Abstract
Using a combination of coarse-grained and atomistic molecular dynamics simulations we have investigated the membrane binding and folding properties of the membrane lytic peptide of Flock House virus (FHV). FHV is an animal virus and an excellent model system for studying cell entry mechanisms in non-enveloped viruses. FHV undergoes a maturation event where the 44 C-terminal amino acids are cleaved from the major capsid protein, forming the membrane lytic (γ) peptides. Under acidic conditions, γ is released from the capsid interior allowing the peptides to bind and disrupt membranes. The first 21 N-terminal residues of γ, termed γ1, have been resolved in the FHV capsid structure and γ1 has been the subject of in vitro studies. γ1 is structurally dynamic as it adopts helical secondary structure inside the capsid and on membranes, but it is disordered in solution. In vitro studies have shown the binding free energies to POPC or POPG membranes are nearly equivalent, but binding to POPC is enthalpically driven, while POPG binding is entropically driven. Through coarse-grained and multiple microsecond all-atom simulations the membrane binding and folding properties of γ1 are investigated against homogeneous and heterogeneous bilayers to elucidate the dependence of the microenvironment on the structural properties of γ1. Our studies provide a rationale for the thermodynamic data and suggest binding of γ1 to POPG bilayers occurs in a disordered state, but γ1 must adopt a helical conformation when binding POPC bilayers.
Keywords: virus infection, molecular dynamics simulations, bilayers, coarse-grained simulations, anti-microbial peptides
1. Introduction
Our current understanding of the mechanisms employed by non-enveloped animal viruses to penetrate the membranes of their host cells is lacking as compared to our knowledge of enveloped virus cell entry mechanisms. Broadly speaking, enveloped animal viruses employ membrane fusion mechanisms mediated via viral surface proteins, called fusion proteins. The fusion proteins of enveloped viruses undergo large conformational changes in response to receptor binding or environmental changes, which allows the fusion proteins to insert into the host membrane. The fusion proteins mediate the merger of the viral and host membrane by reducing the energy barrier to the fused state, in which the virus can shed its membrane and pass through the host membrane.1,2 On the other hand, non-enveloped viruses lack a membrane and therefore cannot perform membrane fusion to enter cells or exit from membrane bound cellular compartments (e.g. endosomes). The mechanisms employed by these non-enveloped viruses to cross cellular membrane barriers are less well understood. Many non-enveloped viruses contain a specialized capsid protein(s) that mediates membrane penetration. It has been proposed that these “penetration proteins” are inactive precursors sequestered inside the capsid. They become activated under specific cellular conditions, leading to their externalization from the capsid interior and exposure to cellular membranes. Recent studies have indicated this general feature of non-enveloped viruses is spread across many non-enveloped viral families including, Picornaviridae, Polyomaviridae, Reoviridae, Parvoviridae, Adenoviridae, Alphatetraviridea and Nodaviridae.3,4
Flock House virus (FHV) is one of the simplest and smallest non-enveloped viruses and provides a model system to study non-enveloped animal viruses.5 FHV infects insects and is a member of the Nodaviridae family and has a small (4.5 kb) bipartite positive sense RNA genome. The FHV capsid protein contains 407 residues, which assemble into a T=3 icosahedral capsid containing 180 subunits. The initial assembled particle is immature and must undergo an autocatalytic cleavage event to mature and become infective.6 The cleavage happens between residue 363 and 364 and is coordinated by residue Asp757 resulting in the major capsid protein β (residues 1–363) and the γ peptide (364–407). The structure of the mature capsid has been determined by X-ray crystallography at 2.7 Å resolution which shows γ is non-covalently associated with the interior capsid surface (Fig. 1A).8 FHV enters cells through the endocytic pathway where the virus encounters an acidic environment. The low pH of the endosome is believed to promote externalization of γ peptides from the capsid interior, which can then disrupt the endosomal membrane creating a channel for the release of the virus or viral RNA into the cytoplasm. This mechanism is supported by in vitro liposome dye leakage experiments which show FHV to have maximal lytic activity at pH 6.9 It has been established that cleavage defective mutants display almost no infective ability,6 and infectivity of non-cleaving mutants can be rescued by coinfection with wild-type virus-like particles (VLPs),10 establishing the criticality of cleaved γ peptides to viral infection.
Figure 1.
FHV Structures. A) A view from inside the FHV capsid, looking down a five-fold axis. The pentameric capsid subunits are shown in blue and the γ-peptides in yellow (PDB ID: 4FTB). B) The γpeptide is displayed with side chains drawn and colored as follows, pink=hydrophobic, green=uncharged polar, blue=positively charged, red=negatively charged. The residue numbering and residue types are listed near the corresponding residues.
Multiple lines of data point toward the γ-peptide being externalized from the FHV capsid and then acting independently (apart from the capsid) to perform membrane disruption.9,11,12 Therefore, several in vitro studies have investigated the properties of the FHV γ-peptide in solution and in the presence of bilayers to provide mechanistic insights into the membrane disruption process. Many of these studies have focused on the region of γ that was resolved in the capsid crystal structure (Fig. 1), which are the first 21 N-terminal residues. This minimal version of γ is termed γ1, and it retains membrane lytic activity, though at a weaker level than the full length γ.13,14 The solution behavior of γ1 has been examined by circular dichroism (CD) and nuclear magnetic resonance (NMR), and it is found that the structure is largely disordered in aqueous environments.12,15 In the NMR study, it was shown that increasing the hydrophobicity of the solvent by addition of trifluoroethanol (TFE) increases the ordering of the peptide, leading to a partially α-helical conformation at 50% TFE.12 Biophysical experimental studies showed γ1 spontaneously partitions into large unilamellar vesicles (LUVs) made of either 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine (POPC) or 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphoglycerol (POPG) lipids.15 CD titration experiments demonstrated that γ1 orders upon addition of POPC or POPG LUVs, increasing in helical content up to 75%. Attenuated total reflectance infrared spectroscopy (ATR-IR)13 as well as thermodynamic analysis,15 support a model in which γ1 orients laterally, burying about half the peptide surface area into the hydrophobic core of the bilayer, while maintaining partial exposure to the solvent. Native tryptophan fluorescence indicates an insertion depth for γ1 to be approximately 1 nm from the bilayer center and deeper insertion occurs into POPC than POPG bilayers.13 Thermodynamic analysis using isothermal titration calorimetry (ITC) revealed that even though the γ1 binding free energies to POPC (−7.5 kcal/mol) or POPG (−8.0 kcal/mol) are nearly equivalent, the association with negatively charged POPG is entropy driven (ΔH =10.2 kcal/mol; ΔS = 61.1 cal/mol·K), whereas the association with neutral POPC is enthalpy dominated (ΔH = −4.1 kcal/mol; ΔS = 11.4 cal/mol·K)13. The activity of the peptides in membrane lysis has been examined by liposome dye release studies and the activity is much higher against POPC than POPG; a ~10-fold higher peptide:lipid ratio was needed to lyse POPG vesicles at comparable levels to POPC vesicles.13 There is minimal pH dependence to the lytic activity of the free peptides, indicating the release mechanism is pH-triggered while the membrane binding/insertion/disruption processes appear pH-independent.9,12
Through these previous studies a link between the membrane composition and the activity of these peptides has been established, though how this relates to the structural and dynamic features of the peptides remains unclear. There are, however, definite specifications involved that govern the binding propensity of the membrane active lytic peptides of the FHV. These governing factors are both kinetic and thermodynamic which are sensitive to the microenvironment. To probe the structural dynamics and energetics of FHV γ1, we have employed multi-scale molecular simulations, using a combination of coarse-grained (CG), all-atom (AA) and enhanced sampling molecular dynamics (MD) simulations. Through this approach we are able to examine 1) the structural and energetic properties of γ1 in solution, 2) the structure and energetics of γ1 binding to bilayers of varying composition, and 3) the structural evolution and membrane insertion properties of γ1 interacting with bilayers of varying composition.
