Вы находитесь на странице: 1из 10

This is an open access article published under an ACS AuthorChoice License, which permits

copying and redistribution of the article or any adaptations for non-commercial purposes.

Article

Cite This: ACS Omega 2018, 3, 6056−6065

Insights into Membrane Translocation of Protegrin Antimicrobial


Peptides by Multistep Molecular Dynamics Simulations
Pin-Kuang Lai* and Yiannis N. Kaznessis*,†
Department of Chemical Engineering and Materials Science, University of Minnesota, 421 Washington Avenue SE, Minneapolis,
Minnesota 55455, United States

ABSTRACT: Protegrin-1 (PG-1) is a cationic arginine-rich antimicro-


bial peptide. It is widely accepted that PG-1 induces membrane
disruption by forming pores that lead to cell death. However, the
insertion mechanism for these highly cationic peptides into the
hydrophobic membrane environment is still poorly understood at the
molecular scale. It has previously been determined that the association of
arginine guanidinium and lipid phosphate groups results in strong
bidentate bonds that stabilize peptide−lipid complexes. It has also been
suggested that arginine residues are able to drag phosphate groups as they
insert inside the membrane to form a toroidal pore. However, whether
bidentate bonds play a significant role in inducing a pore formation
remains unclear. To investigate the role of bidentate complexes in PG-1 translocation, we conducted molecular dynamics
simulations. Two computational electroporation methods were implemented to examine the translocation process. We found
that PG-1 could insert into the membrane, provided the external electric potential is large enough to first induce a water column
or a pore within the lipid bilayer membrane. We also found that the highly charged PG-1 is capable in itself of inducing molecular
electroporation. Substitution of arginines with charge-equivalent lysines showed a markedly reduced tendency for insertion. This
indicates that the guanidinium group likely facilitates PG-1 translocation. Potential of mean force calculations suggests that
peptide insertion inside the hydrophobic environment of the membrane core is not favored. We found that formation of a water
column or a pore might be a prerequisite for PG-1 translocation. We also found that PG-1 can stabilize the pore after insertion.
We suggest that PG-1 could be a pore inducer and stabilizer. This work sheds some light on PG-1 translocation mechanisms at
the molecular level. Methods presented in this study may be extended to other arginine-rich antimicrobial and cell-penetrating
peptides.

1. INTRODUCTION The majority of previous studies focused on certain stages of


Antimicrobial peptides (AMPs) are attracting widespread the pore formation process, such as early membrane surface
interest as an alternative to traditional antibiotic treatments binding and dimerization, or on the final pore structure.
and a potential solution for the emergence of resistance to However, the process of PG-1 membrane translocation remains
antibiotics.1,2 There are over 2000 AMPs discovered in nature enigmatic from a molecular point of view because PG-1 has a
and documented in publicly available databases.3−5 Protegrin-1 +7 net charge at physiological pH and is thus not expected to
(PG-1 RGGRLCYCRRRFCVCVGR-NH2) is a remarkably interact favorably with the hydrophobic core of a lipid bilayer.
potent, naturally occurring cationic AMP, which has 18 In the experiment, the 13C−31P distances for PG-1 in the
amino acid residues and forms a β-hairpin structure (Figure lipid membrane were found to be very short, which suggested
1). PG-1 was first isolated from porcine leukocytes.6 It has two the formation of bidentate complexes between guanidinium
intramolecular disulfide bonds and is rich in arginine (Arg). and lipid phosphate groups (see Figure 2).17−20 This complex
The atomic structure of PG-1 was determined from NMR is also commonly found in other Arg-rich AMPs and cell-
experiments, in monomeric and dimeric forms and then in penetrating peptides (CPPs).21 Therefore, it was suggested that
octameric pores inside lipid bilayers.7−9 the cationic Arg residues could drag the anionic phosphate
The antimicrobial mechanism of PG-1 is proposed to be groups as they insert into the hydrophobic part of the
based on pore formation. It is believed that PG-1 binds onto membrane, thereby facilitating a toroidal pore formation.17
membranes and inserts inside the hydrophobic core of lipid However, it is still uncertain whether the association energy of a
bilayer membranes, where it oligomerizes. Oligomers of PG-1 bidentate complex suffices to overcome the energy barrier for
form pores, through which there is fast, uncontrollable ion membrane insertion. In addition, toroidal pores have also been
transport. The transmembrane (TM) potential collapses rapidly found in peptide-free systems.22,23 The presence of a bidentate
because of this transport, and the ensuing osmotic lysis results
in quick bacterial death.10−12 Received: March 14, 2018
There is a plethora of molecular dynamics (MD) simulations Accepted: May 8, 2018
conducted to investigate the mechanism of action of PG-1.10−16 Published: June 5, 2018

© 2018 American Chemical Society 6056 DOI: 10.1021/acsomega.8b00483


ACS Omega 2018, 3, 6056−6065
ACS Omega Article

proposed that, provided the peptide is sufficiently charged, the


bound peptides on the membrane suffice to trigger electro-
poration.
Herein, we describe a series of MD simulations performed
using different techniques. The adsorption of PG-1 on the
membrane surface was simulated to confirm the presence of
bidentate complexes. In addition, electroporation simulations
were conducted to study the translocation process. This
protocol has been applied before to study pore formation
and transport of small molecules.27−29 Potential of mean force
(PMF) calculations were finally conducted to examine the
effect of bidentate complexes on the insertion energy barrier.

2. RESULTS AND DISCUSSION


2.1. Binding of PG-1 onto the POPE/POPG Bilayers. To
investigate the interactions in the PG-1_POPE:POPG system,
we performed three independent MD simulations (Sim 1, Sim
2, and Sim 3) by placing PG-1 2.0 nm above the bilayer surface
initially. The simulations were carried out for 50 ns. Figure 3

Figure 1. Structure of PG-1. The arrows represent a β-hairpin Figure 3. Distance along the membrane normal (z-axis) between the
structure pointing from the N-terminal to the C-terminal. Arg side PG-1 center of mass and the average location of phosphorus atoms in
chains are in magenta. Disulfide bonds are in yellow. the upper leaflet. Distance calculations are performed using all three
sets of simulations for the PG-1 and POPE/POPG system.

