Physicochemical Properties of Nanoparticles Regulate Translocation across Pulmonary. Surfactant Monolayer and Formation of Lipoprotein Corona

Similar documents
Interaction of Functionalized C 60 Nanoparticles with Lipid Membranes

Effects of graphene oxide nanosheets on ultrastructure and. biophysical properties of pulmonary surfactant film

We parameterized a coarse-grained fullerene consistent with the MARTINI coarse-grained force field

Supplementary Information

Penetration of Gold Nanoparticle through Human Skin: Unraveling Its Mechanisms at the Molecular Scale

The Molecular Mechanism of Monolayer-Bilayer Transformations of Lung Surfactant from Molecular Dynamics Simulations

Supporting information for: Quantifying. lateral inhomogeneity of. cholesterol-containing membranes

Coarse grained simulations of Lipid Bilayer Membranes

MARTINI Coarse-Grained Model of Triton TX-100 in Pure DPPC. Monolayer and Bilayer Interfaces. Supporting Information

Supporting Information

Coarse-Grained Molecular Dynamics for Copolymer- Vesicle Self-Assembly. Case Study: Sterically Stabilized Liposomes.

Supporting material. Membrane permeation induced by aggregates of human islet amyloid polypeptides

Gold nanocrystals at DPPC bilayer. Bo Song, Huajun Yuan, Cynthia J. Jameson, Sohail Murad

Interactions of Polyethylenimines with Zwitterionic and. Anionic Lipid Membranes

Supporting Information: Revisiting partition in. hydrated bilayer systems

Supplementary information for Effects of Stretching Speed on. Mechanical Rupture of Phospholipid/Cholesterol Bilayers: Molecular

Molecular Dynamics Simulations of the Anchoring and Tilting of the Lung-Surfactant Peptide SP-B 1-25 in Palmitic Acid Monolayers

The effect of orientation dynamics in melittin as antimicrobial peptide in lipid bilayer calculated by free energy method

Interaction Between Amyloid-b (1 42) Peptide and Phospholipid Bilayers: A Molecular Dynamics Study

Role of Surface Ligands in Nanoparticle Permeation through a Model. Membrane: A Coarse grained Molecular Dynamics Simulations Study

Molecular modeling of the pathways of vesicle membrane interaction. Tongtao Yue and Xianren Zhang

Nanoparticle Permeation Induces Water Penetration, Ion Transport, and Lipid Flip-Flop

Xianren Zhang ( 张现仁 )

SDS-Assisted Protein Transport Through Solid-State Nanopores

Translocation of C 60 and Its Derivatives Across a Lipid Bilayer

RELEASE OF CONTENT THROUGH MECHANO- SENSITIVE GATES IN PRESSURISED LIPOSOMES. Martti Louhivuori University of Groningen

A proposed mechanism for tear film breakup: a molecular level view by employing in-silico approach

induced inactivation of lung surfactants

Modeling of Nanostructures : Bionanosystems, Polymers, and Surfaces

Phase Behavior of a Phospholipid/Fatty Acid/Water Mixture Studied in Atomic Detail

Biological Membranes. Lipid Membranes. Bilayer Permeability. Common Features of Biological Membranes. A highly selective permeability barrier

Supporting Information for

Chemical Surface Transformation 1

Adaptable Lipid Matrix Promotes Protein Protein Association in Membranes

University of Groningen. Lipids out of equilibrium Tieleman, D. Peter; Marrink, Siewert. Published in: Journal of the American Chemical Society

Coarse-Grain Simulations Reveal Movement of the Synaptobrevin C-Terminus in Response to Piconewton Forces

SUPPLEMENTARY INFORMATION

Supporting Information for Origin of 1/f noise in hydration dynamics on lipid membrane surfaces

Karen Doniza 1*, Hui Lin Ong 2, Al Rey Villagracia 1, Aristotle Ubando 5 Melanie David 1, Nelson Arboleda Jr. 1,3, Alvin Culaba 5, Hideaki Kasai 4

