Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2011 Feb 1.
Published in final edited form as: Cryobiology. 2009 May 19;60(1):80–90. doi: 10.1016/j.cryobiol.2009.05.001

Statistical Thermodynamics of Biomembranes

Ram V Devireddy 1
PMCID: PMC2854501  NIHMSID: NIHMS189188  PMID: 19460363

Abstract

An overview of the major issues involved in the statistical thermodynamic treatment of phospholipid membranes at the atomistic level is summarized: thermodynamic ensembles, initial configuration (or the physical system being modeled), force field representation as well as the representation of long-range interactions. This is followed by a description of the various ways that the simulated ensembles can be analyzed: area of the lipid, mass density profiles, radial distribution functions (RDFs), water orientation profile, Deuteurium order parameter, free energy profiles and void (pore) formation; with particular focus on the results obtained from our recent molecular dynamic (MD) simulations of phospholipids interacting with dimethylsulfoxide (Me2SO), a commonly used cryoprotective agent (CPA).

Keywords: GROMACS, Dimethylsulfoxide, Lipid Bilayer Structure, Mass Density Profiles, Radial Distribution Function, Water Orientation Profile, Deuterium Order Parameter

Introduction

Understanding the basic principles of biomembranes (lipid bilayers), which govern and mediate various biologically relevant processes, on the microscopic level is one of the great challenges in biology. To investigate the characteristics of the membranes and to obtain the intriguing physicochemical aspects of membranes systems many experiments have been (and are still being) performed [21,52,70,108,117,150,153,186]. Recent development of new algorithms and revolutionary advances in the computational power has permitted large-scale molecular dynamic (MD) simulations of interaction(s) between biomembranes and small non-water polar molecules, as well as simulations of two component mixtures of phospholipid membranes and other natural amphiphiles [3,15,40,41,56,60,61,71,72,89,111,124,144,145,159]. Thus, allowing the development of a combination of computer simulations and experiments to analyze biomembrane properties with an even greater degree of detail [5,17,45,47,87,90,170].

Molecular dynamic (MD) simulations are well suited for detailed analysis of the interactions between lipid bilayers and various small molecules, including water, chemicals, co-enzymes, peptides, oligonucleotides and proteins, as evidenced by the extensive body of published literature [1,7,10,14,27,29,36,40,41,45,54,58,59,62,68,73,8183,92,94,96,97,105,111,115,119,122,125127,130,131,135,140,141,147,149,150,155,157,158,163,165,179,182,187,188,190]. Briefly, these studies describe the effect of cholesterol [29,40,68,86,126,127,140,141,158,172], of dimethylsulfoxide (Me2SO) [27,122,157,179,182], of methanol [79,135,139], of ions [54,60,96,119,125,155], of proteins [37,58,76,94,107] and of disaccharides [130,131,163], on lipid bilayers; others describe the permeability coefficients of small organic molecules through lipid bilayers [710,103], the water/bilayer interface [14,41,73,104,106,111], the permeation of water across a lipid bilayer [13,30,81,102,188], lipid-DNA complexes [6,36,56], porating electric fields for various lipid bilayers [59,189], as well as the aquaporin-1 water channel in a lipid bilayer [30,58,62,63,73,187]. Clearly, a detailed description of all these studies is beyond the scope of this communication! The primary aim of this article is to describe the various steps involved in characterizing and analyzing a phospholipid membrane interacting with small molecules (like cryoprotective agents, CPAs) using statistical thermodynamics and large scale atomistic-level computer simulations.

Statistical Thermodynamics

Statistical treatment of a system, in general, assumes a large enough sample that the average behavior of the sample is representative of the full-size system. Thus, the statistical thermodynamic description of a system is based on the fact that the behavior of a system can be explained by sampling a collection of states, rather than a single state. A thermodynamic description of a system can use different sets of independent variables, like the number of particles (N), volume (V), energy (E), absolute temperature (T), chemical potential (μ) and pressure (p). Once the independent variables are chosen, other variables are determined by various thermodynamic relations (e.g., the equation of state, Maxwell relations) [16,18,19,48,66,134,151,160]. Each choice of independent variables, defines a different set of such relationships, and corresponding to each choice of independent variables, there is a statistical thermodynamic ensemble, with their respective formalism [11,23,48,88,98,109,110,151].

Choice of Thermodynamic Ensemble

The thermodynamic ensembles most frequently used, include the canonical (N,V,T), micro-canonical (N,V,E), isothermal-isobaric (N,p,T) and grand-canonical (μ,V,T) ensembles and in some cases, the surface tension (γ) can also be included as an additional variable, leading to simulations in the multiphase ensemble (N,p,γ,T) [3,28,44,61,89,149]. The selection of an ensemble is based on the thermodynamic property that is pre-set to the right value or guaranteed to agree with the experimental value. For example, setting V and N constant, the area/headgroup for lipid bilayers can be set to the known experimental value, although this cannot always be expected to result in the correct value (1 atm) for pressure [3,61,89,105]. Conversely, setting p and T constant requires periodic changes in the volume but allows the area of the lipid bilayer to fluctuate [135,149,172,175]. Using a constant μ, necessitates a change in the number of particles via the use of insertions or deletions but helps to rapidly equilibrate the system [37,94,97,113,177].

Use of Computer Simulations

Analytical theories exist for the characterization of ensembles of simple systems; however, complex systems, like the phospholipid membranes, are not amenable to analytical treatment without extreme simplification. Such complex systems, however, are amenable to be modeled in full atomistic detail using computer simulations [3,61,82,83,89,111,149]. If the size of the computational domain is infinite, the results are independent of the thermodynamic ensemble. However, for a finite computational system, the results can differ by an amount that is proportional to 1/N [16,26,46,50,89]. Clearly, the larger the number of molecules in the model (N), the better the statistical representation, but the associated computational cost is also higher! Fortunately, the accuracy of a computer model with limited number of molecules can be increased significantly by the use of periodic boundary conditions, i.e., a basic computational cell containing the system is surrounded by periodic replicas in all three dimensions [38,71,72,144,145,164,176,184].

Currently two major classes of computational methods are used for simulating soft condensed matter like the lipid bilayer membranes: molecular dynamics (MD) and the Monte Carlo (MC) approach. MD is based on the integration of Newton's law of motion [3,50,61,82,83,89,94,111] while MC uses a mathematical technique called the Markov chain [29,31,35,49,100,112,113,115,123,132,152,162,185]. Of these two techniques, MD has proven to be more popular, since it seems to work well and conversely, the few MC studies do not, as yet, seem to offer any significant advantage or ease of use. Additional techniques like Dissipative Particle Dynamics and Coarse-Grained models, still cannot be considered as predictive as all-atom molecular dynamic simulations [3,61,75,91,105,149,183]. Thus, MD simulations have emerged as one of the principal tools in the theoretical study of biological molecules [3,61,67,82,83,89,105,149].

Molecular dynamics is the science of simulating the motion of a large number of particles using Newton's laws of motion. Since, these simulations provide a description of the individual particle motion as a function of time, they can be analyzed in far more detail than actual experiments and generate a detailed atomistic description (and the physical properties) of the system being modeled. An essential requirement for performing these simulations is an a priori knowledge of the interaction potentials (force fields) for the particles [34,51,69,99,120,147]. The interaction potentials, although approximate, are under the control of the programmer and consequently, can be refined to represent the physical system being modeled as accurately as possible.

Initial Configuration – Biomembrane as a Lipid Bilayer

The cell membrane is a selectively permeable barrier that separates the intracellular components from the extracellular space and actively control the composition of the intracellular fluid. It contains a wide variety of biomolecules, primarily proteins and phospholipids. Generally, phospholipids (double chain amphiphiles) spontaneously arrange themselves in an aqueous solution into a bilayer, with the hydrophobic tail regions of the lipids orientated away from the water and the hydrophilic heads oriented towards the water, while other phospholipids (single chain amphiphiles) will form a closed sphere, i.e. a micelle [55,104,118,128,156,168]. These lipid bilayers (biomembranes) are routinely utilized in computer simulations as an idealized (and simplified) representation of the complex structure of cell membranes. With lipid simulations becoming more and more widespread, reasonably well-equilibrated initial configurations can be obtained from earlier simulations of the same or similar systems [7,10,14,2729,31,34,37,38,40,45,47,58,71,72,92,111,129131,145147,149]. These can be directly imported using commercially and freely available MD computational software, including GROMACS, AMBER and CHARMM; Appendix 1 lists the major features of some of the commonly used MD software [20,24,25,80,93,114,133,174]. Alternatively, the initial structure of the phospholipid membrane can be constructed or created by placing the required number of lipids in a solution (water) with the appropriate force fields (described below) and to letting the system (lipids and water) equilibrate by running an initial simulation; this will result in the formation of the well known lipid bilayer, lipid heads oriented towards the water and tails away from it [41,101,104,118,128,136,137,156]. If the thermodynamic ensemble pre-selects (or inputs) p and T to a constant value, then the value of the area of the lipid will act as a test of the “equilibration” and to the validity of the initial lipid bilayer. Since the area per lipid of an equilibrated lipid bilayer can be measured from experiments [5,17,21,52,70,153] and compared with the simulations [14,29,41,6870,92,97,135,171].

Description of Force Fields

As stated earlier, the generation of an atomistic computer model for a phospholipid bilayer membrane involves the a priori representation of inter-molecular energies and/or forces. The most accurate representation obtained using quantum-mechanical techniques (ab initio calculations) is still prohibitively expensive [53,57,64,74,78,143,169]. Thus, the most commonly used approach to represent inter-molecular forces and/or energies is through the use of molecular mechanical force fields and treating non-bonded interactions in a pairwise additive manner [3,41,61,65,69,89,111,120,147]. Additionally, the intra-molecular interactions are described with bond stretching and bending as well as torsional terms [3,10,16,61,89,111,147,149].

Molecular mechanical force fields express the energy of the system E(XN) as a sum of several terms, and calculate the force acting on each atom as the gradient of this energy: E(XN) = Ebond-stretch + Ebond-bend + Erotate-along-bond + Enon-bonded, where, Ebond-stretch, Ebond-bend, Erotate-along-bond represent the intra-molecular energies summed over all bonds, bond angles and torsions, respectively; Ebond-stretch and Ebond-bend are generally represented with harmonic terms as: Ebond-stretch=kijb(rijrij0)2 and Ebond-bend=kija(αijkαijk0)2, where i and j are the atoms forming the bond of length rij, i, j, and k are the atoms forming the bond angle αijk, the superscript 0 refers to the equilibrium value and the parameters ka and kb are the respective harmonic strengths. In general, the Erotate-along-bond represents two types of terms: the first term, the contribution of the conformational state of a bond to the energy of the molecules is, usually, expressed as, Erotate-along-bend,conformational=kijkl[1+cos(nijklδijkl+δijkl0)], where δ is the torsion angle, the parameter kijkl represents the strength of the interaction, and the parameters nijkl and δoijkl depend on the type of the bond. The other type of Erotate-along-bond is described as improper torsion and is used to enforce either the chirality of an atom or to keep a bond in a plane (e.g., in the case of an aromatic ring). For an atom k with bonded neighbors i, j, and l, it is a harmonic function of the angle between the planes formed by atoms i, j, and k and by atoms j, k, and l [3,50,61,89].

