+ All Categories
Home > Documents > arXiv:1010.0389v1 [cond-mat.soft] 3 Oct 2010arXiv:1010.0389v1 [cond-mat.soft] 3 Oct 2010...

arXiv:1010.0389v1 [cond-mat.soft] 3 Oct 2010arXiv:1010.0389v1 [cond-mat.soft] 3 Oct 2010...

Date post: 23-Jun-2020
Category:
Upload: others
View: 8 times
Download: 0 times
Share this document with a friend
13
arXiv:1010.0389v1 [cond-mat.soft] 3 Oct 2010 Solvent-free coarse-grained lipid model for large-scale simulations Hiroshi Noguchi Institute for Solid State Physics, University of Tokyo, Kashiwa, Chiba 277-8581, Japan (Dated: October 23, 2018) A coarse-grained molecular model, which consists of a spherical particle and an orientation vector, is proposed to simulate lipid membrane on a large length scale. The solvent is implicitly represented by an effective attractive interaction between particles. A bilayer structure is formed by orientation- dependent (tilt and bending) potentials. In this model, the membrane properties (bending rigidity, line tension of membrane edge, area compression modulus, lateral diffusion coefficient, and flip-flop rate) can be varied over broad ranges. The stability of the bilayer membrane is investigated via droplet-vesicle transition. The rupture of the bilayer and worm-like micelle formation can be induced by an increase in the spontaneous curvature of the monolayer membrane. PACS numbers: I. INTRODUCTION Amphiphilic molecules, such as lipids and detergents, self-assemble into various structures depending on the relative size of their hydrophilic parts: spherical or worm-like micelles, bilayer membranes, inverted hexago- nal structures, and inverted micelles. Among these struc- tures, the bilayer membrane of phospholipids has been in- tensively investigated, since it is the basic structure of the plasma membrane and the intracellular compartments of living cells, where the membranes are in a fluid phase and lipid molecules can diffuse in quasi-two-dimensional space. A vesicle (closed membrane) is considered to be a simple model of cells and it has applications in drug- delivery systems as a drug carrier. Bilayer membranes exhibit many interesting phenom- ena such as shape deformation induced by phase sepa- ration or chemical reaction, membrane fusion, and mem- brane fission. The length scale of these phenomena varies from nm to μm, since cells are 10μm in diameter, whereas the thickness of a biomembrane is 5nm. To investigate the morphologies of cells and vesicles, the molecular structure is assumed to be negligible, and the bilayer membrane is described as a smoothly-curved mathematical surface 1–3 . The information about the bi- layer properties is only reflected in the values of the elas- tic parameters. To simulate the membrane with thermal fluctuations, a triangulated surface is widely used 4,5 . An alternative model is a meshless membrane 6–13 , where par- ticles self-assemble into a membrane by anisotripic poten- tial interactions. These models can reproduce μm-scale dynamics of the bilayer membrane well but cannot treat a non-bilayer structure such as the stalk structure of a membrane fusion intermediate 14,15 . To simulate molecular-scale dynamics and the non- bilayer structure, a molecular model is required. Al- though computer technology has grown rapidly, the typ- ical scale for recent simulations of the all-atom models is only 100 ns dynamics of hundreds of lipid molecules. To simulate the membranes on longer and larger scales, vari- ous coarse-grained molecular models have been proposed (see review articles 16–20 ). Recently, the potential param- eters in some of the coarse-grained molecular models are tuned by atomistic simulations 21–25 . In mapping of inter- action parameters, one coarse-grained particle typically represents three or four heavy atoms and their accompa- nying hydrogen atoms. To further reduce the computa- tional costs, larger segments (three or more segment par- ticles per amphiphilic molecule) are employed, and the solvent is implicitly represented by an effective attrac- tive potential between the hydrophobic segments 19,26–30 . Model parameters are chosen to generate a bilayer mem- brane with reasonably realistic values of elastic proper- ties. In this scale, it is difficult to take into account chemical details of lipids, such as an unsaturated bond in hydrocarbon chains. Instead, this type of models can be advantageous to capture the general features in the bilayer membrane, since the simplicity of the model can allow for wide ranges of variation of the membrane prop- erties. In this paper, we propose a solvent-free molecular model to pursue two purposes: (1) to represent the am- phiphilic molecule in a size as small as possible and (2) to allow the variation in the membrane properties for wide ranges. A molecule consisting of many particles has a higher resolution than that with less particles but re- quires a smaller length unit and time step for simulations. Here, we consider a molecule that consists of a spherical particle and an orientation vector. It can reduce compu- tational costs to simulate many molecules. In previous solvent-free models 19,25–30 , the membrane properties are varied only in narrow ranges. On the other hand, in one of the meshless membrane models, the bending rigidity and the line tension of the membrane edge can be inde- pendently varied over wide ranges 7 . This allows the con- ditions of vesicle formation and rupture to be controlled 8 . In addition, it is easy to compare the simulation results with theoretical predictions. Such tuning capability is desired for molecular models. In Sec. II, the lipid model and the simulation method are described. In Sec. III, the results and discussion are provided. The formation of a membrane and its sta- bility for droplet-vesicle transitions are described in Sec. III A. In Sec. III B and III C, the calculation methods
Transcript
Page 1: arXiv:1010.0389v1 [cond-mat.soft] 3 Oct 2010arXiv:1010.0389v1 [cond-mat.soft] 3 Oct 2010 Solvent-free coarse-grained lipid model forlarge-scale simulations Hiroshi Noguchi∗ Institute

arX

iv:1

010.

0389

v1 [

cond

-mat

.sof

t] 3

Oct

201

0

Solvent-free coarse-grained lipid model for large-scale simulations

Hiroshi Noguchi∗

Institute for Solid State Physics, University of Tokyo, Kashiwa, Chiba 277-8581, Japan

(Dated: October 23, 2018)

A coarse-grained molecular model, which consists of a spherical particle and an orientation vector,is proposed to simulate lipid membrane on a large length scale. The solvent is implicitly representedby an effective attractive interaction between particles. A bilayer structure is formed by orientation-dependent (tilt and bending) potentials. In this model, the membrane properties (bending rigidity,line tension of membrane edge, area compression modulus, lateral diffusion coefficient, and flip-floprate) can be varied over broad ranges. The stability of the bilayer membrane is investigated viadroplet-vesicle transition. The rupture of the bilayer and worm-like micelle formation can be inducedby an increase in the spontaneous curvature of the monolayer membrane.

PACS numbers:

I. INTRODUCTION

Amphiphilic molecules, such as lipids and detergents,self-assemble into various structures depending on therelative size of their hydrophilic parts: spherical orworm-like micelles, bilayer membranes, inverted hexago-nal structures, and inverted micelles. Among these struc-tures, the bilayer membrane of phospholipids has been in-tensively investigated, since it is the basic structure of theplasma membrane and the intracellular compartments ofliving cells, where the membranes are in a fluid phaseand lipid molecules can diffuse in quasi-two-dimensionalspace. A vesicle (closed membrane) is considered to bea simple model of cells and it has applications in drug-delivery systems as a drug carrier.

