Abstract
Electrical cell signaling requires adjustment of ion channel, receptor, or transporter function in response to changes in membrane potential. For the majority of such membrane proteins, the molecular details of voltage sensing remain insufficiently understood. Here, we present a molecular dynamics simulation-based method to determine the underlying charge movement across the membrane—the gating charge—by measuring electrical capacitor properties of membrane-embedded proteins. We illustrate the approach by calculating the charge transfer upon membrane insertion of the HIV gp41 fusion peptide, and validate the method on two prototypical voltage-dependent proteins, the Kv1.2 K+ channel and the voltage sensor of the Ciona intestinalis voltage-sensitive phosphatase, against experimental data. We then use the gating charge analysis to study how the T1 domain modifies voltage sensing in Kv1.2 channels and to investigate the voltage dependence of the initial binding of two Na+ ions in Na+-coupled glutamate transporters. Our simulation approach quantifies various mechanisms of voltage sensing, enables direct comparison with experiments, and supports mechanistic interpretation of voltage sensitivity by fractional amino acid contributions.
Introduction
Many ion channels, transporters, neurotransmitter receptors, and enzymes change conformation in response to changes in the transmembrane voltage (Vm); the consequent voltage-dependent changes in function are critical for a large number of processes in excitable and nonexcitable cells. The voltage dependence of membrane protein function is caused by global or localized conformational changes, ion or ligand binding, protonation, or changes in water accessibility that reshape the electric field (1). It can be quantified as transmembrane charge transfer, often denoted as the gating charge (Qg), i.e., the effective number of charges that need to be moved across the whole electric field to provide the electric energy necessary for these changes. In recent years, experimental methods have been developed to quantify Qg and study the molecular determinants of voltage dependence, resulting in great progress in understanding the principles of voltage sensing (2, 3).
Computer simulations based on structures of various voltage-dependent membrane proteins in multiple conformations enable structural information to be linked to function. Poisson-Boltzmann calculations provide satisfying estimates of Qg for classical voltage-gated cation channels (2, 4, 5, 6) or transporters (7, 8). In particular, the recently developed APBSmem program greatly facilitates Poisson-Boltzmann simulations of membrane proteins (9, 10). Unfortunately, its representation of the ionic solution and membrane as a dielectric continuum, and the simplified geometry of the dielectric boundaries limit the suitability of the PB model, for instance when changes in water accessibility refocus the electric field within narrow protein crevices (11, 12). Recent coarse-grained models have been useful in simulating the dynamics of the gating charge transfer, but still carry the inherent risks of implicit solvent representation and a simplified protein description (13). Molecular dynamics (MD) simulations may consider all possible mechanisms of voltage gating (14, 15, 16, 17, 18), but usually come at a high computational cost. To gain insights into the underlying mechanisms and to permit validation by mutagenesis experiments, it is important to quantify the contributions of individual residues to the overall voltage sensitivity. Thus far, two MD approaches have been used to this end. One approach analyzes changes in electrical distance—that is, the fraction of the electric field across the membrane—of certain residues by evaluating the dependency of the local electrostatic potential on Vm (14, 17, 18). This approach has provided important mechanistic insights, but is limited by the need to calculate the electrostatic potential in three dimensions, and the convergence of the local potential and the statistical accuracy of the calculated Qg contributions have not been investigated in these studies. In other studies, perturbative free energy calculations have been used to determine the Qg contributions via the charging free energy of certain amino acids (16, 18). These simulations are computationally expensive, and thus difficult to be routinely applied to a large number of residues.
To overcome these limitations, we developed a user-friendly and efficient method for gating charge calculations using computational electrophysiology (CompEL) simulations (19, 20). Our approach enables accurate estimates of the voltage sensitivity and provides mechanistic insight via directly determining the fractional amino acid contributions to Qg. It is based on the simulation of the membrane/protein capacitor charging process, as pioneered by Treptow et al. (17), and permits direct determination of Qg and the change in capacitance associated with any conformational change, taking all possible components of voltage sensitivity explicitly into account.
Materials and Methods
MD simulations
MD simulations were carried out using GROMACS 5 (21). The AMBER99SB-ILDN force field (22) was used for the voltage-sensing domain of the Ciona intestinalis voltage-sensitive phosphatase (Ci-VSD), Kv1.2, and GltPh, and the GROMOS 43a1 force field (23) was used for the gp41 peptide. All amino acids were modeled in their default ionization state; in Kv1.2 all histidine residues were assigned a protonated state with a charge of +1 e to reflect the most probable state at neutral pH according to pKa calculations on this channel (16). In simulations with the AMBER force field, ion parameters were taken from Joung and Cheatham (24). All simulation boxes contained equilibrated palmitoyloleoyl phosphatidylcholine bilayers (25), except for gp41, where dimyristoyl phosphatidylcholine (26) was used. Ions were added to obtain bulk concentrations in the aqueous solutions of 100 mM KCl for all simulations, except for gp41, which used 70 mM NaCl. Water was modeled using the SPC/E water model (27) in all simulations, except for gp41, where SPC (28) was used. Using virtual sites for hydrogen atoms enabled simulations to be performed with 4 fs time steps.
Membrane proteins were embedded into each membrane using g_membed (29), informed by the Orientations of Proteins in Membranes database (30), while ensuring that S1 and S2 helices of the activated and resting states were superimposed (31). Before production runs, single-bilayer systems were equilibrated in the absence of transmembrane voltage during 100–500 ns simulation with position restraints on the protein heavy atoms, followed by ∼20 ns with backbone-only position restraints.