2. Methods
We performed multi-scale simulations to characterize the structure and dynamics of γ1 peptide (21 residues: ASMWERVKSIIKSSLAAASNI) in aqueous solution and in membrane environments. All system configurations and snapshots were visualized using VMD software.16
2.1. Well-Tempered Metadynamics (WT-MetaD)
To evaluate the conformational dynamics of γ1 in explicit solvent in neutral pH conditions we applied the well-tempered metadynamics method17 implemented in NAMD version 2.10.18 In brief, this method adds time dependent biases along a chosen collective variable(s) (CV) that is used to accelerate the system dynamics in predefined set of CV. The free energy surface (FES) at time t is derived from the histogram of the occupancy of the CV (N(s,t)) in the simulations
| (1) |
where, F is the FES, s is the vector of the CV and ΔT is a bias temperature that regulates the extent of the energy barriers that can be overcome. For our simulations we used a Gaussian width of 0.2 Å, a Gaussian height of 0.2 kcal/mol, a bias temperature ΔT = 3000 K, and an energy deposition frequency of 1 ps−1. We have performed WT-MetaD simulations on FHV γ1 using an end-to-end distance CV, as defined by the difference between the center of geometry of the first two and last two residues backbone atoms. The simulations were run for 190 ns using the CHARMM22 force field19 with CMAP correction.20 The simulations were performed in the NPT ensemble at 300 K and 1 ATM, with period boundary conditions in X,Y, and Z-directions. A time step of 2 fs was used with a non-bonded cutoff distance of 12 Å, which were smoothly shifted to zero after 10 Å, long range electrostatics were computed with the particle mesh Ewald (PME)21 every 4 fs.
2.2. Coarse-Grained (CG) simulations
We performed six different γ1-bilayer CG simulations using the MARTINI force field22–24 with the GROMACS 4.6.5 software package.25 Proteins in MARTINI are not capable of spontaneously forming or breaking secondary structure (SS), therefore we performed simulations both when γ1 was restrained to an α-helical conformation and when γ1 is in a disordered configuration. The description of the systems has been summarized in Table 1. System I and II consist of 270 POPC CG lipid molecules, while System III and IV comprise of 288 POPG CG lipid molecules. Furthermore, system V and VI consist of a 50:50 mix of POPC:POPG, with 100 CG lipid molecules each of POPC and POPG. The bilayer for each system was generated using the insane.py script.26 The bilayers were first energy-minimized and then equilibrated in the NPT ensemble for 400 ns with polarizable water27 for the solvent. The system was ionized with a proper number of Na+ and Cl− ions to make the systems electrically neutral at 0.1 M salt concentration. The simulations were done starting with different initial velocity distributions of γ1 peptide placed ~ 4 nm from the pre-equilibrated bilayer center, with γ1 helical axis parallel to the bilayer. The initial FHV γ1 peptide structure was obtained from the crystal structure, PDB ID: 4FTB.
Table 1.
Simulation System Details. The Initial γ1 Configuration column describes which CG simulation end state was used to initiate the AA simulations.
| System | Lipid Composition | CG Spontaneous Binding Simulations | AA bilayer bound folding simulations | |||
|---|---|---|---|---|---|---|
| SS | γ1 binds | Initial γ1 Configuration | Simulation Time | Final% Helicity | ||
| I | POPC | on | yes | System I | 1 μs | 78 |
| II | POPC | off | no | n/a | n/a | n/a |
| III | POPG | on | yes | System III | 1 μs | 73 |
| IV | POPG | off | yes | System IV | 1 μs | 0 |
| V | POPC/POPG (50/50) | on | yes | System V | 1 μs | 72 |
| VI | POPC/POPG (50/50) | off | yes | System VI | 1 μs | 16 |
| VII | POPC | n/a | n/a | System VI | 1 μs | 63 |
| VIII | POPG | n/a | n/a | System VI | 1 μs | 23 |
After the initial setup, each system was energy minimized using the steepest descent algorithm. All production runs were simulated in the NPT ensemble using Berendsen coupling scheme28 with the temperature maintained at 323 K and pressure kept at 1.0 bar with semi-isotropic coupling. The time constants for the pressure and temperature couplings were 3.0 and 1.0 ps, respectively, and the compressibility value was 3×10−5 bar−1. Electrostatics were computed by the PME method using a relative dielectric constant of εr = 2.5, as is the suggested protocol for MARTINI simulations using polarizable water. The simulation was performed utilizing periodic boundary conditions (PBC), with a time step of 20 fs. The nonbonded Lennard-Jones (LJ) interactions were cut off at 1.2 nm. These simulations were run for 1.5 μs. GROMACS analysis tool g_dist was used to calculate the difference between the center of mass (COM) distance between the γ1 peptide and the bilayer.
To generate systems VII and VIII, which involved converting a mixed bilayer to a pure bilayer, the following protocol was used. In the CG representation both POPC and POPG contain 13 beads, with only one different bead type differentiating between the choline and glycerol moiety in the head group region of POPC and POPG, respectively. Specifically, the bead NC3 (type Q0; charge +1) of POPC was changed to bead GL0 (type P4; charge zero) of POPG and vice-versa. Therefore, the bound state configuration of system VI (mix, SSoff) was converted to either pure POPC (System VII) or pure POPG (System VIII), without altering the peptide configuration. Both systems, VII and VIII were re-neutralized with a proper number of Na+ and Cl− ions to make the systems electrically neutral at 0.1 M salt concentration. Each system was energy minimized using the steepest descent algorithm before being converted to an all-atom representation.
Membrane binding free energies were calculated in the MARTINI model for Systems I–VI. With the exception of System II (POPC, random coil γ1), we were able to observe spontaneous bilayer association when γ1 is initially placed in the bulk solvent. We took the final configuration from the 1.5 μs binding simulation and perform steered MD (SMD) to generate intermediate conformations for umbrella sampling. The SMD runs were performed by applying a harmonic biasing force to the COM Z-coordinate distance between the peptide and the bilayer. A force constant of 1000 kJ mol−1 nm−2 was used and the center of the bias was moved at a rate of 0.001 nm/ps. Positional harmonic restraints with a force constant of 1000 kJ mol−1 nm−2 were applied on the lipid atoms during the SMD and intermediate states were extracted in 0.1 nm spacing ranging up to 3 nm displacement from the bound state. Positional restraints were released and umbrella sampling was performed on the COM Z-coordinate distance between the bilayer and the peptide for 300 ns with an umbrella potential of 2000 kJ mol−1 nm−2 for each window. The data from the last 100 ns of each umbrella window was analyzed using the weighted histogram analysis method (WHAM)29 implemented in the GROMACS g_wham tool, to calculate the potential of mean force (PMF) of peptide-bilayer binding. The PMF for System II was generated in the same manner as all other systems, except that the initial configuration for the SMD was the same as System I, but with the SS restraints turned off (random coil). All reported CG times are the actual simulation times without any scaling factor applied; a common practice is to estimate the “true” time by multiplying the simulation time by a factor of 4.
