Abstract
The influenza A M2 wild-type (WT) proton channel is the target of the anti-influenza drug rimantadine. Rimantadine has two enantiomers, though most investigations into drug binding and inhibition have used a racemic mixture. Solid-state NMR experiments using the full length-M2 WT have shown significant spectral differences that were interpreted to indicate tighter binding for (R)- vs (S)-rimantadine. However, it was unclear if this correlates with a functional difference in drug binding and inhibition. Using X-ray crystallography, we have determined that both (R)- and (S)-rimantadine bind to the M2 WT pore with slight differences in the hydration of each enantiomer. However, this does not result in a difference in potency or binding kinetics, as shown by similar values for kon, koff, and Kd in electrophysiological assays and for EC50 values in cellular assays. We concluded that the slight differences in hydration for the (R)- and (S)-rimantadine enantiomers are not relevant to drug binding or channel inhibition. To further explore the effect of the hydration of the M2 pore on binding affinity, the water structure was evaluated by grand canonical ensemble molecular dynamics simulations as a function of the chemical potential of the water. Initially, the two layers of ordered water molecules between the bound drug and the channel’s gating His37 residues mask the drug’s chirality. As the chemical potential becomes more unfavorable, the drug translocates down to the lower water layer, and the interaction becomes more sensitive to chirality. These studies suggest the feasibility of displacing the upper water layer and specifically recognizing the lower water layers in novel drugs.
Graphical Abstract
The influenza A matrix 2 (M2) protein is the most extensively studied viroporin and is currently considered the most well-validated viroporin antiviral drug target.1 It is an attractive target for anti-influenza medications because it mutates less frequently than other proteins within the virus.2 However, though the M2 channel mutates infrequently, drug-resistant mutations have become prevalent since the early 2000s, and currently, amantadine and rimantadine, which block M2 wild-type (WT) ptoton function, are no longer recommended for treatment of influenza infections.3,4 The rise of adamantane-resistant influenza necessitates an in-depth understanding of the stereochemical requirements for M2 channel blockers so that new drugs can be developed.
The M2 protein is a homotetrameric proton channel5–8 whose minimally functional transmembrane core consists of residues 22–46. 9,10 The gating residues involved in proton conductance are His37 and Trp41.11,12 After an influenza virus particle is endocytosed by the host cell, proton transport through the M2 channel lowers the pH within the viral envelope and allows viral ribonucleoproteins (RNPs) to unpack from M1.13 Blocking proton transport through the M2 WT channel using adamantanes14,15 prevents this dissociation from happening and thus prevents viral replication from occurring.16–18 The proton transport function of M2 plays a second role in the viral life cycle; M2 inhibits the acidification of the trans Golgi compartment, thereby preventing premature activation of the acid-sensitive hemagglutinin protein.19,20
Structural biology studies of M2 WT-drug complexes have been performed using racemic rimantadine,21–23 but in a solid-state NMR (ssNMR) spectra using deuterium-labeled (R)- and (S)-rimantadine, the authors showed differences in chemical shifts when the two enantiomers bind to full-length M2 WT.24 These spectral differences were considered to arise from differences in potency and affinity of the isomers. However, electrophysiological (EP), isothermal titration calorimetry (ITC), and antiviral assays carried out using enantiomerically pure rimantadine indicated that the two enantiomers have equal potency against M2 WT.16,25 Additionally, in vivo experiments in mice have shown similar antiviral activity for both enantiomers.26 One possible explanation for the difference in ssNMR chemical shifts for the two rimantadine enantiomers might relate to differences in the kinetics of binding, which are particularly important for the potency of M2 inhibitors. In the earlier EP study examining rimantadine enantiomer binding kinetics to M2 WT, 16 the time traces were conducted under kinetic control, which did allow sufficient time to determine accurate measurement of koff (Kdiss is the ratio of kon to koff). Here we performed a more stringent kinetic study by recording both the binding and washing curves so that we measured more accurately koff. Using this improved methodology, we did not observe any difference in binding kinetics between the (R)- and (S)- enantiomers, as shown by similar values for kon, koff, and Kd in EP assays. We also tested the potency of the two enantiomers against an amantadine-sensitive strain of influenza in an antiviral plaque assay and found that the EC50 values are indistinguishable.
We also reported the crystal structure of M2(22–46) WT in complex with enantiomerically pure (R)- and (S)-rimantadine. Earlier, a crystal structure had been solved for the complex of M2 with racemic rimantadine, but the density could be interpreted equally well for models involving both enantiomers (as would be expected if they bound with equal affinity) or a single high-affinity enantiomer in multiple conformations. We now showed that, although the crystals of each rimantadine enantiomer complex had different space groups and unit cell dimensions, we observed that the electron density corresponding to the bound drug is similar for the two enantiomers. The density indicates a mechanism of binding in which the rimantadine enantiomer ethylamine can rotate to four positions to form hydrogen bonds with ordered waters in the M2 pore. We observed slight differences between the hydration of the two enantiomers within the pore.
Finally, we employed grand canonical ensemble molecular dynamics simulations (GCMC/MD) to provide further insight into the water network present in the rimantadine binding site. GCMC27–29 can be used in combination with MD simulations in order to bypass the kinetic barriers associated with the binding and unbinding of occluded water sites. This is done by simulating the system at a constant chemical potential, while inserting and deleting waters from a region of interest (defined as the rimantadine enantiomer binding site, in this case). Additionally, carrying out these simulations at a range of chemical potential values, referred to as GCMC titration, allowed the binding free energy of the water network to be rigorously calculated. It is of interest that the titration analysis indicates that a chiral difference is apparent when the upper water layer of the binding site is displaced compared to the two layers, which are sufficiently flexible to eliminate a chiral preference. To validate our GCMC/MD simulations, we first calculated the relative binding free energy of the (R)- and (S) enantiomers to the M2 WT pore.
RESULTS AND DISCUSSION
X-ray Crystal Structures.