The Enon-bonded represents the non-bonded energy, summed over all pairs of atoms Xi and Xj, separated by the distance rij, that are on different molecules or in the same molecule but separated by more than three bonds and includes the contributions of Van der Waals forces and electrostatic interactions as: Enon-bonded (Xi,Xj) = Evan-der-waals + Eelectrostatic. Typically, Evan-der-waals=4ε[(σijrij)12(σijrij)6] where εij and σij are the so-called Lennard-Jones parameters, representing the depth of the attraction due to dispersion forces and the extent of exchange repulsion, respectively; and Eelectrostatic=qiqj4πεorij where qi and qj are the partial charges assigned to atoms i and j to represent the electrostatic interaction between them. Most force fields assign values for each atom type and obtain εij and σij as a combination of the two using the Lorentz-Berthelot rule: arithmetic mean for σij and geometric mean for εij. Partial charges are either obtained from empirical rules or from ab initio calculations, using a fitting procedure that finds partial charges by ensuring the best reproduction of the electric field around a molecule [39,65,77,99,165]. In general, the Lennard-Jones parameters are established independent of the molecule (i.e. specific to the atom), while partial charges are specifically assigned for each molecule [3,24,61,89,111,148,149,174]. Obviously, extreme care should be taken to utilize force fields that have evolved as a whole. Since mixing terms from various sources will lead to an incorrect result.

Treatment of Long-Range Interactions

Accounting for the interactions between all atoms, even assuming pairwise additive potential, as described above, will require enormous computational resources (proportional to N2). The computational time can be significantly reduced by treating the interactions between “distant” atoms separately from interactions between pairs of atoms that are “close” to each other. This observation lead to the introduction of a cutoff distance, Rcutoff, defined as the distance beyond which all interaction energies are assumed to be zero [3,4,24,85,111,174,178]. Obviously choosing a cutoff distance that is “too short” will lead to significant errors, as the contribution of small but numerous distant pairs can add up to a significant amount [32,129]. In addition, the force discontinuity at the cutoff distance, needs to be eliminated by the introduction of a switching function that continuously changes the function to be cut off to zero over a finite interval [32,149,174].

An alternative option to the use of cutoff distance, is to use a simplified representation of interactions between distant pairs, commonly denoted as Ewald sum [22,33,42,43,84,85,121,129,138,142,178]. For simulating lipid membranes this takes the form of using a formalism to obtain the interaction with simplified forms of all periodic cells, extending to infinity. However, the summation of the electrostatic interactions between a simulation cell and its periodic replicas, is nontrivial: the resulting infinite series is only conditionally convergent [3,61,89]. As a consequence, the final sum depends on the order of summation. Ewald introduced a technique that calculates the dipolar sum as two absolute convergent series, one of them in the reciprocal space, hence, the nom de guerre [22,42,43]. The use of Ewald sum has been further facilitated by using the particle mesh Ewald (PME) algorithm that has greatly reduced the associated computational cost and complexity [32,166,178,180].

Statistical Analysis of the Generated Ensemble

A description of the various ways that the simulated ensembles can be analyzed is provided below. To illustrate the analysis, a representative selection of results obtained from recent computer simulations, performed in collaboration with my colleague, Prof. D. Moldovan, and graduate students, D. Pinisetty and R. Alapati, of lipid bilayers interacting with 6 mol% Me2SO are also shown. A detailed description of our MD simulations is provided elsewhere and, in the interest of brevity, will only be briefly described here [116,135]. We have simulated a Di-Myristoyl-Phosphatidyl-Choline (DMPC) lipid system consisting of 96 lipid molecules (48 lipids in each leaflet) in the presence of 5,422 water molecules (full hydration) and 326 Me2SO molecules using the GROMACS 3.2.1. The force field parameters for both bonded and non-bonded interactions were taken from Berger et al. [14] while the partial charges were taken from Saiz and Klein [146]. Periodic boundary conditions were applied along the three space dimensions. The pressure was maintained at 1 atm using the semi-isotropic pressure coupling to a Berendsen barostat with a time constant of 1.0 ps [12,136,137]. The height of the simulation box (z direction) and the cross sectional area (xy-plane) was allowed to vary independently of each other, thereby allowing the area of the bilayers and the distance between the interfaces to fluctuate independently [41]. The nonbonded Lennard-Jones interactions were cut-off at a distance of 1.0 nm and the simulation time step was set to 2 fs. Long-range electrostatics were updated every 10 time steps and handled by particle-mesh Ewald (PME) algorithm [42,43,180]. An energy minimization procedure based on the steepest descent algorithm was initially applied to the initial structure prior to the actual MD run. The atomic coordinates were saved every 2ps for analysis and the total simulation time was 40 ns.

Area Per Lipid

The surface area per lipid is one of the most important quantities characterizing a bilayer membrane and is often monitored in simulations to assess whether or not the system has reached the equilibrium during the atomistic simulation [3,41,61,89,135,165]. Fig. 1 shows the time variation of the area per lipid in the DMPC bilayer systems in the presence and absence of Me2SO over the total simulation time (40 ns). The simulation results show that after an initial transient regime of ~20 ns, the (data not shown). The illustrative figures shown for the mass density profiles, the radial distribution functions, the deuterium order parameters and the water orientation profile are generated by analyzing the simulations between 23 to 27 ns (and represent the equilibrated behavior of the DMPC lipid in the presence of Me2SO at ~25 ns).

Fig. 1.

Fig. 1

Time dependence of the area per lipid for DMPC bilayers in the presence of 6.0 mol% water-Me2SO mixtures. For reference and comparison the temporal evolution in pure water (or 0 mol% Me2SO) is also included. The system equilibration time is shown as a vertical dotted-dashed line.

Mass Density Profiles

One of the most important tools in characterizing the average structure of phopsholipid membranes at different regions along its normal axis is the mass density profile of various atoms or atomic groups. The calculation of density profiles is quite simple and requires only the counting of the average occurrence of the atoms of interest in different lateral slices of the membrane and then dividing it by the volume of the slice (Figs. 2 and 3). Generally, as shown in Figs. 2 and 3, the shape of the mass density profiles in phospholipid membranes shows that the highest density is near the region of the lipid headgroups while the lowest density part is in the middle region of the chain terminal methyl groups. Physically, the density profiles of an atom or atomic group shows how deeply these atoms or atomic groups can (and have) penetrate into the bilayer. Thus, the density at the center of the membranes is a measure of the ability of the component to penetrate into the center of the bilayer. Additionally, the membrane thickness can be estimated by the distance between the density peaks corresponding to the two headgroup regions. Thus, as shown in Fig. 2, the presence of Me2SO in the system causes a decrease in the thickness of the phospholipids membrane (as seen by a comparison between the peaks of the two lines corresponding to the lipid density in the presence and absence of Me2SO). And finally, the comparison of the density distribution of different atoms or atomic groups along the membrane normal axis can also give some indication on the average alignment of various parts of the lipid molecules [7,14,30,106].

Fig. 2.

Fig. 2

Mass density profiles of DMPC lipid (blue) and water (red) across the whole bilayer in the presence (solid line) and absence (dotted line) of 6.0 mol% Me2SO. The size of the lipid bilayer in the z-direction is shown on x-axis and mass density is shown on y-axis. The lipid component is divided into the contribution from two separate leaflets with the center of the lipid bilayer system lying at `0' on the x-axis.

Fig. 3.

Fig. 3

Mass density profiles across the bilayer. The densities given are the nitrogen, N (brown) and phosphorous, P (green) of the DMPC lipid in the presence (solid line) and absence (dotted line) of 6.0 mol% Me2SO. The mass density profile of Me2SO is also shown. The size of the lipid bilayer in the z-direction is shown on x-axis and mass density is shown on y-axis. The center of the lipid bilayer system lies at `0' on the x-axis.

Analysis of the Lipid Head Groups and Lipid Tails

The polar part of the lipid molecules along with the water molecules that might have penetrated into the bilayer constitute the dense head group region of the phospholipids membranes [41,118,173]. Obviously, the intricate interactions between the lipid-lipid and lipid-water are of immense importance in determining the properties of the bilayer membrane [138,168,173]. A description of the headgroup region structure includes the analysis of the lipid headgroups (using radial distribution functions) as well as the structure of the interfacial water (water orientation profiles) while the lipid tails are analyzed via the profile of the methylene group order parameter along the lipid tails (deuterium order parameter).

Radial Distribution Functions (RDFs)

The radial distribution function g(r) is often employed to identify close range ordering of neighboring atoms. In this regard, the radial distribution function, g(r), is defined as: g(r) = N(r)/4πr2ρdr, where, N(r) is the number of atoms in a spherical shell at distance r and thickness dr from a reference atom, and ρ is the number density taken as the ratio of atoms to the volume of the computing box [3,61,82,83,89,94,111]. The calculations of g(r) from a simulation are performed by sorting the neighbors around each atom or molecule into distance “bins” or histograms. The number of neighbors in each bin is then averaged over the entire simulation. Thus, g(r) measures how atoms organize themselves around each other. Specifically, g(r) gives the probability of finding an atom (or molecule, if simulating a molecular fluid) a distance, r, from another atom or molecule compared to the ideal gas distribution. g(r) is thus dimensionless. Higher radial distribution functions (e.g. the triplet radial distribution function) can also be defined but are rarely calculated [3,19,50,89]. The radial distribution function, shown in Fig. 4, shows g(r) is zero for short distances (less than the atomic diameter) and due to the strong repulsive forces. At long distances, g(r), tends to one, indicating that there is no long-range order. In between the peak values represent the increased likelihood of the two molecules being within this distance (for example a g(r) value of 4, means that it is 4 times more likely that two molecules would have this separation than average). Since, RDFs depend on density and temperature, g(r) serves as an indicator of the nature of the phase of the simulated system [29,129,135,145,147,157,163,168]. For example, in regular crystal lattice structures, g(r) will form a sequence of jagged, discrete, unequally spaced columns. If the atoms are vibrating rather than fixed the g(r) will resolve into Gaussians. Note that a primary peak will always occur for any phase at σij (the Lennard-Jones parameter representing the extent of exchange repulsion), since this is the most common radius recorded when two particles collide [3,61,89,94,111]. However, secondary peaks indicate that multiple layers of particles are packed together hinting at a crystalline structure [40]. RDFs can also be measured using X-ray diffraction, allowing us to verify/contradict the usefulness of the MD simulations [5,33,47,51,108,118,168].