illustrates the time trajectory of the center-of-mass distance


between PG-1 and the average position of phosphorus atoms
on the top leaflet. It can be observed that PG-1 rapidly binds
onto the membrane surface in all simulations.
Representative snapshots for Sim 2 are provided in Figure 4.
Figure 2. Schematic of the structure formed between the guanidinium During the simulations, PG-1 freely diffuses in the water
and phosphate groups through combined electrostatic forces and subphase. We observed that the binding process has several
bidentate hydrogen bonding. stages. In the beginning (0−5 ns), PG-1 rotates its axis and
moves toward the lipids rapidly. It interacts with the bilayer
complex might not then be necessary or sufficient for pore with either the two-terminal side or the β-turn side. Both sides
formation. Nevertheless, the experimental distance measure- include a cluster of Arg residues, which indicates that the
ments showed only a result in the octameric pore state. To
better understand the membrane insertion mechanism,
intermediate structures and interactions must be investigated.21
There are a host of studies focusing on the translocation of
peptides across the membrane. Fuertes et al.24 suggested that
membrane-active peptides can serve as inducers and stabilizers
of pores, the formation of which may also be induced by
mechanical or electrical tensions. Sun et al.25 studied the Arg-
rich CPPs of octa-arginine peptides. They applied external Figure 4. Characteristic snapshots of one PG-1 with POPE/POPG
tension on the lipids to induce a pore for peptide translocation. simulation. The Arg residues are represented in sticks. The
They concluded that octa-arginine peptides can slow down the phosphorus atoms of POPE and POPG lipids are colored in gray
kinetic rate for pore closure of naturally occurring pores. Miteva and in purple, respectively. Phosphorus atoms within 0.3 nm are
et al.26 studied NK-lysin associated to a membrane and highlighted with orange color and their lipid tails are shown in sticks.

6057 DOI: 10.1021/acsomega.8b00483


ACS Omega 2018, 3, 6056−6065
ACS Omega Article

positively charged residues are essential for the initial contact because of the hydrophobic environment of the lipids. This
through electrostatic interactions. In the next stage (5−20 ns), process is well-described by the solubility−diffusion mecha-
the remaining part of the molecule fluctuates in water until nism.30 On the contrary, for hydrophilic molecules and ions, it
additional Arg residues interact with the lipids. In the last stage appears that the formation of a water column or a pore (we use
(20−50 ns), PG-1 has barely any conformational change and these two terms for contiguous water structures that span the
inserts fully into the lipid head groups. lipid bilayer) can facilitate lipid flip-flop and subsequent
In line with the aforementioned results, the lipid contact ratio transport of polar molecules.31 However, in conventional
presented in Figure 5 provides clear evidence that Arg residues equilibrium MD simulations, it is challenging to observe pore
formation because of the very long simulation time needed.
There have been some studies where pore formation was
observed, but these studies utilized relatively old force fields
found to underestimate the free energy barrier for lipid flip-flop
across the membrane.32−34 More recent atomistic simulations
using CHARMM36 force fields estimate a much larger energy
cost for lipid flip-flop. For example, in recent μs-long
simulations, AMPs did not disrupt the lipid bilayer, and no
pores were observed.35
It has been suggested that electroporation (i.e., the
establishment of a potential across the membrane along with
subsequent phenomena) is a reasonable method to induce pore
formation. Moreover, there have been reports on the transport
of molecules under TM potential using MD simulations. In this
Figure 5. Lipid contact ratio of each PG-1 residue in the simulation. study, we apply electroporation to investigate the PG-1
This ratio is defined as the number of nonhydrogen PG-1 atoms translocation and attempt to link simulations to experimental
within 0.3 nm of POPE or POPG lipids divided by the total number of results. We have carried out a set of simulations listed in Table
nonhydrogen atoms for each residue. Results are averaged over the last 1 to investigate the effect of charge imbalanced methods, TM
20 ns of all simulations.
Table 1. List of Translocation Simulations Performed in
on both sides of the β-hairpin have the largest number of atoms This Worka
in contact with the lipids. The average lipid contact is defined as
the number of nonhydrogen PG-1 atoms within 0.3 nm of charged peptides
case residues (P/L ratio) ΔQ(e) methods
lipids divided by the total number of nonhydrogen atoms. It is
noticed that the N-terminal Arg residue has the highest contact. Arg6_10e_I Arg 6 (6:160) 10 ions
This might be due to the longer length of the N-terminal side Arg6_7e_I Arg 6 (6:160) 7 ions
compared to the C-terminal side (see Figure 1), which allows Arg6_7e_P Arg 6 (6:160) 7 peptides
the N-terminal Arg residue to insert more deeply into the Arg6_14e_I Arg 6 (6:160) 14 ions
bilayer. Arg6_14e_P Arg 6 (6:160) 14 peptides
To further examine peptide−lipid interactions, Figure 6 Arg0_10e_I Arg 0 (0:160) 10 ions
presents the H-bonds formed between Arg residues and lipids. Arg1_10e_I Arg 1 (1:160) 10 ions
Lys6_10e_I Lys 6 (6:160) 10 ions
a
Arg represents the wild-type PG-1, while Lys indicates the charge-
equivalent mutant. There are three PG-1 dimers in high P/L ratio
simulation for a total of six peptides. The three dimers were placed
parallel to the membrane surface initially. The charge imbalance can be
generated by different numbers of ions (I) or peptides (P) in the two
double bilayer systems. The charge imbalance for the peptide method
(P) is generated by keeping three dimers in one system and
subsequently removing dimers in another systems.

potential, P/L ratio, and bidentate bonds. Each system was


simulated three separate times, starting from the same positions
but using different initial velocities in the beginning of MD.
2.2.1. PG-1 Translocation under TM Potential. The
Arg6_10e_I case displayed PG-1 translocation (Figure 7A)
Figure 6. Number of hydrogen bonds formed between the Arg under ΔQ = 10e charge imbalance. At 1.0 ns, a water pore was
residues and lipids. Results are averaged for all three simulations. formed with some lipids reorienting likely to stabilize the pore.
The peptide started to translocate and at 2.7 ns, the Arg residue
at position 4 (Arg4) reached the bilayer center, followed by the
For the three simulations, the mean value of H-bonds in the last translocation of the N-terminal Arg residue (Arg1). This
20 ns is approximately 12. On average, each Arg residue forms inserted state was also stable until the end of the 10 ns
two H-bonds through the bidentate hydrogen bonding (see simulation. In this case, peptide translocation within the
Figure 2). membrane was only partial. Furthermore, we observed that a
2.2. MD Simulations with TM Potential. Membrane strong bidentate hydrogen bond formed between the inserted
transport of small nonpolar molecules is relatively favorable Arg residue and lipid phosphate groups during the simulation.
6058 DOI: 10.1021/acsomega.8b00483
ACS Omega 2018, 3, 6056−6065
ACS Omega Article

Figure 7. Representative snapshots of the electroporation simulations for different cases. (A) Arg6_10e_I. (B) Arg6_14e_I. (C) Arg6_14e_P. (D)
Arg0_10e_I. (E) Arg1_10e_I. (F) Lys6_10e_I. The description of each case is in Table 1.

