Supporting Material for. Microscopic origin of gating current fluctuations in a potassium channel voltage sensor

Similar documents
Molecular Dynamics Investigation of the ω-current in the Kv1.2 Voltage Sensor Domains

Potassium channel gating and structure!

Ion Channel Structure and Function (part 1)

A SURPRISING CLARIFICATION OF THE MECHANISM OF ION-CHANNEL VOLTAGE- GATING

MARTINI simulation details

Improved Resolution of Tertiary Structure Elasticity in Muscle Protein

Routine access to millisecond timescale events with accelerated molecular dynamics

Supporting Information for. Hydrogen Bonding Structure at Zwitterionic. Lipid/Water Interface

Supporting Information Quantitative characterization of the binding and unbinding of millimolar drug fragments with molecular dynamics simulations

T H E J O U R N A L O F G E N E R A L P H Y S I O L O G Y. jgp

Homology models of the tetramerization domain of six eukaryotic voltage-gated potassium channels Kv1.1-Kv1.6

Closing In on the Resting State of the Shaker K + Channel

Equilibrated atomic models of outward-facing P-glycoprotein and effect of ATP binding on structural dynamics (Supplementary Information)

The Potassium Ion Channel: Rahmat Muhammad

Supplementary Figures for Tong et al.: Structure and function of the intracellular region of the plexin-b1 transmembrane receptor

Supporting Information

Optimization of CHARMM36-UA Force Field Parameters for Phospholipid Bilayer Simulation

Dynamic hydration valve controlled ion permeability and

The Molecular Dynamics Simulation Process

Don t forget to bring your MD tutorial. Potential Energy (hyper)surface

Supplementary Figure S1. Urea-mediated buffering mechanism of H. pylori. Gastric urea is funneled to a cytoplasmic urease that is presumably attached

Supplementary Information

The potassium channel: Structure, selectivity and diffusion

SUPPLEMENTARY INFORMATION

A Model of Voltage Gating Developed Using the KvAP Channel Crystal Structure

Particle-Based Simulation of Bio-Electronic Systems

Why Proteins Fold? (Parts of this presentation are based on work of Ashok Kolaskar) CS490B: Introduction to Bioinformatics Mar.

Calibrated Measurement of Gating-Charge Arginine Displacement in the KvAP Voltage-Dependent K + Channel

SUPPLEMENTARY INFORMATION. doi: /nature07461

Comparing crystal structure of M.HhaI with and without DNA1, 2 (PDBID:1hmy and PDBID:2hmy),

The Molecular Dynamics Method

Supporting Information

Supporting Information

Electro-Mechanical Conductance Modulation of a Nanopore Using a Removable Gate

Introduction The gramicidin A (ga) channel forms by head-to-head association of two monomers at their amino termini, one from each bilayer leaflet. Th

The principle of gating charge movement in a voltage-dependent K 1 channel

SUPPLEMENTARY INFORMATION

Structure Investigation of Fam20C, a Golgi Casein Kinase

Membrane Protein Channels

Replica Exchange with Solute Scaling: A More Efficient Version of Replica Exchange with Solute Tempering (REST2)

All-atom Molecular Mechanics. Trent E. Balius AMS 535 / CHE /27/2010

Universal Repulsive Contribution to the. Solvent-Induced Interaction Between Sizable, Curved Hydrophobes: Supporting Information

Electronic Supplementary Information

THE BIT OF information in nerves is the action potential, a

a Brownian model of the voltage sensor

Supporting Information for Solid-liquid Thermal Transport and its Relationship with Wettability and the Interfacial Liquid Structure

The Molecular Dynamics Method

Enhancing drug residence time by shielding of intra-protein hydrogen bonds: a case study on CCR2 antagonists. Supplementary Information

Probing α/β Balances in Modified Amber Force Fields from a Molecular Dynamics Study on a ββα Model Protein (1FSD)

to Sense Voltage? Francesco Tombola, 1,4 Medha M. Pathak, 2 and Ehud Y. Isacoff 1,2,3, * 1 Department of Molecular and Cell Biology

Voltage Dependence of Conformational Dynamics and Subconducting

Potential Energy (hyper)surface

Molecular Dynamics. A very brief introduction

SUPPLEMENTARY INFORMATION

Supplementary Figures:

The Molecular Dynamics Method

Building a Homology Model of the Transmembrane Domain of the Human Glycine α-1 Receptor

Voltage Sensor of Kv1.2: Structural Basis of Electromechanical Coupling

Computational Modeling of Protein Kinase A and Comparison with Nuclear Magnetic Resonance Data