Fig. 4.

Fig. 4

Radial distribution functions (RDF) between nitrogen, N-N (blue) and phosphorous, P-P (red color) atoms present in the present in the lipid head group in the presence (solid line) and absence (dotted line) of 6.0 mol% Me2SO. The spherical radius `r' is shown on x-axis and radius of gyration (g(r)) is shown on y-axis.

Structure of the Interfacial Water (Water Orientation)

The change of the orientation order of the interfacial water molecules along the membrane normal axis can be characterized by the profiles (i.e., the average values obtained in different lateral membrane slices) of appropriately chosen orientation parameters [3,24,27,58,83]. Thus, the water ordering in the vicinity of the bilayer-water interface (hydration layer) can be obtained from studying the mean cosine value, < cosθ >, of the angle between the water dipolar moment μ and the bilayer normal unit vector n. That is: <cosθ(z)>=1μ(z)<μ(z)n>, where z is the z-coordinate of the centre of mass of the water molecules. The mean cosine value is obtained by averaging, when the system is in the equilibrium regime, over dipolar orientations of all water molecules present in the system and over a large number of equilibrium states. Fig. 5 shows the results from our MD simulations for the ordering of water in the vicinity of DMPC bilayer systems, with and without Me2SO. When generating the water orientation profile, the normal unit vector, n parallel to the z-axis was considered to have the same orientation for the solvent on both sides of the bilayer. Consequently, by symmetry, the cosine average has opposite sign in the two regions. As seen in Fig. 5, the presence of Me2SO decreases both the ordering of water molecules in the hydration layer (as evidenced by the decrease in the peak heights) and the separation distance between the centers of the two hydration layers present on both sides of the membrane.

Fig. 5.

Fig. 5

Water orientation profiles across the vicinity of the DMPC lipid/water interface in the presence (solid line) and absence (dotted line) of 6.0 mol% Me2SO. The size of the lipid bilayer in the z-direction is shown on x-axis and mean cosine angle between water dipolar movement and bilayer normal vector is shown on y-axis. The center of the lipid bilayer system lies at `0' on the x-axis.

Lipid Tails Order Parameter

Generally, the conformation of the hydrocarbon tails of the lipids is highly disordered, since the membrane is in the lamellar or the liquid crystalline (Lα) phase. The conformational and orientation order/disorder can be quantified by various quantities, including the tilt angle of the C-C bonds along the hydrocarbon tails and the profile of the methylene group order parameter along the lipid tails [41,52,111,135,171,172]. The latter quantity (viz. the methylene group order parameter or the deuterium order parameter) can also be measured using nuclear magnetic resonance (NMR) spectroscopy, and therefore, can be used to corroborate or contradict the simulation results [21,51,82,90,170,173,186]. The deuterium order parameter, SCD, is related to the components of the order parameter tensor, Sαβ, defined as: Sαβ=12<3cosθαcosθβδαβ>, where a, b = x, y, z and θα is the angle between the molecular α-axis and the bilayer normal (z-axis). The brackets “< >” denote a time average over an ensemble of configurations. The molecular axes must be defined separately for each segment of an acyl chain. Usually for the Cn methylene group the Cn−1 − Cn+1 direction is taken as z, and the Cn−1 − Cn − Cn+1 plane is the yz plane. Since the bilayer is symmetric with respect to the rotation around z-axis (perpendicular to the bilayer), we have Sxx = Syy and Sxx + Syy + Szz = 0. The relevant order parameter is the diagonal element, Szz, which is related to the deuterium order parameter, SCD, as: SCD=23Sxx13Syy, or SCD=32cos2α12, where α is the angle formed by the C-H bond with the membrane normal. Given the system symmetry we can simply write SCD = 0.5*Szz. Noticeable deviation of the obtained Szz values from −2Sxx or −2Syy indicates rotational anisotropy along the molecular axis joining two C atoms that are separated by two C-C bonds. The deuterium order parameter, SCD, typically ranges from 0 to 0.5. The value of 0.5 is associated with an all transconformation state and a value of 0 is associated with the isotropically disordered state. From the simulations the order parameter is calculated separately for all carbon atoms along the acyl chain and is shown in Fig. 6 [40,136,137,173]. The simulation results clearly show that the presence of Me2SO leads to a substantial decrease of ordering along both lipid tails, with the disordering effect being most pronounced for n being 2 to 7, close to the beginning and the middle of the lipid tails. Additionally, our simulations suggest that the disordering of the carbon atoms along the acyl chain keeps increasing as the concentration of Me2SO is increased from 3 mol% to 11.3 mol% (data not shown).

Fig. 6.

Fig. 6

Ordering of DMPC lipid's acyl sn-1 chain in the presence (solid line) and absence (dotted line) of 6.0 mol% Me2SO. The atom number of carbon atoms is shown on x-axis and deuterium order parameter (Scd) is shown on y-axis.

Free Energy Profile of Small Molecules and Membrane Permeability

It is possible to analyze the free energy profile (and consequently, obtain the membrane permeability) of small, non-polar (neutral) physiologically important molecules like O2, CO2, NO, etc [710,103,119,155,177]. The profile of their solvation free energy across the membrane is determined by inserting the molecules into the system (particle insertion method) or by the method of constrained particles. The particle insertion method is well suited for the less dense parts of the lipid bilayer while the method of constrained particles is well suited for all parts of the lipid bilayer, but in comparison to the particle insertion method is computationally more expensive. The particle insertion method samples the relevant physical space through random insertions of the pre-selected molecule within the biomembrane and obtains the free energy profile of the molecule as the difference between it's chemical potential in the membrane and in the bulk solution. The method of constrained particles constructs a potential of mean force by restricting the molecules to different regions within the biomembrane (and measures the average force, in the direction of the constraint, required to keep the molecule within the proscribed region). When the diffusion constant profile of the solute D(Z) is also determined (e.g., by the force correlation method which relates the time fluctuations in the instantaneous force from the constrained particle method to the local time dependent friction coefficient and to the diffusion constant profile), the experimentally accessible permeability coefficient of the solute, P, can also be calculated using the inhomogeneous solubility-diffusion model of Marrink and Berendsen [103], as: 1P=Z1Z2exp(A(Z)kBT)D(Z)dZ, where A(Z) is the free energy profile along the membrane normal axis, Z; kB is the Boltzmann constant [13,102,105,181]. The inhomogeneous solubility-diffusion model is valid as long as the thermodynamic gradients remain small. Given the extremely inhomogeneous and anisotropic nature of the bilayer, the permeation properties with the solubility-diffusion model of Marrink and Berendsen [103] are to be interpreted in a qualitative nature rather than a strictly quantitative result. However, the results of Marrink and Berendsen [103] do suggest that the resistance to the permeation of small molecules across biomembranes is characterized by the presence of four distinct regions: 1) bulk water: the resistance within this region is negligible (high permeability) for most molecules except for hydrophobic substances; 2) inter phase: a region with high permeability but unlike the bulk phase is characterized by high viscosity and high dielectric constant. This region will only act as a rate limiting step for molecules that exhibit a hydrocarbon/water partition coefficient value of unity; 3) soft polymer/dense alkane: a highly anisotropic region of high density coupled with low solubility of hydrophilic molecule and a largely accessible free volume with an enhanced possibility for the dissolvation of small molecules and is of extreme importance in determining the overall permeability of the molecule in the biomembrane; 4) fluid decane/low density alkane: a region characterized by low dielectric constant and low viscosity, with rapid diffusion.

Void (Pore) Formation

We have recently observed spontaneous formation of small hydrophobic pores across the phospholipids membranes at various times during the simulations (~15 ns for DMPC) in the presence of 11.3 mol% of Me2SO (and also in the presence of 9 mol% of Me2SO; but not at mol% concentrations lower than this) [116]. As there is no external stress applied to the membranes these hydrophobic pores are, presumably, nucleated by thermal fluctuations (Fig. 7). The initial hydrophobic pores transform into hydrophilic pores and continue to grow at an even higher rate, with the corresponding penetration of water and Me2SO into the now hydrophilic pore [116]. The detailed mechanism and the energetics involved in the pore nucleation and growth are still subject to investigation. We postulated an extension to a well accepted Lister model of pore formation [95] by accounting for the entropic contribution to the overall driving force for pore nucleation and growth [154,167]. Utilizing this modified model, we were able to show that the membrane may become unstable to pore nucleation and growth even at zero surface tension provided the pore edge line tension, λ, is lower than the threshold value λ *. The line tensions, λ, of a DMPC bilayer edge, in both water and in Me2SO-water solutions, were calculated from separate “strip bilayer” simulations [167]. Briefly, ribbon-like (strip) bilayers, consisting of 96 lipids surrounded by water (or by Me2SO-water) molecules in the normal (x) and one of the lateral (y) directions, were simulated using periodic boundary conditions in all three directions. The positive line tension, λ, was calculated from the diagonal elements Pij of the pressure tensor, according to the relation, λ=12LxLy[12(Pxx+Pyy)Pzz], where, 〈…〉, stands for time averaging [116,167]. The threshold line tension value, λ *, predicted by this simplified model is temperature and membrane specific, and for DMPC lipid bilayer at T = 300K is, λ * ≈ 5.9 pN. However, the line tension value for a DMPC bilayer edge in the presence of 11.3 mol% Me2SO is ~1.62±1.5 pN; a value that is well below λ *, and hence, pores were able to form and grow (as shown in Fig. 7). Our more recent simulations also suggest that the presence of Me2SO greater than 7 mol% will reduce the line tension below λ * (and the corresponding barrier for pore creation) and facilitate spontaneous and stable pore nucleation and growth in DMPC lipid bilayers (Fig. 8). From a broader perspective of the cryobiological community, these findings suggest that by using certain chemicals one can find ways of bringing biological membranes from the energy dominated stability regions into states in which hydrophilic pores are entropically favored (and consequently, further our understanding of the transport of molecules into and out of the intracellular spaces, a region of extreme interest to cryobiological applications).

Fig. 7.