2.3. All-Atom (AA) simulations
Bound configurations from the CG simulations were transformed to AA descriptions for systems I–VIII to allow for more detailed structural interrogation of the binding mode and folding characteristics of γ1. This was achieved by “back-mapping” to atomic coordinates, utilizing the geometrical approach by Teileman30 to convert from MARTINI to an AA description for simulations in the CHARMM36 forcefield.31,32
For FHV γ1, 21 CG beads were mapped to 328 atoms. The atomistic γ1 had a charge of +2 with a protonated N-terminus and a negatively charged C-terminus. For the POPC lipid, 134 atoms were mapped from the 13 CG beads, and for POPG lipid 127 atoms were mapped from the 13 CG beads. The three bead CG polarizable water was mapped to the TIP3P water model. The ions Na+ and Cl− were also back-mapped into the AA configuration. Each back-mapped system was minimized using the steepest descent algorithm. Following the minimization, the systems were equilibrated in a constant volume, constant temperature (NVT) ensemble for 10 ns. Post equilibration, the production simulations were carried out for 1 μs in the NPT ensemble. Parrinello-Rahman semi-isotropic pressure coupling was used to maintain the pressure at 1.0 bar and the temperature was maintained at 303 K using the velocity (v)-rescale method. The time constants for the pressure and temperature couplings were 1.0 ps and the compressibility value was 4.5 × 10−5 bar−1. Non-bonded short range interactions were cutoff at 1.2 nm with the LJ interaction being force-switched starting at 1.0 nm. Long-range electrostatics were calculated by PME, bonds involving hydrogen atoms lengths were constrained with the LINCS constraint algorithm. The simulations were performed using PBC in X,Y and Z-directions with a time step of 2 fs.
2.4. Trajectory analysis
The trajectories were analyzed using multiple analysis tools from the GROMACS, MDTraj33 and CHARMM34 packages. Total helicity was computed every 1 ns using the COOR SECS function in CHARMM with a cutoff distance of 2.6 Å. Orientation (tilt) angle of γ1 was measured by calculating the arccosine of the dot product between the bilayer normal (0, 0, 1) and the normalized vector along the helical axis of γ1. The helicity per residue was computed using the DSSP35 tool of MDTraj. Distance calculations between COM of bilayer phosphate group and peptide (in Z-direction) were done using the g_traj tool in GROMACS. Also, the g_mindist tool from GROMACS was utilized to calculate the number of contacts γ1 makes with lipid molecules within a 5 Å radius. The ion release entropy calculation was based upon constructing a normalized probability distribution for Na+ ions in the range of 0 to 2.8 nm from the bilayer. This was done by calculating the minimum distance (R) for each Na+ ion to it’s nearest lipid phosphate in the upper leaflet and we assumed that p(R)≈ p(Z) for the histogramming.
3. Results
3.1 γ1 is partially disordered in solution
While the γ1 peptide is known to be well ordered inside the viral capsid, it undergoes an order-disorder transition, when the peptide exits the capsid and is free in solution. However, structural insight into the disordered state or insight into how energetically unfavorable it is to regain structure is lacking. We have conducted AA WT-MetaD simulations17 in explicit solvent to calculate the free energy profile of structural transitions of the γ1 peptide in an aqueous environment. The simulation was conducted for 190 ns and the FES displays good convergence (Fig. S1). We observe that the energy landscape along the end-to-end distance variable, shows a compact structure to be the free energy minimum (Fig. 2). Based upon our definition, a full helix would have an end-to-end distance of ~29 Å, which is not a stable region in the landscape. We have performed structural clustering at several points along the free energy profile and have found the average helicity in the different clusters ranges from 29% to 43%. The cluster centroids, shown in Fig. 2, show that helicity is generally maintained in the N-terminal region, while the C-terminal region adopts a random coil configuration. The results are in reasonable agreement with CD measurements, which have predicted ~20% α-helical and 40% random coil for γ1 in buffer.36
Figure 2.

Free energy profile of γ1 from WT-MetaD. Six cluster centroid structures are shown, the clusters were constructed based upon the end-to-end CV. The following CV ranges were used in constructing the clusters: 16–18 Å, 24–26 Å, 30–32 Å, 35–37 Å, 38–40 Å, 42–44 Å.
3.2. γ1 binding to membranes is dependent on peptide structure and lipid composition
The spontaneous binding of a small peptide from solution to the membrane surface can be a slow process due to dependencies on peptide folding and lipid diffusion rates.37 Therefore the peptide membrane binding process may not be well suited for investigation by AA simulations. The MARTINI CG model22,24 offers an alternative model in which microsecond timescales can be reached with reasonable computational expense. The model has been shown to be effective at predicting membrane partitioning free energies of short peptides,38 but the model is not designed to capture secondary-structure (SS) changes in a protein. Therefore, as a way to circumvent this limitation in MARTINI, we performed membrane binding simulations both when no secondary structure restraints were imposed on γ1 (referred to as SSoff), and when γ1 was restrained to an α-helical conformation (referred to as SSon). We examined the spontaneous membrane binding of γ1 as function of SS restaints and as a function of lipid composition of the bilayer. The six CG binding simulations are described in Table 1, and include bilayers composed purely of POPC (zwitterionic), POPG (negatively charged) or a heterogeneous bilayer containing an equal number of POPC and POPG lipids, referred to as a mixed (mix) bilayer.
The center of mass (COM) separation between γ1 and the bilayer was monitored (Fig. 3) in simulations in which γ1 was initiated in the solvent, ~4 nm away from the bilayer. We find that γ1 only transiently binds to a POPC bilayer when it is in a disordered state (System II). However, upon applying the SS restraints (System I) a helical γ1 binds to the neutral POPC bilayer after ~500 ns, as shown in Fig. 3. We next examined the spontaneous binding of γ1 to charged bilayers as described by Systems III–VI in Table 1. We observed that by introducing negatively charged POPG lipids, γ1 would bind bilayers in both helical and random coil configurations, whereas binding to neutral POPC was dependent on formation of helical structure. Fig. 3 illustrates that γ1 binds pure POPG or a mixed POPC/POPG bilayer much faster (under 200 ns) than it binds a POPC bilayer. This feature can be attributed to the electrostatic attraction between the positively (+2) charged γ1 and the presence of net negative charge in the bilayer. It may also be noted from Fig. 3 that the helical γ1 peptide inserts deeper into the POPC bilayer than the disordered peptides insert into the charged bilayers.
Figure 3.
The COM separation between γ1 and the bilayer. The peptide was modeled in an unstructured state for the binding against PC (red), PG(green) and the mixed bilayer (pink), as well as in a helical state against a PC bilayer (blue). Stable associations were observed for all systems except for the unstructured peptide against PC (red). For clarity purposes the data was smoothed by performing 6 ns running averages.
