energetics of cholesterol-modulated membrane...

40
UPTEC X 11 003 Examensarbete 30 hp Januari 2011 Energetics of cholesterol-modulated membrane permeabilities. A simulation study Christian Wennberg

Upload: others

Post on 19-Oct-2020

0 views

Category:

Documents


0 download

TRANSCRIPT

  • UPTEC X 11 003

    Examensarbete 30 hpJanuari 2011

    Energetics of cholesterol-modulated membrane permeabilities. A simulation study

    Christian Wennberg

  •  

  • Molecular Biotechnology Programme

    Uppsala University School of Engineering

    UPTEC X 11 003

    Date of issue 2011-02

    Author

    Christian Wennberg

    Title (English)

    Energetics of cholesterol-modulated membrane permeabilities. A simulation study

    Title (Swedish)

    Abstract Molecular dynamics simulations were used to study the permeation of four different solutes through different cholesterol containing lipid bilayers. In all bilayers the limiting permeation barrier shifted towards the hydrophobic core, as the cholesterol concentration was increased. Cholesterols reducing effect on the permeation rate was observed, but under certain conditions results indicating an increased permeation rate with increasing cholesterol concentration were also obtained.

    Keywords Molecular dynamics, permeability, cholesterol, lipid bilayer, umbrella simulation, potential of mean force

    Supervisors

    David van der Spoel Uppsala University

    Scientific reviewer

    Johan Åqvist Uppsala University

    Project name

    Sponsors

    Language

    English

    Security

    ISSN 1401-2138

    Classification

    Supplementary bibliographical information Pages

    34

    Biology Education Centre Biomedical Center Husargatan 3 Uppsala Box 592 S-75124 Uppsala Tel +46 (0)18 4710000 Fax +46 (0)18 471 4687

  •  

  • Energetics of cholesterol-modulated membrane permeabilities. A simulation study

    Christian Wennberg

    Populärvetenskaplig sammanfattning

    Skiftande förhållanden i en cells omgivning gör att den måste ha ett försvar som skyddar den från skadliga ämnen. En viktig del i detta försvar är cellens cellmembran som bland annat verkar som ett skyddande skal kring cellen. Cellmembranet skapar en barriär som skadliga ämnen i cellens omgivning måste ta sig igenom, och därmed får cellen ett naturligt skydd mot dessa ämnen. Dock finns en dualitet i detta eftersom de ämnen cellen behöver för att överleva också finns i dess omgivning vilket innebär att cellen måste kunna reglera vilka molekyler som tar sig in och ut. När det gäller stora och/eller laddade molekyler sköts denna reglering av membranbundna transportproteiner som ser till att transportera dessa molekyler genom membranet. Små oladdade molekyler kan dock ta sig genom membranet utan hjälp av transportproteiner. Reglering av denna passiva genomsläpplighet är möjligt för cellen genom inkorporering av kolesterol i membranet. Kolesterol gör membranet stelare vilket försvårar för ämnen att ta sig igenom det. De exakta mekanismerna med vilket detta sker är dock inte fullt kartlagda än, men den stora ökning i datorkraft som skett senaste decenniet har gjort närmare analys av interaktionerna i cellmembran möjliga. I detta arbete har kolesterols effekt på fyra olika membran simulerats. Fyra olika ämnen, utvalda för deras olika egenskaper, har via dator-simuleringar studerats vid sin transport genom de olika membranen. Kolesterolkoncentrationen i membranen har varierats från 0 till 50 % och från simuleringarna har energibarriärer för transporten genom membranen beräknats.

    Examensarbete 30hp Civilingenjörsprogrammet Molekylär bioteknik

    Uppsala universitet februari 2011

  •  

  • 1

    Contents: Page

    1. Introduction

    1.1 The cell membrane 2 1.2 Phospholipids 2

    1.3 Cholesterol 3 1.4 Membrane permeation 4

    1.5 Aim 5

    2. Molecular dynamics 2.1 Basic principles 6

    2.2 Force fields 7 2.3 Algorithms 8

    2.4 Potential of mean force 10 2.5 Umbrella sampling 10

    2.6 Weighted histogram analysis method 11

    3. Methodology 3.1 Lipids and solutes 12

    3.2 Simulation procedure 13

    4. Results 4.1 Lipid patches 15

    4.2 PMFs 15 4.3 Umbrella simulations 15

    5. Discussion

    5.1 Tail region 17 5.2 Head region 18

    5.3 Solutes 19

    6. Conclusion 19

    7. Acknowledgements 20

    8. References 21

    Appendix A 23 Appendix B 27

    Appendix C 31

  • 2

    1. Introduction 1.1 The cell membrane

    A cell’s membrane defines the outer limit of the cell and protects it from the surrounding

    environment by providing a physical barrier for molecules trying to enter or exit the cell’s interior. This allows the cell to maintain certain properties such as pH or ion concentration in its cytoplasm

    despite environmental changes. The membrane also serves as an anchor point for the cells cytoskeleton, thereby also aiding in the maintenance of the cells shape.

    Incorporated into the cell membrane are also membrane bound proteins. These proteins manage a vast number of functions, such as in- and outflow of solutes from the cell’s interior, cell

    communication and production of molecules essential for the cell (e.g. ATP). The cell membrane’s basic structure is built up by phospholipids, which gives the membrane its

    characteristic properties with a hydrophobic core separated from the interior and exterior environment by its hydrophilic surfaces.

    Other lipids existing in the membrane are cholesterol and glycolipids, that aid in the membrane stability (cholesterol) or serve as cell markers (glycolipids).

    1.2 Phospholipids As the major component in cell membranes phospholipids play a crucial role in

    protecting the cells from the outside environment. Phospholipids are built up by a hydrophilic head region which, depending on the type of lipid, contains a

    different number of charged groups of which one always is a phosphate group. The head group is attached to a tail region that consists of long fatty chains,

    making them very hydrophobic (figure 1), and it is this amphipathic character of the phospholipids that enables them to form bilayers in water, where the

    hydrophobic tails group together on the inside of the layer with the head groups outwards forming the boundary to the surrounding water.

    Looking at the various lipids found in cell membranes all but one seems to have glycerol as the “backbone” molecule. The one left out is sphingomyelin,

    which instead is built up from the amino alcohol spingosine. The composition of lipids in the cell membrane differs between the outer and

    inner leaflets. The outer leaflet tends to be populated by phosphatidyl-choline (PC) and spingomyelin while the inner leaflet's major components are

    phosphatidyl-ethanolamine (PE) and phosphatidyl-serine1. The reason for this

    contrast in composition is still unclear, but it has been found that changes in

    this asymmetry are a source for different diseases. Since cells invest energy in trying to maintain this difference in composition between the leaflets the

    subject has a high physiological relevance2.