We have solved two X-ray crystal structures of (R)-rimantadine and (S)-rimantadine bound to M2(22–46) WT through cocrystallization experiments (Table S1). Crystals of (R)-rimantadine bound to M2(22–46) WT (PDB ID 6US9) formed at pH 8.5 in the P21 space group with unit cell dimensions a, b, c (Å) = 48.18, 48.70, 71.67 and α, β, γ (deg) = 90, 90, 90, and the diffraction was limited to 2.00 Å. Crystals of (S)-rimantadine bound to M2(22–46) (PDB ID 6US8) formed at pH 7.5 in the P212121 space group with unit cell dimensions a, b, c (Å) = 49.39, 76.09, 98.63 and α, β, γ (deg) = 90, 90, 90. These crystals diffracted to a resolution of 1.70 Å, making this the highest-resolution structure of M2 bound to a drug or inhibitor.30–32
As in previously solved structures of M2(22–46) WT bound to drugs and inhibitors,31,32 M2(22–46) WT is in the Inwardclosed conformation, and we observed electron density corresponding to the bound drug at the channel’s N-terminus near Ser31, with two layers of ordered water molecules between the bound drug and the channel’s gating His37 residues (Figures 1a–d). These waters form hydrogen bonds to pore-facing carbonyl groups from residues Ala30 and Gly34 and also form a vertical hydrogen bonding network leading down to His37. The X-ray crystal structures remained unchanged during 300 ns MD simulations in hydrated POPC bilayers (Figures 1c,d).
Figure 1.
(a, b) Polder maps are shown here to a contour of 3 σ (dark green mesh) for the X-ray crystal structures of (R)-rimantadine (a) and (S)-rimantadine (b) bound to M2(22–46) WT. Front and back monomer helices have been removed to show the contents of the pore. For both structures, the density corresponding to the rimantadine enantiomer ethylammonium group has been modeled as the average of four rotational conformers. We observed a network of ordered waters between the bound rimantadine enantiomer and the gating His37 residues; the waters form hydrogen bonds with pore-facing carbonyl groups and also form a vertical hydrogen-bond network leading down to the gating His37 residues. (c, d), The radial distribution function (RDF) g(r) for (R)-rimantadine (c) and (S)-rimantadine (d) between the ammonium group and water oxygen atoms from the 300 ns MD simulation shows a strong peak at 2.7 Å, which corresponds to the upper waters layer. The cumulative integrated intensity ∫g(r), shown in the right axis, at 2.7 Å indicates ca. 4 waters which form hydrogen bonds with the ammonium group of rimantadine enantiomer and the carbonyl groups of Ala30. The second broad peak at 4.8 Å corresponds to ca. 5–6 waters of the lower layer, which form hydrogen bonds with the His37 imidazole groups and the carbonyl groups of Gly34.
Though enantiomerically pure (R)- or (S)-rimantadine was used in these crystallization trials, the chirality of the drug is not obvious from the electron density in either structure. This is presumably a result of averaging more than one binding position within the crystal lattice. Thus, we have modeled the bound drug for each of the structures as a superposition of four rotational, by rotation of C1Ad–CEthyl bond, conformers, with the drug ammonium group forming hydrogen bonds to two of the four top layer waters in each position. In our previous crystal structure of racemic rimantadine bound to M2(22–46) WT (PDB ID 6BKL),32 we observed this same ambiguity regarding the position of the rimantadine ethylammonium group, which we interpreted as an effect of cocrystallizing using racemic rimantadine.
The occupancies of the four rotational conformers of (R)- or (S)-rimantadine were allowed to float during refinement; alternate conformers were removed from the model if their occupancy was refined to zero. Each structure contains four tetramers per asymmetric unit; the refined occupancies for the four rimantadine rotational conformers for each tetramer are shown in Table S2. To summarize, for the structure of (R)-rimantadine bound to M2(22–46) WT, three conformers were observed per tetramer with some variations in the occupancy of each (R)-rimantadine conformer depending on the tetramer. In the structure of (S)-rimantadine bound to M2(22–46) WT, three of the tetramers have relatively high (>0.30) occupancy for two of the (S)-rimantadine conformers, with the other two conformers having either zero or <0.05 occupancy. The fourth tetramer has low occupancy (0.01 and 0.11) for two of the conformers and relatively high occupancy (>0.40) for the other two. These data suggest that the binding mechanism remains the same for both enantiomers with the observed slight differences likely due to the hydration of each enantiomer as shown in Figure 1. We subsequently explored the hydration of each enantiomer inside the M2 WT pore using GCMC/MD simulations.
In addition to the density associated with (S)-rimantadine in the pore, we unexpectedly observed electron density that we provisionally assigned to rimantadine in the structure of the complex at 1.7 Å. The additional (S)-rimantadine lies at locations that stabilize the packing of individual tetramers in the crystal lattice (Figures S1 and S2). The two tetramers have an antiparallel orientation, and the rimantadine enantiomer makes contact with residues near the N-terminus of one tetramer and residues near the C-terminus of the other tetramer. The location of the ethylamonnium group is not clear from the electron density, so we have only modeled the adamantyl group. The C-terminal contacts are similar to those proposed for a weak, exterior site observed in solution and ssNMR studies that indicate binding of the adamantanes to the exterior of the channel’s C-terminus at high concentrations of drug (including Leu 40, Ile42, Asp44, Arg45, and Leu46).33,34 However, these previous spectroscopic studies did not identify contacts near the N-terminus (Leu26, Val28, Ala29, Ile32). Thus, it is likely that the exterior site seen here is a crystallographic artifact.
Electrophysiology and Antiviral Plaque Assays.
To study whether chirality has an effect on the functional activity against the M2 WT channel, we determined the association constant (kon), the dissociation constant (koff), and the binding affinity (Kd) for (R)- and (S)-rimantadine in a two-electrode voltage clamp electrophysiological (TEVC) assay using kinetic studies.35–37
Drugs bind slowly to the channel, and dissociate very slowly, so special procedures are required to determine the on and off rates. We first applied relatively high concentrations (100 μM) of the test compound to fully inhibit the channel. The second-order rate constant for inhibition (kon) was determined from the time course of inhibition. The drug is then washed out, and the time required to recover from inhibition was measured to determine koff. Full-length M2 WT from the amantadine-sensitive Udorn strain was expressed in oocytes, and a low pH 5.5 solution was applied to activate the M2 channel. Next, a pH 5.5 solution containing the testing compound was applied to inhibit the M2 channel. Once the current reached the steady state, compound dissociation was initiated by changing the oocyte bathing solution to pH 5.5 without the drug. During the washout, a few pH 8.5 pulses were applied to make sure the current went to baseline to ensure the oocyte quality. The recording traces are shown in Figure 2.
Figure 2.
