Open Access Article
Christopher
Päslack
a,
Lars V.
Schäfer
*a and
Matthias
Heyden
*b
aCenter for Theoretical Chemistry, Faculty of Chemistry and Biochemistry, Ruhr-University Bochum, D-44780 Bochum, Germany. E-mail: lars.schaefer@ruhr-uni-bochum.de; Fax: +49 234 3214045; Tel: +49 234 3221582
bSchool of Molecular Sciences, Arizona State University, Tempe, AZ 85287-1604, USA. E-mail: heyden@asu.edu; Tel: +1 480 965-3980
First published on 15th April 2019
Correlated vibrational motion on the sub-picosecond timescale and associated collective dynamics in a protein–membrane environment are characterized using molecular dynamics simulations. We specifically analyze correlated motion of a membrane-associated protein and a lipid bilayer for distinct separation distances. Correlated vibrations persist up to distances of 25 Å between both biomolecular surfaces. These correlations are mediated by separating layers of water molecules, whose collective properties are altered by the simultaneous presence of protein and lipid bilayer interfaces.
Collective motions in proteins cover a broad range of timescales. They were first investigated by normal mode analysis and linked to low-frequency modes in the density of states of proteins via incoherent neutron scattering.20,21 A study by Hong et al. revealed how specific, collective in-phase dynamic modes can facilitate access for substrates to the binding site of cytochrome P450cam.22 Further, simulations have characterized collective modes that propagate from the surface of a protein into its hydration shell.23–27 Despite this body of knowledge on protein solutions28–30 and the biological significance of membranes,31–33 collective protein–lipid dynamics – and how they are mediated by hydration water – have not yet been investigated.
Here, we characterized correlated low-frequency vibrations between the membrane-associated protein annexin B12 (Anx) and a lipid bilayer consisting of a 7
:
3 mixture of DOPC and DOPS lipids (Fig. 1). We further analyzed correlated vibrations of protein and lipid atoms with water separating both biomolecular surfaces. Atomic vibrations were sampled in all-atom molecular dynamics (MD) simulations and their correlations were analyzed for membrane-bound and unbound states. This analysis revealed the persistence of correlated vibrations between atoms of the Anx protein and membrane lipids up to 25 Å separation distances, which are mediated by the separating shell of hydration water. This finding is supported by the observation of simultaneous modifications of collective protein–water and lipid–water dynamics, which describe the propagation of collective modes from the protein and lipid bilayer surfaces into water layers separating both. A complementary analysis in the time-domain provides an estimate for the exchange of information with respect to the average thermal energy via the collective modes described here.
:
3 mixture of DOPC and DOPS lipids with a total of 512 lipids (256 lipids in each leaflet). We started from an X-ray crystal structure of Anx (PDB ID code 1DM5)34 and from a pre-equilibrated lipid bilayer.35 The simulated systems were prepared following a protocol described in previous work,11 where the initial position of the protein bound to the lipid bilayer was chosen according to experimental data.36
The protein and bilayer were placed in a rectangular box with a size of approximately 135 × 135 × 157 Å3 containing ca. 64
000 water molecules. The net charge of the system was neutralized with Na+ and Cl− ions. Periodic boundary conditions were applied in all three dimensions. The MD simulations were carried out using GROMACS 5.0.637 with the Amber ff99SB*-ILDNP force-field38–41 for Anx and the all-atom Slipids force-field35,42 for the lipids. The TIP4P/2005 model43 was used for water.
We employed the SETTLE and LINCS algorithms44,45 to constrain the internal degrees of freedom of water molecules and the bonds in other molecules, respectively, allowing to integrate the equations of motion with time steps of 2 fs. Short-range non-bonded Coulomb and Lennard-Jones 6,12 interactions were treated with a Verlet buffered pair list46 with potentials smoothly shifted to zero at a 10 Å cut-off. Long-range Coulomb interactions were treated using the smooth particle mesh Ewald (PME) scheme47,48 with a grid spacing of 1.2 Å and cubic spline interpolation. Analytical dispersion corrections were applied for energy and pressure to account for the truncation of the Lennard-Jones interactions.