    Figure 1. Lipid molecule.

    Carbons shown in grey,

    Phosphate, oxygens and

    nitrogen are coloured

    according to atom type.

    Hydrogens attached to carbons not shown.

  • 3

    1.3 Cholesterol Looking further into the composition of lipid membranes, cholesterol is an

    important structural molecule, as it regulates membrane fluidity and thereby also the permeability of solutes across the membranes

    3.

    Cholesterol consists of four rings fused together, a hydrocarbon tail in one end and a hydroxyl group bound at the other. This gives cholesterol an amphiphilic

    character that enables it to position itself in the membrane with the hydroxyl group interacting with the head region and the hydrophobic ring structure

    interacting with the tails of the lipid molecules4 (figure 2). Another important

    feature of the cholesterol molecule is the fact that the two methyl groups bound

    to the ring structure are located on the same side, thereby creating two distinct sides of the molecule denoted the �- and �-face (methyl groups bound to the �-

    face). Molecular dynamics (MD) - simulations have shown that the lipid tails prefer

    interaction with the smooth �-face but interestingly a removal of the methyl groups from the �-face will not increase the effect of cholesterol on the

    membranes but rather decrease it, indicating that the structure of cholesterol has been shaped by evolution to interact in an optimal way with the lipids in

    membranes5. Experiments have also shown that cholesterol prefers interaction

    with saturated lipids compared to unsaturated due to the more favorable van der

    Waals interactions obtained with the saturated tails4,6

    . Increasing the cholesterol concentration in the membrane is known to alter the

    behavior of the lipid tails by inducing a phase transition. A pure lipid bilayer can exist in either the gel-phase (L�) where the tails are elongated as much as possible, or in the

    more fluid liquid-crystalline (L�) phase which exists at higher temperatures. Cholesterol reduces the transition temperature of the L� to L� phase transition. At sufficiently high cholesterol

    concentrations the bilayer enters a phase that resembles both the L� and the L� phase in that it is rigid as the L� phase but it retains fluidity from the L� phase. This state is commonly referred to as

    the liquid ordered (Lo) phase, and the L� phase is usually called the liquid disordered phase7.

    From MD simulations this ordering effect caused by cholesterol can be quantified in a variety of

    ways. Some examples are the decreased amount of gauche rotamers in the lipid tails or the average tilt angle of the chains relative to the membrane normal

    8, and visualization of the bilayers

    is an easy way to see the difference between the different phases (figure 3). The rigidifying effect of cholesterol has also been seen through its ability to greatly reduce the

    amounts of lipids that flip-flop between the two leaflets in bilayers9. This indicates that cholesterol

    could have a role in maintaining the asymmetry between leaflets, found in cell membranes.

    Figure 2. Lipid with

    cholesterol.

    �-face of cholesterol pointing towards reader.

    Figure 3. Snapshots from two different

    simulations.

    Left: Lipid bilayer in L� phase.

    Right: Lipid bilayer with 40% cholesterol

    (cholesterol removed from image for clearer

    view) in the Lo phase

  • 4

    1.4 Membrane permeation In order for a cell to maintain functionality, various solutes have to be transported across the

    membrane to the cell’s interior. This transport is mediated by either membrane bound transport proteins, passive permeation through the membrane by the solutes themselves or by pore

    formation in the membrane. Of these mechanisms the transport by proteins governs mostly ions and other charged molecules, whereas small uncharged molecules can permeate the membrane by

    themselves at a reliable pace. Overton

    10 and Meyer

    11 described this passive permeability through lipid membranes for different

    solutes at the end of the 19th century. They stated that the permeability of a solute through a membrane is directly proportional to the solutes’ solubility in the membrane (or oil/water partition

    coefficient). From this, the so-called solubility-diffusion model, in which the solutes first partition from water into the membrane and then diffuse through it, was developed

    12. It can be expressed by

    a simple equation:

    � ���

    Here P is the permeability coefficient, K is the partition coefficient of the solute between the membrane and water, D is the diffusion constant in the membrane and d is the thickness of the membrane. Despite its simplicity the solubility-diffusion model correctly describes the behavior of

    many solutes, but exceptions have been found and stirred up a debate whether they violate the model or not

    13. The solubility-diffusion model does however over-simplify the relatively complex

    system that is a lipid bilayer and therefore more advanced models have been constructed. Examples of this are the 4-region model by Marrink and Berendsen

    14, which divides the

    membrane into four regions with different characteristics. And also the model constructed by Nagle et al, which divides the membrane into three layers and then computes the total

    permeability by combining the permeation through each of those layers15

    . The first MD simulations of lipid bilayers was performed by Kox et al.

    16, who managed to

    simulate a phase transition in a lipid membrane despite using a very simplified system in which water was excluded. Later on, simulations with water included started to emerge

    17, and also solute

    diffusion in the membranes were studied using MD18

    . These studies revealed that the diffusion constant of solutes in the membrane was not constant but rather depended upon the position of the

    solute within the membrane. This difference in diffusion was proposed to exist because the diffusion process was not linear. Instead the solutes “jump” between different pockets of free

    volume within the membrane, which occurs due to kinks in the lipid tails. In following studies it was observed that the amount of kinks, and thereby also the amount of free volume, was reduced

    as cholesterol was incorporated into the membrane21

    . This was proposed as a reason for the lowered permeation of solutes through membranes containing cholesterol.

    During recent years more comprehensive studies containing many different lipid bilayer and solutes have been done. Mathai et al.

    19 did experimental studies on seven different lipid

    membranes and found that water permeability was strongly correlated with the surface area occupied by the lipids and poorly correlated with the thickness of the membrane, which later on

    resulted in the 3-layer model by Nagle mentioned earlier. Bemporad et al.20

    did all-atom simulations on eight small molecules from which they discovered that the size of the solute had

    less impact on diffusion through the membrane than was previously thought. The partitioning of the solute into the membrane was however proposed to be more size dependent than previously

    thought.