Rimantadine enantiomers and amantadine binding kinetics against Udorn M2 WT were determined using a combined application and washout procedure TEVC assay. (A) Racemic rimantadine, (B) (R)-rimantadine, (C) (S)-rimantadine, or (D) amantadine was applied to oocytes for 5–7 min after the inward current reached its maximum; then a washout protocol was applied to the oocytes. During the washout, pH 8.5 pulses were applied to make sure the current went to baseline to ensure the oocyte quality. The blue bar above the recording trace indicates the period in which the pH 5.5 Barth solution was applied; the red bar indicates the period in which compounds in pH 5.5 Barth solution was applied. Representative recording traces are shown on the left side of each figure. Data extracted from the recording traces were plotted with an association then dissociation equation in GraphPad Prism 5 as shown on the right side of each figure. The best-fit values are shown in Table 1.
Fitting the binding and washing curves with association and dissociation equations yielded the kon, koff, and Kd values for (R)-, (S)-, and racemic rimantadine.37 As shown in Table 1, the binding kinetics parameter (kon, koff, and Kd) values for (R)-, (S)- and racemic rimantadine are not significantly different, which suggests that both the (R)- and (S)-enantiomers of rimantadine bind to the M2 channel with equal potency. We also included amantadine as a control: the kon of amantadine is comparable with that of rimantadine; however, the calculated koff of amantadine is more than 10-fold faster than that of (R)- and (S)-enantiomers of rimantadine. Therefore, the binding affinity of amantadine to Udorn M2 WT is about 10-fold weaker than rimantadine to M2 WT. Overall, there is no significant difference among the Kd values for (R)- and (S)-rimantadine as well as the racemic rimantadine. We note that a previous study16 reported 100-fold higher koff values, resulting in significantly higher Kd values for (R)- and (S)-enantiomers and the racemic mixture of rimantadine with 3.2, 3.9, and 7 μM, respectively, as opposed to the values of 41, 39, and 46 nM, respectively, reported here. In the previous study,16 only the binding curve was recorded and used for data fitting to derive both kon and koff, which might result in significant variations of koff values (personal communication with Dr. David Busath, the lead researcher of the EP assay in the previous study); the time traces were not recorded for sufficient time to allow accurate measurement of koff, and thus, the previous values are an upper limit of the true Kd value. In contrast, we performed more stringent kinetic studies by recording both the binding and washing curves in the current study. As such, the koff values were accurately quantified from the washing curve, which was then used to fit the binding curve to determine kon. The validity of this methodology in determining the Kd values of M2-S31N and M2-V27A inhibitors has been demonstrated in several studies, which showed a positive correlation between the Kd values and the cellular antiviral activity.35–37
Table 1.
Summary of the Binding Affinity of Amantadine and Rimantadine Enantiomers against Udorn M2 WT
rimantadine (racemic) | (R)-rimantadine | (S)-rimantadine | amantadine | |
---|---|---|---|---|
concentration tested | 50 μM | 50 μM | 50 μM | 100 μM |
kon (min−1 M−1) | 19600 ± 300 | 20800 ± 700 | 22500 ± 300 | 20500 ± 300 |
koff (min−1) | (9.1 ± 0.8) × 10−4 | (9 ± 2) × 10−4 | (8.8 ± 0.8) × 10−4 | (119 ± 2) × 10−4 |
Kd = koff/kon (nM) | 46 ± 4 | 41 ± 9 | 39 ± 4 | 580 ± 20 |
To further confirm that the (R)- and (S)-enantiomers of rimantadine have indistinguishable potency in blocking the M2 channel, we tested these compounds against the amantadine-sensitive A/Soloman Island/3/2006 (H1N1) strain (Figure 3). As expected,27 both (R)- and (S)-rimantadine inhibit viral replication with EC50 values of 19.62 and 24.44 nM, respectively. Taken together, results from both the electro-physiological assay and the antiviral plaque assay showed that both (R)- and (S)-rimantadine have equal potency in blocking the M2 WT channel.
Figure 3.
Cellular antiviral assay results of (R)- and (S)-rimantadine against the amantadine-sensitive A/Soloman Island/3/2006 (H1N1) strain. The antiviral potency was determined in a plaque assay. The EC50 values are the mean ± standard deviation of two independent repeats.
Binding Free Energy Calculations.
To validate our experimental results, the relative binding free energies of binding to M2(22-46) WT and hydration of the (R)- and (S)-rimantadine were calculated using the free energy perturbation method38,39 (see Figure S3) coupled with MD simulations (FEP/MD) and the multistate Bennett acceptance ratio (MBAR) method for processing the free energy data.40,41 The free energy perturbations carried out showed that the hydration-free energy difference between the two rimantadine enantiomers (where a positive value would indicate that the (R)-rimantadine is favored, and vice versa) is −0.06 ± 0.03 kcal mol−1, when starting the calculation from the (R)-rimantadine, and 0.000 ± 0.005 kcal mol−1, when starting from the (S)-rimantadine (the uncertainties given represent the standard errors over the three runs). Given that bulk water is an achiral environment, the true value of this hydration free energy difference is zero, and these values, therefore, appeared correct. This result, therefore, supports the view that the free energy protocol used (see Supporting Information) is suitable for this perturbation.
The relative binding free energy difference between the two rimantadine enantiomers was calculated as +0.29 ± 0.04 kcal mol−1 and +0.33 ± 0.06 kcal mol−1, when starting the calculations from the (R)- or (S)-enantiomer, respectively, and the accuracy of the free energy calculations from molecular simulations is typically considered to be around 1 kcal mol−1.42 These values show also that the two rimantadine enantiomers bind equally to M2TM WT without any chiral preference. These FEP/MD calculation results with the MBAR method are also in agreement with previous measurements of enantiomers binding affinities using ITC and calculation of their relative free energy of hydration43 and relative binding free energy to M2(22-46) WT using FEP/MD simulations with the BAR method and different alchemical intermediates.16,44
GCMC Titration.
In the experimental structures, the number of waters in the upper layer (i.e., close to Ala30) is 4 and in the lower layer (i.e., close to Gly34) is 5 (PDB ID 6US9) or 6 (PDB ID 6US8).32
The GCMC titration calculations involve progressively reducing the chemical potential applied to the water molecules in the ligand-binding site. As the chemical potential is reduced, the more weakly bound water molecules are removed from the binding site, until eventually an entirely dry site results (Figure 4). In addition, the binding free energy of the water network may be calculated from the water occupancy as a function of chemical potential data. The resulting free energy data as a function of water occupancy may be used to determine the equilibrium hydration level, which corresponds to the minimum on the curve of the free energy profile (Figure 5).
Figure 4.