Fig. 7

Three snapshots of a DMPC bilayer membrane in the presence of 11.3 mol% Me2SO depicting pore formation due to thermal fluctuations. Both top and selected side view areas of the DMPC bilayer are shown. The left figures show the top view of the DMPC bilayers in which only the lipid molecules as shown as space filling, with red for the head groups and blue for the tails. Water and Me2SO molecules are not shown in the top views. For clarity in addition to the simulation cell three additional periodic images of the cell are also shown. One can clearly see that after about 10 ns a hydrophobic pore is nucleated in the DMPC bilayer. The pore continues to grow and after about 15 ns becomes a hydrophilic pore filled with water. The right figures depict cross-sectional side views of selected bilayer areas crossing the pore (as indicated in the left figures by the white dashed lines). Both the water and lipid molecules are shown in the side view pictures. The waters are shown as space filling with red (oxygen) and white (hydrogen) and the lipids are shown as licorice. The side view figures clearly indicate that the pore nucleated after 10 ns, is indeed of hydrophobic nature. This is substantiated by the fact that no water molecules seem to penetrate inside this pore. However as the nucleated pore grows and the lipid hydrophilic heads rearrange around its rim this evolve into a hydrophilic pore (see the substantial water penetration inside the pore at t = 15 ns).

Fig. 8.

Fig. 8

Variation of DMPC lipid bilayer line tension, λ, with the mol% of Me2SO in the water solution. The temperature (T = 300 K) and membrane (DMPC bilayer) specific value of the threshold line tension value, λ * ≈ 5.9 pN is shown as a dashed-dotted horizontal line. Pores nucleate and growth even at zero surface tension provided the pore edge line tension, λ, is lower than the threshold value λ *; this corresponds to a Me2SO mol% concentration greater than ~7% (shown as the dashed vertical line). DMPC lipid bilayer line tension, λ is shown on the y-axis while the mol% of Me2SO is shown on the x-axis. Note that the error bars represent the standard deviation(s) in the simulated value(s) of the line tension.

Concluding Remarks

Starting with the initial Alder and Wainwright molecular dynamic studies on the interactions of hard spheres [2], to realistic simulations of water [161] and proteins [107], we now routinely find molecular dynamic simulations of solvated proteins, protein-DNA complexes, as well as lipid systems addressing a variety of issues, including, water channels, thermodynamics of ligand binding and permeation of molecules [10,30,37,58,62,73,76,83,94,104,105,120,147,165]. It is to be expected that as the available computational resources keep evolving and improving, the ability of computer simulations to model the complex environment of living cells (and membranes) will also continue to improve. Conceivably, in the not too distant future, computer-based simulations might even enable us to understand and predict the behavior of complex living systems that are beyond the reach of our most powerful experimental probes or our most sophisticated theories.

Acknowledgments

I'm extremely grateful to my colleague, Dr. Dorel Moldovan, for, introducing me to Molecular Dynamic (MD) Simulations. Acknowledgements are also due to my graduate students, D. Pinisetty and R. Alapati, for running the MD simulations and for helping with the formatting the figures. Funding for the work was provided by a variety sources, including the Department of Mechanical Engineering at LSU, National Institute of Health (Grant # 5R03EB005985-02), National Science Foundation under EPSCoR Award No. (EPS-0701491).

APPENDIX 1

AMBER (Assisted Model Building using Energy Refinement) is a suite of programs for molecular mechanics and molecular dynamics simulations and was designed primarily for proteins and nucleic acids [24]. Website: http://ambermd.org/

BOSS (Biochemical and Organic Simulation System is a general purpose molecular modeling system that performs molecular mechanics (MM) calculations, Metropolis Monte Carlo (MC) statistical mechanics simulations, and semi-empirical quantum mechanics (QM) calculations [80]. Website: http://zarbi.chem.yale.edu/software.html

CHARMM (Chemistry at HARvard, Macromolecular mechanics) is a highly regarded and widely used simulation package and has been developed with a primary focus on the study of molecules of biological interest. CHARMM combines standard minimization and dynamics capabilities with features including free energy perturbation (FEP), correlation analysis and combined quantum, and molecular mechanics (QM/MM) methods [20,99,114]. Website: http://www.charmm.org/

GROMACS (GROningen MOlecular Simulation package) is a versatile and popular package to perform molecular dynamics. It is primarily designed for biochemical molecules like proteins and lipids that have a lot of complicated bonded interactions, and is extremely fast at calculating the non-bonded interactions [174]. http://www.gromacs.org/

LAMMPS (Large-scale Atomic/Molecular Massively Parallel Simulator) is a classical molecular dynamics code that is somewhat limited in its ability to interface with visualization software [56]. http://lammps.sandia.gov/

NAMD (NAnoscae Molecular Dynamics) is a molecular dynamics program designed to provide high performance simulations for large biological molecular systems. NAMD uses the popular molecular graphics program Visual Molecular Dynamics (VMD) for simulation setup and trajectory analysis, but is also file-compatible with AMBER and CHARMM [133]. http://www.ks.uiuc.edu/Research/namd/