  • 5

    1.5 Aim In this work a comprehensive study has been done on the effect of cholesterol on solute

    permeation through lipid membranes. The permeation of four different solutes through four different lipid membranes at five different cholesterol concentrations has been studied. The aim of

    the study was to investigate how certain structural properties in the lipids would change the influence of cholesterol on the permeation event, and to see if the different characteristics of the

    solutes had a notable impact on their permeation. These studies were enabled through calculations of the potentials of mean force (PMF) for the permeation events.

  • 6

    2. Molecular dynamics 2.1 Basic principles Molecular dynamics (MD) is a computational method that simulates a system of choice by

    integration of Newton’s law of motion. The reliability of MD to correctly recreate the real systems that it tries to simulate lies in the accuracy of the three main approximations MD is based on:

    • Born-Oppenheimer approximation The Born-Oppenheimer approximation separates the motions of the electrons and nuclei of

    an atom and assumes that the motion of the electrons are so much faster than the nuclei that they instantaneously adjust themselves to the motion of the nuclei. Therefore only the

    motion of the nuclei is accounted for in the MD simulations.

    • Classical treatment Solving the Schrödinger equation for all the nuclei in the systems one desires to study with MD is not possible. Therefore it is assumed that the nuclei can be treated as classical

    particles that behave according to Newton’s second law.

    • Energy surface Exactly as in the case with the nuclei, solving the Schrödinger equation for all the electrons

    in the system in order to obtain the potential energy surface for the molecules is impossible for large (>100 atoms) systems. Therefore the interactions in the system are approximated

    by empirical models (force fields) designed to mimic the real energy surface as closely as possible.

    Using these three approximations MD converts an extremely complex system to a relatively

    simple one, basically representing all particles and molecules with a ball-and-stick model. Starting from Newton’s second law and the simple relation between the particles position and their

    velocities, the core equations of MD can be set up:

    �� � ���

    ���

    � ������

    ��� (1)

    �� ����

    �� (2)

    Here U is the potential energy, F the force acting on particle i, m the mass of particle i and vi the particles velocity. From these equations the positions, accelerations and velocities of all the

    particles in the system can be followed as the system evolves in time. This is done step-by-step following the methodology outlined below:

    1. Calculate forces on each particle based on current positions

    2. Calculate new velocities 3. Calculate new positions by evolving the system in time

    4. Repeat from 1

    This scheme is then repeated until the desired simulation time has been reached, but in order to get useful output a certain amount of criteria has to be met. First of all the potential function, commonly denoted “force field”, has to reproduce the “real” system in a fairly accurate way in

    order to get the correct forces acting on all particles. Then one has to choose a time step, used to calculate the new positions, that is long enough to get a time evolution that makes it possible to

  • 7

    study the wanted properties of the system, but short enough to account for the fastest motion in the

    system in order to avoid numerical and integration errors due to collisions that should not occur throughout the simulation. The “good choice” of a time step is hard to decide but a reasonable

    choice when using flexible molecules is one order of magnitude lower than the fastest motion in the system

    23. This can severely limit the simulation speed in respect to time evolution, and if this

    is the case one can sometimes constrain the fastest motions in order to increase the time step (the fastest motions tends to be bond vibrations which are not of great interest for the properties one

    would like to extract in most cases). One large problem when doing simulations on systems of finite size is boundary effects created at

    the boundaries of the simulation box. This is solved in MD by the use of periodic boundary conditions (PBC), which removes the boundaries in a simulation of a system containing a small

    amount of particles compared to what would be required in order to remove boundary effects if PBC was not used. The way PBC works is basically by replicating the simulation box in all

    directions so that if a molecule exits the box through one of the sides, an identical particle enters on the opposite side. This obviously also puts some limitations on the type of simulation box we

    can use since in order for PBC to work it has to be formed so that it can be replicated in three dimensions without leaving any gaps between the boxes (A sphere is an example of a shape that

    can not be used with PBC).

    2.2 Force fields In order to calculate the forces acting on each particle in a system a force field describing the interactions between particles is required. The force fields utilized in MD is often constructed in

    the same way with different expressions for the energies governing bonded and non-bonded interactions:

    � � � ������� � ����������

    Bonded interactions are composed by three parts describing bonds, angles and dihedrals.

    ������� � ����� � ����

    �����

    � ����� � ����

    ������

    � ��

    ���������

    ��� ��������� � ����

    Here the bond stretching potential is described by the simple harmonic potential:

    � �� � ����� � ����

    But it could also be expressed by the more accurate Morse potential:

    � �� � �� �� ���������� � �����

    This potential, however, is computationally more expensive, since it involves an exponential and also requires more parameters for each bond (dissociation energy De and also the constant a). The

    harmonic potential also agrees well with the Morse potential at short distances from the equilibrium value.

    The dihedral term describes the rotation around bonds, where mi gives an indication of the barrier height for the rotation around the bond, �i is the torsion angle, n is the number of minima when

    rotating the bond 360 degrees and �0 describes the minimum values for the torsion angle (other expressions using multiple cos-terms can also be used).

    The non-bonded part can be divided into Lenard-Jones (LJ) and electrostatic interactions

  • 8

    ������������ � �������

    ���

    ��

    ����

    ���

    �����

    ������������

    Where � is the collision parameter (zero energy) of two particles, � is the interaction energy in the LJ-potential, and the electrostatic interaction is just the Coulomb energy of two point charges

    interacting with each other. Non-bonded interactions only account for pairwise interactions and do not incorporate many-

    body effects. Many-body potentials would increase the computational cost drastically and instead a majority of the many-body effects are incorporated effectively into the pairwise potentials,

    which are often called “effective pair potentials”.

    Calculation of the non-bonded energies takes significantly more computer power than the bonded energies since the non-bonded interactions are calculated between all the particles in the system

    and thereby scale as N2, whereas the bonded interactions grow proportionally to N. In order to

    reduce the cost of non-bonded interactions a cut-off is often utilized which sets all interactions

    with particles outside of this cut-off with particle i to zero. This works well for the LJ interactions since they decrease rapidly with distance, but using a cut-off for electrostatic interactions requires

    care since these decrease very slowly and often other methods are used to deal with these. The one often used in MD is particle mesh Evald (PME), which divides the electrostatic interaction

    potential into a short-ranged term and a long-ranged term. The short-ranged term is quick to calculate in real space, but this is not the case for the long-range part. So this is solved by

    calculating the long range part in the Fourier space instead, which makes the computational cost scale as O(NlogN) instead of O(N

    2) as in the real space calculations.