After an initial energy minimization (500 steps steepest decent) of the Anx protein and the DOPC/DOPS bilayer, the protein and two bound Ca2+ ions were translated along the z-axis (i.e., the direction normal to the membrane) to generate configurations in which Anx is located at various distances R to the bilayer, covering an R-range from 0 to 32 Å. Each of the systems was then solvated with water. Sodium and chloride ions were included at physiological concentrations via substitution of randomly selected water molecules. Care was taken not to introduce water molecules in the hydrophobic part of the bilayer. After a second energy minimization (500 steps steepest decent), the systems were equilibrated for 5 ns in the isothermal–isobaric (NPT) ensemble with harmonic position restraining potentials applied to the heavy atoms of the protein (force constants 1000 kJ mol−1 nm−2). The temperature was kept at 298 K by a velocity-rescaling thermostat49 with a coupling time constant of 0.1 ps. For simulations at constant 1.0 bar pressure, a semi-isotropic Berendsen barostat50 with a coupling time constant of 0.5 ps and a compressibility of 4.5 × 10−5 bar−1 was applied, with lateral (xy) and normal (z) dimensions of the simulation box coupled separately.
For each configuration generated with a specific protein–membrane distance R, simulations of 100 ns were performed in the NPT ensemble. During the equilibrations (but not the production simulations, see below), two harmonic restraining potentials were applied to the angle between the z-axis and the connecting vectors between the Cα-atoms of (1) residues Ser35/Val195 and (2) residues Glu138/His254 (force constant 200 kJ mol−1 rad−2). This was required to avoid overall tumbling of the protein during the simulations. In addition, a harmonic umbrella potential with a force constant of 1500 kJ mol−1 nm−2 was applied in the z-direction between the centers of mass of Anx and selected phosphorus atoms in the lipid head groups of the upper leaflet. The phosphorus atoms were selected in a cylindrical geometry with an inner radius of r0 = 30 Å and an outer radius of r1 = 35 Å, where weighted contributions to the center of mass position were switched to 0 between r0 and r1. The harmonic restraint between the centers of mass of the protein and the phosphorus atoms in the defined cylindrical region kept the protein at the selected distance from the lipid bilayer.
After the 100 ns equilibrations, the NPT equilibrations were extended by another 5 ns for each distance with a stronger angle restraint (1000 kJ mol−1 rad−2) to enforce an ideal orientation of the protein, such that the bottom surface of Anx was parallel to the membrane surface (Fig. 1).
Finally, for selected distances (R = 0 to 30 Å, in 5 Å steps, and two additional distances of 28 Å and 32 Å), we carried out 20 independent 100 ps simulations in the NVE ensemble for analysis of collective motions. The starting structures for the NVE simulations were taken from the final 2 ns of the above NPT equilibrations, separated by 100 ps. Atom positions and velocities were recorded every 8 fs for vibrational analysis. The production simulations in the NVE ensemble were entirely unbiased, i.e., without umbrella potentials or angle restraints, to avoid artifacts in the time-correlation functions and their Fourier transforms due to the restraining potentials. The mass density profiles along the membrane normal (Fig. S1), as well as the analysis of the stability of protein orientation (Fig. S2) and position (Fig. S3 and S4) with respect to the bilayer are provided in the ESI.†
![]() | (1) |
are weighted by the square-root of the atom mass yielding an expression for the average thermal energy for auto-correlations i = j. The brackets 〈…〉τ denote ensemble-averaged time-correlation functions with reference times τ. Eqn (1) allows for direct evaluation of correlated vibrational motion between protein and membrane surface atoms. Here, we restrict the analysis to non-hydrogen atoms, which dominate the vibrational spectra at far-infrared frequencies below 400 cm−1.
In addition to correlated protein–membrane vibrations, we analyzed protein–water and membrane–water correlated vibrations involving water molecules located in the space between Anx and the lipid bilayer. We computed cross-correlation spectra including water oxygens sampled at various distances to the protein or membrane surface, ranging from 2.5–10.0 Å with increments of 0.5 Å. Explicit atomic velocities were replaced by localized velocity densities,
![]() | (2) |
![]() | (3) |
We visualize correlation spectra obtained as a function of the real-space distance using the inverse distance k = 2π/rS,OW, in analogy to coherent scattering data. We identify dispersive collective modes propagating with a wave velocity vmode = dω/dk. The k-resolved spectra, recorded for distinct protein–membrane separations R, are directly related to longitudinal density current spectra I‖J(k,ω) = (ω2/k2)S(k,ω), which describe correlated density fluctuations via the experimentally accessible dynamic structure factor S(k,ω).