Molecular Basis of K + Conduction and Selectivity

Molecular dynamics simulation of Aquaporin-1. 4 nm

The Computational Microscope

Transporters and Membrane Motors Nov 15, 2007

W tot (z) = W (z) k B T ln e i q iφ mp(z i )/k B T, [1]

Nonselective Conduction in a Mutated NaK Channel with Three Cation-Binding Sites

Supplementary Information for Subdiffusion in Membrane Permeation of Small Molecules

Molecular dynamics simulations of a single stranded (ss) DNA

Supplementary Figure 1 Irregular arrangement of E,E-8-mer on TMA. STM height images formed when

Computer simulation methods (2) Dr. Vania Calandrini

Ion Selectivity in the KcsA Potassium Channel from the Perspective of the Ion Binding Site

Tailoring the Properties of Quadruplex Nucleobases for Biological and Nanomaterial Applications

Hands-on Course in Computational Structural Biology and Molecular Simulation BIOP590C/MCB590C. Course Details

Electronic Supplementary Information

STRUCTURE OF IONS AND WATER AROUND A POLYELECTROLYTE IN A POLARIZABLE NANOPORE

Mechanical Proteins. Stretching imunoglobulin and fibronectin. domains of the muscle protein titin. Adhesion Proteins of the Immune System

An introduction to Molecular Dynamics. EMBO, June 2016

SUPPLEMENTARY INFORMATION

This article appeared in a journal published by Elsevier. The attached copy is furnished to the author for internal non-commercial research and

Secondary Structure. Bioch/BIMS 503 Lecture 2. Structure and Function of Proteins. Further Reading. Φ, Ψ angles alone determine protein structure

K versus Na Ions in a K Channel Selectivity Filter: A Simulation Study

Module Membrane Biogenesis and Transport Lecture 15 Ion Channels Dale Sanders

State Key Laboratory of Molecular Reaction Dynamics, Dalian Institute of Chemical Physics, Chinese Academic of Sciences, Dalian , P. R. China.

A Theoretical Model for Calculating Voltage Sensitivity of Ion Channels and the Application on Kv1.2 Potassium Channel

Type II Kinase Inhibitors Show an Unexpected Inhibition Mode against Parkinson s Disease-Linked LRRK2 Mutant G2019S.

Computational Molecular Biophysics. Computational Biophysics, GRS Jülich SS 2013

Peptide folding in non-aqueous environments investigated with molecular dynamics simulations Soto Becerra, Patricia

Towards accurate calculations of Zn 2+ binding free energies in zinc finger proteins

Biophysics 490M Project

arxiv: v1 [physics.chem-ph] 30 Jul 2012

Density Functional Theory: from theory to Applications

NIH Public Access Author Manuscript Nature. Author manuscript; available in PMC 2009 December 18.

Neurobiology Biomed 509 Ion Channels

Computational Predictions of 1-Octanol/Water Partition Coefficient for Imidazolium based Ionic Liquids.

Hyeyoung Shin a, Tod A. Pascal ab, William A. Goddard III abc*, and Hyungjun Kim a* Korea

The Selectivity of K 1 Ion Channels: Testing the Hypotheses

SUPPLEMENTARY INFORMATION

Molecular Dynamics Studies of Human β-glucuronidase

2008 Biowerkzeug Ltd.

Supporting Information Local Density Fluctuations Predict Photoisomerization Quantum Yield of Azobenzene-Modified DNA

Folding of small proteins using a single continuous potential

Transcription:

Supporting Material for Microscopic origin of gating current fluctuations in a potassium channel voltage sensor J. Alfredo Freites, * Eric V. Schow, * Stephen H. White, and Douglas J. Tobias * * Department of Chemistry and Department of Physiology and Biophysics, University of California, Irvine, CA 92697, USA

Figure S1. Fractional membrane potential (Φ m ) calculated along a transmembrane axis passing through the center of the the KvAP VSD. The linear potential drop extends of ~23 Å. This profile is taken from a linearized Poisson-Boltzmann calculation reported in (1).

Figure S2. Conformational dynamics of the S3b-S4 paddle motif, as the C α rmsd from the VSD initial configuration with respect to S1-S2, along the simulation trajectory. The broken red line indentifies the switch from the depolarizing potential (+120 mv) to the hyperpolarizing potential (-120 mv).

Figure S3. Lipid bilayer thickness measured as mean distance between lipid phosphate groups in opposite leaflets. Values are shown for all lipids and those in the first VSD solvation shell (cutoff of 4.5 Å) for the unpolarized simulation (Fig. S5), and the simulation under depolarizing and hyperpolarizing potentials (Fig. 1). No significant differences due to the action of the applied membrane potention are observed. Error bars indicate the statistical uncertainty at the 95% confidence level.