Bilayer membranes exhibit many interesting phenom-ena such as shape deformation induced by phase sepa-ration or chemical reaction, membrane fusion, and mem-brane fission. The length scale of these phenomena variesfrom nm to µm, since cells are ∼ 10µm in diameter,whereas the thickness of a biomembrane is 5nm. Toinvestigate the morphologies of cells and vesicles, themolecular structure is assumed to be negligible, andthe bilayer membrane is described as a smoothly-curvedmathematical surface1–3. The information about the bi-layer properties is only reflected in the values of the elas-tic parameters. To simulate the membrane with thermalfluctuations, a triangulated surface is widely used4,5. Analternative model is a meshless membrane6–13, where par-ticles self-assemble into a membrane by anisotripic poten-tial interactions. These models can reproduce µm-scaledynamics of the bilayer membrane well but cannot treata non-bilayer structure such as the stalk structure of amembrane fusion intermediate14,15.

To simulate molecular-scale dynamics and the non-bilayer structure, a molecular model is required. Al-though computer technology has grown rapidly, the typ-ical scale for recent simulations of the all-atom models isonly 100 ns dynamics of hundreds of lipid molecules. Tosimulate the membranes on longer and larger scales, vari-ous coarse-grained molecular models have been proposed(see review articles16–20). Recently, the potential param-

eters in some of the coarse-grained molecular models aretuned by atomistic simulations21–25. In mapping of inter-action parameters, one coarse-grained particle typicallyrepresents three or four heavy atoms and their accompa-nying hydrogen atoms. To further reduce the computa-tional costs, larger segments (three or more segment par-ticles per amphiphilic molecule) are employed, and thesolvent is implicitly represented by an effective attrac-tive potential between the hydrophobic segments19,26–30.Model parameters are chosen to generate a bilayer mem-brane with reasonably realistic values of elastic proper-ties. In this scale, it is difficult to take into accountchemical details of lipids, such as an unsaturated bondin hydrocarbon chains. Instead, this type of models canbe advantageous to capture the general features in thebilayer membrane, since the simplicity of the model canallow for wide ranges of variation of the membrane prop-erties.

In this paper, we propose a solvent-free molecularmodel to pursue two purposes: (1) to represent the am-phiphilic molecule in a size as small as possible and (2)to allow the variation in the membrane properties forwide ranges. A molecule consisting of many particles hasa higher resolution than that with less particles but re-quires a smaller length unit and time step for simulations.Here, we consider a molecule that consists of a sphericalparticle and an orientation vector. It can reduce compu-tational costs to simulate many molecules. In previoussolvent-free models19,25–30, the membrane properties arevaried only in narrow ranges. On the other hand, in oneof the meshless membrane models, the bending rigidityand the line tension of the membrane edge can be inde-pendently varied over wide ranges7. This allows the con-ditions of vesicle formation and rupture to be controlled8.In addition, it is easy to compare the simulation resultswith theoretical predictions. Such tuning capability isdesired for molecular models.

In Sec. II, the lipid model and the simulation methodare described. In Sec. III, the results and discussionare provided. The formation of a membrane and its sta-bility for droplet-vesicle transitions are described in Sec.III A. In Sec. III B and III C, the calculation methods

Page 2: arXiv:1010.0389v1 [cond-mat.soft] 3 Oct 2010arXiv:1010.0389v1 [cond-mat.soft] 3 Oct 2010 Solvent-free coarse-grained lipid model forlarge-scale simulations Hiroshi Noguchi∗ Institute

2

0

0.5

1

0 1 2 3ri,j/σ

wcv

Urep

fcut

FIG. 1: (Color online) The cutoff function fcut(rij), the com-pact Gaussian weight function wcv(rij), and the repulsive po-tential Urep(rij).

and the dependence of the static and dynamic propertieson model parameters are described, respectively. Thesummary is given in Sec. IV.

II. MODEL AND METHOD

A. Molecular model

In solvent-free lipid models, an amphiphilic molecule istypically represented by three or more particles19. Here,we consider a lipid molecule with minimum (5) degreesof freedom for solvent-free molecular simulations. Each(i-th) molecule has a spherical particle with an orienta-tion vector ui, which represents a direction from the hy-drophobic to the hydrophilic part. There are two pointsof interaction in the molecule: the center of a sphere rsiand a hydrophilic point rei = rsi + ui. The moleculesinteract with each other via the potential,

U

kBT=

i<j

Urep(rsi,j) + ε

i

Uatt(ρi) (1)

+ktilt2

i<j

[

(ui · rsij)2 + (uj · rsij)2]

wcv(reij)

+kbend2

i<j

(

ui − uj − Cbdrsij

)2

wcv(reij),

where ri,j = ri− rj , ri,j = |ri,j |, ri,j = ri,j/ri,j , and kBTis the thermal energy. The molecules have an excludedvolume with a diameter σ via the repulsive potential,

Urep(r) = exp[−20(r/σ − 1)], (2)

with a cutoff at r = 2.4σ.The second term in Eq. (1) represents the attractive

interaction between the molecules. A multibody attrac-tive potential Uatt(ρi) is employed to mimic the “hy-drophobic” interaction. This potential allows the forma-tion of the fluid membrane over wide parameter ranges

and fast lateral diffusion. Similar potentials have beenapplied in the previous membrane models7,26,30 and acoarse-grained protein model31. The potential Uatt(ρi) isgiven by

Uatt(ρi) = 0.25 ln[1 + exp{−4(ρi − ρ∗)}]− C, (3)

where C = 0.25 ln{1 + exp(4ρ∗)} ≃ ρ∗ is chosen suchthat Uatt(0) = 0. The local particle density ρi is approx-imately the number of particles rsi in the sphere whoseradius is ratt.

ρi =∑

j 6=i

fcut(rsi,j), (4)

where fcut(r) is a C∞ cutoff function,

fcut(r) =

{

exp{A(1 + 1(r/rcut)n−1 )} (r < rcut)

0 (r ≥ rcut)(5)

with n = 6, A = ln(2){(rcut/ratt)n − 1}, ratt = 1.9σ(fcut(ratt) = 0.5), and the cutoff radius rcut = 2.4σ (seeFig. 1). The potential Uatt(ρi) acts as a pairwise at-tractive potential (Uatt(ρi) ≃ ρ, so that

i Uatt(ρi) ≃−2

i<j fcut(rsi,j)) for ρi < ρ∗−1 and approaches a con-

stant value (Uatt(ρi) ≃ ρ∗) for ρi > ρ∗ +1. It is assumedthat the hydrophobic parts have no contact with the im-plicit solvent (void space) at ρi & ρ∗.The third and fourth terms in Eq. (1) are dis-

cretized versions of tilt and bending potentials of the tiltmodel32,33, respectively. A smoothly truncated Gaussianfunction7 is employed as a weight function

wcv(r) =

{

exp((r/rga)

2

(r/rcc)n−1 ) (r < rcc)

0 (r ≥ rcc)(6)

with n = 4, rga = 1.5σ, and rcc = 3σ (see Fig. 1). Allorders of derivatives of fcut(r) and wmls(r) are continu-ous at the cutoff radii. The weight is a function of reij(not rsij) to avoid the interaction between the moleculesin the opposite monolayers of the bilayer. The averagedistance between the neighboring molecules in the samemonolayer is rnb ≃ 1.05σ, and the distance to the neigh-boring molecule in the other monolayer is reij ≃ 3σ (seeFig. 2(a)). Thus, these two potentials act between theneighboring molecules in the same monolayer but notbetween the monolayers. The tilt potential has the en-ergy minimum in a completely flat membrane with notilt deformation. Similar tilt potentials have been usedin the meshless membrane models6,9–13. The same typeof bending potential [the fourth term in Eq. (1)] waspreviously used to control the bending rigidity and thespontaneous curvature of the monolayer in the molecu-lar simulations27,30. Positive spontaneous curvature in-dicates that the hydrophilic head is larger than the hy-drophobic tail of amphiphilic molecules. The bendingrigidity is numerically calculated and compared with theestimation from the continuous description of the mem-brane in Sec. III C.