This indicates that the bidentate complex might be crucial for cases, the system changes rapidly from no pore formation to
Arg insertion, which is consistent with experimental findings. one translocated dimer, as shown in Figure 7C. To have a finer
The TM potential for the Arg6_10e_I allows only one or control of the ΔQ, we focused on the charge imbalance method
two Arg residues to translocate at a time. In the following from ions only.
analysis, the Arg6_10e_I case is regarded as a baseline for all 2.2.3. Effect of TM Potential. Figure 7A,B illustrates the
other tests. results for Arg6_10e_I and Arg6_14e_I systems, respectively.
2.2.2. Comparison of Charge Imbalance Methods. For the For the Arg6_14e_I case, a large water pore was formed in the
Arg6_7e_I and Arg6_7e_P cases, there was no pore formation bilayer within several ps, and a few lipid flip-flops occurred. At
during 10 ns simulations, indicating that the TM potential 0.5 ns, one Arg residue side chain started to protrude into the
generated by ΔQ = 7e was too small. Thereby, no other results water pore, and it reached the bilayer center by 1.0 ns. In
are reported here. Generally, if the TM potential is above a addition, one PG-1 dimer initially parallel to the membrane
threshold value, the pore formation can be observed in a few near the pore opening rotated to align along the membrane
hundred picoseconds. normal. Shortly after the insertion of the Arg residue, the whole
Figure 7B,C illustrates PG-1 translocation with two different dimer translocated into the bilayer and spanned the entire pore.
charge imbalance methods, Arg6_14e_I and Arg6_14e_P, The inserted state remained stable until the end of the 10 ns
respectively. Results were similar with both methods. In both simulation. In contrast, the Arg6_10e_I case only shows partial
cases, we observe a dimer translocating into the water pore Arg insertions for PG-1, and the Arg6_7e_I case shows no pore
from the membrane surface under ΔQ = 14e charge imbalance. formation and peptide translocation at all. Clearly, we can
For the Arg6_14e_I case, the TM potential is generated by ions observe a TM voltage dependence on the pore formation and
in the water subphase. In contrast, for the Arg6_14e_P case, PG-1 translocation.
the TM potential is generated by the charges on the PG-1 itself, 2.2.4. Effect of the P/L Ratio. Previous studies have shown
suggesting molecular electroporation.26 In other words, the that AMPs adopt two distinct orientations depending on the
highly charged PG-1 can generate sufficient TM potential to peptide/lipid ratio. In one, the orientation is parallel (at low
induce pore formation and facilitate subsequent translocation as ratio) and in another perpendicular (at high ratio) to the bilayer
if an external electric field is applied. This also provides surface.36 The experimentally determined threshold P/L value
evidence that PG-1 may serve as a pore inducer.24 for PG-1 is 1:30.37 The underlying molecular reasons for this
Removing one peptide dimer from a bath results in a change are unclear. In this work, we performed simulations to examine
of ΔQ = 7e per bilayer. From the Arg6_7e_P and Arg6_14e_P the concentration dependence of PG-1 translocation at low P/
6059 DOI: 10.1021/acsomega.8b00483
ACS Omega 2018, 3, 6056−6065
ACS Omega Article

Figure 8. Change of the charge difference ΔQ across each bilayer in the 10 ns simulations. Charge transfers due to PG-1 translocation are
highlighted. The N-terminal Arg is denoted as Arg1, which has +2 charge. The Arg residue at position 4 is denoted as Arg4. All others are due to
POPG flip-flop. Results are shown for the Arg6_10e_I, Arg1_10e_I and Lys6_10e_I systems. The points are added every 20 ps.

L (1:160) and at high P/L ratios (6:160). The high P/L ratio is It has been shown in the experiment that Arg and Lys
above the threshold value indicated from the experiment. contribute distinctively to cell uptake for CPPs and
Figure 7D shows snapshots for the Arg0_10e_I case (pure antimicrobial potency for AMPs. For example, Mitchell et al.
membrane). With a TM potential, the pure membrane system has shown that a synthetic CPP, polyarginine (Arg)n, enters the
can also form water pores or columns. We noticed that the pore cell more efficiently than polylysine (Lys)n.39 In addition,
size shrinks during the time interval between 3.0 and 10.0 ns. Schmidt et al. reported that cryptdin-4 (Crp4), a potent
The stability of the pore will be discussed later. bactericidal peptide, showed severely reduced antimicrobial
Figure 7E shows snapshots for the Arg1_10e_I case. At 1.0 activity for complete Lys-for-Arg substitution.40 Consistent
ns, small water pores started to form because of the TM with these results, our MD simulations provide molecular
potential. At 3.0 ns, lipid head groups redistributed toward the insight, demonstrating that the difference might be attributed to
membrane interior. By 10.0 ns, a toroidal-shaped pore the impact of bidentate bonds on the propensity of peptide
developed. However, during these simulations, PG-1 remained insertion in a lipid bilayer.
bound in parallel to the membrane surface. This is in contrast 2.2.6. Discharging of TM Potential. As described in the
to the case at the high P/L ratio where PG-1 molecules adopted Methods section, one drawback of electroporation by the
a perpendicular orientation to the bilayer in an inserted state charge imbalance method is that the imbalance is not reset
(Figure 7A,B). during the simulation. The TM potential drops rapidly after the
PG-1 stayed on the membrane surface for a low P/L ratio, formation of a water pore and subsequent ion transfer. To focus
likely because of the fact that the pore was formed relatively far on the interactions between PG-1 and lipids without a
from the peptide. This might be attributed to the stiffness of the collapsing potential, we have added weak constraints to prevent
bilayer in different domains. As previously shown, PG-1 forms ions from passing through the water pore. However, lipid flip-
extensive hydrogen bonds with lipids, which could increase the flop of anionic palmitoyloleoylphosphatidylglycerol (POPG) or
stability of this domain. It may be less likely for lipids in the PG-1 translocation (i.e., Arg transfer) can also result in the TM
vicinity of PG-1 to form a water defect, which may later grow potential being discharged. Because the focus of the study is the
into pores. Similar results have been reported by Reigada38 who translocation of PG-1, along with lipid flip-flop, we did not
simulated the electroporation process for heterogeneous restrict these motions with added constraints.
bilayers composed of liquid-ordered and liquid-disordered It is instructive to show the change of ΔQ in different cases,
regions. These authors found that pore formation occurs as illustrated in Figure 8. One charge transfer will alter ΔQ by
preferably in the disordered region. On the contrary, as the P/L 2e. POPG flip-flop occurred within a few ns, and all observed
ratio increases, the domain for the peptide-free lipid region flip-flop events were directed from Bath 1 (with excess negative
shrinks. Under TM potential, pores may form around PG-1, charge) to Bath 2 (with excess positive charge), (see Figure
and the ensuing lipid movement may trigger PG-1 trans- 12.). Conversely, Arg residues always inserted in the reverse
location. direction. The region of each bath was divided by the bilayer
2.2.5. Effect of Substitution of Arg with Lys. Numerous center. Phosphorus and Cζ atoms were used to determine the
investigations have been conducted to examine Arg-rich location of lipids and Arg residues in the baths.
peptides. One particular analysis involved substituting Arg As can be observed in Figure 8, the discharging rate was
residues with charge-equivalent Lys residues. In our work, we much faster for the Arg6_10e_I system. It took less than 4 ns
substituted Arg residues in the Arg6_10e_I case with Lys for the system to be fully discharged. For the Lys6_10e_I
residues in the Lys6_10e_I case. system, it took about 6 ns to discharge. The Arg1_10e_I
Figure 7F presents representative snapshots of the system discharged the TM potential in more than 10 ns.
Lys6_10e_I system. Compared to wild-type PG-1, a striking Evidently, the discharging rate is faster for systems with the
difference was observed, in that the Lys mutant has a much higher peptide concentration.
reduced tendency for insertion. This finding provides It is also worth mentioning that the ion constraints
compelling evidence that the Arg guanidinium group is key implemented in this work could be relaxed and the ion-water
for PG-1 translocation. It lends support to the assumption that swap techniques may be applied to maintain the constant
bidentate bonds in Arg−phosphate complexes play a significant voltage across the membrane. However, care must be taken
role in aiding PG-1 insertion into the water pore. Indeed, NMR because all transfers of ions, POPG lipids, or charged residues
experiments have shown that Lys−phosphate interactions are result in changes of ΔQ. The user could, in principle, reset the
less stable than Arg−phosphate interactions.21 This also charge imbalance for all of these events. In practice, one must
indicates that PG-1 translocation is not simply driven by the always be mindful of the artificial nature of the simulated
force due to TM potential. systems and processes.
6060 DOI: 10.1021/acsomega.8b00483
ACS Omega 2018, 3, 6056−6065
ACS Omega Article

