Promoting effect of ethanol on dewetting transition in the confined region of melittin tetramer

Similar documents
Water structure near single and multi-layer nanoscopic hydrophobic plates of varying separation and interaction potentials

Supplementary Information. Surface Microstructure Engenders Unusual Hydrophobicity in. Phyllosilicates

arxiv: v1 [cond-mat.soft] 25 Sep 2009

Drying and Hydrophobic Collapse of Paraffin Plates

arxiv: v2 [cond-mat.stat-mech] 23 Sep 2009

Hydrophobic hydration from small to large lengthscales: Understanding and manipulating the crossover

Supporting Information

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

Journal of Pharmacology and Experimental Therapy-JPET#172536

Controllable Water Channel Gating of Nanometer Dimensions

MARTINI simulation details

Hydrophobic Aided Replica Exchange: an Efficient Algorithm for Protein Folding in Explicit Solvent

Segue Between Favorable and Unfavorable Solvation

Unit Cell-Level Thickness Control of Single-Crystalline Zinc Oxide Nanosheets Enabled by Electrical Double Layer Confinement

On the calculation of solvation free energy from Kirkwood- Buff integrals: A large scale molecular dynamics study

Fluctuations of Water near Extended Hydrophobic and Hydrophilic Surfaces

Molecular Wire of Urea in Carbon Nanotube: A Molecular. Dynamics Study

Supporting Information

Hydrogen Bond Kinetics in the Solvation Shell of a Polypeptide

Diffusion of Water and Diatomic Oxygen in Poly(3-hexylthiophene) Melt: A Molecular Dynamics Simulation Study

Are Hydrodynamic Interactions Important in the Kinetics of Hydrophobic Collapse?

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

Ion-Gated Gas Separation through Porous Graphene

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

Proceedings of the ASME 2009 International Mechanical Engineering Congress & Exposition

Supporting Information. for. Influence of Cononsolvency on the. Aggregation of Tertiary Butyl Alcohol in. Methanol-Water Mixtures

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

Modeling Hydrophobic Effects at different lengthscales

MD simulation of methane in nanochannels

Nanoscale Dewetting Transition in Protein Complex Folding

Water Molecules in a Carbon Nanotube under an Applied Electric Field at Various Temperatures and Pressures

The micro-properties of [hmpy+] [Tf 2 N-] Ionic liquid: a simulation. study. 1. Introduction

Level-Set Variational Solvation Coupling Solute Molecular Mechanics with Continuum Solvent

Molecular Dynamics Simulation of a Nanoconfined Water Film

Saturated Sodium Chloride Solution under an External Static Electric Field: a Molecular Dynamics Study

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

sensors ISSN by MDPI

Surface Curvature-Induced Directional Movement of Water Droplets

Hydrophobic and Ionic Interactions in Nanosized Water Droplets

Introduction to molecular dynamics

Weighted density functional theory of the solvophobic effect

6 Hydrophobic interactions

Liquid-vapour oscillations of water in hydrophobic nanopores

Supplemntary Infomation: The nanostructure of. a lithium glyme solvate ionic liquid at electrified. interfaces

Molecular modeling. A fragment sequence of 24 residues encompassing the region of interest of WT-

SUPPLEMENTARY INFORMATION An Empirical IR Frequency Map for Ester C=O Stretching Vibrations

MOLECULAR DYNAMICS SIMULATION OF THE STRUCTURE OF C6 ALKANES INTRODUCTION. A. V. Anikeenko, A. V. Kim, and N. N. Medvedev UDC 544.2: 544.

Molecular Dynamics Simulations. Dr. Noelia Faginas Lago Dipartimento di Chimica,Biologia e Biotecnologie Università di Perugia

WATER PERMEATION THROUGH GRAPHENE NANOSLIT BY MOLECULAR DYNAMICS SIMULATION

Distinct transport properties of O 2 and CH 4 across a carbon nanotube

Supporting Information

Molecular dynamics simulation of Aquaporin-1. 4 nm

(3) E LJ (r ab ) = 4ɛ r ab

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

Supporting information

Dynamics of Water Confined in the Interdomain Region of a Multidomain Protein

The Molecular Dynamics Method

Structure and Interaction in Aqueous Urea-Trimethylamine-N-oxide Solutions

Supporting Information

Can a continuum solvent model reproduce the free energy landscape of a β-hairpin folding in water?