Page 3: arXiv:1010.0389v1 [cond-mat.soft] 3 Oct 2010arXiv:1010.0389v1 [cond-mat.soft] 3 Oct 2010 Solvent-free coarse-grained lipid model forlarge-scale simulations Hiroshi Noguchi∗ Institute

3

10-6

10-5

10-4

10-3

10-2

10-1

-1 0 1

P(u

z)

uz

(b)

ktilt = 1

2

48

0

0.01

0.02

-2 0 2

P(z

)

z/σ

(a)

zs ze

FIG. 2: (Color online) Probability distribution of (a) the po-sitions and (b) the orientation of the molecules in the planarmembrane at N = 512, ρ∗ = 14, ε = 2, kbend = ktilt, andCbd = 0. (a) The z components of rsi and rei are shown forktilt = 8. The dashed line represents rsi of the molecules atuz > 0. (b) The z component uz of the molecular orientationis shown for ktilt = 1, 2, 4, and 8. The error bars are displayedat several data points.

B. Brownian dynamics

We simulated the membrane in the NV T ensemble(constant number of molecules N , volume V , and tem-perature T ) with periodic boundary conditions in a boxwith side length Lx, Ly, and Lz. We employed Brown-ian dynamics (molecular dynamics with Langevin ther-mostat). The motions of the center of the mass rGi =(rsi + rei )/2 and the orientation ui are given by under-damped Langevin equations:

drGidt

= vGi ,

dui

dt= ωi, (7)

mdvG

i

dt= −ζGv

Gi + gG

i (t) + fGi , (8)

Idωi

dt= −ζrωi + (gr

i(t) + f ri )⊥ + λui, (9)

where m and I are the mass and the moment of iner-tia of the molecule, respectively. The forces are given byfGi = −∂U/∂rGi and f ri = −∂U/∂ui with the perpendic-ular component a⊥ = a− (a ·ui)ui and a Lagrange mul-tiplier λ to keep u2

i = 1. According to the fluctuation-dissipation theorem, the friction coefficients ζG, ζr andthe Gaussian white noises gG

i (t), gri(t) obey the follow-

ing relations: the average 〈gβ1

i,α1(t)〉 = 0 and the variance

〈gβ1

i,α1(t)gβ2

j,α2(t′)〉 = 2kBTζβ1

δijδα1α2δβ1β2

δ(t− t′), where

α1, α2 ∈ {x, y, z} and β1, β2 ∈ {G, r}. The Langevinequations are integrated by the leapfrog algorithm34 withvi,n ≡ vi(tn) = (vi(tn+1/2) + vi(tn−1/2))/2. First, thevelocities are updated by

vGi,n+1/2 = a0v

Gi,n−1/2 + a1(g

Gi,n + fGi,n), (10)

ω′i,n+1/2 = b0ωi,n−1/2 + b1(g

ri,n + f ri,n)

⊥ + λ′ui,n,

u′i,n+1/2 = ui,n + ω

′i,n+1/2∆t/2,

ωi,n+1/2 = ω′i,n+1/2 − (ω′

i,n+1/2 · u′i,n+1/2)u

′i,n+1/2,

where

a0 =1− ζG∆t/2m

1 + ζG∆t/2m, a1 =

∆t/m

1 + ζG∆t/2m, (11)

b0 =1− ζr∆t/2I

1 + ζr∆t/2I, b1 =

∆t/I

1 + ζr∆t/2I,

λ′ = λ∆t/I = −2ωi,n−1/2 · ui,n

1 + ζr∆t/2I,

gβ1

i,n = gβ1

i (tn)/√∆t.

Then, the positions are updated by

rGi,n+1 = rGi,n + vGi,n+1/2∆t, (12)

u′i,n+1 = ui,n + ωi,n+1/2∆t,

ui,n+1 = u′i,n+1/|u′

i,n+1|.

We employed m = 1, I = 1, ζG = 1, ζr = 1, kBT =1, ∆t = 0.005, and the total number of the moleculesN = 300 to 8192. The results are displayed with thelength unit σ, the energy unit kBT , and the time unitτ0 = ζGσ

2/kBT . The diffusion coefficientD is normalizedusing the diffusion coefficient D0 = σ2/τ0 of an isolatedmolecule. The error bars of the data are estimated fromthe standard deviations of three to six independent runs.

III. RESULTS AND DISCUSSION

A. Self-assembly and membrane stability

Molecules self-assemble into spherical droplets atkbend = ktilt = 0, i.e., when only the first two termsin Eq. (1) are taken into account. When the third term(the tilt potential) is added, the molecules can sponta-neously form vesicles. Figure 3 shows the self-assemblyof molecules from a random gas state. First, small clus-ters are formed; these clusters merge into disk-like mi-celles. Then, a large disk closes into a vesicle througha bowl-like shape (see Fig. 3(c)). Similar self-assemblyprocesses have been observed in the previous simulationsof molecular26 and meshless8 models.In order to clarify the stability of three-dimensional ag-

gregates and bilayer membranes, the morphologies of theaggregates are investigated as ktilt gradually increases ordecreases. As ktilt increases, a spherical liquid droplet

Page 4: arXiv:1010.0389v1 [cond-mat.soft] 3 Oct 2010arXiv:1010.0389v1 [cond-mat.soft] 3 Oct 2010 Solvent-free coarse-grained lipid model forlarge-scale simulations Hiroshi Noguchi∗ Institute

4

FIG. 3: (Color online) Sequential snapshots of the molecularself-assembly at N = 2000, Lx = Ly = Lz = 40σ, ε = 2,ρ∗ = 14, kbend = 0, ktilt = 8, and Cbd = 0. (a) t/τ0 = 0. (b)t/τ0 = 500. (c) t/τ0 = 16650. (d) t/τ0 = 16700.

FIG. 4: (Color online) Droplet-vesicle transition at N = 500,ε = 2, ρ∗ = 14, kbend = 4, and Cbd = 0. The lower and upperlines represent the radius of gyration Rg in ktilt increasing ordecreasing, respectively. Sliced snapshots are also shown atktilt = 2.5.

transforms into a bilayer vesicle. At the transition point(ktilt = 4), the radius of gyration Rg exhibits an abruptincrease as shown in Fig. 4. As ktilt decreases, thetransition from a vesicle to a droplet occurs, however,the transition point (ktilt = 0.5) is much lower. Thus,a typical hysteresis for the first-order transition is ob-served. The rate of increase or decrease is sufficientlylow (ktilt = 0.0005t/τ0). We checked that the devia-tion of the transition points by the annealing rates isvery small; ∆ktilt = 0.1 between ktilt = 0.000125t/τ0

FIG. 5: (Color online) Formation of vesicles from a dropletat N = 4000, ε = 2, ρ∗ = 14, kbend = 8, and Cbd = 0. Thetilt coefficient ktilt is gradually increased as ktilt = 0.0005t/τ0 .Sliced snapshots are shown at (a) t/τ0 = 12500 (ktilt = 6.25),(b) t/τ0 = 13600, (c) t/τ0 = 13700, (d) t/τ0 = 14000, and (e)t/τ0 = 14500 (ktilt = 7.25). All molecules are also shown fort/τ0 = 13700 in (c’).

and ktilt = 0.001t/τ0 with N = 500, ε = 2, ρ∗ = 14,kbend = ktilt, and Cbd = 0. The transition points are notsensitive to the path of kbend(ktilt), since the difference ofthe results for kbend = ktilt and constant kbend is smallerthan their statistical errors.

