6)<br />

ee<br />

where Dc <strong>and</strong> L n ee are defined in equations 4.3 <strong>and</strong> 4.4, respectively.<br />

4.3 Computational details<br />

We first study the structural properties <strong>of</strong> single-tail <strong>lipid</strong>s, in which we vary the<br />

length <strong>of</strong> the hydrophobic tail, the chain stiffness, <strong>and</strong> the headgroup interaction<br />

parameter.<br />

All the studied <strong>bilayers</strong> consist <strong>of</strong> 400 <strong>lipid</strong>s, <strong>and</strong> approximately 5000 water beads,<br />

<strong>with</strong> a total bead density <strong>of</strong> ρ = 3. The non-bonded interactions between the beads<br />

are represented by the s<strong>of</strong>t repulsion <strong>of</strong> equation 2.2, <strong>with</strong> the parameter set derived<br />

by Groot in [46] <strong>and</strong> reported in table 4.3. The reduced temperature was T ∗ = 1, <strong>and</strong><br />

at this temperature all the considered <strong>bilayers</strong> are in the fluid phase. All the <strong>bilayers</strong><br />

aij w h t<br />

w 25 15 80<br />

h 15 35 (15) 80<br />

t 80 80 25<br />

Table 4.1: Repulsion parameters aij (see equation 2.2) used in our simulations. Water beads<br />

are indicated as w, hydrophilic head beads as h <strong>and</strong> hydrophobic tail beads as t. The parameters<br />

are in units <strong>of</strong> kBT. The value in parenthesis corresponds to a repulsion parameter between<br />

the headgroups which results in a non interdigitated bilayer (see text).


42 Structural characterization <strong>of</strong> <strong>lipid</strong> <strong>bilayers</strong><br />

were first self-assembled from an initial r<strong>and</strong>om configuration <strong>of</strong> the <strong>lipid</strong>s in water.<br />

The self-assembly <strong>of</strong> the <strong>bilayers</strong> required approximately 10000 DPD steps <strong>with</strong><br />

a timestep <strong>of</strong> ∆t = 0.03. To obtain the reference state <strong>of</strong> zero surface tension, 50000<br />

hybrid MC-DPD cycles were performed <strong>with</strong> imposed surface tension γ = 0. The bilayer<br />

remained the stable state throughout the equilibration run. After equilibration,<br />

structural quantities were computed over 50000 hybrid MC-DPD cycles, <strong>with</strong> γ = 0.<br />

To illustrate the <strong>lipid</strong> nomenclature we will use in the text respect to the tail stiffness,<br />

we consider two consecutive bonds in the <strong>lipid</strong> tails, i.e. the bonds between<br />

beads i − 1, i <strong>and</strong> i, i + 1. If no bond-bending potential is defined between the bond<br />

vectors, bead i will be called t, if a bond-bending potential is defined <strong>with</strong> equilibrium<br />

angle θo = 180 o , bead i will be called t (L) , <strong>and</strong> if the equilibrium angle is set to<br />

θo = 135 o (corresponding to a point <strong>of</strong> cis-unsaturation) bead i will be called t (K) .<br />

Figure 4.1 gives an illustration <strong>of</strong> this nomenclature.<br />

t<br />

t (K)<br />

135 o<br />

t t (L) t<br />

180 o<br />

Figure 4.1: Schematic representation to illustrate the nomenclature for the <strong>lipid</strong> beads used in<br />

the text. The head bead is represented by a black particle <strong>and</strong> is denoted as ’h’. The tail beads<br />

are represented by white particles <strong>and</strong> have different names depending on the presence <strong>of</strong> a<br />

bond-bending potential. A bead labeled ’t (L) ’ is the central bead in a bond-bending potential<br />

<strong>with</strong> equilibrium angle θo = 180 o , a bead labeled ’t (K) ’ is the central bead for a bond-bending<br />

potential <strong>with</strong> θo = 135 o , while a bead labeled as ’t’ does not participate to any bond-bending<br />

potential.<br />

4.4 Results <strong>and</strong> discussion<br />

4.4.1 Density pr<strong>of</strong>iles<br />

We first consider single-tail <strong>lipid</strong>s <strong>with</strong> chain length <strong>of</strong> five beads, <strong>and</strong> study the differences<br />

in the bilayer structure between a fully flexible <strong>lipid</strong>, denoted as ht5, <strong>and</strong> a<br />

stiff one, denoted as ht (L)<br />

4 t. Both these <strong>lipid</strong>s self-assemble in a stable bilayer phase,<br />

the internal organization <strong>of</strong> the bilayer, however, strongly depends on the <strong>lipid</strong> architecture,<br />

as shown in figure 4.2. Some general features <strong>of</strong> the density distribution,<br />

such as the <strong>lipid</strong> tails confined in the inner hydrophobic core <strong>and</strong> the higher density<br />

at the interfacial region, where headgroups <strong>and</strong> water pack tightly, are in good qualitative<br />

agreement <strong>with</strong> the packing <strong>of</strong> a phospho<strong>lipid</strong> bilayer. No water permeation<br />

in the bilayer core is observed <strong>and</strong> a partial overlap <strong>of</strong> the headgroup <strong>with</strong> the first


4.4 Results <strong>and</strong> discussion 43<br />

ρ(z)<br />

3<br />

2<br />

1<br />

0<br />

−6 −4 −2 0<br />

Z<br />

2 4 6<br />

(a) ht5<br />

w<br />

h<br />

t (1,..,n−1)<br />

t n<br />

ρ tot<br />

ρ(z)<br />

3<br />

2<br />

1<br />

0<br />

−6 −4 −2 0<br />

Z<br />

2 4 6<br />

(b) ht (L)<br />

4 t<br />

w<br />

h<br />

t (1,..,n−1)<br />

t n<br />

ρ tot<br />

Figure 4.2: Density pr<strong>of</strong>iles across the bilayer as function <strong>of</strong> the distance from the bilayer center<br />

(z = 0) for two single-tail <strong>lipid</strong> <strong>models</strong>: (a) flexible <strong>and</strong> (b) stiff. The density distribution is<br />

shown for water (w, thin solid black line); headgroups (h, thick solid black line); terminal tail<br />

bead tn (thick dot-dashed black line); <strong>and</strong> the remaining tail beads t1,..n−1 (thick dashed black<br />

line). The total density ρtot (thick solid gray line) is also shown.<br />

segments <strong>of</strong> the <strong>lipid</strong> tails is present. These observations are in agreement <strong>with</strong> MD<br />

simulations <strong>of</strong> hydrated phospho<strong>lipid</strong> <strong>bilayers</strong> [63, 77, 87, 88].<br />

The typical electron density pr<strong>of</strong>iles in <strong>lipid</strong> <strong>bilayers</strong> measured experimentally<br />

[78,85], or calculated in atomistic MD simulations [62,89] show a distinct lower density<br />

in the bilayer center compared <strong>with</strong> the tightly packed region in the vicinity <strong>of</strong><br />

the headgroup. Given the coarse-grained nature <strong>of</strong> our model, <strong>and</strong> the s<strong>of</strong>t interactions<br />

between the beads, we find a larger overlap <strong>of</strong> the <strong>lipid</strong>s in the bilayer inner<br />

core compared <strong>with</strong> experimental <strong>and</strong> MD results. This overlap has different causes<br />

depending on the <strong>lipid</strong> type. Although always located in the bilayer core, the tail<br />

beads have very different distributions in the two <strong>bilayers</strong> corresponding to the different<br />

<strong>lipid</strong>s. In the bilayer formed by the flexible <strong>lipid</strong>s, the maximum density for<br />

the terminal tail bead is in the center <strong>of</strong> the bilayer, <strong>and</strong> the total density presents a<br />

small dip in the bilayer center. This indicates that the two monolayers are not very<br />

interdigitated. The distribution <strong>of</strong> the terminal tail-beads, however, shows that the<br />

<strong>lipid</strong>s in the bilayer are very disordered. The <strong>lipid</strong>s can curl, <strong>and</strong> the terminal tail<br />

beads have a non negligible probability to be found near the headgroup region <strong>of</strong><br />

the monolayer to which they belong. It should be expected that stiff <strong>lipid</strong>s, for which<br />

there is an energy barrier to the disordering <strong>of</strong> the tails, would be more localized.<br />

However, as it can be clearly seen from figure 4.2(b), the bilayer <strong>of</strong> stiff <strong>lipid</strong>s has a<br />

completely different structure in the hydrophobic core compared to the bilayer <strong>of</strong><br />

flexible <strong>lipid</strong>s. The terminal tail beads are not located in the midplane region but<br />

rather close to the headgroups <strong>of</strong> the opposing monolayer, <strong>and</strong> their density distri-


44 Structural characterization <strong>of</strong> <strong>lipid</strong> <strong>bilayers</strong><br />

bution presents a clear minimum in the center <strong>of</strong> the bilayer. This void is filled-in<br />

by the non-terminal tails beads, whose density distribution shows a maximum in the<br />

bilayer center. This distribution <strong>of</strong> densities indicates that the stiff bilayer is largely<br />

interdigitated: the molecules in one monolayer are not confined on one side <strong>of</strong> the<br />

bilayer midplane, but extend much further into the opposite monolayer. By decreasing<br />

the headgroup repulsion parameter to a hh = 15 the two monolayers separate,<br />

as can be seen from figure 4.3. The interdigitated phase is experimentally known<br />

to occur in <strong>lipid</strong> membranes below the melting temperature, although it does not<br />

spontaneously occur for double-tail <strong>lipid</strong>s <strong>with</strong> symmetric tails, but it has to be induced,<br />

for example, by changes in the environment or in the molecular structure <strong>of</strong><br />

the <strong>lipid</strong>s [90, 91]. In section 5.2 <strong>of</strong> Chapter 5, where we compute the phase diagram<br />

for single-tail <strong>lipid</strong>s, we will discuss the interdigitated phase in more detail.<br />

The effect <strong>of</strong> increasing the chain stiffness on the structure <strong>of</strong> the bilayer can be<br />

seen by comparing figure 4.3 <strong>with</strong> figure 4.2(a). Increasing the chain stiffness decreases<br />

the disorder in the <strong>lipid</strong> tails as it is also shown by the increase in the value<br />

<strong>of</strong> the order parameter S tail which is reported in table 4.2. The stiff <strong>lipid</strong>s are more<br />

aligned along the bilayer normal, <strong>and</strong> the terminal tail bead is more localized in the<br />

bilayer center, although still <strong>with</strong> a rather broad distribution. Also, the minimum in<br />

the distribution <strong>of</strong> the other tail beads (dashed line in figures 4.2 <strong>and</strong> 4.3) at the bilayer<br />

center is deeper than in the case <strong>of</strong> flexible <strong>lipid</strong>s.<br />

ρ(z)<br />

3<br />

2<br />

1<br />

0<br />

−6 −4 −2 0<br />

Z<br />

2 4 6<br />

w<br />

h<br />

t (1,..,n−1)<br />

t n<br />

ρ tot<br />

Figure 4.3: Density pr<strong>of</strong>iles for a bilayer formed <strong>of</strong> stiff <strong>lipid</strong>s <strong>with</strong> headgroup repulsion parameter<br />

ahh = 15. See also caption <strong>of</strong> figure 4.2.<br />

In table 4.2 we summarize the values <strong>of</strong> the structural properties <strong>of</strong> the three different<br />

<strong>bilayers</strong> considered. The <strong>bilayers</strong> <strong>of</strong> stiff <strong>lipid</strong>s have a larger hydrophobic thickness<br />

compared to the bilayer <strong>of</strong> flexible <strong>lipid</strong>s. Note, however, that, despite stiff <strong>lipid</strong>s have<br />

the same end-to-end length, because <strong>of</strong> interdigitation, the bilayer <strong>with</strong> a hh = 35 has<br />

a smaller thickness than the bilayer <strong>with</strong> a hh = 15. Since the volume occupied by


4.4 Results <strong>and</strong> discussion 45<br />

one <strong>lipid</strong> molecule is about the same in all cases, the area per <strong>lipid</strong> <strong>of</strong> the stiff <strong>lipid</strong>s<br />

is smaller than the area <strong>of</strong> the flexible <strong>lipid</strong>s, <strong>and</strong> the area in the interdigitated bilayer<br />

is larger than in the stiff but non-interdigitated one.<br />

<strong>lipid</strong> type S tail Dc L n ee AL<br />

ht5 (a hh = 35) 0.2 ± 0.1 2.68±0.05 0.99± 0.03 1.01±0.02<br />

ht (L)<br />

4 t (a hh = 35) 0.4 ± 0.1 3.55±0.07 2.04± 0.04 0.82±0.02<br />

ht (L)<br />

4 t (a hh = 15) 0.4 ± 0.1 4.25±0.10 2.08± 0.03 0.68±0.02<br />

Table 4.2: Values <strong>of</strong> bilayer structural properties as function <strong>of</strong> <strong>lipid</strong> type.<br />

4.4.2 Effect <strong>of</strong> chain length<br />

By varying the number <strong>of</strong> beads in the tail <strong>of</strong> the <strong>lipid</strong>s we can investigate the effect <strong>of</strong><br />

chain length on the bilayer structural properties. We consider flexible <strong>and</strong> stiff <strong>lipid</strong><br />

<strong>with</strong> hydrophobic chain lengths <strong>of</strong> 5,6,7,8, <strong>and</strong> 9 beads.<br />

The dependence <strong>of</strong> bilayer area <strong>and</strong> thickness on the <strong>lipid</strong> hydrophobic chain<br />

length has been investigated by Petrache <strong>and</strong> co-workers [92] in a 2 H NMR spectroscopy<br />

study <strong>of</strong> saturated phosphocholines (PC). These authors found that, at fixed<br />

temperature, the area per <strong>lipid</strong> slightly decreases <strong>with</strong> increasing acyl-chain length.<br />

The main effect <strong>of</strong> increasing chain length is on the bilayer thickness which increases<br />

<strong>with</strong> increasing number <strong>of</strong> hydrophobic segments in the tail.<br />

Figure 4.4 shows our results on the dependence <strong>of</strong> the area per <strong>lipid</strong> <strong>and</strong> the bilayer<br />

hydrophobic thickness on <strong>lipid</strong> chain length for the fully flexible <strong>lipid</strong>s <strong>and</strong> the<br />

stiff ones <strong>with</strong> the two headgroup repulsion parameters (a hh = 15 <strong>and</strong> a hh = 35).<br />

In the case <strong>of</strong> the flexible model, the area per <strong>lipid</strong> increases <strong>with</strong> increasing chain<br />

length. This behavior does not reproduce the experimental observed trend. With the<br />

addition <strong>of</strong> chain stiffness, however, the area per <strong>lipid</strong> for a non-interdigitated bilayer<br />

is slightly decreasing <strong>with</strong> increasing chain length, in agreement <strong>with</strong> the results in<br />

[92]; while it is approximately constant for the interdigitated bilayer.<br />

The decrease in area for longer <strong>lipid</strong>s is due to an increase <strong>of</strong> the effective packing<br />

interactions between the tails. In the case <strong>of</strong> flexible <strong>lipid</strong>s this effect is counterbalanced<br />

by the entropic effect that disorders the tails, leading to an increase in the<br />

chain cross sectional area <strong>with</strong> increasing chain length. As a consequence <strong>of</strong> the<br />

larger cross sectional area, at fixed chain length, the bilayer thickness for flexible<br />

<strong>lipid</strong>s is smaller than the thickness for stiff <strong>lipid</strong>s. The increase in bilayer thickness<br />

<strong>with</strong> increasing chain length is in agreement <strong>with</strong> the results <strong>of</strong> Petrache et al. [92].<br />

4.4.3 Lateral pressure pr<strong>of</strong>iles in tensionless <strong>bilayers</strong><br />

In this section we discuss the shape <strong>of</strong> the lateral pressure pr<strong>of</strong>ile in tensionless <strong>bilayers</strong>.<br />

We use the definition <strong>of</strong> the lateral pressure as given in equation 3.13 <strong>of</strong> Chapter


46 Structural characterization <strong>of</strong> <strong>lipid</strong> <strong>bilayers</strong><br />

A L<br />

1.2<br />

1.1<br />

1.0<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

4 5 6 7 8 9 10<br />

chain length (n)<br />

(a)<br />

flexible<br />

stiff a hh =15<br />

stiff a hh =35<br />

D c<br />

9.0<br />

8.0<br />

7.0<br />

6.0<br />

5.0<br />

4.0<br />

3.0<br />

2.0<br />

4 5 6 7 8 9 10<br />

chain length (n)<br />

(b)<br />

flexible<br />

stiff a hh =15<br />

stiff a hh =35<br />

Figure 4.4: Effect <strong>of</strong> the <strong>lipid</strong> topology (stiffness, chain length, <strong>and</strong> headgroup interaction) on<br />

the bilayer structural properties: (a) average area per <strong>lipid</strong>, AL, <strong>and</strong> (b) bilayer hydrophobic<br />

thickness, Dc. The data refer to flexible chains (full circles), stiff chains in a non interdigitated<br />

bilayer (full triangles), <strong>and</strong> stiff chains in an interdigitated bilayer (open squares). The lines are<br />

only a guide to the eye. For the thickness the error bars are smaller than the symbols size.<br />

3, i.e. π(z) = [PL(z) − PT (z)], where PL(z) <strong>and</strong> PN(z) are the lateral <strong>and</strong> normal components<br />

<strong>of</strong> the pressure tensor at position z.<br />

To better describe <strong>and</strong> underst<strong>and</strong> the distribution <strong>of</strong> lateral pressure, we find it<br />

convenient to divide the system in four regions which define three main interfaces,<br />

as illustrated in figure 4.5. A similar approach has been proposed earlier by Marrink<br />

<strong>and</strong> Berendsen [93] to describe permeation <strong>of</strong> water through a <strong>lipid</strong> membrane studied<br />

<strong>with</strong> MD. The three main interfacial regions are: the water/headgroups interface<br />

(WH interface), the headgroups/tails interface (HT interface), <strong>and</strong> the bilayer center<br />

(midplane MP).<br />

We investigate the effect <strong>of</strong> several <strong>lipid</strong> characteristics on the distribution <strong>of</strong> the<br />

pressure pr<strong>of</strong>ile. Namely:<br />

• the effect <strong>of</strong> chain stiffness, by comparing flexible <strong>and</strong> stiff <strong>lipid</strong>s;<br />

• the effect <strong>of</strong> chain packing, by comparing stiff interdigitated <strong>and</strong> not interdigitated<br />

<strong>bilayers</strong>;<br />

• the effect <strong>of</strong> changes in the head group, by comparing the two repulsion parameters<br />

used in the stiff <strong>lipid</strong>s;<br />

• the effect <strong>of</strong> tail length;<br />

• the effect <strong>of</strong> changes in the <strong>lipid</strong> structure at specific positions along the chain.<br />

We start by comparing the model <strong>lipid</strong>s considered in the previous section. The<br />

lateral pressure pr<strong>of</strong>iles for the three different <strong>bilayers</strong> are plotted in figure 4.6. The


4.4 Results <strong>and</strong> discussion 47<br />

WH<br />

HT<br />

MP<br />

Water<br />

Headgroups<br />

Tails<br />

Tails<br />

Headgroups<br />

Water<br />

£££ £££ £££ £££ £££ ££££<br />

££££ ££££ ££££ £££<br />

¢£¢£¢£¢<br />

¤£¤£¤£¤ ¥£¥£¥£¥£¥<br />

¦£¦£¦£¦<br />

§£§£§£§£§<br />

¨£¨£¨£¨ ©£©£©£©<br />

£££ £££ £££ £££ £££<br />

££££ ££££ ££££ ££££ ££££ ££££ ££££ ££££ ££££ ££££<br />

££££ ££££ ££££ ££££ ££££ ££££<br />

£££ £££ £££ £££ £££ £££ £££ £££ £££ £££ £££ £££ £££ £££<br />

£££ £££ £££ £££ £££ £££ £££<br />

£££ £££ £££ £££ £££ £££<br />

¢£¢£¢£¢<br />

££££ ££££ ££££ ££££ ££££ ££££<br />

£££ £££ £££ £££ £££ £££ £££ £££ £££ £££ £££ £££ £££<br />

£££ £££ £££ £££ £££ £££ £££ £££ £££ £££ £££ £££ £££<br />

£££ £££ £££ £££ £££ £££<br />

££££ £££ £££ £££ £££ £££<br />

£££ £££ £££ £££ £££ £££<br />

££££ ££££ ££££ ££££ ££££ ££££<br />

££££ ££££ ££££ ££££ ££££<br />

£££ £££ £££ £££ £££<br />

£££ £££ £££ £££ £££ ££££ ££££ ££££ ££££ ££££ ££££ ££££ ££££ ££££ ££££ ££££ ££££ ££££<br />

££££ ££££ ££££ ££££ ££££ ££££ ££££<br />

£££ £££ £££ £££ £££ £££ £££ £££ £££ £££ £££ £££ £££<br />

£££ £££ £££ £££ £££ £££<br />

£££<br />

Figure 4.5: Schematic representation <strong>of</strong> the four regions <strong>and</strong> three interfaces used in the text<br />

to characterize the lateral pressure distribution in a <strong>lipid</strong> bilayer. The three interfaces are the<br />

water/headgroups interface (WH), the headgroups/tails interface (HT) <strong>and</strong> the interface between<br />

the two opposing monolayers at the bilayer midplane (MP). The arrows represent the<br />

direction, <strong>and</strong> indicative magnitude, <strong>of</strong> the lateral pressure.<br />

£££ £££ £££ £££<br />

££££ ££££<br />

contributions to the lateral pressure from different potentials (non-bonded, spring<br />

<strong>and</strong> angles) are also shown. For all <strong>lipid</strong> types a positive maximum in the pressure<br />

pr<strong>of</strong>ile characterizes the region at the WH interface. The positive lateral pressure indicates<br />

that the net force in this region tends to exp<strong>and</strong> the interface, due to steric<br />

effects <strong>and</strong> hydration <strong>of</strong> the headgroups by the water. The height <strong>of</strong> this maximum<br />

is larger for stiff <strong>lipid</strong>s compared to flexible ones, <strong>and</strong> between the two stiff types<br />

the maximum is higher in the case <strong>of</strong> the interdigitated bilayer, as a consequence <strong>of</strong><br />

the higher value <strong>of</strong> the headgroup repulsion parameter. At the HT interface there<br />

is a strong inward (negative) pressure as the system attempts to limit the contact<br />

between water <strong>and</strong> hydrophobic tails. The depth <strong>of</strong> this minimum shows the same<br />

trend as the height <strong>of</strong> the maximum at the WH interface; i.e. it is deeper for stiff <strong>lipid</strong>s,<br />

<strong>and</strong> slightly deeper in the interdigitated bilayer compared to the non interdigitated<br />

one, but broader in the latter case. It is interesting to observe that the contributions<br />

to both the described peaks arise mainly from the non-bonded <strong>and</strong> spring interactions.<br />

At the sides <strong>of</strong> the box the tension goes to zero indicating that water in these<br />

regions has the characteristics <strong>of</strong> a bulk fluid, <strong>and</strong> that the bilayer is completely hydrated.<br />

The presence <strong>of</strong> the maximum at the WH interface <strong>and</strong> the minimum at the HT<br />

interface are characteristics <strong>of</strong> the pressure pr<strong>of</strong>iles also observed in the atomistic<br />

MD simulations <strong>of</strong> Lindahl <strong>and</strong> Edholm [77] (see also figure 4.11(b)) <strong>and</strong> Gullingsrud<br />

<strong>and</strong> Schulten [81]. Such characteristics are also observed in the CG Lennard-Jones<br />

<strong>models</strong> <strong>of</strong> Groot <strong>and</strong> Rabone [24] <strong>and</strong> DPD <strong>models</strong> <strong>of</strong> Goetz <strong>and</strong> Lipowsky [21] <strong>and</strong>


48 Structural characterization <strong>of</strong> <strong>lipid</strong> <strong>bilayers</strong><br />

0.2<br />

0.1<br />

0.0<br />

−0.1<br />

0.2<br />

0.1<br />

0.0<br />

−0.1<br />

0.2<br />

0.1<br />

0.0<br />

−0.1<br />

WH<br />

HT<br />

MP<br />

−5 −3 −1 1 3 5<br />

Z<br />

WH<br />

(a)<br />

−5 −3 −1 1 3 5<br />

Z<br />

WH<br />

HT MP<br />

HT<br />

(c)<br />

MP<br />

−5 −3 −1 1 3 5<br />

Z<br />

(e)<br />

w<br />

h<br />

t (1,..,n−1)<br />

t n<br />

ρ tot<br />

π(z)<br />

w<br />

h<br />

t (1,..,n−1)<br />

t n<br />

ρ tot<br />

π(z)<br />

w<br />

h<br />

t (1,..,n−1)<br />

t n<br />

ρ tot<br />

π(z)<br />

p L (z)−p N (z)<br />

p L (z)−p N (z)<br />

p L (z)−p N (z)<br />

0.8<br />

0.4<br />

0.0<br />

−0.4<br />

−0.8<br />

−5 −3 −1 1 3 5<br />

Z<br />

0.8<br />

0.4<br />

0.0<br />

−0.4<br />

(b)<br />

−0.8<br />

−5 −3 −1 1 3 5<br />

Z<br />

0.8<br />

0.4<br />

0.0<br />

−0.4<br />

(d)<br />

−0.8<br />

−5 −3 −1 1 3 5<br />

Z<br />

(f)<br />

total<br />

non bonded<br />

springs<br />

total<br />

non bonded<br />

springs<br />

bending<br />

total<br />

non bonded<br />

springs<br />

bending<br />

Figure 4.6: Lateral pressure pr<strong>of</strong>ile π(z) = pL(z) − pT (z), <strong>and</strong> contributions from the different<br />

inter <strong>and</strong> intra molecular potentials as function <strong>of</strong> the distance from the bilayer center z = 0.<br />

The figures refer to <strong>bilayers</strong> <strong>of</strong> fully flexible <strong>lipid</strong>s (top), stiff <strong>lipid</strong>s <strong>with</strong> headgroup repulsion<br />

parameter ahh = 15 (middle), <strong>and</strong> stiff <strong>lipid</strong>s <strong>with</strong> headgroup repulsion parameter ahh = 35<br />

(bottom). The density pr<strong>of</strong>iles ρ(z) for water (w), head (h), <strong>and</strong> tail (t) beads, as well as the<br />

total density, have been plotted in the same graphs to make clear the location <strong>of</strong> the maxima<br />

<strong>and</strong> minima <strong>of</strong> the pressure pr<strong>of</strong>ile <strong>with</strong>in the bilayer. The interfaces described in the text <strong>and</strong><br />

shown in figure 4.5 are also indicated. The density pr<strong>of</strong>iles have been rescaled to adjust to the<br />

pressure scale.


4.4 Results <strong>and</strong> discussion 49<br />

Shillcock <strong>and</strong> Lipowsky [50]. Another common characteristic <strong>of</strong> the pressure pr<strong>of</strong>ile<br />

which was found in the atomistic <strong>and</strong> CG simulations, <strong>and</strong> that we also observe in<br />

our simulations, is that the absolute value <strong>of</strong> the peak at HT interface is larger than<br />

the one at the WH interface. These similarities in the pressure pr<strong>of</strong>iles arise despite<br />

the different forces <strong>and</strong> parameters used in the simulations. For example, Lindahl<br />

<strong>and</strong> Edholm [77] show that the minimum at the HT interface is mainly due to electrostatic<br />

interactions, which are not present in our model nor in the CG <strong>models</strong> in<br />

the cited references. It is also worth commenting that Goetz <strong>and</strong> Lipowsky [21] <strong>and</strong><br />

Shillcock <strong>and</strong> Lipowsky [50] consider <strong>bilayers</strong> <strong>of</strong> single tail <strong>lipid</strong>s (CG Lennard-Jones<br />

MD <strong>and</strong> CG-DPD, respectively) like the ones we use here, while the results <strong>of</strong> Lindahl<br />

<strong>and</strong> Edholm [77], Gullingsrud <strong>and</strong> Schulten [81] <strong>and</strong> Groot <strong>and</strong> Rabone [24] refer to<br />

double tail <strong>lipid</strong>s (atomistic <strong>and</strong> CG-DPD, respectively). It is remarkable that different<br />

<strong>lipid</strong> topology, potentials, <strong>and</strong> parameterizations, produce the same shape <strong>of</strong> the<br />

pressure pr<strong>of</strong>iles. This suggests that the distribution <strong>of</strong> lateral pressure is more sensitive<br />

to the local densities <strong>and</strong> structure <strong>of</strong> the interfaces, rather than to the details<br />

<strong>of</strong> the forces or <strong>of</strong> the model.<br />

The major difference between the flexible <strong>and</strong> stiff <strong>lipid</strong>s considered here is in the<br />

hydrophobic core <strong>of</strong> the bilayer. A common feature is that the lateral pressure in this<br />

region is positive, showing that an outward pressure is compensating for the inward<br />

pressure at the HT interface, <strong>and</strong> that the chains tend to exp<strong>and</strong> the bilayer. However,<br />

the local structure <strong>of</strong> the pressure pr<strong>of</strong>ile depends on the <strong>lipid</strong> characteristics.<br />

For flexible <strong>lipid</strong>s first a local maximum <strong>and</strong> then a local minimum at the center <strong>of</strong><br />

the bilayer (z=0, MP interface) are present. The local minimum at the bilayer center<br />

corresponds to the region where the two monolayers are in contact, <strong>and</strong> where<br />

the density <strong>of</strong> the tail beads is lower. The shape <strong>of</strong> the pressure pr<strong>of</strong>ile for the flexible<br />

bilayer looks very similar to the ones obtained by Goetz <strong>and</strong> Lipowsky [21] <strong>and</strong><br />

Shillcock <strong>and</strong> Lipowsky [50] using MD <strong>and</strong> DPD simulations <strong>of</strong> coarse-grained <strong>lipid</strong>s,<br />

despite the fact that their results refer to stiff (via bond-bending potentials) <strong>lipid</strong>s. In<br />

our case, in the bilayer <strong>of</strong> stiff <strong>lipid</strong>s the pressure pr<strong>of</strong>ile increases monotonously to<br />

a large positive maximum in the bilayer center, which is higher in the non interdigitated<br />

bilayer. This is likely due to the fact that, because <strong>of</strong> interdigitation, the chains<br />

are more tightly <strong>and</strong> orderly packed. As a consequence <strong>of</strong> this packing order, the<br />

thermal collisions, which generate the outward pressure, are damped compared to<br />

the the case <strong>of</strong> a non interdigitated, albeit stiff, bilayer.<br />

It is interesting to observe (figures 4.6(d) <strong>and</strong> 4.6(f)) that, although the additional<br />

stiffness gives a negative contribution to the total lateral pressure, at the same time<br />

the contribution from the spring potential becomes more positive than in the case <strong>of</strong><br />

flexible <strong>lipid</strong>s. The net effect <strong>of</strong> these interactions is an increase in the values <strong>of</strong> the<br />

local pressure. The more positive values <strong>of</strong> the contribution from the spring potential<br />

are due to the fact that stiff <strong>lipid</strong>s have a larger end-to-end distance compared to<br />

flexible ones, hence the bonds along the chain are more stretched.