Comparison of the two titration plots for M2 WT in complex with rimantadine enantiomer with the average number of waters observed plotted against the B value (water-binding free energy profiles). The curve was fitted as the mean from 1000 bootstraps of a sum of four sigmoid curves. Results for rimantadine enantiomers are shown as inset images added to show representative structures from different points on the curve. The (R)-rimantadine is shown in cyan, and the (S)-rimantadine is shown in pink. The m = 4 in the legend refers to the fact that a sum of four sigmoid functions was fitted to the data.
Figure 5.
Water binding free energy plots for both rimantadine enantiomers.
For both of the rimantadine enantiomers, the free energy profile (Figure 5) showed a minimum for N = 9, indicating that this is the most favorable number of water molecules in the binding site, in agreement with our crystal structures. Also, in agreement with the experiments is that there was no significant difference in the calculated binding energy for the (R)-enantiomer(−30 ± 1 kcal mol−1), versus the (S)-enantiomer (−29 ± 1 kcal mol−1). Interestingly, the free energy change associated with adding a tenth water to the binding site for the (R)- and (S)-enantiomers are +0.1 ± 0.1 and +0.3 ± 0.2 kcal mol−1, respectively, indicating that the N = 10 state is only slightly less stable at equilibrium.
We next asked whether there might be a difference if the waters were virtually “squeezed out” at a low potential. This can be observed by a selection of simulation snapshots shown in Figure 6 and the violin plots shown in Figure 7, which show the distribution of the number of waters observed (via their z-coordinate, given that this is the normal to the membrane), as well as that of the ligand position within the channel.
Figure 6.
A selection of representative frames from the simulations performed, showing the structural trends as the waters are removed from the GCMC region. The (R)-rimantadine is shown in cyan, and the (S)-rimantadine is shown in pink.
Figure 7.
Violin plots depicting (a) the water distributions observed at each Adams value during the titration calculations; each water point is considered as its z-coordinate (the z-axis is perpendicular to the membrane), relative to the mean z-coordinate of the Cα atoms of the four His37 residues. (b) The ligand positions observed at each Adams value during the titration calculations; each point is considered as the z-coordinate (the z-axis is perpendicular to the membrane) of the nitrogen atom of the ligand, relative to the mean z-coordinate of the Cα atoms of the four His37 residues. The data for the (R)- and (S)-enantiomer are shown in blue and red, respectively. The plots are normalized such that all violins have the same width.
The general trend observed was that, as the chemical potential or B value decreases and waters are removed from the GCMC region, waters from the upper layer, i.e., close to Ala30, are removed first (see Figure 7a), and the ligand gradually descends further into the protein to interact directly with the lower layer, i.e., close to Gly34 (Figure 6b,f), and then the His37 residues (Figures 6a,e and 7b). It becomes clear that the (R)-rimantadine is more resistant to dehydration (see Figures 4, 6b,f, and 7a) and descended further into the channel than the (S)-rimantadine (see Figures 6a,e and 7b), and the data shown in Figure S5 showed that the (R)-enantiomer appears to show a greater propensity for hydrogen bonding with the His37 residues (see also additional discussion in the Supporting Information). In parallel, when less water is in the binding site, the difference in the computed water network binding free energies between the two enantiomers becomes more significant. For example, a network of 4 water molecules that form the lower water layer appeared to be 2.6 ± 0.6 kcal mol−1 more stable in the presence of the (R)-enantiomer (N = 4 in Table S4). Thus, there is a hypothetical stereochemical difference when the upper water layer is removed from the binding site.
The free energy results obtained also imply that it would be very difficult for a ligand to gain a significant affinity increase by displacing the lower water layer. Given that the binding free energy of the network of 9–10 waters is in the vicinity of ca. 30 kcal mol−1, displacement of the upper layer would result in a network of 4 waters, with a binding free energy in the vicinity of 20 kcal mol−1 (Table S4). This suggests a significant thermodynamic cost associated with the displacement of the lower water layer, in addition to the cost of displacing the upper layer. This is consistent with previous studies of drugs such as spiro-adamantyl amines, which are more extended than rimantadine; these compounds, which bind with high affinity, displace the upper, but not the lower layer of waters in the lumen of the channel.32,45
At lower levels of hydration, due to the chirality of the protein, there is a difference in terms of the computed stability of the water network, which can differentiate the binding of a chiral ligand. Thus, chirality might be an important factor in designing new, novel drugs, which displace the upper layer of waters and tightly engage to tightly bound the lower water layer.
CONCLUSIONS
The environment within the M2 pore is chiral: l-amino acids form right-handed α-helical monomers, which come together to form a left-handed tetrameric helical bundle. Since this chiral environment has the potential to create enantiospecific binding, it was important to investigate whether the chirality of drugs had an effect on drug binding and inhibition. Though the chirality of the M2 WT protein creates the possibility of preferential binding of rimantadine enantiomers, we did not observe this in X-ray crystal structures of M2(22–46) WT bound to (R)- and (S)-rimantadine, in EP kinetics assays using the full-length channel, or in antiviral plaque assays. Instead, the crystal structures showed that both rimantadine enantiomers bind to the M2 pore and that the ethylammonium headgroup of both rimantadine enantiomers averages over four orientations. We observed slight differences in the hydration of each rimantadine enantiomer in the crystal structures. However, these slight differences in hydration did not have an effect on drug binding or channel inhibition. We hypothesize that the two-layer water network, the upper layer adjacent to Ala30 and the lower layer to Gly34, is sufficiently flexible to mask the chirality of the binding site. The GCMC/MD simulation analysis indicated that, as the upper layer of waters is removed, a more substantial chiral difference is observed in the stability of the water networks; i.e., two layers of water are needed to lose this chiral templating effect. This potential enantiomeric selectivity may be of interest for compounds that are able to displace the upper water layer.
MATERIALS AND METHODS
Sample Preparation, Crystallization, Diffraction, and Structure Solving.
Solid-phase peptide synthesis using FMOC chemistry and sample preparation and peptide-drug sample reconstitution into the lipid cubic phase (LCP) were carried out using previously described methods32,46,47 and the Supporting Information. Data collection was carried out at 100 K at the Advanced Light Source (ALS) beam 8.3.1. Data processing was carried out in Mosflm48, and data were scaled in the CCP4 suite.49 The structure was solved using molecular replacement in Phaser-MR50 with structure PDB ID 6BKL32 as a search model. See the Supporting Information for detailed methods regarding crystallization and data collection.
TEVC Assay.
mRNA syntheses expressing the WT of A/Udorn/72 full M2 protein, oocyte culture, microinjection of oocytes, and electrophysiological TEVC recordings were carried out as previously described.35,37,51 The Kd measurement and curve fitting were carried out as previously described.36,37 Briefly, the percentage of current during the application of rimantadine enantiomers and washout protocol was plotted with an association then dissociation equation in GraphPad Prism 5.