    2.3 Algorithms Analytical solutions to the equations obtained in (1) and (2) cannot be found for the large systems that are used in MD, due to the complexity of the potential energy function U. Instead numerical

    solution methods have to be utilized, namely finite difference methods. Of these the two most common ones are the Verlet and Leapfrog algorithms.

    Verlet The Verlet-algorithm used in molecular dynamics is constructed from a third order Taylor

    expansion of the position r(t) at time t+�t and also for t-�t. This results in the following expressions:

    � �� �� � � � ������

    �����

    ������

    �����

    ��

    ������

    �����

    �� ����

    �� (3)

    � �� �� � � � ������

    �����

    ������

    �����

    ��

    ������

    �����

    �� ����

    �� (4)

    From this one notices that������

    �� and

    ������

    ��� is the velocity and acceleration of a particle at time t.

    The third derivative,�������

    ���, is eliminated by adding the two equations and thereby obtaining:

    � � � �� � �� � � � � � �� � � � ��� � � ��� (5)

  • 9

    Where a(t) denotes the acceleration at time t. The acceleration can be obtained from the potential energy using equation (1) The error associated with truncating the Taylor expansion after the third term is of the order �t

    4

    making the Verlet algorithm a fairly accurate method. One problem however is that the velocities are not obtained directly from the equations, but they are needed to calculate important features of

    the system such as the kinetic energy. An easy way to obtain them would be to truncate after the second term and subtract (4) from (3):

    � � �� ���� ��������

    ���� ����

    �� (6)

    But this gives velocities at time t, as opposed to the positions, which are obtained at time t+�t.

    And another notable thing is also that the error is proportional to �t2 instead of �t4.

    In order to get velocities and positions at the same time one could instead use the so called velocity Verlet algorithm which is constructed by a second order Taylor expansion of the position

    r(t):

    � � � �� � � � � � � �� ��

    �� � ��

    �� ����

    �� (7)

    And also a second order expansion of the velocity:

    � �� �� � � � � � � ����

    ������

    �����

    �� ����

    �� (8)

    Using the forward difference approximation ������

    ����

    � ���� �����

    �� one obtains:

    � �� �� � � � ��

    �� � � ���� ��� ��� ����

    �� (9)

    This enables us to calculate the velocities and positions of the particles at time t+�t using the

    values from time t. The acceleration at time t is calculated using equation (1).

    Leapfrog algorithm The other algorithm often used is the leapfrog algorithm, in which the velocities and position “leap” over each other as they are calculated at different time steps:

    � �� �� � � � � �����

    ������ (10)

    � ���

    ��� � � ��

    ��� � ������ (11)

    The leapfrog algorithm is equivalent to the Verlet algorithm, something that is realized by looking at r(t) evaluated in a previous step:

    � � � � �� �� � � ���

    ��� �� (12)

  • 10

    Arranging (12) and inserting into (11) gives:

    � ���

    ��� �

    ��� � � ���� ��� � ������ (13)

    And then (13) and (10) gives:

    � � � �� � � � � � � � � � � �� � � � ���� ������ � � � �� � ������

    � (14) Which is identical to (5).

    The advantage of the leapfrog method over the Verlet is that the algorithm explicitly calculates the velocities, but the drawback is that they are calculated at a different time.

    2.4 Potential of mean force A key concept in computational studies of biological systems is the potential of mean force

    (PMF), G(r), which describes the average force acting upon a particle along a certain reaction coordinate (e.g. rotation around bond, binding events). It is defined as:

    � � � � ��� ��� ��

    ����

    �����

    Where �(r) is the average distribution function along the coordinate and G(r*) and �(r*) are arbitrary constants. It can be shown that the PMF is in fact a free energy profile along the reaction

    coordinate and therefore is closely related to the distribution function along this coordinate24

    . Extracting the distribution function from simulations is however not trivial and the method used in

    this thesis will be described below. 2.5 Umbrella sampling First introduced by Torrie and Valleau in 1976

    25, umbrella sampling is

    a technique that is used to sample a reaction coordinate of a system, which has an energy landscape that makes it hard to use conventional

    MD-sampling for this purpose. This is illustrated in figure 4 where the topmost system will probably only yield configurations either to the

    left or right of the energy barrier since no barrier crossing will occur, but the bottom one is easily sampled by conventional MD.

    The way umbrella sampling solves this problem is by introducing a new potential:

    ��� � � � ����� (15)

    Here W(r) is the introduced potential, and it is designed to bias the simulation around some region in configurational space and thereby

    generate a distribution that is non-Boltzmann. Using figure 4 as an example the potential W(r) is designed in such a way that it reduces (or removes) the barrier in the middle, and thereby enables full sampling of the configurational space.

    Figure 4. Umbrella sampling.

    Top: Sampling of the configurational

    landscape along a reaction coordinate

    Q in the presence of an energy barrier

    Bottom: Umbrella sampling of the

    same area will sample configurations over all space

  • 11

    This means that the system sampled is not the real system, but despite this, Torrie and Valleau showed that it is possible to retrieve the “true” value of the sought thermodynamic property (here

    denoted as S) through the formula:

    � �

    �� �

    �� �

    (16)

    Where ��indicates that the average is taken over the

    distribution created by (15).

    2.6 Weighted histogram analysis method (WHAM) A more specific use of the method of umbrella sampling is when utilizing a harmonic biasing potential:

    � � � ���� ���� (17)

    Which will result in a relatively local sampling of the

    configurational space around the coordinate r0. By doing many simulations with different values of r0 a large set of simulations are obtained, each one centered on a unique r0-value. Recombination of these simulations (or “windows” as they are

    called) can then be done in order to calculate the desired properties of the system (illustrative example can be seen in

    figure 5). A more thorough description of the way of combining the

    windows can be read in reference 26, but the final results are these two equations:

    ���� � �� ���� �� ���������� � � ��

    ����

    ���

    ��

    ���(18)

    ��� ���

    ���� ���������

    �����

    ���� ���� (19)

    Where �(r) is the “real” distribution function, N the number of umbrella simulations, ni and nj the number of independent data points used to construct the biased distributions, ���� � is the ith

    biased distribution function and F represents the free energy associated with the introduction of the potential W(r). As can be seen, equation (18) needs the constants Fj obtained from equation (19) in order to calculate the unbiased distribution function. But equation (19) demands the distribution function

    from (18) in order to calculate them, so an initial guess of the constants F is needed and then the equations are solved self-consistently until the changes in F and p(r) are sufficiently small.

    Figure 5. Schematic overview of the

    WHAM-method.