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

REFERENCES

  • [1].Aird A, Wrachtrup J, Schulten K, Tietz C. Possible pathway for ubiquinone shuttling in Rhodospirillum rubrum revealed by molecular dynamics simulation. Biophys J. 2007;92:23–33. doi: 10.1529/biophysj.106.084715. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [2].Alder BJ, Wainwright TE. Phase transition for a hard sphere system. J. Chem. Phys. 1957;27:1208–1209. [Google Scholar]
  • [3].Allen MP, Tildesley DJ. Computer simulation of liquids. Oxford University Press; Oxford, UK: 1987. [Google Scholar]
  • [4].Alper HE, Bassolino D, Stouch TR. Computer simulation of a phospholipid monolayer-water system: The influence of long range forces on water structure and dynamics. J. Chem. Phys. 1993;98:9798–9807. [Google Scholar]
  • [5].Armen RS, Uitto OD, Feller SE. Phospholipid component volumes: determination and application to bilayer structure calculations. Biophys J. 1998;75:734–44. doi: 10.1016/S0006-3495(98)77563-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [6].Bandyopadhyay S, Tarek M, Klein ML. Molecular dynamics study of a lipid-DNA complex. J. Phys. Chem. B. 1999;103:10075–10080. [Google Scholar]
  • [7].Bassolino-Klimas D, Alper HE, Stouch TR. Solute diffusion in lipid bilayer membranes: an atomic level study by molecular dynamics simulation. Biochemistry. 1993;32:12624–37. doi: 10.1021/bi00210a010. [DOI] [PubMed] [Google Scholar]
  • [8].Bassolino-Klimas D, Alper HE, Stouch TR. Mechanism of solute diffusion through lipid bilayer-membranes by molecular dynamics simulation. J. Am. Chem. Soc. 1995;117:4118–4129. [Google Scholar]
  • [9].Bemporad D, Essex JW, Luttmann C. Permeation of small molecules through a lipid bilayer: A computer simulation study. J. Phys. Chem. B. 2004;108:4875–4884. [Google Scholar]
  • [10].Bemporad D, Luttmann C, Essex JW. Computer simulation of small molecule permeation across a lipid bilayer: dependence on bilayer properties and solute volume, size, and cross-sectional area. Biophys J. 2004;87:1–13. doi: 10.1529/biophysj.103.030601. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [11].Ben-Naim A. Statistical thermodynamics for chemists and biochemists. Plenum Press; New York: 1992. [Google Scholar]
  • [12].Berendsen HJC, Postma JPM, Gunsteren W.F.v., DiNola A, Haak JR. Molecular dynamics with coupling to an external bath. J. Chem. Phys. 1984;81:3684–3690. [Google Scholar]
  • [13].Berendsen HJC, Marrink S-J. Molecular dynamics of water transport through membranes: Water from solvent to solute. Pure and Applied Chem. 1993;65:2513–2520. [Google Scholar]
  • [14].Berger O, Edholm O, Jahnig F. Molecular dynamics simulations of a fluid bilayer of dipalmitoylphosphatidylcholine at full hydration, constant pressure, and constant temperature. Biophys J. 1997;72:2002–13. doi: 10.1016/S0006-3495(97)78845-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [15].Biswas A, Schurmann BL. Molecular dynamics simulation of a dense model bilayer of chain molecules with fixed head groups. J. Chem. Phys. 1991;95:5377–5386. [Google Scholar]
  • [16].Blomberg C. Some fundamental aspects of many-body problems in statistical thermodynamics. 2. Tekniska högskolan; Stockholm: 1966. p. 162. Stockholm. [Google Scholar]
  • [17].Bloom M, Evans E, Mouritsen OG. Physical properties of the fluid lipidbilayer component of cell membranes: a perspective. Q Rev Biophys. 1991;24:293–397. doi: 10.1017/s0033583500003735. [DOI] [PubMed] [Google Scholar]
  • [18].Boltzmann L. Vorlesungen über Gastheorie. J. A. Barth; Leipzig: 1896. [Google Scholar]
  • [19].Boublík T, Nezbeda I, Hlavaty K. Statistical thermodynamics of simple liquids and their mixtures. Academia; Praha: 1980. [Google Scholar]
  • [20].Brooks BR, Bruccoleri RE, Olafson BD, States DJ, Swaminathan S, Karplus M. CHARMM: A Program for Macromolecular Energy, Minimization, and Dynamics Calculations. J. Comp. Chem. 1983;4:187–217. [Google Scholar]
  • [21].Brown MF, Ribeiro AA, Williams GD. New view of lipid bilayer dynamics from 2H and 13C NMR relaxation time measurements. Proc Natl Acad Sci U S A. 1983;80:4325–9. doi: 10.1073/pnas.80.14.4325. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [22].Campbell ES. Existence of a “well defined” specific energy for an ionic crystal: Justification of Ewald's formulae and their use to deduce equations for multipole lattices. J. Phys. Chem. Solids. 1963;24:197–208. [Google Scholar]
  • [23].Carey VP. Statistical thermodynamics and microscale thermophysics. Cambridge University Press; Cambridge, UK; New York: 1999. [Google Scholar]
  • [24].Case DA, Cheatham TE, 3rd, Darden T, Gohlke H, Luo R, Merz KM, Jr., Onufriev A, Simmerling C, Wang B, Woods RJ. The Amber biomolecular simulation programs. J Comput Chem. 2005;26:1668–88. doi: 10.1002/jcc.20290. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [25].Case DA, Darden TA, C. TE, III, Simmerling CL, Wang J, Duke RE, Luo R, Crowley M, Walker RC, Zhang W, Merz KM, Wang B, Hayik S, Roitberg A, Seabra G, Kolossváry I, Wong KF, Paesani F, Vanicek J, Wu X, Brozell SR, Steinbrecher T, Gohlke H, Yang L, Tan C, Mongan J, Hornak V, Cui G, Mathews DH, Seetin MG, Sagui C, Babin V, Kollman PA. Amber 10. University of California; San Francisco: 2008. [Google Scholar]
  • [26].Castro-Roman F, Benz RW, White SH, Tobias DJ. Investigation of finite system-size effects in molecular dynamics simulations of lipid bilayers. J Phys Chem B. 2006;110:24157–64. doi: 10.1021/jp064746g. [DOI] [PubMed] [Google Scholar]
  • [27].Chang HH, Dea PK. Insights into the dynamics of DMSO in phosphatidylcholine bilayers. Biophys Chem. 2001;94:33–40. doi: 10.1016/s0301-4622(01)00206-x. [DOI] [PubMed] [Google Scholar]
  • [28].Chiu SW, Clark M, Balaji V, Subramaniam S, Scott HL, Jakobsson E. Incorporation of surface tension into molecular dynamics simulation of an interface: a fluid phase lipid bilayer membrane. Biophys J. 1995;69:1230–45. doi: 10.1016/S0006-3495(95)80005-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [29].Chiu SW, Jakobsson E, Mashl RJ, Scott HL. Cholesterol-induced modifications in lipid bilayers: a simulation study. Biophys J. 2002;83:1842–53. doi: 10.1016/S0006-3495(02)73949-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [30].Choe YK, Tsuchida E, Ikeshoji T, Yamakawa S, Hyodo SA. Nature of water transport and electro-osmosis in nafion: insights from first-principles molecular dynamics simulations under an electric field. J Phys Chem B. 2008;112:11586–94. doi: 10.1021/jp8041878. [DOI] [PubMed] [Google Scholar]
  • [31].Coppock PS, Kindt JT. Atomistic simulations of mixed-lipid bilayers in gel and fluid phases. Langmuir. 2009;25:352–9. doi: 10.1021/la802712q. [DOI] [PubMed] [Google Scholar]
  • [32].Cordomi A, Edholm O, Perez JJ. Effect of different treatments of long-range interactions and sampling conditions in molecular dynamic simulations of rhodopsin embedded in a dipalmitoyl phosphatidylcholine bilayer. J Comput Chem. 2007;28:1017–30. doi: 10.1002/jcc.20579. [DOI] [PubMed] [Google Scholar]
  • [33].Darden T, Perera L, Li L, Pedersen L. New tricks for modelers from the crystallography toolkit: the particle mesh Ewald algorithm and its use in nucleic acid simulations. Structure. 1999;7:R55–60. doi: 10.1016/s0969-2126(99)80033-1. [DOI] [PubMed] [Google Scholar]
  • [34].Davis JE, Rahaman O, Patel S. Molecular dynamics simulations of a DMPC bilayer using nonadditive interaction models. Biophys J. 2009;96:385–402. doi: 10.1016/j.bpj.2008.09.048. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [35].Davis JR, Panagiotopoulos AZ. Monte Carlo simulations of amphiphilic nanoparticle self-assembly. J Chem Phys. 2008;129:194706. doi: 10.1063/1.3009183. [DOI] [PubMed] [Google Scholar]
  • [36].De Luca L, Vistoli G, Pedretti A, Barreca ML, Chimirri A. Molecular dynamics studies of the full-length integrase-DNA complex. Biochem Biophys Res Commun. 2005;336:1010–6. doi: 10.1016/j.bbrc.2005.08.211. [DOI] [PubMed] [Google Scholar]
  • [37].de Meyer FJ, Venturoli M, Smit B. Molecular simulations of lipid-mediated protein-protein interactions. Biophys J. 2008;95:1851–65. doi: 10.1529/biophysj.107.124164. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [38].Dolan EA, Venable RM, Pastor RW, Brooks BR. Simulations of membranes and other interfacial systems using P2(1) and Pc periodic boundary conditions. Biophys J. 2002;82:2317–25. doi: 10.1016/S0006-3495(02)75577-X. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [39].Dubois V, Pasquarello A. Ab initio molecular dynamics of liquid hydrogen chloride. J Chem Phys. 2005;122:114512. doi: 10.1063/1.1869972. [DOI] [PubMed] [Google Scholar]
  • [40].Edholm O, Nyberg AM. Cholesterol in model membranes. A molecular dynamics simulation. Biophys J. 1992;63:1081–9. doi: 10.1016/S0006-3495(92)81678-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [41].Egberts E, Marrink SJ, Berendsen HJ. Molecular dynamics simulation of a phospholipid membrane. Eur Biophys J. 1994;22:423–36. doi: 10.1007/BF00180163. [DOI] [PubMed] [Google Scholar]
  • [42].Essman U, Perera L, Berkowitz ML, Darden T, Lee H, Pedersen LG. A smooth particle mesh Ewald method. J. Chem. Phys. 1995;103:8577–8593. [Google Scholar]
  • [43].Ewald P. Die Berechnung optischer und elektrostatischer Gitterpotentiale. Ann. Phys. 1924;64:253–287. [Google Scholar]
  • [44].Feller SE, Zhang Y, Pastor RW. Computer simulation of liquid interfaces II: surface tension-area dependence of a bilayer and monolayer. J. Chem. Phys. 1995;103:10267–10273. [Google Scholar]
  • [45].Feller SE, Yin D, Pastor RW, MacKerell AD., Jr. Molecular dynamics simulation of unsaturated lipid bilayers at low hydration: parameterization and comparison with diffraction studies. Biophys J. 1997;73:2269–79. doi: 10.1016/S0006-3495(97)78259-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [46].Feller SE. Molecular dynamics simulations of lipid bilayers. J. Curr Opin. Coll. Int. Sci. 2000;5:217–223. [Google Scholar]
  • [47].Feller SE. Molecular dynamics simulations as a complement to nuclear magnetic resonance and X-ray diffraction measurements. Methods Mol Biol. 2007;400:89–102. doi: 10.1007/978-1-59745-519-0_7. [DOI] [PubMed] [Google Scholar]
  • [48].Fermi E. Notes on thermodynamics and statistics. University of Chicago Press; Chicago: 1966. [Google Scholar]