50 Structural characterization <strong>of</strong> <strong>lipid</strong> <strong>bilayers</strong><br />

Effect <strong>of</strong> chain length<br />

In figure 4.7 we compare the pressure pr<strong>of</strong>ile for flexible <strong>and</strong> stiff (interdigitated)<br />

<strong>lipid</strong>s <strong>of</strong> different chain lengths. We observe that in both cases <strong>with</strong> increasing chain<br />

π(z)<br />

0.1<br />

0.0<br />

−0.1<br />

−0.2<br />

−6 −4 −2 0<br />

Z<br />

2 4 6<br />

(a) flexible<br />

n=5<br />

n=7<br />

n=9<br />

π(z)<br />

0.1<br />

0.0<br />

−0.1<br />

−0.2<br />

−6 −4 −2 0<br />

Z<br />

2 4 6<br />

(b) stiff<br />

Figure 4.7: Pressure pr<strong>of</strong>ile as function <strong>of</strong> the distance z from the bilayer center (z = 0) for<br />

flexible (a) <strong>and</strong> stiff (b) <strong>lipid</strong>s <strong>with</strong> different hydrophobic chain length (n=5,7,9).<br />

length, the first maximum (WH interface) decreases in depth while the first minimum<br />

(HT interface) increases, to ensure that the integral <strong>of</strong> the pressure pr<strong>of</strong>ile (i.e.<br />

the total surface tension) remains zero. The width <strong>of</strong> both these peaks remains the<br />

same for different chain lengths since the interfacial regions to which they correspond<br />

is not chain length dependent. The pressure pr<strong>of</strong>ile in the inner hydrophobic<br />

part <strong>of</strong> the bilayer is instead chain length dependent; the longer the hydrophobic<br />

section <strong>of</strong> the <strong>lipid</strong>s, the broader the corresponding pressure pr<strong>of</strong>ile. This is a consequence<br />

<strong>of</strong> the increase in hydrophobic thickness <strong>with</strong> increasing chain length. For<br />

the bilayer to be stable, the net outward pressure in the inner core should balance<br />

the net inward pressure at the interface. Since the latter, as we have seen, is not<br />

chain length dependent, in a thicker bilayer the pressure should be more broadly<br />

distributed. This effect is more pronounced in the case <strong>of</strong> the stiff <strong>lipid</strong>s for which<br />

there is a larger increase <strong>of</strong> thickness <strong>with</strong> increasing chain length than in the case<br />

<strong>of</strong> the flexible bilayer. The spreading <strong>of</strong> the pressure <strong>with</strong> increasing chain length,<br />

as well as the almost constant width <strong>of</strong> the peaks at the interfacial region, are trends<br />

which are also observed in the mean-field calculations <strong>of</strong> Cantor [94, 95]. This author<br />

used a mean-field statistical thermodynamic theory on a lattice model to describe<br />

the chain conformational contributions to the free energy, <strong>and</strong> calculated the<br />

equilibrium properties <strong>and</strong> lateral pressure pr<strong>of</strong>iles in <strong>lipid</strong> <strong>bilayers</strong> as function <strong>of</strong><br />

<strong>lipid</strong> composition. Cantor predicted that large redistributions <strong>of</strong> lateral pressure can<br />

be induced by variations in the <strong>lipid</strong> chain length <strong>and</strong> in the degree <strong>and</strong> position <strong>of</strong><br />

n=5<br />

n=7<br />

n=9


4.4 Results <strong>and</strong> discussion 51<br />

chain unsaturation, as well as by changes in the headgroup repulsion.<br />

Effect <strong>of</strong> local changes in the <strong>lipid</strong> chain<br />

To investigate if the effect <strong>of</strong> a modification in the <strong>lipid</strong> architecture on the distribution<br />

<strong>of</strong> the pressure pr<strong>of</strong>ile depends on the position <strong>of</strong> the change along the chain,<br />

we compare fully flexible <strong>lipid</strong>s <strong>with</strong> flexible <strong>lipid</strong>s in which a bond-bending potential<br />

<strong>with</strong> equilibrium angle θo = 180 o is introduced either close to the headgroup<br />

(ht (L) tn−1) or toward the end <strong>of</strong> the tail (htn−2t (L) t). We also compare stiff <strong>lipid</strong>s <strong>with</strong><br />

<strong>lipid</strong>s in which a kink is introduced either between the first two bonds (ht (K) t (L)<br />

n−2 t),<br />

or the last two bonds <strong>of</strong> the chain (ht (L)<br />

n−2 t(K) t). In both cases short <strong>lipid</strong>s <strong>with</strong> n = 5<br />

tail beads <strong>and</strong> long ones <strong>with</strong> n = 9 tail beads are studied.<br />

The effect <strong>of</strong> a local stiffness in the bilayer <strong>of</strong> flexible amphiphiles <strong>of</strong> short <strong>and</strong><br />

long chains is shown in figure 4.8. From these plots it can be observed that an in-<br />

π(z)<br />

0.1<br />

0.0<br />

−0.1<br />

−0.2<br />

−4 −2 0 2 4<br />

Z<br />

(a)<br />

ht 5<br />

ht 3 t (L)<br />

t<br />

ht (L)<br />

t 4<br />

π(z)<br />

0.1<br />

0.0<br />

−0.1<br />

−0.2<br />

−4 −2 0 2 4<br />

Z<br />

Figure 4.8: Pressure pr<strong>of</strong>ile, π(z), as function <strong>of</strong> the distance z from the bilayer center (z=0).<br />

Fully flexible <strong>lipid</strong>s <strong>and</strong> <strong>lipid</strong>s <strong>with</strong> a stiffness near the headgroup or at the tail end are compared<br />

for (a) short <strong>lipid</strong>s <strong>with</strong> 5 tail beads, <strong>and</strong> (b) long <strong>lipid</strong>s <strong>with</strong> 9 tail beads.<br />

creased stiffness at the tail-end does not significantly change the pressure pr<strong>of</strong>ile, in<br />

both cases <strong>of</strong> long <strong>and</strong> short <strong>lipid</strong>s. However, if the stiffness is close to the head group<br />

we observe a larger effect. The maximum at the WH interface increases in magnitude<br />

<strong>and</strong> the minimum at the HT interface deepens compared to fully flexible <strong>lipid</strong>s.<br />

Also, the first maximum in the hydrophobic core increases in height, while the central<br />

minimum does not change. The stiffness reduces the disorder <strong>of</strong> the tails <strong>and</strong> the<br />

headgroups pack more tightly, increasing the density at the interface <strong>and</strong> in the first<br />

region <strong>of</strong> the hydrophobic core. The higher density results in an increase <strong>of</strong> the lateral<br />

pressures in these regions, but since the effect is local, it does not propagate as far as<br />

the midplane. On the other h<strong>and</strong>, an increase in stiffness in the midplane region,<br />

(b)<br />

ht 9<br />

ht 7 t (L)<br />

t<br />

ht (L)<br />

t 8


52 Structural characterization <strong>of</strong> <strong>lipid</strong> <strong>bilayers</strong><br />

where the tails are already very disordered, does not change the packing structure <strong>of</strong><br />

the bilayer <strong>and</strong> has no effect on the pressure pr<strong>of</strong>ile.<br />

As in the case <strong>of</strong> flexible <strong>lipid</strong>s, a modification in the tail end <strong>of</strong> stiff <strong>lipid</strong>s (figure<br />

4.9) does not change the pressure pr<strong>of</strong>ile, while a kink near the headgroup has a large<br />

effect. It is interesting to note that, while for flexible <strong>lipid</strong>s the main effect is in the<br />

interfacial region, for stiff <strong>lipid</strong>s the main effect is in the hydrophobic core. This difference<br />

can be easily explained. The considered <strong>bilayers</strong> formed by stiff, linear <strong>lipid</strong>s,<br />

are largely interdigitated, <strong>and</strong> interdigitated <strong>bilayers</strong> are more compact <strong>and</strong> ordered<br />

than <strong>bilayers</strong> formed by flexible <strong>lipid</strong>s. A kink near the <strong>lipid</strong>s headgroup destabilizes<br />

the interfacial region, decreasing the extent <strong>of</strong> interdigitation <strong>and</strong> increasing the disorder<br />

in the terminal region <strong>of</strong> the hydrophobic chains. This leads to a pressure pr<strong>of</strong>ile<br />

more similar to the one <strong>of</strong> fully flexible chains, <strong>with</strong> a minimum <strong>of</strong> the lateral<br />

pressure at the bilayer center. It is interesting to note that this effect is larger for the<br />

shorter chain.<br />

Our results are in general agreement <strong>with</strong> the mean field calculations <strong>of</strong> Cantor<br />

[94]. In [94] the effects <strong>of</strong> unsaturated bonds was studied by imposing a preference<br />

for a 90 o angle between three subsequent sites compared to 180 o used for a saturated<br />

chain. Also Cantor found that the largest changes in the lateral pressure distribution<br />

are obtained by modifications in the <strong>lipid</strong> topology close to the head group, <strong>and</strong> that<br />

the changes are more pronounced in short chains than in long ones.<br />

π(z)<br />

0.1<br />

0.0<br />

−0.1<br />

−0.2<br />

−7 −5 −3 −1 1 3 5 7<br />

Z<br />

(a)<br />

ht (L)<br />

4 t<br />

ht (L)<br />

3 t(K) t<br />

ht (K) t (L)<br />

3 t<br />

π(z)<br />

0.1<br />

0.0<br />

−0.1<br />

−0.2<br />

−7 −5 −3 −1 1 3 5 7<br />

Z<br />

(b)<br />

ht (L)<br />

8 t<br />

ht (L)<br />

7 t(K) t<br />

ht (K) t (L)<br />

7 t<br />

Figure 4.9: Pressure pr<strong>of</strong>ile, π(z), as function <strong>of</strong> the distance z from the bilayer center (z=0).<br />

Stiff <strong>lipid</strong>s <strong>and</strong> <strong>lipid</strong>s <strong>with</strong> a kink near the headgroup or at the tail end are compared for (a)<br />

short <strong>lipid</strong>s <strong>with</strong> 5 tail beads, <strong>and</strong> (b) long <strong>lipid</strong>s <strong>with</strong> 9 tail beads.


4.4 Results <strong>and</strong> discussion 53<br />

4.4.4 Double-tail <strong>lipid</strong>s<br />

We have so far investigated the structure <strong>of</strong> <strong>bilayers</strong> formed by single-tail <strong>lipid</strong>s. We<br />

have shown that <strong>lipid</strong> stiffness is an important characteristics in order to reproduce<br />

the bilayer structure <strong>and</strong> <strong>lipid</strong> chain order as displayed in phospho<strong>lipid</strong> <strong>bilayers</strong>. Furthermore,<br />

we have shown that by tuning the interaction parameter between the <strong>lipid</strong><br />

headgroups we obtain either an interdigitated or a non interdigitated bilayer. To<br />

further investigate the dependence <strong>of</strong> the bilayer structure on the <strong>lipid</strong> topology we<br />

can increase the complexity <strong>of</strong> the model by considering model <strong>lipid</strong>s <strong>with</strong> two hydrophobic<br />

tails. We will focus the study to <strong>lipid</strong>s <strong>with</strong> two symmetric tails, i.e. two<br />

tails consisting <strong>of</strong> the same number <strong>of</strong> hydrophilic beads.<br />

Here we consider the <strong>lipid</strong> denoted as h3(t5)2, i.e. <strong>with</strong> three head-beads <strong>and</strong><br />

two tails <strong>of</strong> five hydrophobic beads each, as already described in the previous Chapters.<br />

The <strong>lipid</strong> tails are made stiffer by introducing a bond-bending potential between<br />

each consecutive bond, <strong>with</strong> bending constant Kθ = 6 <strong>and</strong> equilibrium angle<br />

θo = 180 o . An extra bond-bending potential is defined between the bonds connecting<br />

the two <strong>lipid</strong> tails to the headgroup, <strong>with</strong> Kθ = 3 <strong>and</strong> θo = 90 o . We considered a<br />

bilayer <strong>of</strong> 400 <strong>lipid</strong>s, at the condition <strong>of</strong> zero surface tension, <strong>and</strong> in the fluid phase.<br />

In figure 4.10 we compare the density distribution in the bilayer <strong>of</strong> double-tail<br />

<strong>lipid</strong>s (black lines) <strong>with</strong> the density distribution in the bilayer <strong>of</strong> stiff, single-tail <strong>lipid</strong>s<br />

(gray lines), <strong>with</strong> five tail beads, <strong>and</strong> the headgroup repulsion parameter for which<br />

the bilayer is not interdigitated.<br />

ρ(z)<br />

4.0<br />

3.0<br />

2.0<br />

1.0<br />

w<br />

tot<br />

t (1,.,n−1)<br />

h h<br />

tn 0.0<br />

−6 −4 −2 0 2 4 6<br />

z<br />

Figure 4.10: Density distribution in <strong>bilayers</strong> <strong>of</strong> single-tail <strong>lipid</strong>s (gray lines) <strong>and</strong> double-tail<br />

<strong>lipid</strong>s (black-lines). The black dashed lines are the density <strong>of</strong> each <strong>of</strong> the three head-beads <strong>of</strong><br />

the double-tail <strong>lipid</strong>.<br />

A noticeable difference between the two <strong>bilayers</strong> is at the WH interface, where in<br />

the case <strong>of</strong> the double-tail <strong>lipid</strong>s the density <strong>of</strong> water shows a small plateau, which is<br />

not present in the case <strong>of</strong> single-tail <strong>lipid</strong>s. This plateau corresponds to a thicker re-<br />

w


54 Structural characterization <strong>of</strong> <strong>lipid</strong> <strong>bilayers</strong><br />

gion where the <strong>lipid</strong> headgroup are hydrated by the water, <strong>and</strong> is due to the fact that<br />

double-tail <strong>lipid</strong>s have a larger headgroup (three beads) respect to the one <strong>of</strong> singletail<br />

<strong>lipid</strong>s (one bead). Also, at the HT interface the total density shows more marked<br />

minima in the case <strong>of</strong> double-tail <strong>lipid</strong>s, <strong>and</strong> a small difference between the two <strong>bilayers</strong><br />

is observed in the bilayer hydrophobic core where the total density <strong>of</strong> the tail<br />

beads is slightly larger for the single tail <strong>lipid</strong>s. A remarkable fact is that, despite the<br />

π(z)<br />

0.1<br />

0.0<br />

−0.1<br />

−0.2<br />

−0.3<br />

−6 −4 −2 0 2 4 6<br />

z<br />

(a) (b)<br />

Figure 4.11: Figure (a) shows the lateral pressure pr<strong>of</strong>ile distribution in a bilayer <strong>of</strong> singletail<br />

<strong>lipid</strong>s (gray line) <strong>and</strong> a bilayer <strong>of</strong> double-tail <strong>lipid</strong>s (black) line. In figure (b) the pressure<br />

pr<strong>of</strong>ile from atomistic MD simulations <strong>of</strong> a DPPC bilayer is shown. Figure (b) is reproduced<br />

from reference [77] <strong>with</strong> kind permission <strong>of</strong> the authors.<br />

differences in the density pr<strong>of</strong>iles <strong>of</strong> the two <strong>bilayers</strong> being not too large, the distribution<br />

<strong>of</strong> lateral pressure in the double-tail <strong>lipid</strong> bilayer is very different compared<br />

<strong>with</strong> the single-tail <strong>lipid</strong> bilayer, as shown in figure 4.11(a). In the case <strong>of</strong> double-tail<br />

<strong>lipid</strong>s, the peaks at the WH <strong>and</strong> HT interface become more pronounced, indicating<br />

that the <strong>lipid</strong>s are more densely packed. Also, because <strong>of</strong> the larger headgroup size<br />

<strong>of</strong> the double-tail <strong>lipid</strong>s, the peak at the WH interface broadens. The central peak<br />

(at the bilayer midplane) decreases in size, due to the lower density in the bilayer<br />

center. Most noticeably, we observe the appearance <strong>of</strong> two secondary peaks at the<br />

edge <strong>of</strong> the hydrophobic core, which are not present in the case <strong>of</strong> the single-tail<br />

<strong>lipid</strong>s. These peaks correspond to the region <strong>of</strong> maximum density for the tail beads<br />

(excluded the terminal one). This shape <strong>of</strong> the pressure pr<strong>of</strong>ile is remarkably similar<br />

to the one calculated by Lindahl <strong>and</strong> Edholm in atomistic MD simulations <strong>of</strong> a<br />

dipalmitoylphosphatidylcholine bilayer [77], <strong>and</strong> shown in figure 4.11(b).<br />

In Chapter 2 we have shown that, if a DPD bead is chosen to represent three<br />

methyl groups, the <strong>lipid</strong> h3(t5)2 can be mapped onto the DMPC phospho<strong>lipid</strong>. The<br />

structure <strong>of</strong> DPPC is very similar to the one <strong>of</strong> DMPC, both having a phosphocholine


4.4 Results <strong>and</strong> discussion 55<br />

headgroup <strong>and</strong> two saturated hydrocarbon chains, <strong>with</strong> DPPC having 16 carbons in<br />

each tail, <strong>and</strong> DMPC 14. This result indicates that the coarse-grained model we have<br />

developed for double-tail <strong>lipid</strong>s well reproduces the distribution <strong>of</strong> pressure in a bilayer<br />

<strong>of</strong> saturated phosphocholines. However, care must be used in such a conclusion.<br />

We want to point out that, given the differences between the atomistic <strong>and</strong> the<br />

CG representations, <strong>and</strong> the different way in which the inter-<strong>and</strong> intra-molecular interactions<br />

are implemented in the two <strong>models</strong>, the remarkable similarity between the<br />

pressure pr<strong>of</strong>iles in the two model might be fortuitous. To further address this issue,<br />

it would be interesting to compare in a more systematic way pressure pr<strong>of</strong>iles calculated<br />

from MD <strong>and</strong> CG simulations. Unfortunately, the published pressure pr<strong>of</strong>iles<br />

in <strong>lipid</strong> <strong>bilayers</strong> from MD simulations are still very few.


V<br />

Phase behavior <strong>of</strong> coarse-grained <strong>lipid</strong> <strong>bilayers</strong>


58 Phase behavior <strong>of</strong> coarse-grained <strong>lipid</strong> <strong>bilayers</strong><br />

5.1 Introduction<br />

As we have discussed in the previous Chapter, an important question in the development<br />

<strong>of</strong> a mesoscopic model is how much chemical detail should be included in the<br />

model. We have described in some detail the structural characteristics <strong>of</strong> <strong>bilayers</strong> <strong>of</strong><br />

single-tail <strong>lipid</strong>s. In this Chapter, we investigate if such a model correctly describes<br />

the (phase) behavior <strong>of</strong> double tail phospho<strong>lipid</strong>s, <strong>and</strong> we compare the results <strong>of</strong> our<br />

simulations <strong>with</strong> experiments that are performed on single-tail <strong>lipid</strong>s.<br />

In the second part <strong>of</strong> this Chapter we increase the complexity <strong>of</strong> the model <strong>lipid</strong>s,<br />

by studying the phase behavior <strong>of</strong> a double-tail <strong>lipid</strong> <strong>with</strong> three hydrophilic headbeads<br />

<strong>and</strong> five hydrophobic beads in each <strong>of</strong> the tails. Furthermore, we show that<br />

a correspondence can be made between reduced <strong>and</strong> physical temperature. With<br />

this mapping, <strong>and</strong> the mapping <strong>of</strong> reduced length scales onto physical lengths, the<br />

typical bilayer characteristics, like the area per <strong>lipid</strong> <strong>and</strong> bilayer thickness, computed<br />

in the simulations are found to be in good quantitative agreement <strong>with</strong> experimental<br />

results. In particular, we show that this <strong>lipid</strong> model correctly reproduces the phase<br />

behavior <strong>of</strong> dimyristoylphosphatidylcholine (DMPC) <strong>lipid</strong> <strong>bilayers</strong>.<br />

The phase behavior <strong>of</strong> different phosphocholines (PC’s) has been determined experimentally<br />

(see [96] for a review). All PC’s have a low temperature Lβ ′ phase (see<br />

figure 5.1(a)). In this phase the bilayer is a gel: the chains <strong>of</strong> the phospho<strong>lipid</strong>s are<br />

ordered <strong>and</strong> show a tilt relative to the bilayer normal. At higher temperature the Lα<br />

phase is the stable phase. This phase is the liquid crystalline state <strong>of</strong> the bilayer in<br />

which the chains are disordered <strong>and</strong> tail overlap due to this thermal disorder is possible.<br />

This phase is physiologically the most relevant [97].<br />

(a) L β ′ (b) P β ′ (c) Lα (d) LβI<br />

Figure 5.1: Schematic drawings <strong>of</strong> the various bilayer phases. The characteristics <strong>of</strong> these<br />

phases are explained in the text. The filled circles represent the hydrophilic headgroup <strong>of</strong> a<br />

phospho<strong>lipid</strong> <strong>and</strong> the lines represent the hydrophobic tails.<br />

Under normal conditions the two monolayers <strong>of</strong> a bilayer contact each other<br />

at the terminal methyl group <strong>of</strong> their hydrophobic chains, while their hydrophilic<br />

headgroups are in contact <strong>with</strong> water. However, it is known experimentally that at<br />

low temperatures an interdigitated state, in which the terminal methyl groups <strong>of</strong><br />

one monolayer interpenetrate the opposing layer, is also possible. This LβI phase


5.2 Single-tail <strong>lipid</strong> <strong>bilayers</strong> 59<br />

does not spontaneously form in <strong>bilayers</strong> <strong>of</strong> symmetrical chain phospho<strong>lipid</strong>s, like<br />

dipalmitoylphosphatidylcholines (DPPC), [98] but can be induced by changes in the<br />

environment, like hydrostatic pressure or changes in the pH <strong>of</strong> the solution [99], or<br />

by incorporation, at the membrane interface <strong>of</strong> small amphiphilic molecules, like<br />

alcohols [90, 100, 101], or anesthetics [102]. Interdigitation can also be induced by<br />

changes in the <strong>lipid</strong> structure, for example by introducing an ester-linkage in the<br />

headgroup <strong>of</strong> the phospho<strong>lipid</strong>s [103, 104].<br />

Interdigitation reduces the bilayer thickness, <strong>and</strong> this can, for example, affect the<br />

diffusion <strong>of</strong> ions across the bilayer or influence the activity <strong>of</strong> membrane proteins. It<br />

has been proposed [91, 99] that specific interactions are not important in the formation<br />

<strong>of</strong> an interdigitated phase, <strong>and</strong> that the main driving force that induces interdigitation<br />

is an increase in the headgroup surface area, which results in the creation <strong>of</strong><br />

voids between the molecules. Since voids in the bilayer core are energetically unfavorable,<br />

they are filled up by molecules <strong>of</strong> the opposite monolayer.<br />

This mechanism suggests that the formation <strong>of</strong> an interdigitated phase should be<br />

a general phenomenon. This would imply that an interdigitated phase could also<br />

be induced in <strong>bilayers</strong> <strong>of</strong>, for example, single-tail <strong>lipid</strong>s. The fact that for single-tail<br />

<strong>lipid</strong>s the interdigitated phase has not been observed experimentally, is one <strong>of</strong> the<br />

motivations to investigate the molecular aspects underlying the formation <strong>of</strong> an interdigitated<br />

phase in more detail.<br />

Our simulations correctly describe the hydrophobic tail length dependence <strong>of</strong><br />

this transition <strong>and</strong> the effect <strong>of</strong> adding salt. In addition, the simulations predict that<br />

both the interdigitated <strong>and</strong> non-interdigitated phases can be formed in systems <strong>with</strong><br />

single-tail <strong>lipid</strong>s. Conversely, we do not find any spontaneous interdigitation is <strong>bilayers</strong><br />

<strong>of</strong> double-tail <strong>lipid</strong>s.<br />

5.2 Single-tail <strong>lipid</strong> <strong>bilayers</strong><br />

5.2.1 Computational details<br />

In this investigation we consider <strong>lipid</strong>s <strong>with</strong> one head segment connected to a single<br />

tail <strong>with</strong> variable length (see figure 5.2). Two consecutive beads are connected by<br />

harmonic springs <strong>with</strong> spring constant Kr = 100 <strong>and</strong> ro = 0.7. A harmonic bond<br />

bending potential between three consecutive beads is added <strong>with</strong> a bending constant<br />

Kθ = 10 <strong>and</strong> an equilibrium angle θ0 = 180 ◦ .


60 Phase behavior <strong>of</strong> coarse-grained <strong>lipid</strong> <strong>bilayers</strong><br />

ht 6<br />

ht7<br />

ht8<br />

ht9<br />

Figure 5.2: Schematic drawing <strong>of</strong> the model <strong>lipid</strong>s used in this study <strong>with</strong> their nomenclature.<br />

The black particles represent the head beads <strong>and</strong> the white particles the tail beads.<br />

The repulsion parameters used are aww = att = 25, a wh = 15, <strong>and</strong> awt = 80 (see<br />

equation 2.2, Chapter 2). In addition, we vary the head-head interaction parameter<br />

(a hh) to study the effect <strong>of</strong> changing the interactions between the hydrophilic segments<br />

<strong>of</strong> a <strong>lipid</strong>. In a real system the head-head interactions can be changed by, for<br />

example, adding salt to the system.<br />

All our simulations are performed on tensionless <strong>bilayers</strong> <strong>of</strong> 200 <strong>lipid</strong>s. The total<br />

number <strong>of</strong> particles was 3500. The overall density <strong>of</strong> the system is ρ = 3. We initialize<br />

our system by distributing <strong>lipid</strong>s r<strong>and</strong>omly in water <strong>and</strong> we observe the self-assembly<br />

<strong>of</strong> a bilayer using DPD simulation only. After the bilayer is formed, we perform, in addition<br />

to the DPD moves, Monte Carlo moves in which we change the area as well.<br />

A typical simulation required 100,000 cycles <strong>of</strong> which 20,000 cycles were needed for<br />

equilibration. All the results are expressed in the usual reduced units, i.e. using Rc as<br />

the unit <strong>of</strong> length <strong>and</strong> kBTo = 1, <strong>with</strong> To room temperature, as unit <strong>of</strong> the energy. In<br />

the following we will denote as T ∗ the temperature expressed in this unit.<br />

5.2.2 Results <strong>and</strong> Discussion<br />

In this section we first describe in detail the different phases <strong>of</strong> a bilayer formed by<br />

single tail <strong>lipid</strong>s consisting <strong>of</strong> one head bead <strong>and</strong> nine tail beads (ht9), which we study<br />

at different reduced temperatures, from T ∗ = 0.8 to T ∗ = 1.5. At a fixed head-head repulsion<br />

<strong>of</strong> a hh = 35 an interdigitated gel phase is formed at low temperatures, while<br />

at a hh = 15 the non-interdigitated Lβ phase is formed. We then investigate the influence<br />

<strong>of</strong> changing the interactions between the headgroups <strong>and</strong> we show that we<br />

obtain the non-interdigitated Lβ phase or the interdigitated LβI phase dependent on<br />

the head-head repulsion parameter (see figure 5.3). Finally, we investigate the influence<br />

<strong>of</strong> tail length <strong>and</strong> we compare our results <strong>with</strong> experimental data on single-tail<br />

<strong>lipid</strong>s.


5.2 Single-tail <strong>lipid</strong> <strong>bilayers</strong> 61<br />

ahh<br />

40<br />

30<br />

20<br />

L EI<br />

L E<br />

10<br />

0.7 0.8 0.9 1.0 1.1<br />

Figure 5.3: Computed phase diagram <strong>of</strong> the <strong>lipid</strong> ht9 as function <strong>of</strong> head-head repulsion parameter<br />

ahh <strong>and</strong> reduced temperature T ∗ . At high values <strong>of</strong> the head-head repulsion parameters<br />

the interdigitated LβI phase is formed, while at low values the non-interdigitated Lβ phase<br />

is formed. Increasing temperature causes the melting <strong>of</strong> the bilayer to the Lα phase.<br />

The <strong>lipid</strong> ht9<br />

Head-head repulsion a hh = 35<br />

In figure 5.4 the average area per <strong>lipid</strong> AL <strong>and</strong> the bilayer thickness, Dc, are plotted as<br />

function <strong>of</strong> temperature. The error bars have been calculated <strong>with</strong> the block averages<br />

method [76, 105]. In all the other plots <strong>of</strong> the area per <strong>lipid</strong> or bilayer thickness, we<br />

will not include error bars, which, however, have been estimated as ≤ 5%.<br />

A L<br />

0.95<br />

0.90<br />

0.85<br />

0.80<br />

0.8 1.0 1.2<br />

T*<br />

1.4 1.6<br />

(a)<br />

D c<br />

T*<br />

8.20<br />

7.80<br />

7.40<br />

L D<br />

7.00<br />

0.8 1.0 1.2<br />

T*<br />

1.4 1.6<br />

Figure 5.4: Area per <strong>lipid</strong> AL <strong>and</strong> (b) bilayer thickness Dc as function <strong>of</strong> reduced temperature<br />

T ∗ for <strong>lipid</strong> type ht9.<br />

(b)


62 Phase behavior <strong>of</strong> coarse-grained <strong>lipid</strong> <strong>bilayers</strong><br />

In both figures 5.4(a) <strong>and</strong> 5.4(b) we can distinguish two regions: at low temperatures,<br />

the area per <strong>lipid</strong> decreases <strong>with</strong> increasing temperature <strong>and</strong> the thickness is<br />

increasing, while at high temperatures the area is increasing <strong>with</strong> increasing temperature,<br />

<strong>and</strong> the thickness is decreasing. At the lowest temperature studied (T ∗ = 0.8)<br />

the area is larger than the area at the highest temperature studied (T ∗ = 1.5) while<br />

the thickness at T ∗ = 0.8 is smaller than the thickness at T ∗ = 1.5. This different<br />

temperature dependence <strong>of</strong> AL <strong>and</strong> Dc suggests that the bilayer undergoes a phase<br />

transition. Before discussing this transition in detail we will first characterize the low<br />

<strong>and</strong> the high temperature phases.<br />

To characterize the ordering <strong>of</strong> the <strong>lipid</strong>s in the bilayer we use the order parameters<br />

S tail <strong>and</strong> Sn. In figure 5.5 the values <strong>of</strong> both S tail <strong>and</strong> Sn are plotted as a function <strong>of</strong><br />

temperature. The high values <strong>of</strong> Sn at temperatures below T ∗ = 0.95 indicate that the<br />

bonds are ordered along the bilayer normal. This order persists even for bonds far<br />

from the head-group region, decreasing slightly <strong>with</strong> increasing temperature. Above<br />

T ∗ = 0.95 the values <strong>of</strong> Sn further decrease <strong>with</strong> increasing temperature, <strong>and</strong> the order<br />

along the chain is lost.<br />

(a) (b)<br />

Figure 5.5: (a) Local order parameter Sn <strong>and</strong> (b) tail order parameter Stail, as function <strong>of</strong> reduced<br />

temperature T ∗ .<br />

The overall order <strong>of</strong> the tails (S tail) shows a similar behavior (figure 5.5(b)). Also<br />

here we can distinguish two regions: below T ∗ = 0.95 where S tail has values higher<br />

than 0.5 indicating that the chains are ordered along the bilayer normal, <strong>and</strong> above<br />

T ∗ = 0.95 where the values <strong>of</strong> S tail decrease below 0.5, showing an increase in the<br />

disorder <strong>of</strong> the chains.<br />

To further characterize the structure <strong>of</strong> the bilayer in the low <strong>and</strong> high temperature<br />

regions, we compare in figure 5.6 the in-plane radial distribution function g(r) <strong>of</strong><br />

the head beads <strong>of</strong> the <strong>lipid</strong>s at one interface, for two different temperatures: T ∗ = 0.8<br />

<strong>and</strong> T ∗ = 1.5.At T ∗ = 0.8, the radial distribution function shows more pronounced


5.2 Single-tail <strong>lipid</strong> <strong>bilayers</strong> 63<br />

peaks compared to the g(r) at T ∗ = 1.5, which corresponds to a more structured organization<br />

<strong>of</strong> the <strong>lipid</strong>s headgroups in the bilayer plane. The structure in the radial<br />

distribution function <strong>and</strong> the high values <strong>of</strong> the order parameters for low temperatures,<br />

suggest that the low temperature phase is the ordered gel phase, while at high<br />

temperatures the bilayer is in the disordered liquid crystalline phase.<br />

Figure 5.6: Two dimensional radial distribution function g(r) in the bilayer plane for the headgroups<br />

at T ∗ = 0.8 <strong>and</strong> T ∗ = 1.5.<br />

In figure 5.7 we show the density pr<strong>of</strong>iles in the direction normal to the bilayer for<br />

the system components at different reduced temperatures. Figures 5.7(a) <strong>and</strong> 5.7(b)<br />

correspond to a bilayer in the gel phase, while 5.7(c) <strong>and</strong> 5.7(d) correspond to a bilayer<br />

in the liquid crystalline phase. It is clearly visible that, in the low temperature<br />

region, the two monolayers are interdigitated. At T ∗ = 0.8 the overlap extends up to<br />

the 8 th bead in the tail <strong>and</strong> the peaks <strong>of</strong> the density pr<strong>of</strong>iles for the <strong>lipid</strong>s tail beads in<br />

one monolayer (black full lines) are exactly alternating <strong>with</strong> the peaks <strong>of</strong> the opposite<br />

monolayer (gray full lines), showing an optimal packing <strong>of</strong> the tails. This structure resembles<br />

the experimentally observed interdigitated phase LβI.<br />

We can now explain the temperature dependence <strong>of</strong> the area per <strong>lipid</strong> (figure 5.4).<br />

The low temperature phase is the interdigitated gel LβI. In this phase the ordering<br />

<strong>of</strong> the chains is the dominating effect. The <strong>lipid</strong>s stretch out in the direction normal<br />

to the bilayer, inducing interdigitation. This packing results in a larger average<br />

distance between the <strong>lipid</strong>s headgroups in each monolayer <strong>and</strong> in a larger area. In<br />

this region an increase <strong>of</strong> temperature reduces the values <strong>of</strong> the order parameter (figure<br />

5.5(b)), but along the chain the order persists (figure 5.5(a)). Thus interdigitation<br />

is still present, but is decreasing in depth, resulting in an increase <strong>of</strong> the bilayer<br />

thickness <strong>and</strong> a decrease <strong>of</strong> the area per <strong>lipid</strong>. Above the transition temperature, the<br />

chains loose the persisting order <strong>and</strong> are not interdigitated. Only the terminal tail<br />

beads overlap, due to thermal disorder. In this temperature region an increase in<br />

temperature increases the effective volume occupied by the molecules, but the extent<br />

<strong>of</strong> tail overlap does not depend significantly <strong>of</strong> temperature. As a result the area


64 Phase behavior <strong>of</strong> coarse-grained <strong>lipid</strong> <strong>bilayers</strong><br />

(a) (b)<br />

(c) (d)<br />

Figure 5.7: Density pr<strong>of</strong>iles ρ(z) along the bilayer normal z for different reduced temperatures<br />

T ∗ . Each line is the density pr<strong>of</strong>ile for a different bead: full lines are the densities <strong>of</strong> the tail<br />

beads, dashed lines are the densities <strong>of</strong> the head beads, <strong>and</strong> the thin solid line is the density<br />

<strong>of</strong> water. The black lines correspond to the <strong>lipid</strong>s in one monolayer, while the gray lines correspond<br />

to the <strong>lipid</strong>s in the opposite monolayer. The big dots correspond to the maxima in<br />

the bead density distributions <strong>and</strong> illustrate the position <strong>of</strong> the beads in the bilayer. The full<br />

circles correspond to tail beads <strong>and</strong> the open circles to head beads.


5.2 Single-tail <strong>lipid</strong> <strong>bilayers</strong> 65<br />

per molecule increases while the bilayer thickness decreases.<br />

Head-head repulsion a hh = 15<br />

In the previous section we have seen that single tail <strong>lipid</strong>s spontaneously form an<br />

interdigitated phase at low temperatures, while the most common organization <strong>of</strong><br />

(symmetric) phospho<strong>lipid</strong>s in membranes is a bilayer formed by two separate monolayers<br />

[106]. It is therefore interesting to investigate whether we can adapt the single<br />

tail model to reproduce the phase behavior <strong>of</strong> real membranes, <strong>and</strong> in particular if we<br />

are able to obtain a non-interdigitated gel phase. If the main cause <strong>of</strong> interdigitation<br />

is an increase in the head-groups surface area [91, 99], we can test this mechanism<br />

by changing the value <strong>of</strong> the head-group repulsion parameter, a hh, in our model.<br />

Taking as initial condition the interdigitated bilayer at T ∗ = 0.85, we decrease the<br />

head-group repulsion parameter from a hh = 35 to a hh = 15, the latter being the same<br />

repulsion parameter as between an hydrophilic bead <strong>and</strong> a water-bead. Experimentally,<br />

changing the head-head interactions corresponds to, for example, adding salt<br />

to the system. It is important to recall that, <strong>with</strong> the zero surface tension scheme, the<br />

system can evolve to the optimum area per <strong>lipid</strong> even if the bilayer undergoes structural<br />

rearrangements.<br />

(a) (b)<br />

Figure 5.8: Comparison <strong>of</strong> (a) the area per <strong>lipid</strong> AL as function <strong>of</strong> reduced temperature T ∗ , <strong>and</strong><br />

(b) two dimension radial distribution function g(r) in the plane <strong>of</strong> the bilayer at T ∗ = 0.85, for<br />

two different repulsion parameters between the <strong>lipid</strong> headgroups: ahh = 15 (circles, solid line)<br />

<strong>and</strong> ahh = 35 (squares, dashed line).<br />

Figure 5.8(a) shows the temperature dependence <strong>of</strong> the area per <strong>lipid</strong> for the repulsion<br />

parameters a hh = 35 <strong>and</strong> a hh = 15. We observe that the behavior in temperature<br />

<strong>of</strong> the area per <strong>lipid</strong> for the two values <strong>of</strong> a hh is very different. At low temperatures<br />

the area at a hh = 35 is almost twice the value <strong>of</strong> the area at a hh = 15. The decrease


66 Phase behavior <strong>of</strong> coarse-grained <strong>lipid</strong> <strong>bilayers</strong><br />

<strong>of</strong> the head-group surface area is also shown in figure 5.8(b), where we compare the<br />

radial distribution functions <strong>of</strong> the headgroups in the bilayer plane at T ∗ = 0.85 (see<br />

also figure 5.9 for snapshots <strong>of</strong> the two systems). The peaks in the radial distribution<br />

function for the system <strong>with</strong> a hh = 15 (solid line) are shifted to the left compared to<br />

the system <strong>with</strong> a hh = 35 (dashed line), showing a decrease <strong>of</strong> the distance between<br />

the headgroups. This is a strong indication that at low temperature, <strong>with</strong> the lower<br />

repulsion parameter, the bilayer is in the Lβ phase.<br />

(a) (b)<br />

Figure 5.9: Snapshots <strong>of</strong> the simulations <strong>of</strong> a bilayer consisting <strong>of</strong> the <strong>lipid</strong> ht9 at T ∗ = 0.85.<br />

(a) The non-interdigitated gel phase Lβ at ahh = 15 <strong>and</strong> (b) the interdigitated gel phase Lβ at<br />

ahh = 15. Black represents the hydrophilic headgroup <strong>and</strong> gray represents the hydrophobic<br />

tails.<br />

(a) (b)<br />

Figure 5.10: (a) Local order parameter Sn <strong>and</strong> (b) tail order parameter Stail, as function <strong>of</strong><br />

reduced temperature T ∗ for a bilayer formed by <strong>lipid</strong>s <strong>with</strong> ahh = 15.<br />

To further characterize the bilayer structure for a hh = 15, we study the order parameters<br />

Sn<strong>and</strong> S tail, which are plotted in figure 5.10. At temperatures T ∗ ≤ 0.95 the<br />

chains are locally ordered (values <strong>of</strong> Sn above 0.5), <strong>and</strong> the order does not decrease<br />

significantly going through the hydrophobic core. Also the overall order <strong>of</strong> the chains<br />

S tail is high in this temperature region. Above T ∗ = 0.95 we observe a decrease in both


5.2 Single-tail <strong>lipid</strong> <strong>bilayers</strong> 67<br />

the order parameters. The chains become disordered <strong>and</strong> the persistence <strong>of</strong> order<br />

along the chain is lost. This trend is analogous to the one observed for a hh = 35. In<br />

both cases the low temperature region is characterized by the ordering <strong>of</strong> the chains,<br />

while at high temperatures the chains are disordered. However, while for a hh = 35<br />

the two monolayers are interdigitated in the ordered phase, for a hh = 15 the ordered<br />

phase is a bilayer formed by two separated leaflets. This can clearly be seen from<br />

the density pr<strong>of</strong>iles, which we plot as function <strong>of</strong> reduced temperature in figure 5.11.<br />

This figure shows that the melting <strong>of</strong> the bilayer results in a broader shape <strong>of</strong> the den-<br />

(a) T ∗ = 0.8 (b) T ∗ = 0.9<br />

(c) T ∗ = 0.95 (d) T ∗ = 1.0<br />

Figure 5.11: Density pr<strong>of</strong>iles as function <strong>of</strong> temperature for a bilayer formed by <strong>lipid</strong>s <strong>with</strong><br />

ahh = 15 (see also the caption to figure 5.7).<br />

sity pr<strong>of</strong>iles. The increase <strong>of</strong> disorder in the chains (see figure 5.10) results in a partial<br />

overlap <strong>of</strong> the two monolayers. This transition upon heating is also reflected in the<br />

trend <strong>of</strong> the area per <strong>lipid</strong> <strong>with</strong> temperature (figure 5.8(a)), which shows a sharp increase<br />

between T ∗ = 0.95 <strong>and</strong> T ∗ = 1.0. We can then conclude that a transition from<br />

an ordered to a disordered phase takes place at a temperature 0.95 < Tm < 1.0.