    Top: The unknown potential G(�) along the

    coordinate �. Middle: Umbrella simulations centered on different �-values will create a set of

    distributions each one influenced by the

    local shape of the potential G(�).

    Bottom: Recombination of the different

    umbrella windows will yield an estimate of

    the unknown potential G(�).

    Picture taken and modified, with permission, from reference 24

  • 12

    3. Methodology

    3.1 Lipids and solutes The lipids chosen for the project were:

    • 1-Palmitoyl-2-oleoyl-sn-glycero-3-phosphoethanolamine (POPE)

    • 1-Palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine (POPC)

    • 1,2-Dimyristoyl-sn-glycero-3-phosphocholine (DMPC)

    • 1,2-Dipalmitoyl-sn-glycero-3-phosphocholine (DPPC)

    POPE and POPC have the same structural tail region with one unsaturated and one saturated tail. The head region of POPC is however larger than in POPE due to the three methyl groups which

    have replaced the hydrogens in the amine-group of POPE. DPPC and DMPC have the same head region as POPC but contain two saturated tails. The

    lengths of the tails in DPPC are 16 carbons and those in DMPC are 14 carbons. The different structures of the four lipids enabled studies of how cholesterol would influence the

    permeation across lipid membranes with structural differences:

    • Head groups: small: POPE large: POPC, DPPC, DMPC

    • Tails: saturated: DPPC, DMPC unsaturated: POPE, POPC

    long: DPPC short: DMPC

  • 13

    Solutes were chosen to be nitric oxide (NO), propane, benzene and neopentane in order to study

    the cholesterol effect on hydrophobic (propane, benzene, neopentane) and amphiphilic (NO) solutes. Earlier data from hydrophilic solutes (ammonia in this case) was already available, and is

    included in the analysis. The three different hydrophobic solutes also enabled studies of how bulkiness would effect the permeation across the membrane.

    3.2 Simulation procedure

    Gromacs The simulations performed in this work were done using the Gromacs

    31 simulation package.

    Gromacs is a MD-software designed for biochemical simulations and it is also the fastest software

    available since it handles the non-bonded interactions extremely well.

    Membranes The membranes were built by first putting lipids at evenly distributed positions in a quadratic grid, and then copying and inverting this grid in order to get a bilayer. For the membranes containing

    cholesterol, lipids at random positions in the grid were exchanged for cholesterol molecules in order to obtain four different cholesterol concentrations for each membrane type in addition to the

    pure membranes. Table 1 shows a schematic over the composition of each system. The OPLS (Optimized Potentials for Liquid Simulations) united atom force field

    32 and TIP4P

    33

    model was used for the solutes and water molecules. All lipids except DPPC used the Berger force field

    34. For DPPC the parameters from Ulmschneider et al.

    35 was used instead. The time step

    utilized was 2 fs. Bonds were constrained using the LINCS36

    (Linear Constraint Solver) algorithm. Electrostatic interactions were calculated by particle-mesh Ewald and dispersion interactions were

    described by a Lennard-Jones potential with a 1 nm cut-off. Three-dimensional periodic boundary conditions were used in all simulations.

    Equilibration The membranes were solvated in a box of water and equilibrated for 200 ns (20-40 ns for the pure lipid membranes). The temperature was kept at 300 K (323 K for DPPC in order to avoid a phase

    transition to the gel-phase) using velocity rescale, and pressure was kept at 1 bar by the Berendsen barostat.

    Table 1. Simulated systems. Composition of each bilayer at each concentration of cholesterol.

    X/Y notation shows the number of lipids (X) and the number of cholesterol molecules (Y) in each system. Percentages is based on number of molecules.

    Nitric oxide Propane Benzene Neopentane

  • 14

    After the equilibration, random snapshots where extracted from the last 20 ns in the trajectory

    (last 10 ns for the pure membranes). These were then used as starting structures in the umbrella simulations.

    Umbrella simulations The umbrella simulations were done with a spacing of 0.25 Å

    between umbrella windows using a harmonic umbrella potential with a force constant of 1000 kJ/mol/nm

    2. Temperature was

    kept at 300 K by a nose-hoover thermostat, and the pressure was set to 1 bar by a Parrinello-Rahman barostat.

    In order to increase the sampling, four solute molecules were constrained in each step, at a distance of one half of the

    simulation box away from each other. This ensured that they did not interact with each other. Each simulation also contained 4-5

    “slices” with different umbrella windows in order to reduce the total simulation time (example in figure 6). These slices where

    put 15 Å apart for the small solutes (propane, NO) and 20 Å apart for the larger ones (benzene, neopentane).

    The constrained atoms in the umbrella simulations for each

    solute was as follows:

    • NO: A dummy atom centered between the oxygen and nitrogen atom

    • Propane: The central carbon atom

    • Benzene: A dummy atom placed at the center of the benzene ring

    • Neopentane: The central carbon atom The small solutes were inserted into the membrane at their respective z-positions making sure that

    they are not placed within 1.2 Å of a lipid molecule, in order to avoid bad interactions. The larger solutes could not be inserted in this way so instead they were placed within the membrane using a

    smaller tolerance radius (0.3 Å). Thereafter, a short free energy perturbation (FEP) -simulation was done in order to “grow” the solute into the membrane.

    Equilibration was done using a conjugated gradient energy minimization, followed by a 1 ns MD simulation.

    PMFs The first 200 ps of the umbrella simulations were removed for equilibration purposes and then the

    PMFs were calculated by the WHAM-method using Gromacs inbuilt g_wham-tool37

    .

    Figure 6. Snapshot from umbrella simulation.

    Pure POPE bilayer simulated with propane

    (black) in 5 different umbrella windows (see

    WHAM-section), water molecules are colored blue.

  • 15

    4. Results 4.1 Lipid patches The final equilibration of the lipid patches were verified to be converged through the stabilized

    total energy plotted against time. All these plots can be seen in appendix A (DPPC with 30% cholesterol and DMPC with 40% cholesterol shown here in figure 7 & 8). Almost all patches seem

    to have converged at the end of the equilibration. The exception to this is DPPC with 30% cholesterol, which still shows a downward slope in the energy plot indicating a slow convergence

    and that a longer equilibration time is needed. The data from these simulations will be presented in the plots despite this, so the reader is asked to remember that the 30% cholesterol data for

    DPPC-membranes could look different if a longer equilibration is done.