  • [49].Freire E, Snyder B. Monte Carlo studies of the lateral organization of molecules in two-component lipid bilayers. Biochim Biophys Acta. 1980;600:643–54. doi: 10.1016/0005-2736(80)90468-x. [DOI] [PubMed] [Google Scholar]
  • [50].Frenkel D, Smit B. Understanding molecular simulation : from algorithms to applications. Academic Press; San Diego: 2002. [Google Scholar]
  • [51].Gawrisch K, Ruston D, Zimmerberg J, Parsegian VA, Rand RP, Fuller N. Membrane dipole potentials, hydration forces, and the ordering of water at membrane surfaces. Biophys J. 1992;61:1213–23. doi: 10.1016/S0006-3495(92)81931-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [52].Gawrisch K, Gaede HC, Mihailescu M, White SH. Hydration of POPC bilayers studied by 1H-PFG-MAS-NOESY and neutron diffraction. Eur Biophys J. 2007;36:281–91. doi: 10.1007/s00249-007-0142-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [53].Gee RH, Roszak S, Balasubramanian K, Fried LE. Ab initio based force field and molecular dynamics simulations of crystalline TATB. J Chem Phys. 2004;120:7059–66. doi: 10.1063/1.1676120. [DOI] [PubMed] [Google Scholar]
  • [54].Giner Casares JJ, Camacho L, Martin-Romero MT, Lopez Cascales JJ. Effect of Na+ and Ca2+ ions on a lipid Langmuir monolayer: an atomistic description by molecular dynamics simulations. Chemphyschem. 2008;9:2538–43. doi: 10.1002/cphc.200800321. [DOI] [PubMed] [Google Scholar]
  • [55].Goetz R, Lipowsky R. Computer simulation of bilayer membranes: Self-assembly and interfacial tension. J. Chem. Phys. 1998;108:7397–7409. [Google Scholar]
  • [56].Grindon C, Harris S, Evans T, Novik K, Coveney P, Laughton C. Large-scale molecular dynamics simulation of DNA: implementation and validation of the AMBER98 force field in LAMMPS. Philos Transact A Math Phys Eng Sci. 2004;362:1373–86. doi: 10.1098/rsta.2004.1381. [DOI] [PubMed] [Google Scholar]
  • [57].Guidon M, Schiffmann F, Hutter J, VandeVondele J. Ab initio molecular dynamics using hybrid density functionals. J Chem Phys. 2008;128:214104. doi: 10.1063/1.2931945. [DOI] [PubMed] [Google Scholar]
  • [58].Gumbart J, Wang Y, Aksimentiev A, Tajkhorshid E, Schulten K. Molecular dynamics simulations of proteins in lipid bilayers. Curr Opin Struct Biol. 2005;15:423–31. doi: 10.1016/j.sbi.2005.07.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [59].Gurtovenko AA, Vattulainen I. Pore formation coupled to ion transport through lipid membranes as induced by transmembrane ionic charge imbalance: atomistic molecular dynamics study. J Am Chem Soc. 2005;127:17570–1. doi: 10.1021/ja053129n. [DOI] [PubMed] [Google Scholar]
  • [60].Gurtovenko AA, Vattulainen I. Effect of NaCl and KCl on phosphatidylcholine and phosphatidylethanolamine lipid membranes: insight from atomic-scale simulations for understanding salt-induced effects in the plasma membrane. J Phys Chem B. 2008;112:1953–62. doi: 10.1021/jp0750708. [DOI] [PubMed] [Google Scholar]
  • [61].Haile JM. Molecular dynamics simulation : elementary methods. Wiley; New York: 1992. [Google Scholar]
  • [62].Han BG, Guliaev AB, Walian PJ, Jap BK. Water transport in AQP0 aquaporin: molecular dynamics studies. J Mol Biol. 2006;360:285–96. doi: 10.1016/j.jmb.2006.04.039. [DOI] [PubMed] [Google Scholar]
  • [63].Hashido M, Kidera A, Ikeguchi M. Water transport in aquaporins: osmotic permeability matrix analysis of molecular dynamics simulations. Biophys J. 2007;93:373–85. doi: 10.1529/biophysj.106.101170. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [64].Hayashi S, Tajkhorshid E, Schulten K. Molecular dynamics simulation of bacteriorhodopsin's photoisomerization using ab initio forces for the excited chromophore. Biophys J. 2003;85:1440–9. doi: 10.1016/S0006-3495(03)74576-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [65].Hayes JM, Greer JC, Morton-Blake DA. A force-field description of short-range repulsions for high density alkane molecular dynamics simulations. J Comput Chem. 2004;25:1953–66. doi: 10.1002/jcc.20116. [DOI] [PubMed] [Google Scholar]
  • [66].Hill TL. An introduction to statistical thermodynamics. Addison-Wesley Pub. Co.; Reading, Mass.: 1960. [Google Scholar]
  • [67].Hockney RW, Eastwood JW. Computer simulations using particles. McGraw Hill; London: 1981. [Google Scholar]
  • [68].Hofsass C, Lindahl E, Edholm O. Molecular dynamics simulations of phospholipid bilayers with cholesterol. Biophys J. 2003;84:2192–206. doi: 10.1016/S0006-3495(03)75025-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [69].Hogberg CJ, Nikitin AM, Lyubartsev AP. Modification of the CHARMM force field for DMPC lipid bilayer. J Comput Chem. 2008;29:2359–69. doi: 10.1002/jcc.20974. [DOI] [PubMed] [Google Scholar]
  • [70].Hristova K, White SH. Determination of the hydrocarbon core structure of fluid dioleoylphosphocholine (DOPC) bilayers by x-ray diffraction using specific bromination of the double-bonds: effect of hydration. Biophys J. 1998;74:2419–33. doi: 10.1016/S0006-3495(98)77950-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [71].Huang P, Bertaccini E, Loew GH. Molecular dynamics simulation of anesthetic-phospholipid bilayer interactions. J Biomol Struct Dyn. 1995;12:725–54. doi: 10.1080/07391102.1995.10508773. [DOI] [PubMed] [Google Scholar]
  • [72].Huang P, Loew GH. Interaction of an amphiphilic peptide with a phospholipid bilayer surface by molecular dynamics simulation study. J Biomol Struct Dyn. 1995;12:937–56. doi: 10.1080/07391102.1995.10508789. [DOI] [PubMed] [Google Scholar]
  • [73].Hub JS, Grubmuller H, de Groot BL. Dynamics and energetics of permeation through aquaporins. What do we learn from molecular dynamics simulations? Handb Exp Pharmacol. 2009:57–76. doi: 10.1007/978-3-540-79885-9_3. [DOI] [PubMed] [Google Scholar]
  • [74].Iftimie R, Minary P, Tuckerman ME. Ab initio molecular dynamics: concepts, recent developments, and future trends. Proc Natl Acad Sci U S A. 2005;102:6654–9. doi: 10.1073/pnas.0500193102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [75].Illya G, Lipowsky R, Shillcock JC. Two-component membrane material properties and domain formation from dissipative particle dynamics. J Chem Phys. 2006;125:114710. doi: 10.1063/1.2353114. [DOI] [PubMed] [Google Scholar]
  • [76].Isralewitz B, Baudry J, Gullingsrud J, Kosztin D, Schulten K. Steered molecular dynamics investigations of protein function. J Mol Graph Model. 2001;19:13–25. doi: 10.1016/s1093-3263(00)00133-9. [DOI] [PubMed] [Google Scholar]
  • [77].Izvekov S, Parrinello M, Burnham CJ, Voth GA. Effective force fields for condensed phase systems from ab initio molecular dynamics simulation: a new method for force-matching. J Chem Phys. 2004;120:10896–913. doi: 10.1063/1.1739396. [DOI] [PubMed] [Google Scholar]
  • [78].Izvekov S, Voth GA. Effective force field for liquid hydrogen fluoride from ab initio molecular dynamics simulation using the force-matching method. J Phys Chem B. 2005;109:6573–86. doi: 10.1021/jp0456685. [DOI] [PubMed] [Google Scholar]
  • [79].Jin B, Hopfinger AJ. Characterization of lipid membrane dynamics by simulation: 3. Probing molecular transport across the phospholipid bilayer. Pharm Res. 1996;13:1786–94. doi: 10.1023/a:1016016621185. [DOI] [PubMed] [Google Scholar]
  • [80].Jorgensen WL, Tirado-Rives J. Molecular modeling of organic and biomolecular systems using BOSS and MCPRO. J Comput Chem. 2005;26:1689–700. doi: 10.1002/jcc.20297. [DOI] [PubMed] [Google Scholar]
  • [81].Kalko SG, Hernandez JA, Grigera JR, Fischbarg J. Osmotic permeability in a molecular dynamics simulation of water transport through a single-occupancy pore. Biochim Biophys Acta. 1995;1240:159–66. doi: 10.1016/0005-2736(95)00187-5. [DOI] [PubMed] [Google Scholar]
  • [82].Karplus M, Petsko GA. Molecular dynamics simulations in biology. Nature. 1990;347:631–9. doi: 10.1038/347631a0. [DOI] [PubMed] [Google Scholar]
  • [83].Karplus M, McCammon JA. Molecular dynamics simulations of biomolecules. Nat Struct Biol. 2002;9:646–52. doi: 10.1038/nsb0902-646. [DOI] [PubMed] [Google Scholar]
  • [84].Khelashvili GA, Scott HL. Combined Monte Carlo and molecular dynamics simulation of hydrated 18:0 sphingomyelin-cholesterol lipid bilayers. J Chem Phys. 2004;120:9841–7. doi: 10.1063/1.1724814. [DOI] [PubMed] [Google Scholar]
  • [85].Klauda JB, Wu X, Pastor RW, Brooks BR. Long-range Lennard-Jones and electrostatic interactions in interfaces: application of the isotropic periodic sum method. J Phys Chem B. 2007;111:4393–400. doi: 10.1021/jp068767m. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [86].Kornilov VV, Rabinovich AL, Balabaev NK, Bessonov VV. Effect of cholesterol on the structure and dynamic properties of unsaturated phospholipid bilayers. Biofizika. 2008;53:84–92. [PubMed] [Google Scholar]
  • [87].Kucerka N, Perlmutter JD, Pan J, Tristram-Nagle S, Katsaras J, Sachs JN. The effect of cholesterol on short- and long-chain monounsaturated lipid bilayers as determined by molecular dynamics simulations and X-ray scattering. Biophys J. 2008;95:2792–805. doi: 10.1529/biophysj.107.122465. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [88].Laurendeau NM. Statistical thermodynamics : fundamentals and applications. Cambridge University Press; New York: 2005. [Google Scholar]
  • [89].Leach AR. Molecular modelling : principles and applications. Prentice Hall; Harlow, England ; New York: 2001. [Google Scholar]
  • [90].Levy GC, Craik DJ, Kumar A, London RE. A critical evaluation of models for complex molecular dynamics: application of NMR studies of double- and single-stranded DNA. Biopolymers. 1983;22:2703–26. doi: 10.1002/bip.360221214. [DOI] [PubMed] [Google Scholar]
  • [91].Li DW, Liu XY. Examination of membrane fusion by dissipative particle dynamics simulation and comparison with continuum elastic models. J Chem Phys. 2005;122:174909. doi: 10.1063/1.1889433. [DOI] [PubMed] [Google Scholar]
  • [92].Lindahl E, Edholm O. Mesoscopic undulations and thickness fluctuations in lipid bilayers from molecular dynamics simulations. Biophys J. 2000;79:426–33. doi: 10.1016/S0006-3495(00)76304-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [93].Lindahl E, Hess B, Spoel D.v.d. GROMACS 3.0: A package for molecular simulation and trajectory analysis. J. Mol. Mod. 2001;7:306–317. [Google Scholar]
  • [94].Lindahl E, Sansom MS. Membrane proteins: molecular dynamics simulations. Curr Opin Struct Biol. 2008;18:425–31. doi: 10.1016/j.sbi.2008.02.003. [DOI] [PubMed] [Google Scholar]
  • [95].Lister JD. Stability of lipid bilayers and red blood cell membranes. Phys. Lett. 1975;53A:193–194. [Google Scholar]