For large aggregates with N = 4000, two vesicles ora vesicle with disks are formed instead of a single vesi-cle. Figure 5 shows an example of the formation of twovesicles. A void space is opened in the droplet, and abilayer skirt is formed. Then, it is separated into twoparts and forms two vesicles. If the separated membraneis small, a disk is formed. Since the larger droplets canhave a clearer molecular layer on the surface, which pre-vents the shape change, higher ktilt is needed to triggerthe shape transition [see Fig. 6(a)]. At N = 300, the co-existence region of the droplets and the vesicles is narrow,since the number of the molecules is not sufficient to formthe surface and inside layers. The points of the droplet-bilayer transition are also dependent on Cbd, while they

Page 5: arXiv:1010.0389v1 [cond-mat.soft] 3 Oct 2010arXiv:1010.0389v1 [cond-mat.soft] 3 Oct 2010 Solvent-free coarse-grained lipid model forlarge-scale simulations Hiroshi Noguchi∗ Institute

5

0

2

4

6

0 2 4 6 8

k tilt

kbend

(d)16

14

12

ρ* =

0

2

4

6

k tilt

(c)

8

42

ε =

0

2

4

6

k tilt

(b)

- 0.4

- 0.2

00.2

0.40.6

Cbd =

0

2

4

6k t

ilt(a)

300

400

5001000

20004000N =

FIG. 6: (Color online) Shape transition points of molecu-lar aggregates between the droplet and the bilayer membrane(vesicles and disks) for various (a) N , (b) Cbd, (c) ε, and (d)ρ∗. If not specified, N = 500, ε = 2, ρ∗ = 14, and Cbd = 0.The solid and dashed lines represent data with increasing anddecreasing ktilt, respectively. The open and filled symbols rep-resent data with fixed kbend and kbend = ktilt, respectively.

are almost independent of ε and ρ∗. Thus, the bilayerstability is determined by the tilt and bending potentialsbut not by the attractive potential.

Let us discuss the condition required to form a stablebilayer. When all molecules have the same orientationui = uj , the bending potential energy becomes zero atCbd = 0. Thus, the bending potential with Cbd = 0

0

10

1.1 1.2 1.3

γσ2 /k

BT

a

(b)

ktilt = 2

4 8

0

10

1 1.1 1.2 1.3

γσ2 /k

BT

axy

(a)

ktilt = 2

4 8

FIG. 7: (Color online) Surface tension γ of a flat membraneat N = 512, ρ∗ = 14, ε = 2, kbend = ktilt, and Cbd =0. Dependence of γ on (a) the projected area per moleculeaxy = 2Axy/Nσ2 and (b) the intrinsic area per molecule a =2A/Nσ2. The squares, triangles, and circles represent γ forktilt = 2, 4, and 8, respectively. The error bars are smallerthan the line thickness.

can have the minimum energy for any structure of theaggregate. Therefore, the spherical liquid droplet, whichhas the minimum surface area, would be the equilibriumstate at ktilt = 0 instead of the bilayer. However, largevesicles with N ≥ 1000 and planar membranes can main-tain their bilayer structure as a metastable state even atktilt = 0 with finite kbend. Note that the capability tokeep a pre-formed bilayer membrane does not guaranteethe self-assembly to the bilayer. In particular, the pe-riodic boundary condition is a strong constraint, whichcan keep the bilayer membrane as a thin liquid layer evenat ktilt = kbend = 0. In order to obtain the spontaneousformation of the bilayer membrane, ktilt > 2 is required.

B. Calculation of membrane properties

To investigate the membrane properties, we formed anearly planar membrane without edges or pores. Themembrane area and the surface tension are varied by in-creasing or decreasing the projected area Axy = LxLy,where Lx = Ly. The membrane has a clear bilayer struc-ture (see Fig. 2). The intrinsic area of the tensionlessmembrane per molecule a0 = 2A0/Nσ2, the area com-pression modulus KA, and the half lifetime τff of the flip-flop motion are calculated from the flat membranes withN = 512. The bending rigidity κ and the diffusion coef-

Page 6: arXiv:1010.0389v1 [cond-mat.soft] 3 Oct 2010arXiv:1010.0389v1 [cond-mat.soft] 3 Oct 2010 Solvent-free coarse-grained lipid model forlarge-scale simulations Hiroshi Noguchi∗ Institute

6

0.001

0.01

0.1

1

10

100

1000

0.1 1

<h(

q)2 >

q/π

q-4

0

10

0 0.04

1/<

h(q)

2 >

(q/π)2

FIG. 8: (Color online) Spectra of undulation modes 〈|h(q)|2〉of nearly planar, tensionless membranes (γ = 0) at N = 8192,ρ∗ = 14, ε = 2, kbend = ktilt = 8, and Cbd = 0. Results for〈|h(q)|2〉 calculated from the molecular positions (+) and fromthe averaged positions on a square mesh (×) are shown. Theinset shows the dependence of 1/〈|h(q)|2〉 on q2, which is usedto extract the bending rigidity κ.

9

9.2

6 8

Γσ/k

BT

rp0 /σ

FIG. 9: (Color online) Line tension Γ of membrane edge atε = 2, ρ∗ = 14, kbend = 4, ktilt = 4, and Cbd = 0. The circlesrepresent Γ calculated from a pore on the flat membrane atN = 2048. The solid line represents Γ calculated from thestriped membrane at N = 512: Γσ/kBT = 9.08 ±0.06.

ficient D are calculated at larger tensionless membraneswith N = 8192. The line tension Γ of membrane edgeis calculated from the strip of the flat membrane withN = 512.The surface tension γ is given by34,35

γ = 〈Pzz − (Pxx + Pyy)/2〉Lz, (13)

with the diagonal components of the pressure tensor

Pαα = (NkBT −∑

i

αi∂U

∂αi)/V, (14)

where α ∈ {x, y, z}. When the potential interactioncrosses the periodic boundary, the periodic image αi +nLα nearest to the other interacting molecules is em-ployed. The intrinsic area A of the membrane is largerthan the projected area Axy in the xy-plane due to the

0.1

1

0 200 400

Pup

- P

dow

n

t/τ0

ktilt =

46

8

FIG. 10: (Color online) Time development of the probabilitydifference Pup−Pdown of the molecules in the upper and lowermonolayers at N = 512, ρ∗ = 14, ε = 2, kbend = 0, andCbd = 0. At the initial states (t = 0), Pup = 1 and Pdown = 0.The error bars are displayed at several data points.