68 Phase behavior <strong>of</strong> coarse-grained <strong>lipid</strong> <strong>bilayers</strong><br />

We have shown that the bilayer structure in the low temperature region depends<br />

on the repulsion between the <strong>lipid</strong> headgroups. By tuning this parameter, we can obtain<br />

both the gel phase Lβ <strong>and</strong> the interdigitated gel phase LβI. Experimentally, both<br />

in the liquid crystalline phase [92] <strong>and</strong> in the gel phase [107], a monotonic increase <strong>of</strong><br />

the area per <strong>lipid</strong> is observed when the temperature is increased. This is caused by an<br />

increase in the disorder <strong>of</strong> the tails [92]. For the low repulsion parameter <strong>of</strong> ahh = 15<br />

we reproduce the experimental observed trends. It is worth mentioning that, in most<br />

cases, in the gel phase the phospho<strong>lipid</strong> chains are tilted <strong>with</strong> respect to the bilayer<br />

normal [96]. While for single tail <strong>lipid</strong>s we do not observe any tilt, we will see in the<br />

next section that the double tail <strong>lipid</strong>s are tilted in the gel phase (Lβ ′ phase).<br />

Phase behavior as a function <strong>of</strong> head-head repulsion<br />

It is now interesting to do a more systematic study <strong>of</strong> these phase transitions for a<br />

range <strong>of</strong> repulsion parameters. The phase transitions we consider are:<br />

1. transition from interdigitated gel to gel (LβI → Lβ)<br />

2. transition from interdigitated gel to liquid crystalline (LβI → Lα)<br />

3. transition from gel to liquid crystalline (Lβ → Lα).<br />

As we have shown, the first transition is induced by a decrease in the repulsion parameter<br />

a hh, while the latter ones are temperature dependent.<br />

We use three quantities to distinguish among the different phases: the area per<br />

<strong>lipid</strong> AL, the extent <strong>of</strong> tail overlap D overlap, <strong>and</strong> the ordering <strong>of</strong> the tails S tail. By studying<br />

the behavior <strong>of</strong> these quantities as function <strong>of</strong> temperature <strong>and</strong> head-head repulsion<br />

parameter we can determine the phase diagram <strong>of</strong> ht9 as shown in figure 5.3.<br />

In figure 5.12 we plot the area per <strong>lipid</strong> AL, the extent <strong>of</strong> tail overlap D overlap, <strong>and</strong><br />

the chain order parameter S tail as function <strong>of</strong> temperature <strong>and</strong> head-head repulsion<br />

parameter. For repulsion parameters a hh ≤ 18, the low temperature phase is the<br />

bilayer gel Lβ phase, while for repulsion parameters a hh > 18, the low temperature<br />

phase is the interdigitated gel LβI.


5.2 Single-tail <strong>lipid</strong> <strong>bilayers</strong> 69<br />

A L<br />

1.0<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

a hh =10<br />

a hh =15<br />

a hh =20<br />

a hh =25<br />

a hh =35<br />

0.5<br />

0.8 0.9 1.0<br />

T*<br />

1.1 1.2<br />

(a) (b)<br />

(c)<br />

Figure 5.12: (a) Area per <strong>lipid</strong> AL, (b) extent <strong>of</strong> chain overlap Doverlap, <strong>and</strong> (c) tail order parameter<br />

Stail as function <strong>of</strong> reduced temperature T ∗ for different repulsion parameters ahh. Dashed<br />

curves show a transition from the Lβ to the Lα phase, solid curves show the transition from the<br />

LβI to the Lα phase.<br />

By increasing temperature all <strong>bilayers</strong> melt from an ordered into a disordered<br />

phase. For <strong>bilayers</strong> in the Lβ phase, the area per molecule <strong>and</strong> chain overlap increase<br />

upon melting, while for <strong>bilayers</strong> in the LβI phase the area per molecule <strong>and</strong> chain<br />

overlap decrease.<br />

The curves in figure 5.12(c) show that the transition from an ordered phase to a<br />

disordered one is very gradual. Much larger systems might be required to observe<br />

a sharp transition in these quasi two-dimensional systems. This gradual transition<br />

makes it difficult to determine the exact location <strong>of</strong> the phase boundaries <strong>and</strong> therefore<br />

we used the inflection point as our definition <strong>of</strong> the phase boundary. The temperature<br />

at which the chains get disordered is the same as the temperature <strong>of</strong> the<br />

inflection point in AL <strong>and</strong> D overlap. We define as the main transition temperature Tm<br />

the value <strong>of</strong> temperature at the inflection point <strong>of</strong> the shown curves. Tm is higher for


70 Phase behavior <strong>of</strong> coarse-grained <strong>lipid</strong> <strong>bilayers</strong><br />

<strong>bilayers</strong> in the Lβ phase than for <strong>bilayers</strong> in the LβI phase. This is in agreement <strong>with</strong><br />

experimental results [99].<br />

Phase behavior as a function <strong>of</strong> tail length<br />

Besides investigating the effect <strong>of</strong> changing the head-head repulsion parameter, it is<br />

also interesting to vary the tail length <strong>of</strong> the <strong>lipid</strong>. A similar analysis, as was presented<br />

for the <strong>lipid</strong> ht9, has been carried out for <strong>lipid</strong> types ht6, ht7, <strong>and</strong> ht8 (see figure 5.13).<br />

Depending on the repulsion parameter we obtain two gel phases LβI <strong>and</strong> Lβ for all<br />

tail lengths. For high head-head repulsion the system can gain energy by adding water<br />

particles in between the heads. As a result the distance between the head groups<br />

increases <strong>and</strong> the interdigitated phase is stabilized. For low values <strong>of</strong> a hh the headgroups<br />

expel water <strong>and</strong> the stable phase is the non-interdigitated phase. In between<br />

we find a ∗ hh for which the transition from LβI to Lβ occurs. The difference between<br />

the two phases is that in the LβI phase the tail ends are in direct contact <strong>with</strong> water,<br />

whereas in the Lβ phase the tail ends face each other. Therefore, the critical value a ∗ hh<br />

to induce interdigitation is higher than the value <strong>of</strong> a wh.<br />

We observe hysteresis if we change a hh at a constant temperature: the bilayer can<br />

be both in the Lβ or in the LβI phase, depending on the initial dimension <strong>of</strong> the area.<br />

The range <strong>of</strong> a hh, in which hysteresis occurs, increases <strong>with</strong> decreasing temperature<br />

(see figure 5.14). This suggests that the transition Lβ to LβI is a first order transition.<br />

In the phase diagrams <strong>of</strong> figure 5.13 we define the phase found during decreasing<br />

temperature at a constant head-head repulsion parameter as the stable phase.<br />

Figure 5.14: Hysteresis curves for the area per <strong>lipid</strong>, AL <strong>of</strong> the <strong>lipid</strong> type ht8, as function <strong>of</strong><br />

the head-head repulsion parameter ahh at constant temperature T ∗ = 0.75 (solid line) <strong>and</strong><br />

T ∗ = 0.8 (dashed line). Configurations <strong>of</strong> both phases Lβ (at ahh = 6) <strong>and</strong> LβI (at ahh = 30) are<br />

taken as initial conditions, <strong>and</strong> ahh is slowly increased or decreased respectively.


5.2 Single-tail <strong>lipid</strong> <strong>bilayers</strong> 71<br />

a hh<br />

a hh<br />

a a hh a hh<br />

40<br />

40<br />

40<br />

30<br />

30<br />

20<br />

20<br />

L E,<br />

E,<br />

L E<br />

L D<br />

10 L E<br />

10<br />

0.7 0.7 0.8 0.9 1.0 1.1<br />

0.7 0.8 0.9 T*<br />

(a)<br />

T*<br />

1.0 1.1<br />

40<br />

40<br />

30<br />

30<br />

20<br />

L E,<br />

L E,<br />

L E,<br />

L D<br />

L D<br />

L D<br />

20<br />

L E<br />

10 L E<br />

10<br />

0.7<br />

0.7 0.8<br />

0.8<br />

0.9<br />

0.9<br />

1.0<br />

1.0<br />

1.1<br />

1.1<br />

0.7 0.8 0.9 T*<br />

T*<br />

T*<br />

1.0 1.1<br />

(c)<br />

a hh<br />

hh<br />

a hh<br />

a hh a hh<br />

hh<br />

40<br />

40<br />

40<br />

30 30<br />

30<br />

20 20<br />

20<br />

LL E,<br />

E,<br />

LEL E<br />

L L<br />

D D<br />

10 10 L E<br />

10<br />

0.7 0.8 0.8 0.9 0.9 1.0 1.0 1.1 1.1<br />

0.7 0.8 0.9 T* T*<br />

(b) T*<br />

1.0 1.1<br />

40 40<br />

40<br />

30 30<br />

30<br />

L E,<br />

LE, LE, L E,<br />

L D<br />

20 20<br />

LDL D<br />

L D<br />

20<br />

LEL E<br />

10<br />

10<br />

L E<br />

10<br />

0.7<br />

0.7<br />

0.8<br />

0.8<br />

0.9<br />

0.9<br />

1.0<br />

1.0<br />

1.1<br />

1.1<br />

0.7 0.8 0.9 T*<br />

T*<br />

T*<br />

1.0 1.1<br />

Figure 5.13: Phase diagrams as a function <strong>of</strong> the head-head repulsion parameter ahh <strong>and</strong> reduced<br />

temperature T ∗ for <strong>lipid</strong>s <strong>of</strong> different chain lengths: (a) ht6 (b) ht7 (c) ht8, <strong>and</strong> (d) ht9.<br />

(d)


72 Phase behavior <strong>of</strong> coarse-grained <strong>lipid</strong> <strong>bilayers</strong><br />

As we increase the tail length the gel phases are stabilized <strong>and</strong> the transition shifts<br />

to higher temperatures. The effect <strong>of</strong> increasing the head-head repulsion on the gel to<br />

liquid crystalline transition temperature is much more pronounced for the Lβ → Lα<br />

compared to LβI → Lα. This can be understood from the fact that in the interdigitated<br />

phase the average distance between the heads is already much larger compared<br />

to the non-interdigitated phase, <strong>and</strong> a further increase in this distance does not have<br />

a dramatic effect on the stability <strong>of</strong> the gel phase.<br />

For <strong>lipid</strong>s ht8 <strong>and</strong> ht9 the LβI phase occurs at slightly lower repulsion parameters<br />

than for <strong>lipid</strong>s ht6 <strong>and</strong> ht7. This is consistent <strong>with</strong> experimental results [91]. Since<br />

the interdigitated phase is more closely packed than the non-interdigitated phase,<br />

the van der Waals energy is greater. This energy gain is proportional to the number<br />

<strong>of</strong> carbon atoms in the phospho<strong>lipid</strong> chain <strong>and</strong> thus interdigitation becomes energetically<br />

more favorable for longer chains. Also in our simulations we observe that<br />

the interdigitated phase is more compact <strong>and</strong> hence a∗ hh decreases slightly <strong>with</strong> increasing<br />

tail length.<br />

It is interesting to compare these results <strong>with</strong> the experimental data. Misquitta<br />

<strong>and</strong> Caffrey in [108] systematically investigate the phase diagrams <strong>of</strong> monoacylglycerols,<br />

a single-tail <strong>lipid</strong>, <strong>and</strong> show a similar tail length dependence for the Lβ →Lα<br />

transition. Interestingly, as we will show later, for a similar model <strong>of</strong> a double-tail<br />

<strong>lipid</strong> we do not observe the spontaneous formation <strong>of</strong> an interdigitated phase. This<br />

corresponds to the experimental observation that for the most common double-tail<br />

<strong>lipid</strong>s the interdigitated phase does not form spontaneously, but should be induced<br />

by the addition <strong>of</strong>, for example, alcohol [98].<br />

The effect <strong>of</strong> adding salt on the gel to liquid crystalline transition has been studied<br />

for double-tail <strong>lipid</strong>s [109] <strong>and</strong> recently for single-tail <strong>lipid</strong>s [110]. These studies<br />

show that adding so-called kosmotropic salts increases the Lβ →Lα transition temperature,<br />

while chaotropic salts decrease this transition temperature. Similar effects<br />

have been observed for nonionic single-tail <strong>lipid</strong>s [111]. Takahashi et al. [110] explain<br />

these observations by assuming that kosmotropes tend to be excluded from the interfacial<br />

region <strong>and</strong> hence reduce the amount <strong>of</strong> interfacial water, while chaotropic<br />

salts have the inverse effects, i.e. are adsorbed at the interfacial region <strong>and</strong> increase<br />

the amount <strong>of</strong> interfacial water. In our model a similar effect can be achieved by<br />

changing the head-head interactions; increasing or decreasing a hh corresponds to<br />

adding chaotropes or kosmotropes, respectively. Our simulations show that decreasing<br />

the head-head repulsion stabilizes the Lβ phase, which corresponds to the case<br />

that water is excluded from the interface. Adding chaotropic salts has the reverse<br />

effect: it increases the head-head repulsion <strong>and</strong> stabilizes the Lα phase. Our simulations<br />

show that at sufficiently high head-head repulsion the interdigitated phase<br />

(LβI) is stable. This suggests that it might be possible in experiments to induce the<br />

Lβ →LβI phase transition by adding chaotropic salts to the systems.


5.3 Double-tail <strong>lipid</strong> <strong>bilayers</strong> 73<br />

5.3 Double-tail <strong>lipid</strong> <strong>bilayers</strong><br />

In the previous section we have discussed the phase behavior <strong>of</strong> single-tail <strong>lipid</strong> <strong>bilayers</strong>.<br />

In this section we extend the model <strong>and</strong> investigate the phase behavior <strong>of</strong><br />

double-tail <strong>lipid</strong>s. In particular, we consider a coarse-grained <strong>lipid</strong> <strong>with</strong> three headbeads<br />

<strong>and</strong> two tails <strong>of</strong> five beads each. This <strong>lipid</strong> will be denoted as h3(t5)2. As<br />

discussed in section 2.2 <strong>of</strong> Chapter 2, if a DPD bead is taken to represent the volume<br />

<strong>of</strong> three water molecules, i.e. 90 ˚A 3 , then the coarse-grained <strong>lipid</strong> h3(t5)2 can be<br />

considered as a model for dimyristoylphosphatidylcholine (DMPC) (see figure 2.2 in<br />

Chapter 2).<br />

5.3.1 Computational details<br />

For this study we consider <strong>bilayers</strong> <strong>of</strong> 900 h3(t5)2 <strong>lipid</strong>s, <strong>and</strong> 25 water particles per<br />

<strong>lipid</strong>, corresponding to fully hydrated conditions (i.e. no interaction <strong>of</strong> the bilayer<br />

<strong>with</strong> its periodic images).<br />

The values <strong>of</strong> the interaction parameters are aww = att = 25, a hh = 35, awt = a ht =<br />

80. Two consecutive beads in the <strong>lipid</strong> chain are connected by a harmonic spring,<br />

<strong>with</strong> equilibrium distance ro = 0.7, <strong>and</strong> elastic constant Kr = 100.<br />

Consecutive bonds in the <strong>lipid</strong> chain are subjected to a harmonic bond-bending<br />

potential. The values <strong>of</strong> the parameters related to this bond-bending potential were<br />

derived from the comparison <strong>of</strong> the CG model <strong>with</strong> MD simulations on all-atom<br />

model for a DMPC <strong>lipid</strong> bilayer [112]. The resulting values for the bending constant<br />

<strong>and</strong> the equilibrium angle in the <strong>lipid</strong> tails are Kθ=6 <strong>and</strong> θo=180 o , respectively. About<br />

the bond-bending potential between the head-bead connected to the <strong>lipid</strong> tails <strong>and</strong><br />

the first beads in the tails values <strong>of</strong> Kθ=3 <strong>and</strong> θo=90 o were found to reproduce the<br />

correct configurational distribution <strong>and</strong> structure <strong>of</strong> the atomistic detailed phospho<strong>lipid</strong>.<br />

All the simulations were carried at zero surface tension conditions. To explore the<br />

phase diagram <strong>of</strong> the bilayer, the temperature <strong>of</strong> the system was gradually decreased<br />

from T ∗ = 1.0 to T ∗ = 0.2 . At each temperature, a total <strong>of</strong> 100,000 DPD-MC cycles was<br />

performed, <strong>of</strong> which the first 20,000 cycles were needed for equilibration. Statistical<br />

averages were then collected over the next 80,000 DPD-MC cycles.<br />

The phase boundaries were detected, as described in the previous sections for<br />

the single-tail <strong>lipid</strong>s, by monitoring the temperature behavior <strong>of</strong> the area per <strong>lipid</strong>,<br />

the bilayer thickness <strong>and</strong> the order <strong>of</strong> the tails.<br />

5.3.2 Results <strong>and</strong> Discussion<br />

The area per <strong>lipid</strong>, AL, bilayer hydrophobic thickness, Dc, <strong>and</strong> tail order parameter<br />

S tail, as function <strong>of</strong> reduced temperature, T ∗ , are shown in Figure 5.15.


74 Phase behavior <strong>of</strong> coarse-grained <strong>lipid</strong> <strong>bilayers</strong><br />

No interdigitation was found for the chosen value <strong>of</strong> the repulsion parameter between<br />

the headgroups, a hh=35. Also, an increase <strong>of</strong> a hh up to 55 does not lead to any<br />

interdigitation (data not shown). This result is consistent <strong>with</strong> the experimentally<br />

observed structure <strong>of</strong> symmetric PC’s <strong>bilayers</strong>, for which no spontaneous interdigitation<br />

is found. The transition temperature was derived from the inflection points<br />

A L<br />

1.8<br />

1.6<br />

1.4<br />

1.2<br />

1.0<br />

0.1 0.3 0.5 0.7<br />

T*<br />

0.9 1.1<br />

(a)<br />

S tail<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

D c<br />

6.0<br />

5.5<br />

5.0<br />

4.5<br />

4.0<br />

3.5<br />

0.0<br />

0.1 0.3 0.5 0.7<br />

T*<br />

0.9 1.1<br />

(c)<br />

3.0<br />

0.1 0.3 0.5 0.7<br />

T*<br />

0.9 1.1<br />

Figure 5.15: Area per <strong>lipid</strong> AL (a), bilayer thickness Dc (b), <strong>and</strong> tail order parameter, Stail (c), as<br />

function <strong>of</strong> reduced temperature T ∗ for <strong>lipid</strong> type h3(t5)2. The error bars are smaller than the<br />

symbol size.<br />

<strong>of</strong> the shown curves. The system undergoes a main transition at a reduced temperature<br />

T∗ m=0.425. Above the reduced temperature T∗ m, the <strong>lipid</strong> chains are in the melted<br />

state (hence a low value <strong>of</strong> the bilayer thickness <strong>and</strong> <strong>of</strong> the tail order parameter) <strong>and</strong><br />

the system is in the Lα fluid phase. The snapshot in figure 5.16(c) shows a typical<br />

configuration <strong>of</strong> the system in the fluid phase.<br />

At very low temperatures the system is in the Lβ ′ gel phase, which is characterized<br />

by having ordered chains, hence a high value <strong>of</strong> the bilayer thickness <strong>and</strong> <strong>of</strong> the<br />

tail order parameter. While single-tail <strong>lipid</strong>s are not tilted in the gel phase, for the<br />

(b)


5.3 Double-tail <strong>lipid</strong> <strong>bilayers</strong> 75<br />

(a) L β ′ (b) P β ′<br />

(c) Lα<br />

Figure 5.16: Snapshots <strong>of</strong> typical configurations <strong>of</strong> the h3(t5)2 bilayer simulated at reduced<br />

temperatures: (a) T ∗ < 0.35, corresponding to the gel phase, or Lβ ′; (b) 0.35 ≤ T∗ < 0.425<br />

corresponding to the ripple-like ’striated’ phase, or Pβ ′; <strong>and</strong> (c) T∗ > 0.425 corresponding to<br />

the fluid phase, or Lα. The <strong>lipid</strong> headgroups are represented by black lines <strong>and</strong> the <strong>lipid</strong> tails<br />

by gray lines, <strong>with</strong> the terminal tail beads darker gray. The water is not shown.<br />

double-tail <strong>lipid</strong> we observe that the <strong>lipid</strong> chains are tilted <strong>with</strong> respect to the bilayer<br />

normal. We find a tilt angle <strong>of</strong> 25 o , which is slightly lower than the value <strong>of</strong> ≈ 32 o<br />

measured experimentally for DMPC <strong>lipid</strong> <strong>bilayers</strong> [113]. A typical configuration at<br />

this temperature can be see in the snapshot in figure 5.16(a).<br />

Between the Lα <strong>and</strong> the Lβ ′ phases, when the temperature is increased above<br />

T∗ =0.35, we observed a third phase. This phase, which disappears again as the temperature<br />

reaches the main-transition temperature, is characterized by having striated<br />

regions made <strong>of</strong> <strong>lipid</strong>s in the gel-state intercalated by regions made <strong>of</strong> <strong>lipid</strong>s<br />

in the fluid-state. This modulated structure can be seen in the snapshot in figure<br />

5.16(b). This phase resembles the Pβ ′, or ripple-phase. The ripple-phase occurs in<br />

phospho<strong>lipid</strong> <strong>bilayers</strong> at the so-called pre-transition temperature, <strong>and</strong> is characterized<br />

by a rippling <strong>of</strong> the bilayer, <strong>with</strong> a wave length <strong>of</strong> the order <strong>of</strong> 150 ˚A [114].<br />

As we have discussed in Chapters 2 <strong>and</strong> 3, the double-tail <strong>lipid</strong> h3(t5)2 can be<br />

mapped onto DMPC, if a coarse-grained representation is used in which one DPD<br />

bead has a volume <strong>of</strong> 90 ˚A 3 . This correspondence between the <strong>lipid</strong> h3(t5)2 <strong>and</strong><br />

DMPC allows us to quantitatively compare the values <strong>of</strong> structural quantities found<br />

in our simulations <strong>with</strong> experimentally measured values. Besides the unit <strong>of</strong> length,<br />

which is derived, as discussed in Chapter 2, from the volume <strong>of</strong> one DPD bead, <strong>and</strong><br />

is equal to Rc = 6.4633˚A, we need to map the reduced temperature onto real temper-


76 Phase behavior <strong>of</strong> coarse-grained <strong>lipid</strong> <strong>bilayers</strong><br />

ature.<br />

To give an estimate <strong>of</strong> the values <strong>of</strong> the reduced temperatures, T ∗ , in terms <strong>of</strong><br />

physical temperatures T, we have mapped T ∗ onto T according to the following linear<br />

relation<br />

T = aT ∗ + b. (5.1)<br />

The values <strong>of</strong> the coefficients a <strong>and</strong> b were found by solving the system <strong>of</strong> linear equations<br />

obtained by substituting in equation 5.1 the reduced <strong>and</strong> physical values <strong>of</strong> the<br />

main- <strong>and</strong> pre-transition temperatures for a pure DMPC phospho<strong>lipid</strong> bilayer. From<br />

the values ≈ 24 o C <strong>and</strong> ≈ 14 o C [96] <strong>of</strong> the main- <strong>and</strong> pre-transition temperatures,<br />

respectively, we obtained a = 133 o C, <strong>and</strong> b = −33 o C.<br />

The values <strong>of</strong> the bilayer hydrophobic thickness <strong>and</strong> the area per <strong>lipid</strong> obtained<br />

from our simulations can now be compared <strong>with</strong> the corresponding experimental<br />

values for fully hydrated DMPC <strong>bilayers</strong>, as shown in Table 5.1. The values from sim-<br />

Table 5.1: Values from simulation <strong>of</strong> the bilayer hydrophobic thickness, Dc, <strong>and</strong> area per <strong>lipid</strong>,<br />

AL, at different temperatures, compared <strong>with</strong> experimental values. The error on the simulation<br />

data is 0.2 ˚A for Dc <strong>and</strong> 0.4 ˚A 2 for AL.<br />

T [ o C] Phase Dc [˚A] AL [˚A 2 ]<br />

sim exper sim exper<br />

10 Lβ ′ 34.3 30.3† 48.6 47.2 †<br />

30 Lα 26.3 25.6 ‡ 60.4 60.0 ‡<br />

50 Lα 24.3 24.0 ‡ 64.4 65.4 ‡<br />

65 Lα 23.6 23.4 ‡ 65.7 68.5 ‡<br />

† From [113]. The error for Dc is 0.2 ˚A, <strong>and</strong> for AL 0.5 ˚A 2 .<br />

‡ From [92]. The error is not reported in the cited reference.<br />

ulations are in good quantitative agreement <strong>with</strong> the experimental data, although<br />

some deviations from the experimental values are observed for the area per <strong>lipid</strong> at<br />

high temperature (65 o C), <strong>and</strong> for the bilayer hydrophobic thickness in the gel phase<br />

(10 o C). The larger thickness, compared <strong>with</strong> the experimental value, found in our<br />

simulations in the gel phase, could be due to the smaller tilt angle shown by the<br />

coarse grained <strong>lipid</strong>s compared to the DMPC <strong>lipid</strong>s.<br />

5.4 Conclusion<br />

We performed simulations on the most simple representation <strong>of</strong> a phospho<strong>lipid</strong>. The<br />

model consists <strong>of</strong> a single hydrophilic head bead connected to one tail <strong>of</strong> hydrophobic<br />

beads, which can vary in length. Using the area per <strong>lipid</strong>, the hydrophobic thick-


5.4 Conclusion 77<br />

ness, the order parameter, <strong>and</strong> the extent <strong>of</strong> chain overlap, we were able to characterize<br />

various bilayer phases.<br />

The simulations showed that different stable phases are obtained for a wide range<br />

<strong>of</strong> temperatures. We characterized the low temperature phase as a gel phase, <strong>and</strong><br />

we reproduced the main order/disorder phase transition from a gel to a liquid crystalline<br />

phase. This transition temperature to the Lα phase increases <strong>with</strong> increasing<br />

tail length, as was also found experimentally.<br />

In <strong>bilayers</strong> consisting <strong>of</strong> single-tail <strong>lipid</strong>s, only the non-interdigitated Lβ phase<br />

<strong>and</strong> the fluid Lα phase are observed. However, experiments show that if chaotropic<br />

salts are added to the system, the distance between the headgroups is increased, stabilizing<br />

the Lα phase. We show that at high enough head-head repulsion the Lα<br />

phase is indeed stabilized <strong>and</strong> that the low temperature phase is the interdigitated<br />

LβI phase. This suggests that it is possible to induce an interdigitated phase in <strong>bilayers</strong><br />

<strong>of</strong> single-tail <strong>lipid</strong>s by adding chaotropic salts to the system.<br />

From our results, we can conclude that a model consisting <strong>of</strong> a head bead connected<br />

to a single tail does not describe the phase behavior <strong>of</strong> a double-tail <strong>lipid</strong> correctly.<br />

Single-tail <strong>lipid</strong>s spontaneously form a low temperature interdigitated phase<br />

for high enough values <strong>of</strong> the repulsion parameter between headgroups, while experimentally<br />

this phase is observed in <strong>bilayers</strong> consisting <strong>of</strong> double-tail <strong>lipid</strong>s, but<br />

has to be induced. By lowering the value <strong>of</strong> the head-head repulsion the low temperature<br />

phase is the Lβ phase. In this phase the chains <strong>of</strong> single-tails <strong>lipid</strong>s are ordered<br />

parallel to the bilayer normal, while for most common double-tail phospho<strong>lipid</strong>s the<br />

hydrocarbon tails show a tilt <strong>with</strong> respect to the bilayer normal (the Lβ ′).<br />

We have hence increased the topological complexity <strong>of</strong> the model by considering<br />

model <strong>lipid</strong>s <strong>with</strong> two hydrophobic tails. We have shown that the <strong>bilayers</strong> formed<br />

by these <strong>lipid</strong>s display a phase behavior as function <strong>of</strong> temperature that well reproduces<br />

the phase behavior <strong>of</strong> phosphatidylcholine <strong>bilayers</strong>. Moreover, by defining a<br />

coarse-graining level in which a DPD bead corresponds to a volume <strong>of</strong> 90 ˚A 3 , i.e. to<br />

three water molecules or three methyl groups, we have defined a correspondence between<br />

a coarse-grained <strong>lipid</strong> <strong>with</strong> three head-beads <strong>and</strong> two tails <strong>of</strong> five beads each<br />

<strong>and</strong> the DMPC phospho<strong>lipid</strong>. By also mapping the model reduced temperature, via<br />

the phase transition temperatures <strong>of</strong> DMPC <strong>bilayers</strong>, onto physical units, we were<br />

able to compare the simulated bilayer structural properties <strong>with</strong> their corresponding<br />

experimental values. A good agreement in the trends <strong>and</strong> values was found. This<br />

agreement validates our mesoscopic model for <strong>lipid</strong> <strong>bilayers</strong>.


VI<br />

Interaction <strong>of</strong> small molecules <strong>with</strong> <strong>bilayers</strong>


80 Interaction <strong>of</strong> small molecules <strong>with</strong> <strong>bilayers</strong><br />

6.1 Introduction<br />

The role <strong>of</strong> the pressure pr<strong>of</strong>iles in <strong>lipid</strong> <strong>bilayers</strong> is an important, <strong>and</strong> controversial,<br />

issue in the study <strong>of</strong> <strong>lipid</strong>-protein interactions. Non-bilayer <strong>lipid</strong>s, i.e. <strong>lipid</strong>s which<br />

preferentially do not aggregate into a bilayer phase [115], <strong>and</strong> molecules like alcohols<br />

or anesthetics, may alter <strong>and</strong> shift the distribution <strong>of</strong> pressure across a bilayer, <strong>and</strong><br />

in this way the activity <strong>of</strong> so-called mechanosensitive transmembrane proteins [116]<br />

can be regulated. To date, the way anesthetics work is not yet fully understood, <strong>and</strong><br />

the controversy as to whether the primary mechanism <strong>of</strong> action <strong>of</strong> general anesthetics<br />

is through a <strong>lipid</strong> mediated, non-specific, interaction, or via a direct action on<br />

the membrane proteins is still a matter <strong>of</strong> debate (for an exhaustive discussion <strong>of</strong><br />

the problem see [117] <strong>and</strong> references therein). An interesting hypothesis was proposed<br />

by Cantor [29, 30, 114, 118]. Cantor suggested that a purely mechanical, <strong>lipid</strong>mediated,<br />

effect might be responsible for the action <strong>of</strong> general anesthetics through<br />

an anesthetic-induced redistribution <strong>of</strong> the lateral pressure across the bilayer. Consider<br />

an intrinsic membrane protein that can exist in two principal configurations, <strong>of</strong><br />

which one is active <strong>and</strong> the other one inactive. If the two configurations display a different<br />

distribution <strong>of</strong> the cross-sectional area as function <strong>of</strong> the depth in the bilayer,<br />

a redistribution <strong>of</strong> lateral pressure across the bilayer might shift the conformational<br />

equilibria <strong>of</strong> the protein.<br />

In Chapter 3 we have shown that pressure pr<strong>of</strong>iles can be directly calculated in<br />

molecular simulations, <strong>and</strong> in Chapter 4 we have discussed how the <strong>lipid</strong>s characteristics,<br />

like chain length <strong>and</strong> stiffness, can determine the distribution <strong>of</strong> the lateral<br />

pressure across a bilayer. In this Chapter we investigate the influence <strong>of</strong> the bilayer<br />

local density <strong>and</strong> pressure on the partitioning <strong>of</strong> small, dumb-bell like, solutes in <strong>bilayers</strong><br />

<strong>of</strong> <strong>lipid</strong>s <strong>with</strong> different topology. We also study the modifications induced by<br />

the solutes on the bilayer structural properties, <strong>and</strong> in particular on the distribution<br />

<strong>of</strong> local pressures.<br />

6.2 Computational details<br />

A dumb-bell molecule consists <strong>of</strong> two beads tied together by a harmonic spring. Different<br />

chemical properties <strong>of</strong> these dumb-bells can be represented by varying the<br />

strength <strong>of</strong> the interaction parameters <strong>of</strong> the dumb-bell beads <strong>with</strong> the other components<br />

<strong>of</strong> the system, i.e. water <strong>and</strong> <strong>lipid</strong>s.<br />

We consider three types <strong>of</strong> dumb-bells: purely hydrophobic (indicated in the following<br />

as dbH), in which both beads have a large repulsion parameter <strong>with</strong> water <strong>and</strong><br />

the <strong>lipid</strong> headgroup; neutral dumb-bells (dbN), in which both beads interact <strong>with</strong><br />

the other components as if they were the same specie; <strong>and</strong> amphiphilic dumb-bells<br />

(dbA), in which one bead is hydrophilic <strong>and</strong> the other one is hydrophobic. The values<br />

<strong>of</strong> the interaction parameters for the different dumb-bells are reported in table 6.1.


