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

Similar documents
Structuring of hydrophobic and hydrophilic polymers at interfaces Stephen Donaldson ChE 210D Final Project Abstract

Kobe-Brown Simulation Summer School 2015 Project: DPD Simulation of a Membrane

Dissipative Particle Dynamics: Foundation, Evolution and Applications

Mesoscopic simulation for the structural change of a surfactant solution using dissipative particle dynamics

Introduction to molecular dynamics

Dynamics of Polymers and Membranes P. B. Sunil Kumar

Directed Assembly of Functionalized Nanoparticles with Amphiphilic Diblock Copolymers. Contents

Effect of surfactant structure on interfacial properties

arxiv: v1 [physics.chem-ph] 11 Feb 2014

Introduction to Computer Simulations of Soft Matter Methodologies and Applications Boulder July, 19-20, 2012

Supplementary Information for Spontaneous symmetry breaking induced unidirectional rotation of a chain-grafted colloidal particle in the active bath

Origins of Mechanical and Rheological Properties of Polymer Nanocomposites. Venkat Ganesan

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

Adaptive resolution molecular-dynamics simulation: Changing the degrees of freedom on the fly

Supporting Information Soft Nanoparticles: Nano Ionic Networks of Associated Ionic Polymers

Supporting information

Lecture 2+3: Simulations of Soft Matter. 1. Why Lecture 1 was irrelevant 2. Coarse graining 3. Phase equilibria 4. Applications

Distribution of chains in polymer brushes produced by a grafting from mechanism

Dissipative particle dynamics simulations of weak polyelectrolyte adsorption on charged and neutral surfaces as a function of the degree of ionization

Chain Stiffness and Attachment-Dependent Attraction between Polyelectrolyte-Grafted Colloids

Dissipative particle dynamics: Bridging the gap between atomistic and mesoscopic simulation

Comparative Study of the Water Response to External Force at Nanoscale and Mesoscale

Computer Simulations of Soft Nanoparticles and Their Interactions with DNA-Like Polyelectrolytes. STOLL, Serge. Abstract

Theory, Algorithms and Applications of Dissipative Particle Dynamics. Sept , Shanghai

Supplementary Information

Coarse-graining, applications and mesoscopic molecular dynamics

Coarse-grained Models for Oligomer-grafted Silica Nanoparticles

MARTINI simulation details

Randomly Triangulated Surfaces as Models for Fluid and Crystalline Membranes. G. Gompper Institut für Festkörperforschung, Forschungszentrum Jülich

Coarse-graining limits in open and wall-bounded dissipative particle dynamics systems

Structure and Dynamics of Polymeric Canopies in Nanoscale Ionic Materials: An Electrical Double Layer Perspective

Potentials, periodicity

Supplementary Materials

Stability of colloidal systems

Supporting Information

A Molecular Dynamics Simulation of a Homogeneous Organic-Inorganic Hybrid Silica Membrane

From Molecular Dynamics to hydrodynamics a novel Galilean invariant thermostat

Specific ion effects on the interaction of. hydrophobic and hydrophilic self assembled

Protein Synthetic Lipid Interactions

Structure-Property Relationships of Porous Materials for Carbon Dioxide Separation and Capture

Electrostatics of membrane adhesion

Scientific Computing II

Multimedia : Boundary Lubrication Podcast, Briscoe, et al. Nature , ( )

LAMMPS Performance Benchmark on VSC-1 and VSC-2

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

A MOLECULAR DYNAMICS STUDY OF POLYMER/GRAPHENE NANOCOMPOSITES

Lecture 3. Phenomena at Liquid-gas and Liquid-Liquid interfaces. I

Diffuse Interface Field Approach (DIFA) to Modeling and Simulation of Particle-based Materials Processes

Supporting Information for

On the Dynamics and Disentanglement in Thin and Two-Dimensional Polymer Films

Molecular Dynamics Simulations of Polyelectrolyte-Polyampholyte Complexes. Effect of Solvent Quality and Salt Concentration

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

Investigation of Surfactant Efficiency Using Dissipative Particle Dynamics

Colloidal Suspension Rheology Chapter 1 Study Questions

Colloid Chemistry. La chimica moderna e la sua comunicazione Silvia Gross.