membrane undulations. We calculate A from a√

N/2×√

N/2 square mesh with (xmh, ymh) = (dmhi, dmhj). Theheight zmh of a mesh point is obtained from the weightedaverage of molecular position rsi in the four neighborcells, with zmh =

i ziwmh(xi, yi)/(∑

i wmh(xi, yi)) andwmh(xi, yi) = (1 − |xi − xmh|/dmh)(1 − |yi − ymh|/dmh).Figure 7 shows the area dependence on the surface ten-sion γ. The tension γ exhibits a roughly linear increasewith the molecular area at γ & 0. The compressed mem-brane with γ < 0 buckles out of plane and has the largerintrinsic area A than the projected area Axy. Similar γdependence and buckling are obtained in the simulationsof other molecular models29,36 and meshless models7,10.The area A0 of the tensionless membrane (γ = 0) is

obtained by the minimization of γ, where the projectedarea is updated as Anew

xy = Axy − bγ∆tγ every ∆tγ in-terval, where γ is the time average for ∆tγ . We usebτ0 = 0.00025 to 0.005 and ∆tγ/τ0 = 50 or 100. Thearea compression modulus KA is defined as

KA = A0∂γ/∂A|A=A0. (15)

We calculate KA from the slope of a-γ lines shown in Fig.7(b).The bending rigidity κ is calculated from the spectra

of undulation modes 〈|h(q)|2〉 of the planar membranesin Fourier space1,37,38,

〈|h(q)|2〉 = kBT

γq2 + κq4. (16)

Figure 8 clearly shows the q−4 dependence of the ten-sionless membrane. We calculate |h(q)|2 from the rawdata (the particle position rsi), as well as from the squaremesh with the same mesh-points which were used for theestimation of the intrinsic area A. Averaging over themesh removes most of the effects of the molecular pro-trusions. The bending rigidity κ is estimated from a fitof 1/〈|h(q)|2〉 = (κ/kBT )(q

2)2 for (q/π)2 < 0.015, where

Page 7: arXiv:1010.0389v1 [cond-mat.soft] 3 Oct 2010arXiv:1010.0389v1 [cond-mat.soft] 3 Oct 2010 Solvent-free coarse-grained lipid model forlarge-scale simulations Hiroshi Noguchi∗ Institute

7

the difference of two spectra is very small (see the insetof Fig. 8).The line tension Γ of the membrane edge is calculated

from the strip of the flat membrane as39,40

Γ = 〈Pxx − (Pyy + Pzz)/2〉LyLz/2, (17)

since Γ is the energy per unit length of the membraneedge, and the length of the membrane edge is 2Lx. Sincethe striped membrane is tensionless, 〈Pyy〉 = 〈Pzz〉 ≃ 0for solvent-free simulations. The tension Γ and its errorbar are estimated from the average and standard devi-ations for Lx/σ = 14, 15, 16, and 18. Alternatively, Γcan be also calculated from a circular pore on the flatmembrane7,39. In this case, Γ is balanced with the sur-face tension γ as Γ = γrp0. Since one has to estimate thepore radius rp0, this method gives larger statistical errorsas shown in Fig. 9. Therefore, we used the membranestrip for the calculation of Γ.In bilayer membranes, molecules can move laterally on

a monolayer and transversely between upper and lowermonolayers. The lateral diffusion coefficient D of themolecules is calculated from the diffusion of the molecu-lar projections in the xy plane; D = 〈(xi(t) − xi(0))

2 +(yi(t) − yi(0))

2〉/4t. In the fluid phase, the moleculesexhibit a fast diffusion rate D/D0 = 0.05 to 0.1.The relaxation time of the transverse motion (flip-flop)

between the upper and lower monolayers is measuredfrom the relaxation of the labeled molecules26,41. The dif-ferential equation of the probability Pup(t) (Pdown(t)) ofthe molecules, which belong to the upper (lower) mono-layer, is given by dPup/dt = −kuPup + kdPdown, wherePup + Pdown = 1. For planar membranes, ku = kd, andPup = Pdown = 1/2 at t → ∞. When the molecules inthe upper monolayer are initially labeled (Pup(0) = 1),the probability decays as

Pup(t)− Pdown(t) = exp[−(ku + kd)t] (18)

with the half lifetime τff = ln(2)/(ku + kd). Figure 10shows that Pup(t)−Pdown(t) indeed follows the exponen-tial decay in our simulations. Either the z component ofthe position rei or orientation ui can be used to detect thenomolayer, to which a molecule belongs. Both rei and ui

give the same probability distribution (entirely same formost of the parameters), since both the quantities havea clear minimum between two peaks (see Fig. 2).

C. Parameter dependence of membrane properties

Figures 11–18 show the parameter dependence of theproperties of the tensionless membrane. The bilayermembrane is formed in the fluid phase over broad rangesof the parameters due to the multibody attractive poten-tial. A gel phase is obtained only at a large value of ρ∗ inEq. (3). The fluid-gel transition occurs at ρ∗ = 16 andρ∗ = 15 for ε = 2 and 8, respectively [see jumps of D inFig. 11(e)]. As ρ∗ decreases, the lateral diffusion and flip-flop motion become faster, and the membrane elasticities

0

0.05

0.1

12 14 16

D/D

0

ρ*

(e)2

8

ε =

0

10

20

Γσ/k

BT

(d)

2

8ε =

20

30

40

κ/k B

T (c)

2

8ε =

0

200

400

KAσ2 /k

BT

(b)2

8ε =

1

1.2

1.4

a 0

(a)2

8

ε =

FIG. 11: (Color online) Parameter ρ∗ dependence of (a) theintrinsic area a0 = 2A0/Nσ2 per molecule, (b) area compres-sion modulus KA, (c) bending rigidity κ, (d) line tension Γ,and (e) diffusion coefficient D for the tensionless membraneat ktilt = 4, kbend = 4, and Cbd = 0. The circles and squaresrepresent data for ε = 2 and 8, respectively.

(KA, κ, Γ) decrease [see Figs. 11 and 18(d)]. The intrin-sic area a0 = 2A0/Nσ2 per molecule (N/2 molecules inone of monolayers) decreases with increasing ρ∗ and ap-

proaches the closest-packing area√3σ2/2 ≃ 0.87σ2 in

the gel phase. In this paper, we focus on the fluid mem-

Page 8: arXiv:1010.0389v1 [cond-mat.soft] 3 Oct 2010arXiv:1010.0389v1 [cond-mat.soft] 3 Oct 2010 Solvent-free coarse-grained lipid model forlarge-scale simulations Hiroshi Noguchi∗ Institute

8

0.04

0.06

0.08

0 2 4 6 8 10

D/D

0

ε

(e)

2

48

ktilt =

0

10

20

Γσ/k

BT

(d)

2

48ktilt =

0

20

40

κ/k B

T

(c)2

4

8ktilt =

0

100

200

KAσ2 /k

BT

(b)2

4

8

ktilt =

1.15

1.2

1.25a 0

(a)

2

4

8ktilt =

FIG. 12: (Color online) Parameter ε dependence of (a)a0 = 2A0/Nσ2, (b) KA, (c) κ, (d) Γ, and (e) D for ρ∗ = 14,kbend = ktilt, and Cbd = 0. The triangles, circles, and squaresrepresent data for ktilt = 2, 4, and 8, respectively.

brane and set ρ∗ = 14, hereafter.