  • [96].Lopez Cascales JJ, Garcia de la Torre J. Effect of lithium and sodium ions on a charged membrane of dipalmitoylphosphatidylserine: a study by molecular dynamics simulation. Biochim Biophys Acta. 1997;1330:145–56. doi: 10.1016/s0005-2736(97)00156-9. [DOI] [PubMed] [Google Scholar]
  • [97].Lopez Cascales JJ, Huertas ML, Garcia de la Torre J. Molecular dynamics simulation of a dye molecule in the interior of a bilayer: 1,6-diphenyl-1,3,5-hexatriene in dipalmitoylphosphatidylcholine. Biophys Chem. 1997;69:1–8. doi: 10.1016/s0301-4622(96)02267-3. [DOI] [PubMed] [Google Scholar]
  • [98].Lustig R. Statistical thermodynamics in the classical molecular dynamics ensemble. I. Fundamentals. J. Chem. Phys. 1994;100:3048–3059. [Google Scholar]
  • [99].MacKerell AD, Jr., Banavali N, Foloppe N. Development and current status of the CHARMM force field for nucleic acids. Biopolymers. 2000;56:257–65. doi: 10.1002/1097-0282(2000)56:4<257::AID-BIP10029>3.0.CO;2-W. [DOI] [PubMed] [Google Scholar]
  • [100].Markov AA, Nagorny NM. Theory of Algorithms. Springer-Verlag; London: 1988. [Google Scholar]
  • [101].Marrink J, M. AE, Tieleman DP. Molecular dynamics simulations of the kinetics of spontaneous micelle formation. J. Phys. Chem. B. 2000;104:12165–12173. [Google Scholar]
  • [102].Marrink SJ, Berendsen HJC. Simulation of water transport through a lipid-membrane. J. Phys. Chem. 1994;98:4155–4168. [Google Scholar]
  • [103].Marrink SJ, Berendsen HJC. Permeation process of small molecules across lipid membranes studied by molecular dynamics simulations. J. Phys. Chem. 1996;100:16729–16738. [Google Scholar]
  • [104].Marrink SJ, Lindahl E, Edholm O, Mark AE. Simulation of the spontaneous aggregation of phospholipids into bilayers. J Am Chem Soc. 2001;123:8638–9. doi: 10.1021/ja0159618. [DOI] [PubMed] [Google Scholar]
  • [105].Marrink SJ, de Vries AH, Tieleman DP. Lipids on the move: Simulations of membrane pores, domains, stalks and curves. Biochim Biophys Acta. 2009;1788:149–68. doi: 10.1016/j.bbamem.2008.10.006. [DOI] [PubMed] [Google Scholar]
  • [106].Mashl RJ, Scott HL, Subramaniam S, Jakobsson E. Molecular simulation of dioleoylphosphatidylcholine lipid bilayers at differing levels of hydration. Biophys J. 2001;81:3005–15. doi: 10.1016/S0006-3495(01)75941-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [107].McCammon JA, Gelin BR, Karplus M. Dynamics of folded proteins. Nature. 1977;267:585–90. doi: 10.1038/267585a0. [DOI] [PubMed] [Google Scholar]
  • [108].McIntosh TJ, Simon SA. Contributions of hydration and steric (entropic) pressures to the interactions between phosphatidylcholine bilayers: experiments with the subgel phase. Biochemistry. 1993;32:8374–84. doi: 10.1021/bi00083a042. [DOI] [PubMed] [Google Scholar]
  • [109].McMillan WG, Mayer JE. The statistical thermodynamics of multicomponent systems. New York city: 1945. [Google Scholar]
  • [110].McQuarrie DA. Statistical thermodynamics. University Science Books; Mill Valley, Calif.: 1973. [Google Scholar]
  • [111].Merz KM., Jr. Molecular dynamics simulations of lipid bilayers. Curr Opin Struct Biol. 1997;7:511–7. doi: 10.1016/s0959-440x(97)80115-7. [DOI] [PubMed] [Google Scholar]
  • [112].Meyn SP, Tweedie RL. Markov Chains and Stochastic Stability. Springer-Verlag; London: 1993. [Google Scholar]
  • [113].Milik M, Skolnick J. Insertion of peptide chains into lipid membranes: an off-lattice Monte Carlo dynamics model. Proteins. 1993;15:10–25. doi: 10.1002/prot.340150104. [DOI] [PubMed] [Google Scholar]
  • [114].Miller BT, Singh RP, Klauda JB, Hodoscek M, Brooks BR, Woodcock HL., 3rd CHARMMing: a new, flexible web portal for CHARMM. J Chem Inf Model. 2008;48:1920–9. doi: 10.1021/ci800133b. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [115].Miloshevsky GV, Jordan PC. Water and ion permeation in bAQP1 and GlpF channels: a kinetic Monte Carlo study. Biophys J. 2004;87:3690–702. doi: 10.1529/biophysj.104.043315. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [116].Moldovan D, Pinisetty D, Devireddy RV. Molecular dynamics simulation of pore growth in lipid bilayer membranes in the presence of edge-active agents. Appl. Phys. Lett. 2007;91:204104–204107. [Google Scholar]
  • [117].Montal M. Experimental membranes and mechanisms of bioenergy transductions. Annu Rev Biophys Bioeng. 1976;5:119–75. doi: 10.1146/annurev.bb.05.060176.001003. [DOI] [PubMed] [Google Scholar]
  • [118].Nagle JF, Tristram-Nagle S. Structure of lipid bilayers. Biochim Biophys Acta. 2000;1469:159–95. doi: 10.1016/s0304-4157(00)00016-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [119].Neamtu A, Suciu D. Modelling of ion permeation in K+ channels by nonequilibrium molecular dynamics simulations: I. Permeation energetics and structure stability. Rom J Physiol. 2004;41:69–82. [PubMed] [Google Scholar]
  • [120].Norberg J, Nilsson L. Advances in biomolecular simulations: methodology and recent applications. Q Rev Biophys. 2003;36:257–306. doi: 10.1017/s0033583503003895. [DOI] [PubMed] [Google Scholar]
  • [121].Norberto de Souza O, Ornstein RL. Molecular dynamics simulations of a protein-protein dimer: particle-mesh Ewald electrostatic model yields far superior results to standard cutoff model. J Biomol Struct Dyn. 1999;16:1205–18. doi: 10.1080/07391102.1999.10508328. [DOI] [PubMed] [Google Scholar]
  • [122].Notman R, den Otter WK, Noro MG, Briels WJ, Anwar J. The permeability enhancing mechanism of DMSO in ceramide bilayers simulated by molecular dynamics. Biophys J. 2007;93:2056–68. doi: 10.1529/biophysj.107.104703. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [123].Okulski W, Sujak A, Gruszecki WI. Dipalmitoylphosphatidylcholine membranes modified with carotenoid pigment lutein: experiment versus Monte Carlo simulation study of the membrane organization. Biochim Biophys Acta. 2008;1778:2105–18. doi: 10.1016/j.bbamem.2008.04.002. [DOI] [PubMed] [Google Scholar]
  • [124].Paci E, Marchi M. Membrane crossing by a polar molecule: A molecular dynamics simulation. Mol. Simul. 1994;14:1–10. [Google Scholar]
  • [125].Pandit SA, Berkowitz ML. Molecular dynamics simulation of dipalmitoylphosphatidylserine bilayer with Na+ counterions. Biophys J. 2002;82:1818–27. doi: 10.1016/S0006-3495(02)75532-X. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [126].Pandit SA, Vasudevan S, Chiu SW, Mashl RJ, Jakobsson E, Scott HL. Sphingomyelin-cholesterol domains in phospholipid membranes: atomistic simulation. Biophys J. 2004;87:1092–100. doi: 10.1529/biophysj.104.041939. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [127].Pasenkiewicz-Gierula M, Rog T, Kitamura K, Kusumi A. Cholesterol effects on the phosphatidylcholine bilayer polar region: a molecular simulation study. Biophys J. 2000;78:1376–89. doi: 10.1016/S0006-3495(00)76691-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [128].Pastor RW, Venable RM, Karplus M. Model for the structure of the lipid bilayer. Proc Natl Acad Sci U S A. 1991;88:892–6. doi: 10.1073/pnas.88.3.892. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [129].Patra M, Karttunen M, Hyvonen MT, Falck E, Lindqvist P, Vattulainen I. Molecular dynamics simulations of lipid bilayers: major artifacts due to truncating electrostatic interactions. Biophys J. 2003;84:3636–45. doi: 10.1016/S0006-3495(03)75094-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [130].Pereira CS, Hunenberger PH. Interaction of the sugars trehalose, maltose and glucose with a phospholipid bilayer: a comparative molecular dynamics study. J Phys Chem B. 2006;110:15572–81. doi: 10.1021/jp060789l. [DOI] [PubMed] [Google Scholar]
  • [131].Pereira CS, Hunenberger PH. Effect of trehalose on a phospholipid membrane under mechanical stress. Biophys J. 2008;95:3525–34. doi: 10.1529/biophysj.108.131656. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [132].Petrie T. Probabilistic Functions of Finite-State Markov Chains. Proc Natl Acad Sci U S A. 1967;57:580–581. doi: 10.1073/pnas.57.3.580. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [133].Phillips JC, Braun R, Wang W, Gumbart J, Tajkhorshid E, Villa E, Chipot C, Skeel RD, Kale L, Schulten K. Scalable molecular dynamics with NAMD. J Comput Chem. 2005;26:1781–802. doi: 10.1002/jcc.20289. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [134].Pica Ciamarra M, Coniglio A, Nicodemi M. Thermodynamics and statistical mechanics of dense granular media. Phys Rev Lett. 2006;97:158001. doi: 10.1103/PhysRevLett.97.158001. [DOI] [PubMed] [Google Scholar]
  • [135].Pinisetty D, Moldovan D, Devireddy R. The effect of methanol on lipid bilayers: an atomistic investigation. Ann Biomed Eng. 2006;34:1442–51. doi: 10.1007/s10439-006-9148-y. [DOI] [PubMed] [Google Scholar]
  • [136].Ploeg P.v.d., Berendsen HJC. Molecular dynamics simulation of a bilayer membrane. J. Chem. Phys. 1982;76:3271–3276. [Google Scholar]
  • [137].Ploeg P.v.d., Berendsen HJC. Molecular dynamics of a bilayer membrane. Mol. Phys. 1983;49:233–248. [Google Scholar]
  • [138].Polyansky AA, Volynsky PE, Nolde DE, Arseniev AS, Efremov RG. Role of lipid charge in organization of water/lipid bilayer interface: insights via computer simulations. J Phys Chem B. 2005;109:15052–9. doi: 10.1021/jp0510185. [DOI] [PubMed] [Google Scholar]
  • [139].Rao BG, Singh UC. A free-energy perturbation study of solvation in methanol and dimethylsulfoxide. J. Am. Chem. Soc. 1990;112:3803–3811. [Google Scholar]
  • [140].Rog T, Pasenkiewicz-Gierula M. Cholesterol effects on the phospholipid condensation and packing in the bilayer: a molecular simulation study. FEBS Lett. 2001;502:68–71. doi: 10.1016/s0014-5793(01)02668-0. [DOI] [PubMed] [Google Scholar]
  • [141].Rog T, Pasenkiewicz-Gierula M. Cholesterol effects on the phosphatidylcholine bilayer nonpolar region: a molecular simulation study. Biophys J. 2001;81:2190–202. doi: 10.1016/S0006-3495(01)75867-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [142].Rog T, Murzyn K, Pasenkiewicz-Gierula M. Molecular dynamics simulations of charged and neutral lipid bilayers: treatment of electrostatic interactions. Acta Biochim Pol. 2003;50:789–98. [PubMed] [Google Scholar]
  • [143].Rojas AV, Liwo A, Scheraga HA. Molecular dynamics with the United-residue force field: ab initio folding simulations of multichain proteins. J Phys Chem B. 2007;111:293–309. doi: 10.1021/jp065810x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [144].Roy D, Mukhopadhyay C. GD1a in phospholipid bilayer: a molecular dynamics simulation. J Biomol Struct Dyn. 2001;18:639–46. doi: 10.1080/07391102.2001.10506695. [DOI] [PubMed] [Google Scholar]