Initial structures used for the simulations of Ci-VSD in the down- and up-states were taken from the Protein Data Bank (PDB): 4G80 and 4G7V (15), respectively, using the residue range 106–236. The up-state structure of Ci-VSD (PDB: 4G7V) carried the mutation R217E, which was reverted to wild-type using the software MODELLER 9.10 (32). Kv1.2 structures were taken from Pathak et al. (5), using the residue range 32–421. The gp41 fusion peptide simulation system is derived from a single-layered MD setup in Gapsys et al. (33), and the peptide structure originates from Charloteaux et al. (34). The GltPh crystal structures used in this study are symmetric, outward-facing trimers either in an apo (PDB: 4OYE) or in a sodium-only bound state (PDB: 4OYF), using the residue range 6–416 (35). The sodium-only structure shows the Na1 ion. So far, the second Na+ ion, termed Na3, has not been resolved in any GltPh crystal structure. However, it has been predicted by electrophysiology (36) and MD simulations (37), and was recently resolved in a novel aspartate-bound structure of another prokaryotic transporter homolog, GltTk (38). However, because no sodium-only bound GltTk structure is available, we decided to use GltPh in this study. GltPh and GltTk share highly similar sequence and structure, and the residues forming the Na3 site are identical. We manually inserted the Na3 ion in the sodium-only GltPh structure according to the GltTk position (38), and found this configuration to be stable in free MD simulations on a timescale of 1 μs.
Computational electrophysiology setup
For the gating charge calculations, we used an antiparallel configuration of two lipid bilayers, both containing the same membrane protein in the same state, adopting the CompEL method (19, 20). The simulation system is created by duplicating a fully equilibrated single-bilayer/protein configuration along the membrane normal, with the copied bilayer being inverted. The size of the simulation box was chosen such that the minimum distance between the two membrane proteins, within the box and across periodic boundaries, was always >40 Å. Thus, interactions between the proteins are negligible, stable density profiles in the bulk region are established, and Vm is well defined. Electroporation through the lipid membrane was not observed in any simulations in this study.
Calculation of the gating charge
After setting up the CompEL systems, the activated/up-state and resting/down-state proteins were further equilibrated—without position restraints—in the presence of +100 or −100 mV, respectively, for ∼500 ns. Position restraints on the protein heavy atoms were then turned on again, using spring constants of 1000 kJ mol−1 nm−2. Finally, the charge titration simulations, i.e., multiple CompEL runs with different ionic charge imbalances, were conducted for ∼80–100 ns for each condition. The use of position restraints is not always required, as Vm(qexc,sol) converges within less than a nanosecond (Fig. S1 b in the Supporting Material), i.e., faster than most voltage-dependent conformational changes of the protein might take. Nevertheless, the position restraints ensure that the reported gating charges can be assigned to specific conformational states of the protein. Finally, the analyses were performed as described in the main text using Eqs. 3 and 4.
Integration of the charge densities and calculation of Vm by Eq. 2 were performed individually on each frame of the MD trajectory. Consequently, the Vm(qexc,sol) relationships were fitted globally by Eq. 3 to determine qexc,p,λ and Cmp,λ; bootstrap sampling was used to determine standard deviations of the fit parameters.
Numerical errors during the double integration of the charge densities led to small deviations in the electrostatic potential across the periodic boundary. To circumvent this problem––assuming a neutral simulation system—the floating-point sum of all charges divided by the number of all nonzero charges was subtracted from the charges of all charged atoms in the system. This procedure effectively ensures that the electrostatic potential is the same at the upper and lower edge of the simulation box, and is, for example, available in the gmx potential module of GROMACS (21, 39). If certain charged atoms are excluded from the analysis, e.g., to investigate the influence of an amino acid on Vm and Qg, then this procedure will implicitly offset all other nonzero charges by a very small value, thereby keeping the remaining system neutral.
Results
Calculation of gating charges by charge titration simulations using CompEL
The CompEL protocol (19, 20), introduced into GROMACS (21), enables simulations under realistic transmembrane potentials by all-atom MD (40). Using an ion/water exchange procedure, CompEL generates electrochemical potential gradients by maintaining slight differences in ion numbers between two aqueous compartments in a double-bilayer simulation system (Fig. 1 a). CompEL induces excess charges qexc,sol in the aqueous compartments (positive sign, intracellular; negative, extracellular), which result in a well-defined Vm that can be determined by solving Poisson’s equation (41)
(1) |
(2) |
where ψ(z) is the electrostatic potential profile along the membrane normal, and qi and ρi(z) are the charge and the charge density, respectively, of atom i in the MD trajectory. When the membrane proteins are centered in the simulation box, ψ(z) and the first integral of ρi(z), the electric field, will be set to zero at z = 0 as boundary conditions. Because the CompEL setup behaves as an ideal electrical capacitor (Fig. S1), Vm depends on qexc,sol as
(3) |
where Cmp,λ is the capacitance of each of the two lipid bilayers including the protein in state λ. The charge distribution within membrane proteins is usually uneven along the membrane normal, and will be hereafter represented by excess charges ±qexc,p,λ for each of the two identical proteins (positive sign, intracellular side; negative, extracellular). qexc,p,λ describes how the protein in state λ contributes to the total capacitor charge on each side of the membrane. To obtain qexc,p,λ from MD simulations, we constructed CompEL simulation systems for the membrane protein in both the assumed active and resting states, starting from an equilibrated single-bilayer system. The membrane was then duplicated along the z axis, with one bilayer inverted to obtain an antiparallel orientation. This results in application of identical voltages to the two membrane proteins such that the effects of the excess charges of both proteins are additive. Subsequent charge titration simulations—i.e., CompEL simulations at various ionic charge imbalances—provide the Vm(qexc,sol) relationship that is fitted by Eq. 3 to determine qexc,p,λ and Cmp,λ. The charge titration simulations for resting and activated states of the protein provide the gating charge via Eq. 4:
(4) |
According to Eq. 2, the electrostatic potential depends on the additive effects of individual charges. Thus, the specific contribution of, for example, certain amino acids to the voltage sensitivity can be determined by excluding the charges of a group of atoms from the sum over the charge densities in Eq. 2. When a charged residue is excluded, the charges of all other atoms in the system are slightly offset to maintain neutrality of the total system and to increase the accuracy of the numerical integration (see Materials and Methods). Subsequent analysis by Eq. 3 will then provide the altered Qg as if the residue under consideration were neutral; subtraction from the total Qg of the protein yields the specific contribution of that residue. Thereby, the contributions of all amino acids, bound ligands, or lipid molecules can be determined without running new simulations. This implies a linear decomposition of the charge contributions and is thus an approximation. We evaluated this approach by additional MD simulations of neutralizing mutations on the S4 segment of a VSD and found that our charge-exclusion method provides a good approximation of the gating charge contributions (Fig. S2). Furthermore, we note that this approach requires the gating states to be robust against modification of individual residues. This assumption is reasonable, and is implied in every experimental approach for probing individual residue contributions to voltage sensing.