Hydrophobic Compounds Reshape Membrane Domains

PROCEEDINGS OF THE YEREVAN STATE UNIVERSITY

Supplementary Figures

in-silico Design of Nanoparticles for Transdermal Drug Delivery Application

Bilayer Deformation, Pores & Micellation Induced by Oxidized Lipids

The Transport and Organization of Cholesterol in Planar Solid-Supported Lipid Bilayer Depend on the Phospholipid Flip-Flop Rate

A: All atom molecular simulation systems

MOLECULAR DYNAMICS SIMULATION OF MIXED LIPID BILAYER WITH DPPC AND MPPC: EFFECT OF CONFIGURATIONS IN GEL-PHASE

BIOPHYSICS II. By Prof. Xiang Yang Liu Department of Physics,

Molecular Dynamics Simulations of a Pulmonary Surfactant Protein B Peptide in a Lipid Monolayer

Massive oxidation of phospholipid membranes. leads to pore creation and bilayer. disintegration

Coarse-Grained Molecular Dynamics Study of Cyclic Peptide Nanotube Insertion into a Lipid Bilayer

John M. Leveritt III, Almudena Pino-Angeles, Themis Lazaridis*.

Lecture 15. Membrane Proteins I

Brazilian Journal of Physics ISSN: Sociedade Brasileira de Física Brasil

Supplementary Figures

RSC Advances.

Simulationen von Lipidmembranen

Transient β-hairpin Formation in α-synuclein Monomer Revealed by Coarse-grained Molecular Dynamics Simulation

Application of Molecular Modelling and EPR Spectroscopy to Lipid Membranes a Combined Approach

Physical Cell Biology Lecture 10: membranes elasticity and geometry. Hydrophobicity as an entropic effect

Molecular Dynamics Simulation of a Dipalmitoylphosphatidylcholine Bilayer with NaCl

Conformational Flexibility of the Peptide Hormone Ghrelin in Solution and Lipid Membrane Bound: A Molecular Dynamics Study

In this article we address a seemingly simple thermodynamic

Comparing Simulations of Lipid Bilayers to Scattering Data: The GROMOS 43A1-S3 Force Field

Interactions between Bisphosphate. Geminis and Sodium Lauryl Ether

SUPPLEMENTARY INFORMATION

Physical Pharmacy. Interfacial phenomena. Khalid T Maaroof MSc. Pharmaceutical sciences School of pharmacy Pharmaceutics department

Supporting Information: Atomistic Model for Near-Quantitative. Simulations of Langmuir Monolayers

Supporting Information Identification of Amino Acids with Sensitive Nanoporous MoS 2 : Towards Machine Learning-Based Prediction

Supplementary Information: A Critical. Comparison of Biomembrane Force Fields: Structure and Dynamics of Model DMPC, POPC, and POPE Bilayers

Structural Characterization on the Gel to Liquid-Crystal Phase Transition of Fully Hydrated DSPC and DSPE Bilayers

Cellular Neurophysiology I Membranes and Ion Channels

Model of an Asymmetric DPPC/DPPS Membrane: Effect of Asymmetry on the Lipid Properties. A Molecular Dynamics Simulation Study

Distribution of Amino Acids in a Lipid Bilayer from Computer Simulations

Supplementary Materials for

Supplementary Figure 1 (previous page). EM analysis of full-length GCGR. (a) Exemplary tilt pair images of the GCGR mab23 complex acquired for Random

Interactions Between a Voltage Sensor and a Toxin via Multiscale Simulations

Simulation of Self-Assembly of Ampiphiles Using Molecular Dynamics

Supporting Information. to the article. Conical lipids in flat bilayers induce packing defects similar to that induced by positive curvature

Permeability of Small Molecules through a Lipid Bilayer: A Multiscale Simulation Study

Ethanol Induces the Formation of Water-Permeable Defects in Model. Bilayers of Skin Lipids

Molecular Structure and Permeability at the Interface between Phase-Separated Membrane Domains

University of Groningen

PCCP Accepted Manuscript

Effects of Epicholesterol on the Phosphatidylcholine Bilayer: A Molecular Simulation Study

Molecular Dynamics Simulation of. Amphiphilic Aggregates

Asymmetry of lipid bilayers induced by monovalent salt: Atomistic molecular-dynamics study

level and yield the relevant dynamical and structural information.