    4.2 PMFs The PMFs computed for each respective bilayer were put together in plots as seen in

    figure 9 (which shows propane in POPE). A plot contains all the PMFs for one solute at

    each cholesterol concentration in one bilayer. The center of the bilayer is at zero and the

    two distinct regions on each side including the barriers represent the head group region

    (at around +/- 2 on the z-axis) and the tail region (further towards the center).

    4.3 Umbrella simulations The PMFs calculated from the umbrella

    simulations can be found in appendix B, and in order to do a more qualitative analysis the output were divided into data from the head

    Figure 8. Equilibration of DMPC.

    Plot of total energy from equilibration of DMPC with 40% cholesterol

    Figure 7. Equilibration of DPPC.

    Plot of total energy from equilibration of DPPC with 30 %

    cholesterol

    Figure 9. Permeation of propane through POPE.

    PMFs for the permeation of propane through POPE. Definition of head- and tail-

    region highlighted by boxes. Vertical line shows how values for tail-region-peaks at lower concentrations were approximated for the plots in appendix C.

  • 16

    group and tail regions of the lipids. From each of these regions, the highest point of the barrier was

    plotted against the cholesterol concentration and these plots can be found in appendix C. As can be seen in figure 9 the tail region at the lower cholesterol concentrations does not have a

    clear peak, so in these cases the “highest point” was chosen at the plateau-like region or in the cases where this was not possible the point were chosen by just drawing a vertical line as seen in

    figure 9. This of course introduces an uncertainty in the results but since the data analysis is more focused on the profile of the plots rather than the absolute values of each point this should not

    impact the end result drastically. In order to get an idea of how the overall permeability is influenced by the cholesterol presence

    the highest point of the PMF, the limiting barrier in the permeation, was plotted against cholesterol concentration for all the solutes in all the membranes (figure 10).

    Figure 10. Permeation barriers in all membranes.

    Plots showing the highest peak in the PMFs plotted against the cholesterol concentration for each solute, in each membrane. Errorbars not shown

    in order to get an easy overview.

  • 17

    5. Discussion 5.1 Tail region Comparison of the different results from the POPC,

    DMPC and DPPC simulations (figure 10) shows that the main difference is how the barrier height

    changes with increasing cholesterol content. In the patches with saturated tails (DMPC and DPPC)

    there seems to be a fairly constant increase in barrier height with increasing cholesterol concentration, whereas in the unsaturated POPC tail decrease in

    barrier height is seen at a cholesterol concentration of 30%. From the plots in appendix C (as an example plots for propane is shown in figure 11) it can be

    seen that it is after this cholesterol concentration that the “tail-barrier” overcomes the barrier in the head group region and becomes the rate-limiting barrier that the solute has to cross. In DMPC this

    barrier continues to increase as one adds more cholesterol, but in POPC the increase in barrier height is slower when increasing the cholesterol concentration above 30%. Comparing this with

    phase transition diagrams for DMPC27

    and POPC28

    the explanation for the faster increase in DMPC could be a phase transition between the L� and the Lo phase of DMPC that occurs

    approximately around 30% cholesterol, whereas the transition in POPC occurs at 50 % cholesterol. A phase transition to the more rigid Lo would mean that the solute has less free

    volume available during the permeation and thereby we see this continuing increase in the barrier in DMPC compared to POPC. In DPPC, which undergo a phase transition at approximately 20 %

    cholesterol29

    , the effect of this phase transition can be seen clearly in figure 10 when one looks at the permeation barrier for ammonia. The unsaturation of one of the tails in POPC also means that

    the kink in this tail introduces more free volume for the solutes to use during their diffusion in the membrane.

    Figure 11. Head- and tail-barriers for propane

    Barrier height at the head- and tail-region plotted against cholesterol content

    in the membrane.

    Solute: propane

  • 18

    5.2 Head region Comparison between POPE and POPC showed no large difference in the barrier at the tail region

    (as one should expect since their tail-regions are identical) but a striking difference is the continuous reduction of the barrier in the head region of POPE. The same region in the bilayers

    containing PC-head groups showed an increase in the barrier until ~30% cholesterol where it starts to decrease. The larger PC head groups of POPC allow more space between the tails for

    cholesterol to position itself within. POPE on the other hand has a rather small head group and therefore the space between the tails is more constricted. An analogy would be to see POPC as a

    football with two ropes attached, as tails, and POPE would then be the same construction but with a handball instead.

    Inserting cholesterol into the POPE bilayer would then mean that the head groups would be pushed apart from each other as more and more cholesterol are added. In POPC the cholesterol

    has additional space between the head groups, that it can populate before “saturation” is achieved and the head groups are pushed apart. This increase in the barrier at the head region, which is seen

    in the PC head groups, is due to the condensing effect that cholesterol has on the membrane. This effect can be seen in figure 12, which shows the area per lipid plotted against cholesterol

    concentration in the membrane. The area per lipid for a pure membrane is calculated by dividing the total area of the bilayer with the number of lipids in one leaflet. When cholesterol is present

    this is however not as simple, since the area occupied by cholesterol has to be accounted for. This is solved by subtracting the area occupied by cholesterol from the total area of the membrane. The

    average area of cholesterol in bilayers does however vary a lot between different studies17

    , but the values calculated for cholesterol in bilayers vary between 22-29 Å

    2, which does not change the

    main features of the plot. So for this analysis the average value of 27 Å2 calculated by Hofsäß et

    al.30

    was used. To see the different behavior of the PC head group compared to the PE head group

    figure 12 also contains the area per lipid without correction for the cholesterol in the membrane in order to see the swelling of the membrane as cholesterol is added. POPE starts to swell directly as

    cholesterol is added whereas the PC-containing lipids show an increased “resistance” towards swelling as cholesterol is incorporated.

    Figure 12. Area per lipid in all membranes. Left: Area per lipid for all membranes with correction for the area occupied by cholesterol. Right: Area per lipid

    without correction for cholesterol.

  • 19

    5.3 Solutes Comparing the different PMFs for all the solutes (appendix A) the largest difference between the

    hydrophilic ammonia and the hydrophobic solutes was the single barrier over each leaflet for ammonia, whereas the hydrophobic solutes starts out with a barrier over the head group region and

    then, at higher cholesterol concentrations, actually have the rate limiting barrier located in the hydrophobic tail region. Ammonia has the largest barrier at all cholesterol concentrations in all the

    membranes except in pure POPE where the head group barrier for neopentane is even higher (figure 10). This is quite interesting since it shows that the hydrophilic region at the lipid-water

    interface can actually be as rate limiting for hydrophobic solutes as the hydrophobic interior is to hydrophilic solutes.