In the following, we illustrate gating charge calculations using CompEL on the membrane insertion of the gp41 fusion peptide, and then test it on two proteins, for which a large body of experimental data exist: the VSD of the Ciona intestinalis voltage-sensitive phosphatase (Ci-VSP), and the Kv1.2 K+ channel. Finally, we employ the approach to quantify the role of the cytoplasmic T1 domain in Kv1.2 voltage sensing and the voltage dependence of Na+ binding to the glutamate transporter homolog GltPh.
Voltage dependence of membrane insertion of HIV gp41
The N-terminal fusion peptide of the HIV protein gp41 consists of a 12-amino-acid α-helix with a positive N-terminus and mediates HIV envelope fusion with the T-cell membrane (34). We assumed that gp41 may insert into the target membrane in two states defined by 180° rotation of the α-helix, then constructed two CompEL systems (Fig. 1 b) and performed charge titration simulations for each conformation (Fig. 1, c and d). Fitting the charge titration plots with Eq. 3 reveals a gating charge of 1.79 e associated with this flip. Voltage may thus alter gp41 membrane insertion and influence susceptibility to HIV infection. Decomposition of the gating charge reveals that contributions are made by the N-terminus and the backbone (Fig. 1 e), but also by an asymmetric curvature of the bilayer around the charged N-terminus (Fig. 1 f). This weakens the influence of the dipole contributions of nearby lipid molecules to Vm and contributes to the total Qg by 0.32 e. By decomposing the gating charge into molecular contributions, the charge titration approach thus reveals additional mechanisms of voltage sensing.
Gating charges conferred by voltage-sensing domains
We next determined gating charges for the VSD of the Ciona intestinalis Ci-VSP and the Kv1.2 K+ channel. In Ci-VSP, the VSD transduces depolarization to activation of an enzyme that dephosphorylates cellular signaling molecules such as phosphoinositides (42). Membrane depolarization induces opening of the K+-selective pore of Kv1.2, and the resulting K+ efflux contributes to repolarization during action potentials in various cell types. Ci-VSD and Kv1.2 have been extensively studied to understand voltage sensing by voltage-gated Na+, K+, and Ca2+ channels in excitable tissues. For both proteins, high-resolution structures for an activated and a resting state (Fig. 2, a and d) have been determined by x-ray crystallography (15) or computational structure prediction and refinement based on x-ray and functional data (5, 31, 43). We decided to use the Kv1.2 models from Pathak et al. (5); these models also include the intracellular T1 domain, and have been thoroughly tested against experimental data and shown to be consistent with other structure predictions (31, 43). Both Ci-VSP and Kv1.2 contain a VSD made up of four helices (S1–S4), with S4 containing several positive side chains, but differ in the extent of voltage-dependent rearrangements (15).
Fig. 2, b and e, depicts averaged electrostatic potential profiles at varying ionic charge imbalances, and Fig. 2, c and f, provides the corresponding charge titration plots for the two proteins in their resting and activated states. This analysis yields a gating charge of 0.95 ± 0.01 e for Ci-VSD, and of 10.10 ± 0.01 e for Kv1.2. These values are in good agreement with experimental values (42, 44, 45). Ci-VSD gating charges, as determined by fitting the voltage dependence of time-integrated gating currents using either a Boltzmann relationship or a five-state model, are scattered around one elementary charge (1.10, 1.26, and 1.41 e) (42, 45). Similarly, gating charges of individual Kv1.2 channels were obtained by limiting slope analysis (9.6 e), as well as by normalizing macroscopic gating currents to the number of channels determined through noise analysis (9.96 e) (44).
The Kv1.2 analysis illustrates a further advantage of our charge titration method. Previous MD-based gating charge calculations assumed constant total protein/bilayer capacitances (14, 16, 17, 18). However, our method reveals that the large conformational changes of the channel upon activation result in a decrease in the total capacitance of ∼5%; as the simulation systems used here are larger than those commonly used, the change in Cmp,λ in smaller systems could be even more significant. By fitting both qexc,p,λ and Cmp,λ (Eq. 3), we prevent any possible bias in the Qg estimate due to changes in membrane/protein dielectric properties.
We then investigated the influence of individual residues on Qg by neutralizing individual side chains, i.e., by excluding them from the charge integral in Eq. 2. This approach determined the contribution of every individual amino acid to the total gating charge (Fig. 3; Data S1 and S2). As an internal control, we calculated the cumulative sum of per residue contributions—starting with the gating charge contributions of the non-protein part of the system, e.g., solvent and the membrane—and found these values to converge to the total Qg of the full system (Fig. 3, a and c). As expected, the most prominent positive contributions came from residues located at the S4 segment.