The solute water interactions that determine the solubility

Effect of Surfactant Shape on Solvophobicity and Surface Activity in Alcohol-Water Systems

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

arxiv: v2 [physics.chem-ph] 23 Aug 2013

Water and Aqueous Solutions. 2. Solvation and Hydrophobicity. Solvation

Supporting Information: Improved Parametrization. of Lithium, Sodium, Potassium, and Magnesium ions. for All-Atom Molecular Dynamics Simulations of

Atomistic nature of NaCl nucleation at the solid-liquid interface

Temperature and Pressure Dependence of the Interfacial Free Energy against a Hard. Surface in Contact with Water and Decane. J.

Crystal nucleation and growth both from melts and

Water Dynamics at Interfaces and Solutes: Disentangling Free Energy and Diffusivity Contributions

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

Supplementary Information for: Controlling Cellular Uptake of Nanoparticles with ph-sensitive Polymers

Interfaces and the Driving Force of Hydrophobic Assembly

Supporting information

Lower critical solution temperature (LCST) phase. behaviour of an ionic liquid and its control by. supramolecular host-guest interactions

Supporting Information for: Physics Behind the Water Transport through. Nanoporous Graphene and Boron Nitride

Equilibrium Study of Protein Denaturation by Urea

On the accurate calculation of the dielectric constant and the diffusion coefficient from molecular dynamics simulations: the case of SPC/E water

arxiv: v1 [q-bio.bm] 9 Jul 2007

The dependence of hydrophobic interactions on the solute size

Correlation between local structure and dynamic heterogeneity in a metallic glass-forming liquid

Structural and mechanistic insight into the substrate. binding from the conformational dynamics in apo. and substrate-bound DapE enzyme

Exploring the Changes in the Structure of α-helical Peptides Adsorbed onto Carbon and Boron Nitride based Nanomaterials

arxiv: v1 [physics.chem-ph] 8 Mar 2010

Molecular dynamics simulations of anti-aggregation effect of ibuprofen. Wenling E. Chang, Takako Takeda, E. Prabhu Raman, and Dmitri Klimov

Molecular Dynamics Simulation of Methanol-Water Mixture

Choosing weights for simulated tempering

Interface Resistance and Thermal Transport in Nano-Confined Liquids

Molecular dynamics study of the lifetime of nanobubbles on the substrate

Characterizing hydrophobicity of interfaces by using cavity formation, solute binding, and water correlations

Bioengineering 215. An Introduction to Molecular Dynamics for Biomolecules

Ionic Behavior in Highly Concentrated Aqueous Solutions Nanoconfined between Discretely Charged Silicon Surfaces

Statistical Theory and Learning from Molecular Simulations

Supporting Information

Molecular Dynamics. A very brief introduction

Filling and emptying kinetics of carbon nanotubes in water

Free energy, electrostatics, and the hydrophobic effect

MOLECULAR DYNAMICS STUDIES OF THE STABILITY OF CO 2 AND CH 4 HYDRATES IN THE PRESENCE OF UNDERSATURATED FLUIDS

Urea Impedes the Hydrophobic Collapse of Partially Unfolded Proteins

SUPPLEMENTARY INFORMATION. Very Low Temperature Membrane Free Desalination. by Directional Solvent Extraction

Transcription:

Nuclear Science and Techniques 23 (2012) 252 256 Promoting effect of ethanol on dewetting transition in the confined region of melittin tetramer REN Xiuping 1,2 ZHOU Bo 1,2 WANG Chunlei 1 1 Shanghai Institute of Applied Physics, Chinese Academy of Sciences, Shanghai 201800, China 2 Graduate School of the Chinese Academy of Sciences, Beijing 100080, China Abstract To study the influence of ethanol molecules on the melittin tetramer folding, we investigated the dewetting transition of the melittin tetramer immersed in pure water and 8% aqueous ethanol solution (mass fraction) by the molecular dynamics simulations. We found that the marked dewetting transitions occurred inside a nanoscale channel of the melittin tetramer both in pure water and in aqueous ethanol solution. Also, ethanol molecules promoted this dewetting transition. We attributed this promoting effect to ethanol molecules which prefer to locate at the liquidvapor interface and decrease the liquid-vapor surface energy. The results provide insight into the effect of ethanol on the water dewetting phenomena. Key words Dewetting transition, Melittin tetramer, Ethanol, Molecular dynamics simulation 1 Introduction Hydrophobic interactions play an important role in many physico-chemical processes, such as protein folding [1], micelle formation [2,3], water permeation in membrane channels [4,5] and separation of hydrophobic particles [6 8]. In 1973, Stillinger F. H. [9] argued that water molecules at ambient conditions did not wet a hard wall, and formed a water-vapor interface near the wall. Later, many groups studied this phenomenon using computer simulations and theoretical analysis [10 23]. It was found that large hydrophobic solutes or surfaces can cause reorganizations of water molecules, and there is a nanoscale dewetting transition, when the hydrophobic interaction is extremely strong [24 26]. This happens when two nanoscale hydrophobic plates approach each other and reach a critical distance which is still large enough to accommodate several layers of water molecules, so the water trapped between the two hydrophobic surfaces evaporates spontaneously in a very short period of time before the collapse of the two plates. This transition occurring on a microscopic length scale is analogous to a first-order phase transition from liquid to vapor. Proteins are * Corresponding author. E-mail address: wangchunlei@sinap.ac.cn Received date: 2012-02-25 much more complicated than purely hydrophobic systems by exerting electrostatic forces on water [27 29]. Because of the electrostatic forces, there is no strong dewetting transition observed in the collapse of the BphC enzyme [30]. However, Berne et al. [13,31] studied the collapse of the melittin tetramer in water by computer simulations, and observed a marked water dewetting transition inside a nanoscale channel of the tetramer with a channel size of up to two or three water- molecule diameters. The presence of small solute molecules, such as denaturants and alcohols, may greatly affect the hydrophobic/hydrophilic interaction in aqueous solutions, and thus may influence the thermal stability or solubility of proteins [32 34]. The mechanism of how cosolvent molecules affect hydrophobic/hydrophilic interactions is a basic issue in the protein chemistry field, but remains a matter of some controversy [23,35,36]. Thus, examining the impact of these solutes on the dewetting transition phenomenon can provide a new perspective regarding the microscopic mechanisms [36]. In order to study the influence of ethanol molecules on the melittin tetramer folding, we performed molecular dynamics simulations on the