Figure S4. Interpretation of the KvAP biotin-avidin accessibility data for the S3b-S4 paddle motif (2, 3) in terms of positions of the C atoms along the transmembrane direction measured as mean values over the last 4.5 s of trajectory under a depolarizing potential. The upper red band represents the location of the extracellular interface (as the mean of the lipid carbonyl distribution) and the lower band represents the effective length of the biotin tether (10 Å). Residues whose C atoms fall within the two bands or over the upper band are considered to be accessible from the extracellular side, and residues located under the lower band are considered not to be accessible from the extracellular side under a depolarizing potential. The simulation data is in complete agreement with the experimental results reported in (2, 3).

Figure S5. VSD configuration snapshot from a 5 μs unpolarized control simulation (after 4.2 μs). Molecular representations and color scheme are as in Fig. 1. The only significant change with respect to the initial configuration (reported in (4)) is the conversion of the R6-D62 salt-bridge to a water-mediated interaction and the formation of a watermediated interaction between R6 and E93.

Figure S6. Per-residue contributions to the gating charge. The gating charge associated to each VSD basic or acidic side chain was calculated as difference between the average charge displacement over the last 3 μs of the depolarized trajectory and the average charge displacement 16 μs of the hyperpolarized trajectory. Conserved positions for acidic side chains in S1-S3 and for basic side chains in S4 and the S4-S5 linker are labeled as in Fig. 1. Also indicated is E107 a side chain in S3b that forms salt-bridges with R3 and R4.

Figure S7. Evolution of the charge displacement and solvation environment for R4 (top two graphs) and K7 (bottom two graphs). Similar to R6, the contributions to the total charge displacement for R4 and K7 follow closely the changes in their solvation environments. R4, besides forming salt-bridges and water-mediated interactions with conserved acidic side chains (E45 and D62), also forms a flickering salt-bridge with E107 (in S3b) which is not part of the set of S1-S3 conserved positions for acidic side chains in Kv channels.

Figure S8. Comparison of the KvAP VSD structure (orange; residues 24 to 147 of the KvAP full channel model proposed by Lee et al. (5)) with the X-ray crystal structure of the Kv1.2 paddle-chimera channel VSD (purple; residues 162 to 322) (6). The superposition was performed by minimizing the backbone rmsd of the S1-S2 transmembrane segments. The difference in length of the S4 segments in both structures causes an apparent relative shift on the spatial location of the S4 basic side chains. Thus, the R6 position in KvAP and the R4 position in Kv1.2 share a common location close to the VSD constriction region separating the extracellular and intracellular crevices. Note that the S1-S2 alignment places the S4-S5 linker C-terminal backbones at similar locations along the transmembrane direction. KvAP side chains are colored by atom type (cf. Fig. 1) and those from Kv1.2 are colored by residue type (basic, blue; acidic, red). KvAP residue numbers are indicated in plain font, and Kv1.2 residue numbers are in italics. The Kv1.2 S1-S2 loop was omitted for clarity.

Figure S9. VSD configuration snapshot from the end of a 14 μs trajectory under a -120 mv hyperpolarizing potential started from an unpolarized configuration. In contrast to the configuration of the charged side chains at the end of the hyperpolarized trajectory started from a depolarized configuration (Fig. 1C), the R6 center of charge remains at center of the VSD (as in the unpolarized configuration, cf. Fig. S5) and K7 is exposed to the VSD intracellular crevice. Molecular representations and color scheme are as in Fig. 1.