The dependence on the strength of attraction ε in Eq.(1) is shown in Fig. 12. It has a tendency similar to ρ∗

dependence. The line tension Γ can be varied by ε. Atε = 1, Γ is close to kBT/σ and the bilayer membraneis accompanied by free molecules (gas) with the average

0

20

40

0 5 10 15 20

κ/k B

T

ktilt+kbend

(c)

2(Kbend+ktilt)+6

0

20

40

0 5 10

κ/k B

T

kbend

(b)

ktilt = 8

ktilt = 4

0

20

40

0 5 10

κ/k B

T

ktilt

(a)

kbend = 8

kbend = ktilt

kbend = 0

FIG. 13: (Color online) Bending rigidity κ dependence on (a)ktilt, (b) kbend, and (c) ktilt + kbend for ρ∗ = 14, ε = 2, andCbd = 0. The solid lines with squares, circles, and trianglesrepresent data for kbend = 0, kbend = ktilt, and kbend = 8,respectively. The dashed lines with crosses and diamondsrepresent data for ktilt = 4 and 8, respectively.

density of the gas ∼ 0.001/σ3. The molecules departfrom the bilayer membrane and return. At ε = 0.75 withkbend = ktilt = 4 and Cbd = 0, the bilayer membranebreaks and small micelles are formed (≃ 60 moleculesper micelle for N/V = 0.08/σ3). On the other hand, nofree molecules are seen at ε ≥ 2. The critical micelle con-centration (CMC) of lipids is very low, and their chemicalpotential difference in solution and in membrane is typ-ically more than 10kBT per lipid42. Thus, the numberof lipid molecules on vesicles is conserved in typical ex-periments. In order to keep the number of molecules onmembrane constant during simulations, a sufficiently lowCMC is required. We mainly use ε = 2 and 8, where thefluid membranes without free molecules are obtained.The bending rigidity κ of the bilayer membrane can

be controlled by ktilt or kbend. in Eq. (1). Figures13(a) and (b) show the linear dependence of κ on ktilt

Page 9: arXiv:1010.0389v1 [cond-mat.soft] 3 Oct 2010arXiv:1010.0389v1 [cond-mat.soft] 3 Oct 2010 Solvent-free coarse-grained lipid model forlarge-scale simulations Hiroshi Noguchi∗ Institute

9

0.06

0.08

0 2 4 6 8 10

D/D

0

ktilt

(d)0

ktilt

8

kbend =

8

10

Γσ/k

BT

(c)0

ktilt

8kbend =

0

100

KAσ2 /k

BT

(b)

0ktilt

8kbend =

1.1

1.2a 0

(a)0

ktilt

8kbend =

FIG. 14: Parameter ktilt dependence of (a) a0 = 2A0/Nσ2,(b)KA, (c) Γ, and (d)D for ρ∗ = 14, ε = 2, and Cbd = 0. Thesquares, circles, and triangles represent data for kbend = 0,kbend = ktilt, and kbend = 8, respectively.

and kbend, respectively. When they are plotted togetherfor ktilt + kbend, all lines are roughly overlapped on aline κ/kBT = 2(ktilt + kbend) + 6 [see Fig. 13(c)]. Alarge deviation from the line is seen only at one datapoint at ktilt = 0 and kbend = 8 (the leftmost triangle),where the bilayer structure is metastable. The bend-ing rigidity κ is also weakly dependent on ε. The κ-εcurve maintains its shape and shifts upward with in-creasing ktilt as shown in Fig. 12(c). Thus, it is ex-pressed by κ/kBT = 2(ktilt + kbend) + bε(ε), where bε(ε)is an increasing function as bε(2) = 6, bε(4) = 14, andbε(8) = 18. The area compression modulus KA increaseswith increasing ktilt, while KA shows only slight depen-dence on kbend for large ktilt (see Figs. 14 and 15). Theother membrane properties a, Γ, and D show weak de-

0.06

0.07

0 2 4 6 8 10

D/D

0

kbend

(d)

4

8

ktilt =

8

9

10

Γσ/k

BT

(c)4

8ktilt =

100

150

KAσ2 /k

BT

(b)

4

8ktilt =

1.18

1.2

1.22

a 0

(a)

8

4

ktilt =

FIG. 15: (Color online) Parameter kbend dependence of (a)a0 = 2A0/Nσ2, (b) KA, (c) Γ, and (d) D for ρ∗ = 14, ε = 2,and Cbd = 0. The squares and circles represent data forktilt = 4 and 8, respectively.

pendence on ktilt and kbend. Thus, κ can be varied byktilt or kbend without a large variation in Γ. The modulusKA can be varied by ktilt.

The bending elastisity generated by the orientation-dependent potentials can be derived from the Helfrichtheory for monolayer membranes. When the orientationvectors ui are equal to the normal vectors of the monolay-ers without tilt deformation, the bending and tilt energiesare written by

Ucv =

dAκ′bend

2[(C1 − C′

0)2 + (C2 − C′

0)2]

+κ′tilt

2[(C2

1 + C22 )] (19)

Page 10: arXiv:1010.0389v1 [cond-mat.soft] 3 Oct 2010arXiv:1010.0389v1 [cond-mat.soft] 3 Oct 2010 Solvent-free coarse-grained lipid model forlarge-scale simulations Hiroshi Noguchi∗ Institute

10

0

5

10

-0.4 0 0.4 0.8

Γσ/k

BT

Cbd

(b)

2

81ktilt =

kbend = 8

0

5

10Γσ

/kBT (a)

4

2

8 6

ktilt =

kbend = ktilt

FIG. 16: (Color online) Line tension Γ dependence on Cbd

for (a) kbend = ktilt and (b) kbend = 8 at ρ∗ = 14 and ε = 2.(a) The diamonds, squares, triangles, and circles representdata for ktilt = 2, 4, 6, and 8, respectively. (b) The squares,triangles, and circles represent data for ktilt = 1, 2, and 8,respectively.

=

dAκ′bend + κ′

tilt

2[(C1 + C2 − C0)

2

−(κ′bend + κ′

tilt)C1C2 + U0 (20)

in the continuum limit, where C1 and C2 are two prin-cipal curvatures of the monolayer. The first and secondterms in Eq. (19) are the contributions of the bendingand tilt potentials, respectively. The spontaneous curva-ture of the bending potential is given by C′

0 = Cbd/rnb.27

The nearest-neighbor distance rnb ≃ 1.05σ is obtainedfrom the radial distribution function. By assuming thehexagonal packing of the molecules in the monolayers,the monolayer bending rigidities generated by the bend-ing and tilt potentials are estimated as κ′

bend/kBT =√3kbendwcv(rnb) and κ′

tilt/kBT =√3ktiltwcv(rnb)/2, re-

spectively. The bending rigidity of the monolayer is givenby the sum of these κmono = κ′

bend + κ′tilt. For the

monolayer membrane, Eq. (20) gives the saddle-splaymodulus κmono = −κmono and the spontaneous curva-ture C0 = {κ′

bend/(κ′bend + κ′

tilt)}Cbd/rnb with U0 =(κ′

bend + κ′tilt)(1/2 + κ′

tilt/κ′bend)C

20 . Thus, the bending

rigidity κ of the bilayer is estimated as κ = 2κmono ≃(2.1kbend + 1.1ktilt)kBT from wcv(1.05σ) = 0.61. Thisestimation of κ supports the linear dependence of theobtained simulation results on kbend and ktilt. The pref-actor (≃ 2) of kbend gives the quantitative agreement,whereas the prefactor of ktilt is half of the numerical es-timation. In the simulation, the thermal fluctuations in-