Kv1.2 voltage sensing is modulated by the T1 domain
The Kv1.2 channel has a large cytoplasmic T1 domain that is expected to reside outside of the electric field. During equilibration of the Kv1.2 simulation systems, in both the activated and the resting state, we reproducibly observed conformational rearrangement of the T1 domain and the T1–S1 linker relative to the transmembrane part of the channel, in which the T1/linker approached the lipid bilayer, assisted by the transient formation of salt bridges between both parts of the channel (Fig. S3, a and b). Comparison of the equilibrated resting and activated conformations demonstrates an upward movement of ∼3 Å of the T1/linker during channel activation (Fig. S3 c). Moreover, the T1 domain can sense a small fraction of the transmembrane field (Figs. 2 e and 3 d, inset), and the T1 domain, including the T1–S1 linker, contains a total charge of −5 e per monomer. Movement of these charges within the electric field results in a contribution of ∼−0.6 e per monomer by the T1/linker to the total Qg (Fig. 3 c; Data S2). Earlier electrophysiological experiments demonstrated point mutations in T1 cause voltage-dependent rearrangements of this domain and changes in the apparent Qg in Kv1.2 (46, 47, 48). Our new analysis shows that the T1 domain could affect K+ channel voltage sensing at least partially by directly modifying the gating charge.
Voltage-dependent Na+ binding to the glutamate/aspartate transporter GltPh
Glutamate is the major excitatory neurotransmitter in the mammalian central nervous system. Excitatory amino acid transporters (EAATs) mediate secondary active transport of glutamate or aspartate across the cell membrane by coupling it to the cotransport of three Na+, and one H+, and to the antiport of one K+ (49). This stoichiometry results in the net inward movement of two positive elementary charges per transported glutamate. Na+ binding is a crucial determinant of glutamate or aspartate association to its binding site and substrate translocation. In the apo outward-facing state, the external hairpin HP2 is closed, preventing glutamate/aspartate association (35, 38). The initial binding of two Na+ ions (the Na1 and the Na3 ions) induces or stabilizes HP2 opening (35, 50), such that glutamate can subsequently bind (51, 52) (Fig. 4 a). The bound substrate induces closure of HP2 (51), which is stabilized by the final binding of the third Na+ (the Na2 ion). When HP2 is closed, the fully bound transporter can translocate to the inward-facing conformation (53).
Whereas Na2 binds after the neurotransmitter, Na1 and Na3 need to be occupied in advance to initiate the transport cycle and thus represent the key binding sites that mediate Na+-dependent gating of the substrate binding site (35). Because initial Na+ binding promotes HP2 opening and primes the transporter for substrate binding, the voltage dependence of Na+ binding and Na+-associated conformational changes before glutamate association defines the efficiency of EAAT transporters to capture the free neurotransmitter (54, 55). Recently, a novel crystal structure of a prokaryotic homolog, GltTk, allowed localization of all Na+ binding sites (38) (Fig. 4 b). Comparison of apo and Na+-only bound crystal structures shows that initial Na+ binding is coupled to HP2 opening (35); however, it is unclear if this conformational change occurs before or after binding of the Na1 and Na3 ions. In microsecond-long MD simulations of outward-facing apo GltPh (that is, before spontaneous Na+ binding events), HP2 mainly assumed a closed conformation, as in the apo crystal structure; but spontaneous, reversible HP2 openings—as in the Na+-only bound structure (35)—were reproducibly observed. In contrast, the Na+-only bound GltPh structure resides in its HP2-open conformation in MD simulations (data not shown). We thus hypothesize that Na+-associated HP2 opening occurs via a conformational selection mechanism, where binding of Na+ ions stabilizes HP2 in its open state: HP2 opens spontaneously and permits Na+ binding to the Na1 site. This Na+ ion then hops to the Na3 site; followed by binding of a second ion to the Na1 site. Such a Na+ binding sequence was first proposed by Huang and Tajkhorshid (56) and further corroborated by Bastug et al. (37).
We constructed single-bilayer simulation setups of GltPh in these states and observed reproducible changes of the transmembrane potential profile within the electric field across the membrane, suggesting that each partial reaction is electrogenic (Fig. 4 c). To further quantify the voltage dependence of Na1 and Na3 binding, we performed gating charge calculations as described above. Fitting the charge titration plots for the four consecutive processes permitted the assignment of a gating charge of +0.114 e to HP2 opening before Na+ binding. Subsequent Na+ binding to Na1 confers −0.292 e, and its translocation from Na1 to Na3 −0.084 e. The binding of an additional Na+ to Na1 results in a doubly occupied transporter and is again associated with the transfer of −0.290 e (Fig. 4 d). In total, HP2 opening and binding of Na1 and Na3 are thus associated with a charge transfer of ∼−0.55 e. The voltage dependence of Na+ binding has been experimentally determined by direct measurements of EAAT2-associated charge movements and by voltage-clamp fluorometry on EAAT3 to be in the range of −0.4 to −0.5 e (57, 58, 59). In all experimental studies, single Boltzmann functions have been used to quantify this process, which is expected to slightly underestimate the true voltage dependence. These experiments are thus in good agreement with our gating charge calculations.
Discussion
Voltage-dependent changes in protein function are essential for electrical signal generation and propagation, and regulate many other cellular processes. Thus far, the molecular mechanisms of voltage sensing have mainly been studied using a combination of site-directed mutagenesis and electrophysiological recordings, including sophisticated techniques such as direct measurements of gating currents, noise analysis, and limited slope analysis (1, 2, 3, 4, 60). In many transporters or receptors, substrate binding and translocation contribute to voltage sensing, which can be reliably measured, but the underlying mechanisms are more challenging to be resolved experimentally.
Molecular dynamics-based gating charge calculations permit the direct and efficient assessment of all sources of voltage sensitivity and the contributions of defined groups of atoms such as residues, ligands, or bound ions. While most experiments can only report on the total voltage sensitivity, gating charge calculations can be used to identify functionally relevant and voltage-dependent partial reactions in channels and transporters—which in turn can inform further experiments for validation. In recent years, an increasing number of membrane proteins utilizing noncanonical voltage sensing mechanisms (61, 62, 63, 64, 65) have been reported, and MD simulations can also reliably quantify the gating charge also in these cases. However, all simulations still require structural models of the protein states of interest, and the reliability of the insights from gating charge calculations depends on the quality of the input model. We assume the combination of experiments and computational gating charge calculations to be the most efficient way to obtain a complete understanding of the voltage sensing mechanisms in biological membrane proteins.