Simulationen von Lipidmembranen

Cholesterol modulates amyloid beta peptide 1-42 channel formation in planar lipid membranes

Paper 4. Biomolecules and their interactions Module 22: Aggregates of lipids: micelles, liposomes and their applications OBJECTIVE

University of Groningen

Biochemistry 1 Recitation1 Cell & Water

Peptide Nanopores and Lipid Bilayers: Interactions by Coarse-Grained Molecular-Dynamics Simulations

Molecular dynamics simulation of a polysorbate 80 micelle in water

The main biological functions of the many varied types of lipids include: energy storage protection insulation regulation of physiological processes

Coarse-grained simulation studies of mesoscopic membrane phenomena

Biology 2E- Zimmer Protein structure- amino acid kit

Main Functions maintain homeostasis

Neutron reflectivity in biology and medicine. Jayne Lawrence

Transcription:

Physicochemical Properties of Nanoparticles Regulate Translocation across Pulmonary Surfactant Monolayer and Formation of Lipoprotein Corona Supplementary Information Guoqing Hu 1 *, Bao Jiao 1, Xinghua Shi 1, Russell P. Valle 2, Qihui Fan 2, and Yi Y. Zuo 2 * 1. The State Key Laboratory of Nonlinear Mechanics (LNM), Institute of Mechanics, Chinese Academy of Sciences, Beijing, 119, China 2. Department of Mechanical Engineering, University of Hawaii at Manoa, Honolulu, HI, 96822 * e-mail: guoqing.hu@imech.ac.cn; yzuo@hawaii.edu S1

S1. Additional experimental data Surface pressure (mn/m) 8 7 6 5 4 3 2 1 a Pure Infasurf control hr 1 hr 2 hrs 3 hrs 7 hrs 1 day Surface pressue (mn/m) 8 7 6 5 4 3 2 1 b 5 1 15 2 25 3 Surface area (cm 2 ) 5 1 15 2 25 3 Pure Infasurf Infasurf exposed to 1% HA-NPs for: hr 1 hr 2 hr 3 hr 7 hr 1 day Surface area (cm 2 ) Surface pressure (mn/m) 8 7 6 5 4 3 2 1 Pure Infasurf Infasurf exposed to 1% PST-NPs for hr.5 hr 1 hour 2 hours 3 hours 5 1 15 2 25 3 Surface area (cm 2 ) Figure S1. Time-dependent inhibition of Infasurf films by NPs of different hydrophobicity. a, Control experiment: time-dependent compression isotherms of pure Infasurf. Pure Infasurf is very stable, indicated by consistent compression isotherms without time-dependent shifting. 1 b, Infasurf exposed to 5 µg/ml (i.e., 1% surfactant phospholipids) hydrophilic hydroxyapatite NPs (HA-NPs) shows time-dependent inhibition, indicated by isotherm shifting to the left. Surfactant inhibition stabilizes after 7 hours of particle exposure. 1 c, Infasurf exposed to 5 µg/ml hydrophobic polystyrene NPs (PST-NPs) stabilizes after 1 hour of particle exposure. This comparison demonstrates the effect of particle hydrophobicity on kinetics of surfactant inhibition. The hydrophobic NPs tend to inhibit pulmonary surfactant much quicker than hydrophilic NPs. c S2

S2. Detailed molecular dynamics simulations S2.1. Overall system setup As shown in Fig. S2, our initial simulation system consisted of a water slab sandwiched by two symmetric pulmonary surfactant monolayers. The water slab was set to be 2 nm in thickness (approximately 1 times the thickness of a lipid monolayer), large enough to separate the two monolayers without interaction with each other. In other words, during our simulations these two leaflets behaved as independent monolayers instead of a bilayer. Single nanoparticles were introduced symmetrically from the airside modeled by a vacuum. This symmetric arrangement allows us to conveniently set up system boundaries without compromising the accuracy of simulating monolayers or significantly increasing the computational load. Only the top leaflet (at z=) of these two monolayers was illustrated in the paper. To study translocation of nanoparticles across static monolayers (Fig. 2), the system was studied within a simulation box with a constant volume and constant surface area of the monolayer. To simulate interactions between nanoparticles and dynamic pulmonary surfactant monolayers (Figs. 3-5), the simulation box and the surface area of the monolayers were compressed. S3