  • [145].Roy D, Mukhopadhyay C. Molecular dynamics simulation of GM1 in phospholipid bilayer. J Biomol Struct Dyn. 2002;19:1121–32. doi: 10.1080/07391102.2002.10506815. [DOI] [PubMed] [Google Scholar]
  • [146].Saiz L, Klein ML. Structural properties of a highly polyunsaturated lipid bilayer from molecular dynamics simulations. Biophys J. 2001;81:204–16. doi: 10.1016/S0006-3495(01)75692-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [147].Saiz L, Bandyopadhyay S, Klein ML. Towards an understanding of complex biological membranes from atomistic molecular dynamics simulations. Biosci Rep. 2002;22:151–73. doi: 10.1023/a:1020130420869. [DOI] [PubMed] [Google Scholar]
  • [148].Saiz L, Klein ML. Electrostatic interactions in neutral model phospholipid bilayer by molecular dynamics simulations. J. Chem. Phys. 2002;116:3052–3057. [Google Scholar]
  • [149].Saiz L, Klein ML. Computer simulation studies of model biological membranes. Acc Chem Res. 2002;35:482–9. doi: 10.1021/ar010167c. [DOI] [PubMed] [Google Scholar]
  • [150].Scheidt HA, Huster D. The interaction of small molecules with phospholipid membranes studied by 1H NOESY NMR under magic-angle spinning. Acta Pharmacol Sin. 2008;29:35–49. doi: 10.1111/j.1745-7254.2008.00726.x. [DOI] [PubMed] [Google Scholar]
  • [151].Schrödinger E, Dublin Institute for Advanced Studies . School of Theoretical Physics., Statistical thermodynamics; a course of seminar lectures delivered in January-March 1944, at the School of Theoretical Physics. Dublin Institute for Advanced Studies, University Press; Cambridge [Eng.]: 1946. [Google Scholar]
  • [152].Scott HL, Kalaskar S. Lipid chains and cholesterol in model membranes: a Monte Carlo Study. Biochemistry. 1989;28:3687–91. doi: 10.1021/bi00435a010. [DOI] [PubMed] [Google Scholar]
  • [153].Seelig J. Deuterium magnetic resonance: theory and application to lipid membranes. Q Rev Biophys. 1977;10:353–418. doi: 10.1017/s0033583500002948. [DOI] [PubMed] [Google Scholar]
  • [154].Shillcock JC, Boal DH. Entropy-driven instability and rupture of fluid membranes. Biophys J. 1996;71:317–26. doi: 10.1016/S0006-3495(96)79227-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [155].Shrivastava IH, Sansom MS. Simulations of ion permeation through a potassium channel: molecular dynamics of KcsA in a phospholipid bilayer. Biophys J. 2000;78:557–70. doi: 10.1016/S0006-3495(00)76616-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [156].Singer SJ, Nicolson GL. The fluid mosaic model of the structure of cell membranes. Science. 1972;175:720–31. doi: 10.1126/science.175.4023.720. [DOI] [PubMed] [Google Scholar]
  • [157].Smondyrev AM, Berkowitz ML. Molecular dynamics simulation of DPPC bilayer in DMSO. Biophys J. 1999;76:2472–8. doi: 10.1016/S0006-3495(99)77402-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [158].Smondyrev AM, Berkowitz ML. Structure of dipalmitoylphosphatidylcholine/cholesterol bilayer at low and high cholesterol concentrations: molecular dynamics simulation. Biophys J. 1999;77:2075–89. doi: 10.1016/S0006-3495(99)77049-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [159].Söderhäll JA, Laaksonen A. Molecular dynamics simulation of ubiquinone inside a lipid bilayer. J. Phys. Chem. B. 2001;105:9308–9315. [Google Scholar]
  • [160].Sonntag RE, Van Wylen GJ. Fundamentals of statistical thermodynamics. Wiley; New York: 1966. [Google Scholar]
  • [161].Stillinger FH, Rahman A. Improved simulation of liquid water by molecular dynamics. J. Chem. Phys. 1974;60:1545–1557. [Google Scholar]
  • [162].Sugar IP, Biltonen RL. Lateral diffusion of molecules in two-component lipid bilayer: a Monte Carlo simulation study. J Phys Chem B. 2005;109:7373–86. doi: 10.1021/jp045669x. [DOI] [PubMed] [Google Scholar]
  • [163].Sum AK, Faller R, de Pablo JJ. Molecular simulation study of phospholipid bilayers and insights of the interactions with disaccharides. Biophys J. 2003;85:2830–44. doi: 10.1016/s0006-3495(03)74706-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [164].ter Laak AM, Kuhne R. Bacteriorhodopsin in a periodic boundary water-vacuum-water box as an example towards stable molecular dynamics simulations of G-protein coupled receptors. Receptors Channels. 1999;6:295–308. [PubMed] [Google Scholar]
  • [165].Tieleman DP, Marrink SJ, Berendsen HJ. A computer perspective of membranes: molecular dynamics studies of lipid bilayer systems. Biochim Biophys Acta. 1997;1331:235–70. doi: 10.1016/s0304-4157(97)00008-7. [DOI] [PubMed] [Google Scholar]
  • [166].Tieleman DP, Hess B, Sansom MS. Analysis and evaluation of channel models: simulations of alamethicin. Biophys J. 2002;83:2393–407. doi: 10.1016/s0006-3495(02)75253-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [167].Tolpekina TV, den Otter WK, Briels WJ. Simulations of stable pores in membranes: system size dependence and line tension. J Chem Phys. 2004;121:8014–20. doi: 10.1063/1.1796254. [DOI] [PubMed] [Google Scholar]
  • [168].Tristram-Nagle S, Nagle JF. Lipid bilayers: thermodynamics, structure, fluctuations, and interactions. Chem Phys Lipids. 2004;127:3–14. doi: 10.1016/j.chemphyslip.2003.09.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [169].Tse JS. Ab initio molecular dynamics with density functional theory. Annu Rev Phys Chem. 2002;53:249–90. doi: 10.1146/annurev.physchem.53.090401.105737. [DOI] [PubMed] [Google Scholar]
  • [170].Tsui V, Radhakrishnan I, Wright PE, Case DA. NMR and molecular dynamics studies of the hydration of a zinc finger-DNA complex. J Mol Biol. 2000;302:1101–17. doi: 10.1006/jmbi.2000.4108. [DOI] [PubMed] [Google Scholar]
  • [171].Tu K, Tobias DJ, Blasie JK, Klein ML. Molecular dynamics investigation of the structure of a fully hydrated gel-phase dipalmitoylphosphatidylcholine bilayer. Biophys J. 1996;70:595–608. doi: 10.1016/S0006-3495(96)79623-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [172].Tu K, Klein ML, Tobias DJ. Constant-pressure molecular dynamics investigation of cholesterol effects in a dipalmitoylphosphatidylcholine bilayer. Biophys J. 1998;75:2147–56. doi: 10.1016/S0006-3495(98)77657-X. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [173].Ulrich AS, Watts A. Molecular response of the lipid headgroup to bilayer hydration monitored by 2H-NMR. Biophys J. 1994;66:1441–9. doi: 10.1016/S0006-3495(94)80934-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [174].Van Der Spoel D, Lindahl E, Hess B, Groenhof G, Mark AE, Berendsen HJ. GROMACS: fast, flexible, and free. J Comput Chem. 2005;26:1701–18. doi: 10.1002/jcc.20291. [DOI] [PubMed] [Google Scholar]
  • [175].Venable RM, Brooks BR, Pastor RW. Molecular dynamics simulations of gel phase lipid bilayers in constant pressure and constant surface area ensembles. J. Chem. Phys. 2000;112:4822–4832. [Google Scholar]
  • [176].Watanabe MS. Percolation with a periodic boundary condition: The effect of system size for crystallization in molecular dynamics. Phys Rev E Stat Phys Plasmas Fluids Relat Interdiscip Topics. 1995;51:3945–3951. doi: 10.1103/physreve.51.3945. [DOI] [PubMed] [Google Scholar]
  • [177].Wilson MA, Pohorille A. Mechanism of unassisted ion transport across membrane bilayers. J Am Chem Soc. 1996;118:6580–7. doi: 10.1021/ja9540381. [DOI] [PubMed] [Google Scholar]
  • [178].Wohlert J, Edholm O. The range and shielding of dipole-dipole interactions in phospholipid bilayers. Biophys J. 2004;87:2433–45. doi: 10.1529/biophysj.104.044222. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [179].Yamashita Y, Kinoshita K, Yamazaki M. Low concentration of DMSO stabilizes the bilayer gel phase rather than the interdigitated gel phase in dihexadecylphosphatidylcholine membrane. Biochim Biophys Acta. 2000;1467:395–405. doi: 10.1016/s0005-2736(00)00237-6. [DOI] [PubMed] [Google Scholar]
  • [180].Yeh IC, Berkowitz ML. Ewald summation for systems with slab geometry. J. Chem. Phys. 1999;18:1463–1472. [Google Scholar]
  • [181].Yoshii N, Okazaki S. Free energy of water permeation into hydrophobic core of sodium dodecyl sulfate micelle by molecular dynamics calculation. J Chem Phys. 2007;126:096101. doi: 10.1063/1.2464000. [DOI] [PubMed] [Google Scholar]
  • [182].Yu ZW, Quinn PJ. The effect of dimethyl sulphoxide on the structure and phase behaviour of palmitoleoylphosphatidylethanolamine. Biochim Biophys Acta. 2000;1509:440–50. doi: 10.1016/s0005-2736(00)00326-6. [DOI] [PubMed] [Google Scholar]
  • [183].Zaman MH. Statistical mechanics of cellular systems and processes. Cambridge University Press; Cambridge; New York: 2009. [Google Scholar]
  • [184].Zhang Z, Davis HT, Kroll DM. Molecular dynamics simulations of tethered membranes with periodic boundary conditions. Phys Rev E Stat Phys Plasmas Fluids Relat Interdiscip Topics. 1996;53:1422–1429. doi: 10.1103/physreve.53.1422. [DOI] [PubMed] [Google Scholar]
  • [185].Zheng Z, Stroumpoulis D, Parra A, Petzold L, Tirrell M. A Monte Carlo simulation study of lipid bilayer formation on hydrophilic substrates from vesicle solutions. J Chem Phys. 2006;124:64904. doi: 10.1063/1.2166392. [DOI] [PubMed] [Google Scholar]
  • [186].Zhou Z, Sayer BG, Hughes DW, Stark RE, Epand RM. Studies of phospholipid hydration by high-resolution magic-angle spinning nuclear magnetic resonance. Biophys J. 1999;76:387–99. doi: 10.1016/S0006-3495(99)77205-X. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [187].Zhu F, Tajkhorshid E, Schulten K. Molecular dynamics study of aquaporin-1 water channel in a lipid bilayer. FEBS Lett. 2001;504:212–8. doi: 10.1016/s0014-5793(01)02749-1. [DOI] [PubMed] [Google Scholar]
  • [188].Zhu F, Tajkhorshid E, Schulten K. Pressure-induced water transport in membrane channels studied by molecular dynamics. Biophys J. 2002;83:154–60. doi: 10.1016/S0006-3495(02)75157-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [189].Ziegler MJ, Vernier PT. Interface water dynamics and porating electric fields for phospholipid bilayers. J Phys Chem B. 2008;112:13588–96. doi: 10.1021/jp8027726. [DOI] [PubMed] [Google Scholar]
  • [190].Zubrzycki IZ. Molecular dynamics study on an adipokinetic hormone peptide in aqueous solution. Z Naturforsch [C] 2000;55:125–8. doi: 10.1515/znc-2000-1-223. [DOI] [PubMed] [Google Scholar]

RESOURCES