We further characterized the γ1 membrane binding properties by performing umbrella sampling simulations to obtain the binding free energy for Systems I–VI listed in Table 1. Since disordered γ1 (SSoff) never forms a stable complex with a POPC bilayer, the bound state of System I was used as the starting configuration for the umbrella sampling of System II. For all other systems the initial configuration and SS restraints were consistent with the spontaneous binding simulations. Fig. 4 presents the 1D free energy profile (potential of mean force, PMF) for the six systems that were umbrella sampled. We observe a ΔG of binding of −1.5 kcal/mol, −11.0 kcal/mol, −7.5 kcal/mol with POPC, POPG and the mixed bilayers, respectively, when SS restraints are not imposed (SSoff). However, the bilayer-γ1 binding with a helical γ1 (SSon) show considerably stronger binding for all systems. With SSon we observe a ΔG of binding of −15.0 kcal/mol, −22.5 kcal/mol, −18.5 kcal/mol with POPC, POPG and the mixed bilayer, respectively.
Figure 4.
PMF of γ1 binding to membranes from umbrella sampling the COM displacement from the bound configuration. For all membrane compositions, the umbrella sampling was performed both with γ1 in a helical configuration (SSon) and in an unstructured state (SSoff). When the peptide was in a helical state, the binding free energy increased, though the binding to PC as a helix (red solid) and PG as a coil (blue dashed) show compared binding free energies.
Experimentally the γ1 adsorption free energies have been measured against POPC and POPG bilayers by ITC and nearly equivalent values are observed (−7.5 and −8.0 kcal/mol, respectively).15 While the magnitudes of our PMFs are larger, we observed similar magnitude ΔGs (−11.0 vs −15.0 kcal/mol) when comparing ordered γ1 binding to POPC and disordered γ1 binding to POPG. This observation appears reasonable in consideration that the calorimetry studies also reported that binding to POPC membranes is enthalpically driven while binding to POPG membranes is entropically driven. Our studies provide a rationale for the thermodynamic measurements, that γ1 can bind to POPG in a disordered state but must adopt a helical conformation when binding POPC. Energetically, association of γ1 to POPG offers a lower conformational entropic loss thus providing a means to compensate for the decreased enthalpic contribution to binding.
Solvent effects can also have a significant contribution to the protein binding thermodynamics,39 and we investigate an aspect of this by computing the entropic gain from cation desorption from the bilayer surface when γ1 binds to the bilayer. We have computed the sodium ion probability distributions (p(Z)) in the direction normal to the peptide binding bilayer (upper) leaflet, for POPC and POPG bilayers when γ1 is bound and when the peptide is in solution. From the probability distributions the entropy can be computed as
| (2) |
where kB is Boltzmann’s constant, and N is the number of ions in the system.40 The change in ion translational entropy upon peptide binding can be computed as ΔS = Sbound-Sunbound, and the contribution to the free energy of binding as −TΔS. The changes to p(Z) upon binding γ1 are slight for both the POPC and POPG systems, however due to the larger number of cations in the POPG simulations to neutralize the system, we find that the entropic gain due to ion release is approximately 10-fold larger in the POPG system than the POPC system. With a histogram bin width of 0.5 Å, we compute −TΔSPOPG = −6.5 kcal/mol and −TΔSPOPC = −0.7 kcal/mol, indicating that the ion release entropy gain is a significant contribution in the binding of γ1 to POPG bilayers. Regarding the magnitudes of the computed binding ΔGs it is possible the CG model overestimates these quantities. However a more recent study measuring surface pressure changes in a Langmuir film balance, calculated a Kd = 2.8 nM for γ1 binding to DPPC/DPPS (4:1) monolayers.41 This Kd would correspond to a ΔG = −14 kcal/mol, according to eq (3)
| (3) |
which is more consistent with our estimates of the binding free energy magnitudes.
3.3. Influence of lipid composition on the membrane bound folding properties of γ1
To understand the effect of lipid composition on the structural aspects of γ1 binding, we move to AA simulations to study the structural evolution of γ1 when interacting with bilayers. We first analyze Systems I and IV, which displayed similar binding free energies despite significant differences in the secondary structure content of the peptide. We converted CG structures from the end of the 1.5 μs spontaneous binding simulations to AA configurations and then performed 1μs AA simulations. The initial and final snapshots of these AA simulations are shown in Fig. 5. Our final configuration for POPC (System I) and POPG (System IV) shows that γ1 remains helical on POPC and embeds into the hydrophobic region of the bilayer, while γ1 on POPG remains disordered with 0% helicity on the bilayer periphery. From these detailed AA simulations, the difference in conformational state of γ1 on POPC (ordered/low entropy) as compared to POPG (disordered/high entropy) again corroborates with the thermodynamic decomposition of γ1 binding to POPC and POPG bilayers.
Figure 5.

AA membrane bound simulations. Simulations were initiated from CG bound structures to POPC (A) and POPG (C) bilayers. Simulations were run for 1 μs on each bilayer and end states are shown in (B) and (D). We observe that γ1 in the simulation on POPC remains helical and embeds deeper in the membrane, while the simulation on POPG shows that γ1 remains unstructured and bound to the membrane surface without partitioning into the hydrophobic core.
The AA simulations are supported by the strong agreement with experimental measurements of the configuration of γ1 on POPC bilayers. We have computed the helicity, TRP4 (reside 367 by capsid numbering) insertion depth, and helical orientation relative to the bilayer normal during the final 100 ns of γ1-POPC simulation (System I) and find all these quantities to be well correlated with experimental measurements.13 The helicity in our simulations is calculated to be 78 +/− 4%, which matches the 75% helicity as determined from CD. The TRP4 distance to the bilayer mid-plane is measured to be approximately 1 nm from the center of a POPC bilayer and our simulations show TRP4 inserts to a distance of 1.2 +/− 0.2 nm from the bilayer center. The angle the helical axis makes with the bilayer normal was measured to between 66° – 78°, while in the simulation the peptide samples a wider range of angles between 65° – 90°, with an average value of 81° +/− 6° This excellent agreement with the structural and thermodynamic experimental measurements gives us confidence in predicting and analyzing the binding, folding and insertion properties of FHV γ1 in systems which have not been experimentally examined.
The simulations described above support a conformational dependence of γ1 as a function of bilayer composition. However, even during 1 μs simulations, we cannot discount that kinetic barriers may exist, to either nucleate folding or unfolding, which are not overcome on the timescale of the simulations. Therefore, we further explored the membrane composition sensitive folding propensities, by starting simulations on POPC, POPG and the mixed bilayer that are initiated in a consistent, partially folded configuration. To do this we back-mapped the CG configuration of γ1 bound to the mixed bilayer (Fig. 6A; System VI) and transformed the bilayer to either a pure POPC (System VII) or pure POPG bilayer (System VIII). This approach was adopted to try to eliminate any bias introduced by differing initial peptide configurations in the folding behavior of γ1.
Figure 6.