Supplementary Information 4 a b Air 2 z (nm) W ater -2-4 Air -6 5 1 15 2 25 3 3 Density of the lipid headgroup (kg/m ) Figure S2. Initial setup of the molecular dynamics simulation system. a, The initial density profile of lipid headgroups along the vertical direction of the simulation box; b, The simulated symmetric system with a water slab sandwiched by two non-interacting surfactant monolayers with individual nanoparticles introduced from the airside. S4

SP-B 1-25 DPPC POPG SP-C Neutral particle Cationic particle Anionic particle Figure S3. Schematics of coarse-grained (CG) molecules and nanoparticles involved in our simulations. DPPC in green; POPG in yellow; SP-B 1-25 in purple; SP-C in orange; neutral particle in black. Surface of the cationic particle was modified with randomly distributed positive charges shown in red. Surface of the anionic particle was modified with randomly distributed negative charges shown in blue. For clarification, cationic and anionic particles in text are shown in solid red and solid blue, respectively. S5

S2.2. Coarse-grained models The entire system was simulated with coarse-grained (CG) models which allow molecular simulations at a larger length-scale and longer time-scale than all-atom models. Each pulmonary surfactant monolayer contained 1,12 CG DPPC (green) and 48 CG POPG (yellow) molecules (7:3), doped with 7 CG SP-B 1-25 (purple) and 7 CG SP-C (orange) molecules. The simulation parameters of CG DPPC and CG POPG were taken from the MARTINI force field. 2-4 All-atom models for SP-C (1SPF.pdb) and SP-B 1-25 (1DFW.pdb) were obtained from the protein data bank. 5 CG models of these two protein peptides were derived from these all-atom models using martinize.py by analyzing the secondary structure and the sequence of residues of these protein peptides. The CG models for nanoparticles were built to be compatible with the MARTINI CG force field for proteins and lipids. 2-4 Each nanoparticle contained 1,721 CG beads arranged in a facecentered cubic arrangement. The diameter of the final nanoparticle was approximately 5 nm. The CG bead types P2 and C1 were used for constructing the neutral (black) hydrophilic and hydrophobic nanoparticles, respectively. Qda and Qa were used to modify the surface charges of the particle. Cationic particle was shown in red and anionic particle was shown in blue. The water slab contained 158,418 CG water beads and 862 Na + ions to neutralize the system. Fig. S3 shows the schematics and color schemes of the simulated CG molecules and particles. S1.3. Molecular dynamics simulations All simulations were performed using the GROMACS 4.5.4. 6 All simulations were conducted at a constant temperature of 31 K. For the study of particle translocation across a static surfactant monolayer (Fig. 2), the simulation box was set to be at a constant volume of 31 31 1 nm, at which the surfactant monolayer had a molecular area of approximately 6 Å 2 S6

per lipid. The nanoparticle (NP) was introduced to the pre-equilibrium monolayer and the interaction was simulated up to 2 ns. The potential of mean force (PMF) was calculated using the weighted histogram analysis method (WHAM) 7 and the umbrella sampling technique 8 with.2 nm spacing between the umbrella windows. For each window the system underwent 2 ns constant volume and temperature equilibration simulations and 5 ns umbrella sampling runs. For the study of particle interaction with dynamic surfactant film (Figs. 3-5), the simulation box and the surface area of the monolayers were compressed by applying positive lateral pressure using the semi-isotropic pressure coupling scheme (isotropic coupling in the x-y plane and anisotropic coupling in the z-direction). The Berendsen barostat was used for the semiisotropic pressure coupling with a coupling constant P 4 ps. The system compressibility was set to be 5 1-5 bar -1 in the x-y plane and bar -1 in the z-direction. The temperature was maintained at 31 K by Berendsen temperature coupling with a coupling constant T 1 ps. The time step was 2 fs and the neighbor list was updated every 1 steps. S7