Modeling of Micro-Fluidics by a Dissipative Particle Dynamics Method. Justyna Czerwinska

Simulations of Self-Assembly of Polypeptide-Based Copolymers

NANO 243/CENG 207 Course Use Only

Electronic Supplementary Information (ESI)

Surface Forces & Liquid Films (Answers to Exercise Problems)

Colloid stability. Lyophobic sols. Stabilization of colloids.

Smoothed Dissipative Particle Dynamics: theory and applications to complex fluids

Intermolecular and Surface Forces

APMA 2811T. By Zhen Li. Today s topic: Lecture 2: Theoretical foundation and parameterization. Sep. 15, 2016

minimal models for lipid membranes: local modifications around fusion objects

arxiv:cond-mat/ v3 [cond-mat.soft] 12 Sep 2001

An electrokinetic LB based model for ion transport and macromolecular electrophoresis

Contents. Preface XI Symbols and Abbreviations XIII. 1 Introduction 1

Supplementary Information for Atomistic Simulation of Spinodal Phase Separation Preceding Polymer Crystallization

The effect of surface dipoles and of the field generated by a polarization gradient on the repulsive force

TECHNOLOGIES THAT TRANSFORM POLLUTANTS TO INNOCUOUS COMPONENTS: CHEMICAL AND PHYSICOCHEMICAL METHODS

Experimental Soft Matter (M. Durand, G. Foffi)

Systematic Coarse-Graining and Concurrent Multiresolution Simulation of Molecular Liquids

Nucleation rate (m -3 s -1 ) Radius of water nano droplet (Å) 1e+00 1e-64 1e-128 1e-192 1e-256

Supplementary information for

Stick and Slip Behaviour of Confined Oligomer Melts under Shear. A Molecular-Dynamics Study.

Property Prediction with Multiscale Simulations of Silicon Containing Polymer Composites

Supporting information. The role of halide ions in the anisotropic growth. of gold nanoparticles: a microscopic, atomistic.

Direct Oil Recovery from Saturated Carbon Nanotube Sponges

Budding Dynamics of Individual Domains in Multicomponent Membranes Simulated by N-Varied Dissipative Particle Dynamics

Van Gunsteren et al. Angew. Chem. Int. Ed. 45, 4064 (2006)

Modeling and Simulating Gold Nanoparticle Interactions on a Liquid-Air Interface

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

Glass-Transition and Side-Chain Dynamics in Thin Films: Explaining. Dissimilar Free Surface Effects for Polystyrene and Poly(methyl methacrylate)

Liquid-vapour oscillations of water in hydrophobic nanopores

Sanitary Engineering. Coagulation and Flocculation. Week 3

Dissipative Particle Dynamics

Measurements of interaction forces in (biological) model systems

Introduction to Molecular Dynamics

Chiral selection in wrapping, crossover, and braiding of DNA mediated by asymmetric bend-writhe elasticity

Coupling Atomistic and Continuum Hydrodynamics

This semester. Books

Next layer is called diffuse layer----ions not held tightly----thickness is > 1000 angstroms-- exact distance depends on ionic strength of soln


Contents. Preface XIII

Ionic Liquids simulations : obtention of structural and transport properties from molecular dynamics. C. J. F. Solano, D. Beljonne, R.

Lecture 3 Charged interfaces

On the size and shape of self-assembled micelles

An alternative approach to dissipative particle dynamics

Transcription:

Supplementary Information for: Controlling Cellular Uptake of Nanoparticles with ph-sensitive Polymers Hong-ming Ding 1 & Yu-qiang Ma 1,2, 1 National Laboratory of Solid State Microstructures and Department of Physics, Nanjing University, Nanjing 2193, China 2 Center for Soft Condensed Matter Physics and Interdisciplinary Research, Soochow University, Suzhou 2156, China E-mail address: myqiang@nju.edu.cn I. Supplementary Methods The DPD is a coarse-grained simulation technique with hydrodynamic interaction 1. The dynamics of the elementary units which are so-called DPD beads, is governed by Newton s equation of motion dv i /dt = f i /m. Typically, in the DPD, there are three types of pairwise forces acting on bead i by bead j: the conservative force, dissipative force, and random force. In the present work, the electrostatic force is introduced to take into account the electrostatic interactions between charged beads. The conservative force F C ij = a ij (1 r ij /r c )ê ij is used to model the repulsive interaction of beads i and j, where r ij = r ij is the distance between beads i and j, ê ij = r ij /r ij is the unit vector, r c is the cutoff radius of the force, and a ij represents the maximum repulsion interaction of beads i and j. For any two beads of the same type, we take the repulsive parameter a ii = 25, and for any two beads of different types, we set the interaction parameter a W H = a W P = a W I = a W C = a HP = a HI = a HC = a P I = a P C = a IC = 25 and a W T = a P T = a HT = a IT = a CT = 1 (W stands for water bead, and I represents counterion bead which is introduced into the system to ensure the charge neutrality) to denote the hydrophilic/hydrophobic property of the beads 2,3. The dissipative force F D ij = γ(1 r ij /r c )(ê ij v ij )ê ij and random force F R ij = 2γk B T (1 r ij /r c )ζ ij t 1/2 ê ij are for thermostat, where v ij = v i v j is relative velocity between beads i and j, γ is the strength of friction, ζ ij is a symmetric random variable with zero mean and unit variance, and t is the time step of simulation 1.