REN Xiuping et al. / Nuclear Science and Techniques 23 (2012) 252 256 dewetting transition of the melittin tetramer immersed in pure water and aqueous ethanol solutions, and found marked dewetting transitions inside a nanoscale channel of the melittin tetramer in both systems. It was also found that the ethanol molecules promoted the dewetting transition. The vapor phase is preferred because ethanol molecules locate at the liquid-vapor interface, and decrease the liquid-vapor surface energy. We believe that the findings will be of help for understanding the dewetting behavior of the mixed solution in the nanoscale space, including that in the nano-devices and biosystems. 2 Simulation method Protein melittin with a 26-residue polypeptide is a small toxic protein in honeybee venom, and often self-assembles into a tetramer, as shown in Fig.1. The starting structure of the melittin tetramer was taken from the crystal structure deposited in PDB[37]. The two dimers of the tetramer were separated by a distance D, ranging from 0.55 to 0.75 nm to create a nanoscale channel, and were solvated in pure water or 8% aqueous ethanol solution. The channel was filled with water molecules or ethanol molecules, and protein atoms were constrained with a harmonic potential with a force constant of 1000 kj mol 1 nm 2. The simulation process of the melittin tetramer was the same as the dewetting simulation used in the previous work of Berne et al[31]. Twenty Cl counterions were added to all solvated water or mixture boxes to make the system electrically neutral (Fig.1). Fig.1 Schematics of imulation system. Melittin tetramer, Cl, water and ethanol are shown as ribbons, and van der Waals spheres, thin lines and thick bonds, respectively. Parameters for ethanol molecules were taken 253 from the all-atom OPLS (optimized potentials for liquid simulations) force field[38]. The rigid potential SPC/E (extended simple point charge) was used for water[39], and the SETTLE algorithm was used to keep the O H distance fixed at 0.1 nm and the H O H angle at 109.47º[40]. The GROMACS[41] MD package was used for all simulations with a time step of 1 fs. Geometric combining rules were applied to calculate the Lennard Jones interactions between different particles: εij=(εiiεjj)1/2 and σij=(σiiσjj)1/2, where εii and σii are the parameters of atom (i) for the Lennard Jones diameter and well depth, respectively. The systems were run at a constant pressure and temperature (NPT) of 105 Pa and 298 K for 5 ns, after the energy minimization with a steepest-descent algorithm. The NPT ensemble was maintained at 298 K by the Nosé Hoover thermostat[42] and 105 Pa by the Parrinello Rahman barostat[43]. The Particle-Mesh Ewald method[44] was used to treat long-range electrostatic interactions, and van der Waals interactions were treated with a cutoff distance of 1.20 nm. The 3-ns simulations were carried out for each separation with three different initial configurations at the constant temperature and pressure using the same methods mentioned above, to explore the critical distance for dewetting DC below which vapor is the stable phase. 3 Results and Discussion Figure 2 shows snapshots from two dewetting simulations with an initial separation distance of D= 0.55 nm. One sees from Fig.2a that water molecules inside the gap of the melittin tetramer in pure water are quickly expelled within about 100 ps. Then, a strong, cooperative water dewetting transition occurs inside the nanoscale channel in a large enough size to accommodate two or three layers of water molecules.. The entire dewetting process is about 300 ps, with a fluctuation for a short period before all the water molecules inside the channel are expelled. In Fig.2b, water molecules in the gap of melittin tetramer in the ethanol solutions are expelled, too, but it is slower than that in pure water. The ethanol molecules enter the gap of the melittin tetramer first, and then are expelled with the water molecules. In the aqueous ethanol solutions, there is also a dewetting transition, which takes about 500 ps.

254 REN Xiuping et al. / Nuclear Science and Techniques 23 (2012) 252 256 Fig.2 Snapshots of water molecules inside the gap of the melittin tetramer in pure water (a) and of water or ethanol molecules inside the gap of the melittin tetramer in aqueous ethanol solutions (b). Only water and ethanol molecules near the center of the channel were plotted, which was defined as the region with a spherical radius of 1 nm from the center of the enlarged tetramer. For clarity, the melittin tetramer, water and ethanol are shown as transparent gray ribbons, gray bonds, and van der Waals spheres, respectively. Fig.3 Time traces of the atom number inside the channel of melittin tetramer in pure water (a) and aqueous ethanol solution (b) with D=0.55, 0.6, 0.65, 0.7 and 0.75 nm. Figure 3 plots a few time traces of the atom number inside the channel of the melittin tetramer, starting from different initial distances with one initial configuration. Fig.3a shows the results in pure water: at D=0.5 and 0.55 nm, water molecules inside the channel of the melittin tetramer are expelled and the vapor phase is dominant; at D=0.6 nm, both the liquid phase and vapor phase exist inside the channel of the melittin tetramer; at D>0.6 nm, the number of water molecules inside the channel of the melittin tetramer keeps almost equal to that with the wet initial conditions, and the liquid phase is dominant. The simulation results show that the critical distance for dewetting DC in pure water is about 0.6 nm, which is equivalent to two water-molecule diameters. Fig.3b shows the results in aqueous ethanol solutions: water and ethanol molecules inside the channel of the melittin tetramer are expelled at D<0.65 nm, and the