2.3. PMF Calculation. It has been proposed that It is proposed that formation of transient water pores is the
interactions between charged protein residues and lipid rate-limiting step in the process of TM lipid flip-flop.41,42 Once
phosphate groups or water can provide the energy needed to a pore has been formed, the ensuing lipid translocation takes
overcome the insertion energy barrier for charged residues to place on timescales of nanoseconds as observed from our
insert inside the membrane.21 It is therefore interesting to study electroporation simulations. We reason that water pore
the thermodynamic driving forces of PG-1 insertion. formation can not only increase lipid flip-flop rates but also
A total of 3 μs umbrella sampling simulations were facilitate the PG-1 translocation.
conducted to mimic the insertion processes of lipid, PG-1, It should be stressed that the insertion processes used for
and the lipid/PG-1 complex (Figure 9). The PMF profiles for PMF calculations are not directly comparable to those of the
these processes were obtained and compared (Figure 10). electroporation processes. In the latter, a water pore forms first,
and lipids reorient themselves before PG-1 insertion takes
place. In the former, a water defect is formed as the forced
insertion process takes place.
It should also be stressed that the presented PMF
calculations are of artificially built systems. The definition of
the PMF zero level is entirely operational and only pertains to
the presented cases. Consequently, these PMFs alone, without
Figure 9. Illustration of the pulling groups of (A) lipid, (B) PG-1, and reactive trajectory sampling, can offer only very limited
(C) lipid/PG-1 complex at the bilayer center. The green spheres mechanistic interpretations. In other words, in the absence of
highlight the pulling atoms. For (A), the pulling atom is the simulations that accurately capture the vast phase space during
phosphorus atom. For (B), the pulling atom is the Cζ atom of the Arg. the entire process of peptides moving from the bulk water
For (C), the pulling atom is the phosphorus atom, while the bidentate
subphase onto membranes and then inside the lipid core, any
hydrogen bonds are restrained by harmonic forces shown in yellow
dashes. proposal of molecular mechanisms underlying biological
phenomena must be considered tentative. Unfortunately,
despite amazing improvements in both simulation software
and hardware, such simulations are currently implausible.
Nonetheless, MD simulations, for all their limitations, are often
the only scientifically agreeable approach to capturing the vast
complexity of biomolecular interactions in a way fit for forming
sound hypotheses.
2.4. Stability of Water Pores with PG-1 Inserted. It is
reported from the experiment that membrane pores induced by
membrane-active peptides have a much longer lifetime than
transient pores induced by thermal energy. Computer
simulations also showed that many membrane-active peptides
can stabilize membrane pores. This motivates us to study the
role of PG-1 in pore stability.
Figure 11A shows the dynamics of water pores for the pure
Figure 10. PMFs for moving the pulling atom into the center of a membrane system (Arg0_10e_I) after we release the
POPE/POPG bilayer for the lipid, peptide, and bidentate complex.
The PMFs were obtained from umbrella sampling calculations, while
the pulling atom was restrained at a varying distance from the bilayer
center of mass, in the direction normal to the plane of the bilayer. The
error bars were determined by bootstrap analysis.