The results shown in Fig. 2A report on correlated vibrations between atoms of the protein and membrane surfaces. To investigate the propagation of collective modes, i.e., correlated vibrational motion between the lipid and protein surfaces analyzed in Fig. 2A into the interior of the protein, we repeated the analysis using mass-weighted atomic velocities of atoms selected within a hemisphere (oriented towards the membrane-facing protein surface) of 10 Å radius around the protein center of mass. The results are shown in Fig. 2B and indicate a decreased overall intensity of the correlations, as expected due to the increased separation distance between the membrane surface and the protein core in comparison to the separation between the binding interfaces. In addition, Fig. 2B shows that the correlation intensities decrease significantly faster with increasing R. Relative to the cross-correlation spectrum for R = 0 Å, both, the positive and negative peak intensities (apart from the dispersive shift in frequency) decrease to just 40% of their initial intensity for R = 10 Å in Fig. 2B. When intensities in the cross-correlation spectra are compared between membrane and protein surface atoms for R = 0 Å and R = 10 Å in Fig. 2A, the positive peak decreases only to approximately 50% of its original intensity, while the negative intensity even retains roughly 80%. Correlated vibrations between the protein core and the membrane surface are therefore more sensitive to the separation between both binding interfaces. Hence, collective modes that propagate through the HB network of water separating the protein and membrane surfaces seem to dissipate more efficiently within the protein.
![]() | ||
| Fig. 3 Cross-correlation spectra of mass-weighted atomic velocities and localized velocity densities (eqn (3)) for vibrations of non-hydrogen protein or membrane surface atoms with hydration water oxygen atoms as a function of reciprocal distance k = 2π/rS,OW. Shown are results for Anx in water (left panel; without a lipid bilayer) and for a DOPC/DOPS bilayer (right panel; without a protein). Crosses trace the peak-intensity of the negative-intensity modes for which linear dispersion curves (dashed lines) were determined. | ||
![]() | ||
| Fig. 4 Wave propagation velocities of longitudinal, collective protein–water (red) and membrane–water (gray) modes for distinct protein–membrane separation distances R (see Fig. S6 in the ESI† for corresponding cross-correlation spectra). Error bars show standard errors of the mean from 20 individual simulations. Dashed lines indicate reference values for Anx in water (no membrane) and the lipid bilayer in water (no protein); insets visualize the onset of overlapping protein and membrane hydration layers of 10 Å thickness. | ||
The maximum change Δvmode is about +927 m s−1 for the lipid bilayer and +307 m s−1 for Anx. The propagation of collective modes involving the lipid bilayer and its hydration water becomes similar to fast sound in bulk water (≈3000 m s−1).52,53 However, we do not observe a strictly monotonous trend for neither protein nor lipid bilayer besides the increase for distances R < 25 Å, which coincides with the gradually increasing overlap of the protein and membrane hydration shells. The inset of Fig. 4 shows that the protein surface is not planar and its hydration shell has a conical shape. Assuming a 10 Å thickness of the hydration layer, there is almost no overlap of the protein and membrane hydration shells at a separation distance of R = 30 Å. However, at 25 Å distance the hydration shells begin to partially overlap. When the protein binds to the membrane, i.e., for short distances, the protein and lipids share one single hydration shell. Single particle dynamics, such as water self-diffusion and rotational relaxation, are slowed down in the hydration shell and the hydrogen bond network becomes effectively more rigid.6,11,54,55 This is correlated to changes in local thermodynamic properties as observed in our previous work,11 but the effects are short-ranged and mainly affect a single hydration layer within distances of 3–5 Å. However, here we analyze collective modes involving protein and membrane surface atoms and surrounding water molecules, which tend to extend over distances of at least 10 Å as observed in Fig. 3 and in previous studies on protein–water systems.24,26
In addition to these previous observations, we observe that the presence of two biomolecular surfaces with separation distances of 25 Å or less modifies collective protein–water and membrane–water dynamics and results in the onset of correlated motion of protein and lipid atoms. Our observations can be compared to abrupt dynamical transitions in water hydrogen bond network dynamics for crowded protein solutions observed experimentally for protein–protein distances of 30–40 Å and 20–25 Å in accompanying simulations.56 Indeed, the increased propagation velocity of protein–water and membrane–water collective modes for separation distances R < 25 Å may serve as a sensitive indicator of long-lived water hydrogen bond networks connecting both biomolecular surfaces, which are likely to feature collective properties similar to low-temperature water or ice.57,58
To support this assumption, the lifetimes of water–water hydrogen bonds involving the water molecules between the protein and the membrane (see Fig. 1B, dark blue) were analyzed. As described in the ESI,† we computed the hydrogen bond correlation functions CHB(t) and estimated the HB lifetimes τHB as CHB(τHB) = 1/e. The average lifetimes for all water molecules between the protein and the membrane are shown in Fig. 5A.