Membrane folding of γ1 on different composition bilayers. A) The initial configuration for all simulations is derived from the CG simulation of γ1 binding to a mixed bilayer with SSoff. The final configuration after 1 μs of simulation is shown when the configuration from (A) is run on mixed POPC/POPG bilayer, pure POPG bilayer and pure POPC bilayer in (B), (C) and (D), respectively. The progression of helicity is shown in (E), and the per-residue helicity averaged over the final 100 ns is shown in (F). The residue insertions with respect to the lipid phosphate groups is shown in (G), which shows insertion depth and folding (F) are strongly correlated. In (A)–(D) the POPG lipids are colored greenm the POPC lipids are gray and the peptide is blue.
The results from these simulations initiated from a consistent peptide conformation are presented in Fig. 6. Consistent with the AA simulations of System I and System IV, the simulation on POPC gains helical content reaching 63% helicity during the final 100 ns (Fig. 6D,E). Whereas γ1 on POPG maintains a relatively low helical content of only 23% helicity over the final 100 ns (Fig. 6C,E). These “lipid-substitution” simulations, re-affirm our understanding that the lipid composition affects the bound conformational state of γ1 as the peptide folds or remains helical on a POPC bilayer (System I and VII). On the other hand, when bound to an anionic POPG bilayer, γ1 remains in a low helicity disordered state (System IV and VIII).
However, the behavior of γ1 on the mixed POPC/POPG bilayer cannot be as simply interpreted. There does not appear to be a linear relationship between the charge content in the membrane and the folding capacity of the peptide, as the mixed bilayer does not display an intermediate degree of helicity between POPC and POPG. After 1 μs of equilibrium simulation, as shown in Fig. 6B γ1 appears as a random coil structure with a low helical content of 16%. To comprehend the structural dynamics of γ1 in this system we focus on the micro-environment around γ1 during the simulation. The local environment around the initial conformation of γ1 on the mixed bilayer (Fig. 6A) is dominated by POPG contacts. During the first 100 ns of the 1 μs simulation the average ratio of PG to PC contacts is 4.6:1. The contacts evolve to a more equalized ratio, through the simulation, but PG contacts still dominate, the average ratio during the final 100 ns is 1.4:1::PG:PC. The ratio of PC:PG contacts during the 1 μs simulation are presented in Fig. S2. One possibility why a greater degree of folding is not seen in the mixed system is that slow diffusion of the lipids prevents formation of a POPC dominated microdomain to nucleate γ1 folding and insertion.
The peptide insertion depth does appear to be an important quantity as it correlates strongly with formation of secondary structure in γ1. In Systems VI–VIII simulations, helix formation is observed in the central region (residue 9–13) of peptide, but only in system VII (pure POPC) does helicity develop in the C-terminal region (Fig. 6F). We also observe that γ1 inserts into the hydrophobic region to a much greater extent, especially in the C-terminal region, when comparing POPC against POPG for the simulations converted from the mixed bilayer system (Fig. 6G). The average insertion depth of γ1 on POPC is 6.9 Å below the headgroups, while the helical residues (6–19) insert further to an average depth of 8.6 Å. Similar trends are seen on POPG where the overall average residue insertion depth is 1.3 Å and 3.6 Å for the helical residues (9–14). Likewise, on the mixed bilayer γ1 has an average insertion depth of 0.1 Å above the headgroups while the helical residue (9–13) are inserted to an average depth of 2.5 Å. The decreased insertion depth into POPG bilayers has been noted in the experimental literature as well.15
3.4 To fold or not to fold?
Functionally we expect folding to be beneficial to the membrane disrupting activity of the peptide to create transmembrane structures and hence an important component to viral infection. The observed folding characteristics in this study may help in understanding why γ1 elicits more robust membrane disruption of POPC vs. POPG vesicles in-vitro.13 However, from an energetic perspective, our CG simulations indicate that the folded state is always thermodynamically more favorable, regardless of membrane composition (Fig. 4). The nature of the MARTINI model and its inability to correctly model the protein dynamics, as well as the noted enthalpy/entropy imbalance42 could be causes for this discrepancy between the CG and AA simulation. Another possibility is inadequate sampling in the AA simulations due to large energy barriers in the folding pathway of γ1. To address these possibilities we conducted additional simulations (System III and V) to explore if helical conformations were stable on the POPG or mixed bilayers in the AA model. If the presences of anionic lipids in the bilayer were highly destabilizing to the helical conformation, we would expect to see rapid unfolding of γ1 when initiated in a helical conformation. However, we do not observe unfolding in these simulations, and γ1 maintains high helicity content through the simulation.
A plot showing the insertion depth for each residue with respect to the COM of the top leaflet phosphates is shown in Fig. 7. For each system, we have compared the initial 100 ns to the final 100 ns of the 1 μs simulations, to judge the extend of insertion stability. We observe that the insertion depth is stable in POPC (Fig. 7A) and POPG (Fig. 7B), while the mixed system shows the N-terminus inserting deeper while the C-terminus moves toward the surface (Fig. 7C). The insertion profile for POPC shows the deepest insertion and is consistent with the insertion profile of System VII (Fig. 6G). The average insertion depth per residue during the final 100 ns residue is 6.0 Å, 3.3 Å, and 4.9 Å, for the POPC, POPG and mix bilayer systems, respectively. The increased insertion depth into the POPC and mixed bilayers compared to the POPG system may indicate a lower energy barrier to transition to a transmembrane configuration.
Figure 7.
Residual membrane insertion depth for systems with γ1 initiated in a helical conformation. The depths are averaged over the first (red) and last (blue) 100 ns. Simulations were performed on A) POPC (System I), B) POPG (System III) and C) the mixed bilayer (System V).
4. Discussion
The results from these simulation studies provide qualitative explanations of the relatively high ΔG values when helical γ1 binds to the membrane, (Fig. 4) and underscores the time-scale limitation of AA MD (ns to μs) simulations to access processes like membrane insertion and folding of a peptide. An illustrative example is System IV (Fig. 6, S2), which shows that towards the end of 1 μs simulation, the micro-environment around γ1 has a comparable amount of POPC and POPG molecules. However, due to the limitation of the time-scales for folding, membrane insertion and/or lipid diffusion the peptide does not encounter a more concentrated POPC environment to undergo an entropic loss and gain helicity. Integrating our findings with previous studies we suggest an energetic pathway as shown in Fig. 8, that once a membrane-γ1 complex is formed a “folding” barrier must be overcome to gain helicity and insert into the membrane. This barrier appears to be a function of membrane charge (and other possible factors), and may be non-existent for neutral membranes where the binding and folding process can be coupled, which is schematically represented in Fig. 8. It is possible that the barrier is too high for the POPG system, and the peptide remains trapped in the disordered state, resulting in the observed low lytic activity.13
Figure 8.
Proposed energy landscape for γ1 to transition from an unstructured, solvated state to a helical, membrane bound state. We propose the pathway is modified based upon the lipid composition, and specifically the increase in anionic lipids may introduce/increase an energy barrier to reach the folded state.