REN Xiuping et al. / Nuclear Science and Techniques 23 (2012) 252 256 255 vapor phase is dominant; the number of water and ethanol inside the channel of the melittin tetramer at D >0.7 nm keeps about equal to that with the wet initial conditions, and the liquid phase is dominant; the channel of the melittin tetramer vacillates between the liquid phase and vapor phase at D=0.65 or 0.7 nm. The simulation results show that the critical distance for dewetting D C in aqueous ethanol solutions is 0.65 0.7 nm, which is larger than that in pure water. Comparing Fig.3b with Fig.3a, one finds that the melittin tetramer collapses more easily in aqueous ethanol solutions at low ethanol concentration. We note that there are many atoms inside the channel of the melittin tetramer in aqueous ethanol solutions than that in pure water. D=0.55 nm, water inside the channel of melittin tetramer are expelled very quickly, with only a small fluctuation in water number, while the number of inner ethanol increase with a large fluctuation. The liquid phase is dominant at D=0.75 nm, both water and ethanol have a large fluctuation. From Fig.4b, it could be deduced that ethanol preferentially locate in the vicinity of the melittin tetramer. That is to say, ethanol molecules prefer to stay at the interfaces. Vazquez et al. [45] measured the surface tension of aqueous alcohol solutions at the liquid-vapor interface and found that surface tension decreases as the ethanol concentration increases for a given temperature. It can be concluded that the ethanol molecules locate at the liquid-vapor interface, and decrease the liquid-vapor surface energy through the previous results [45,46]. Then, the vapor phase is preferred in the gap of the melittin tetramer, and the critical distance for dewetting becomes larger. 4 Conclusions Fig.4 Time traces of the atom number of water (a) and ethanol (b) inside the channel of melittin tetramer in aqueous ethanol solutions with D=0.55, 0.65 and 0.75 nm. To explore mechanisms of the promoting effect of ethanol on the dewetting transition, we calculated the time traces of the atom number of water and ethanol inside the channel of melittin tetramer in aqueous ethanol solutions with D=0.55 nm (in vapor phase), 0.65 nm (vacillating between two phases) and 0.75 nm (in liquid phase), respectively (Fig.4). As mentioned above, the vapor phase is dominant at The dewetting transition of the melittin tetramer immersed in 8% aqueous ethanol solutions was investigated by the molecular dynamics simulations. The obvious dewetting transitions happen inside a nanoscale channel of the melittin tetramer in pure water and aqueous ethanol solution. The dewetting process is slightly faster in pure water (~300 ps) than in aqueous ethanol solution (~500 ps). But the addition of ethanol molecules increases the critical distance for dewetting from 0.6 nm to 0.65 0.7 nm, and promotes this dewetting transition. Also, the ethanol molecules preferentially locate in the vicinity of the melittin tetramer. It could be concluded that the ethanol molecules locate at the liquid-vapor interface, and decrease the liquid-vapor surface energy. Therefore, the vapor phase is preferred in the gap of the melittin tetramer, and the critical distance for dewetting becomes larger. This observation is helpful to understanding the dewetting behavior of the cosolvent solutions. Acknowledgments We thank Professors Haiping Fang, Jun Hu, and Bin Li, and Dr. Wenpeng Qi and Dr. Guanghong Zuo, for their helpful suggestions. This work was partly