Plaque Reduction Assay.
Plaque reduction assays were performed in MDCK cells with the A/Soloman Island/3/2006 (H1N1) virus as previously described.36,51–53 A/Soloman Island/3/2006 (H1N1) M2 contains the same sequence of M2 WT as A/Udorn/72. Briefly, confluent cells were washed with phosphate buffered saline (PBS) and infected with virus diluted in Dulbecco’s Modified Eagle’s Medium (DMEM) medium supplemented with 0.5% BSA for a final concentration of approximately 100 plaque forming units (PFU) per well. Viral infection was synchronized for 30 min at 4 °C and then incubated for 1 h at 37 °C in a 5% CO2 atmosphere. The inoculum was aspirated, and cells were washed and incubated in a DMEM overlay media containing different concentrations of compound, 2 μg/mL N-acetyl trypsin, and 1.2% Avicel microcrystalline cellulose (FMC BioPolymer, Philadelphia, PA) at 37 °C in a 5% CO2 atmosphere. Cells were stained 2 days post infection with 0.2% crystal violet dye. EC50 values were calculated by plotting the plaque area per well against the rimantadine concentration applied with a dose–response function in Prism 5.
Molecular Dynamics Simulations.
The complex of M2(22-46) WT with (S)-rimantadine (PDB ID 6US8) or (R)-rimantadine (PDB ID 6US9) was embedded in a POPC lipid bilayer extending 30 Å beyond the solutes, and the system was placed in an orthorhombic box (90 × 90 × 105 Å3). The number of lipids added was ca. 200. The bilayer was then solvated by a 30 Å thick layer of water. Na+ and Cl− ions were placed in the water phase to neutralize the systems and to reach the experimental salt concentration of 0.150 M NaCl. The total number of atoms was ca. 80 000. Membrane generation and system solvation was carried out using the “System Builder” utility of Maestro-Desmond 202054, and periodic boundary conditions were applied. We performed 300 ns MD simulations using the amber99sb force field for the protein, CHARMM36 force field to lipids, while the generalized Amber force field (GAFF) was used for the ligand parameterization with the antechamber module of AMBER18 software; intermolecular interactions were calculated with amber99sb force field. The TIP3P force field was used for the water. The MD simulations were performef with the Desmond 202055,56 software for each of the two rimantadine enantiomers in complex with M2(22-46) WT (in the X-ray crystal structures PDB ID 6US8 or PDB ID 6US9) embedded in a hydrated lipid bilayer to test the stability of the experimental structures. See the Supporting Information for a detailed description of the simulation conditions, the protocol and related references.
GCMC Titration Calculations.
GCMC method was used to allow the number of waters within a simulation to vary according to the chemical potential, in order to simulate the grand canonical (μVT) ensemble. This allows the waters within the M2 pore to exchange with bulk water much more rapidly, bypassing the kinetic barriers which can limit water sampling in conventional MD simulations (see Supporting Information and Figure S4). GCMC is adequate for the simulation of the hydration of M2 pore at equilibrium with bulk water using the chemical potential reflected by the Bequil value. Description of the theory upon which GCMC is based and a detailed description of the simulation conditions can be found in the Supporting Information. The GCMC simulations were carried out using version 7.3.1 of OpenMM.57 Version 1.0.0 of the grand Python module58 was used with OpenMM to carry out the GCMC sampling, with the MD sampling carried out using the native OpenMM functionality.
For the free energy calculations reported here, the AMBER ff14sb force field was used to describe the protein and intermolecular interactions, the lipid17 force field for the membrane, TIP3P for the water and the Joung-Cheatham parameters were used for the ions. The ligands were described using the GAFF with AM1-BCC charges, generated using antechamber. All non-bonded interactions within 12 Å were calculated directly, Lennard-Jones interactions were switched to zero from 10–12 Å, and long-range electrostatic interactions were calculated using the Particle Mesh Ewald method (see the Supporting Information for details and references).
The region in which GCMC moves were carried out was defined as a sphere with a radius of 6.0 Å (for which Bequil = −6.820), centered on the average coordinate of the Cα atoms of the Gly34 residues. This was chosen to ensure that the two water layers of interest are covered by the GCMC sampling. Version 1.0.0 of the grand Python module58 was used alongside OpenMM to carry out the GCMC sampling, with the MD sampling carried out using the native OpenMM functionality.
For each enantiomer, an initial equilibration procedure was applied, beginning with a very short minimization, following which the waters present in the GCMC region were deleted. This was followed by GCMC/MD equilibration at Bequil, to allow the waters in the GCMC region to be repopulated without bias, and then constant pressure equilibration to allow the volume of the system to equilibrate. The GCMC/MD stage was carried out in three phases to ensure rigorous equilibration of the waters before significant movement of the protein complex: first 1 ps of MD, with 5k moves attempted every 100 fs; then 10 ps of MD, with 100 moves every 200 fs; and finally, 50 ps of MD, with 500 moves every 250 fs. The constant pressure equilibration was carried out for 500 ps. The resulting equilibrated structure for each enantiomer was used as a starting point for each set of independent repeats in order to ensure that they all had the same system volume and, therefore, sampled the same ensemble.
For each independent set of repeats, for each enantiomer, a further GCMC/MD equilibration was carried out at Bequil for 500 ps, with 50 GCMC moves executed every 250 fs. From this, multiple simulations were run at 21 equally spaced B values, from −24.820 to −4.820. Each of these simulations was equilibrated for 500 ps, with 50 GCMC moves every 250 fs, to allow the number of waters to adjust to the new B value. This was followed by a production run of 2.5 ns, with 20 GCMC moves per 500 fs, with simulation frames written out every 5 ps. The average number of waters observed in the GCMC region, 〈N〉, was calculated over these production runs. Three sets of these simulations were carried out for each enantiomer, generating three values of 〈N〉 per B value, per enantiomer (Table S3).
Free Energy Calculations.
In order to determine the relative free energies of both hydration and binding for the two enantiomers, free energy calculations were performed in OpenMM, using version 0.7.1 of the perses Python module,59 using the force fields and non bonded cutoff values described previously. This involved perturbing each of the enantiomers into an achiral intermediate compound, in which the methyl group of rimantadine is replaced by a hydrogen atom. Full details of these calculations and references are provided in the Supporting Information.