6.2 Computational details 81<br />

For reference purpose we also report the interaction parameters between water <strong>and</strong><br />

<strong>lipid</strong>s. By choosing these three different types <strong>of</strong> solute molecules we can discriminate<br />

between the chemical affinity <strong>and</strong> the mechanic effects on the distribution <strong>of</strong><br />

the solutes in the bilayer.<br />

dbA dbH dbN<br />

a (db1,w) 25 80 25 a (w,w) 25<br />

a (db1,h) 25 80 35 a (w,h) 15<br />

a (db1,t) 50 25 25 a (w,t) 80<br />

a (db2,w) 80 80 25 a (h,h) 35<br />

a (db2,h) 80 80 25 a (h,t) 80<br />

a (db2,t) 25 25 25 a (t,t) 25<br />

a (db1,db1) 25 25 25<br />

a (db2,db2) 25 25 25<br />

a (db1,db2) 25 25 25<br />

Table 6.1: Repulsion parameters aij (for the force <strong>of</strong> equation F C ij = aij(1−rij/Rc)^rij) between<br />

dumb-bells, water, <strong>and</strong> <strong>lipid</strong>s. Water beads are indicated as w, <strong>lipid</strong> hydrophilic head-beads as<br />

h, <strong>lipid</strong> hydrophobic tail-beads as t, <strong>and</strong> the two beads <strong>of</strong> a dumb-bell as db1 <strong>and</strong> db2 respectively.<br />

Three types <strong>of</strong> dumb-bells are considered: hydrophobic (dbH), amphiphilic (dbA) <strong>and</strong><br />

neutral (dbN).<br />

To investigate the influence <strong>of</strong> the <strong>lipid</strong> characteristics on the distribution <strong>of</strong> the<br />

solutes, we consider three <strong>lipid</strong> types whose density distribution <strong>and</strong> lateral pressure<br />

pr<strong>of</strong>ile show different shapes. All the considered <strong>lipid</strong>s have a single tail <strong>of</strong> 7<br />

hydrophobic beads attached to one hydrophilic head-bead. The <strong>lipid</strong> tail can have<br />

different stiffness. We consider <strong>bilayers</strong> formed by fully flexible <strong>lipid</strong>s (ht7), by linear,<br />

t), <strong>and</strong> by poly-unsaturated <strong>lipid</strong>s, where the unsaturation points are<br />

stiff <strong>lipid</strong>s (ht (L)<br />

6<br />

represented by kinks along the chain (ht (L) t (K)<br />

4 t(L) t). The <strong>lipid</strong>s nomenclature follows<br />

the convention introduced in section 4.3 <strong>of</strong> Chapter 4, <strong>and</strong> the different <strong>lipid</strong> topologies<br />

used in this work are shown in figure 6.1. In all cases, the solutes were added<br />

to equilibrated, tensionless <strong>bilayers</strong> consisting <strong>of</strong> 200 <strong>lipid</strong>s (100 in each monolayer),<br />

<strong>and</strong> approximately 2000 water beads, at an overall bead density <strong>of</strong> 3. The dumb-bells<br />

mole fraction (respect to the number <strong>of</strong> <strong>lipid</strong>s) was 25% (the same used in [119]),<br />

which results in 50 dumb-bells for the <strong>bilayers</strong> considered here.<br />

The initial positions <strong>of</strong> the dumb-bells were chosen at r<strong>and</strong>om <strong>with</strong>in the hydrophobic<br />

part <strong>of</strong> the bilayer. The system was then equilibrated for 5000 DPD steps<br />

to allow the added molecules to diffuse in the bilayer. After this equilibration period<br />

the molecules were already located in the bilayer regions where they would reside<br />

throughout the rest <strong>of</strong> the simulation. Further 20000 MC-DPD cycles <strong>with</strong> imposed<br />

zero surface tension concluded the equilibration. Density <strong>and</strong> pressure pr<strong>of</strong>iles, <strong>and</strong><br />

bilayer structural characteristics were measured over 50000 MC-DPD cycles (at zero<br />

surface tension) <strong>with</strong> sampling frequency <strong>of</strong> 10 cycles.


82 Interaction <strong>of</strong> small molecules <strong>with</strong> <strong>bilayers</strong><br />

t<br />

t<br />

t<br />

t<br />

ht7<br />

t<br />

t<br />

t<br />

t (L)<br />

t (L)<br />

t (L)<br />

t (L)<br />

t (L)<br />

t (L)<br />

t<br />

ht 6t<br />

(L)<br />

t (L)<br />

t (K)<br />

t (K)<br />

t (L)<br />

t (K)<br />

t (K)<br />

h t 4<br />

(K)<br />

t (L)<br />

t (L)<br />

Figure 6.1: Schematic representation <strong>of</strong> the three <strong>lipid</strong> types used in this study <strong>with</strong> their<br />

nomenclature. The head-bead is represented by a black particle <strong>and</strong> is denoted as ’h’. The<br />

tail-beads are represented by white particles <strong>and</strong> have different names depending on the presence,<br />

<strong>and</strong> type, <strong>of</strong> a bond-bending potential. A bead labeled ’t (L) ’ is the central bead in a<br />

bond-bending potential <strong>with</strong> equilibrium angle θo = 180 o , a bead labeled ’t (K) ’ is the central<br />

bead in a bond-bending potential <strong>with</strong> θo = 135 o , while a bead labeled as ’t’ does not participate<br />

to any bond-bending potential. The subscript indicates the number <strong>of</strong> beads <strong>of</strong> a given<br />

type that are tied together.<br />

6.3 Results <strong>and</strong> Discussion<br />

6.3.1 Partitioning <strong>of</strong> the solutes in the bilayer<br />

Figure 6.2 shows the snapshots <strong>of</strong> the equilibrium configuration <strong>of</strong> the dbA, dbH <strong>and</strong><br />

dbN molecules in the poly-unsaturated bilayer. The distributions <strong>of</strong> the solutes in<br />

the other bilayer types are similar to the ones shown here. The partitioning at equilibrium<br />

<strong>of</strong> the solutes molecules depends on the chemical interactions <strong>with</strong> the bilayer<br />

components <strong>and</strong> <strong>with</strong> water: the amphiphilic molecules are found near the<br />

<strong>lipid</strong> headgroup (figure 6.2(a)) while the hydrophobic molecules are located in the<br />

bilayer core (figure 6.2(b)). Only the neutral molecules (figure 6.2(c)) have significantly<br />

diffused into the water phase. As a general observation, we see that the dumbbells<br />

have higher density in the energetically more favorable bilayer region, i.e. where<br />

the repulsion is lower. However, the neutral molecules that have not diffused in the<br />

water, are found predominantly near the <strong>lipid</strong> headgroup region (figure 6.2(c)). Because<br />

<strong>of</strong> their neutrality, the preference for these molecules to absorb at the interface<br />

must be due to thermodynamic factors, like density <strong>and</strong> pressure distributions<br />

t<br />

t


6.3 Results <strong>and</strong> Discussion 83<br />

<strong>with</strong>in the bilayer. The neutral dumb-bells locate in fact in the bilayer region <strong>with</strong><br />

the lowest density <strong>and</strong> pressure, which corresponds to the interface between the hydrophilic<br />

headgroups <strong>and</strong> the hydrophobic tails <strong>of</strong> the <strong>lipid</strong>s. These observations<br />

are supported by the shape <strong>of</strong> the density pr<strong>of</strong>iles as shown in figures 6.3, 6.4, <strong>and</strong> 6.5.<br />

Considering the interaction energy only, purely hydrophobic dumb-bells should be<br />

found <strong>with</strong> uniform distribution <strong>with</strong>in the bilayer core. However, when we compare<br />

the location <strong>of</strong> the hydrophobic dumb-bells <strong>with</strong> the shape <strong>of</strong> the lateral pressure<br />

<strong>with</strong>in the pure bilayer (also shown in the figures), we observe that the distribution<br />

<strong>of</strong> the solute molecules is not uniform, but follows the distribution <strong>of</strong> the <strong>lipid</strong> density<br />

<strong>and</strong> pressure <strong>with</strong>in the bilayer core. The hydrophobic dumb-bells locate in the<br />

regions <strong>of</strong> lower hydrocarbon density <strong>and</strong> pressure. These observations suggest that<br />

the distributions <strong>of</strong> density <strong>and</strong> pressure in the bilayer is an important factor in predicting<br />

the absorption sites for small solute molecules.<br />

6.3.2 Effect on the bilayer properties<br />

By comparing the density pr<strong>of</strong>iles in presence <strong>of</strong> solutes <strong>with</strong> the ones for the pure<br />

bilayer, we can study how the bilayer structure is modified by the addition <strong>of</strong> these<br />

molecules.<br />

As shown in Chapter 4, for stiff <strong>lipid</strong>s <strong>with</strong> the headgroup repulsion parameter<br />

used here, the structure <strong>of</strong> the pure system is a largely interdigitated bilayer. If the<br />

dumb-bells are located at the interface (amphiphilic <strong>and</strong> neutral) in this kind <strong>of</strong> bilayer<br />

(figures 6.3(c) <strong>and</strong> 6.3(d)) the central density peak in the bilayer hydrophobic<br />

core increases in magnitude, while the secondary peaks at the edge <strong>of</strong> the hydrophobic<br />

core decrease in magnitude. Both these features in the density pr<strong>of</strong>ile are a signature<br />

for an increase in interdigitation. The opposite happens when hydrophobic<br />

dumb-bells are added (figure 6.3(b)). Since these molecules adsorb in the bilayer<br />

hydrophobic core, they decrease the packing <strong>of</strong> the <strong>lipid</strong> tails <strong>and</strong> increase their disorder.<br />

As a consequence, the bilayer is less interdigitated; the peaks in the density<br />

<strong>of</strong> the last bead <strong>of</strong> the tail shift toward the bilayer center, <strong>and</strong> the overall density in<br />

the bilayer core becomes more uniform. For the flexible <strong>and</strong> the unsaturated <strong>bilayers</strong><br />

(figures 6.5 <strong>and</strong> 6.4), the addition <strong>of</strong> solutes has a smaller effect on the <strong>lipid</strong>s density<br />

distribution. A general observation for both the flexible <strong>and</strong> the unsaturated bilayer<br />

is that the addition <strong>of</strong> hydrophobic molecules decreases the central minimum (at<br />

z = 0) <strong>of</strong> the <strong>lipid</strong> tail density. This effect can be easily explained considering that the<br />

hydrophobic dumb-bells adsorb in the bilayer inner core, <strong>and</strong> hence decrease the<br />

space available to the <strong>lipid</strong> tails in this region.<br />

To further characterize the modifications induced by the solute molecules on the<br />

bilayer, we compare the structural quantities <strong>of</strong> the mixed bilayer <strong>with</strong> the ones <strong>of</strong>


84 Interaction <strong>of</strong> small molecules <strong>with</strong> <strong>bilayers</strong><br />

(a) dbA (b) dbH<br />

(c) dbN<br />

Figure 6.2: Snapshots <strong>of</strong> configurations taken from simulations <strong>of</strong> absorption <strong>of</strong> dumb-bell<br />

molecules in a bilayer <strong>of</strong> poly-unsaturated ht (L) t (K)<br />

4 t (L) t <strong>lipid</strong>s. (a) Amphiphilic, (b) hydrophobic,<br />

(c) neutral dumb-bells. The dumb-bells are represented as large spheres, the <strong>lipid</strong> tails as<br />

lines, the <strong>lipid</strong> headgroups as dark small spheres, <strong>and</strong> the water particles as light small spheres.<br />

The simulation box is periodic in all three Cartesian directions.


6.3 Results <strong>and</strong> Discussion 85<br />

ρ(z)<br />

0.3<br />

0.2<br />

0.1<br />

0.0<br />

−0.1<br />

−6 −4 −2 0<br />

Z<br />

2 4 6<br />

0.0<br />

−0.1<br />

(a) Pure bilayer<br />

−0.2<br />

−6 −4 −2 0<br />

Z<br />

2 4 6<br />

w<br />

h<br />

t (1,..,n−1)<br />

t n<br />

ρ tot<br />

0.3 w<br />

h<br />

0.2<br />

t (1,..,n−1)<br />

tn db<br />

0.1<br />

π(z)<br />

(c) dbA<br />

0.3 w<br />

h<br />

0.2<br />

t (1,..,n−1)<br />

tn db<br />

0.1<br />

π(z)<br />

0.0<br />

−0.1<br />

−0.2<br />

−6 −4 −2 0<br />

Z<br />

2 4 6<br />

0.0<br />

−0.1<br />

(b) dbH<br />

0.3 w<br />

h<br />

0.2<br />

t (1,..,n−1)<br />

tn db<br />

0.1<br />

π(z)<br />

−0.2<br />

−6 −4 −2 0<br />

Z<br />

2 4 6<br />

(d) dbN<br />

Figure 6.3: Density pr<strong>of</strong>iles for the pure ht (L)<br />

6 t bilayer (a), <strong>and</strong> for the same bilayer <strong>with</strong> (b)<br />

hydrophobic (dbH), (c) amphiphilic (dbA), <strong>and</strong> (d) neutral dumb-bells (dbN). In figures (b),<br />

(c) <strong>and</strong> (d) the lateral pressure pr<strong>of</strong>ile π(z) <strong>of</strong> the pure bilayer is also plotted. Note that the<br />

densities have been rescaled to fit in the same scale as the lateral pressure (hence the values<br />

on the graphs ordinate refer to π(z) only). The scaling factor for the density <strong>of</strong> the dumb-bells<br />

is two times smaller than the scaling for water <strong>and</strong> <strong>lipid</strong> densities. The densities refer to water<br />

(w), <strong>lipid</strong> headgroup (h), <strong>lipid</strong> hydrophobic tail-beads excluding the last one (t(1,...,n−1)), last<br />

tail-bead (tn), <strong>and</strong> dumb-bells (db). For the pure bilayer the total density ρtot is also plotted.


86 Interaction <strong>of</strong> small molecules <strong>with</strong> <strong>bilayers</strong><br />

ρ(z)<br />

0.3<br />

0.2<br />

0.1<br />

0.0<br />

−0.1<br />

−6 −4 −2 0<br />

Z<br />

2 4 6<br />

0.0<br />

−0.1<br />

(a) Pure bilayer<br />

−0.2<br />

−6 −4 −2 0<br />

Z<br />

2 4 6<br />

w<br />

h<br />

t (1,..,n−1)<br />

t n<br />

ρ tot<br />

0.3 w<br />

h<br />

0.2<br />

t (1,..,n−1)<br />

tn db<br />

0.1<br />

π(z)<br />

(c) dbA<br />

0.3 w<br />

h<br />

0.2<br />

t (1,..,n−1)<br />

tn db<br />

0.1<br />

π(z)<br />

0.0<br />

−0.1<br />

−0.2<br />

−6 −4 −2 0<br />

Z<br />

2 4 6<br />

0.0<br />

−0.1<br />

(b) dbH<br />

0.3 w<br />

h<br />

0.2<br />

t (1,..,n−1)<br />

tn db<br />

0.1<br />

π(z)<br />

−0.2<br />

−6 −4 −2 0<br />

Z<br />

2 4 6<br />

(d) dbN<br />

Figure 6.4: ht (L) t (K)<br />

4 t (L) t bilayer. See caption <strong>of</strong> figure 6.3.


6.3 Results <strong>and</strong> Discussion 87<br />

ρ(z)<br />

0.3<br />

0.2<br />

0.1<br />

0.0<br />

−0.1<br />

−6 −4 −2 0<br />

Z<br />

2 4 6<br />

0.0<br />

−0.1<br />

(a) Pure bilayer<br />

−0.2<br />

−6 −4 −2 0<br />

Z<br />

2 4 6<br />

w<br />

h<br />

t (1,..,n−1)<br />

t n<br />

ρ tot<br />

0.3 w<br />

h<br />

0.2<br />

t (1,..,n−1)<br />

tn db<br />

0.1<br />

π(z)<br />

(c) dbA<br />

0.3 w<br />

h<br />

0.2<br />

t (1,..,n−1)<br />

tn db<br />

0.1<br />

π(z)<br />

0.0<br />

−0.1<br />

−0.2<br />

−6 −4 −2 0<br />

Z<br />

2 4 6<br />

0.0<br />

−0.1<br />

(b) dbH<br />

0.3 w<br />

h<br />

0.2<br />

t (1,..,n−1)<br />

tn db<br />

0.1<br />

π(z)<br />

−0.2<br />

−6 −4 −2 0<br />

Z<br />

2 4 6<br />

(d) dbN<br />

Figure 6.5: ht7 bilayer. See caption <strong>of</strong> figure 6.3.


88 Interaction <strong>of</strong> small molecules <strong>with</strong> <strong>bilayers</strong><br />

the pure bilayer. We consider the parameter<br />

∆Q = Qdb − Qo<br />

, (6.1)<br />

Qo<br />

where Q can be one <strong>of</strong> the following: bilayer area, bilayer thickness, <strong>lipid</strong> hydrophobic<br />

length (projected along the bilayer normal). Qo <strong>and</strong> Q db refer to any <strong>of</strong> these<br />

quantities in the pure bilayer <strong>and</strong> in the bilayer <strong>with</strong> dumb-bells, respectively. For<br />

the definition <strong>and</strong> method <strong>of</strong> calculation <strong>of</strong> these quantities we refer the reader to<br />

section 4.2 <strong>of</strong> Chapter 4. The values <strong>of</strong> these differences for the different dumb-bells<br />

<strong>and</strong> bilayer types are shown in figure 6.6. By addition <strong>of</strong> the solute, the bilayer area<br />

∆Q %<br />

15<br />

10<br />

5<br />

0<br />

−5<br />

−10<br />

A<br />

D c<br />

n<br />

Lee A<br />

D c<br />

n<br />

Lee D c<br />

dbA dbH dbN<br />

(a) ht (L)<br />

6 t<br />

A<br />

n<br />

Lee ∆Q %<br />

15<br />

10<br />

5<br />

0<br />

−5<br />

−10<br />

A<br />

n<br />

Dc Lee A<br />

Dc<br />

n<br />

Lee n<br />

Dc Lee dbA dbH dbN<br />

(b) ht (L) t (K)<br />

4 t (L) t<br />

A<br />

∆Q %<br />

15<br />

10<br />

5<br />

0<br />

−5<br />

−10<br />

A<br />

n<br />

Dc Lee A<br />

Dc n<br />

Lee dbA dbH dbN<br />

(c) ht7<br />

Figure 6.6: Difference (as percent), <strong>with</strong> respect to the pure bilayer, <strong>of</strong> bilayer structural quantities<br />

(Q) by addition <strong>of</strong> different solute molecules in different bilayer types. The quantities<br />

considered are: bilayer area, A (black), bilayer hydrophobic thickness, Dc (black <strong>and</strong> white),<br />

<strong>and</strong> <strong>lipid</strong> hydrophobic end-to-end distance projected along the bilayer normal, L n ee (gray).<br />

(in black in figures 6.6) increases for all bilayer types, although the increase is much<br />

larger when the molecules absorb at the interface (amphiphilic <strong>and</strong> neutral dumbbells).<br />

Because we are considering the total area, <strong>and</strong> because the <strong>bilayers</strong> are tensionless,<br />

a larger number <strong>of</strong> molecules in the system results in a larger area compared<br />

to the pure bilayer. Moreover, if the solutes adsorb at the headgroup interfacial region<br />

this increase is larger. This increase in the bilayer area results in a lower interfacial<br />

density <strong>of</strong> <strong>lipid</strong>s, <strong>and</strong> in a potential creations <strong>of</strong> voids in the bilayer core. The<br />

rearrangement <strong>of</strong> the bilayer to compensate for this is very much dependent on the<br />

bilayer structure <strong>and</strong> on the type <strong>of</strong> solute. We will first consider the case <strong>of</strong> the ht (L)<br />

6 t<br />

bilayer, <strong>with</strong> either the neutral or the amphiphilic solute molecules, which both absorb<br />

at the interface near the <strong>lipid</strong> headgroups. Since the dumb-bells at the interface<br />

are far too short to fill-in the voids created in the bilayer interior by an increase <strong>of</strong> the<br />

surface area, two compensating mechanisms could happen. The <strong>lipid</strong> tails could become<br />

more disordered, or the <strong>lipid</strong>s in each opposing monolayer could interdigitate<br />

to fill-in the extra free volume. The decrease in bilayer hydrophobic thickness (<strong>and</strong><br />

the described shape <strong>of</strong> the density pr<strong>of</strong>iles) show that the latter mechanism is taking<br />

A<br />

D n<br />

cLee


6.3 Results <strong>and</strong> Discussion 89<br />

S m<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

1 2 3 4 5 6 7<br />

m<br />

(a)<br />

pure bilayer<br />

dbA<br />

dbH<br />

dbN<br />

P(φ)<br />

0.03<br />

0.02<br />

0.01<br />

0.00<br />

0 20 40<br />

φ<br />

60 80 100<br />

(b)<br />

pure bilayer<br />

dbA<br />

dbH<br />

dbN<br />

Figure 6.7: Modification <strong>of</strong> the ordering <strong>of</strong> the <strong>lipid</strong>s induced by addition <strong>of</strong> solute dumb-bells<br />

in a ht (L)<br />

6 t bilayer. (a) Order parameter Sm between the vector connecting each two consecutive<br />

beads in the <strong>lipid</strong> tail <strong>and</strong> the bilayer normal, m is the index <strong>of</strong> the vector (starting from<br />

the head group). (b) Distribution <strong>of</strong> the angle (in degrees) formed by the vector between the<br />

head-group <strong>and</strong> the last bead in the <strong>lipid</strong> tail <strong>with</strong> the bilayer normal.<br />

place. This conclusion is further supported by the observed increase <strong>of</strong> the <strong>lipid</strong> endto-end<br />

distance along the bilayer normal. If the two monolayers were well separated,<br />

an increase <strong>of</strong> the <strong>lipid</strong> tail length would correspond to an increase <strong>of</strong> the bilayer<br />

thickness, while we observe a decrease <strong>of</strong> the latter quantity. Because <strong>of</strong> increased<br />

interdigitation, the <strong>lipid</strong> tails are more ordered. This can be seen (figure 6.7(a)) from<br />

the increase in the tail order parameters respect to the pure bilayer <strong>and</strong> to the bilayer<br />

<strong>with</strong> hydrophobic dumb-bells. Also, in the case where the solutes adsorb at<br />

the interface, the distribution <strong>of</strong> the angle between the <strong>lipid</strong> tail <strong>and</strong> the bilayer normal<br />

(figure 6.7(b)) is less broad, <strong>and</strong> <strong>with</strong> the maximum at lower values <strong>of</strong> the angle.<br />

Also in the case <strong>of</strong> flexible (ht7) <strong>and</strong> unsaturated (ht (L) t (K)<br />

4 t(L) t) <strong>bilayers</strong> <strong>with</strong> added<br />

surface-absorbed molecules, to an increase in surface area corresponds a decrease<br />

<strong>of</strong> bilayer thickness. However, this decrease can be accounted for by a shortening <strong>of</strong><br />

the <strong>lipid</strong> tails, as can be seen by the decrease <strong>of</strong> the end-to-end distance. For these<br />

more flexible <strong>bilayers</strong>, the effect <strong>of</strong> molecules absorbed at the interface is opposite<br />

than in the case <strong>of</strong> the more stiff bilayer, i.e. the <strong>lipid</strong> chains become more disordered<br />

to compensate for the larger volume available due to the increase in surface area.<br />

When hydrophobic dumb-bells are added, the increase in surface area is much<br />

smaller for all the bilayer types considered, <strong>and</strong> the bilayer hydrophobic thickness<br />

is always larger than the thickness <strong>of</strong> the pure bilayer. Since the hydrophobic solutes<br />

locate in the bilayer center, they increase the bulk volume <strong>of</strong> the bilayer. The<br />

largest increase in thickness is observed for the stiff bilayer. The presence <strong>of</strong> the solutes<br />

in the hydrophobic core disrupts the ordering <strong>of</strong> the <strong>lipid</strong>s tail (as can be seen


90 Interaction <strong>of</strong> small molecules <strong>with</strong> <strong>bilayers</strong><br />

from figure 6.7). In the ht (L)<br />

6 t bilayer, this decrease in chain order leads to a decrease<br />

in interdigitation, <strong>and</strong> the two monolayers become more separated. This results in<br />

an increase <strong>of</strong> the bilayer thickness, even if the end-to-end distance decreases. In the<br />

case <strong>of</strong> the flexible <strong>and</strong> unsaturated bilayer, the increase in thickness is much smaller,<br />

<strong>and</strong> it can be accounted for by both the presence <strong>of</strong> the solute molecules in the bilayer<br />

interior <strong>and</strong> by a (small) increase in the end-to-end distance. Since the <strong>lipid</strong>s<br />

in these bilayer types are already much less ordered than in the stiff bilayer, the addition<br />

<strong>of</strong> the solutes does not have a significant effect on the order parameters (data<br />

not shown).<br />

It is important to observe that the effect <strong>of</strong> the neutral dumb-bells on the bilayer<br />

properties is the same than the effect <strong>of</strong> the amphiphilic dumb-bells, because they<br />

both absorb at the bilayer interface. However, since the neutral molecules largely<br />

diffuse in the water phase (<strong>and</strong> sometimes even in the bilayer inner core) they have<br />

an effective lower concentration in the bilayer compared to the amphiphilic ones.<br />

For this reason, they have a lesser effect on the bilayer structure compared to the<br />

amphiphilic molecules.<br />

6.3.3 Effect on the pressure distribution<br />

The effect <strong>of</strong> addition <strong>of</strong> solute molecules on the lateral pressure pr<strong>of</strong>ile across the<br />

bilayer is shown in figure 6.8, where we plot the pressure pr<strong>of</strong>iles in the <strong>bilayers</strong> <strong>with</strong><br />

added the solutes to compare them to the pressure pr<strong>of</strong>ile in the pure <strong>bilayers</strong>. As<br />

a general observation, the peaks (both positive <strong>and</strong> negative) <strong>of</strong> the lateral pressure<br />

decrease in magnitude by addition <strong>of</strong> solutes, irrespective <strong>of</strong> the bilayer type, <strong>and</strong> <strong>of</strong><br />

the nature <strong>of</strong> the added molecules. In this sense, the solutes can be seen as interfacial<br />

active molecules, that, like soap, have the effect <strong>of</strong> shifting to zero the local pressure<br />

(<strong>and</strong> the surface tension) <strong>of</strong> the interface where it locates.<br />

However, since a bilayer is not a simple interface, but has a complex structure, different<br />

molecules, at different position <strong>with</strong>in the bilayer, change the lateral pressure<br />

in different ways. To describe the characteristics <strong>of</strong> the pressure pr<strong>of</strong>ile, we make use<br />

<strong>of</strong> the interfaces defined in Chapter 4 (see figure 4.5 therein).<br />

It is important to remind that the integral <strong>of</strong> the lateral pressure across the bilayer<br />

is always zero, since the bilayer is in a tensionless state. Hence, a positive or negative<br />

change in the lateral pressure in one region <strong>of</strong> the bilayer, should be compensated by<br />

opposite changes in other regions <strong>of</strong> the bilayer.<br />

The largest shift in the pressure is at the headgroup/tails interface (first negative<br />

peak in figures 6.8), but the magnitude <strong>of</strong> this shift largely depends on the bilayer<br />

structure. For the stiff bilayer (figure 6.8(a)) both hydrophilic <strong>and</strong> hydrophobic solutes<br />

increase the local pressure by approximately the same amount, while for the<br />

more flexible <strong>bilayers</strong> (figure 6.8(b) <strong>and</strong> 6.8(c)) the amphiphilic molecules have a<br />

much larger effect than the hydrophobic ones. This shift <strong>of</strong> lateral pressure is com-


6.3 Results <strong>and</strong> Discussion 91<br />

π(z)<br />

0.1<br />

0.0<br />

−0.1<br />

−0.2<br />

−5 −3 −1<br />

z<br />

1 3 5<br />

(a) ht (L)<br />

6 t<br />

π(z)<br />

0.1<br />

0.0<br />

−0.1<br />

pure<br />

dbA<br />

dbH<br />

dbN<br />

π(z)<br />

0.1<br />

0.0<br />

−0.1<br />

−0.2<br />

−5 −3 −1<br />

z<br />

1 3 5<br />

(c) ht7<br />

−0.2<br />

−5 −3 −1<br />

z<br />

1 3 5<br />

(b) ht (L) t (K)<br />

4 t (L) t<br />

Figure 6.8: Pressure pr<strong>of</strong>iles in <strong>bilayers</strong> <strong>with</strong> added dumb-bells compared to the pure bilayer,<br />

for the three bilayer types, <strong>and</strong> the three dumb-bell types considered in this study.<br />

pure<br />

dbA<br />

dbH<br />

dbN<br />

pure<br />

dbA<br />

dbH<br />

dbN


92 Interaction <strong>of</strong> small molecules <strong>with</strong> <strong>bilayers</strong><br />

pensated by an opposite change <strong>of</strong> pressure at the water/headgroups interface (first<br />

maximum in figure 6.8).<br />

Also the shift in lateral pressure in the bilayer hydrophobic core depends on the<br />

bilayer topology <strong>and</strong> solute characteristics. In the ht (L) t (K)<br />

4 t(L) t <strong>and</strong> ht7 <strong>bilayers</strong>, the<br />

hydrophilic (<strong>and</strong> the neutral) dumb-bells have little effect in the bilayer center (z =<br />

0). In the ht (L)<br />

6 t bilayer, because <strong>of</strong> the decrease in interdigitation, <strong>and</strong> <strong>of</strong> packing<br />

density, induced by the interface-adsorbed solutes, the effect is a decrease <strong>of</strong> the local<br />

pressure in the bilayer center.<br />

To further characterize the changes in the local pressure induced by the solute<br />

molecules we define the following quantity<br />

∆π(z) = π db(z) − πo(z) (6.2)<br />

where π db(z) is the pressure for a bilayer <strong>with</strong> dumb-bells <strong>and</strong> πo(z) is the pressure<br />

for the pure bilayer.<br />

Since the most common organization <strong>of</strong> <strong>lipid</strong> <strong>bilayers</strong> is the non interdigitated<br />

state, we focus on the pressure shift ∆π(z) for the <strong>lipid</strong> type ht (L) t (K)<br />

4 t(L) t, as shown in<br />

figure 6.9. The shift in pressure for the fully flexible bilayer ht7 is very similar to the<br />

one shown. Because <strong>of</strong> symmetry, just half <strong>of</strong> the pressure pr<strong>of</strong>ile is represented, <strong>with</strong><br />

the bilayer center at z = 0. Also, z has been normalized by the hydrophobic thickness<br />

<strong>of</strong> the bilayer. From the figure it can be seen that the hydrophobic dumb-bells shift<br />

Figure 6.9: Difference in pressure pr<strong>of</strong>iles ∆π(z) (equation 6.2) in a ht (L) t (K)<br />

4 t (L) t bilayer <strong>with</strong><br />

added different dumb-bells respect to the pure bilayer. Only half <strong>of</strong> the pressure pr<strong>of</strong>ile is<br />

shown. The bilayer center is at z = 0, <strong>and</strong> z has been rescaled by the bilayer hydrophobic thickness,<br />

Dc. A schematic representation <strong>of</strong> the <strong>lipid</strong>s is also shown for a better characterization<br />

<strong>of</strong> the position <strong>of</strong> the peaks <strong>of</strong> ∆π(z). The first peak in the graphs (at z ≈ −0.6) corresponds<br />

to the water/headgroups interface, <strong>and</strong> the second peak (z ≈ −0.5) to the headgroups/tails<br />

interface.


6.3 Results <strong>and</strong> Discussion 93<br />

the local pressure less than the dumb-bells which adsorb at the headgroup interface.<br />

The difference between the amphiphilic <strong>and</strong> the neutral solutes is due to the smaller<br />

concentration <strong>of</strong> the latter at the interface, because they have partially diffused into<br />

the water phase. The molecules adsorbed at the interface decrease the pressure in<br />

the water/headgroup region <strong>and</strong> increase the pressure at the headgroup/tail interface.<br />

Also, a decrease in pressure going through the hydrophobic core, in the acyl<br />

chain region, is observed, while no change in pressure is present in the bilayer center<br />

(z=0). This local redistribution <strong>of</strong> pressure, <strong>and</strong> in particular in the headgroup<br />

interfacial region, could influence the gating mechanism <strong>of</strong> integral channel proteins<br />

[116], or affect the binding <strong>of</strong> peripheral membrane proteins [120].<br />

It is interesting to compare our results <strong>with</strong> other simulation studies reported<br />

in literature. For example, using atomistic MD simulations, Klein <strong>and</strong> co-workers<br />

investigated the interaction <strong>of</strong> anesthetic molecules (halotane) [117, 121] <strong>and</strong> their<br />

non-anesthetic (or non-immobilizer) analogues (hexafluoroethane) [122] <strong>with</strong> saturated<br />

[123] <strong>and</strong> poly-unsaturated <strong>lipid</strong> <strong>bilayers</strong> [119]. They have shown that the<br />

anesthetic molecules are preferentially absorbed at the head-group region, increasing<br />

the surface area per <strong>lipid</strong> <strong>and</strong> inducing structural modifications <strong>of</strong> the <strong>lipid</strong> bilayer,<br />

while the non-anesthetic molecules are mainly located in the bilayer core, <strong>and</strong><br />

the <strong>lipid</strong> bilayer exhibits almost no structural changes compared to the pure <strong>lipid</strong>.<br />

Our model hydrophobic <strong>and</strong> neutral molecules have the same solubility in oil since<br />

they both have the same value <strong>of</strong> repulsion parameter (a = 25) <strong>with</strong> the tails. Although<br />

the neutral molecules largely diffuse in the water phase, if they remain in the<br />

bilayer region they locate predominantly near the headgroups, while the hydrophobic<br />

molecules are found mainly in the bilayer hydrophobic core. This is consistent<br />

<strong>with</strong> the findings in [117,119,121,122] for the location <strong>of</strong> the anesthetic halotane <strong>and</strong><br />

the non-anesthetic hexafluoroethane in <strong>lipid</strong> <strong>bilayers</strong>.<br />

Our results indicate that the small molecules that locate preferentially in the bilayer<br />

headgroup interfacial region are the most effective in modifying the bilayer<br />

properties, <strong>and</strong> that the partitioning in the bilayer <strong>of</strong> solutes which have the same<br />

size <strong>and</strong> chemical nature is largely driven by lateral stresses <strong>and</strong> density differences<br />

<strong>with</strong>in the bilayer. Molecules adsorbed in a bilayer have as well a large effect on the<br />

distribution <strong>of</strong> the lateral pressures across the bilayer, this effect being very much dependent<br />

on bilayer composition, solute properties, <strong>and</strong> location <strong>of</strong> the solutes <strong>with</strong>in<br />

the bilayer.