We here describe computational charge titration using CompEL all-atom MD simulations to calculate gating charges. CompEL permits membrane protein simulations with electrochemical gradients under most realistic conditions and permits optimal determination of the Vm(qexc,sol) relationships. CompEL enables the user to easily set up multiple charge titration simulations in presence of different ionic charge imbalances for reliable statistical evaluation of the fit. Because the capacitance is treated as a second state-dependent variable, the method can generally be applied to all kinds of membrane proteins. This implies another advantage for the user, because the simulation setups for the different protein states do not need to be of identical atomic composition, and might even be reused from other projects. Although the use of classical MD implies simplifications, e.g., lack of polarizability and fixed protonation states, the agreement of our and other MD data with experiments shows that this falls within acceptable levels. However, such features could be easily provided by alternative force fields and extensions to the MD code.
Before the introduction of CompEL (19, 20), another approach had been developed to enable simulations of single-bilayer MD systems under electrochemical gradients using an artificial vacuum slab to split the aqueous compartment into two parts (66). Treptow et al. (17) applied this vacuum slab technique to gating charge calculations on Kv1.2. In this work, the Vm(qexc,sol) relationships have been derived from two simulations with different charge imbalances with the channel in two states to estimate the gating charge. A potential drawback of the vacuum slab method is that water evaporation at the air/water interface can induce electrical noise, which can bias the transmembrane potential calculation (41). Furthermore, Treptow et al. (17) focused on an initial partial deactivation transition of Kv1.2, and based on these simulations, the authors concluded that the total membrane/protein capacitance remains constant upon the conformational transition of the channel (17). Therefore, the gating charge calculations in this and subsequent studies on Kv1.2 assumed constant capacitance (14, 17). In contrast, for the full deactivation transition of Kv1.2, we observe a statistically significant increase in the total capacitance of ∼5% upon deactivation. Because the conformational changes of the channel are rather large, such a capacitance change is not unexpected. It depends on the protein density in the membrane, and is thus not detectable in electrophysiological experiments. However, to obtain an unbiased estimate of the gating charge from MD simulations, it is crucial to treat the capacitance as a second state-dependent variable. This also applies to another gating charge calculation method that—as opposed to analyzing Vm as a function of applied charge imbalances—analyzes the amount of charge transferred upon the conformational transition, the so-called displacement charge, in MD simulations at constant Vm (18). This approach has proven useful (15, 16), but its general applicability is again limited by its underlying assumption of constant membrane/protein capacitance.
We furthermore describe an efficient approach to estimate the fractional contributions of individual amino acids to the total Qg by investigating the effects of excluding these partial charges from the charge density for the electrostatic potential calculation. Several earlier studies employed another method: The fractional contribution of an individual residue to the total Qg is given by the change in electrical distance upon the gating transition multiplied by the partial charges of this residue. The electrical distance was determined by evaluating the dependency of the local electrostatic potential at this residue on Vm (14, 17, 18); this method thus requires knowledge of the electrostatic potential distribution in three dimensions. In contrast, our charge-exclusion approach evaluates the effect of charge neutralizations on Vm and therefore only requires a one-dimensional potential profile, which is expected to converge faster and thus provide higher statistical accuracy (Fig. S1). The underlying linear decomposition of the charge contributions represents a simplified treatment of the interactions between individual residues and the rest of the system. We tested for potential influences of this simplification in MD simulations of S4 mutations in Ci-VSD that demonstrated a reasonable approximation of the gating charge contributions (Fig. S2). Thus, the charge-exclusion approach determines the contributions of all constituents of the membrane protein system such as amino acids, lipids interacting with the protein, ions, or ligands, for just a few minutes of additional computer time required to reintegrate the charge densities upon charge exclusion.
The double-bilayer setup—and thus the use of two identical copies of the same protein—appears at first glance computationally inefficient. This is not true. The presence of two identical proteins leads to twice as much sampling per simulation time. Thus, less total simulation time is required for the Vm(qexc,sol) relationship to be converged (Fig. S1 b). As commonly done in biomolecular simulations, we evaluate long-range electrostatics by the particle mesh Ewald (PME) method, which scales with an N⋅log(N) relationship, where N is the number of particles. Thus, PME could result in a slight computational overhead when calculating gating charges with CompEL. However, this effect is negligible (because it just applies to the fast Fourier transformation calculations of PME, which represent ∼10% of the computational time needed for one time step in GROMACS) and is usually fully overcompensated by the increase in parallel efficiency of the larger CompEL simulations systems on modern hardware (20). For the initial single-bilayer setups of Kv1.2 or Ci-VSD, we typically achieve simulation performances of ∼36 or 185 ns/day on a single computer node equipped with two standard multicore CPUs and two high-end consumer-class GPUs (67) (see Materials and Methods). In practice, a few hundreds of nanoseconds may be initially required for the equilibration of the simulation systems, followed by a few CompEL charge titration simulations for just a couple of nanoseconds, which is typically sufficient to reach acceptable statistics (Fig. S1 b). Thus, the computational requirements are relatively modest, and even users without access to high-performance computing clusters might be able to conduct the gating charge calculations on just one or two modern desktop computers.
Conclusions
Advances in structural biology have provided high-resolution structures for multiple membrane proteins that now have to be linked with functional data. At present, only a few characteristics of channels, transporters, and receptors can be used for such comparisons. Transport rates are often difficult to obtain experimentally and in simulations, while differences in ionic conditions, membrane background, and force-field inaccuracies make comparing binding affinities and pharmacological signatures problematic. In contrast, voltage-dependent changes in protein function can be easily and reliably quantified for channels, transporters, and receptors in electrophysiological experiments. Computational charge titration by CompEL simulations, as a method to calculate gating charges of voltage-dependent proteins, provides the opportunity to quantify the voltage dependence of multiple individual protein functions for comparison with experimental data. We believe that this approach will not only be useful for further understanding voltage sensing and gating in atomistic detail, but also to test the correctness and biological significance of existing and future structural models by validating theoretical gating charges with experimental results.