Supplementary Material
ACKNOWLEDGMENTS
The authors thank Pil Seok Chae (Hanyang University, Seoul, South Korea) for providing the MNG detergent for crystallization trials. Data collection was carried out at ALS 8.3.1. Beamline 8.3.1 at the Advanced Light Source operated by the University of California Office of the President, Multicampus Research Programs and Initiatives grant MR-15-328599 and NIGMS grants P30 GM124169 and R01 GM124149. The authors thank George Meigs and James Holton at ALS 8.3.1 for support during data collection. A.K. acknowledges the access to ARIS Supercomputer for the MD simulations using Desmond software. The authors acknowledge the use of the IRIDIS High-Performance Computing Facility and associated support services at the University of Southampton in the completion of this work.
Funding
J.L.T. and W.F.D. were supported by NIH grants R35-GM122603 and R01-GM117593. M.L.S. is supported by the EPSRC-funded CDT in Next Generation Computational Modeling, under grant EP/L015382/1. H.E.B.M acknowledges support from a Molecular Sciences Software Institute Fellowship and Relay Therapeutics. J.W. was supported by NIH grants AI119187 and AI144887. A.K. acknowledges support from Chiesi Hellas (SARG grant no. 10354). The use of the LCP crystallization robot was made possible by the National Center for Research Resources Grant 1S10RR027234-01.
Footnotes
Supporting Information
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.biochem.1c00437.
Details for sample preparation and crystallization data, for equilibrium molecular dynamics simulations, alchemical free energy calculations, GCMC theory and simulation, and additional results for GCMC titration of waters (PDF)
Accession Codes
PDB IDs: 6US9 and 6US8.
The authors declare no competing financial interest.
Contributor Information
Jessica L. Thomaston, Laboratory of Medicinal Chemistry, Section of Pharmaceutical Chemistry, Department of Pharmacy, University of California, San Francisco, California 94158, United States.
Marley L. Samways, School of Chemistry, University of Southampton, Southampton SO17 1BJ, United Kingdom.
Athina Konstantinidi, Department of Pharmaceutical Chemistry, School of Pharmacy, National and Kapodistrian University of Athens, 15771 Athens, Greece.
Chunlong Ma, Department of Pharmacology and Toxicology, College of Pharmacy, University of Arizona, Tucson, Arizona 85721, United States.
Yanmei Hu, Department of Pharmacology and Toxicology, College of Pharmacy, University of Arizona, Tucson, Arizona 85721, United States.
Hannah E. Bruce Macdonald, Computational and Systems Biology Program, Memorial Sloan Kettering Cancer Center, New York, New York 10065, United States.
Jun Wang, Department of Pharmacology and Toxicology, College of Pharmacy, University of Arizona, Tucson, Arizona 85721, United States.
Jonathan W. Essex, School of Chemistry, University of Southampton, Southampton SO17 1BJ, United Kingdom.
William F. DeGrado, Laboratory of Medicinal Chemistry, Section of Pharmaceutical Chemistry, Department of Pharmacy, University of California, San Francisco, California 94158, United States
Antonios Kolocouris, Department of Pharmaceutical Chemistry, School of Pharmacy, National and Kapodistrian University of Athens, 15771 Athens, Greece.
REFERENCES
- (1).To J, Surya W, and Torres J, (2016) Targeting the Channel Activity of Viroporins. Adv. Protein Chem. Struct. Biol 104, 307–355. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (2).Dong G, Peng C, Luo J, Wang C, Han L, Wu B, Ji G, and He H, (2015) Adamantane-resistant influenza A viruses in the world (1902—2013): frequency and distribution of M2 gene mutations. PLoS One 10, No. e0119115. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (3).Bright RA, Medina MJ, Xu XY, Perez-Oronoz G, Wallis TR, Davis XHM, Povinelli L, Cox NJ, and Klimov AI, (2005) Incidence of adamantane resistance among influenza A (H3N2) viruses isolated worldwide from 1994 to 2005: a cause for concern. Lancet 366, 1175–1181. [DOI] [PubMed] [Google Scholar]
- (4).Lampejo T, (2020) Influenza and antiviral resistance: an overview. Eur. J. Clin. Microbiol. Infect. Dis 39, 1201–1208. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (5).Chizhmakov IV, Geraghty FM, Ogden DC, Hayhurst A, Antoniou M, and Hay AJ, (1996) Selective proton permeability and pH regulation of the influenza virus M2 channel expressed in mouse erythroleukaemia cells. J. Physiol 494, 329–336. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (6).Lin TI, and Schroeder C, (2001) Definitive assignment of proton selectivity and attoampere unitary current to the M2 ion channel protein of influenza A virus. J. Virol 75, 3647–3656. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (7).Mould JA, Drury JE, Frings SM, Kaupp UB, Pekosz A, Lamb RA, and Pinto LH, (2000) Permeation and activation of the M-2 ion channel of influenza A virus. J. Biol. Chem 275, 31038–31050. [DOI] [PubMed] [Google Scholar]
- (8).Shimbo K, Brassard DL, Lamb RA, and Pinto LH, (1996) Ion selectivity and activation of the M2 ion channel of influenza virus. Biophys. J 70, 1335–1346. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (9).Duff KC, and Ashley RH, (1992) The transmembrane domain of influenza A M2 protein forms amantadine-sensitive proton channels in planar lipid bilayers. Virology 190, 485–489. [DOI] [PubMed] [Google Scholar]
- (10).Ma CL, Polishchuk AL, Ohigashi Y, Stouffer AL, Schon A, Magavern E, Jing XH, Lear JD, Freire E, Lamb RA, DeGrado WF, and Pinto LH, (2009) Identification of the functional core of the influenza A virus A/M2 proton-selective ion channel. Proc. Natl. Acad. Sci. U. S. A 106, 12283–12288. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (11).Tang YJ, Zaitseva F, Lamb RA, and Pinto LH, (2002) The gate of the influenza virus M-2 proton channel is formed by a single tryptophan residue. J. Biol. Chem 277, 39880–39886. [DOI] [PubMed] [Google Scholar]