For the pure palmitoyloleoylphosphatidylethanolamine


(POPE)/POPG bilayer, we obtained a relative free energy
barrier for a single POPE flip-flop of 81.1 kJ/mol. The relative Figure 11. Representative snapshots of the MD simulations for (A)
free energy barrier for a PG-1 Arg residue insertion was 94.9 pure membrane pores Arg0_10e_I and (B) membrane pores with PG-
kJ/mol. Figure 9A,B illustrates the lipid head and the Arg 1 inserted from the Arg6_10e_I system.
residue at bilayer center, respectively. We notice that there was
no spontaneous pore formation. Only a water-filled membrane
defect (water column from one side) occurred.
The relative free energy barrier for the lipid/PG-1 complex constraints on ions. It is observed that the transient water
was 124.8 kJ/mol. This free energy is still higher than that for a pore was resealed within 30 ns. In contrast, if there is a PG-1
single lipid flip-flop. The actual timescale required for lipid flip- inserted in a water pore (Arg6_10e_I), the pore remained open
flops ranges between minutes and hours.23 With a higher free for the subsequent 200 ns, as shown in Figure 11B. This is
energy barrier, it is expected that the timescale for the bidentate similar to the kinetic stabilization of octa-arginine peptides
complex translocation may take even longer. Consequently, the reported by Sun et al.25 Therefore, we reason that PG-1 is able
argument that the bidentate complex facilitates pore formation to stabilize and slow down the kinetic rate of water pore
by dragging phosphate groups as Arg residue inserts into the closure. This also provides evidence that PG-1 can serve as a
hydrophobic part of the membrane appears less convincing.17 pore stabilizer after insertion.24
6061 DOI: 10.1021/acsomega.8b00483
ACS Omega 2018, 3, 6056−6065
ACS Omega Article

2.5. Proposed Mechanism for PG-1 Translocation. We 4. METHODS AND THEORY


have applied a computational electroporation method to study 4.1. PG-1 Simulation System Preparation. The solvated
PG-1 translocation. The highly charged PG-1 generates peptide−membrane systems for various simulations were
molecular electroporation as if an external electric field is constructed using CHARMM-GUI.44−47 The initial structures
applied. Under conditions of a sufficient TM potential, a water of the PG-1 monomer and dimer were taken from the protein
pore or column will form across the membrane. data bank (PDB IDs: 1PG1 and 1ZY6, respectively.7,8) The
There have been extensive studies on the collective motions bilayer was modeled using the 3:1 mixture of POPE and POPG
of pore formation and lipid flip-flop.41,42 It has been proposed lipids, mimicking the components of the inner membrane of
that formation of pores unavoidably results in lipid flip-flop. Gram-negative bacteria.48 There were 120 POPE molecules
This is called the pore-mediated lipid flip-flop mechanism. In and 40 POPG molecules in the system. POPE is neutral, while
this study, we found that with the strong interactions of the anionic POPG has a net (−1) charge. TIP3P waters49 were
bidentate bonds in the guanidinium−phosphate complex, it is added on each side of the membrane with a thickness of
likely that the Arg residues might be dragged by the lipids approximately 3 nm. The systems were neutralized by adding
during the diffusive TM transport through the water pores. potassium and chloride as counter ions. There were 40
This in turn may lead to PG-1 insertion and translocation. The potassium atoms, and the number of chloride atoms depended
charge-equivalent lysine mutant showed reduced permeability. on the number of simulated peptides.
This shows that PG-1 insertion is not simply driven by strong To investigate the interactions of PG-1 with the POPE and
TM potential, and the bidentate complex may play an POPG lipid bilayers, three sets of 50 ns simulations were
important role in translocation. performed. The PG-1 monomer was initially placed in the
Although we applied electroporation by imposing a charge water subphase, approximately 2.0 nm away from the
imbalance to induce pore formation, it should be emphasized phosphorus atoms of the upper membrane leaflet. This system
that transient pores may form under different circumstances. is denoted as PG-1_POPE:POPG.
For example, the mechanical stress created by peptide 4.2. Computational Electroporation Setup. The electro-
aggregation is also considered to produce pores to relieve poration simulations were carried out to mimic a TM potential
membrane line tension.25,43 and to facilitate PG-1 translocation. The system was set up with
In addition, we also showed that PG-1 insertion can stabilize a double bilayer configuration by duplicating a second bilayer
water pores and slow down the kinetic rate of pore closure. We on top of the original bilayer (Figure 12). PG-1 molecules were
therefore propose that PG-1 may be capable of inducing pore initially placed at the equilibrium position on the membrane
formation by electrical or mechanical tension. The pore- (approximately 2.2 nm from the bilayer center) as determined
mediated, lipid flip-flop driven PG-1 insertion process facilitates by PG-1 membrane simulations described in section 2.1. A TM
potential was generated by two different methods. The first one
PG-1 translocation. The inserted PG-1 can further increase the
is by swapping ions and waters in Bath 1 and in Bath 2, creating
stability of the water pores. The whole process from PG-1
a charge imbalance.50 The second one is to keep PG-1 in Bath 1
inducing to stabilizing water pores may form a feedback loop,
and partially remove PG-1 in Bath 2 to create a charge
thereby highlighting a great example of collective motions in imbalance. The number of chloride is adjusted to maintain a
biological membranes. neutral system. The number of ions remain the same for both
baths. The former is considered as applying an external electric
3. CONCLUSIONS field, while the latter is regarded as an intrinsic electric field
We have applied various MD protocols to study the generated by the charges on the PG-1 itself. The charge
interactions between PG-1 and lipids that underlie the peptide difference is denoted as ΔQ. The whole system maintained zero
translocation process. Arg residues are crucial for PG-1 to bind net charge.
on the membrane surface. An extensive network of bidentate Nine test cases were considered (listed in Table 1) to
bonds is formed between Arg guanidinium and lipid phosphate investigate the effect of four different factors on the PG-1
groups. Electroporation simulations illustrate that a strong TM translocation: (1) the comparison of the two charge imbalance
methods. (2) The influence of TM potential was studied by
potential results in water pore formation and the disruption of
varying the charge difference ΔQ in the double bilayer system.
the membrane. The charge on the PG-1 is large enough to
(3) The concentration dependence of peptides was examined
create TM potential to induce water pores. Lipid flip-flops and
by changing the peptide to lipid (P/L) ratio. (4) The
PG-1 translocation then take place in a few ns near the water importance of the guanidinium group was investigated by
pore. In addition, a threshold P/L appears to be necessary for substituting Arg with charge-equivalent Lys residues.
the water pores to form in the vicinity of PG-1. Furthermore, One important drawback of the charge imbalance method is
the Arg to Lys substitution significantly reduces the insertion that the imbalance is not reset during the simulation. Once a
propensity, which strongly indicates that the bidentate charged particle transfers to other sides of the membrane, the
hydrogen bonds facilitate PG-1 translocation. Moreover, the system ΔQ changes. A protocol using a swapping method has
inserted PG-1 can stabilize water pores. PMF calculations show been proposed to maintain a constant charge imbalance.51 In
that the barrier for the bidentate complex is still high for this procedure, the number of ions in the two solution baths is
insertion into the hydrophobic environment. We reason that counted frequently and, if the number differs from the initial
PG-1 might be a pore inducer to facilitate water pore formation, setup, a swapping action is carried out: an ion of one solution is
which is a rate-limiting step for lipid flip-flop and for PG-1 exchanged by a water molecule of the other solution bath.
translocation. After pore formation, lipids that flip-flop may However, this method only accounts for the change of charge
drag PG-1 by virtue of the bidentate bond strength. The imbalance due to ion flow. This approach has some limitations
inserted PG-1 serves as a pore stabilizer after translocation. in our system. In addition to ions, there are anionic lipids,
6062 DOI: 10.1021/acsomega.8b00483
ACS Omega 2018, 3, 6056−6065
ACS Omega Article