0.06

0.065

-0.4 0 0.4 0.8

D/D

0

Cbd

(d)

4

8

ktilt =

0

20

40

60

κ/k B

T(c)

4

8ktilt =

120

160

KAσ2 /k

BT

(b)

8

4

ktilt =

1.2

1.25

a 0

(a)

8

4

ktilt =

FIG. 17: (Color online) Parameter Cbd dependence of (a)a0 = 2A0/Nσ2, (b) KA, (c) κ, and (d) D for ρ∗ = 14, ε = 2,and kbend = ktilt. The squares and circles represent data forktilt = 4 and 8, respectively.

duce molecular protrusion and tilt. These tilt fluctua-tions likely change the prefactor of ktilt to twice its value(κ′

tilt/κ′bend = ktilt/kbend). The attractive potential also

adds a small bending resistance, bε(ε).

The flip-flop time τff shows exponential dependence onthe parameters of the tilt and bending potentials, while ithas weak dependence on ε (see Fig. 18). The free-energybarrier between two monolayers would be the main fac-tor to determine τff . It can be roughly estimated fromthe probability distribution of the molecular orientationas Fff = kBT (ln[P (uz = 1)] − ln[P (uz = 0)]) [see Fig.2(b)]. The dashed lines in Fig. 18(a) represent Fff cal-culated by this method. It gives very good agreementswith ln(τff/τ0), and thus, the flip-flop time is written byτff/τ0 = bff exp(Fff/kBT ) with bff ≃ 1. The barrier Fff is

Page 11: arXiv:1010.0389v1 [cond-mat.soft] 3 Oct 2010arXiv:1010.0389v1 [cond-mat.soft] 3 Oct 2010 Solvent-free coarse-grained lipid model forlarge-scale simulations Hiroshi Noguchi∗ Institute

11

103

104

105

12 14 16

τ ff/τ

0

ρ*

(d)

ε = 8

ε = 2

102

103

104

0 2 4 6 8 10

τ ff/τ

0

ε

(c)

ktilt = 4

ktilt = 2

103

104

-0.4 0 0.4 0.8

τ ff/τ

0

Cbd

(b)

ktilt = 4kbend = 4

101

102

103

104

105

106

0 2 4 6 8 10

5

10

τ ff/τ

0

ln(τ

ff/τ 0

), F

/kBT

ktilt

(a)kbend = ktilt

kbend = 0

FIG. 18: (Color online) Half lifetime τff of flip-flop motion.(a) Dependence on ktilt at ρ∗ = 14, ε = 2, and Cbd = 0.The dashed lines represent the free-energy barrier Fff/kBTestimated by the orientation distribution shown in Fig. 2(b).(b) Dependence on Cbd at ρ∗ = 14 and ε = 2. (c) Dependenceon ε at ρ∗ = 14, kbend = ktilt, and Cbd = 0. (d) Dependenceon ρ∗ at kbend = ktilt = 4 and Cbd = 0.

linear with the orientation-dependent potentials, while itis almost independent of the attractive potentials. Theflip-flop rate can be tuned by ktilt or kbend. In typicalexperimental conditions, the flip-flop motion of phospho-lipids is very slow, and τff is several hours or days41.At a sufficiently high ktilt or kbend, no flip-flop occurs in

FIG. 19: (Color online) Sequential snapshots of vesicle rup-ture at N = 2000, ε = 2, ρ∗ = 14, kbend = 8, ktilt = 8, andCbd = 0.85. (a) t/τ0 = 500. (b) t/τ0 = 640. (c) t/τ0 = 3000.

typical simulation time scales. This result agrees withthe experimental observations. On the other hand, atsmall ktilt and kbend, the molecule shows very fast flip-flop, which is advantageous to equilibrate the membranesystem quickly during simulations. In the present model,one can choose slow or fast flip-flop condition to matchone’s simulation purpose.The spontaneous curvature of the monolayer is var-

ied by the parameter Cbd. According to the above con-tinuous theory, C0 = {kbend/σ(kbend + ktilt)}Cbd. Athigh spontaneous curvature C0 ∼ 1/σ, a worm-like mi-celle is stabilized. As Cbd increases, the line tension Γdecreases, and the bilayer structure becomes unstableat Γ ∼ kBT/σ. The Cbd-Γ curves in Fig. 16 have aparabolic shape with a maximum at Cbd ∼ −0.1. Thebending rigidity κ and the flip-flop time τff increase withCbd, while a0, KA, and D do not depend significantlyon Cbd [see Figs. 17 and 18(b)]. At negative C0, themolecules easily stay at the middle of the bilayer struc-ture (low Fff), and the less clear bilayer would generatelower κ.Membrane rupture is observed when Cbd is increased.

Figure 19 shows the rupture process of a vesicle at Cbd =0.85. The initial vesicle is metastable and maintains itsshape for 500τ0. Then, a pore opens and grows to cracks[see Fig. 19(b)]. Finally, a branched worm-like micelle isformed [see Fig. 19(c)]. Thus, the line tension and thestable structure can be varied by Cbd.

IV. SUMMARY

We have proposed a simple molecular model of bilayermembranes. The molecule has five degrees of freedom,three translational degrees and two orientational degrees.Since this molecular model consists of one spherical par-

Page 12: arXiv:1010.0389v1 [cond-mat.soft] 3 Oct 2010arXiv:1010.0389v1 [cond-mat.soft] 3 Oct 2010 Solvent-free coarse-grained lipid model forlarge-scale simulations Hiroshi Noguchi∗ Institute

12

ticle for the excluded volume, it is smaller than previousmolecular models. Thus, this model is more efficient forlarge-scale simulations. Despite the extreme simplifica-tion, this model reproduces many aspects of lipid bilayermembranes.In the present model, the properties of the fluid

membranes can be controlled in broad ranges includingmetastable bilayer membranes. The bending rigidity κ islinearly dependent on ktilt and kbend. The line tension Γof the membrane edge can be varied by ε and Cbd. Themembrane has a wide range of fluid phase, and the fluid-gel transition point can be controlled by ρ∗. The areacompression modulus KA can be varied by ktilt. Theflip-flop time τff can be varied by ktilt and kbend.The corresponding time and length scales can be

mapped by the membrane thickness 5nm and the lateraldiffusion coefficient ∼ 10−8cm2/s for phospholipids43.Thus, the unit length and times are estimated as σ = 2nm and τ0 ∼ 0.1µs, respectively.Our model is suitable to study the details of

topological-change processes of the membrane, such asmolecular self-assembly, pore formation, membrane fu-sion, and membrane fission. Although meshless mem-brane models can also be used, they cannot treat detailedstructures such as the fusion intermediates14,15. Re-cently, fusion dynamics have been investigated by molec-ular simulations44–51. However, the condition to deter-mine the fusion pathways has not be clarified so far.

The ability to vary the membrane properties in wideranges would be an advantage of this model for quan-titative investigation of the membrane fusion pathways.On the other hand, this model is not suitable for ap-plication to phenomena, in which the stretching of hy-drophobic chains or atomistic details play an importantrole, e.g., in the interactions of membrane proteins viahydrophibic mismatch, where molecular stretching is notnegligible52–55.