For short protein–membrane distances R ≤ 10 Å, the average lifetimes deviate strongly from the bulk lifetime of about 3 ps. Water molecules are confined and H-bonds persist on average up to almost 20 ps, i.e., almost one order of magnitude longer than in bulk. For increasing distances R, the correlation times drop to almost the bulk lifetime, however, not reaching the bulk value since water molecules in the first hydration shells are still included in the average. These findings agree with previous results from two-dimensional infrared spectroscopy and MD simulations,56 where a dynamic transition of HB lifetimes of water molecules between proteins in crowded solutions occurred for comparable separation distances between protein surfaces. In addition, we analyzed the average HB lifetimes involving water molecules selected as a function of distance from either the protein or membrane surface (Fig. 5B, top and bottom panels, respectively); these analyses were carried out for simulations with varying protein–membrane distances R. Notably, only at protein–membrane distances of R ≥ 25 Å, i.e., in the center between protein and membrane, the dynamics of water become similar to bulk, marking the onset of overlapping solvation shells. These results corroborate the interpretation of shifted mode velocities vmode (Fig. 4) due to hydration shell overlap. Notably, collective modes with a propagation velocity of 2500 m s−1 traverse a 25 Å separation distance between the protein and the membrane in just 1 ps.
The observation that correlated vibrations between the protein and membrane persist up to 25 Å separation distance may have important implications for crowded environments such as the cytoplasm. The wave velocities of protein–water and membrane–water collective modes are affected by the combined influence of both biomolecular surfaces. Importantly, the presence of a protein substantially increases the propagation velocities obtained from membrane–water cross-correlation spectra, in this respect resembling fast sound-like dynamics as observed in solid forms of water with slow structural relaxation times. However, the propagation velocities of protein–water collective vibrations are enhanced to a much smaller extent upon hydration shell overlap. While the protein–water collective vibrations propagate inherently faster than the membrane–water vibrations in the infinite dilution limit, membrane–water collective vibrations exhibit faster propagation velocities for protein-membrane separation distances R < 25.
![]() | ||
| Fig. 6 Time domain representation of the longitudinal mass-weighted velocity cross-correlation functions (TCCF) between (A) protein–membrane, (B) protein–water, and (C) membrane–water atoms for distinct separation distances. The data correspond to the inverse Fourier transform of the frequency spectra shown in Fig. 2A and 3. The cross-correlation functions are normalized by the value of the corresponding mass-weighted velocity time auto-correlation function at t = 0, which describes the thermal energy per degree of freedom at the simulation temperature. | ||
The peak intensities of the cross-correlation functions indicate that, in this single degree of freedom, correlated particles exchange between 0.25% and 1% of their energy with each other on a 50 fs timescale. Since our simulations are carried out under equilibrium conditions, no net flow of energy takes place. However, the amount of exchanged energy can be interpreted in terms of dynamic information relative to the thermal fluctuations.
Further, we follow the peak intensities of the time-resolved cross-correlation functions for distinct separation distances in Fig. 7, which describe the spatial dissipation of this dynamic information. While the maximum peak intensities observed at the shortest correlation distances, i.e., for particles in direct contact with each other, are only about 1% of the thermal energy, we note that the exchange of dynamic information and its propagation to larger distances is very fast (ca. 10–100 fs), as seen from the position of the peaks in Fig. 6. Therefore, the amount of energy exchanged between particles quickly accumulates on picosecond timescales.
![]() | ||
| Fig. 7 Peak intensity of mass-weighted velocity time cross correlation functions between protein–membrane atoms (black, inset) and protein–water (blue) and membrane–water atoms (red) as a function of separation distance. Numerical values are equivalent to the y-axis scales in Fig. 6 and describe fractions of the average thermal energy per degree of freedom. | ||
Footnote |
| † Electronic supplementary information (ESI) available: Simulation protocol, details on analysis of correlated motions. See DOI: 10.1039/c9cp00725c |
| This journal is © the Owner Societies 2019 |