S3. Interactions between the cationic hydrophilic NP and lipids For the hydrophilic NPs, we found that the cationic NP selectively adsorbs POPG. Fig. S4 shows the molar ratio of POPG to DPPC lipids adsorbed onto the NP surface. A higher POPG:DPPC ratio is found in the lipid monolayer adsorbed on the particle surface than that of the interfacial monolayer (POPG:DPPC=3:7). This is likely due to electrostatic interactions between the cationic NP and the anionic POPG molecules. 2. 1.5 POPG:DPPC 1..5 POPG:DPPC =.43 (3:7). 5 1 15 2 Time (ns) Figure S4. Time dependent variations of POPG:DPPC molecular ratio adsorbed on the surface of a cationic hydrophilic NP. S8

S4. Interactions between the anionic hydrophilic NP and surfactant proteins To study the binding specificity between the anionic hydrophilic NP and the protein molecules, Fig. S5a shows the interaction distance between the NP and the protein molecules (SP-B 1-25 and SP-C) during the simulation time. It is found that SP-B 1-25 selectively binds to the NP. However, despite being in contact with the NP from 25 to 125 ns, SP-C is eventually separated from the NP at equilibrium. Similarly, Fig. S5b shows the potential of mean force (PMF) required to "pull" a protein molecule out of the interfacial surfactant monolayer. It is found that the PMF of "pulling" a SP-C molecule is more than double of the PMF of "pulling" a SP-B 1-25 molecule, indicating a higher affinity between the surfactant monolayer and the SP-C molecule than with the SP-B 1-25 molecule. Interaction distance (nm) 14 12 1 8 6 4 2 SP-B 1-25 SP-C 5 1 15 2 Time (ns) a PMF (KJ/mol) 16 14 12 1 8 6 4 2-2 b SP-B 1-25 SP-C -4-3 -2-1 Distance (nm) Figure S5. Additional MD simulations that confirm the binding specificity between an anionic hydrophilic NP and surfactant proteins (SP-B 1-25 and SP-C). a, Interaction distance as a function of the simulation time; b, Potential of mean force (PMF) for interaction between the surfactant monolayer and the protein molecules. S9

S5. Amino acid sequences of SP-B 1-25 and SP-C The human SP-B monomer contains 79 amino acids. The detailed three dimensional structure of this protein is still unclear. In our MD simulations, we used SP-B 1-25, which is a SP-B peptide of 1-25 amino-terminal residues, to replace SP-B. Despite the simplicity, both experimental 9-15 and MD studies 16-18 have shown that SP-B 1-25 represents certain properties of the full-length SP- B. Fig. S6. compares the charge and polarity of amino acid sequences for SP-B 1-25 and SP-C. SP-B 1-25 contain 4 net positive changes and 16 non-polar amino acids out of the 25 amino acid residue. The human SP-C consists of 35 amino-acid polypeptide enriched in valine (V), leucine (L), and isoleucine (I), which make SP-C one of the most hydrophobic proteins in the proteome. Specifically, SP-C contains 3 net positive charges and 27 non-polar amino acids out of the 35 amino acid residue. Hence, compared to SP-C, SP-B 1-25 has a higher positive charge density and is significantly less hydrophobic. 1 1 11 2 21 25 FPIPLPYCWL CRALIKRIQA MIPKG a. SP-B 1-25 1 1 11 2 21 3 31 35 LRIPCCPVNL KRLLVVVVVV VLVVVVIVGA LLMGL b. SP-C Figure S6. Amino acid sequence of a. SP-B 1-25 ; and b. SP-C molecules. For each amino acid, the charge at ph 7 is indicated by different colors: positive (red), negative (blue), and neutral (black); the polarity is indicated by the background: polar (clear) and non-polar (gray). S1