VII<br />

<strong>Mesoscopic</strong> model for <strong>lipid</strong> <strong>bilayers</strong> <strong>with</strong><br />

<strong>embedded</strong> proteins


96 <strong>Mesoscopic</strong> model for <strong>lipid</strong> <strong>bilayers</strong> <strong>with</strong> <strong>embedded</strong> proteins<br />

7.1 Introduction<br />

The hydrophobic matching between the <strong>lipid</strong> bilayer hydrophobic thickness <strong>and</strong> the<br />

hydrophobic length <strong>of</strong> integral membrane proteins has been proposed as a generic<br />

physical principle on which the <strong>lipid</strong>-protein interaction in biomembranes is based<br />

[14, 20, 124–127]. The energy cost <strong>of</strong> exposing polar moieties, from either hydrocarbon<br />

chains or protein residues, is so high that the hydrophobic part <strong>of</strong> the <strong>lipid</strong><br />

bilayer should match the hydrophobic domain <strong>of</strong> membrane proteins. The results<br />

from a number <strong>of</strong> investigations have indeed pointed out the relevance <strong>of</strong> the hydrophobic<br />

matching in relation to the <strong>lipid</strong>-protein interactions, hence to membrane<br />

organization <strong>and</strong> biological function. It is now known that hydrophobic matching<br />

is used in cell membrane organization. For example, the membranes <strong>of</strong> the Golgi<br />

have different thicknesses; along their secretory pathway, proteins that pass through<br />

the Golgi undergo changes <strong>of</strong> their hydrophobic length to match the membrane hydrophobic<br />

thickness <strong>of</strong> the Golgi [128–131]. Hydrophobic matching seems also to<br />

play a role in sequestering proteins <strong>with</strong> long transmembrane regions [132] into sphingo<strong>lipid</strong>s-cholesterol<br />

biomembrane domains denoted as ‘rafts’ [133,134]. The biological<br />

importance <strong>of</strong> rafts <strong>and</strong> their involvement in Alzheimer’s <strong>and</strong> prion diseases is<br />

nowadays an intensively investigated subject [135].<br />

Biological membranes have at disposal a number <strong>of</strong> ways, which may be used individually<br />

or simultaneously, to compensate for hydrophobic mismatch [136]. These<br />

ways may imply changes <strong>of</strong> the membrane structure <strong>and</strong> dynamics on a microscopic,<br />

as well as on a macroscopic scale, <strong>and</strong> therefore can affect the membrane biological<br />

function [137–139]. To adjust to hydrophobic mismatch a membrane protein<br />

may cause a change <strong>of</strong> the <strong>lipid</strong> bilayer hydrophobic thickness in its vicinity. Experimental<br />

studies on reconstituted systems show that the range <strong>of</strong> the perturbation<br />

induced by proteins on the membrane thickness varies considerably from system to<br />

system [140–146]. A <strong>lipid</strong> sorting at the <strong>lipid</strong>-protein interface may also occur, where<br />

the protein prefers, on a statistical basis, to be associated <strong>with</strong> the <strong>lipid</strong> type that best<br />

matches its hydrophobic surface [147–150]. Another way in which a protein could<br />

adapt to a too thin bilayer matrix is to tilt [20, 151–155]. Besides the protein as a<br />

whole, also the individual helices <strong>of</strong> which a protein could be composed may experience<br />

a tilt; there is indeed some experimental evidence that the latter phenomenon<br />

may occur in channel proteins [19], <strong>and</strong> that a change <strong>of</strong> the tilt angle <strong>of</strong> the individual<br />

helices could be the cause <strong>of</strong> changed protein activity. Long <strong>and</strong> single-spanning<br />

membrane proteins might also bend to adapt to a too thin bilayer. Spectroscopic<br />

measurements on phospho<strong>lipid</strong> <strong>bilayers</strong> <strong>with</strong> <strong>embedded</strong> poly(leucine-alanine) αhelices<br />

suggest that the conformation <strong>of</strong> long peptides deviates from a straight helical<br />

end-to-end conformation [155, 156]. A protein may also undergo structural changes<br />

to adapt to a mismatched <strong>lipid</strong> bilayer. Spectroscopy measurements indicate that, indeed,<br />

long hydrophobic polyleucine peptides might distort in the C- <strong>and</strong> N-terminus


7.1 Introduction 97<br />

to reduce their hydrophobic length <strong>and</strong> thus match the thickness <strong>of</strong> the <strong>lipid</strong> bilayer<br />

in the gel-phase [157]. Lipid-mediated protein aggregation could also occur to reduce<br />

the stress caused by hydrophobic mismatch [145,150,158]. Domains [134] may<br />

thus form whose functional properties differ from those <strong>of</strong> the ‘bulk’, i.e. the unperturbed<br />

bilayer [159–161]. When the degree <strong>of</strong> mismatch is too large to be compensated<br />

for by the adaptations just described, the proteins might partition between a<br />

in-plane <strong>and</strong> a transmembrane orientation, or even avoid incorporation in the membrane<br />

[156, 162, 163]. The phenomena just mentioned refer to local microscopic<br />

changes related to mismatch adjustment. Perturbations <strong>of</strong> the membrane on the<br />

macroscopic scale may also occur; these can range from in-plane protein segregation<br />

<strong>and</strong> crystallization, <strong>and</strong> gel-fluid phase separation [14,147,149,164], to changes <strong>of</strong> the<br />

three-dimensional structure <strong>of</strong> the membrane. The formation <strong>of</strong> non-bilayer phases<br />

upon protein incorporation is an example <strong>of</strong> the latter type <strong>of</strong> phenomena [165,166].<br />

In an effort to elucidate the effects caused at the molecular level by the <strong>lipid</strong>protein<br />

hydrophobic mismatch, <strong>and</strong> even their implications for the formation <strong>of</strong> biologically<br />

relevant domain-like structures such as rafts, a number <strong>of</strong> theoretical studies<br />

have been done <strong>with</strong> the help <strong>of</strong> different types <strong>of</strong> theoretical <strong>models</strong> [7–11, 13,<br />

15–18, 167–171]. One <strong>of</strong> the quantities that has drawn considerable attention in the<br />

recent years is the extension <strong>of</strong> the domain size, which is determined by the the coherence<br />

length <strong>of</strong> the spatial fluctuations occurring in the system. Such fluctuations,<br />

which depend on the thermodynamic state <strong>of</strong> the system, can be induced, as well as<br />

sorted, by proteins. In the past, computer simulations have been made on a lattice<br />

model to compute the extent <strong>of</strong> the perturbation induced by a protein on the surrounding<br />

<strong>lipid</strong> bilayer [167]. The results from these simulations indicated that the<br />

extension <strong>of</strong> the perturbation depends on factors such as the degree <strong>of</strong> hydrophobic<br />

mismatch, the size <strong>of</strong> the protein (i.e. the curvature <strong>of</strong> the protein hydrophobic<br />

surface in contact <strong>with</strong> the <strong>lipid</strong> hydrocarbon chains), <strong>and</strong> on the temperature <strong>of</strong> the<br />

investigated system. Also, it was found that, away from the protein, the perturbation<br />

decays in a exponential manner, <strong>and</strong> can therefore by characterized by a decay<br />

length, ξP. This length is a coherence length which is a measure <strong>of</strong> the extension <strong>of</strong><br />

the range over which the <strong>lipid</strong>-mediated interaction between proteins may operate.<br />

The decay length is also a measure <strong>of</strong> the size <strong>of</strong> small-scale inhomogeneities (i.e.<br />

domains) experienced by proteins when <strong>embedded</strong> in the <strong>lipid</strong> bilayer. Results from<br />

model studies <strong>of</strong> a phenomenological interfacial model for protein-like objects in a<br />

bilayer-like system, suggest that, under well defined thermodynamic conditions, the<br />

protein-induced perturbation may propagate <strong>with</strong>out decay over a number <strong>of</strong> <strong>lipid</strong><br />

shells around the protein (the number <strong>of</strong> <strong>lipid</strong> shells being dependent, among others,<br />

on the size <strong>of</strong> the protein), may extend over long ranges, <strong>and</strong> might eventually<br />

establish a thermodynamic phase [14,15]. The phase <strong>of</strong> the multi-layered region that<br />

the protein prefers to be surrounded <strong>with</strong> is thus said to ’wet’ the protein [14, 15].<br />

The type <strong>of</strong> <strong>models</strong> briefly mentioned above are relatively crude (in the sense that


98 <strong>Mesoscopic</strong> model for <strong>lipid</strong> <strong>bilayers</strong> <strong>with</strong> <strong>embedded</strong> proteins<br />

they do not take into full account the three dimensional structure <strong>of</strong> the bilayer), <strong>and</strong>,<br />

for example, they can not be used to investigate the <strong>lipid</strong>-induced protein tilting.<br />

Simulations on more realistic <strong>models</strong>, such as all-atom <strong>models</strong> for <strong>lipid</strong> <strong>bilayers</strong> <strong>with</strong><br />

<strong>embedded</strong> proteins, have anyway confirmed that, at least <strong>with</strong>in a time <strong>of</strong> the order<br />

<strong>of</strong> the nanoseconds, a mismatched protein may induce a deformation <strong>of</strong> the <strong>lipid</strong> bilayer<br />

structure [10, 63, 170], <strong>and</strong> that the deformation is <strong>of</strong> the exponential type [11].<br />

The same type <strong>of</strong> studies have also shown that—although to reduce a possible hydrophobic<br />

mismatch synthetic peptides may prefer to deform the <strong>lipid</strong> bilayer, rather<br />

than undergo tilting [171]—tilting may also occur for membrane peptides [7,8]. Incidentally,<br />

the results from these studies indicated that the helical-peptides experience<br />

a slight bend in the middle <strong>of</strong> the helix.<br />

No matter the huge body <strong>of</strong> experimental <strong>and</strong> theoretical studies on <strong>lipid</strong> <strong>bilayers</strong><br />

<strong>with</strong> <strong>embedded</strong> proteins, issues such as the range <strong>of</strong> the protein-induced <strong>lipid</strong> bilayer<br />

perturbation, its dependence on protein size, <strong>and</strong> the occurrence <strong>of</strong> protein tilting (or<br />

even bending) to adjust for hydrophobic mismatch, are still a matter <strong>of</strong> debate. Here<br />

we want to focus on these issues by adopting the DPD simulation method to study<br />

the behavior <strong>of</strong> a mesoscopic model for <strong>lipid</strong> <strong>bilayers</strong> <strong>with</strong> <strong>embedded</strong> proteins. We<br />

have focused our attention on the perturbation caused by a membrane protein on<br />

the surrounding <strong>lipid</strong>s, its possible dependence on hydrophobic mismatch, protein<br />

size, <strong>and</strong> on temperature. We have investigated whether <strong>and</strong> to which extent—due to<br />

hydrophobic mismatch <strong>and</strong> via the cooperative nature <strong>of</strong> the system—a protein may<br />

prefer to tilt (<strong>with</strong> respect to the normal to the bilayer plane), rather than to induce a<br />

bilayer deformation <strong>with</strong>out (or even <strong>with</strong>) tilting.<br />

7.2 Computational details<br />

7.2.1 Lipid <strong>and</strong> protein <strong>models</strong><br />

Within the mesoscopic approach, each molecule <strong>of</strong> the system (or groups <strong>of</strong> molecules)<br />

is coarse-grained by a set <strong>of</strong> beads. In particular, to model the bilayer <strong>and</strong> the<br />

<strong>embedded</strong> proteins, we consider three types <strong>of</strong> beads: a water-like bead, labeled ’w’;<br />

a hydrophilic bead, labeled ’h’, which <strong>models</strong> a part <strong>of</strong> the headgroup <strong>of</strong> either the<br />

<strong>lipid</strong> or the protein; <strong>and</strong> a hydrophobic bead, labeled either ’tL’ or ’tP’, depending on<br />

whether it refers to a portion <strong>of</strong> the <strong>lipid</strong> hydrocarbon chain or a portion <strong>of</strong> the hydrophobic<br />

region <strong>of</strong> protein, respectively. The systems that we have simulated are<br />

made <strong>of</strong> model <strong>lipid</strong>s having three headgroup beads <strong>and</strong> two tails <strong>of</strong> five beads each;<br />

this corresponds to the case <strong>of</strong> acyl chains <strong>with</strong> fourteen carbon atoms, namely to a<br />

model for a dimyristoylphosphatidylcholine (DMPC) phospho<strong>lipid</strong>, as illustrated in<br />

figure 2.2 <strong>of</strong> Chapter 2, <strong>and</strong> in figure 7.1(a). Within the model formulation, a protein<br />

is considered as a rod-like object, <strong>with</strong> no appreciable internal flexibility, <strong>and</strong> characterized<br />

by a hydrophobic length dP. The model for the transmembrane protein is


7.2 Computational details 99<br />

dL o<br />

(a) (b)<br />

(c)<br />

eff<br />

dL (r) dP<br />

(d)<br />

φ tilt<br />

Figure 7.1: Schematic representation <strong>of</strong> a model-<strong>lipid</strong> (a), <strong>and</strong> a model protein (NP=43 <strong>and</strong><br />

˜dP=41˚A) (b). A typical configuration <strong>of</strong> the assembled bilayer <strong>with</strong> <strong>embedded</strong> a model-protein<br />

(as results from the simulations) is shown in the snapshot (c). The drawing in (d) shows to<br />

which part <strong>of</strong> the system the following quantities refer to: the pure <strong>lipid</strong> bilayer hydrophobic<br />

thickness, d o L, the perturbed <strong>lipid</strong> bilayer hydrophobic thickness, dL(r), the protein hydrophobic<br />

length, dP, the tilted-protein hydrophobic length, d eff<br />

P , <strong>and</strong> the tilt-angle, φ tilt .<br />

dP


100 <strong>Mesoscopic</strong> model for <strong>lipid</strong> <strong>bilayers</strong> <strong>with</strong> <strong>embedded</strong> proteins<br />

built by connecting ntP hydrophobic beads into a chain, to which ends n h headgrouplike<br />

beads are attached. NP <strong>of</strong> these amphiphatic chains are linked together into a<br />

bundle. In each model protein, all the NP chains are linked to the neighboring ones<br />

by springs, to form a relatively rigid body. We have considered three typical modelprotein<br />

sizes, two <strong>of</strong> them referring to a “skinny” peptide-like molecule, <strong>and</strong> the third<br />

type to a “fat” protein. These model proteins consist <strong>of</strong> NP=4, 7, or 43 chains linked<br />

together in a bundle. The bundle <strong>of</strong> NP=7 chains is formed by a central chain surrounded<br />

by a single layer <strong>of</strong> six other chains. The NP=43 bundle is made <strong>of</strong> three layers<br />

arranged concentrically around a central chain, <strong>and</strong> containing each six, twelve,<br />

<strong>and</strong> twenty four amphiphatic chains, respectively. The number <strong>of</strong> beads at each hydrophilic<br />

end <strong>of</strong> the bead-chains forming the protein is set equal to three. Each protein<br />

hydrophobic bead, tP, corresponds to a section <strong>of</strong> a α- or β-helical membrane<br />

protein. The distance spanned by a bead corresponds approximately to that spanned<br />

by a helix turn. About the chosen protein sizes, NP=4, 7, <strong>and</strong> 43, <strong>and</strong> their relation to<br />

the ones <strong>of</strong> actual proteins, the hydrophobic section <strong>of</strong> single-spanning membrane<br />

proteins like Glycophorin [172], <strong>and</strong> the M13 major coat protein from phage [173]<br />

or α-helical synthetic peptides [164] may be modeled by a skinny NP=4 type. β-helix<br />

proteins like gramicidin A [165], may be modeled by a NP=7 type. The fat protein may<br />

be a model for larger proteins consisting <strong>of</strong> transmembrane α-helical peptides that<br />

associate in bundles, or β-barrel proteins [174]. Specific examples could be bacteriorhodopsin<br />

[175], lactose permease [176], the photosynthetic reaction center [177],<br />

cytochrome c oxidase [178], or aquaglycerolporin [179]. Because we were interested<br />

in mismatch dependent effects, we have chosen protein hydrophobic sections composed<br />

<strong>of</strong> chains <strong>with</strong> different number <strong>of</strong> hydrophobic beads: ntP =2, 4, 6, 8, 10, <strong>and</strong><br />

12. To have an idea <strong>of</strong> what these numbers correspond to in terms <strong>of</strong> protein hydrophobic<br />

length, one can consider that the equilibrium distance between the beads<br />

is req=0.7Rc; assuming a value <strong>of</strong> Rc=6.4633 ˚A, the resulting values for the protein hydrophobic<br />

lengths will be, dP=14 ˜ ˚A (4 beads), 18 ˚A (5 beads), 23 ˚A (6 beads), 32 ˚A (8<br />

beads), 41 ˚A (10 beads), <strong>and</strong> 50 ˚A (12 beads). It is worth mentioning that these estimated<br />

protein hydrophobic lengths (which we denoted by dP, ˜ to distinguish from<br />

the lengths calculated from the simulations), are only meant to be indicative. Because<br />

<strong>of</strong> the s<strong>of</strong>t interactions involved in DPD, the value <strong>of</strong> the protein hydrophobic<br />

length that results from the simulations (<strong>and</strong> which, in the following, is denoted by<br />

dP) might be subjected to a small deviation <strong>of</strong> the order <strong>of</strong> 1 ˚A <strong>with</strong> respect to the<br />

values given above. Figure 7.1(b) shows a cartoon <strong>of</strong> a model protein <strong>of</strong> size NP=43,<br />

while figure 7.1(c) shows a snapshot <strong>of</strong> a typical configuration <strong>of</strong> the assembled bilayer<br />

<strong>with</strong> the <strong>embedded</strong> protein, as results from the simulations.<br />

Model parameters<br />

The non-bonded interactions between the beads are represented by the s<strong>of</strong>t-repulsive<br />

conservative force F C ij = aij(1 − rij/Rc)^rij. The values <strong>of</strong> the parameters referring


7.2 Computational details 101<br />

to the <strong>lipid</strong>-<strong>lipid</strong> <strong>and</strong> <strong>lipid</strong>-water interactions have been chosen equal to the ones<br />

used for modeling the pure <strong>lipid</strong> bilayer system. The numerical values <strong>of</strong> the interaction<br />

parameters between different bead types are shown in Table 7.1. Regarding<br />

Table 7.1: Repulsion parameters aij for different bead-types.<br />

aij w h tL tP<br />

w 25 15 80 120<br />

h 15 35 80 80<br />

tL 80 80 25 25<br />

tP 120 80 25 25<br />

the protein-protein interactions, the parameters related to the repulsive interactions<br />

between the beads forming the hydrophilic part <strong>of</strong> the protein, as well as the ones<br />

between the hydrophobic beads, have been chosen <strong>with</strong> the same values as the interaction<br />

parameters between hydrophilic <strong>and</strong> hydrophobic beads <strong>of</strong> the <strong>lipid</strong>, respectively,<br />

i.e. ahh = 35 <strong>and</strong> atPtP = atLtL = 25. About the parameter related to the<br />

interaction between the protein hydrophobic beads <strong>and</strong> the water, we have chosen a<br />

value to ensure that the hydrophobic section <strong>of</strong> the protein was sufficiently shielded<br />

from the water environment. This was done by calculating the interaction energy<br />

between the protein outer hydrophobic chains <strong>and</strong> the water, for different values <strong>of</strong><br />

the repulsion parameter awtP . The trend in the energy shows the onset <strong>of</strong> a plateau<br />

for awtP ≥ 120, which was hence the chosen value for the interaction parameter. Two<br />

consecutive beads in the <strong>lipid</strong> or in the protein are connected by harmonic springs<br />

(equation 2.13) <strong>with</strong> spring constant Kr=100 <strong>and</strong> equilibrium distance ro=0.7. To control<br />

the <strong>lipid</strong> flexibility, a harmonic bond-bending potential between two consecutive<br />

bonds in the <strong>lipid</strong> tails was added <strong>with</strong> bending constant Kθ=6 <strong>and</strong> equilibrium angle<br />

θo=180 o . An additional bond-bending potential was applied between the vectors<br />

connecting the <strong>lipid</strong> tails to the headgroup, <strong>with</strong> Kθ=3 <strong>and</strong> θo=90 o . Compared to the<br />

<strong>lipid</strong> hydrocarbon chains, the hydrophobic part <strong>of</strong> membrane proteins can be considered<br />

fairly rigid; therefore the value <strong>of</strong> the bending constant in the protein chains<br />

was set to Kθ=100, i.e. about an order <strong>of</strong> magnitude larger than the one used for the<br />

<strong>lipid</strong> chains. At this point it is important to mention that, although these parameters<br />

are chosen such that the system behavior is in reasonable agreement <strong>with</strong> the experimental<br />

system, a one to one link to all specific properties is not always possible to<br />

make.<br />

7.2.2 Method <strong>of</strong> calculation <strong>of</strong> statistical quantities<br />

We have studied the physical properties <strong>of</strong> the system both in the absence <strong>and</strong> in the<br />

presence <strong>of</strong> the proteins. The pure <strong>lipid</strong> bilayer hydrophobic thickness, do L , was estimated<br />

by calculating, for each sampled configuration, the difference between the<br />

average position along the bilayer normal (z direction, considering the bilayer paral-


102 <strong>Mesoscopic</strong> model for <strong>lipid</strong> <strong>bilayers</strong> <strong>with</strong> <strong>embedded</strong> proteins<br />

lel to the xy-plane) <strong>of</strong> the tail-beads attached to the headgroup <strong>of</strong> the <strong>lipid</strong>s in one<br />

(top) monolayer, <strong>and</strong> <strong>of</strong> the <strong>lipid</strong>s in the opposite (bottom) monolayer:<br />

d o L = 〈zt (top) − zt (bottom) 〉 (7.1)<br />

where zt is the z position <strong>of</strong> the beads attached to the <strong>lipid</strong> headgroup. The over-line<br />

represents an average over the two chains for each <strong>lipid</strong> <strong>and</strong> over the total number<br />

<strong>of</strong> <strong>lipid</strong>s in each monolayer, <strong>and</strong> the brackets indicate that the difference is further<br />

averaged over the number <strong>of</strong> configurations chosen for the statistical sampling.<br />

To study the effect <strong>of</strong> a protein on the surrounding bilayer structure, we have calculated<br />

the <strong>lipid</strong> bilayer hydrophobic thickness, dL(r), as function <strong>of</strong> the radial distance<br />

r from the protein hydrophobic surface, namely from the interface <strong>with</strong> the<br />

<strong>lipid</strong> hydrocarbon chains, as schematically illustrated in figure 7.1(d). The method<br />

<strong>of</strong> calculation <strong>of</strong> dL(r) is similar to the method <strong>of</strong> calculation <strong>of</strong> do L as it is explained<br />

below <strong>and</strong> illustrated in figure 7.2.<br />

dL(r)<br />

3<br />

3<br />

2<br />

∆r<br />

2 1<br />

r<br />

1<br />

(a)<br />

top monolayer<br />

x<br />

z<br />

y<br />

bottom monolayer<br />

Figure 7.2: Schematic drawing to illustrate the method <strong>of</strong> calculation <strong>of</strong> dL(r), which is described<br />

in details in the text. The protein is represented by a gray cylinder. The figure shows<br />

the case when the protein is parallel to the bilayer normal (a), <strong>and</strong> the case when the protein<br />

is tilted <strong>with</strong> respect to the bilayer normal (b).<br />

For each sampled configuration, we have first calculated the circularly averaged<br />

value <strong>of</strong> the position along the bilayer normal <strong>of</strong> the tail-beads attached to the headgroup<br />

<strong>of</strong> the <strong>lipid</strong>s <strong>with</strong>in each circular sector k at distance r = k∆r from the protein<br />

surface, in the bottom <strong>and</strong> top monolayer separately, i.e. zt (bottom) (r), <strong>and</strong> zt (top) (r).<br />

The instantaneous value <strong>of</strong> the bilayer hydrophobic thickness at distance r from the<br />

protein surface is then given by the difference <strong>of</strong> these two values. This difference<br />

has been further averaged over the sampled configurations to give dL(r):<br />

dL(r) = 〈zt (top) (r) − zt (bottom) (r)〉, r = k∆r, ∀ k = 1, 2, 3, . . . (7.2)<br />

The bin size, ∆r, was chosen to be <strong>of</strong> the order <strong>of</strong> the diameter <strong>of</strong> the <strong>lipid</strong>-chain<br />

projected area on the bilayer plane. It is important to note that, if the protein is tilted<br />

d L<br />

(r)<br />

3<br />

2<br />

r<br />

1<br />

3<br />

2<br />

1<br />

(b)


7.2 Computational details 103<br />

(figure 7.2b), the circular sectors at distance r from the protein surface in the top<br />

<strong>and</strong> bottom monolayer are shifted respect to each other in the bilayer plane, <strong>and</strong> the<br />

value <strong>of</strong> dL(r) calculated <strong>with</strong> the described method is an approximated value <strong>of</strong> the<br />

actual thickness in the vicinity <strong>of</strong> the tilted peptide. However, this value converges to<br />

the correct value in the bulk at distances sufficiently far from the protein. Moreover,<br />

because <strong>of</strong> the tilt, the conformation <strong>of</strong> the <strong>lipid</strong>s around the protein might not be<br />

symmetric. We want to point out that these possible effects due to the asymmetry <strong>of</strong><br />

the protein orientation in the bilayer have been averaged out.<br />

The behavior <strong>of</strong> dL(r) allowed us to access the extension <strong>of</strong> the protein-mediated<br />

perturbation on the bilayer. Based on previous theoretical finding [167], we first assumed<br />

that the perturbation induced by the protein on the surrounding <strong>lipid</strong>s is <strong>of</strong><br />

exponential type. We have then verified this assumption later by analyzing the deviation<br />

<strong>of</strong> the functional form <strong>of</strong> the calculated dL(r) from the assumed one. If the<br />

behavior <strong>of</strong> dL(r) is exponential, the protein-induced perturbation can be expressed<br />

in terms <strong>of</strong> a typical coherence length, the decay length ξP:<br />

dL(r) = d o L + (dP − d o L)e −r/ξP . (7.3)<br />

where do L is the mean hydrophobic thickness <strong>of</strong> the unperturbed pure <strong>lipid</strong> bilayer,<br />

<strong>and</strong> dP is the protein hydrophobic length. The above equation expresses the fact<br />

that away from the protein surface, <strong>and</strong> at distances at least <strong>of</strong> the order <strong>of</strong> ξP, the<br />

perturbed dL(r) decays to the bulk value do L , namely the value corresponding to that<br />

<strong>of</strong> the pure <strong>lipid</strong> system at the considered temperature, if no finite-size effects are<br />

present. In principle, by knowing dL(r), dP, <strong>and</strong> do L <strong>and</strong> by using equation 7.3 one<br />

can estimate ξP. In our case, we have determined the value <strong>of</strong> ξP by best-fitting the<br />

values dL(r) resulting from the simulations <strong>with</strong> equation 7.3, where ξP <strong>and</strong> do L are<br />

the fitting parameters. About the resulting value <strong>of</strong> the parameter do L obtained by the<br />

best-fitting, we have verified that this is equal, <strong>with</strong>in the statistical accuracy, to the<br />

value <strong>of</strong> the <strong>lipid</strong> bilayer hydrophobic thickness in the bulk, <strong>and</strong> directly calculated<br />

from the simulations.<br />

Since the protein can be subjected to tilt, the input parameter we used for the fit<br />

is not the actual hydrophobic length <strong>of</strong> the model-protein, dP (or even the a priori<br />

estimate <strong>of</strong> it, dP), ˜ but instead an effective length, deff P . This effective length is defined<br />

as the projection onto the normal <strong>of</strong> the bilayer plane <strong>of</strong> the protein hydrophobic<br />

length directly obtained from the simulations: d eff<br />

P = dP cos(φ tilt ), where φ tilt is the<br />

tilt angle (see figure 7.1d). To calculate the degree <strong>of</strong> tilting <strong>of</strong> a protein <strong>with</strong> respect<br />

to the bilayer normal we have considered, for each chain <strong>of</strong> the protein, the vector<br />

that connects the position <strong>of</strong> the two hydrophobic beads bound to the protein hydrophilic<br />

beads (i.e. close to the <strong>lipid</strong>-water interface), one located in one monolayer<br />

<strong>of</strong> the bilayer, <strong>and</strong> the other in the opposite monolayer. The tilt angle, φ tilt , is defined<br />

as the average value, over all the chains <strong>of</strong> a protein, <strong>of</strong> the angle between this vector<br />

<strong>and</strong> the bilayer normal.


104 <strong>Mesoscopic</strong> model for <strong>lipid</strong> <strong>bilayers</strong> <strong>with</strong> <strong>embedded</strong> proteins<br />

In some cases, to facilitate the interpretation <strong>of</strong> the data obtained from the simulations,<br />

it was necessary to determine the degree <strong>of</strong> order/disorder <strong>of</strong> the <strong>lipid</strong> chains<br />

in the vicinity <strong>of</strong> the protein, <strong>and</strong> eventually to compare it <strong>with</strong> the one <strong>of</strong> the pure<br />

<strong>lipid</strong> bilayer, i.e. in the bulk, away from the protein-induced perturbation. Therefore,<br />

we have calculated the value <strong>of</strong> the <strong>lipid</strong> chain order parameter, S(r), which is defined<br />

as follows:<br />

S = 1<br />

2 (3 cos2 φS − 1), (7.4)<br />

<strong>with</strong><br />

cos φS = rij · ^n<br />

rij<br />

= zij<br />

, (7.5)<br />

rij<br />

where φS is the angle between the orientation <strong>of</strong> the vector rij = rj − ri (rij = |rij|)<br />

along two consecutive <strong>lipid</strong> chain beads, i, j, <strong>and</strong> the bilayer normal unit vector, ^n.<br />

S(r) has been independently calculated for each <strong>of</strong> the two monolayers <strong>of</strong> the bilayer,<br />

as well as averaged over all the bonds <strong>of</strong> the <strong>lipid</strong> chains at distance r from the surface<br />

<strong>of</strong> the protein.<br />

7.3 Results <strong>and</strong> discussion<br />

In this section we present the results from the simulations <strong>of</strong> the <strong>lipid</strong> bilayer <strong>models</strong>ystem<br />

<strong>with</strong> <strong>embedded</strong> proteins. We focus on the low protein-concentration regime,<br />

where the correlation between different proteins can be neglected, <strong>and</strong> hence consider<br />

<strong>bilayers</strong> <strong>with</strong> <strong>embedded</strong> a single protein.<br />

To investigate the dependence on mismatch <strong>and</strong> protein size on the <strong>lipid</strong> bilayer<br />

around an <strong>embedded</strong> protein we first studied the behavior <strong>of</strong> the system at a constant<br />

temperature, well above the pure <strong>lipid</strong> bilayer transition temperature. Since<br />

the bilayer hydrophobic thickness varies <strong>with</strong> temperature, a way to change the hydrophobic<br />

mismatch is by changing temperature. Hence, we studied the behavior <strong>of</strong><br />

ξP <strong>and</strong> φ tilt in the temperature range above the melting temperature <strong>of</strong> the pure system,<br />

i.e. in the fluid phase, <strong>and</strong> for a number <strong>of</strong> <strong>lipid</strong>-protein model-systems. Finally,<br />

for few hydrophobic mismatch conditions <strong>and</strong> for few selected temperatures below<br />

the melting temperature, we investigated how the phase behavior <strong>of</strong> the pure <strong>lipid</strong><br />

system in the gel-phase affects the localization <strong>and</strong> orientation <strong>of</strong> the proteins in this<br />

phase.<br />

The results discussed next refer to <strong>lipid</strong> <strong>bilayers</strong> <strong>of</strong> 900 <strong>lipid</strong>s <strong>and</strong> 22500 water<br />

beads (i.e. 25 water beads per <strong>lipid</strong>), corresponding to a fully hydrated bilayer. We<br />

have made calculations for smaller systems, <strong>and</strong> we have found that a size corresponding<br />

to 900 <strong>lipid</strong> molecules was sufficient to avoid finite-size effects, at least for<br />

temperatures close or above the main-transition temperature. Before collecting the<br />

data used to estimate statistical quantities, we have first equilibrated each bilayer<br />

system for 20,000 DPD-MC cycles. In each cycle it was chosen (<strong>with</strong> a probability <strong>of</strong>


7.3 Results <strong>and</strong> discussion 105<br />

70%) whether to perform a number <strong>of</strong> DPD steps or an attempt to change the box<br />

aspect ratio according to the imposed value <strong>of</strong> the surface tension, γ = 0. After equilibration,<br />

data were collected over 50,000 DPD-MC cycles, at γ = 0. The statistical<br />

averages were then made over configurations that were separated from one another<br />

by 50 DPD steps. On average, 10,000 independent configurations were considered<br />

for the statistical averages.<br />

7.3.1 Protein-induced bilayer perturbation<br />

The results shown next refer to the reduced temperature T ∗ =0.7, well above the melting<br />

temperature <strong>of</strong> the system, i.e. in the fluid phase. The pure <strong>lipid</strong> bilayer hydrophobic<br />

thickness calculated at this reduced temperature is d o L =(23.6±0.2) ˚A. Figure<br />

7.3 shows the calculated bilayer hydrophobic thickness pr<strong>of</strong>ile, dL(r), as a function<br />

<strong>of</strong> the distance r from the protein surface. The data refer to different values <strong>of</strong><br />

dP, resulting in different values <strong>of</strong> hydrophobic mismatch ∆d, ranging from ∆d=-8 ˚A<br />

to 28˚A, <strong>and</strong> to the three protein sizes, which correspond to NP =4, 7, <strong>and</strong> 43. Because<br />

the probability <strong>of</strong> finding a <strong>lipid</strong> molecule in the <strong>lipid</strong>-shell closest to the protein is<br />

much lower than in the other <strong>lipid</strong>-shells, the data collected at a distance r = ∆r from<br />

the protein surface have not been considered for the statistics.<br />

One can clearly see from the curves in figure 7.3 that the protein induces a perturbation<br />

<strong>of</strong> the <strong>lipid</strong> bilayer in its vicinity. The perturbation decays in a manner that<br />

depends on the hydrophobic mismatch <strong>and</strong> on protein size. If the protein hydrophobic<br />

length is smaller than the unperturbed bilayer hydrophobic thickness (dP < d o L ),<br />

i.e. negative mismatch, ∆d < 0 (open symbols), the <strong>lipid</strong>s around the protein shrink<br />

to match the protein hydrophobic surface. By choosing the peptide hydrophobic<br />

length to approximately match the value <strong>of</strong> the hydrophobic thickness <strong>of</strong> the unperturbed<br />

<strong>lipid</strong> bilayer, i.e. ∆d ≈0 (crosses), one can clearly see from figure 7.3 that the<br />

perturbation induced by the protein on the surrounding <strong>lipid</strong>s becomes negligible.<br />

Instead, when the chosen protein is such that dP > do L , i.e. ∆d > 0 (full symbols), to<br />

match the protein hydrophobic surface the <strong>lipid</strong>s in the vicinity <strong>of</strong> the protein stretch<br />

<strong>and</strong> become more gel-like than the bulk <strong>lipid</strong>s, far away from the protein.<br />

Figure 7.4 shows the thickness pr<strong>of</strong>iles for two values <strong>of</strong> mismatch ∆d=-10 ˚A, <strong>and</strong><br />

∆d=17 ˚A, <strong>and</strong> for the three considered protein sizes. The symbols indicate the data<br />

obtained directly from the simulations, while the continuum line is obtained by best-<br />

fitting <strong>with</strong> the function in equation 7.3, where do L <strong>and</strong> ξP are the fitting parameters<br />

(<strong>and</strong> where deff P is the input parameter). For convenience, we have drawn a horizon-<br />

tal dashed line to indicate the value <strong>of</strong> the pure <strong>lipid</strong> bilayer hydrophobic thickness,<br />

do L , calculated at the same reduced temperature considered for the simulations <strong>of</strong><br />

the mixed systems. To help the interpretation <strong>of</strong> the data, the values <strong>of</strong> the protein<br />

hydrophobic length, dP, directly calculated from the simulations, <strong>and</strong> <strong>of</strong> deff P , the projected<br />

protein hydrophobic length onto the normal to the bilayer plane, are also plot-


106 <strong>Mesoscopic</strong> model for <strong>lipid</strong> <strong>bilayers</strong> <strong>with</strong> <strong>embedded</strong> proteins<br />

d L (r) [Å]<br />

45.0<br />

35.0<br />

25.0<br />

15.0<br />

0.0 20.0 40.0 60.0 80.0<br />

(a) NP=4<br />

d L (r) [Å]<br />

r [Å]<br />

45.0<br />

35.0<br />

25.0<br />

∆d=−10<br />

∆d=−6<br />

∆d=−1<br />

∆d= 8<br />

∆d=17<br />

∆d=26<br />

d L (r) [Å]<br />

45.0<br />

35.0<br />

25.0<br />

15.0<br />

0.0 20.0 40.0 60.0 80.0<br />

r [Å]<br />

(c) NP=43<br />

15.0<br />

0.0 20.0 40.0 60.0 80.0<br />

∆d=−10<br />

∆d=−6<br />

∆d=−1<br />

∆d= 8<br />

∆d=17<br />

∆d=26<br />

(b) NP=7<br />

r [Å]<br />

∆d=−10<br />

∆d=−6<br />

∆d=−1<br />

∆d= 8<br />

∆d=17<br />

∆d=26<br />

Figure 7.3: Lipid bilayer hydrophobic thickness pr<strong>of</strong>iles, dL(r) as function <strong>of</strong> the distance r<br />

from the protein surface, for different hydrophobic mismatch ∆d = ˜dP − d o L, <strong>and</strong> for the three<br />

protein sizes NP=4, 7 <strong>and</strong> 43. The data refer to the reduced temperature T ∗ =0.7, which is well<br />

above the main-transition temperature <strong>of</strong> the pure system (T ∗ m=0.425). The calculated pure<br />

<strong>lipid</strong> bilayer hydrophobic thickness at T ∗ =0.7 is d o L =(23.6±0.2) ˚A. The symbols for ∆d refer to<br />

the following estimated values <strong>of</strong> the protein hydrophobic length: the open circle to ˜dP=14 ˚A<br />

(negative ∆d), the open triangle to ˜dP=18 ˚A (negative ∆d), the plus to ˜dP=23 ˚A (∆d ≈ 0), the full<br />

triangle to ˜dP=32 ˚A (positive ∆d), the full circle to ˜dP=41 ˚A (positive ∆d), <strong>and</strong> the full diamonds<br />

to ˜dP=50 ˚A (positive ∆d).