window size was 0.1 nm. An umbrella potential of a force


constant 3000 kJ mol−1 nm−2 was used for each window to hold
the pulling groups together. A production run of 50 ns was
performed for each window. The PMFs were evaluated using
the g_wham tool in GROMACS.52
4.4. MD Simulation Details. All simulations in this work
were conducted using the GROMACS 5.1 package.53 All-atom
simulations were run at a temperature of 310.15 K using a
Nosé−Hoover thermostat54,55 with a coupling time constant of
1.0 ps. The semi-isotropic pressure coupling method was used
with both lateral and perpendicular pressures (both 1 atm)
independently coupled to the Parrinello−Rahman barostat56
with a coupling time constant of 5 ps and a compressibility of
4.5 × 10−5. Periodic boundary conditions were employed. The
simulation time step was 2 fs. The CHARMM36 force field57
was used for proteins and lipids, and the TIP3P model49 was
used for waters. The cutoffs for the van der Waals interactions
and electrostatic interactions calculated using the particle-mesh
Ewald method58 were both 1.2 nm. The van der Waals
interactions were smoothly turned off between 1.0 and 1.2 nm
with the force-switching method. The LINCS algorithm59 was
used to constrain all of the hydrogen bonds.

■ AUTHOR INFORMATION
Corresponding Authors
*E-mail: laixx266@umn.edu (P.-K.L.).
*E-mail: yiannis@umn.edu. Phone: 651-503-2696. Fax: 612-
626-7246 (Y.N.K.).
Figure 12. Schematic representation of the double bilayer system with ORCID
a charge imbalance ΔQ across each bilayer. The classical 3D periodic Pin-Kuang Lai: 0000-0003-2894-3900
boundary condition is implemented. POPG lipids are shown in purple. Yiannis N. Kaznessis: 0000-0002-5088-1104
POPE lipids are shown in gray. Potassium ions are shown in orange.
Chloride ions are shown in green. Peptides on the membrane surface Present Address

are shown in spherical beads. Waters are represented with a General Probiotics Inc., 1000 Westgate Dr, Ste 122, St. Paul,
semitransparent surface. MN 55114 (Y.N.K.).
Notes
The authors declare no competing financial interest.


POPG and cationic PG-1 molecules, in the system. Any POPG
lipid flip-flop or peptide transport through the membrane alters ACKNOWLEDGMENTS
the imbalance as well. It is not clear how to best deal with some
of these events. In this study, to avoid such situations, we added This work was supported by grants from the National Institutes
a weak one-dimensional constraint (force constant 100 kJ of Health (grant no. GM111358) and by a grant from the
mol−1 nm−2) on the ion positions along the membrane normal National Science Foundation (grant no. CBET-1412283). We
(z-direction). This constraint still allows ions to diffuse in their acknowledge computational support from the Minnesota
original baths but prevents them from passing across the Supercomputing Institute (MSI) and from the Extreme Science
membrane. Although the system could still discharge because of and Engineering Discovery Environment (XSEDE), which is
translocation of lipids or peptides, time scales for such events supported by National Science Foundation grant no. ACI-
are slow enough for the dynamic events of interest to take place 10535753. Support from the University of Minnesota Digital
well before the system is fully discharged, as shown in the Technology Center and the University of Minnesota Institute
for Engineering in Medicine is also acknowledged.


Results and Discussion section.
4.3. PMF Calculation. To calculate the free energy of PG-1
translocation and to shed light on the importance of the REFERENCES
bidentate complex, we performed umbrella sampling simu- (1) Zasloff, M. Antimicrobial peptides of multicellular organisms.
lations for three systems. In system A, the lipid flip-flop free Nature 2002, 415, 389−395.
energy was obtained by pulling a phosphorus atom of a single (2) Brogden, K. A. Antimicrobial peptides: pore formers or metabolic
lipid into the bilayer center. In system B, the energy cost of PG- inhibitors in bacteria? Nat. Rev. Microbiol. 2005, 3, 238−250.
1 insertion was calculated by pulling a Cζ atom on an Arg (3) Wang, Z.; Wang, G. APD: the antimicrobial peptide database.
Nucleic Acids Res. 2004, 32, D590−D592.
residue into the bilayer center. In system C, the free energy of (4) Wang, G.; Li, X.; Wang, Z. APD2: the updated antimicrobial
the combined lipid and PG-1 system was measured by pulling peptide database and its application in peptide design. Nucleic Acids
into the bilayer center a phosphorus atom on the lipid that has Res. 2008, 37, D933−D937.
harmonic constraints on the bidentate hydrogen bonds. (5) Wang, G.; Li, X.; Wang, Z. APD3: the antimicrobial peptide
The reaction coordinate of the PMF was defined as the database as a tool for research and education. Nucleic Acids Res. 2016,
distance to the bilayer center along the bilayer normal. The 44, D1087−D1093.