Author Contributions
J.-P.M. conceived the project; J.-P.M., R.B., and C.F. designed research; J.-P.M., C.A., and R.B. carried out simulations and analyses; B.L.d.G. advised on theory development and simulation design; and all authors wrote the article.
Acknowledgments
We thank Ulrich Zachariae, Carsten Kutzner, David Köpfer, and Daniel Kortzak for helpful discussions. The authors gratefully acknowledge the computing time granted by the JARA-HPC Vergabegremium and VSR Commission on the supercomputer JURECA (68) at Forschungszentrum Jülich.
Editor: Jose Faraldo-Gomez.
Footnotes
Two data files and three figures are available at http://www.biophysj.org/biophysj/supplemental/S0006-3495(17)30226-6.
Supporting Material
References
- 1.Bezanilla F. How membrane proteins sense voltage. Nat. Rev. Mol. Cell Biol. 2008;9:323–332. doi: 10.1038/nrm2376. [DOI] [PubMed] [Google Scholar]
- 2.Chanda B., Asamoah O.K., Bezanilla F. Gating charge displacement in voltage-gated ion channels involves limited transmembrane movement. Nature. 2005;436:852–856. doi: 10.1038/nature03888. [DOI] [PubMed] [Google Scholar]
- 3.Aggarwal S.K., MacKinnon R. Contribution of the S4 segment to gating charge in the Shaker K+ channel. Neuron. 1996;16:1169–1177. doi: 10.1016/s0896-6273(00)80143-9. [DOI] [PubMed] [Google Scholar]
- 4.Roux B. Influence of the membrane potential on the free energy of an intrinsic protein. Biophys. J. 1997;73:2980–2989. doi: 10.1016/S0006-3495(97)78327-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Pathak M.M., Yarov-Yarovoy V., Isacoff E.Y. Closing in on the resting state of the Shaker K+ channel. Neuron. 2007;56:124–140. doi: 10.1016/j.neuron.2007.09.023. [DOI] [PubMed] [Google Scholar]
- 6.Grabe M., Lecar H., Jan L.Y. A quantitative assessment of models for voltage-dependent gating of ion channels. Proc. Natl. Acad. Sci. USA. 2004;101:17640–17645. doi: 10.1073/pnas.0408116101. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Grewer C., Zhang Z., Gameiro A. Charge compensation mechanism of a Na+-coupled, secondary active glutamate transporter. J. Biol. Chem. 2012;287:26921–26931. doi: 10.1074/jbc.M112.364059. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Tanui R., Tao Z., Grewer C. Electrogenic steps associated with substrate binding to the neuronal glutamate transporter EAAC1. J. Biol. Chem. 2016;291:11852–11864. doi: 10.1074/jbc.M116.722470. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Callenberg K.M., Choudhary O.P., Grabe M. APBSmem: a graphical interface for electrostatic calculations at the membrane. PLoS One. 2010;5:e12722. doi: 10.1371/journal.pone.0012722. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Marcoline F.V., Bethel N., Grabe M. Membrane protein properties revealed through data-rich electrostatics calculations. Structure. 2015;23:1526–1537. doi: 10.1016/j.str.2015.05.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Sansom M.S., Smith G.R., Biggin P.C. The dielectric properties of water within model transbilayer pores. Biophys. J. 1997;73:2404–2415. doi: 10.1016/S0006-3495(97)78269-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Moy G., Corry B., Chung S.H. Tests of continuum theories as models of ion channels. I. Poisson-Boltzmann theory versus Brownian dynamics. Biophys. J. 2000;78:2349–2363. doi: 10.1016/S0006-3495(00)76780-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Kim I., Warshel A. A microscopic capacitor model of voltage coupling in membrane proteins: gating charge fluctuations in Ci-VSD. J. Phys. Chem. B. 2016;120:418–432. doi: 10.1021/acs.jpcb.5b10956. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Delemotte L., Tarek M., Treptow W. Intermediate states of the Kv1.2 voltage sensor from atomistic molecular dynamics simulations. Proc. Natl. Acad. Sci. USA. 2011;108:6109–6114. doi: 10.1073/pnas.1102724108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Li Q., Wanderling S., Perozo E. Structural mechanism of voltage-dependent gating in an isolated voltage-sensing domain. Nat. Struct. Mol. Biol. 2014;21:244–252. doi: 10.1038/nsmb.2768. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Khalili-Araghi F., Jogini V., Schulten K. Calculation of the gating charge for the Kv1.2 voltage-activated potassium channel. Biophys. J. 2010;98:2189–2198. doi: 10.1016/j.bpj.2010.02.056. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Treptow W., Tarek M., Klein M.L. Initial response of the potassium channel voltage sensor to a transmembrane potential. J. Am. Chem. Soc. 2009;131:2107–2109. doi: 10.1021/ja807330g. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Roux B. The membrane potential and its representation by a constant electric field in computer simulations. Biophys. J. 2008;95:4205–4216. doi: 10.1529/biophysj.108.136499. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Kutzner C., Grubmüller H., Zachariae U. Computational electrophysiology: the molecular dynamics of ion channel permeation and selectivity in atomistic detail. Biophys. J. 2011;101:809–817. doi: 10.1016/j.bpj.2011.06.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Kutzner C., Köpfer D.A., Zachariae U. Insights into the function of ion channels by computational electrophysiology simulations. Biochim. Biophys. Acta. 2016;1858(7 Pt. B):1741–1752. doi: 10.1016/j.bbamem.2016.02.006. [DOI] [PubMed] [Google Scholar]
- 21.Abraham M.J., Murtola T., Lindahl E. GROMACS: high performance molecular simulations through multi-level parallelism from laptops to supercomputers. SoftwareX. 2015;1–2:19–25. [Google Scholar]