- (12).Wang C, Lamb RA, and Pinto LH, (1995) Activation of the M2 ion channel of influenza virus: a role for the transmembrane domain histidine residue. Biophys. J 69, 1363–1371. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (13).Martin K, and Heleniust A, (1991) Nuclear transport of influenza virus ribonucleoproteins: The viral matrix protein (M1) promotes export and inhibits import. Cell 67, 117–130. [DOI] [PubMed] [Google Scholar]
- (14).Hay AJ, Wolstenholme AJ, Skehel JJ, and Smith MH, (1985) The molecular basis of the specific anti-influenza action of amantadine. EMBO J. 4, 3021–3024. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (15).Wang C, Takeuchi K, Pinto LH, and Lamb RA, (1993) Ion-channel activity of influenza-A virus M(2) protein - characterization of the amantadine block. J. Virol 67, 5585–5594. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (16).Drakopoulos A, Tzitzoglaki C, McGuire K, Hoffmann A, Konstantinidi A, Kolokouris D, Ma C, Freudenberger K, Hutterer J, Gauglitz G, Wang J, Schmidtke M, Busath DD, and Kolocouris A, (2018) Unraveling the Binding, Proton Blockage, and Inhibition of Influenza M2 WT and S31N by Rimantadine Variants. ACS Med. Chem. Lett 9, 198–203. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (17).Duque MD, Ma C, Torres E, Wang J, Naesens L, Juárez-Jiménez J, Camps P, Luque FJ, DeGrado WF, Lamb RA, Pinto LH, and Vázquez S, (2011) Exploring the Size Limit of Templates for Inhibitors of the M2 Ion Channel of Influenza A Virus. J. Med. Chem 54, 2646–2657. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (18).Skehel JJ, Hay AJ, and Armstrong JA, (1978) On the mechanism of inhibition of influenza virus replication by amantadine hydrochloride. J. Gen. Virol 38, 97–110. [DOI] [PubMed] [Google Scholar]
- (19).Ciampor F, Bayley PM, Nermut MV, Hirst EMA, Sugrue RJ, and Hay AJ, (1992) Evidence that the amantadine-induced, M2-mediated conversion of influenza-A virus hemagglutinin to the low pH conformation occurs in an acidic transgolgi compartment. Virology 188, 14–24. [DOI] [PubMed] [Google Scholar]
- (20).Sugrue RJ, Bahadur G, Zambon MC, Hall-Smith M, Douglas AR, and Hay AJ, (1990) Specific structural alteration of the influenza haemagglutinin by amantadine. EMBO J. 9, 3469–3476. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (21).Cady SD, Schmidt-Rohr K, Wang J, Soto CS, Degrado WF, and Hong M, (2010) Structure of the amantadine binding site of influenza M2 proton channels in lipid bilayers. Nature 463, 689–692. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (22).Andreas LB, Barnes AB, Corzilius B, Chou JJ, Miller EA, Caporini M, Rosay M, and Griffin RG, (2013) Dynamic nuclear polarization study of inhibitor binding to the M2(18–60) proton transporter from influenza A. Biochemistry 52, 2774–2782. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (23).Thomaston JL, Polizzi NF, Konstantinidi A, Wang J, Kolocouris A, and DeGrado WF, (2018) Inhibitors of the M2 proton channel engage and disrupt transmembrane networks of hydrogen-bonded waters. J. Am. Chem. Soc 140, 15219–15226. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (24).Wright AK, Batsomboon P, Dai J, Hung I, Zhou H-X, Dudley GB, and Cross TA, (2016) Differential Binding of Rimantadine Enantiomers to Influenza A M2 Proton Channel. J. Am. Chem. Soc 138, 1506–1509. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (25).Drakopoulos A, Tzitzoglaki C, Ma C, Freudenberger K, Hoffmann A, Hu Y, Gauglitz G, Schmidtke M, Wang J, and Kolocouris A, (2017) Affinity of rimantadine enantiomers against influenza A/M2 protein revisited. ACS Med. Chem. Lett 8, 145–150. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (26).Aldrich PE, Hermann EC, Meier WE, Paulshock M, Prichard WW, Synder JA, and Watts JC, (1971) Antiviral agents. 2. Structure-activity relations of compounds related to 1-adamantanamine. J. Med. Chem 14, 535–543. [DOI] [PubMed] [Google Scholar]
- (27).Bruce Macdonald HE, Cave-Ayland C, Ross GA, and Essex JW, (2018) Ligand Binding Free Energies with Adaptive Water Networks: Two-Dimensional Grand Canonical Alchemical Perturbations. J. Chem. Theory Comput 14, 6586–6597. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (28).Ross GA, Bodnarchuk MS, and Essex JW, (2015) Water Sites, Networks, And Free Energies with Grand Canonical Monte Carlo. J. Am. Chem. Soc 137, 14930–14943. [DOI] [PubMed] [Google Scholar]
- (29).Ross GA, Bruce Macdonald HE, Cave-Ayland C, Cabedo Martinez AI, and Essex JW, (2017) Replica-Exchange and Standard State Binding Free Energies with Grand Canonical Monte Carlo. J. Chem. Theory Comput 13, 6373–6381. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (30).Stouffer AL, Acharya R, Salom D, Levine AS, Di Costanzo L, Soto CS, Tereshko V, Nanda V, Stayrook S, and DeGrado WF, (2008) Structural basis for the function and inhibition of an influenza virus proton channel. Nature 451, 596–599. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (31).Thomaston JL, Konstantinidi A, Liu L, Lambrinidis G, Tan J, Caffrey M, Wang J, Degrado WF, and Kolocouris A, (2020) X-ray Crystal Structures of the Influenza M2 Proton Channel Drug-Resistant V27A Mutant Bound to a Spiro-Adamantyl Amine Inhibitor Reveal the Mechanism of Adamantane Resistance. Biochemistry 59, 627–634. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (32).Thomaston JL, Polizzi NF, Konstantinidi A, Wang J, Kolocouris A, and DeGrado WF, (2018) Inhibitors of the M2 proton channel engage and disrupt transmembrane networks of hydrogen-bonded waters. J. Am. Chem. Soc 140, 15219–15226. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (33).Schnell JR, and Chou JJ, (2008) Structure and mechanism of the M2 proton channel of influenza A virus. Nature 451, 591–U512. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (34).Cady SD, Schmidt-Rohr K, Wang J, Soto CS, DeGrado WF, and Hong M, (2010) Structure of the amantadine binding site of influenza M2 proton channels in lipid bilayers. Nature 463, 689–U127. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (35).Musharrafieh R, Lagarias P, Ma C, Hau R, Romano A, Lambrinidis G, Kolocouris A, and Wang J, (2020) Investigation of the Drug Resistance Mechanism of M2-S31N Channel Blockers through Biomolecular Simulations and Viral Passage Experiments. ACS Pharmacology & Translational Science 3, 666–675. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (36).Musharrafieh R, Ma C, and Wang J, (2018) Profiling the in vitro drug-resistance mechanism of influenza A viruses towards the AM2-S31N proton channel blockers. Antiviral Res. 153, 10–22. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (37).Wang Y, Hu Y, Xu S, Zhang Y, Musharrafieh R, Hau RK, Ma C, and Wang J, (2018) In Vitro Pharmacokinetic Optimizations of AM2-S31N Channel Blockers Led to the Discovery of Slow-Binding Inhibitors with Potent Antiviral Activity against Drug-Resistant Influenza A Viruses. J. Med. Chem 61, 1074–1085. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (38).Tembre BL, and Mc Cammon JA, (1984) Ligand-receptor interactions. Comput. Chem 8, 281–283. [Google Scholar]