6063 DOI: 10.1021/acsomega.8b00483


ACS Omega 2018, 3, 6056−6065
ACS Omega Article

(6) Steinberg, D. A.; Hurst, M. A.; Fujii, C. A.; Kung, A. H.; Ho, J. F.; (25) Sun, D.; Forsman, J.; Woodward, C. E. Atomistic molecular
Cheng, F. C.; Loury, D. J.; Fiddes, J. C. Protegrin-1: a broad-spectrum, simulations suggest a kinetic model for membrane translocation by
rapidly microbicidal peptide with in vivo activity. Antimicrob. Agents arginine-rich peptides. J. Phys. Chem. B 2015, 119, 14413−14420.
Chemother. 1997, 41, 1738−1742. (26) Miteva, M.; Andersson, M.; Karshikoff, A.; Otting, G. Molecular
(7) Fahrner, R. L.; Dieckmann, T.; Harwig, S. S. L.; Lehrer, R. I.; electroporation: a unifying concept for the description of membrane
Eisenberg, D.; Feigon, J. Solution structure of protegrin-1, a broad- pore formation by antibacterial peptides, exemplified with NK-lysin.
spectrum antimicrobial peptide from porcine leukocytes. Chem. Biol. FEBS Lett. 1999, 462, 155−158.
1996, 3, 543−550. (27) Breton, M.; Delemotte, L.; Silve, A.; Mir, L. M.; Tarek, M.
(8) Mani, R.; Tang, M.; Wu, X.; Buffy, J. J.; Waring, A. J.; Sherman, Transport of siRNA through lipid membranes driven by nanosecond
M. A.; Hong, M. Membrane-bound dimer structure of a β-hairpin electric pulses: an experimental and computational study. J. Am. Chem.
antimicrobial peptide from rotational-echo double-resonance solid- Soc. 2012, 134, 13938−13941.
state NMR. Biochemistry 2006, 45, 8341−8349. (28) Salomone, F.; Breton, M.; Leray, I.; Cardarelli, F.; Boccardi, C.;
(9) Mani, R.; Cady, S. D.; Tang, M.; Waring, A. J.; Lehrer, R. I.; Bonhenry, D.; Tarek, M.; Mir, L. M.; Beltram, F. High-yield nontoxic
Hong, M. Membrane-dependent oligomeric structure and pore gene transfer through conjugation of the CM18-Tat11 chimeric
peptide with nanosecond electric pulses. Mol. Pharmaceutics 2014, 11,
formation of a beta-hairpin antimicrobial peptide in lipid bilayers
2466−2474.
from solid-state NMR. Proc. Natl. Acad. Sci. U.S.A. 2006, 103, 16242−
(29) Casciola, M.; Tarek, M. A molecular insight into the electro-
16247.
transfer of small molecules through electropores driven by electric
(10) Langham, A. A.; Ahmad, A. S.; Kaznessis, Y. N. On the nature of
fields. Biochim. Biophys. Acta, Biomembr. 2016, 1858, 2278−2289.
antimicrobial activity: a model for protegrin-1 pores. J. Am. Chem. Soc. (30) Missner, A.; Pohl, P. 110 years of the Meyer−Overton rule:
2008, 130, 4338−4346. predicting membrane permeability of gases and other small
(11) Bolintineanu, D.; Hazrati, E.; Davis, H. T.; Lehrer, R. I.; compounds. ChemPhysChem 2009, 10, 1405−1414.
Kaznessis, Y. N. Antimicrobial mechanism of pore-forming protegrin (31) Gurtovenko, A. A.; Anwar, J.; Vattulainen, I. Defect-mediated
peptides: 100 pores to kill E. coli. Peptides 2010, 31, 1−8. trafficking across cell membranes: insights from in silico modeling.
(12) Bolintineanu, D. S.; Kaznessis, Y. N. Computational studies of Chem. Rev. 2010, 110, 6077−6103.
protegrin antimicrobial peptides: a review. Peptides 2011, 32, 188−201. (32) Leontiadou, H.; Mark, A. E.; Marrink, S. J. Antimicrobial
(13) Lipkin, R.; Lazaridis, T. Computational prediction of the peptides in action. J. Am. Chem. Soc. 2006, 128, 12156−12161.
optimal oligomeric state for membrane-inserted β-barrels of protegrin- (33) Herce, H. D.; Garcia, A. E. Molecular dynamics simulations
1 and related mutants. J. Pept. Sci. 2017, 23, 334−345. suggest a mechanism for translocation of the HIV-1 TAT peptide
(14) Lipkin, R.; Lazaridis, T. Computational studies of peptide- across lipid membranes. Proc. Natl. Acad. Sci. U.S.A. 2007, 104,
induced membrane pore formation. Philos. Trans. R. Soc., B 2017, 372, 20805−20810.
20160219. (34) Sengupta, D.; Leontiadou, H.; Mark, A. E.; Marrink, S.-J.
(15) Lipkin, R.; Pino-Angeles, A.; Lazaridis, T. Transmembrane Pore Toroidal pores formed by antimicrobial peptides show significant
Structures of β-Hairpin Antimicrobial Peptides by All-Atom disorder. Biochim. Biophys. Acta, Biomembr. 2008, 1778, 2308−2317.
Simulations. J. Phys. Chem. B 2017, 121, 9126−9140. (35) Bennett, W. F. D.; Hong, C. K.; Wang, Y.; Tieleman, D. P.
(16) Lipkin, R. B.; Lazaridis, T. Implicit membrane investigation of Antimicrobial peptide simulations and the influence of force field on
the stability of antimicrobial peptide β-barrels and arcs. J. Membr. Biol. the free energy for pore formation in lipid bilayers. J. Chem. Theory
2015, 248, 469−486. Comput. 2016, 12, 4524−4533.
(17) Tang, M.; Waring, A. J.; Hong, M. Phosphate-mediated arginine (36) Huang, H. W. Action of antimicrobial peptides: two-state
insertion into lipid membranes and pore formation by a cationic model. Biochemistry 2000, 39, 8347−8352.
membrane peptide from solid-state NMR. J. Am. Chem. Soc. 2007, 129, (37) Yamaguchi, S.; Hong, T.; Waring, A.; Lehrer, R. I.; Hong, M.
11438−11446. Solid-State NMR Investigations of Peptide- Lipid Interaction and
(18) Rothbard, J. B.; Jessop, T. C.; Lewis, R. S.; Murray, B. A.; Orientation of a β-Sheet Antimicrobial Peptide, Protegrin. Biochemistry
Wender, P. A. Role of membrane potential and hydrogen bonding in 2002, 41, 9852−9862.
the mechanism of translocation of guanidinium-rich peptides into cells. (38) Reigada, R. Electroporation of heterogeneous lipid membranes.
J. Am. Chem. Soc. 2004, 126, 9506−9507. Biochim. Biophys. Acta, Biomembr. 2014, 1838, 814−821.
(19) Rothbard, J.; Jessop, T.; Wender, P. Adaptive translocation: the (39) Mitchell, D. J.; Steinman, L.; Kim, D. T.; Fathman, C. G.;
role of hydrogen bonding and membrane potential in the uptake of Rothbard, J. B. Polyarginine enters cells more efficiently than other
guanidinium-rich transporters into cells. Adv. Drug Delivery Rev. 2005, polycationic homopolymers. J. Pept. Res. 2000, 56, 318−325.
(40) Schmidt, N. W.; Tai, K. P.; Kamdar, K.; Mishra, A.; Lai, G. H.;
57, 495−504.
Zhao, K.; Ouellette, A. J.; Wong, G. C. L. Arginine in α-defensins:
(20) Pantos, A.; Tsogas, I.; Paleos, C. M. Guanidinium group: a
differential effects on bactericidal activity correspond to geometry of
versatile moiety inducing transport and multicompartmentalization in
membrane curvature generation and peptide-lipid phase behavior. J.
complementary membranes. Biochim. Biophys. Acta, Biomembr. 2008,
Biol. Chem. 2012, 287, 21866−21872.
1778, 811−823. (41) Gurtovenko, A. A.; Vattulainen, I. Molecular mechanism for
(21) Su, Y.; Li, S.; Hong, M. Cationic membrane peptides: atomic- lipid flip-flops. J. Phys. Chem. B 2007, 111, 13554−13559.
level insight of structure-activity relationships from solid-state NMR. (42) Tieleman, D. P.; Marrink, S.-J. Lipids out of equilibrium:
Amino Acids 2013, 44, 821−833. Energetics of desorption and pore mediated flip-flop. J. Am. Chem. Soc.
(22) Gurtovenko, A. A.; Vattulainen, I. Ion Leakage through 2006, 128, 12462−12467.
Transient Water Pores in Protein-Free Lipid Membranes Driven by (43) Henderson, J. M.; Waring, A. J.; Separovic, F.; Lee, K. Y. C.
Transmembrane Ionic Charge Imbalance. Biophys. J. 2007, 92, 1878− Antimicrobial peptides share a common interaction driven by
1890. membrane line tension reduction. Biophys. J. 2016, 111, 2176−2189.
(23) Gurtovenko, A.; Vattulainen, I. In Biomembrane Frontiers, (44) Jo, S.; Kim, T.; Iyer, V. G.; Im, W. CHARMM-GUI: a web-
Collective Dynamics in Lipid Membranes: From Pore Formation to Flip- based graphical user interface for CHARMM. J. Comput. Chem. 2008,
Flops; Jue, T., Risbud, S., Longo, M., Faller, R., Eds.; Humana Press: 29, 1859−1865.
New York, 2009; Chapter 5, pp 121−139. (45) Brooks, B. R.; Brooks, C. L.; MacKerell, A. D.; Nilsson, L.;
(24) Fuertes, G.; Giménez, D.; Esteban-Martín, S.; Sánchez-Munoz, Petrella, R. J.; Roux, B.; Won, Y.; Archontis, G.; Bartels, C.; Boresch, S.
O. L.; Salgado, J. A lipocentric view of peptide-induced pores. Eur. CHARMM: the biomolecular simulation program. J. Comput. Chem.
Biophys. J. 2011, 40, 399−415. 2009, 30, 1545−1614.