7.3 Results <strong>and</strong> discussion 107<br />

ted (see gray <strong>and</strong> white rectangular areas). The best fit is obtained <strong>with</strong> the values <strong>of</strong><br />

the fitting parameters given in Table 7.2, for the three chosen protein sizes, <strong>and</strong> for<br />

varying values <strong>of</strong> mismatch, i.e. protein hydrophobic thickness.<br />

Table 7.2: Values <strong>of</strong> the decay length, ξP, the pure <strong>lipid</strong> bilayer hydrophobic thickness d o L (both<br />

derived from fitting the thickness pr<strong>of</strong>iles dL(r) by using Eq. 7.3), the protein hydrophobic<br />

length, dP, <strong>and</strong> the effective protein hydrophobic length, d eff<br />

P (both calculated from the simulations)<br />

given for different values <strong>of</strong> hydrophobic mismatch, ∆d, <strong>and</strong> for the three protein sizes<br />

corresponding to NP=4, 7 <strong>and</strong> 43. The data refer to simulations made at the reduced temperature<br />

T ∗ =0.7, well above the main-transition temperature <strong>of</strong> the pure <strong>lipid</strong> bilayer system. The<br />

pure <strong>lipid</strong> bilayer hydrophobic thickness calculated at this temperature is d o L =(23.6±0.2) ˚A. In<br />

the case <strong>of</strong> approximately zero mismatch (∆d=-1 ˚A), the values <strong>of</strong> ξP <strong>and</strong> d o L are not calculated.<br />

∆d [˚A] ξP [˚A] d o L [˚A] dP [˚A] d eff<br />

P [˚A]<br />

Protein fitted fitted computed computed<br />

NP=4 -10 9.3 ± 0.3 24.0 ± 0.1 15 ± 1 15 ± 1<br />

-6 11.9 ± 0.3 23.9 ± 0.1 20 ± 1 19 ± 1<br />

-1 * * 24 ± 1 24 ± 1<br />

8 9.6 ± 0.7 23.4 ± 0.1 34 ± 1 32 ± 1<br />

17 9.7 ± 0.7 23.4 ± 0.2 43 ± 1 35 ± 3<br />

26 12.3 ± 0.6 23.2 ± 0.1 53 ± 1 36 ± 3<br />

NP= 7 -10 10.1 ± 0.4 24.2 ± 0.1 15 ± 1 14 ± 1<br />

-6 12.4 ± 0.7 24.0 ± 0.1 19 ± 1 19 ± 1<br />

-1 * * 24 ± 1 23 ± 1<br />

8 9.4 ± 0.8 23.5 ± 0.1 33 ± 1 32 ± 1<br />

17 11.8 ± 0.7 23.2 ± 0.2 42 ± 1 39 ± 2<br />

28 12.4 ± 0.8 23.3 ± 0.2 51 ± 1 39 ± 3<br />

NP= 43 -10 12.8 ± 0.8 24.2 ± 0.1 14 ± 1 14 ± 1<br />

-6 17 ± 2 24.3 ± 0.2 19 ± 1 19 ± 1<br />

-1 * * 24 ± 1 24 ± 1<br />

8 10 ± 1 23.2 ± 0.3 33 ± 1 33 ± 1<br />

17 12 ± 1 22.5 ± 0.6 43 ± 1 43 ± 1<br />

26 12 ± 1 22.1 ± 0.8 52 ± 1 51 ± 2<br />

At negative mismatch, no difference is observed between dP <strong>and</strong> deff P , as can be<br />

seen by looking at figure 7.4(a,c,e). This means that the orientation <strong>of</strong> the protein is<br />

perpendicular to the bilayer plane, hence dP=deff P . However, for a positive mismatch<br />

too large to be compensated for by fully stretching the <strong>lipid</strong>s in the vicinity <strong>of</strong> the<br />

protein, another effect can be observed; the peptide tilts in order to decrease its effective<br />

hydrophobic length. The effect is much more pronounced in the case <strong>of</strong> the<br />

skinny protein (NP=4) than for the larger protein, as can be seen by comparing figures<br />

is much shorter than the ac-<br />

7.4b,d <strong>with</strong> figure 7.4(f); where in the former cases deff P<br />

tual protein hydrophobic length. The values <strong>of</strong> deff P <strong>and</strong> dP shown in Table 7.2 confirm


108 <strong>Mesoscopic</strong> model for <strong>lipid</strong> <strong>bilayers</strong> <strong>with</strong> <strong>embedded</strong> proteins<br />

30.0<br />

d L (r) [Å]<br />

25.0<br />

20.0<br />

15.0<br />

10.0<br />

5.0<br />

0.0<br />

30.0<br />

d L (r) [Å]<br />

25.0<br />

20.0<br />

15.0<br />

10.0<br />

5.0<br />

0.0<br />

30.0<br />

d L (r) [Å]<br />

25.0<br />

20.0<br />

15.0<br />

10.0<br />

5.0<br />

0.0<br />

eff<br />

dP dP eff<br />

dP dP eff<br />

dP dP ∆d0<br />

0 10 20 30 40 50 60 70 80 90<br />

r [Å]<br />

(f) NP = 43,<br />

Figure 7.4: Calculated values <strong>of</strong> dL(r) (open circles) <strong>and</strong> fitting ones using expression 7.3 (solid<br />

line) as a function <strong>of</strong> the distance r from the protein surface. The data refer to the three protein<br />

sizes NP=4, 7, <strong>and</strong> 43; <strong>and</strong> to two values <strong>of</strong> the protein hydrophobic length dP=14˚A (∆d=-10˚A)<br />

<strong>and</strong> 41˚A (∆d=17˚A). Also shown is the ‘level’ value <strong>of</strong> the unperturbed <strong>lipid</strong> bilayer thickness<br />

d o L =(23.6±0.2) ˚A (dashed line), the calculated protein hydrophobic length dP (gray area), <strong>and</strong><br />

the effective protein hydrophobic length d eff<br />

P (white area), which is defined as the projection <strong>of</strong><br />

dP onto the normal to the bilayer plane. The data refer to simulations at the reduced temperature<br />

T ∗ =0.7.<br />

o<br />

dL dL dL fit<br />

o<br />

dL dL dL fit<br />

o<br />

dL dL dL fit


7.3 Results <strong>and</strong> discussion 109<br />

that this is also the case for the other values <strong>of</strong> positive mismatch. The derived values<br />

<strong>of</strong> ξP (using equation 7.3) as function <strong>of</strong> protein size <strong>and</strong> hydrophobic mismatch,<br />

which are shown in Table 7.2, indicate that there is a mismatch dependence <strong>of</strong> the<br />

perturbation caused by the protein on the surrounding <strong>lipid</strong>s. For a given protein<br />

size, NP, if the mismatch is negative, the correlation length increases <strong>with</strong> decreasing<br />

mismatch (absolute value), while for positive mismatch the opposite happens,<br />

<strong>and</strong> the correlation length increases <strong>with</strong> increasing mismatch. Also, in the case <strong>of</strong><br />

negative mismatch the decay length increases by increasing the protein size. Instead<br />

there is no detectable ξP dependence on NP in the case <strong>of</strong> ∆d > 0, at least at the considered<br />

temperature, i.e. well above the melting temperature <strong>of</strong> the pure system. The<br />

scenario is somehow different when the temperature decreases <strong>and</strong> approaches the<br />

transition temperature, as it is discussed in the next section.<br />

Figure 7.4f shows that, for the large protein (NP=43), the <strong>lipid</strong> thickness pr<strong>of</strong>ile<br />

dL(r) differs from an exponential one. The effect is even more pronounced at lower<br />

temperatures (data not shown), at least in the case <strong>of</strong> negative mismatch (since lower<br />

temperature means larger negative mismatch). This non-exponential behavior, <strong>and</strong><br />

the possible reason for it, will be discussed later.<br />

Table 7.2 also gives the values <strong>of</strong> the pure <strong>lipid</strong> bilayer hydrophobic thickness d o L<br />

obtained from the best fit <strong>of</strong> dL(r) using equation 7.3. For all the considered cases, the<br />

(fitted) compare well <strong>with</strong> the value <strong>of</strong> the pure <strong>lipid</strong> bilayer hydrophobic<br />

values <strong>of</strong> do L<br />

thickness, do L =23.6 ˚A, obtained directly from the simulation at the considered temperature.<br />

7.3.2 Lipid-induced protein tilting<br />

The protein tilt angle <strong>with</strong> respect to the bilayer normal as function <strong>of</strong> ∆d <strong>and</strong> protein<br />

size NP is shown in figure 7.5. The snapshots on the right show typical configurations<br />

<strong>of</strong> the system, for a fixed value <strong>of</strong> the protein hydrophobic length, for the<br />

three protein sizes, NP=4, 7, <strong>and</strong> 43, <strong>and</strong> for the largest (positive) value <strong>of</strong> mismatch,<br />

∆d=26 ˚A, at the considered temperature, T ∗ =0.7. For ∆d < 0 the tilt angle is very<br />

small, <strong>and</strong> is <strong>with</strong>in the statistical tilt-fluctuations to which the protein is subject<br />

in the bilayer; as the protein hydrophobic length increases (<strong>and</strong> the mismatch becomes<br />

positive), the protein undergoes a significant tilting. Also, for equal values <strong>of</strong><br />

hydrophobic mismatch, the “thinner” protein (NP=4) is much more tilted than the<br />

“fatter” one (NP=43). These results, combined <strong>with</strong> the one discussed above, suggest<br />

that in the case <strong>of</strong> proteins <strong>with</strong> small surface area, the main mechanism to compensate<br />

for a large hydrophobic mismatch is the tilt, while in the case <strong>of</strong> proteins<br />

<strong>with</strong> a large surface area, that cannot accommodate a too large tilt, the mismatch is<br />

mainly compensated for by an increase <strong>of</strong> the bilayer thickness around the protein,<br />

as is clearly illustrated by the snapshot in figure 7.5 (NP=43).


110 <strong>Mesoscopic</strong> model for <strong>lipid</strong> <strong>bilayers</strong> <strong>with</strong> <strong>embedded</strong> proteins<br />

φ tilt<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

N p =4<br />

N p =7<br />

N p =43<br />

0<br />

−15 −10 −5 0 5 10 15 20 25 30<br />

∆d [Å]<br />

Figure 7.5: Protein tilt-angle φ tilt as function <strong>of</strong> mismatch, ∆d. The data refer to a reduced<br />

temperature <strong>of</strong> T ∗ =0.7 <strong>and</strong> to the three considered protein sizes NP=4, 7 <strong>and</strong> 43. The dashed<br />

lines are only meant to be a guideline for the eye. Typical configurations <strong>of</strong> the systems resulting<br />

from the simulations are shown on the right. Starting from the top, the snapshots refer to<br />

proteins sizes NP=4, 7 <strong>and</strong> 43. In the three cases the protein hydrophobic length is ˜dP=50 ˚A,<br />

hence the hydrophobic mismatch is ∆d=26 ˚A.<br />

Theoretical studies based on MD simulations on all-atom <strong>models</strong> have pointed<br />

out the possibility that α-helical hydrophobic peptides may tilt when subjected to<br />

positive mismatch conditions [7, 8, 171], the degree <strong>of</strong> tilting depending on the specific<br />

system. The occurrence <strong>of</strong> protein tilting has also been confirmed experimentally.<br />

In fact, the results from very recent experimental investigations by solid state<br />

NMR spectroscopy [155] show that α-helical model peptides — <strong>of</strong> fixed hydrophobic<br />

length <strong>and</strong> <strong>with</strong> a hydrophobic leucine-alanine core, <strong>and</strong> tryptophan flanked ends<br />

— experience tilt when <strong>embedded</strong> in phospho<strong>lipid</strong> <strong>bilayers</strong> <strong>of</strong> varying hydrophobic<br />

thickness (such that dP ≥ do L , i.e. ∆d > 0). It was found that the tilt angle increases by<br />

systematically increasing hydrophobic mismatch; however, the tilt dependence on<br />

hydrophobic mismatch was not as pronounced as one would have expected, given<br />

the degree <strong>of</strong> mismatch. This result brought the authors to conclude that the tilt <strong>of</strong><br />

these peptides is energetically unfavorable, <strong>and</strong> to suggest that the (anchoring) effects<br />

by specific residues such as tryptophans are more dominant than mismatch<br />

effect. A large tilt is instead experienced by the M13 coat protein peptide when <strong>embedded</strong><br />

in phospho<strong>lipid</strong> bilayer <strong>of</strong> varying hydrophobic thickness [154]. For values<br />

<strong>of</strong> mismatch <strong>of</strong> the same order <strong>of</strong> the one experienced by the synthetic peptides just


7.3 Results <strong>and</strong> discussion 111<br />

mentioned [155], the degree <strong>of</strong> tilting experienced by M13 peptide is much higher.<br />

Our simulation data indicate a dependence <strong>of</strong> the protein-tilt angle on mismatch in<br />

agreement <strong>with</strong> the experimental data just discussed. Incidentally, the results from<br />

our simulations suggest that, when a skinny peptide (Np =4) is subjected to a large<br />

positive mismatch (dP > do L ), it might bend—besides to experience a tilt—as can<br />

be seen by looking at the snapshot shown on the top-right <strong>of</strong> figure 7.5. Also, as<br />

soon as the positive mismatch decreases, the bending disappears, although the peptide<br />

still tends to remain tilted. Results from MD simulations on a all-atom model<br />

<strong>of</strong> a poly(32)alanine helical peptide <strong>embedded</strong> in a dimyristoylphosphatidylcholine<br />

(DMPC) bilayer show that this type <strong>of</strong> helix, not only tilts by 30o as a whole <strong>with</strong><br />

respect to the bilayer normal, but it also experiences a bend at its middle [7]; MD<br />

simulations have also shown a similar tendency for a poly(16)leucine helical peptide<br />

<strong>embedded</strong> in a DMPC bilayer [8]. From the experimental point <strong>of</strong> view, it is now<br />

possible to detect peptide/protein bending by NMR spectroscopy [155,180]. Indeed,<br />

the data from Str<strong>and</strong>berg et al. [155] on the behavior <strong>of</strong> a synthetic leucine-alanine<br />

α-helical peptide in <strong>lipid</strong> <strong>bilayers</strong> <strong>of</strong> varying thickness, do indicate that, for large positive<br />

mismatch, the peptides might experience bending (besides tilting), in agreement<br />

<strong>with</strong> our observations. However, we must point out that the occurrence, or extent, <strong>of</strong><br />

bending <strong>of</strong> the small peptide (NP=4) might very well be dependent on the value <strong>of</strong><br />

the bending constant, Kθ, chosen to model the stiffness <strong>of</strong> the protein chains.<br />

7.3.3 Thermotropic behavior<br />

Fluid phase<br />

We now discuss the response <strong>of</strong> the <strong>lipid</strong>-protein system when the temperature decreases<br />

<strong>and</strong> approaches the main-transition temperature, T ∗ m. The dependence <strong>of</strong> ξP<br />

on the reduced temperature, T ∗ , where T ∗ > T ∗ m, is shown in figure 7.6, for two values<br />

<strong>of</strong> protein hydrophobic length ˜dP=14 ˚A <strong>and</strong> ˜dP=41 ˚A. These values were chosen<br />

to fulfill the condition that either ∆d < 0 or ∆d > 0, respectively, even if by changing<br />

temperature the <strong>lipid</strong> bilayer hydrophobic thickness, <strong>and</strong> consequently ∆d, may<br />

change. The data refer to two protein sizes, NP=7 <strong>and</strong> 43. The behavior <strong>of</strong> ξP shown<br />

in figure 7.6 indicates that the closer the temperature is to the main-transition temperature,<br />

the longer is the perturbation caused by the protein on the surrounding<br />

<strong>lipid</strong>s. This is probably due to the enhanced density fluctuations that occur in the<br />

pure system close to the transitions temperature. Also, for negative mismatch, there<br />

is a pronounced dependence <strong>of</strong> ξP on the protein size, the larger the protein is, the<br />

longer the correlation length becomes. The dependence on protein size can be qualitatively<br />

explained by the fact that the larger the protein, the smaller its curvature, <strong>and</strong><br />

therefore the influence <strong>of</strong> a given portion <strong>of</strong> the protein hydrophobic surface extends<br />

to more <strong>lipid</strong>s. Although for positive mismatch the dependence on protein size is not<br />

that pronounced, the values <strong>of</strong> ξP in the case <strong>of</strong> the large protein are systematically


112 <strong>Mesoscopic</strong> model for <strong>lipid</strong> <strong>bilayers</strong> <strong>with</strong> <strong>embedded</strong> proteins<br />

ξ P [Å]<br />

25<br />

20<br />

15<br />

10<br />

∆d0<br />

N P =7<br />

N P =43<br />

5<br />

0.4 0.5 0.6 0.7 0.8 0.9 1.0<br />

T*<br />

Figure 7.6: Decay length ξP dependence on reduced temperature T ∗ (T ∗ > T ∗ m), for two values<br />

<strong>of</strong> protein hydrophobic length ˜dP=14 ˚A (a) <strong>and</strong> ˜dP=41 ˚A (b). These values were chosen to fulfill<br />

the condition that either ∆d < 0 (˜dP=14 ˚A) or ∆d > 0 (˜dP=41 ˚A), even if by changing temperature<br />

the <strong>lipid</strong> bilayer hydrophobic thickness may change, <strong>and</strong> consequently ∆d. The data refer<br />

to two protein sizes, NP=7 <strong>and</strong> 43. The dashed lines are only a guideline for the eye.<br />

higher than the ones related to the small protein.<br />

These findings are consistent <strong>with</strong> the results obtained from Monte Carlo simulations<br />

on lattice <strong>models</strong> [167], which predicted that the temperature dependence<br />

<strong>of</strong> the decay length has a dramatic peak at the transition temperature. They are<br />

also consistent <strong>with</strong> results from MD simulations on all-atom <strong>models</strong>. In fact, data<br />

from recent MD simulations on <strong>bilayers</strong> <strong>of</strong> fluid POPE <strong>and</strong> POPC <strong>with</strong> <strong>embedded</strong> the<br />

membrane channel aquaglycerolporin [11] show that when the hydrophobic length<br />

<strong>of</strong> the protein is shorter than the pure <strong>lipid</strong> bilayer hydrophobic thickness, the <strong>lipid</strong>s<br />

close to the <strong>lipid</strong>-protein interface compress to favor hydrophobic matching, thus<br />

inducing a curvature in the bilayer. Also, the mismatch-induced perturbation is <strong>of</strong><br />

exponential type, <strong>and</strong> can be characterized by a decay length around ξP =10 ˚A. Simulations<br />

on POPC <strong>bilayers</strong> <strong>with</strong> <strong>embedded</strong> a much smaller protein than aquaglycerolporin,<br />

the membrane channel gramicidin A, suggest that this channel induces perturbation<br />

having a coherence length that is smaller than that in the case <strong>of</strong> aquaglycerolporin<br />

[181].<br />

At the present stage <strong>of</strong> experimental development, estimates <strong>of</strong> the range <strong>of</strong> the<br />

perturbation that proteins such as Bacteriorhodopsin [143, 146], lactose permease<br />

[148], <strong>and</strong> the synthetic alpha-helical peptides [182,183], induce on the surrounding<br />

<strong>lipid</strong>s, suggest that the perturbation might be mismatch <strong>and</strong> protein-size dependent,<br />

the larger the protein the more long range the perturbation, consistent <strong>with</strong> the theo-<br />

(b)


7.3 Results <strong>and</strong> discussion 113<br />

retical predictions discussed here. If the coherence length associated to the proteininduced<br />

perturbation is dependent on protein size, one would expect that bilayer<br />

activities affected by changes <strong>of</strong> the coherence length, might be thus be affected by<br />

protein sizes. This is indeed the case for the phenomenon <strong>of</strong> flip-flop <strong>of</strong> phospho<strong>lipid</strong>s<br />

in <strong>bilayers</strong>. In fact, experimental data on flip-flop suggest that the larger the<br />

protein size (hence the smaller its curvature at the interface <strong>with</strong> the <strong>lipid</strong> chains) the<br />

more reduced is the ability <strong>of</strong> the protein to cause flip-flop [184].<br />

It is worth mentioning that the larger the protein size is, the more the behavior<br />

<strong>of</strong> dL(r) obtained from our DPD simulations differs from the one <strong>of</strong> the exponential<br />

function used for the best fitting. This could already be seen by looking at figure 7.4f<br />

(at a temperature well above the main-transition temperature) for NP=43 <strong>and</strong> in the<br />

case <strong>of</strong> negative mismatch. Figure 7.7 illustrates more in details the non-exponential<br />

dependence <strong>of</strong> the <strong>lipid</strong> bilayer thickness pr<strong>of</strong>ile on r. The figure shows the calculated<br />

values <strong>of</strong> dL(r) (open circles) <strong>and</strong> the fitting ones using expression 7.3 (solid line). The<br />

30.0<br />

d L (r) [Å]<br />

25.0<br />

20.0<br />

15.0<br />

10.0<br />

5.0<br />

0.0<br />

eff<br />

dP dP ∆d0<br />

0 10 20 30 40 50 60 70 80 90<br />

r [Å]<br />

(b)<br />

o<br />

dL dL dL fit<br />

Figure 7.7: The calculated values <strong>of</strong> dL(r) (open circles) <strong>and</strong> the fitting ones using expression<br />

7.3 (solid line) as a function <strong>of</strong> the distance r from the protein surface. The data refer to a protein<br />

size NP=43, <strong>and</strong> to the following cases: (a) ∆d=-12 ˚A (˜dP=14 ˚A) <strong>and</strong> T ∗ =0.5, <strong>and</strong> (b) ∆d=19<br />

˚A (˜dP=41 ˚A) <strong>and</strong> T ∗ =1.0, where in both cases the temperature is above the melting temperature<br />

<strong>of</strong> the pure system. The dashed line indicates the value <strong>of</strong> the pure <strong>lipid</strong> bilayer hydrophobic<br />

thickness d o L at the considered temperature. Also shown are the calculated protein hydropho-<br />

bic length dP (gray area), <strong>and</strong> the effective protein hydrophobic length d eff<br />

P (white area), which<br />

is defined as the projection <strong>of</strong> dP onto the normal to the bilayer plane.<br />

data refer to a protein size NP=43, <strong>and</strong> to the following two cases: (a) ∆d=-12 ˚A (˜dP=14<br />

˚A) <strong>and</strong> T∗ =0.5, <strong>and</strong> (b) ∆d=19 ˚A (˜dP=41 ˚A) <strong>and</strong> T∗ =1.0. In the case <strong>of</strong> positive <strong>and</strong> large<br />

mismatch (dP > do L ) (but low enough to avoid protein tilting), figure 7.7(b) indicates<br />

that the <strong>lipid</strong>s in the layers closest to the protein surface are characterized by gellike<br />

chain in order to minimize the hydrophobic mismatch; surprisingly, next to this


114 <strong>Mesoscopic</strong> model for <strong>lipid</strong> <strong>bilayers</strong> <strong>with</strong> <strong>embedded</strong> proteins<br />

<strong>lipid</strong>-ordered region, there appears a few-layer thick region (which for convenience<br />

we denote ‘undershooting’ region) where the bilayer has an hydrophobic thickness<br />

that is less that the one in the bulk. This undershooting phenomenon is probably due<br />

to the fact that, on the one h<strong>and</strong>, the system has to satisfy the boundary-constraint<br />

imposed by the value <strong>of</strong> the hydrophobic bilayer thickness <strong>of</strong> the bulk; <strong>and</strong>, on the<br />

other h<strong>and</strong>, the system has to adjust to the perturbation caused by the protein in<br />

such a way to hold the bilayer density constant, also close to the protein. We suggest<br />

that, if the protein is large enough that tilting is unfavorable, <strong>and</strong> if the mismatch is<br />

large enough that even the ordered (gel-like) <strong>lipid</strong>s closest to the protein are not able<br />

to match its hydrophobic surface, a void is formed in the center <strong>of</strong> the bilayer. To fill<br />

this void <strong>and</strong> maintain the constraint <strong>of</strong> uniform density in the bilayer core, the <strong>lipid</strong><br />

chains in the ‘undershooting’ region might tilt <strong>and</strong> bend toward the protein. This hypothesis<br />

is indeed consistent <strong>with</strong> the fact that, although the end-to-end distance <strong>of</strong><br />

the <strong>lipid</strong>s in the undershooting region is approximately equal to the one <strong>of</strong> the <strong>lipid</strong>s<br />

in the bulk, the projected length on the bilayer normal is shorter than the one <strong>of</strong> the<br />

<strong>lipid</strong>s in the bulk. Therefore, in the undershooting region, the bilayer hydrophobic<br />

thickness is smaller than in the bulk.<br />

In the case <strong>of</strong> dP < do L (see figure 7.7a), the bilayer in a neighboring region (which<br />

in this case we call ‘overshooting’ region) to the one closest to the protein interface<br />

has an hydrophobic thickness that is higher than the one <strong>of</strong> the <strong>lipid</strong> bilayer in the<br />

bulk. This might be explained by the fact that the <strong>lipid</strong> chains nearest to the protein<br />

might tilt (<strong>and</strong> possibly bend) to satisfy the matching constraint. At the same time the<br />

<strong>lipid</strong>s in the ‘overshooting’ region stretch their chains (i.e. become more gel-like) to<br />

satisfy the constant-density constraint. This is consistent <strong>with</strong> the fact that, in the region<br />

closest to the protein (the so-called ‘annulus’), the values <strong>of</strong> the order parameter<br />

<strong>of</strong> the <strong>lipid</strong> chains, <strong>and</strong> <strong>of</strong> both the <strong>lipid</strong> end-to-end distance <strong>and</strong> projected length,<br />

are smaller than the values in the bulk. This shows that the <strong>lipid</strong>s closest to the protein<br />

are more disordered <strong>and</strong> might bend to match the protein hydrophobic length<br />

(which is shorter than the bilayer thickness at the considered temperature). On the<br />

other h<strong>and</strong>, the projected length on the bilayer normal <strong>of</strong> the <strong>lipid</strong>s in the overshooting<br />

region is slightly longer than the one <strong>of</strong> the <strong>lipid</strong>s in the bulk; also the order parameter,<br />

S, in the overshooting region is higher than in the bulk. This indicates that<br />

the <strong>lipid</strong>s in this region are more stretched <strong>and</strong> ordered (i.e. gel-like) than the <strong>lipid</strong>s<br />

in the bulk. Finally, we want to point out that the overshooting/undershooting effect<br />

seems not to be due to finite-size effects.<br />

There could be a number <strong>of</strong> interesting biological implications if a curved structure<br />

resulting from the overshooting/undershooting effect occurs around proteins<br />

<strong>embedded</strong> in biological membranes. Its presence could affect the permeability properties<br />

<strong>of</strong> the membrane in the vicinity <strong>of</strong> each protein; it could be a basis for <strong>lipid</strong>sorting<br />

in the vicinity <strong>of</strong> the protein <strong>and</strong> when more than one <strong>lipid</strong> species is present<br />

in the system; it could also be a way to regulate protein-protein contacts, hence pro-


7.3 Results <strong>and</strong> discussion 115<br />

tein lateral distribution; finally, it could create a fertile (or adverse) ground for the<br />

attachment <strong>of</strong> fusion peptides, which are known to enter the bilayer in a oblique<br />

manner [185,186], <strong>and</strong> thus could be favored by the presence <strong>of</strong> tilted <strong>lipid</strong>s. It could<br />

also cause a change in the lateral pressure pr<strong>of</strong>ile around each protein, which in turn<br />

could induce conformational changes in the proteins. Furthermore, if there exists<br />

an overshooting/undershooting effect, one should be careful in the interpretation<br />

<strong>of</strong> results obtained from spectroscopic measurements <strong>of</strong> the <strong>lipid</strong> order parameter<br />

S; if the protein concentration is high enough that there is a sufficient number <strong>of</strong><br />

overshooting/undershooting <strong>lipid</strong>s around the isolated proteins, calculation <strong>of</strong> the<br />

<strong>lipid</strong> bilayer hydrophobic thickness from measurements <strong>of</strong> S could give underestimation/overestimation<br />

<strong>of</strong> the value <strong>of</strong> the hydrophobic thickness.<br />

Gel phase<br />

Finally, we discuss the behavior <strong>of</strong> the <strong>lipid</strong>-protein system below the melting temperature,<br />

i.e. in the gel phase. The results from the simulations are illustrated in two<br />

figures. Figure 7.8 shows the snapshots <strong>of</strong> two typical configurations obtained from<br />

the simulation <strong>of</strong> <strong>bilayers</strong> <strong>with</strong> <strong>embedded</strong> a NP=43 protein. The simulations are done<br />

(a) ∆d > 0 (b) ∆d < 0<br />

Figure 7.8: Snapshots <strong>of</strong> two typical configurations <strong>of</strong> <strong>lipid</strong>-protein <strong>bilayers</strong> at a reduced tem-<br />

perature T ∗ =0.25, corresponding to the Lβ ′ or gel-phase. The calculated value <strong>of</strong> the pure <strong>lipid</strong><br />

bilayer thickness is d o L =(36.6±0.1) ˚A. In (a) is shown the case <strong>of</strong> a protein <strong>of</strong> size NP=43 <strong>and</strong><br />

˜dP=50 ˚A, thus subjected to a positive mismatch ∆d=13 ˚A, while in (b) is shown the case <strong>of</strong> a<br />

NP=43 size protein, but <strong>with</strong> ˜dP=14 ˚A, thus subjected to negative mismatch ∆d=-23 ˚A.<br />

at a reduced temperature T∗ =0.25, corresponding to the Lβ ′ or gel-phase. The pure<br />

<strong>lipid</strong> bilayer in the Lβ ′ is characterized by a tilted orientation <strong>of</strong> the <strong>lipid</strong>s, as previously<br />

discussed. The calculated value <strong>of</strong> the pure <strong>lipid</strong> bilayer thickness at this reduced<br />

temperature is do L =(36.6±0.1) ˚A. In figure 7.8(a) is shown the case <strong>of</strong> a protein<br />

<strong>with</strong> ˜dP=50 ˚A, thus subjected to a positive mismatch ∆d=13 ˚A. The response <strong>of</strong> the<br />

protein to the positive mismatch is to tilt slightly to decrease its effective length, <strong>and</strong><br />

thus orient parallel to the tilted <strong>lipid</strong>s in the gel-state.