256 REN Xiuping et al. / Nuclear Science and Techniques 23 (2012) 252 256 supported by the National Science Foundation of China (No. 10975175, 90923002, 21073222) and Chinese Academy of Sciences (No. KJCX2-EW-N03). References 1 Levy Y, Onuchic J N. Annu Rev Biophys Biomol Struct, 2006, 35: 389 415. 2 Hummer G, Garde S, Garcia A E, et al. Chem Phys, 2000, 258: 349 370. 3 Oro J R D. J Bio Phys, 2001, 27: 73 79. 4 Beckstein O, Sansom M S P. Proc Natl Acad Sci U S A, 2003, 100: 7063 7068. 5 Beckstein O, Biggin P C, Sansom M S P. J Phys Chem B, 2001, 105: 12902 12905. 6 Rosta E, Buchete N-V, Hummer G. J Chem Theory Comput, 2009, 5: 1393 1399. 7 Wang C L, Lu H J, Wang Z G, et al. Phys Rev Lett, 2009, 103: 137801. 8 Wang C L, Zhou B, Xiu P, et al. J Phys Chem C, 2011, 115: 3018 3024. 9 Stillinger F H. J Solution Chem, 1973, 2: 141 158. 10 Lum K, Luzar A. Phys Rev E, 1997, 56: R6283 R6286. 11 Li X, Li J Y, Eleftheriou M, et al. J Am Chem Soc, 2006, 128: 12439 12447. 12 Choudhury N, Pettitt B M. J Am Chem Soc, 2005, 127: 3556 3567. 13 Cheng Y K, Rossky P J. Nat, 1998, 392: 696 699. 14 Ashbaugh H S, Paulaitis M E. J Am Chem Soc, 2001, 123: 10721 10728. 15 Wallqvist A, Gallicchio E, Levy R M. J Phys Chem B, 2001, 105: 6745 6753. 16 Werder T, Walther J H, Jaffe R L, et al. J Phys Chem B, 2003, 107: 1345 1352. 17 Granick S, Bae S C. Sci, 2008, 322: 1477 1478. 18 Kaiser A B, Gómez-Navarro C, Sundaram R S, et al. Nano Lett, 2009, 9: 1787 1792. 19 Sansom M S P, Biggin P C. Nat, 2001, 414: 156 158. 20 Hummer G, Rasaiah J C, Noworyta J P. Nat, 2001, 414: 188 190. 21 Leung K, Luzar A, Bratko D. Phys Rev Lett, 2003, 90: 065502. 22 Rasaiah J C, Garde S, Hummer G. Annu Rev Phys Chem, 2008, 59: 713 740. 23 Xiu P, Yang Z X, Zhou B, et al. J Phys Chem B, 2011, 115: 2988 2994. 24 Lum K, Chandler D, Weeks J D. J Phys Chem B, 1999, 103: 4570 4577. 25 ten Wolde P R, Chandler D. Proc Natl Acad Sci U S A, 2002, 99: 6539 6543. 26 Huang X, Margulis C J, Berne B J. Proc Natl Acad Sci U S A, 2003, 100: 11953 11958. 27 Giovambattista N, Lopez C F, Rossky P J, et al. Proc Natl Acad Sci U S A, 2008, 105: 2274 2279. 28 Berne B J, Weeks J D, Zhou R H. Annu Rev Phys Chem, 2009, 60: 85 103. 29 Yang Z, Shi B, Lu H, et al. J Phys Chem B, 2011, 115: 11137 11144. 30 Zhou R H, Huang X H, Margulis C J, et al. Sci, 2004, 305: 1605 1609. 31 Liu P, Huang X H, Zhou R H, et al. Nat, 2005, 437: 159 162. 32 Timasheff S N. Annu Rev Biophys Biomol Struct, 1993, 22: 67 97. 33 Bolen D W, Rose G D. Annu Rev Biochem, 2008,77: 339 362. 34 Herskovi.Tt, Gadegbek.B, Jaillet H. J Biol Chem, 1970, 245: 2588 2598. 35 Das P, Zhou R. J Phys Chem B, 2010, 114: 5427 5430. 36 England J L, Pande V S, Haran G. J Am Chem Soc, 2008, 130: 11854 11855. 37 Terwilliger T C, Eisenberg D. J Biol Chem, 1982, 257: 6010 6015. 38 Jorgensen W L, Maxwell D S, TiradoRives J. J Am Chem Soc, 1996, 118: 11225 11236. 39 Berendsen H J C, Grigera J R, Straatsma T P. J Phys Chem, 1987, 91: 6269 6271. 40 Miyamoto S, Kollman P A. J Comput Chem, 1992, 13: 952 962. 41 Lindahl E, Hess B, van der Spoel D. J Mol Model, 2001, 7: 306 317. 42 Shuichi N. J Chem Phys, 1984, 81: 511 519. 43 Parrinello M, Rahman A. J Appl Phys, 1981, 52: 7182 7190. 44 Essmann U, Perera L, Berkowitz M L, et al. J Chem Phys, 1995, 103: 8577 8593. 45 Vazquez G, Alvarez E, Navaza J M. J Chem Eng Data, 1995, 40: 611 614. 46 Lundgren M, Allan N L, Cosgrove T. Langmuir, 2002, 18: 10462 10466.