6064 DOI: 10.1021/acsomega.8b00483


ACS Omega 2018, 3, 6056−6065
ACS Omega Article

(46) Wu, E. L.; Cheng, X.; Jo, S.; Rui, H.; Song, K. C.; Dávila-
Contreras, E. M.; Qi, Y.; Lee, J.; Monje-Galvan, V.; Venable, R. M.
CHARMM-GUI Membrane Builder toward realistic biological
membrane simulations. J. Comput. Chem. 2014, 35, 1997−2004.
(47) Lee, J.; Cheng, X.; Swails, J. M.; Yeom, M. S.; Eastman, P. K.;
Lemkul, J. A.; Wei, S.; Buckner, J.; Jeong, J. C.; Qi, Y. CHARMM-GUI
Input Generator for NAMD, GROMACS, AMBER, OpenMM, and
CHARMM/OpenMM Simulations Using the CHARMM36 Additive
Force Field. J. Chem. Theory Comput. 2015, 12, 405−413.
(48) Murzyn, K.; Róg, T.; Pasenkiewicz-Gierula, M. Phosphatidyle-
thanolamine phosphatidylglycerol bilayer as a model of the inner
bacterial membrane. Biophys. J. 2005, 88, 1091−1103.
(49) Jorgensen, W. L.; Chandrasekhar, J.; Madura, J. D.; Impey, R.
W.; Klein, M. L. Comparison of simple potential functions for
simulating liquid water. J. Chem. Phys. 1983, 79, 926−935.
(50) Sachs, J. N.; Crozier, P. S.; Woolf, T. B. Atomistic simulations of
biologically realistic transmembrane potential gradients. J. Chem. Phys.
2004, 121, 10847−10851.
(51) Kutzner, C.; Grubmüller, H.; de Groot, B. L.; Zachariae, U.
Computational electrophysiology: the molecular dynamics of ion
channel permeation and selectivity in atomistic detail. Biophys. J. 2011,
101, 809−817.
(52) Hub, J. S.; de Groot, B. L.; van der Spoel, D. g_wham - A Free
Weighted Histogram Analysis Implementation Including Robust Error
and Autocorrelation Estimates. J. Chem. Theory Comput. 2010, 6,
3713−3720.
(53) Abraham, M. J.; Murtola, T.; Schulz, R.; Páll, S.; Smith, J. C.;
Hess, B.; Lindahl, E. GROMACS: High performance molecular
simulations through multi-level parallelism from laptops to super-
computers. SoftwareX 2015, 1, 19−25.
(54) Hoover, W. G.; Ladd, A. J. C.; Moran, B. High-strain-rate plastic
flow studied via nonequilibrium molecular dynamics. Phys. Rev. Lett.
1982, 48, 1818.
(55) Nosé, S. A molecular dynamics method for simulations in the
canonical ensemble. Mol. Phys. 1984, 52, 255−268.
(56) Parrinello, M.; Rahman, A. Polymorphic transitions in single
crystals: A new molecular dynamics method. J. Appl. Phys. 1981, 52,
7182−7190.
(57) Klauda, J. B.; Venable, R. M.; Freites, J. A.; O’Connor, J. W.;
Tobias, D. J.; Mondragon-Ramirez, C.; Vorobyov, I.; MacKerell, A. D.,
Jr.; Pastor, R. W. Update of the CHARMM all-atom additive force
field for lipids: validation on six lipid types. J. Phys. Chem. B 2010, 114,
7830−7843.
(58) Essmann, U.; Perera, L.; Berkowitz, M. L.; Darden, T.; Lee, H.;
Pedersen, L. G. A smooth particle mesh Ewald method. J. Chem. Phys.
1995, 103, 8577−8593.
(59) Hess, B.; Bekker, H.; Berendsen, H. J. C.; Fraaije, J. G. E. M.
LINCS: a linear constraint solver for molecular simulations. J. Comput.
Chem. 1997, 18, 1463−1472.

6065 DOI: 10.1021/acsomega.8b00483


ACS Omega 2018, 3, 6056−6065

Вам также может понравиться