- 22.Lindorff-Larsen K., Piana S., Shaw D.E. Improved side-chain torsion potentials for the Amber ff99SB protein force field. Proteins. 2010;78:1950–1958. doi: 10.1002/prot.22711. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.van Gunsteren W.F. VdF Hochschulverlag AG an der ETH Zürich; Zürich, Switzerland: 1996. Biomolecular Simulation: The GROMOS96 Manual und User Guide. [Google Scholar]
- 24.Joung I.S., Cheatham T.E., 3rd Determination of alkali and halide monovalent ion parameters for use in explicitly solvated biomolecular simulations. J. Phys. Chem. B. 2008;112:9020–9041. doi: 10.1021/jp8001614. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Berger O., Edholm O., Jähnig F. Molecular dynamics simulations of a fluid bilayer of dipalmitoylphosphatidylcholine at full hydration, constant pressure, and constant temperature. Biophys. J. 1997;72:2002–2013. doi: 10.1016/S0006-3495(97)78845-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Gurtovenko A.A., Patra M., Vattulainen I. Cationic DMPC/DMTAP lipid bilayers: molecular dynamics study. Biophys. J. 2004;86:3461–3472. doi: 10.1529/biophysj.103.038760. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Berendsen H.J.C., Grigera J.R., Straatsma T.P. The missing term in effective pair potentials. J. Phys. Chem. 1987;91:6269–6271. [Google Scholar]
- 28.Berendsen H.J.C., Postma J.P.M., Hermans J. Interaction models for water in relation to protein hydration. In: Pullman, editor B., editor. Intermolecular Forces: Proceedings of the Fourteenth Jerusalem Symposium on Quantum Chemistry and Biochemistry Held in Jerusalem, Israel, April 13–16, 1981. Springer; Dordrecht, The Netherlands: 1981. pp. 331–342. [Google Scholar]
- 29.Wolf M.G., Hoefling M., Groenhof G. g_membed: efficient insertion of a membrane protein into an equilibrated lipid bilayer with minimal perturbation. J. Comput. Chem. 2010;31:2169–2174. doi: 10.1002/jcc.21507. [DOI] [PubMed] [Google Scholar]
- 30.Lomize M.A., Lomize A.L., Mosberg H.I. OPM: orientations of proteins in membranes database. Bioinformatics. 2006;22:623–625. doi: 10.1093/bioinformatics/btk023. [DOI] [PubMed] [Google Scholar]
- 31.Vargas E., Bezanilla F., Roux B. In search of a consensus model of the resting state of a voltage-sensing domain. Neuron. 2011;72:713–720. doi: 10.1016/j.neuron.2011.09.024. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Webb B., Sali A. Comparative protein structure modeling using MODELLER. Curr. Protoc. Bioinform. 2014;47 doi: 10.1002/0471250953.bi0506s47. 5.6.1–32. [DOI] [PubMed] [Google Scholar]
- 33.Gapsys V., de Groot B.L., Briones R. Computational analysis of local membrane properties. J. Comput. Aided Mol. Des. 2013;27:845–858. doi: 10.1007/s10822-013-9684-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Charloteaux B., Lorin A., Brasseur R. The N-terminal 12 residue long peptide of HIV gp41 is the minimal peptide sufficient to induce significant T-cell-like membrane destabilization in vitro. J. Mol. Biol. 2006;359:597–609. doi: 10.1016/j.jmb.2006.04.018. [DOI] [PubMed] [Google Scholar]
- 35.Verdon G., Oh S., Boudker O. Coupled ion binding and structural transitions along the transport cycle of glutamate transporters. eLife. 2014;3:e02283. doi: 10.7554/eLife.02283. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Tao Z., Rosental N., Grewer C. Mechanism of cation binding to the glutamate transporter EAAC1 probed with mutation of the conserved amino acid residue Thr101. J. Biol. Chem. 2010;285:17725–17733. doi: 10.1074/jbc.M110.121798. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Bastug T., Heinzelmann G., Ryan R.M. Position of the third Na+ site in the aspartate transporter GltPh and the human glutamate transporter, EAAT1. PLoS One. 2012;7:e33058. doi: 10.1371/journal.pone.0033058. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Guskov A., Jensen S., Slotboom D.J. Coupled binding mechanism of three sodium ions and aspartate in the glutamate transporter homologue GltTk. Nat. Commun. 2016;7:13420. doi: 10.1038/ncomms13420. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Tieleman D.P., Berendsen H.J.C. Molecular dynamics simulations of a fully hydrated dipalmitoyl phosphatidylcholine bilayer with different macroscopic boundary conditions and parameters. J. Chem. Phys. 1996;105:4871–4880. [Google Scholar]
- 40.Machtens J.P., Kortzak D., Fahlke C. Mechanisms of anion conduction by coupled glutamate transporters. Cell. 2015;160:542–553. doi: 10.1016/j.cell.2014.12.035. [DOI] [PubMed] [Google Scholar]
- 41.Gurtovenko A.A., Vattulainen I. Calculation of the electrostatic potential of lipid bilayers from molecular dynamics simulations: methodological issues. J. Chem. Phys. 2009;130:215107. doi: 10.1063/1.3148885. [DOI] [PubMed] [Google Scholar]
- 42.Murata Y., Iwasaki H., Okamura Y. Phosphoinositide phosphatase activity coupled to an intrinsic voltage sensor. Nature. 2005;435:1239–1243. doi: 10.1038/nature03650. [DOI] [PubMed] [Google Scholar]
- 43.Vargas E., Yarov-Yarovoy V., Roux B. An emerging consensus on voltage-dependent gating from computational modeling and molecular dynamics simulations. J. Gen. Physiol. 2012;140:587–594. doi: 10.1085/jgp.201210873. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Ishida I.G., Rangel-Yescas G.E., Islas L.D. Voltage-dependent gating and gating charge measurements in the Kv1.2 potassium channel. J. Gen. Physiol. 2015;145:345–358. doi: 10.1085/jgp.201411300. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Villalba-Galea C.A., Sandtner W., Bezanilla F. S4-based voltage sensors have three major conformations. Proc. Natl. Acad. Sci. USA. 2008;105:17600–17607. doi: 10.1073/pnas.0807387105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Minor D.L., Lin Y.F., Berger J.M. The polar T1 interface is linked to conformational changes that open the voltage-gated potassium channel. Cell. 2000;102:657–670. doi: 10.1016/s0092-8674(00)00088-x. [DOI] [PubMed] [Google Scholar]