116 <strong>Mesoscopic</strong> model for <strong>lipid</strong> <strong>bilayers</strong> <strong>with</strong> <strong>embedded</strong> proteins<br />

A somehow different scenario arises when a protein is subjected to a negative<br />

mismatch, as illustrated in the snapshot in figure 7.8(b), which shows a typical configuration<br />

obtained from the simulations on a bilayer <strong>with</strong> <strong>embedded</strong> a protein <strong>with</strong><br />

˜dP=14 ˚A, thus subjected to negative mismatch ∆d=-23 ˚A. In this case the protein orients<br />

antiparallel to the orientation <strong>of</strong> the <strong>lipid</strong>s; also, the <strong>lipid</strong>s in the vicinity <strong>of</strong> the<br />

protein interdigitate to decrease the bilayer hydrophobic thickness <strong>and</strong> thus fulfill<br />

the matching condition. To better illustrate this phenomenon, we have depicted the<br />

end-bead <strong>of</strong> <strong>of</strong> the <strong>lipid</strong> chains in darker color. From the snapshot in figure 7.8(b) it<br />

can be seen that, in the vicinity <strong>of</strong> the protein, the end-tail <strong>of</strong> the <strong>lipid</strong>s in one monolayer<br />

are close to the headgroups <strong>of</strong> the <strong>lipid</strong>s in the opposite monolayer.<br />

As described in section 5.3 <strong>of</strong> Chapter 5, by increasing the temperature above<br />

T∗ =0.35, the pure <strong>lipid</strong> bilayer undergoes a transition from the Lβ ′ phase to a ‘striated’<br />

phase that resembles the Pβ ′ phase <strong>of</strong> phospho<strong>lipid</strong> <strong>bilayers</strong>. Interestingly,<br />

upon incorporation <strong>of</strong> a protein, <strong>and</strong> depending on the mismatch condition, the protein<br />

segregates into the striated region whose hydrophobic thickness better matches<br />

the protein hydrophobic length. This is clearly shown in the snapshots in figure<br />

7.9, which depict two typical configurations <strong>of</strong> the systems at a reduced temperature<br />

T∗ =0.4, <strong>and</strong> in the case <strong>of</strong> proteins <strong>with</strong> ˜dP=41 ˚A <strong>and</strong> 18 ˚A, respectively. The calculated<br />

value <strong>of</strong> the pure <strong>lipid</strong> bilayer hydrophobic thickness is do L =(30.0±0.3) ˚A. Both cases<br />

refer to a protein <strong>of</strong> size NP=7. For positive values <strong>of</strong> hydrophobic mismatch (see fig-<br />

(a) ∆d > 0 (b) ∆d < 0<br />

Figure 7.9: Snapshots <strong>of</strong> two typical configurations <strong>of</strong> <strong>lipid</strong>-protein <strong>bilayers</strong> at a reduced tem-<br />

perature T ∗ =0.4, in the ‘striated’ gel phase, which resembles the Pβ ′ phase in phospho<strong>lipid</strong><br />

<strong>bilayers</strong>. The calculated value <strong>of</strong> the pure <strong>lipid</strong> bilayer hydrophobic thickness is d o L =(30.0±0.3)<br />

˚A. In (a) is shown the case <strong>of</strong> a protein <strong>of</strong> size NP=7 <strong>and</strong> ˜dP=41 ˚A, thus subjected to a positive<br />

mismatch ∆d=11 ˚A, while in (b) is shown the case <strong>of</strong> a NP=7 size protein, but <strong>with</strong> ˜dP=18 ˚A,<br />

thus subjected to negative mismatch ∆d=-12 ˚A.<br />

ure 7.9(a)), the protein prefers to segregate in the striated region formed by <strong>lipid</strong>s in<br />

the gel-like state. Instead, in the case <strong>of</strong> negative mismatch, the protein prefers the<br />

region where the chains are fluid-like (see figure 7.9(b)). The interplay between the<br />

underlying structure <strong>of</strong> the striated phase <strong>and</strong> the mismatch-induced perturbation<br />

could provide a mean to tune the lateral organization <strong>of</strong> membrane proteins, <strong>and</strong><br />

thus control their segregation in the two-dimensional ordered structure. To underst<strong>and</strong><br />

how ordered structures might form is important because scattering methods


7.4 Conclusion 117<br />

make use <strong>of</strong> ordered structure as a matrix to determine the three-dimensional structure<br />

<strong>of</strong> proteins [187, 188].<br />

7.4 Conclusion<br />

We have presented a mesoscopic model for <strong>lipid</strong> <strong>bilayers</strong> <strong>with</strong> <strong>embedded</strong> proteins,<br />

which we have studied <strong>with</strong> the Dissipative Particle Dynamics simulation method.<br />

One <strong>of</strong> our aims was to point out the advantages <strong>of</strong> the DPD-simulation-CG-model<br />

approach by addressing some simple issues related to the collective nature <strong>of</strong> a threedimensional<br />

membrane system, a <strong>lipid</strong> bilayer containing just one <strong>lipid</strong> species <strong>and</strong><br />

an <strong>embedded</strong> protein. More specifically, we have investigated the effect due to mismatch<br />

<strong>and</strong> protein size on the perturbation induced by the protein on the surrounding<br />

<strong>lipid</strong> bilayer. The perturbation around the protein was quantified in terms <strong>of</strong> the<br />

bilayer hydrophobic thickness pr<strong>of</strong>ile. We found that the pr<strong>of</strong>ile may have an exponential<br />

form, decays to the value <strong>of</strong> the thickness <strong>of</strong> the unperturbed system (i.e.<br />

<strong>with</strong>out protein), <strong>and</strong> can be characterized by the coherence length, ξP, <strong>of</strong> the spatial<br />

fluctuation around the protein. We found that, under well defined thermodynamic<br />

conditions, the value <strong>of</strong> ξP may depend on mismatch <strong>and</strong> protein size, the larger the<br />

mismatch/size the larger ξP. Also, we found that to adapt to a too thin bilayer the<br />

protein may tilt (or even bend) in a manner which is mismatch <strong>and</strong> protein-size dependent.<br />

We have found that the model predictions are in qualitative agreement <strong>with</strong><br />

previous theoretical <strong>and</strong> experimental findings. We want to stress that the phenomena<br />

that we have investigated <strong>with</strong> the DPD simulation method involve molecular<br />

rearrangements in the membrane plane via, among others, diffusion <strong>of</strong> molecules<br />

whose time scale might be outside the range <strong>of</strong> investigation <strong>of</strong> more ‘traditional’<br />

simulation techniques, such as MD. The results discussed above refer to a model for<br />

DMPC <strong>bilayers</strong>. The trend shown by these results can also be applied to <strong>lipid</strong> <strong>bilayers</strong><br />

<strong>with</strong> other types <strong>of</strong> phospho<strong>lipid</strong>s, i.e. <strong>with</strong> longer or shorter hydrocarbon chains<br />

then the ones <strong>of</strong> DMPC. We would like to conclude by saying that the predictions that<br />

arise from numerical simulation studies <strong>of</strong> model systems, such as the one presented<br />

here, may be used as a complementary tool to experimental studies to reveal information<br />

not otherwise accessible; also, results from numerical studies can provide a<br />

framework for the interpretation <strong>of</strong> experimental data, as well as serve as a source <strong>of</strong><br />

inspiration for future experiments.


References<br />

[1] Tanford, C. Science 1978, 200, 1012–1018.<br />

[2] Tanford, C. Hydrophobic Effect: Formation <strong>of</strong> Micelles <strong>and</strong> Biological Membranes; Wiley,<br />

New York, 1980.<br />

[3] Sackmann, E. In Lipowsky, R., Sackmann, E., Eds., Structure <strong>and</strong> dynamics <strong>of</strong> membranes,<br />

pages 1–65. Elsevier, 1995.<br />

[4] König, S.; Sackmann, E. Current Opinion in Colloid <strong>and</strong> Interface Sci. 1996, 1, 78–82.<br />

[5] Tieleman, D.; Marrink, S.; Berendsen, H. Biochim. Biophys. Acta 1997, 1331, 235–270.<br />

[6] Marrink, S.; Edholm, O.; Mark, A. J. Am. Chem. Soc. 2001, 123, 8638–8639.<br />

[7] Shen, L.; Bassolino, D.; Stouch, T. Biophys. J. 1997, 73, 3–20.<br />

[8] Belohorcová, K.; Davis, J.; Woolf, T.; Roux, B. Biophys. J. 1997, 73, 3039–3055.<br />

[9] Belohorcová, K.; Qian, J.; .; Davis, J. Biophys. J. 2000, 79, 3201–3016.<br />

[10] Jensen, M.; Tajkhorshid, E.; Schulten, K. Structure 2001, 9, 1083–1093.<br />

[11] Jensen, M.; Mouritsen, O. Biochim. Biophys. Acta: Biomembranes 2004, in press.<br />

[12] Stouch, T.; Bassolino, D. In Mertz, J., B, R., Eds., Biological Membranes, pages 255–277,<br />

Boston, 1996. Birkhauser.<br />

[13] Mouritsen, O.; Sperotto, M.; Risbo, J.; Zhang, Z.; Zuckermann, M. Adv. Comp. Biol. 1996,<br />

2, 15–64.<br />

[14] Gil, T.; Ipsen, J.; Mouritsen, O.; Sabra, M.; Sperotto, M.; Zuckermann, M. Biochim. Biophys.<br />

Acta 1998, 1376, 245–266.<br />

[15] Gil, T.; Ipsen, J. Phys. Rev. E 1997, 55, 1713–1721.<br />

[16] S., M. Curr. Opin. Colloid Interface Sci. 2000, 5, 244–249.<br />

[17] Fattal, D.; Ben-Shaul, A. Biophys. J. 1993, 65, 1795–1809.<br />

[18] Nielsen, C.; Goulian, M.; Andersen, O. Biophys. J. 1998, 74, 1966–1983.<br />

[19] Lee, A. Biochim. Biophys. Acta 2003, 1612, 1–40.<br />

[20] Killian, J. Biochim. Biophys. Acta 1998, 1376, 401–416.<br />

[21] Goetz, R.; Lipowsky, R. J. Chem. Phys. 1998, 108, 7397–7409.<br />

[22] Shelley, J.; Shelley, M.; Reeder, R.; B<strong>and</strong>yopadhyay, S.; Klein, M. J. Phys. Chem. B 2001,<br />

105, 4464–4470.<br />

[23] Shelley, J.; Shelley, M.; Reeder, R.; B<strong>and</strong>yopadhyay, S.; Moore, P.; Klein, M. J. Phys. Chem.<br />

B 2001, 105, 9785–9792.<br />

[24] Groot, R.; Rabone, K. Biophys. J. 2001, 81, 725–736.<br />

[25] Marrink, S.; de Vries, A.; Mark, A. J. Phys. Chem. B 2004, 108, 750–760.<br />

[26] Hoogerbrugge, P.; Koelman, J. Europhys. Lett. 1992, 19, 155–160.<br />

[27] Español, P.; Warren, P. Europhys. Lett. 1995, 30, 191–196.


120 <strong>Mesoscopic</strong> model for <strong>lipid</strong> <strong>bilayers</strong> <strong>with</strong> <strong>embedded</strong> proteins<br />

[28] Groot, R.; Warren, P. J. Chem. Phys. 1997, 107, 4423–4435.<br />

[29] Cantor, R. Biochemistry 1997, 36, 2339–2344.<br />

[30] Cantor, R. J. Phys. Chem. B 1997, 101, 1723–1725.<br />

[31] Coveney, P.; Novik, K. Phys. Rev. E 1996, 54, 5134–5141.<br />

[32] Coveney, P.; Novik, K. Phys. Rev. E 1997, 55, 4831.<br />

[33] Novik, K.; Coveney, P. Int. J. Mod. Phys. C 1997, 8, 909–918.<br />

[34] Groot, R.; Madden, T. J. Chem. Phys. 1998, 108, 8713–8724.<br />

[35] Groot, R.; Madden, T.; Tildesley, D. J. Chem. Phys. 1999, 110, 9739–9749.<br />

[36] Schlijper, A.; Hoogerbrugge, P.; Manke, C. J. Rehol. 1995, 39, 567–579.<br />

[37] Schlijper, P.; Manke, C.; Madden, W.; Kong, Y. Int. J. Mod. Phys. C 1997, 8, 919–929.<br />

[38] Kong, Y.; Manke, C.; Madden, W.; Schlijper, A. J. Chem. Phys. 1997, 107, 592–602.<br />

[39] Spenley, N. Europhys. Lett. 2000, 49, 534–540.<br />

[40] Koelman, J.; Hoogerbrugge, P. Europhys. Lett. 1993, 21, 363–368.<br />

[41] Boek, E.; Coveney, P.; Lekkerkekker, H. J.Phys.: Condens. Matter 1996, 8, 9509–9512.<br />

[42] Boek, E.; Coveney, P.; Lekkerkekker, H.; van Schoot, P. Phys. Rev. E 1997, 55, 3124–3133.<br />

[43] Wijmans, C.; Smit, B.; Groot, R. J. Chem. Phys. 2001, 114, 7644–7654.<br />

[44] Wijmans, C.; Smit, B. Macromolecules 2002, 35, 7138–7148.<br />

[45] Jury, S.; Bladon, P.; Cates, M.; Krishna, S.; Hagen, M.; Ruddock, N.; Warren, P. Phys. Chem.<br />

Chem. Phys. 1999, 1, 2051–2056.<br />

[46] Groot, R. Langmuir 2000, 16, 7493–7502.<br />

[47] Warren, P. Curr. Opin. Colloid. Interf. Sci. 1998, 3, 620–624.<br />

[48] Español, P. SIMU, Challenges in molecular simulations 2002, 4, 59–77.<br />

[49] Willemsen, S.; Vlugt, T.; Hoefsloot, H.; Smit, B. J. Comp. Phys. 1998, 147, 507–517.<br />

[50] Shillcock, J.; Lipowsky, R. J. Chem. Phys. 2002, 117, 5048–5061.<br />

[51] Gibson, J.; Chen, K.; Chynoweth, S. J. Colloid Interface Sci. 1998, 206, 464–474.<br />

[52] Alberts, B.; Johnson, A.; Lewis, J.; Raff, M.; Roberts, K.; Walters, P. Molecular biology <strong>of</strong><br />

the cell; Garlan Science: New York, 2002.<br />

[53] Rowlinson, J.; Widom, B. Molecular Theory <strong>of</strong> Capillarity; Clarendon Press: Oxford, 1982.<br />

[54] Jähnig, F. Biophys. J. 1996, 71, 1348–1349.<br />

[55] Kwok, R.; Evans, E. Biophys. J. 1981, 35, 637–652.<br />

[56] Chiu, S.; Clark, M.; Balaji, V.; Subramaniam, S.; Scott, H.; Jakobsson, E. Biophys. J. 1995,<br />

69, 1230–1245.<br />

[57] Zhang, Y.; Feller, S.; Brooks, B.; Pastor, R. J. Chem. Phys. 1995, 103, 10252–10266.<br />

[58] Feller, S.; Pastor, R. Biophys. J. 1996, 71, 1350–1355.<br />

[59] Feller, S.; Pastor, R. J. Chem. Phys. 1999, 111, 1281–1287.<br />

[60] Marrink, S.; Mark, A. J. Phys. Chem. B 2001, 105, 6122–6127.<br />

[61] Feller, S.; Zhang, Y.; Pastor, R. J. Chem. Phys. 1995, 103, 10267–10276.<br />

[62] Tieleman, D.; Berendsen, H. J. Chem. Phys. 1996, 105, 4871–4880.<br />

[63] Chiu, S.; Jakobsson, E.; Subramaniam, S.; Scott, H. Biophys. J. 1999, 77, 2462–2469.<br />

[64] Venable, R.; Brooks, R.; Pastor, R. J. Chem. Phys. 2000, 112, 4822–4832.<br />

[65] Sun, F. Biophys. J. 2002, 85, 2511 –2519.<br />

[66] Varnik, F.; Baschnagel, J.; Binder, K. J. Chem. Phys. 2000, 113, 4444–4453.<br />

[67] Kirkwood, J.; Buff, F. J. Chem. Phys. 1949, 17, 338.


7.4 Conclusion 121<br />

[68] Ono, S.; Kondo, S. Encyclopedia <strong>of</strong> physics, Vol. 10; S. Flugge, Springer: Berlin, 1960.<br />

[69] Hansen, J.; McDonald, I. Theory <strong>of</strong> Simple Liquids; Academic Press: London, 2nd ed.,<br />

1986.<br />

[70] Allen, M.; Tildesley, D. Computer Simulation <strong>of</strong> Liquids; Clarendon Press: Oxford, 1987.<br />

[71] Walton, J.; Tildesley, D.; Rowlinson, J. Mol. Phys. 1983, 48, 1357–1368.<br />

[72] Irving, J.; Kirkwood, J. J. Chem. Phys. 1950, 18, 817–829.<br />

[73] Harasima, A. Adv. Chem. Phys. 1958, 1, 203.<br />

[74] Nijmeijer, M.; Bakker, A.; Bruin, C.; Sikkenk, J. J. Chem. Phys. 1988, 89, 3789–3792.<br />

[75] Venturoli, M.; Smit, B. Phys.Chem.Comm. 1999, 10, 45–49.<br />

[76] Frenkel, D.; Smit, B. Underst<strong>and</strong>ing Molecular Simulations: from Algorithms to Applications;<br />

Academic Press: San Diego, 2002.<br />

[77] Lindahl, E.; Edholm, O. J. Chem. Phys. 2000, 113, 3882–3893.<br />

[78] Rawicz, W.; Olbrich, K.; McIntosh, T.; Needham, D.; Evans, E. Biophys. J. 2000, 79, 328–<br />

339.<br />

[79] Miller, D. M. Biochim. Biophys. Acta 1986, 856, 27–35.<br />

[80] Templer, R.; Castle, S.; Curran, A.; Rumbles, G.; Klug, D. Faraday Discuss. 1998, 111,<br />

41–53.<br />

[81] Gullingsrud, J.; Schulten, K. Biophys. J. 2004, 86, 3496–3509.<br />

[82] Harries, D.; Ben-Shaul, A. J. Chem. Phys. 1997, 106, 1609–1619.<br />

[83] Seelig, J.; Seelig, A. Q. Rev. Biophys. 1980, 13, 19–61.<br />

[84] Wiener, M.; White, S. Biophys. J. 1992, 61, 434–447.<br />

[85] Nagle, J.; Zhang, R.; Tristram-Nagle, S.; Sun, W.; Petrache, H.; Suter, R. Biophys. J. 1996,<br />

70, 1419–1431.<br />

[86] Nagle, J. Biophys. J. 1993, 64, 1476–1481.<br />

[87] Petrache, H.; Feller, S.; Nagle, J. Biophys. J. 1997, 70, 2237–2242.<br />

[88] Petrache, H.; Tu, K.; Nagle, J. Biophys. J. 1999, 76, 2749–2487.<br />

[89] Takaoka, Y.; Pasenkiewicz-Gierula, M.; ¡iyagawa, H.; Kitamura, K.; Tamura, Y.; Kusumi,<br />

A. Biophys. J. 2000, 79, 3118–3138.<br />

[90] McIntosh, T.; McDaniel, R.; Simon, S. Biochim. Biophys. Acta 1983, 731, 109–114.<br />

[91] Simon, S.; McIntosh, T. Biochim. Biophys. Acta 1984, 773, 169–172.<br />

[92] Petrache, H.; Dodd, S.; Brown, M. Biophys. J. 2000, 79, 3172–3192.<br />

[93] Marrink, S.; Berendsen, H. J. Phys. Chem. 1994, 98, 4155–4168.<br />

[94] Cantor, R. Biophys. J. 1999, 76, 2625–2639.<br />

[95] Cantor, S. Biophys. J. 2001, 80, 2284–2297.<br />

[96] Koynova, R.; Caffrey, M. Biochim. Biophys. Acta 1998, 1376, 91–145.<br />

[97] Katsaras, J.; Gutberlet, T. Lipid Bilayers. Structure <strong>and</strong> interactions; Springer-Verlag: New<br />

York, 2001.<br />

[98] Slater, J.; Huang, C.-H. Prog. Lipid. Res. 1988, 27, 325–359.<br />

[99] Furuike, S.; Levadny, V.; Li, S.; Yamazaki, M. Biophys. J. 1999, 77, 2015–2023.<br />

[100] McIntosh, T.; Lin, H.; Li, S.; Huang, C.-H. Biochim. Biophys. Acta 2001, 1510, 219–230.<br />

[101] Adachi, T.; Takahashi, H.; Ohki, K.; Hatta, I. Biophys. J. 1995, 68, 1850–1855.<br />

[102] Hata, T.; Matsuki, H.; Kaneshina, S. Biophys. Chem. 2000, 87, 25–36.<br />

[103] Lewis, R.; Winter, I.; Kriechbaum, M.; Lohner, K.; McElhaney, R. Biophys. J. 2001, 80,<br />

1329–1342.


122 <strong>Mesoscopic</strong> model for <strong>lipid</strong> <strong>bilayers</strong> <strong>with</strong> <strong>embedded</strong> proteins<br />

[104] Winter, I.; Pabst, G.; Rappolt, M.; Lohner, K. Chem. Phys. Lip. 2001, 112, 137–150.<br />

[105] Flyvbjerg, H.; Petersen, H. J. Chem. Phys. 1989, 91, 461–466.<br />

[106] Cevc, G.; Marsh, D. Phospho<strong>lipid</strong> <strong>bilayers</strong>: physical properties <strong>and</strong> <strong>models</strong>; John Wiley<br />

<strong>and</strong> Sons: New York, 1987.<br />

[107] Sun, W.-J.; Tristram-Nagle, S.; Suter, R.; Nagle, J. Biophys. J. 1996, 71, 885–891.<br />

[108] Misquitta, Y.; Caffrey, M. Biophys. J. 2001, 81, 1047–1058.<br />

[109] Koynova, R.; Brankov, J.; Tenchov, B. Eur. Biophys. J. 1997, 25, 261–275.<br />

[110] Takahashi, H.; Matsuo, A.; Hatta, I. Phys. Chem. Chem. Phys. 2002, 4, 2365–2370.<br />

[111] Iwanaga, T.; Suzuki, M.; Kunieda, H. Langmuir 1998, 14, 5775–5781.<br />

[112] Kranenburg, M.; Nicolas, J.; Smit, B. Phys. Chem. Chem. Phys. 2004, 6, 4142–4151.<br />

[113] S., T.-N.; Liu, Y.; Legleiter, J.; Nagle, J. F. Biophys. J. 2002, 83, 3324–3335.<br />

[114] Cantor, R. Toxicology Lett. 1998, 101, 451–458.<br />

[115] de Kruijff, B. Nature 1997, 386, 129–130.<br />

[116] Perozo, E.; Kloda, A.; Cortes, D.; Martinac, B. Nature Structural Biology 2002, 9, 696–703.<br />

[117] Tu, K.; Tarek, M.; Klein, M.; Scharf, D. Biophys. J. 1998, 75, 2123–2134.<br />

[118] Cantor, R. Chemistry <strong>and</strong> Physics <strong>of</strong> Lipids 1999, 101, 45–56.<br />

[119] Koubi, L.; Saiz, L.; Tarek, M.; Scharf, D.; Klein, M. J. Phys. Chem. B 2003, 107, 14500–<br />

14508.<br />

[120] van den Brink-van der Laan, E.; Dalbey, R. E.; Demel, R. A.; Killian, J. A.; de Kruijff, B.<br />

Biochemistry 2001, 40, 9677–9684.<br />

[121] Koubi, L.; Tarek, M.; Klein, M.; Scharf, D. Biophys. J. 2000, 78, 800–811.<br />

[122] Koubi, L.; Tarek, M.; B<strong>and</strong>yopadhyay, S.; Klein, M.; Scharf, D. Anesthesiology 2002, 97,<br />

848–855.<br />

[123] Koubi, L.; Tarek, M.; B<strong>and</strong>yopadhyay, S.; Klein, M.; Scharf, D. Biophys. J. 2001, 81, 3339–<br />

3345.<br />

[124] Mouritsen, O.; Blom, M. Biophys. J 1984, 46, 141–153.<br />

[125] Sackmann, E. In Biological membranes, Vol. 5, pages 105–143, London, 1984. Academic<br />

Press.<br />

[126] Mouritsen, O.; Sperotto, M. In Jackson, M., Ed., hermodynamics <strong>of</strong> Membrane Receptors<br />

<strong>and</strong> Channels, pages 127–181, Boca Raton, Florida, 1993. CRC Press, Inc.<br />

[127] Dumas, F.; Lebrum, M.; Tocanne, J.-F. FEBS Lett. 1999, 485, 271–277.<br />

[128] Munro, S. EMBO J. 1995, 14, 4695–4704.<br />

[129] Munro, S. Trends Cell Biol. 1998, 8, 11–15.<br />

[130] Bretscher, M.; Munro, S. Science 1993, 261, 1280–1281.<br />

[131] Pelham, H.; Munro, S. Cell 1993, 75, 603–605.<br />

[132] McIntosh, T.; Vidal, A.; Simon, S. Biophys. J 2003, 85, 1656–1666.<br />

[133] K., S.; Ikonen, E. Nature 1997, 387, 569–572.<br />

[134] Binder, W.; Barragan, V.; Menger, F. Angew. Chem. Int. Ed. 2003, 42, 5802–5827.<br />

[135] Fantini, J.; R., N. G.; Mahfoud.; Yahi, N. In Expert reviews in Molecular medicine, pages<br />

1–22. Cambridge University Press, 2002.<br />

[136] de Planque, M.; Killian, J. A. Mol. Membr. Biol. 2003, 20, 271–284.<br />

[137] Johansson, A.; Smith, G.; Metcalfe, J. Biochim. Biophys. Acta 1981, 641, 416–421.<br />

[138] In’t Veld, G.; Driessen, A.; den Kamp, J. O.; Konings, W. Biochim. Biophys. Acta 1991,<br />

1065, 203–212.


7.4 Conclusion 123<br />

[139] Lee, A. Biochim. Biophys. Acta 1998, 1376, 381–390.<br />

[140] Jost, P.; Hayes Griffith, O.; Capaldi, R.; V<strong>and</strong>erkooi, G. Proc. Natl. Acad. Sci. USA 1973, 70,<br />

480–484.<br />

[141] Hesketh, T.; Smith, G.; Houslay, M.; McGill, K.; Birdsall, N.; Metcalfe, J.; Warren, G. Biochemistry<br />

1976, 15, 4145–4151.<br />

[142] Jost, P.; Griffith, O. H. Ann. N.Y. Acad. Sci. USA 1980, 348, 391–405.<br />

[143] Rehorek, M.; Dencher, N.; Heyn, M. Biochemistry 1985, 24, 5980–5988.<br />

[144] Piknová, B.; Pérochon, E.; Tocanne, J.-F. Eur. J. Biochem. 1993, 218, 385–396.<br />

[145] Harroun, T. A.; Heller, W. T.; Weiss, T. M.; Yang, L.; Huang, H. W. Biophys. J. 1999, 76,<br />

937–945.<br />

[146] Bryl, K.; Yoshihara, K. Eur. Biophys. J. Biophys. 2001, 29, 628–640.<br />

[147] Dumas, F.; Sperotto, M.; Lebrum, M.; Tocanne, J.-F.; Mouritsen, O. Biophys. J 1997, 73,<br />

1940–1953.<br />

[148] Lehtonen, J.; Kinnunen, P. K. Biophys. J 1997, 72, 1247–1257.<br />

[149] Fahsel, S.; Pospiech, E.-M.; Zein, M.; Hazlet, T.; Gratton, E.; Winter, R. Biophys. J. 2002,<br />

83, 334–344.<br />

[150] Fern<strong>and</strong>es, F.; Loura, L.; Prieto, M.; Koehorst, R.; Spruijt, R.; Hemminga, M. Biophys. J.<br />

2003, 85, 2430–2441.<br />

[151] Glaubitz, C.; Grobner, G.; Watts, A. Biochim. Biophys. Acta 2000, 1463, 151–161.<br />

[152] Sharpe, S.; Barber, K.; Grant, C.; Goodyear, D.; Morrow, M. Biophys. J 2002, 83, 345–358.<br />

[153] van der Wel, P.; Str<strong>and</strong>berg, E.; Killian, J.; II, R. K. Biophys. J 2002, 83, 1479–1488.<br />

[154] Koehorst, R.; Spruijt, R.; Vergeldt, F.; Hemminga, M. Biophys. J. 2004, 87, 1445–1455.<br />

[155] Str<strong>and</strong>berg, E.; Özdirekcan, S.; Rijkers, D.; van der Wel, P.; II, R. K.; Liskamp, R.; Killian, J.<br />

Biophys. J. 2004, 86, 3709–3721.<br />

[156] Harzer, U.; Bechinger, B. Biochemistry 2000, 39, 13106–13114.<br />

[157] Liu, F.; Lewis, R.; Hodges, R.; McElhaney, R. Biochemistry 2002, 41, 9197–9207.<br />

[158] Mall, S.; Broadbridge, R.; Sharma, R.; East, J.; Lee, A. Biochemistry 2001, 40, 12379–12386.<br />

[159] Tocanne, J.-F. Comm. Mol. Cell. Biophys. 1992, 8, 53–72.<br />

[160] Tocanne, J.-F.; Cézanne, L.; Lopaz, A.; Piknová, B.; Schram, V.; Turnier, J.; Welby., M.<br />

Chem. Phys. Lipids 1994, 73, 139–159.<br />

[161] Thomson, T.; Sankaram, M.; Biltonen, R.; Marsh, D.; Vaz, W. Mol. Membr. Biol. 1995, 12,<br />

157–162.<br />

[162] de Planque, M.; Goormaghtigh, E.; Greathouse, D.; II, R. K.; Kruijtzer, J.; Liskamp, R.;<br />

de Kruijff, B.; Killian, J. A. Biochemistry 2001, 40, 5000–5010.<br />

[163] Ridder, A.; Hoef, W. V. D.; Stam, J.; Kuhn, A.; Kruijff, B. D.; Killian, J. Biochemistry 2002,<br />

41, 4946–4952.<br />

[164] Morein, S.; J.A., K.; Sperotto, M. Biophys. J 2002, 82, 1405–1417.<br />

[165] Killian, J. Biochim. Biophys. Acta 1992, 1113, 391–425.<br />

[166] Ep<strong>and</strong>, R. M. Biochim. Biophys. Acta 1998, 1376, 353–368.<br />

[167] Sperotto, M.; Mouritsen, O. Biophys. J. 1991, 59, 261–270.<br />

[168] T., S.; Baumgärtner, A. Biophys. J 1997, 73, 2251–2259.<br />

[169] Duque, D.; Li, X.; Katsov, K.; Schick, M. J. Chem. Phys. 2002, 116, 10478–10484.<br />

[170] Petrache, H. I.; Grossfield, A.; MacKenzie, K. R.; Engelman, D. M.; Woolf, T. J. Mol. Biol.<br />

2000, 302, 727–746.


124 <strong>Mesoscopic</strong> model for <strong>lipid</strong> <strong>bilayers</strong> <strong>with</strong> <strong>embedded</strong> proteins<br />

[171] Petrache, H. I.; Zuckerman, D.; Sachs, J.; Killian, J.; II, R. K.; Woolf, T. Langmuir 2002, 18,<br />

1340–1351.<br />

[172] MacKenzie, K.; Prestegard, J.; Engelmann, D. Science 1997, 276, 131–133.<br />

[173] Stopar, D.; Spruijt, R.; Wolfs, C.; Hemminga, M. Biochim. Biophys. Acta 2003, 1611, 5–15.<br />

[174] von Heijne, G.; Manoil, M. Protein Engineering 1990, 4, 109–112.<br />

[175] Henderson, R.; Unwin, P. Nature 1975, 257, 28–32.<br />

[176] Foster, D.; Boublik, M.; Kaback, H. J. Biol. Chem 1983, 258, 31–34.<br />

[177] Deisenh<strong>of</strong>er, J.; Epp, O.; Miki, K.; Huber, R.; Michael., H. Nature 1985, 318, 618–624.<br />

[178] Iwata, S.; Ostermeier, C.; Ludwig, B.; Michael, H. Nature 1995, 376, 660–669.<br />

[179] Fu, D.; Libson, A.; Miercke, L.; Weitzman, C.; <strong>and</strong>J. Krucinski, P. N.; Stround, R. Science<br />

2000, 290, 481–486.<br />

[180] A.A., N.; Mesleh, M.; Opella, S. Magn. Reson. Chem. 2004, 42, 162–171.<br />

[181] M.Ø. Jensen, private communication.<br />

[182] Ridder, A.; Spelbrink, E.; Demmers, J.; Rijkers, D.; Liskamp, R.; Brunner, J.; Heck, A.;<br />

Kruijff, B. D.; Killian, J. Biochemistry 2004, 43, 4482–4489.<br />

[183] T.M., W.; van der Wel, P.; Killian, J.; II., R. K.; Huang, H. Biophys. J 2003, 84, 379–385.<br />

[184] Kol, M.; van Delen, A.; de Kroon, A.; de Kruijff, B. J. Biol. Chem. 2003, 278, 24586–24593.<br />

[185] Brasseur, R. Mol. Membr. Biol. 2000, 17, 31–40.<br />

[186] Lins, L.; Charloteaux, B.; Thomas, A.; Brasseur, R. Proteins: Struct., Funct., Genet. 2001,<br />

44, 435–447.<br />

[187] Pebay-Peyroula, E.; Rummel, G.; Rosenbusch, J.; L<strong>and</strong>au, E. Science 1997, 277, 1676–<br />

1681.<br />

[188] Ostermeier, C.; Michael, H. Curr. Opinion Struct. Biol. 1997, 7, 697–701.