2 Electrostatic interactions were incorporated into the DPD simulations by Groot 4. Since soft potential in the DPD allows for the overlap between DPD beads, when the charged DPD beads are modeled, this can lead to the formation of artificial ion pairs and cause the divergence of the electrostatic potential. To avoid this problem, Groot chose to spread out the charges using the distribution 4 : ρ e (r) = 3 πr e 3 (1 r/r e ) with r < r e, where r e is the electrostatic smearing radius, and is typically set as 1.6r c (for details of the method, see Ref. 4). Moreover, in order to mimic the receptor-ligand interaction, we use a modified LJ potential 5 : U LJ ij = 4ϵ[( σ r ij ) 12 ( σ r ij ) 6 ] +.22ϵ, where r ij r cut, σ =.624r c, and ϵ is used to identify the strength of the receptor-ligand interaction (we fix it as 1 k B T ). The cutoff r cut of the potential is the same as that in the DPD (i.e., r c ) unless otherwise stated. Additionally, to guarantee the proper running of the DPD technology, the repulsive force is set to be 25k B T/r cut if it is larger than 25k B T/r cut, therefore this potential becomes soft. Further, we use a harmonic bond U s = k s (1 r i,i+1 /l ) 2 (here we choose k s = 64, l =.5r c ) between the neighboring beads to ensure the integrality of lipids and polymers, and insert a weaker bond (k s = 1, l =.5r c ) between the first hydrophobic beads on two tails of the lipid to keep the tails oriented in the same direction 3. We also use a three-body bond angle potential U a = k a (1 cos(ϕ ϕ )) to depict the rigidity of lipid tails (k a = 1, ϕ = 18 ). Our simulations apply the velocity-verlet integration algorithm and the integration time step t =.15τ. In addition, we choose the cutoff radius r c, bead mass m, energy k B T as the simulation units. All simulations are performed in the NVT ensembles. The size of the simulation box is 65r c 65r c 4r c with the number density of ρ = 3/rc. 3 The area (A ) per lipid when the membrane is under zero tension at the beginning of the simulations is about 1.28nm 2. During the simulations, to keep the membrane surface under zero tension, the box shape changes with the area (A b ) per lipid on the boundary, i.e., if A b > A, the box will be compressed in X-Y plane until A b = A ; while if A b < A, the box will be stretched in X-Y plane until A b = A. At the same time, the box length in membrane-normal direction will change correspondingly to keep the box volume fixed 2. And we perform the above operation every 1 time steps. The periodic boundary conditions are adopted in three directions. We can convert the DPD units into SI units via examining the membrane thickness and the lipid diffusion coefficient 5. Usually, the thickness of DPPC bilayer is about 4nm and

3 a b 4 3 Lipid Tail Lipid head (no charge) Water Lipid head (positive) Lipid head (negtive) Receptor nm -2 2 1-4 -2 2 4 Z (nm) Figure S1: Additional information of nanoparticle-polymer solutions and membranes-water systems. (a) Snapshot of the initial state of placing nanoparticle in polymer solutions, where there are 1 polymers (the water and counterion beads are omitted). (b) Density profiles of water beads, lipid head beads, lipid tail beads and receptor head beads. the in-plane diffusion constant of lipids is about 5.µm 2 /s in experiment. Comparing that the thickness is about 4.r c (the distance between two peaks in density profiles describing the first two beads of lipid molecules) and diffusion constant is about.12r 2 c /τ, we can yield r c = 1.nm and τ = 2.4ns. All simulations in this work are carried out by using the modified soft package Lammps (12 Jun 211) 6. II. Supplementary Figures and Discussions Figures S2a-d show the snapshots of the final equilibrium of nanoparticle-polymers complex (NPC) under different ionized degrees (i.e., external ph value). With the increase of ionized monomer number, the electrostatic attractive interaction becomes stronger, and the number of adsorbed polymers on the nanoparticle will increase. However, with a further increase of ionized monomer number, the number of adsorbed polymers will decrease, i.e., the maximum number of adsorbed polymers corresponds to the middle range of N (from 4 to 6), as shown in Fig. S2e. This is due to an increase of the electrostatic repulsive interaction between the adsorbed polymers on the particle surface with ionized monomer number N, while the averaged attractive interaction energy be-