- 47.Börjesson S.I., Elinder F. Structure, function, and modification of the voltage sensor in voltage-gated ion channels. Cell Biochem. Biophys. 2008;52:149–174. doi: 10.1007/s12013-008-9032-5. [DOI] [PubMed] [Google Scholar]
- 48.Wang G., Covarrubias M. Voltage-dependent gating rearrangements in the intracellular T1-T1 interface of a K+ channel. J. Gen. Physiol. 2006;127:391–400. doi: 10.1085/jgp.200509442. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Zerangue N., Kavanaugh M.P. Flux coupling in a neuronal glutamate transporter. Nature. 1996;383:634–637. doi: 10.1038/383634a0. [DOI] [PubMed] [Google Scholar]
- 50.Focke P.J., Moenne-Loccoz P., Larsson H.P. Opposite movement of the external gate of a glutamate transporter homolog upon binding cotransported sodium compared with substrate. J. Neurosci. 2011;31:6255–6262. doi: 10.1523/JNEUROSCI.6096-10.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Ewers D., Becher T., Fahlke C. Induced fit substrate binding to an archeal glutamate transporter homologue. Proc. Natl. Acad. Sci. USA. 2013;110:12486–12491. doi: 10.1073/pnas.1300772110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Hänelt I., Jensen S., Slotboom D.J. Low affinity and slow Na+ binding precedes high affinity aspartate binding in the secondary-active transporter GltPh. J. Biol. Chem. 2015;290:15962–15972. doi: 10.1074/jbc.M115.656876. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Reyes N., Ginter C., Boudker O. Transport mechanism of a bacterial homologue of glutamate transporters. Nature. 2009;462:880–885. doi: 10.1038/nature08616. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Watzke N., Bamberg E., Grewer C. Early intermediates in the transport cycle of the neuronal excitatory amino acid carrier EAAC1. J. Gen. Physiol. 2001;117:547–562. doi: 10.1085/jgp.117.6.547. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Machtens J.P., Kovermann P., Fahlke C. Substrate-dependent gating of anion channels associated with excitatory amino acid transporter 4. J. Biol. Chem. 2011;286:23780–23788. doi: 10.1074/jbc.M110.207514. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Huang Z., Tajkhorshid E. Identification of the third Na+ site and the sequence of extracellular binding events in the glutamate transporter. Biophys. J. 2010;99:1416–1425. doi: 10.1016/j.bpj.2010.06.052. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Larsson H.P., Tzingounis A.V., Kavanaugh M.P. Fluorometric measurements of conformational changes in glutamate transporters. Proc. Natl. Acad. Sci. USA. 2004;101:3951–3956. doi: 10.1073/pnas.0306737101. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Wadiche J.I., Arriza J.L., Kavanaugh M.P. Kinetics of a human glutamate transporter. Neuron. 1995;14:1019–1027. doi: 10.1016/0896-6273(95)90340-2. [DOI] [PubMed] [Google Scholar]
- 59.Hotzy J., Machtens J.P., Fahlke C. Neutralizing aspartate 83 modifies substrate translocation of excitatory amino acid transporter 3 (EAAT3) glutamate transporters. J. Biol. Chem. 2012;287:20016–20026. doi: 10.1074/jbc.M112.344077. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Villalba-Galea C.A., Frezza L., Bezanilla F. Sensing charges of the Ciona intestinalis voltage-sensing phosphatase. J. Gen. Physiol. 2013;142:543–555. doi: 10.1085/jgp.201310993. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Schewe M., Nematian-Ardestani E., Baukrowitz T. A non-canonical voltage-sensing mechanism controls gating in K2P K+ channels. Cell. 2016;164:937–949. doi: 10.1016/j.cell.2016.02.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.Pusch M., Ludewig U., Jentsch T.J. Gating of the voltage-dependent chloride channel CIC-0 by the permeant anion. Nature. 1995;373:527–531. doi: 10.1038/373527a0. [DOI] [PubMed] [Google Scholar]
- 63.Grieschat M., Alekov A.K. Multiple discrete transitions underlie voltage-dependent activation in CLC Cl−/H+ antiporters. Biophys. J. 2014;107:L13–L15. doi: 10.1016/j.bpj.2014.07.063. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Vickery O.N., Machtens J.P., Zachariae U. Structural mechanisms of voltage sensing in G protein-coupled receptors. Structure. 2016;24:997–1007. doi: 10.1016/j.str.2016.04.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Lísal J., Maduke M. The ClC-0 chloride channel is a ‘broken’ Cl−/H+ antiporter. Nat. Struct. Mol. Biol. 2008;15:805–810. doi: 10.1038/nsmb.1466. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Delemotte L., Dehez F., Tarek M. Modeling membranes under a transmembrane potential. J. Phys. Chem. B. 2008;112:5547–5550. doi: 10.1021/jp710846y. [DOI] [PubMed] [Google Scholar]
- 67.Kutzner C., Páll S., Grubmüller H. Best bang for your buck: GPU nodes for GROMACS biomolecular simulations. J. Comput. Chem. 2015;36:1990–2008. doi: 10.1002/jcc.24030. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Krause D., Thörnig P. JURECA: general-purpose supercomputer at Jülich supercomputing centre. J. Large-Scale Res. Facilities. 2016;2:A62. [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.