Summary<br />

Biological membranes, as they are found in all living cells, are complex systems.<br />

They contain many different molecules <strong>and</strong> display dynamic <strong>and</strong> structural properties<br />

which span many orders <strong>of</strong> magnitude, both in length <strong>and</strong> time scales. The<br />

characterization <strong>of</strong> the structure <strong>of</strong> <strong>lipid</strong> <strong>bilayers</strong>, their thermodynamic <strong>and</strong> dynamic<br />

behavior, the dependence <strong>of</strong> these properties on the membrane composition, as well<br />

as the interaction <strong>of</strong> the <strong>lipid</strong> bilayer <strong>with</strong> other molecules, are key questions in the<br />

underst<strong>and</strong>ing <strong>of</strong> how a membrane works <strong>and</strong> how specific functions are carried out.<br />

Different approaches can be adopted to address these questions. Disciplines like biology,<br />

biochemistry, physics, have applied different, <strong>and</strong> complementary, tools to the<br />

study <strong>of</strong> these complex systems. Computer simulations are a relatively new tool that<br />

has proved useful in the study <strong>of</strong> condensed matter systems. From the point <strong>of</strong> view<br />

<strong>of</strong> a physicist, a membrane is a s<strong>of</strong>t-matter, quasi two-dimensional, aggregate. Due to<br />

the non covalent interactions between the <strong>lipid</strong>s forming the bilayer matrix, a membrane<br />

is fluid. Lipids <strong>and</strong> other molecules can diffuse in the bilayer plane, or flip-flop<br />

from one monolayer to the other, small ions or molecules can cross it, <strong>and</strong> proteins<br />

can be incorporated. But a membrane is not a fluid in the sense that it constitutes a<br />

barrier, thought semi-permeable, which cannot be crossed by large molecules, <strong>and</strong><br />

which has the properties <strong>of</strong> an elastic sheet, it can bend, it has a specific rigidity, it<br />

can be compressed or stretched. Furthermore, because the <strong>lipid</strong>s <strong>of</strong> which it is constituted<br />

are oriented, a membrane has an internal structure. The <strong>lipid</strong> hydrophilic<br />

headgroups stick into the aqueous environment, <strong>and</strong> the <strong>lipid</strong> hydrophobic tails extend<br />

in the bilayer core.<br />

Computer simulations as a tool to study <strong>lipid</strong> <strong>bilayers</strong><br />

The subject <strong>of</strong> this thesis is the study <strong>of</strong> <strong>lipid</strong> <strong>bilayers</strong> <strong>with</strong> computer simulations, using<br />

a mesoscopic approach. This mesoscopic approach consists in coarse-graining<br />

the molecules that form the membrane, i.e. the <strong>lipid</strong>s. Instead <strong>of</strong> considering an<br />

atomistic representations <strong>of</strong> the <strong>lipid</strong>s, groups <strong>of</strong> atoms are lumped together to form<br />

beads, which are connected by springs to build the <strong>lipid</strong> molecules. These beads interact<br />

via simplified forces, which are described in Chapter 2. In this Chapter we have<br />

also introduced the Dissipative Particle Dynamics technique which we use to simulate<br />

the movement <strong>of</strong> these particles. A first, important result <strong>of</strong> this approach, is<br />

that, despite the purely repulsive nature <strong>of</strong> the interactions between particles <strong>of</strong> different<br />

type, we find that the model <strong>lipid</strong>s can phase separate <strong>and</strong> form <strong>bilayers</strong>, <strong>and</strong><br />

that the formation <strong>of</strong> the bilayer depends on a single parameter, i.e. on the relative<br />

strength <strong>of</strong> the repulsion between hydrophilic <strong>and</strong> hydrophobic beads.<br />

It is known experimentally, <strong>and</strong> by thermodynamic considerations, that unconstrained,<br />

self-assembled, <strong>lipid</strong> <strong>bilayers</strong> are in a tensionless state. To mimic such a


126 Summary<br />

state, we introduce in Chapter 3 a fast <strong>and</strong> efficient simulation technique, based on<br />

the Monte Carlo method, to impose on the bilayer a chosen value <strong>of</strong> the surface tension,<br />

<strong>of</strong> which the zero value is a particular case. By means <strong>of</strong> this simulation method<br />

we are able to study the area compressibility <strong>of</strong> the bilayer, <strong>and</strong> investigate finite-size<br />

effects as function <strong>of</strong> the applied surface tension. We found that for tensionless or<br />

stretched <strong>bilayers</strong> the finite-size effects are very small.<br />

Phase behavior <strong>of</strong> <strong>lipid</strong> <strong>bilayers</strong><br />

In Chapter 4 we optimize the model by studying the characteristics <strong>of</strong> the <strong>lipid</strong> architecture<br />

that are required to correctly reproduce the structural <strong>and</strong> thermodynamic<br />

properties <strong>of</strong> <strong>lipid</strong> <strong>bilayers</strong>. We find that chain stiffness is an important parameter in<br />

reproducing the correct density distribution <strong>and</strong> area dependence on chain length.<br />

We find that for single-tail <strong>lipid</strong>s the value <strong>of</strong> the interaction parameter between the<br />

headgroups is a key factor in determining the internal structure <strong>of</strong> a bilayer. Due<br />

to the small hydrophobic section <strong>of</strong> this <strong>lipid</strong> model, a large repulsion between the<br />

headgroups results in an interdigitated bilayer; interdigitation disappears when the<br />

repulsion parameter is decreased.<br />

Interdigitated phases are observed experimentally for phospho<strong>lipid</strong>s in the gel<br />

phase. They do not occur spontaneously, but have to be induced. We propose that<br />

the general mechanism that is responsible for inducing interdigitation in a bilayer is<br />

a decrease in the interfacial density <strong>of</strong> the <strong>lipid</strong> headgroups. We investigate this hypothesis<br />

in Chapter 5 where we study the phase diagram <strong>of</strong> single-tail <strong>lipid</strong>s. We find<br />

a fluid <strong>and</strong> a gel phase for these <strong>lipid</strong>s, <strong>and</strong> in the gel phase we find a transition from<br />

the interdigitated to the non-interdigitated state which is induced by a decrease in<br />

the headgroup repulsion. This suggests that it is possible to induce an interdigitated<br />

phase in <strong>bilayers</strong> <strong>of</strong> single-tail <strong>lipid</strong>s by adding chaotropic salts, which are adsorbed<br />

at the bilayer interfacial region <strong>and</strong> increase the amount <strong>of</strong> interfacial water.<br />

Next we study the phase diagram <strong>of</strong> double-tail <strong>lipid</strong>s, <strong>and</strong> in particular <strong>of</strong> a <strong>lipid</strong><br />

<strong>with</strong> five hydrophobic beads in each tail <strong>and</strong> three hydrophilic head-beads, which<br />

can be seen as a coarse-grained representation for the DMPC phospho<strong>lipid</strong>. For<br />

this <strong>lipid</strong> model, we find the phases which are experimentally observed in phosphocholine<br />

<strong>bilayers</strong>. These phases are the low temperature gel phase, or Lβ ′, in which the<br />

<strong>lipid</strong> tails show a tilt respect to the normal to the bilayer plane <strong>and</strong> have a low diffusion<br />

coefficient, <strong>and</strong> the high temperature fluid phase, or Lα, in which the <strong>lipid</strong> tails<br />

are partially disordered, <strong>and</strong> the <strong>lipid</strong>s can diffuse in the bilayer plane. Between these<br />

two phases we find a narrow region <strong>of</strong> temperature in which a striated phase is observed,<br />

which resembles the ripple, or Pβ ′, phase <strong>of</strong> the phosphatidylcholine <strong>bilayers</strong>.<br />

By mapping the reduced phase transition temperatures <strong>of</strong> the simulated system<br />

onto the experimental main ( Pβ ′ → Lα) <strong>and</strong> pre-transition (Lβ ′ → Pβ ′) temperatures<br />

for DMPC <strong>bilayers</strong>, we are able to quantitatively compare bilayer properties like the<br />

area per <strong>lipid</strong> or the bilayer thickness <strong>with</strong> their experimental values. The very good


agreement we find gives us confidence in the ability <strong>of</strong> our coarse-grained approach<br />

to model real phospho<strong>lipid</strong>s systems.<br />

Lateral pressure in <strong>lipid</strong> <strong>bilayers</strong><br />

While the surface tension <strong>of</strong> a <strong>lipid</strong> bilayer, or monolayer, can be measured experimentally,<br />

there are no experimental techniques to measure the local distribution <strong>of</strong><br />

pressure. Simulations can then be an helpful tool to characterize the shape <strong>of</strong> the<br />

pressure pr<strong>of</strong>ile in <strong>lipid</strong> <strong>bilayers</strong>. The study <strong>of</strong> the effect <strong>of</strong> local redistribution <strong>of</strong><br />

lateral pressure in <strong>bilayers</strong> induced by site dependent changes in the <strong>lipid</strong> topology<br />

can help to answer some important questions. Can shifts in the lateral pressure be<br />

induced by changes in the <strong>lipid</strong> structure, like different level <strong>of</strong> unsaturation in the<br />

chain? How do these changes depend on the position <strong>and</strong> nature <strong>of</strong> the modification<br />

in the <strong>lipid</strong> topology? A related <strong>and</strong> interesting question is the possible correlation<br />

between the local distribution <strong>of</strong> the pressure pr<strong>of</strong>ile <strong>and</strong> the partitioning <strong>of</strong> small<br />

molecules in the bilayer: where do small solutes preferentially adsorb in a <strong>lipid</strong> bilayer?<br />

And, is the partitioning <strong>of</strong> solutes correlated to the distribution <strong>of</strong> pressure<br />

<strong>with</strong>in the bilayer? To address these questions, in Chapter 4 we describe the shape<br />

<strong>of</strong> the lateral pressure distribution in <strong>lipid</strong> <strong>bilayers</strong> for different <strong>lipid</strong> topologies. We<br />

show that the pressure distribution is tightly correlated <strong>with</strong> the density distribution,<br />

i.e. <strong>with</strong> the localization <strong>of</strong> the <strong>lipid</strong> segments <strong>with</strong>in the bilayer, while it is less sensitive<br />

to the functional form <strong>and</strong> parameterization <strong>of</strong> the interaction potentials. To<br />

support the latter observation, we compare the pressure pr<strong>of</strong>ile in a bilayer <strong>of</strong> coarsegrained<br />

double-tail <strong>lipid</strong>s <strong>with</strong> the pressure pr<strong>of</strong>ile computed in atomistic molecular<br />

dynamics simulations <strong>of</strong> a DPPC bilayer. We show that the shapes <strong>of</strong> the pressure<br />

pr<strong>of</strong>iles calculated <strong>with</strong> the two different approaches have remarkable similarities.<br />

In Chapter 6 we investigate the interaction <strong>of</strong> small solutes, which can model<br />

anesthetics, <strong>with</strong> the <strong>lipid</strong> <strong>bilayers</strong>. We show that the partitioning <strong>of</strong> small solutes<br />

molecules in a bilayer is mainly driven by chemical affinity, i.e. by the degree <strong>of</strong> hydrophobicity<br />

<strong>of</strong> the solutes. But also another effect is found; these molecules locate<br />

preferentially in the region <strong>of</strong> the bilayer where the pressure pr<strong>of</strong>ile has local minima.<br />

These observations can help to predict the partitioning <strong>of</strong> molecules like anesthetics<br />

in the bilayer. Furthermore, we show that, if the molecules are located in<br />

the headgroups region <strong>of</strong> the bilayer, they significantly decrease the local pressure in<br />

this region. Since anesthetics are also incorporated in the bilayer interfacial region,<br />

we postulate that these molecules could regulate the activity <strong>of</strong> membrane proteins<br />

trough changes is the pressure distribution.<br />

The hydrophobic mismatch<br />

In Chapter 7 we extend the coarse-grained model, <strong>and</strong> apply it to the study <strong>of</strong> <strong>lipid</strong><br />

<strong>bilayers</strong> <strong>with</strong> <strong>embedded</strong> proteins, at low protein-to-<strong>lipid</strong> ratios. In particular we in-<br />

127


128 Summary<br />

vestigate hydrophobic mismatch. The matching between the <strong>lipid</strong> bilayer hydrophobic<br />

thickness <strong>and</strong> the hydrophobic length <strong>of</strong> integral membrane proteins has been<br />

proposed as a generic physical principle on which the <strong>lipid</strong>-protein interaction in<br />

biomembranes is based. The energy cost <strong>of</strong> exposing polar moieties, from either hydrocarbon<br />

chains or protein residues, is so high that the hydrophobic part <strong>of</strong> the <strong>lipid</strong><br />

bilayer should match the hydrophobic domain <strong>of</strong> membrane proteins. The results<br />

from a number <strong>of</strong> investigations have indeed pointed out the relevance <strong>of</strong> the hydrophobic<br />

matching in relation to the <strong>lipid</strong>-protein interactions, hence to membrane<br />

organization <strong>and</strong> biological function. To address these issues, we study <strong>bilayers</strong> <strong>with</strong><br />

<strong>embedded</strong> a single protein, which can vary in size <strong>and</strong> in hydrophobic length.<br />

We compute the coherence length for the protein-induced spatial fluctuations,<br />

<strong>and</strong> quantify it in terms <strong>of</strong> a decay length <strong>of</strong> the <strong>lipid</strong> bilayer hydrophobic thickness<br />

pr<strong>of</strong>ile around the protein. This quantity is not accessible to experimental measurements,<br />

but can be important in determining the range <strong>of</strong> the protein-induced perturbation,<br />

<strong>and</strong> can be relevant in predicting <strong>lipid</strong>-mediated protein-protein interaction<br />

<strong>and</strong> aggregation. We find that the protein-induced <strong>lipid</strong> perturbation depends on the<br />

mismatch <strong>and</strong> on the temperature <strong>of</strong> the system, <strong>and</strong> that a protein-size dependence<br />

appears for values <strong>of</strong> temperatures approaching from above the transition temperature<br />

<strong>of</strong> the pure <strong>lipid</strong> bilayer, i.e. in the fluid phase <strong>of</strong> the system.<br />

An interesting <strong>and</strong> unexpected result <strong>of</strong> this study, is that we find that for large<br />

mismatch conditions, either positive or negative, <strong>and</strong> for large proteins, the decay <strong>of</strong><br />

the protein-induced perturbation on the bilayer thickness is not <strong>of</strong> the exponential<br />

type, as previously predicted by lattice <strong>models</strong>. We find, instead, a non-monotonous<br />

decay. In the region next to the one closest to the protein hydrophobic surface <strong>and</strong><br />

where hydrophobic matching occurs, the bilayer hydrophobic thickness is lower (undershooting)<br />

or higher (overshooting) than the unperturbed equilibrium thickness,<br />

depending on the sign <strong>of</strong> the mismatch. These localized changes in bilayer hydrophobic<br />

thickness can be <strong>of</strong> biological relevance for phenomena like cell adhesions, fusion,<br />

membrane rupture, or membrane permeability.<br />

We also consider the effect <strong>of</strong> the bilayer on the proteins, <strong>and</strong> we investigate the<br />

mismatch conditions under which membrane proteins may assume a tilted conformation<br />

<strong>with</strong> respect to the <strong>lipid</strong> bilayer normal. We find a size dependence <strong>of</strong> the<br />

tilt; large proteins, due to energetic <strong>and</strong> topological reasons, do not undergo relevant<br />

tilt, even for large (positive) mismatch, while smaller proteins undergo a tilt <strong>with</strong> an<br />

angle that increases <strong>with</strong> increasing mismatch.<br />

At temperatures below the main transition temperature, but above the pre-transition<br />

temperature <strong>of</strong> the pure system, i.e. in the ripple or striated phase, our results<br />

show that the <strong>embedded</strong> model-protein prefers to segregate along those stripes<br />

where the hydrophobic bilayer thickness matches the protein hydrophobic length.


Samenvatting<br />

Biologische membranen, zoals die gevonden worden in alle levende cellen, zijn complexe<br />

systemen. Ze bevatten veel verschillende molekulen en vertonen dynamische<br />

en strukturele eigenschappen die vele ordes van grootte omvatten, zowel in lengteals<br />

tijdschaal. De karakterisatie van de struktuur van <strong>lipid</strong>e bilagen, hun thermodynamisch<br />

<strong>and</strong> dynamisch gedrag, de afhankelijkheid van deze eigenschappen van<br />

de samenstelling van het membraan, zowel als de wisselwerking van de <strong>lipid</strong>e bilaag<br />

met <strong>and</strong>ere molekulen, zijn sleutelvragen om te begrijpen hoe een membraan<br />

werkt en hoe specifieke funkties uitgevoerd worden. Verschillende benaderingen<br />

kunnen worden gebruikt om deze vragen aan te pakken. Disciplines zoals biologie,<br />

biochemie en natuurkunde, hebben verschillende en elkaar complementerende<br />

methoden gebruikt om deze complexe systemen te onderzoeken. Computer simulatie<br />

is een relatief nieuwe methode, die nuttig is gebleken bij het bestuderen van<br />

gecondenseerde materie. Vanuit het gezichtspunt van de fysicus is een membraan<br />

een zacht, quasi twee-dimensionaal aggregaat. Ten gevolge van de niet-covalente<br />

wisselwerking tussen de <strong>lipid</strong>en die de bilaag vormen, is een membraan vloeibaar.<br />

Lipiden en <strong>and</strong>ere molekulen kunnen in het vlak van de bilaag diffunderen, <strong>of</strong> van de<br />

ene monolaag naar de <strong>and</strong>ere springen, kleine ionen <strong>of</strong> molekulen kunnen erdoor<br />

heen en eiwitten kunnen worden opgenomen. Maar een membraan is geen vloeist<strong>of</strong><br />

in de zin dat het een barriere vormt, naar men aanneemt semi-permeabel, waar grote<br />

molekulen niet doorheen kunnen, en dat de eigenschappen heeft van een elastische<br />

plaat. Het kan buigen, heeft een specifieke stijfheid en het kan samengedrukt <strong>of</strong><br />

uitgerekt worden. Bovendien, omdat de <strong>lipid</strong>en waaruit het membraan bestaat een<br />

bepaalde orientatie hebben, heeft een membraan een interne struktuur. De <strong>lipid</strong>e<br />

hydr<strong>of</strong>iele kopgroepen steken in de waterige omgeving en hydr<strong>of</strong>obe staarten steken<br />

in de kern van de bilaag.<br />

Computer simulaties als een middel om <strong>lipid</strong>e bilagen te onderzoeken<br />

Het onderwerp van dit proefschrift is het bestuderen van <strong>lipid</strong>e bilagen met computer<br />

simulaties op mesoscopische schaal. Deze mesoscopische benadering bestaat<br />

uit het “coarse-grainen” van de <strong>lipid</strong>e molekulen die de bilaag vormen. In plaats<br />

van een atomaire representatie van de <strong>lipid</strong>en, worden groepen atomen samengesmolten<br />

tot “kralen”, die zijn verbonden in een kralensnoer door middel van veertjes.<br />

Deze kralen interakteren door middel van een versimpeld krachtveld, dat wordt<br />

beschreven in ho<strong>of</strong>dstuk 2. In dit ho<strong>of</strong>dstuk introduceren we ook de Dissipative Particle<br />

Dynamics techniek die we gebruiken om de beweging van deze deeltjes na te<br />

bootsen.<br />

Het eerste, belangrijke resultaat van deze aanpak is dat, ondanks de uitsluitend<br />

repulsieve wisselwerking tussen deeltjes van verschillende types, we vinden dat de


130 Samenvatting<br />

model <strong>lipid</strong>en kunnen fasescheiden en bilagen vormen, en dat de vorming van de<br />

bilaag afhangt van één enkele parameter, n.l. de relatieve sterkte van de repulsie<br />

tussen hydr<strong>of</strong>iele en hydr<strong>of</strong>obe kralen.<br />

Het is experimenteel bekend, en ook uit thermodynamische overwegingen, dat<br />

vrij-zwevende, zelf-geassembleerde <strong>lipid</strong>e bilagen in een spanningsloze toest<strong>and</strong> verkeren.<br />

Om zo’n toest<strong>and</strong> na te bootsen, introduceren we in Ho<strong>of</strong>dstuk 3 een snelle<br />

en efficiente simulatie techniek, gebaseerd op de Monte Carlo methode, om op de<br />

bilaag een bepaalde oppervlakte-spanning uit te oefenen, waarvan de nul-waarde<br />

een speciaal geval is. Door middel van deze simulatiemethode zijn we in staat om<br />

de kompressibiliteit van het oppervlak te bestuderen, en finite-size effekten als funktie<br />

van de opgelegde oppervlakte spanning. We vonden dat, voor spanningsloze <strong>of</strong><br />

uitgerekte bilagen, de finite-size effekten erg klein zijn.<br />

Fase gedrag van <strong>lipid</strong>e bilagen<br />

In Ho<strong>of</strong>dstuk 4 optimaliseren we het model door het bestuderen van de karakteristieken<br />

van de <strong>lipid</strong>e architektuur, die nodig zijn om de strukturele en thermodynamiche<br />

eigenschappen van <strong>lipid</strong>e bilagen te reproduceren. We vonden dat de<br />

stijfheid van de keten een belangrijke parameter is om de juiste dichtheidsverdeling<br />

en oppervlakte afhanke-lijkheid van de ketenlengte te reproduceren. We vinden dat<br />

voor <strong>lipid</strong>en met één enkele staart, de waarde van de interaktie parameter tussen de<br />

kopgroepen een sleutelrol speelt in het bepalen van de interne struktuur van een bilaag.<br />

Als gevolg van de kleine hydr<strong>of</strong>obe doorsnede van dit <strong>lipid</strong>e model, resulteert<br />

een grote repulsie tussen de koppen in een geïnterdigiteerde bilaag. Deze interdigitatie<br />

verdwijnt als de waarde van deze parameter kleiner wordt.<br />

Geïnterdigiteerde fasen worden experimenteel geobserveerd voor phospho<strong>lipid</strong>en<br />

in de gelfase. Ze treden niet spontaan op maar moeten geïnduceerd worden.<br />

Wij stellen voor dat het algemene mechanisme dat verantwoordelijk is voor het induceren<br />

van interdigitatie in een bilaag, een afname van de dichtheid van de <strong>lipid</strong>e<br />

kopgroepen aan het grensvlak is. We onderzoeken deze hypothese in Ho<strong>of</strong>dstuk 5<br />

waar we het fasendiagram van <strong>lipid</strong>en met een enkele staart bestuderen. We vinden<br />

een vloeibare en een gelfase voor deze <strong>lipid</strong>en, en in de gel fase vinden we een overgang<br />

van de geïnterdigiteerde naar de niet-geïnterdigiteerde toest<strong>and</strong> geïnduceerd<br />

door een afname in de kopgroeprepulsie. Dit suggereert dat het mogelijk is om een<br />

geïnterdigiteerde fase te induceren in bilagen van <strong>lipid</strong>en met één enkele staart door<br />

het toevoegen van chaotrope zouten, die geadsorbeerd worden aan het grensvlak en<br />

de hoeveelheid water aan het grensvlak doen toenemen.<br />

Vervolgens bestuderen we het fasendiagram van <strong>lipid</strong>en met een dubbele staart,<br />

en in het bijzonder van een <strong>lipid</strong>e met vijf hydr<strong>of</strong>obe kralen in elke staart en drie<br />

hydr<strong>of</strong>iele kop-kralen, dat beschouwd kan worden als een gr<strong>of</strong>korrelige representatie<br />

van het DMPC phospho<strong>lipid</strong>e. Voor dit <strong>lipid</strong>e model vinden we de fasen die<br />

experimenteel geobserveerd worden in phosphocholine bilagen. Deze fasen zijn a)


de lage temperatuur gel fase <strong>of</strong> Lβ ′, waarin de <strong>lipid</strong>e staarten een tilt vertonen ten<br />

opzichte van de normaal op het vlak van de bilaag en een kleine diffusie coëfficient<br />

hebben, en b) de hoge temperatuur fase, <strong>of</strong> Lα, waarin the <strong>lipid</strong>e staarten gedeeltelijk<br />

wanordelijk zijn, en de <strong>lipid</strong>en kunnen diffunderen in het vlak van de bilaag. Tussen<br />

deze twee fasen vinden we een smal gebied in temperatuur waarin een “gestreepte-<br />

fase” wordt geobserveerd, die lijkt op de “ribbel” <strong>of</strong> Pβ ′ fase van de fosfatidylcholine<br />

bilagen. Door de gereduceerde fase-overgangstemperatuur van het gesimuleerde<br />

systeem te mappen op de experimentele ho<strong>of</strong>d- ( Pβ ′ → Lα) en pre-overgangs- (Lβ ′<br />

→ Pβ ′) temperaturen voor DMPC bilagen, zijn we in staat om een kwantitatieve<br />

vergelijking te maken wat betreft de bilaag eigenschappen, zoals de oppervlakte per<br />

<strong>lipid</strong>e <strong>of</strong> de dikte van de bilaag, met hun experimentele waarden. We vinden een uitstekende<br />

overeenkomst en dit geeft ons vertrouwen dat onze gr<strong>of</strong>korrelige benadering<br />

een goed model is voor realistische phospho<strong>lipid</strong>e systemen.<br />

Laterale druk in <strong>lipid</strong>e bilagen<br />

Terwijl de oppervlakte spanning van een <strong>lipid</strong>e bilaag, <strong>of</strong> monolaag, experimenteel<br />

gemeten kan worden, zijn er geen experimentele technieken om de lokale drukverdeling<br />

te meten. Simulaties kunnen dan een hulpmiddel zijn om de vorm van het drukpr<strong>of</strong>iel<br />

in <strong>lipid</strong>e bilagen te karakteriseren. De studie van het effekt van lokale herverdeling<br />

van laterale druk in bilagen, geinduceerd door plaatsafhankelijke ver<strong>and</strong>eringen<br />

in de topologie van het <strong>lipid</strong>e, kan van pas komen om enige belangrijke vragen<br />

op te lossen. Kunnen ver<strong>and</strong>eringen in de laterale druk geinduceerd worden door<br />

ver<strong>and</strong>eringen in de struktuur van het <strong>lipid</strong>e, zoals verschillende gradaties van onverzadigdheid<br />

in de keten? Hoe hangen deze ver<strong>and</strong>eringen af van de positie en aard<br />

van de ver<strong>and</strong>ering in de topologie van het <strong>lipid</strong>e?<br />

Een gerelateerde en interessante vraag is <strong>of</strong> er een correlatie is tussen de lokale<br />

verdeling van het drukpr<strong>of</strong>iel en de verdeling van kleine molekulen in de bilaag: waar<br />

adsorberen kleine opgeloste molekulen bij voorkeur in een <strong>lipid</strong>e bilaag? En is de<br />

verdeling van opgeloste molekulen gecorreleerd met de vorm van de drukverdeling<br />

in de dubbellaag?<br />

Om deze vragen aan te pakken, beschrijven we in Ho<strong>of</strong>dstuk 4 de vorm van de laterale<br />

drukverdeling in <strong>lipid</strong>e bilagen voor verschillende topologieën van het <strong>lipid</strong>e.<br />

We laten zien dat de drukverdeling nauw samenhangt met de dichtheidverdeling,<br />

d.w.z. met de lokalisatie van de <strong>lipid</strong>e segmenten in de bilaag, terwijl de drukverdeling<br />

minder gevoelig is voor de funktionele vorm en parameterisatie van de interaktie<br />

potentialen. Om deze observatie te ondersteunen, vergelijken we het drukpr<strong>of</strong>iel in<br />

een bilaag van gr<strong>of</strong>korrelige twee-staartige <strong>lipid</strong>en met het druk pr<strong>of</strong>iel berekend uit<br />

atomistische Molekulaire Dynamica simulaties van een DPPC bilaag. We laten zien<br />

dat de vormen van de drukpr<strong>of</strong>ielen, berekend met de twee verschillende benaderingen,<br />

opmerkelijke overeenkomsten hebben.<br />

In Ho<strong>of</strong>dstuk 6 onderzoeken we de wisselwerking van kleine opgeloste molekulen,<br />

131


132 Samenvatting<br />

als model voor verdovingsmiddelen, met de <strong>lipid</strong>e bilagen. We laten zien dat de<br />

verdeling van kleine opgeloste molekulen in een bilaag ho<strong>of</strong>dzakelijk bepaald wordt<br />

door chemische affiniteit, d.w.z. door de mate van hydr<strong>of</strong>obiciteit van de opgeloste<br />

st<strong>of</strong>fen. Maar we vonden ook een <strong>and</strong>er effekt; deze molekulen zitten bij voorkeur<br />

op een plek in de bilaag waar het drukpr<strong>of</strong>iel lokale minima heeft. Deze observaties<br />

kunnen helpen om de verdeling van molekulen, zoals narcotica, in de dubbellaag<br />

te voorspellen. Bovendien laten we zien dat, als de molekulen in de kopgroep regio<br />

van de bilaag zitten, ze de lokale druk in dit gebied sterk verlagen. Aangezien<br />

verdovingsmiddelen ook opgenomen worden in het grensvlak gebied van de bilaag,<br />

postuleren we dat deze molekulen de aktiviteit van membraan eiwitten kunnen reguleren<br />

door ver<strong>and</strong>eringen in de drukverdeling.<br />

De hydr<strong>of</strong>obe “mismatch”<br />

In Ho<strong>of</strong>dstuk 7 breiden we het gr<strong>of</strong>korrelige model uit en passen het toe op de studie<br />

van <strong>lipid</strong>e bilagen met ingesloten eiwitten, bij lage verhoudingen van eiwit : <strong>lipid</strong>e.<br />

In het bijzonder onderzoeken we de hydr<strong>of</strong>obe “mismatch”.<br />

Het is voorgesteld dat het bij elkaar passen van de hydr<strong>of</strong>obe dikte van de <strong>lipid</strong>e<br />

bilaag en de hydr<strong>of</strong>obe lengte van integrale membraan eiwitten, een algemeen fysisch<br />

principe is, waarop de <strong>lipid</strong>e-eiwit interaktie in celmembranen is gebaseerd.<br />

De energetische kosten van het blootstellen van polaire groepen aan koolwaterst<strong>of</strong><br />

ketens <strong>of</strong> eiwit residuen, is zo hoog dat het hydr<strong>of</strong>obe deel van de <strong>lipid</strong>e bilaag zou<br />

moeten passen bij het hydr<strong>of</strong>obe domein van membraan-eiwitten. De resultaten van<br />

een aantal onderzoeken hebben inderdaad uitgewezen dat de hydr<strong>of</strong>obe “matching”,<br />

in relatie met de <strong>lipid</strong>e eiwit interakties, en daarom met membraan organisatie en biologische<br />

funktie, relevant is. Om deze zaken aan te pakken, bestuderen we bilagen<br />

met één ingesloten eiwit, dat kan varieren in grootte en hydr<strong>of</strong>obe lengte.<br />

We berekenen de coherentie lengte voor de eiwit-geinduceerde ruimtelijke fluktuaties,<br />

<strong>and</strong> kwantificeren deze in termen van een verval lengte van het hydr<strong>of</strong>obe<br />

dikte pr<strong>of</strong>iel van de <strong>lipid</strong>e bilaag rond het eiwit. Deze grootheid is niet toegankelijk<br />

voor experi-mentele metingen, maar kan belangrijk zijn bij het bepalen van de<br />

strekking van de eiwit-geinduceerded perturbatie, en kan relevant zijn bij het voorspellen<br />

van <strong>lipid</strong>e-bemiddelde eiwit-eiwit interaktie en aggregatie. We vinden dat<br />

de eiwit-geinduceerde <strong>lipid</strong>e perturbatie afhangt van de mismatch en van de temperatuur<br />

van het systeem, en dat een eiwit-grootte afhankelijkheid verschijnt voor<br />

waarden van de temperatuur die van boven de overgangstemperatuur van de pure<br />

<strong>lipid</strong>e bilaag benaderen, d.w.z. in de vloeibare fase van het systeem.<br />

Een interessant en onverwacht resultaat van dit onderzoek is dat we vinden dat,<br />

voor grote “mismatch” condities, hetzij positief, hetzij negatief, en voor grote eiwitten,<br />

het verval van de eiwit-geinduceerde perturbatie met de dikte van de bilaag,<br />

niet exponentieel van aard is, zoals vroeger werd voorspeld door roostermodellen.<br />

In plaats daarvan vinden we een niet-monotoon verval. In het gebied naast het ge-


ied dat het dichtst bij het hydr<strong>of</strong>obe oppervlak van het eiwit is, en waar hydr<strong>of</strong>obe<br />

“matching” optreedt, is de hydr<strong>of</strong>obe dikte van de bilaag kleiner (undershooting)<br />

<strong>of</strong> groter (overshooting) dan de onverstoorde evenwichtsdikte, afhankelijk van het<br />

teken van de mismatch. Deze plaatselijke ver<strong>and</strong>eringen in de hydr<strong>of</strong>obe dikte van<br />

de bilaag kunnen van belang zijn voor biologische verschijnselen zoals cel-adhesie,<br />

fusie, membraan scheuring <strong>of</strong> membraan permeabiliteit.<br />

We beschouwen ook het effekt van de bilaag op de eiwitten, en we onderzoeken<br />

de mismatch condities waaronder aangenomen kan worden dat membraaneiwitten<br />

een scheve conformatie aannemen ten opzichte van de normaal op het bilaag oppervlak.<br />

We vinden dat dit afhangt van de grootte: grote eiwitten gaan niet scheef<br />

staan, ten gevolge van energetische en topologische effekten, zelfs voor een grote<br />

(positieve) mismatch, terwijl kleinere eiwitten scheef gaan staan onder een hoek die<br />

toeneemt met toenemende mismatch.<br />

Bij temperaturen onder de ho<strong>of</strong>d-overgangstemperatuur, maar boven de pre-transitie<br />

temperatuur van het pure systeem, d.w.z. in de ribbel <strong>of</strong> gestreepte fase, laten<br />

onze resultaten zien dat de ingesloten model-eiwitten de voorkeur geven aan een<br />

scheiding langs de strepen waar de hydr<strong>of</strong>obe dikte van de bilaag overeenkomt met<br />

de hydr<strong>of</strong>obe lengte.<br />

133


Curriculum Vitae<br />

Maddalena Venturoli was born in Rome, in 1970. In 1989 she got her diploma in classical<br />

studies at “Liceo Classico Giulio Cesare” in Rome. In 1998 she graduated cum<br />

laude in physics at “Università La Sapienza” <strong>of</strong> Rome, <strong>with</strong> a thesis on Molecular Dynamics<br />

<strong>of</strong> Myoglobin, under the supervision <strong>of</strong> pr<strong>of</strong>. dr. Giovanni Ciccotti. Between<br />

1998 <strong>and</strong> 2002 she did her doctoral research at the “Universiteit van Amsterdam”,<br />

<strong>with</strong> the supervision <strong>of</strong> pr<strong>of</strong>. dr. ir. Berend Smit. The results <strong>of</strong> this research are the<br />

subject <strong>of</strong> this thesis. In 2003 she started to work as a post-doc at the University College<br />

<strong>of</strong> London in the group <strong>of</strong> pr<strong>of</strong>. dr. Peter V. Coveney <strong>with</strong>in the UK RealityGrid<br />

project, <strong>and</strong> in collaboration <strong>with</strong> Schlumberger Cambridge Research, on the topic<br />

<strong>of</strong> lattice-Bolzmann simulations <strong>of</strong> flow in porous media.


This thesis is based on the following publications:<br />

• M. Venturoli <strong>and</strong> B. Smit, Phys. Chem. Comm., 1999, 10, 45-49,<br />

”Simulating the self-assembly <strong>of</strong> model membranes”<br />

Publications<br />

• M. Kranenburg, M. Venturoli <strong>and</strong> B. Smit, Phys. Rev. E, 2003, 67, art.nr. 060901(R),<br />

”Molecular simulations <strong>of</strong> mesoscopic bilayer phases”<br />

• M. Kranenburg, M. Venturoli <strong>and</strong> B. Smit, J. Phys. Chem. B., 2003, 107, 11491-<br />

11501,<br />

”Phase behavior <strong>and</strong> induced interdigitation in <strong>bilayers</strong> studied <strong>with</strong> dissipative<br />

particle dynamics”<br />

• M. Venturoli, B. Smit <strong>and</strong> M.M. Sperotto, accepted for publication in Biophys. J.<br />

”Simulation studies <strong>of</strong> protein-induced bilayer deformations, <strong>and</strong> <strong>lipid</strong>-induced<br />

protein tilting, on a mesoscopic model for <strong>lipid</strong> <strong>bilayers</strong> <strong>with</strong> <strong>embedded</strong> proteins”


Acknowledgments<br />

This thesis would not have been possible <strong>with</strong>out the help <strong>and</strong> support <strong>of</strong> a great<br />

number <strong>of</strong> people. First <strong>of</strong> all, I would like to thank my supervisor, Berend Smit, for<br />

giving me the opportunity to start a PhD in his group, for his constant enthusiasm,<br />

encouragement <strong>and</strong> support, <strong>and</strong> for his unshakable trust in my work, which helped<br />

me a great deal to overcome my own doubts. I am thankful beyond words to Maria<br />

Maddalena Sperotto, for everything she taught me about membranes, for giving me<br />

energy <strong>and</strong> motivation for my research, <strong>and</strong> for her constant support <strong>and</strong> friendship.<br />

Without her this thesis would probably have never been completed. I also want<br />

to thank Marieke Kranenburg for the very fruitful collaboration <strong>and</strong> the stimulating<br />

discussions, <strong>and</strong> Thijs Vlugt for everything he taught me about simulations <strong>and</strong> for<br />

the nice time we had while sharing the same <strong>of</strong>fice. I am also extremely grateful to<br />

Gooitzen Zwanenburg for his precious help <strong>with</strong> computer related problems <strong>and</strong> for<br />

being always nice. I would like to thank Antoinette Killian, Els van der Brink van der<br />

Laan, Ben de Kruijff, <strong>and</strong> all the people in the Biochemistry <strong>of</strong> Membranes group<br />

at Utrecht University for introducing me to the world <strong>of</strong> “real” membranes <strong>and</strong> for<br />

making me feel always welcome whenever I visited their group. I am also grateful to<br />

Daan Frenkel for the rare, but extremely instructive, discussions. I wish to thank Edo<br />

Boek for his careful translation <strong>of</strong> the Dutch samenvatting, <strong>and</strong> Johan Padding for<br />

his help <strong>with</strong> the thesis cover. Thanks to Peter Coveney for letting me have the extra<br />

time I needed to finish the writing <strong>of</strong> this thesis. A sincere thank to CPe <strong>and</strong> Colin for<br />

their encouragement during the writing <strong>of</strong> this thesis. I also wish to thank the people<br />

that made my days at the University enjoyable, <strong>and</strong> that helped me in many different<br />

ways: Jan Pierre, Titus, Vincent, Ranieri, Tim, Nicole, Renate, Maureen, Fred. A<br />

special thank also to Lucia Minozzi <strong>and</strong> Elio Cecchetto for being my paranimfen, <strong>and</strong><br />

friends. Finally, words are not enough to thank the people that, <strong>with</strong> their friendship,<br />

made the four years I spent in Amsterdam a very happy <strong>and</strong> special time for me: to<br />

Lucia <strong>and</strong> Arjan, for their invaluable friendship <strong>and</strong> their constant support, to Elio,<br />

Daniele, Guido, Barbara, Massimo, Marjolen, Marco, Jos, Fabrizio, Annamaria, Luca,<br />

Elly, Hans, Miki, Roberto, Frank, Roberta. This thesis is dedicated to all the people in<br />

my family, for their love <strong>and</strong> trust, <strong>and</strong> for everything, grazie.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!