In this paper, we used Brownian dynamics but onecan also use the Monte Carlo method and molecular dy-namics with another thermostat. When the solvent-freemolecular model is combined with a particle-based hy-drodynamics method, multi-particle collision dynamics(MPC)56,57, the hydrodynamic interaction can be takeninto account as demonstrated for a meshless membranemodel8. Thus, the present coarse-grained molecularmodel is efficient for large-scale dynamics of a biomem-brane with or without hydrodynamic interactions and isapplicable to many kinds of phenomena.

Acknowledgments

This work is supported by KAKENHI (21740308) fromthe Ministry of Education, Culture, Sports, Science, andTechnology of Japan.

[email protected] S. A. Safran, Statistical Thermodynamics of Surfaces, In-

terfaces, and Membranes (Addison-Wesley, Reading, MA,1994).

2 R. Lipowsky and E. Sackmann, eds., Structure and

Dynamics of Membranes (Elsevier Science, Amsterdam,1995).

3 U. Seifert, Adv. Phys. 46, 13 (1997).4 G. Gompper and D. M. Kroll, in Statistical Mechanics of

Membranes and Surfaces, edited by D. R. Nelson, T. Piran,and S. Weinberg (World Scientific, Singapore, 2004), 2nded.

5 G. Gompper and D. M. Kroll, J. Phys. Condens. Matter9, 8795 (1997).

6 J. M. Drouffe, A. C. Maggs, and S. Leibler, Science 254,1353 (1991).

7 H. Noguchi and G. Gompper, Phys. Rev. E 73, 021903(2006).

8 H. Noguchi and G. Gompper, J. Chem. Phys. 125, 164908(2006).

9 M. G. Del Popolo and P. Ballone, J. Chem. Phys. 128,024705 (2008).

10 T. Kohyama, Physica A 388, 3334 (2009).11 P. Liu, J. Li, and Y. W. Zhang, Appl. Phys. Lett. 95,

143104 (2009).12 R. M. Fuchslin, T. Maeke, and J. S. McCaskill, Eur. Phys.

J. E 29, 431 (2009).13 H. Yuan, C. Huang, J. Li, G. Lykotrafitis, and S. Zhang,

Phys. Rev. E 82, 011905 (2010).

14 R. Jahn and H. Grubmuller, Curr. Opin. Cell Biol. 14, 488(2002).

15 L. V. Chernomordik and M. M. Kozlov, Nat. Struct. Mol.Biol. 15, 675 (2008).

16 M. Muller, K. Katsov, and M. Schick, Phys. Rep. 434, 113(2006).

17 M. Venturoli, M. M. Sperotto, M. Kranenburg, andB. Smit, Phys. Rep. 437, 1 (2006).

18 M. L. Klein and W. Shinoda, Science 321, 798 (2008).19 H. Noguchi, J. Phys. Soc. Jpn. 78, 041007 (2009).20 S. J. Marrink, A. H. de Vries, and D. P. Tieleman, Biochim.

Biophys. Acta 1788, 149 (2009).21 S. J. Marrink, A. H. de Vries, and A. E. Mark, J. Phys.

Chem. B 108, 750 (2004).22 S. Izvekov and G. A. Voth, J. Phys. Chem. B 109, 2469

(2005).23 A. Arkhipov, Y. Yin, and K. Schulten, Biophys. J. 95,

2806 (2008).24 W. Shinoda, R. DeVane, and M. L. Klein, Soft Matter 4,

2454 (2008).25 Z. J. Wang and M. Deserno, J. Phys. Chem. B 114, 11207

(2010).26 H. Noguchi and M. Takasu, Phys. Rev. E 64, 041913

(2001).27 H. Noguchi, Phys. Rev. E 67, 041901 (2003).28 G. Brannigan, L. C. L. Lin, and F. L. H. Brown, Eur.

Biophys. J. 35, 104 (2006).29 M. Deserno, Macromol. Rapid. Commun. 30, 752 (2009).30 O. Farago and N. Grønbech-Jensen, J. Am. Chem. Soc.

Page 13: arXiv:1010.0389v1 [cond-mat.soft] 3 Oct 2010arXiv:1010.0389v1 [cond-mat.soft] 3 Oct 2010 Solvent-free coarse-grained lipid model forlarge-scale simulations Hiroshi Noguchi∗ Institute

13

131, 2875 (2009).31 S. Takada, Z. Luthey-Schulten, and P. G. Wolynes, J.

Chem. Phys. 110, 11616 (1999).32 M. Hamm and M. M. Kozlov, Eur. Phys. J. B 6, 519

(1998).33 M. Hamm and M. M. Kozlov, Eur. Phys. J. E 3, 323 (2000).34 M. P. Allen and D. J. Tildesley, Computer simulation of

liquids (Clarendon Press, Oxford, 1987).35 J. S. Rowlinson and B. Widom, Molecular Theory of Cap-

illarity (Clarendon, Oxford, 1982).36 W. K. den Otter, J. Chem. Phys. 123, 214906 (2005).37 R. Goetz, G. Gompper, and R. Lipowsky, Phys. Rev. Lett.

82, 221 (1999).38 E. Lindahl and O. Edholm, Biophys. J. 79, 426 (2000).39 T. V. Tolpekina, W. K. den Otter, and W. J. Briels, J.

Chem. Phys. 121, 8014 (2004).40 B. J. Reynwar and M. Deserno, Bioinerphases 3, FA117

(2008).41 R. D. Kornberg and H. M. McConnell, Biochemistry 10,

1111 (1971).42 C. Tanford, The hydrophobic effect: formation of micelles

and biological membranes (Wiley, New York, 1980).43 E. S. Wu, K. Jacobson, and D. Papahadjopoulos, Biochem-

istry 16, 3936 (1977).

44 H. Noguchi and M. Takasu, J. Chem. Phys. 115, 9547(2001).

45 H. Noguchi and M. Takasu, Biophys. J. 83, 299 (2002).46 H. Noguchi, J. Chem. Phys. 117, 8130 (2002).47 M. Muller, K. Katsov, and M. Schick, J. Chem. Phys. 116,

2342 (2002).48 D. W. Li and X. Y. Liu, J. Chem. Phys. 122, 174909

(2005).49 J. Shillcock and R. Lipowsky, Nat. Mater. 4, 225 (2005).50 A. F. Smeijers, A. J. Markvoort, K. Pieterse, and P. A. J.

Hilbers, J. Phys. Chem. B 110, 13212 (2006).51 V. Knecht and S. J. Marrink, Biophys. J. 92, 4254 (2007).52 R. Phillips, T. Ursell, P. Wiggins, and P. Sens, Nature 459,

379 (2009).53 J. B. Fournier, Eur. Phys. J. B 11, 261 (1999).54 F. J. de Meyer, M. Venturoli, and B. Smit, Biophys. J. 95,

1851 (2008).55 B. West, F. L. H. Brown, and F. Schmid, Biophys. J. 96,

101 (2009).56 R. Kapral, Adv. Chem. Phys. 140, 89 (2008).57 G. Gompper, T. Ihle, D. M. Kroll, and R. G. Winkler,

Adv. Polym. Sci. 221, 1 (2009).


Recommended