Methods Simulation system. The initial atomistic model of the unpolarized KvAP VSD corresponds to residues 24 to 147 of the KvAP full channel model proposed by Lee et al. (5). The initial simulation system configuration was derived from the end of a previously equilibrated, unpolarized simulation of the KvAP VSD in a POPC bilayer in excess water (4). The system consisted of 10,048 waters, 232 lipids, 2 chloride counterions, and 1 VSD, for a total of 63,247 atoms. Molecular dynamics simulations. The μs-timescale simulations were performed on Anton, a special-purpose computer for molecular dynamics simulations of biomolecules (7). The system was equilibrated for 12 ns using the Desmond Molecular Dynamics System, version 2.4 (D. E. Shaw Research, New York, NY, 2008) on a conventional high performance cluster before being transferred to Anton. The CHARMM22 (8) and CHARMM36 (9) force fields were used for the protein and lipids, respectively, and the TIP3P model was used for water. A reversible multiple-timestep algorithm was employed to integrate the equations of motion with a time step of 6 fs for the long-range nonbonded forces, and 2 fs for short-range non- bonded and bonded forces. All bond lengths involving hydrogen atoms were held fixed using the SHAKE algorithm. The k-space Gaussian split Ewald method (10) with a 32 x 32 x 32 grid was used to calculate longrange electrostatic interactions. A cutoff of 14 Å was used for the Lennard-Jones and short-range electrostatic interactions. The simulations were performed at constant temperature (300 K) and pressure (1 atm), using a Berendsen thermostat and semiisotropic Berendsen barostat. Analyses and visualization were performed with VMD 1.9 (11). Membrane potential and gating charge. To model the applied membrane potential, we imposed a constant electric field with a magnitude equal to 120 mv/23 Å. The specific value for the thickness of the dielectric barrier was taken from a linearized Poisson- Boltzmann calculation of the electrostatic potential along a transmembrane axis passing through the center of the VSD (see Fig. S1), as previously described (1). The total charge displacement with respect to the initial configuration (Δq) was calculated as q q i m [z i (t)] m [z i (0)] i where t is time, Φ m is the fractional membrane potential shown in Fig. S1, and z i and q i are the position along the transmembrane direction and the partial charge of the i th atom, respectively. The sum was taken over the partial charges of the atoms that constitute the centers of charge of all the VSD basic and acidic side chains. The gating-charge displacement (Q) between the depolarized and hyperpolarized trajectories was calculated as Q q d q h

where < Δq > d is the average charge displacement over the last 3 μs of the depolarized trajectory and < Δq > h is the average charge displacement over the last 16 μs of the hyperpolarized trajectory. References 1. Krepkiy, D., M. Mihailescu, J. A. Freites, E. V. Schow, D. L. Worcester, K. Gawrisch, D. J. Tobias, S. H. White, and K. J. Swartz. 2009. Structure and hydration of membranes embedded with voltage-sensing domains. Nature 462:473-479. 2. Jiang, Y. X., V. Ruta, J. Y. Chen, A. Lee, and R. MacKinnon. 2003. The principle of gating charge movement in a voltage-dependent K + channel. Nature 423:42-48. 3. Ruta, V., J. Chen, and R. MacKinnon. 2005. Calibrated measurement of gating-charge arginine. Cell 123:463-475. 4. Freites, J. A., D. J. Tobias, and S. H. White. 2006. A voltage-sensor water pore. Biophys. J. 91:L90-L92. 5. Lee, S.-Y., A. Lee, J. Chen, and R. MacKinnon. 2005. Structure of the KvAP voltagedependent K + channel and its dependence on the lipid membrane. Proc. Natl. Acad. Sci. U.S.A. 102:15441-15446. 6. Long, S. B., X. Tao, E. B. Campbell, and R. Mackinnon. 2007. Atomic structure of a voltage-dependent K + channel in a lipid membrane-like environment. Nature 450:376-382. 7. Shaw, D. E., M. M. Deneroff, R. O. Dror, J. S. Kuskin, R. H. Larson, J. K. Salmon, C. Young, B. Batson, K. J. Bowers, J. C. Chao, M. P. Eastwood, J. Gagliardo, J. P. Grossman, C. R. Ho, D. J. Lerardi, I. Kolossváry, J. L. Klepeis, T. Layman, C. McLeavey, M. A. Moraes, R. Mueller, E. C. Priest, Y. Shan, J. Spengler, M. Theobald, B. Towles, and S. C. Wang. 2007. Anton, a special-purpose machine for molecular dynamics simulation. Proc. 34th Annu. Internat. Sym. Computer Architect. 8. MacKerell, A. D., Jr., M. Feig, and C. L. Brooks, II. 2004. Extending the treatment of backbone energetics in protein force fields: Limitations of gas-phase quantum mechanics in reproducing conformational distributions in molecular dynamics simulations. J. Comput. Chem. 25:1400-1415. 9. Klauda, J. B., R. M. Venable, J. A. Freites, J. W. O'Connor, D. J. Tobias, C. Mondragon-Ramirez, I. Vorobyov, A. D. MacKerell, Jr., and R. W. Pastor. 2010. Update of the CHARMM all-atom additive force field for lipids: validation on six lipid types. J. Phys. Chem. B 114:7830-7843. 10. Shan, Y., J. L. Klepeis, M. P. Eastwood, R. O. Dror, and D. E. Shaw. 2005. Gaussian split Ewald: a fast Ewald mesh method for molecular simulation. J. Chem. Phys. 122:054101-054101 to 054101-054113. 11. Humphrey, W., W. Dalke, and K. Schulten. 1996. VMD: Visual molecular dynamics. J. Mol. Graph. 14:33-38.