- (39).Zwanzig RW, (1954) High-Temperature Equation of State by a Perturbation Method. I. Nonpolar Gases. J. Chem. Phys 22, 1420–1426. [Google Scholar]
- (40).Bennett CH, (1976) Efficient estimation of free energy differences from Monte Carlo data. J. Comput. Phys 22, 245–268. [Google Scholar]
- (41).Shirts MR, and Chodera JD, (2008) Statistically optimal analysis of samples from multiple equilibrium states. J. Chem. Phys 129, 124105–124105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (42).Cournia Z, Allen B, and Sherman W, (2017) Relative Binding Free Energy Calculations in Drug Discovery: Recent Advances and Practical Considerations. J. Chem. Inf. Model 57, 2911–2937. [DOI] [PubMed] [Google Scholar]
- (43).Gkeka P, Eleftheratos S, Kolocouris A, and Cournia Z, (2013) Free energy calculations reveal the origin of binding preference for aminoadamantane blockers of influenza A/M2TM pore. J. Chem. Theory Comput 9, 1272–1281. [DOI] [PubMed] [Google Scholar]
- (44).Ioannidis H, Drakopoulos A, Tzitzoglaki C, Homeyer N, Kolarov F, Gkeka P, Freudenberger K, Liolios C, Gauglitz G, Cournia Z, Gohlke H, and Kolocouris A, (2016) Alchemical Free Energy Calculations and Isothermal Titration Calorimetry Measurements of Aminoadamantanes Bound to the Closed State of Influenza A/M2TM. J. Chem. Inf. Model 56, 862–876. [DOI] [PubMed] [Google Scholar]
- (45).Wang J, Ma C, Fiorin G, Carnevale V, Wang T, Hu F, Lamb RA, Pinto LH, Hong M, Klein ML, and DeGrado WF, (2011) Molecular dynamics simulation directed rational design of inhibitors targeting drug-resistant mutants of influenza A virus M2. J. Am. Chem. Soc 133, 12834–12841. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (46).Thomaston JL, Alfonso-Prieto M, Woldeyes RA, Fraser JS, Klein ML, Fiorin G, and DeGrado WF, (2015) High-resolution structures of the M2 channel from influenza A virus reveal dynamic pathways for proton stabilization and transduction. Proc. Natl. Acad. Sci U. S. A 112, 14260–14265. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (47).Thomaston JL, and DeGrado WF, (2016) Crystal structure of the drug-resistant S31N influenza M2 proton channel. Protein Sci. 25, 1551–1554. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (48).Battye TGG, Kontogiannis L, Johnson O, Powell HR, and Leslie AGW, (2011) iMOSFLM: a new graphical interface for diffraction-image processing with MOSFLM. Acta Crystallogr. Sect. D: Biol. Crystallogr 67, 271–281. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (49).Winn MD, Ballard CC, Cowtan KD, Dodson EJ, Emsley P, Evans PR, Keegan RM, Krissinel EB, Leslie AGW, McCoy A, McNicholas SJ, Murshudov GN, Pannu NS, Potterton EA, Powell HR, Read RJ, Vagin A, and Wilson KS, (2011) Overview of the CCP4 suite and current developments. Acta Crystallogr., Sect. D: Biol. Crystallogr 67, 235–242. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (50).McCoy AJ, Grosse-Kunstleve RW, Adams PD, Winn MD, Storoni LC, and Read RJ, (2007) Phaser crystallographic software. J. Appl. Crystallogr 40, 658–674. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (51).Hu Y, Musharrafieh R, Ma C, Zhang J, Smee DF, DeGrado WF, and Wang J, (2017) An M2-V27A channel blocker demonstrates potent in vitro and in vivo antiviral activities against amantadine-sensitive and -resistant influenza A viruses. Antiviral Res. 140, 45–54. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (52).Ma C, Li F, Musharrafieh RG, and Wang J, (2016) Discovery of cyclosporine A and its analogs as broad-spectrum anti-influenza drugs with a high in vitro genetic barrier of drug resistance. Antiviral Res. 133, 62–72. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (53).Ma C, Zhang J, and Wang J, (2016) Pharmacological Characterization of the Spectrum of Antiviral Activity and Genetic Barrier to Drug Resistance of M2-S31N Channel Blockers. Mol. Pharmacol 90, 188–198. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (54).Schrödinger L, (2020) Maestro-Desmond Interoperability Tools. [Google Scholar]
- (55).Bowers KJ, Chow DE, Xu H, Dror RO, Eastwood MP, Gregersen BA, Klepeis JL, Kolossvary I, Moraes MA, Sacerdoti FD, Salmon JK, Shan Y, and Shaw DE, (2006) Scalable Algorithms for Molecular Dynamics Simulations on Commodity Clusters. In SC ’06: Proceedings of the 2006 ACM/IEEE Conference on Supercomputing, pp 43–43. [Google Scholar]
- (56).Schrodinger L, (2020) Release 2020–4: Desmond Molecular Dynamics System, D.E. Shaw Research, New York, NY. [Google Scholar]
- (57).Eastman P, Swails J, Chodera JD, McGibbon RT, Zhao Y, Beauchamp KA, Wang LP, Simmonett AC, Harrigan MP, Stern CD, Wiewiora RP, Brooks BR, and Pande VS, (2017) OpenMM 7: Rapid development of high performance algorithms for molecular dynamics. PLoS Comput. Biol 13, No. e1005659. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (58).Samways ML, Bruce Macdonald HE, and Essex JW, (2020) grand: A Python Module for Grand Canonical Water Sampling in OpenMM. J. Chem. Inf. Model 60, 4436–4441. [DOI] [PubMed] [Google Scholar]
- (59).Rufa DA, Bruce Macdonald HE, Fass J, Wieder M, Grinaway PB, Roitberg AE, Isayev O, and Chodera JD, (2020) Towards chemical accuracy for alchemical free energy calculations with hybrid physics-based machine learning/molecular mechanics potentials, bioRxiv, 2020.2007.2029.227959. [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.