4 a b c d e Adsorption number 1 8 6 4 2 4 6 8 1 12 N f Averaged energy (k B T) 6 45 3 15 2 4 6 8 1 12 N Figure S2: Final equilibrium of charged nanoparticles in polymer solutions. (a) N=1, (b) N=4, (c) N=8, and (d) N=11. The bead colors are the same as those of the snapshot in Fig. 1 in the main text. (e) The total adsorption number of polymers on the nanoparticles surface as a function of ionized number N. (f) The averaged attractive interaction energy (absolute value) between nanoparticles and polymers increase monotonously with ionized number N, indicating that the adsorbed polymers on the nanoparticle surface become more and more stable. The red line is the linear fitted line. Error bars are obtained by taking the standard derivation of ten independent simulations. tween single polymer and nanoparticle increase linearly with the increase of N (see Fig. S2f). In order to determine the zeta potential ζ of the nanoparticle-polymer complex, we first calculate the mean electrostatic potential ϕ at the distance r from the nanoparticle center 7 : ϕ(r) = dr E(r ) = q dr P (r ), where E(r) is the electric field and P(r) is the integrated 4πε r 2 r r charge in units of q within the distance r, and ε is the dielectric constant of water. The mean electrostatic potential as functions of the distance r under different ph conditions is shown in Fig. S3. Then we can obtain the zeta potential in various cases. Interestingly, as shown in the inset of Fig. S3, it is found that the zeta potential will drop below zero when N 6, indicating that the surface charge sign of nanoparticle could be changed because of

5 (mv) 5 4 3 2 1 N=1 N=2 N=4 N=6 N=8 N=1 N=11 (mv) 8 4-4 2 4 6 N 8 1 12-1 4 6 8 1 12 r (nm) Figure S3: Mean electrostatic potential (ϕ) as a function of distance (r) from nanoparticle center. The inset shows the zeta potential (ζ) of nanoparticle-polymer complex as a function of the ionized number N. Note that the zeta potential cannot be obtained exactly in simulations because the shear plane is not easy to be determined 7, here we just use the electrostatic potential corresponding to r=5.5nm as the zeta potential 8. Error bars are obtained by taking the standard derivation of ten independent simulations. the associated anionic polymers on its surface. Supplementary References 1 Groot, R. D. & Warren, P. B. Dissipative particle dynamics: Bridging the gap between atomistic and mesoscopic simulations. J. Chem. Phys. 17, 4423-4435 (1997). 2 Smith, K. A., Jasnow, D. & Balazs, A. C. Designing synthetic vesicles that engulf nanoscopic particles. J. Chem. Phys. 127, 8473 (27). 3 Ding, H. M., Tian, W. D. & Ma, Y. Q. Designing nanoparticle translocation through membranes by computer simulations. ACS Nano 6, 123-1238 (212). 4 Groot, R. D. Electrostatic interactions in dissipative particle dynamics simulation of polyelectrolytes and anionic surfactants. J. Chem. Phys. 118, 11265 (23).

6 5 Yang, K. & Ma, Y. Q. Computer simulation of the translocation of nanoparticles with different shapes across a lipid bilayer. Nat. Nanotechnol. 5, 579-583 (21). 6 Plimpton, S. J. Fast parallel algorithms for short-range molecular dynamics. J. Comput. Phys. 117, 1-19 (1995). 7 Diehl, A. & Levin, Y. Smoluchowski equation and the colloidal charge reversal. J. Chem. Phys. 125, 5492 (26). 8 Wang, Z. Y. & Ma, Y. Q. Insights from Monte Carlo simulations on charge inversion of planar electric double layers in mixtures of asymmetric electrolytes. J. Chem. Phys. 133, 6474 (21).