References 1. Fan, Q.; Wang, Y. E.; Zhao, X.; Loo, J. S.; Zuo, Y. Y. Adverse Biophysical Effects of Hydroxyapatite Nanoparticles on Natural Pulmonary Surfactant. ACS. Nano. 211, 5, 641-6416. 2. Marrink, S. J.; de Vries, A. H.; Mark, A. E., Coarse Grained Model for Semiquantitative Lipid Simulations. J. Phys. Chem. B. 24, 18, 75-76. 3. Marrink, S. J.; Risselada, H. J.; Yefimov, S.; Tieleman, D. P.; De Vries, A. H. The MARTINI Force Field: Coarse Grained Model for Biomolecular Simulations. J. Phys. Chem. B. 27, 111, 7812-7824. 4. Monticelli, L.; Kandasamy, S. K.; Periole, X.; Larson, R. G.; Tieleman, D. P.; Marrink, S. J. The MARTINI Coarse-Grained Force Field: Extension to Proteins. J. Chem. Theory Comput. 28, 4, 819-834. 5. Berman, H. M.; Westbrook, J.; Feng, Z.; Gilliland, G.; Bhat, T.; Weissig, H.; Shindyalov, I. N.; Bourne, P. E. The Protein Data Bank. Nucleic Acids Res. 2, 28, 235-242. 6. Lindahl, E.; Hess, B.; van der Spoel, D. GROMACS 3.: A Package for Molecular Simulation and Trajectory Analysis. J. Mol. Model. 21, 7, 36-317. 7. Kumar, S.; Rosenberg, J. M.; Bouzida, D.; Swendsen, R. H.; Kollman, P. A. The Weighted Histogram Analysis Method for Free Energy Calculations on Biomolecules. I. The Method. J. Comput. Chem. 1992, 13, 111-121. 8. Torrie, G.; Valleau, J. Nonphysical Sampling Distributions in Monte Carlo Free-Energy Estimation: Umbrella Sampling. J. Comput. Phys. 1977, 23, 187-199. 9. Farver, R. S.; Mills, F. D.; Antharam, V. C.; Chebukati, J. N.; Fanucci, G. E.; Long, J. R. Lipid Polymorphism Induced by Surfactant Peptide SP-B(1-25). Biophys. J. 21, 99, 1773-1782. 1. Longo, M.; Bisagno, A.; Zasadzinski, J.; Bruni, R.; Waring, A. A Function of Lung Surfactant Protein SP-B. Science. 1993, 261, 453-456. 11. Lee, K. Y. C.; Lipp, M. M.; Zasadzinski, J. A.; Waring, A. J. Effects of Lung Surfactant Specific Protein SP-B and Model SP-B Peptide on Lipid Monolayers at the Air-Water Interface. Colloids Surf., A. 1997, 128, 225-242. 12. Veldhuizen, E. J. A.; Waring, A. J.; Walther, F. J.; Batenburg, J. J.; van Golde, L. M. G.; Haagsman, H. P. Dimeric N-terminal Segment of Human Surfactant Protein B (dsp-b1-25) Has Enhanced Surface Properties Compared to Monomeric SP-B1-25. Biophys. J. 2, 79, 377-384. 13. Gupta, M.; Hernandez-Juviel, J.; Waring, A.; Bruni, R.; Walther, F. Comparison of Functional Efficacy of Surfactant Protein B Analogues in Lavaged Rats. Eur.Respir. J. 2, 16, 1129-1133. 14. Gupta, M.; Hernandez-Juviel, J.; Waring, A.; Walther, F. Function and Inhibition Sensitivity of the N-terminal Segment of Surfactant Protein B (SP-B1-25) in Preterm Rabbits. Thorax. 21, 56, 871-876. 15. Flanders, B. N.; Vickery, S. A.; Dunn, R. C. Divergent Fluctuations in the Molar Area of a Model Lung Surfactant. J. Phys. Chem. B. 22, 16, 353-3533. 16. Choe, S.; Chang, R.; Jeon, J.; Violi, A. Molecular Dynamics Simulation Study of a Pulmonary Surfactant Film Interacting with a Carbonaceous Nanoparticle. Biophys. J. 28, 95, 412-4114. 17. Freites, J. A.; Choi, Y.; Tobias, D. J. Molecular Dynamics Simulations of a Pulmonary Surfactant Protein B Peptide in a Lipid Monolayer. Biophys. J. 23, 84, 2169-218. S11

18. Kaznessis, Y. N.; Kim, S.; Larson, R. G. Specific Mode of Interaction Between Components of Model Pulmonary Surfactants Using Computer Simulations. J. Mol. Biol. 22, 322, 569-582. S12