Investigations into the molecular dynamics of membrane active peptides from non-enveloped viruses are quite limited as compared with anti-microbial peptides (AMP). However, a recent study has also examined the FHV γ-peptide using MD and experimental methods.43 In that study γ1, the full-length γ-peptide and a construct peptide in which the 8 C-terminal residues of γ were fused to γ1 (termed Δ385–399), were investigated to characterize their interactions with membranes. The Δ385–399 construct contained three Phe residues in the C-terminal region, which had been previously implicated to be critical for membrane disruption,14 yet Δ385–399 was less effective in disrupting membranes than γ1 or even a non-cleaving maturation defective mutant. Based upon 100 ns simulations of the peptides in the presence of a POPC bilayer, it was proposed that the full C-terminal region is critical in stabilizing the helicity in the N-terminal region. This proposal was based upon their simulations that showed i) γ1 displayed low helicity when bound to the bilayer and ii) a homology model of the full length γ peptide, which formed a helical hairpin structure, displayed a stabilized N-terminal helix on the bilayer. Both our simulations and previous experimental results indicate that γ1 should retain a high degree of helicity on a POPC bilayer and therefore the proposal that the C-terminal region acts as a stabilizer of the N-terminal helix may be disputable. Furthermore, the initial helical hairpin predicted structure may deviate from the authentic structure and the initial condition could significantly bias the simulation, especially on the 100 ns timescale.
While the literature on non-enveloped membrane active viral peptides is small, there has been a tremendous amount of work investigating the mechanism of AMP interactions with membranes We will briefly expand upon some studies on AMP systems which have a degree of similarity to FHV γ1. The size of FHV γ1 (21 amino acids) and it’s amphipathic character is similar to several AMPs such as alamethicin, piscidin, magainin and melittin, which have 20, 22, 23 and 26 amino acids, respectively. These AMPs also exhibit a similar structural behavior to γ1 as they are disordered in the solution phase but gain helicity upon interactions with a membrane.44–47 γ1 is a cationic peptide with a net charge of +2 at neutral pH, which is consistent with most AMPs,48 including melittin (+6), piscidin (+3) and magainin (+3). The thermodynamics of membrane binding of of the aforementioned AMPs is also similar to FHV γ1 (≈−7.5 kcal/mol), for example, the free energy of melittin binding to a 80/20 POPC:POPG bilayer is −8.7 kcal/mol49 and −6.2 kcal/mol for alamethicin binding to a POPC bilayer.50 Numerous mechanisms of action have been proposed for AMPs and for many systems the mode of membrane disruption remains unclear. Recent long time scale (multi μs) have been conducted on pore structures for melittin, piscidin 1, magainin (MAG2 and MG-H2) and PGLa, and the results vary between the systems. The peptides generally show dynamic pores, with the peptides being able to transition between a transmembrane (T) and surface bound state (S). For melittin, magainin, and PGLa a toroidal pore structure is supported, in which lipid head groups are pulled into the bilayer interior and (some) peptides are stable in the T-state.51,52 However, for piscidin 1 a pore structure was not stable as the peptides preferred S-state orientations.53 Transient membrane defects were observed for piscidin 1, which were characterized by the deep insertion of the C-terminal region of the peptide.53 Based upon the size and charge similarity of γ1 to piscidin 1, and that we also observe deep C-terminal insertion (Fig. 6G), we would predict these peptides to share a similar mechanism of action, but further studies will be required to test this prediction.
Enveloped virus membrane fusion is a heavily studied field and these systems have key differences but also share some commonalities with non-enveloped virus membrane disruption. 54 Influenza virus is best characterized system and several molecular dynamics studies have investigated the fusion peptide-membrane interactions to understand the structural and energetic aspects of this interaction.55–59 The role of enveloped virus fusion proteins is not to directly disrupt the cellular membrane, as is the case for non-enveloped viruses, but to lower the energy barrier associated with fusing the viral and cellular membranes. Enveloped viruses can have multiple surface glycoproteins, one of which is the fusion protein. There are three classes of fusion proteins: class I consist of mainly α-helical, class II are largely β-structures and class III contain both α-helical and β-structures. Fusion proteins require a priming step which involves a proteolytic cleavage event, to make them fusion competent, this is analogous to the maturation process in FHV which cleaves the γ-peptide from the capsid protein. Enveloped fusion is often triggered by exposure to low pH. In many cases the low-pH environment is required but not sufficient for fusion, while in other cases such as influenza and dengue virus, low-pH alone is sufficient to trigger fusion, similar to FHV. Fusion proteins have either an N-terminal fusion peptide or an internal fusion loop, generally on the tips of an extended β-sheet. The fusion peptides are conformationally diverse and dynamics, as the undergo conformational changes when they transition to the membrane pre-fusion state. A key difference between FHV and enveloped viruses is that there is a well defined oligomeric state in the post-fusion conformation, as all known cases exist in a trimer of hairpins, whereas the oligomeric state of FHV is unknown and may not be well-defined. Presenting the fusion pepetides/loops to the cellular membrane and the folding-back process to fuse the two membranes both involve large conformational changes to the fusion proteins, which is not analogous to the FHV mechanism. Overall enveloped virus fusion machinery is larger and the mechanisms are more complex than FHV membrane disruption.
In this work we have extended our understanding of FHV γ1 membrane interactions through a multi-scale computational approach, which combined CG MARTINI, AA metadynamics and 8 μs of equilibrium AA simulations. We were able to observe that γ1 membrane binding is consistent with several experimental measurement and provides a justification of experimentally known thermodynamic energy decomposition of γ1 binding to POPC and POPG bilayers. Binding of γ1 to a negatively charged POPG membrane is entropically driven and strongly influenced by electrostatic effects, including the release of counter ions form the membrane surface. However, γ1 adopts a helical conformation while undergoing an entropic loss when binding to neutral POPC bilayers, which is an exothermically driven binding process. Our simulations also revealed that the structural dynamics of γ1 are governed not only by the bulk lipid composition, but the micro-environment in the vicinity of γ1 also plays a crucial role. Assimilating the CG and AA simulations conducted in this study we have proposed an energetic pathway of γ1 binding. γ1 will be rapidly attracted to negatively charged membranes and will adsorb in a disordered state, where the peptide has to overcome an energy barrier in order to fold and insert into the membrane, and the size of this barrier may limit the membrane activity of the peptide. Overall this study provides insights into the conformational dynamics of FHV γ1 and suggests the virus may be evolved for efficient binding and folding against the endocytic membranes the peptides would encounter during the virus infection pathway. Further investigations utilizing enhanced sampling methods and a wider range of lipid compositions are warranted to further tease apart these details.
Supplementary Material
Highlights.
Thermodynamic driving force of γ1 membrane binding is dissected
Membrane composition affects peptide folding rates
Increased insertion depth in neutral membranes may relate to increased lysis
Acknowledgments
The work has been supported by the National Institutes of Health through grants K22AI099163 and R35GM119762. Computational resources have been provided through the University of Connecticut Booth Engineering Center for Advanced Technology.