    Propane and benzene behaves roughly the same in all membranes except DMPC where the reduced free volume, due to the phase transition, at higher cholesterol concentrations seemed to

    influence the larger benzene molecule more than propane. This is however not the case in DPPC where they again behave roughly similar, with benzene having some large outliers.

    The permeation barriers for neopentane is however larger than those for benzene and propane in almost all simulations (one large outlier in POPC with 40% cholesterol), and the overall behavior

    for the different cholesterol concentrations is the same. This similarity between propane and benzene, but dissimilarity when comparing to neopentane is interesting. It indicates that benzene

    despite its more bulky structure still manages to permeate the membrane with relative ease compared to propane. Neopentane shows a similar permeation barrier in the pure PC-membranes,

    but with added cholesterol the barrier directly increases more than those of propane and benzene. The reason for this should be that neopentane, being even more bulky than benzene, is directly

    affected of the more condensed head group regions created by the addition of cholesterol, something that can be seen in the plots in appendix C. This is also illustrated by neopentane’s

    permeation barrier in pure POPE, which has a fairly condensed head group region in the pure membrane compared to the PC-membranes.

    The small amphipathic nitric oxide has an overall low permeability barrier, as could be expected since it should be able to cross both the head and tail region with relative ease. Comparing it to the

    hydrophobic solutes the only real difference is the permeation through pure POPE, were NO has a much smaller permeation barrier. This is due to the hydrophilic barrier created by POPE, which is

    larger than those of the PC-bilayers.

    6. Conclusion Four different solutes were used in simulations in four different bilayers at five different

    cholesterol concentrations. The effect of cholesterol on the permeation through the lipid bilayers was studied through the calculated PMFs of the permeation event. Incorporation of cholesterol

    was found to shift the rate-limiting barrier of permeation for the hydrophobic solutes; a barrier at the tail region replaced the hydrophilic barrier at the head groups and became the rate-limiting

    factor. In the three PC-bilayers, incorporation of cholesterol increased the permeation barrier for hydrophobic solutes by condensing the membranes and thereby reduce the amount of free volume

    available. In POPE, where the head groups are much smaller than in PC-bilayers, an opposite effect was observed as the permeation barrier was decreased upon addition of up to 30%

    cholesterol. This shows that the head group size is a deciding factor of how cholesterol will change the behavior of a lipid bilayer. It also demonstrates that cholesterol can have an increasing

    effect on the permeability of hydrophobic solutes across specific lipid bilayers, which is very interesting since research often is focused towards cholesterols ability to decrease permeability

    across membranes. Lipid bilayers using a single type of lipids can be utilized to study cell permeability to some

  • 20

    extent, but the membrane in a cell consists of a variety of lipids, distributed in an asymmetric way

    between the outer and inner leaflet. Future simulations focused on membranes with an asymmetric lipid composition are of interest, since the effect of cholesterol on those membranes could give

    some new insights to why this asymmetry is maintained by the cells.

    7. Acknowledgements I would like to thank Dr. Jochen Hub and Prof. David van der Spoel for their guidance and help throughout the whole project.

  • 21

    7. References

    (1) M. Bretscher, Assymetrical Lipid Bilayer Structure for Biological Membranes, Nature New Biology 1972 236:11-12

    (2) R. F.A. Zwaal, A. J. Schroit, Pathophysiologic Implications of Membrane Phospholipid Asymmetry in Blood Cells, Blood 1997 89:1121-1132

    (3) E. Ikonen, Cellular Cholesterol Trafficking and Compartmentalization, Nature reviews Mol. Cell Biol. 2008 9:125-138

    (4) H. Ohvo-Rekilä, B. Ramstedt, P. Leppimäki, J. P. Slotte, Cholesterol interactions with phospholipids in membranes, Progress in Lipid Research 2002 41:66-97

    (5) T. Róg , M. Pasenkiewicz-Gierula, I. Vattulainen, M. Karttunen, What Happens if Cholesterol Is Made Smoother: Importance of Methyl Substituents in Cholesterol Ring Structure on Phosphatidylcholine–Sterol

    Interaction, Biophysical Journal 2007 92:3346-3357

    (6) W. J. van Blitterswijk, B. W. van der Meer, H. Hilkmann, Quantitative Contributions of Cholesterol and the Individual Classes of Phospholipids and Their Degree of Fatty Acyl (Un)Saturation to Membrane Fluidity

    Measured by Fluorescence Polarization, Biochemistry 1987 26:1746-1756

    (7) T. P.W. McMullen, R. N.A.H. Lewis, R. N. McElhaney, Cholesterol–phospholipid interactions, the liquid-ordered phase and lipidrafts in model and biological membranes, Current Opinion in Colloid and Interface Science 2004 8:459–468

    (8) T. Róg , M. Pasenkiewicz-Gierula, I. Vattulainen, M. Karttunen, Ordering effects of cholesterol and its analogues, Biochimica et Biophysica Acta 2009 1788:97–121

    (9) W. F. D. Bennett, J. L. MacCallum, D. P. Tieleman, Thermodynamic Analysis of the Effect of Cholesterol on Dipalmitoylphosphatidylcholine Lipid Membranes, J. Am. Chem. Soc 2009 131:1972–1978

    (10) E. Overton, Vierteljahrsschr. Naturforsch. Ges. Zuerich 1896, 41, 383 (11) H. Meyer, Arch. Exp. Pathol. Pharmakol 1899, 42, 109-118 (12) A. Finkelstein, Water and Nonelectrolyte Permeability of Lipid Bilayer Membranes, J. Gen. Phys. 1976

    68:127-135

    (13) A. Missner, P. Pohl, 110 Years of the Meyer-Overton Rule: Predicting Membrane Permeability of Gases and Other Small Compounds, ChemPhysChem 2009 10:1405-1414

    (14) S. Marrink, H. J. C. Berendsen, Simulation of Water Transport through a Lipid Membrane, J. Phys. Chem. 1994 98:4155-4168

    (15) J. F. Nagle, J. C. Mathai, M. L. Zeidel, S. Tristram-Nagle, Theory of Passive Permeability through Lipid Bilayers, J. Gen. Physiol 2008 131:77-85

    (16) A. J. Kox, J. P. J. Michels, F. W. Wiegel, Simulation of a lipid monolayer using molecular dynamics, Nature 1980 287:317-319