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.Harrison SC. Viral membrane fusion. Nat Struct Mol Biol. 2008;15:690–698. doi: 10.1038/nsmb.1456. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Weissenhorn W, Hinz A, Gaudin Y. Virus membrane fusion. FEBS Letters. 2007;581:2150–2155. doi: 10.1016/j.febslet.2007.01.093. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Moyer CL, Nemerow GR. Viral weapons of membrane destruction: variable modes of membrane penetration by non-enveloped viruses. Current Opinion in Virology. 2011;1:44–49. doi: 10.1016/j.coviro.2011.05.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Banerjee M, Johnson JE. Activation, exposure and penetration of virally encoded, membrane-active polypeptides during non-enveloped virus entry. Curr Protein Pept Sci. 2008;9:16–27. doi: 10.2174/138920308783565732. [DOI] [PubMed] [Google Scholar]
- 5.Odegard A, Banerjee M, Johnson JE. Flock house virus: a model system for understanding non-enveloped virus entry and membrane penetration. Curr Top Microbiol Immunol. 2010;343:1–22. doi: 10.1007/82_2010_35. [DOI] [PubMed] [Google Scholar]
- 6.Schneemann A, Zhong W, Gallagher TM, Rueckert RR. Maturation cleavage required for infectivity of a nodavirus. Journal of Virology. 1992;66:6728–6734. doi: 10.1128/jvi.66.11.6728-6734.1992. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Zlotnick A, et al. Capsid assembly in a family of animal viruses primes an autoproteolytic maturation that depends on a single aspartic acid residue. J Biol Chem. 1994;269:13680–13684. [PubMed] [Google Scholar]
- 8.Fisher AJ, Johnson JE. Ordered duplex RNA controls capsid architecture in an icosahedral animal virus. Nature. 1993;361:176–179. doi: 10.1038/361176a0. [DOI] [PubMed] [Google Scholar]
- 9.Odegard AL, et al. Low Endocytic pH and Capsid Protein Autocleavage Are Critical Components of Flock House Virus Cell Entry. Journal of Virology. 2009;83:8628–8637. doi: 10.1128/JVI.00873-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Walukiewicz HE, Banerjee M, Schneemann A, Johnson JE. Rescue of maturation-defective flock house virus infectivity with noninfectious, mature, viruslike particles. Journal of Virology. 2008;82:2025–2027. doi: 10.1128/JVI.02278-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Bothner B, Dong XF, Bibbs L, Johnson JE, Siuzdak G. Evidence of viral capsid dynamics using limited proteolysis and mass spectrometry. J Biol Chem. 1998;273:673–676. doi: 10.1074/jbc.273.2.673. [DOI] [PubMed] [Google Scholar]
- 12.Maia LF, et al. Structure of a membrane-binding domain from a non-enveloped animal virus: insights into the mechanism of membrane permeability and cellular entry. J Biol Chem. 2006;281:29278–29286. doi: 10.1074/jbc.M604689200. [DOI] [PubMed] [Google Scholar]
- 13.Bong DT, Steinem C, Janshoff A, Johnson JE, Reza Ghadiri M. A highly membrane-active peptide in Flock House virus: implications for the mechanism of nodavirus infection. Chemistry & Biology. 1999;6:473–481. doi: 10.1016/s1074-5521(99)80065-9. [DOI] [PubMed] [Google Scholar]
- 14.Banerjee M, et al. Dissecting the functional domains of a nonenveloped virus membrane penetration peptide. Journal of Virology. 2009;83:6929–6933. doi: 10.1128/JVI.02299-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Bong DT, Janshoff A, Steinem C, Ghadiri MR. Membrane partitioning of the cleavage peptide in flock house virus. Biophys J. 2000;78:839–845. doi: 10.1016/S0006-3495(00)76641-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Humphrey W, Dalke A, Schulten K. VMD: visual molecular dynamics. J Mol Graph. 1996;14:33–8. 27–8. doi: 10.1016/0263-7855(96)00018-5. [DOI] [PubMed] [Google Scholar]
- 17.Barducci A, Bussi G, Parrinello M. Well-Tempered Metadynamics: A Smoothly Converging and Tunable Free-Energy Method. Phys Rev Lett. 2008;100:020603. doi: 10.1103/PhysRevLett.100.020603. [DOI] [PubMed] [Google Scholar]
- 18.Phillips JC, et al. Scalable molecular dynamics with NAMD. J Comput Chem. 2005;26:1781–1802. doi: 10.1002/jcc.20289. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.MacKerell AD, et al. All-Atom Empirical Potential for Molecular Modeling and Dynamics Studies of Proteins †. J Phys Chem B. 1998;102:3586–3616. doi: 10.1021/jp973084f. [DOI] [PubMed] [Google Scholar]
- 20.MacKerell AD, Jr, Feig M, Brooks CL., III Extending the treatment of backbone energetics in protein force fields: Limitations of gas-phase quantum mechanics in reproducing protein conformational distributions in molecular dynamics simulations. J Comput Chem. 2004;25:1400–1415. doi: 10.1002/jcc.20065. [DOI] [PubMed] [Google Scholar]
- 21.Essmann U, et al. A smooth particle mesh Ewald method. J Chem Phys. 1995;103:8577–8593. [Google Scholar]
- 22.Marrink SJ, Risselada HJ, Yefimov S, Tieleman DP, de Vries AH. The MARTINI Force Field: Coarse Grained Model for Biomolecular Simulations. J Phys Chem B. 2007;111:7812–7824. doi: 10.1021/jp071097f. [DOI] [PubMed] [Google Scholar]
- 23.Monticelli L, et al. The MARTINI Coarse-Grained Force Field: Extension to Proteins. J Chem Theory Comput. 2008;4:819–834. doi: 10.1021/ct700324x. [DOI] [PubMed] [Google Scholar]
- 24.de Jong DH, et al. Improved Parameters for the Martini Coarse-Grained Protein Force Field. J Chem Theory Comput. 2013;9:687–697. doi: 10.1021/ct300646g. [DOI] [PubMed] [Google Scholar]
- 25.Hess B, Kutzner C, Van Der Spoel D, Lindahl E. GROMACS 4: Algorithms for Highly Efficient, Load-Balanced, and Scalable Molecular Simulation. J Chem Theory Comput. 2008;4:435–447. doi: 10.1021/ct700301q. [DOI] [PubMed] [Google Scholar]
- 26.Wassenaar TA, Ingólfsson HI, Böckmann RA, Tieleman DP, Marrink SJ. Computational Lipidomics with insane: A Versatile Tool for Generating Custom Membranes for Molecular Simulations. J Chem Theory Comput. 2015;11:2144–2155. doi: 10.1021/acs.jctc.5b00209. [DOI] [PubMed] [Google Scholar]
- 27.Yesylevskyy SO, Schäfer LV, Sengupta D, Marrink SJ. Polarizable Water Model for the Coarse-Grained MARTINI Force Field. PLoS Comput Biol. 2010;6:e1000810. doi: 10.1371/journal.pcbi.1000810. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Berendsen HJC, Postma JPM, van Gunsteren WF, Dinola A, Haak JR. Molecular dynamics with coupling to an external bath. J Chem Phys. 1984;81:3684–3690. [Google Scholar]
- 29.Roux B. The calculation of the potential of mean force using computer simulations. Comp Phys Comm. 1995;91:275–282. [Google Scholar]
- 30.Wassenaar TA, Pluhackova K, Böckmann RA, Marrink SJ, Tieleman DP. Going Backward: A Flexible Geometric Approach to Reverse Transformation from Coarse Grained to Atomistic Models. J Chem Theory Comput. 2014;10:676–690. doi: 10.1021/ct400617g. [DOI] [PubMed] [Google Scholar]