    (17) B. Jönsson, O. Edholm, O. Teleman, Molecular dynamics simulations of a sodium octanoate micelle in aqueous solution, J. Chem. Phys. 1986 85:2259-2271

    (18) D. Bassolino-Klimas, H. E. Alper, T. R. Stouch, Solute diffusion in Lipid Bilayer Membranes: An Atomistic Level Study by Molecular Dynamics simulation, Biochemistry 1993 32:12624-12637

    (19) J. C. Mathai, S. Tristram-Nagle, J. F. Nagle, M. L. Zeidel, Structural Determinants of Water Permeability through the Lipid Membrane, J. Gen. Physiol. 2008 131:69-76

    (20) D. Bemporad, C. Luttmann, J. W. Essex, Computer Simulation of Small Molecule Permeation across a Lipid Bilayer: Dependence on Bilayer Properties and Solute Volume, Size, and Cross-Sectional Area, Biophysical

    Journal 2004 87:1–13

    (21) A. J. Robinson, W. G. Richards, P. J. Thomas, M. M. Hann, Behavior of Cholesterol and Its Effect on Head Group and Chain Conformations in Lipid Bilayers: A Molecular Dynamics Study, Biophysical Journal 1995

    68:164-170

    (22) D. Bassolino-Klimas, H. E. Alper, T. R. Stouch, Mechanism of Solute Diffusion through Lipid Bilayer Membranes by Molecular Dynamics Simulation, J. Am. Chem. Soc. 1995 117:4118-4129

    (23) A. R. Leach, (2001). Molecular Modelling Principles and Applications Essex: Pearson Education Limited

    (24) J. S. Hub Ph.D. thesis, University of Göttingen 2008 (25) G. M. Torrie, J. P. Valleau, Nonphysical Sampling Distributions in Monte Carlo Free-Energy Estimation:

    Umbrella Sampling, J. Comp. Physics 1977 23:187-199

    (26) B. Roux, The Calculation of the Potential of mean force using computer simulations, Computer Physics Communication 1995 91:275-282

  • 22

    (27) E. J. Shimshick, H. M. McConnell, Lateral Phase Separations in Binary Mixtures of Cholesterol and Phospholipids, Biochemical and Biophysical Research Communications 1973 53:446-451

    (28) R. F. M. de Almeida, A. Fedorov, M. Prieto, Sphingomyelin/Phosphatidylcholine/Cholesterol Phase Diagram: Boundaries and Composition of Lipid Rafts, Biophysical Journal 2003 85:2406–2416

    (29) T. P.W. McMullen, R. N. McElhaney, New aspects of the interaction of cholesterol with dipalmitoylphosphatidylcholine bilayers as revealed by high-sensitivity differential scanning calorimetry,

    Biochimica et Biophysica Acta 1995 1234:90-98

    (30) C. Hofsäß, E. Lindahl, O. Edholm, Molecular Dynamics Simulations of Phospholipid Bilayers with Cholesterol, Biophysical Journal 2003 84:2192–2206

    (31) D. van der Spoel, E. Lindahl, B. Hess, G. Groenhof, A. E. Mark, H. J. C. Berendsen, GROMACS: Fast, Flexible and Free, J. Comp. Chem. 2005 26:1701-1719

    (32) W. L. Jorgensen, D. S. Maxwell, and J. Tirado-Rives, Development and Testing of the OPLS All-Atom Force Field on Conformational Energetics and Properties of Organic Liquids, J. Am. Chem. Soc. 1996

    118:11225–11236

    (33) W. L. Jorgensen, J. Chandrasekhar, J. D. Madura, R. W. Impey, and M. L. Klein, Comparison of simple potential functions for simulating liquid water, J. Chem. Phys. 1983 79:926-935

    (34) O. Berger, O. Edholm, F. Jahnig, Molecular dynamics simulations of a fluid bilayer of dipalmitoylphosphatidylcholine at full hydration, constant pressure, and constant temperature, Biophysical

    Journal 1997 72:2002-2013

    (35) J. P. Ulmschneider, M. B. Ulmschneider, United Atom Lipid Parameters for Combination with the Optimized Potentials for Liquid Simulations All-Atom Force Field, J. Chem. Theory Comput. 2009 5:1803–

    1813

    (36) B. Hess, H. Bekker, H. J. C. Berendsen, J. G. E. M. Fraaije, LINCS: A linear constraint solver for molecular simulations, J. Comp. Chem. 1997 18:1463-1472

    (37) J. S. Hub, B. L. de Groot, D. van der Spoel, g_wham—A Free Weighted Histogram Analysis Implementation Including Robust Error and Autocorrelation Estimates, J. Chem. Theory Comput. 2010 6: 3713–3720

  • 23

    Appendix A Shown here is the total energy during the equilibration of the lipid patches. Energy is shown in

    kJ/mol and is plotted against the simulation time in ps. Running average over 1000 data points shown in red.

    Pure POPE: POPE with 20% cholesterol

    POPE with 30% choelsterol POPE with 40% cholesterol

    POPE with 50% cholesterol

  • 24

    Pure POPC POPC with 20% cholesterol

    POPC with 30% cholesterol POPC with 40% cholesterol

    POPC with 50% cholesterol

  • 25

    Pure DMPC DMPC with 20% cholesterol

    DMPC with 30% cholesterol DMPC with 40% cholesterol

    DMPC with 50% cholesterol

  • 26

    Pure DPPC DPPC with 20% cholesterol

    DPPC with 30% cholesterol DPPC with 40% cholesterol

    DPPC with 50% cholesterol

  • 27

    Appendix B PMFs for all solutes in all membranes. Each plot shows five PMFs for each solute in each

    membrane, one for each cholesterol concentration. Center of the membrane is located at z=0, and head group regions at approximately +/- 2.

    POPE

    Nitric oxide Propane

    Benzene Neopentane

    Ammonia

  • 28

    POPC

    Nitric oxide Propane

    Benzene Neopentane

    Ammonia

  • 29

    DMPC

    Propane Nitric oxide

    Benzene Neopentane

    Ammonia

  • 30

    DPPC

    Nitric oxide Propane

    Benzene Neopentane

    Ammonia

  • 31

    Appendix C Plots showing changes in the permeation barrier in the head- and tail-region of the lipid patches,

    for each solute, at different cholesterol concentrations. Head group region shown in orange, tail region in yellow.

    Nitric oxide

  • 32

    Propane

  • 33

    Benzene

  • 34

    Neopentane