- 31.Klauda JB, et al. Update of the CHARMM all-atom additive force field for lipids: validation on six lipid types. J Phys Chem B. 2010;114:7830–7843. doi: 10.1021/jp101759q. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Best RB, et al. Optimization of the additive CHARMM all-atom protein force field targeting improved sampling of the backbone φ, ψ and side-chain χ(1) and χ(2) dihedral angles. J Chem Theory Comput. 2012;8:3257–3273. doi: 10.1021/ct300400x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.McGibbon RT, et al. MDTraj: A Modern Open Library for the Analysis of Molecular Dynamics Trajectories. Biophys J. 2015;109:1528–1532. doi: 10.1016/j.bpj.2015.08.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Brooks BR, et al. CHARMM: The biomolecular simulation program. J Comput Chem. 2009;30:1545–1614. doi: 10.1002/jcc.21287. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Kabsch W, Sander C. Dictionary of protein secondary structure: pattern recognition of hydrogen-bonded and geometrical features. Biopolymers. 1983;22:2577–2637. doi: 10.1002/bip.360221211. [DOI] [PubMed] [Google Scholar]
- 36.Janshoff A, Bong DT, Steinem C, Johnson JE, Ghadiri MR. An Animal Virus-Derived Peptide Switches Membrane Morphology: Possible Relevance to Nodaviral Transfection Processes †. Biochemistry. 1999;38:5328–5336. doi: 10.1021/bi982976i. [DOI] [PubMed] [Google Scholar]
- 37.Lin D, Grossfield A. Thermodynamics of antimicrobial lipopeptide binding to membranes: origins of affinity and selectivity. Biophys J. 2014;107:1862–1872. doi: 10.1016/j.bpj.2014.08.026. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Singh G, Tieleman DP. Using the Wimley–White Hydrophobicity Scale as a Direct Quantitative Test of Force Fields: The MARTINI Coarse-Grained Model. J Chem Theory Comput. 2011;7:2316–2324. doi: 10.1021/ct2002623. [DOI] [PubMed] [Google Scholar]
- 39.Leavitt S, Freire E. Direct measurement of protein binding energetics by isothermal titration calorimetry. Current Opinion in Structural Biology. 2001;11:560–566. doi: 10.1016/s0959-440x(00)00248-7. [DOI] [PubMed] [Google Scholar]
- 40.von Deuster CIE, Knecht V. Antimicrobial selectivity based on zwitterionic lipids and underlying balance of interactions. BBA - Biomembranes. 2012;1818:2192–2201. doi: 10.1016/j.bbamem.2012.05.012. [DOI] [PubMed] [Google Scholar]
- 41.Hinz A, Galla HJ. Viral membrane penetration: lytic activity of a nodaviral fusion peptide. Eur Biophys J. 2005;34:285–293. doi: 10.1007/s00249-004-0450-z. [DOI] [PubMed] [Google Scholar]
- 42.Marrink SJ, Tieleman DP. Perspective on the Martini model. Chem Soc Rev. 2013;42:6801. doi: 10.1039/c3cs60093a. [DOI] [PubMed] [Google Scholar]
- 43.Bajaj S, Dey D, Bhukar R, Kumar M, Banerjee M. Non-Enveloped Virus Entry: Structural Determinants and Mechanism of Functioning of a Viral Lytic Peptide. Journal of Molecular Biology. 2016:1–17. doi: 10.1016/j.jmb.2016.06.006. [DOI] [PubMed] [Google Scholar]
- 44.Tieleman DP, Berendsen H, Sansom M. Surface binding of alamethicin stabilizes its helical structure: molecular dynamics simulations. Biophys J. 1999 doi: 10.1016/S0006-3495(99)77470-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Irudayam SJ, Berkowitz ML. Binding and reorientation of melittin in a POPC bilayer: computer simulations. Biochim Biophys Acta. 2012;1818:2975–2981. doi: 10.1016/j.bbamem.2012.07.026. [DOI] [PubMed] [Google Scholar]
- 46.Perrin BS, Jr, et al. High-Resolution Structures and Orientations of Antimicrobial Peptides Piscidin 1 and Piscidin 3 in Fluid Bilayers Reveal Tilting, Kinking, and Bilayer Immersion. J Am Chem Soc. 2014;136:3491–3504. doi: 10.1021/ja411119m. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Leontiadou H, Mark AE, Marrink SJ. Antimicrobial Peptides in Action. J Am Chem Soc. 2006;128:12156–12161. doi: 10.1021/ja062927q. [DOI] [PubMed] [Google Scholar]
- 48.Last NB, Schlamadinger DE, Miranker AD. A common landscape for membrane-active peptides. Protein Science. 2013;22:870–882. doi: 10.1002/pro.2274. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Beschiaschvili G, Seelig J. Melittin binding to mixed phosphatidylglycerol/phosphatidylcholine membranes. Biochemistry. 1990;29:52–58. doi: 10.1021/bi00453a007. [DOI] [PubMed] [Google Scholar]
- 50.Lewis JR, Cafiso DS. Correlation between the Free Energy of a Channel-Forming Voltage-Gated Peptide and the Spontaneous Curvature of Bilayer Lipids †. Biochemistry. 1999;38:5932–5938. doi: 10.1021/bi9828167. [DOI] [PubMed] [Google Scholar]
- 51.Leveritt JM, III, Pino-Angeles A, Lazaridis T. The Structure of a Melittin-Stabilized Pore. Biophys J. 2015;108:2424–2426. doi: 10.1016/j.bpj.2015.04.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Pino-Angeles A, Leveritt JM, Lazaridis T. Pore Structure and Synergy in Antimicrobial Peptides of the Magainin Family. PLoS Comput Biol. 2016;12:e1004570–17. doi: 10.1371/journal.pcbi.1004570. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Perrin BS, Fu R, Cotten ML, Pastor RW. Simulations of Membrane-Disrupting Peptides II: AMP Piscidin 1 Favors Surface Defects over Pores. Biophys J. 2016;111:1258–1266. doi: 10.1016/j.bpj.2016.08.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.White JM, Whittaker GR. Fusion of Enveloped Viruses in Endosomes. Traffic. 2016;17:593–614. doi: 10.1111/tra.12389. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Larsson P, Kasson PM. Lipid Tail Protrusion in Simulations Predicts Fusogenic Activity of Influenza Fusion Peptide Mutants and Conformational Models. PLoS Comput Biol. 2013;9:e1002950–9. doi: 10.1371/journal.pcbi.1002950. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Brice AR, Lazaridis T. Structure and Dynamics of a Fusion Peptide Helical Hairpin on the Membrane Surface: Comparison of Molecular Simulations and NMR. J Phys Chem B. 2014;118:4461–4470. doi: 10.1021/jp409412g. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Panahi A, Feig M. Conformational Sampling of Influenza Fusion Peptide in Membrane Bilayers as a Function of Termini and Protonation States. J Phys Chem B. 2010;114:1407–1416. doi: 10.1021/jp907366g. [DOI] [PubMed] [Google Scholar]
- 58.Baylon JL, Tajkhorshid E. Capturing Spontaneous Membrane Insertion of the Influenza Virus Hemagglutinin Fusion Peptide. J Phys Chem B. 2015;119:7882–7893. doi: 10.1021/acs.jpcb.5b02135. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Ward MD, Nangia S, May ER. Evaluation of the hybrid resolution PACE model for the study of folding, insertion, and pore formation of membrane associated peptides. J Comput Chem. 2017 doi: 10.1002/jcc.24694. [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.






