+ All Categories
Home > Documents > BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink,...

BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink,...

Date post: 03-Oct-2020
Category:
Upload: others
View: 0 times
Download: 0 times
Share this document with a friend
179
BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX GEOMETRIES by Yu Zhang A dissertation submitted in partial fulfillment of the requirements for the degree of Doctor of Philosophy (Chemical Engineering) at the UNIVERSITY OF WISCONSIN – MADISON 2011
Transcript
Page 1: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

BROWNIAN DYNAMICS SIMULATION OF

DNA IN COMPLEX GEOMETRIES

by

Yu Zhang

A dissertation submitted in partial

fulfillment of the requirements for the degree of

Doctor of Philosophy

(Chemical Engineering)

at the

UNIVERSITY OF WISCONSIN – MADISON

2011

Page 2: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and
Page 3: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

i

To my wife, Xi

For your love, your support,

and a wonderful child.

To my son, Alexander

For your laughters, which brighten my life.

To my parents, Liangzhu Zhang and Zhenglan Yu

For your unconditional love.

Page 4: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

ii

Acknowledgments

Thanks to my advisors, Michael Graham and Juan de Pablo for their support, pa-

tience, and guidance.

Thanks to former colleague Hongbo Ma, Juan Hernandez-Ortiz, and Patrick Un-

derhill for many useful discussions on fluid mechanics and polymer physics.

Thanks to current colleague Amit Kumar and Pieter Janssen for their helps and

many discussions.

Thanks also to other former and current colleagues of the Graham group – Wei

Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik

Pranay, Kushal Sinha, Rafael Henriquez Rivera, and Friedemann Hahn – for all your

helps and many discussions.

Research projects presented in this dissertation are financially supported by the

University of Wisconsin-Madison Nanoscale Science and Engineering Center funded

by National Science Foundation.

Page 5: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

iii

Abstract

This dissertation is concerned with the dynamics of a long DNA molecule in complex

geometries, driven by either electrostatic field or flow field. This is accomplished pri-

marily through the use of Brownian dynamics simulation, which captures the essential

physics at mesoscopic length scale, and allows us to simulate events happening on

long time scale, such as DNA pore translocation and cyclic dynamics of a tethered

DNA molecule in shear flow. Our work is novel in that, accurate and efficient algo-

rithms have been designed and developed for both electric field-driven and flow-driven

systems in complex geometries. We have focused on three distinct problems, which

we describe below.

In the electric field-driven case, we propose a novel class of electric field-actuated

soft mechanical control element for microfluidics. This type of element employs the

idea that under confinement a single polymer molecule is essentially a nanoscale

porous media. The chain could block the passageway of relatively large analytes such

as cells. At the same time, polyelectrolyte molecules, such as DNA, could deform

and squeeze through narrow pores when a sufficiently strong electric field is applied.

Brownian dynamics (BD)/Finite Element Method (FEM) simulation efficiently ex-

plore the design space, and results demonstrate that the On/Off switching could be

Page 6: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

iv

achieved within a proper parameter space.

To study the effects of a solid impenetrable wall on the dynamics of a nearby

DNA molecule, we examine the cyclic dynamics of a single DNA molecule tethered

to a hard wall in shear flow, using Brownian dynamics simulation. We focus on the

dynamics of the free end (last bead) of the tethered chain and we examined the cross-

correlation function and power spectral density of the chain extensions in the flow

and gradient directions as a function of chain length N and dimensionless shear rate

Wi. Extensive simulation results suggest a classical fluctuation-dissipation stochastic

process and question the existence of periodicity of the cyclic dynamics, as previously

claimed. We support our numerical findings with a simple analytical calculation for

a harmonic dimer in shear flow.

In the case of flow-driven DNA molecule in micro/nano-fluidics, one big challenge

is that an efficient algorithm is required to calculate fluctuating hydrodynamic in-

teractions (HI) in complex geometries. We have developed an accelerated immersed

boundary method that allows fast calculation of Brownian motion of polymer chains

and other particles in complex geometries with HI. With this new method, the first

detailed analysis of a recent set of interesting nanofluidic experiments involving DNA

dynamics in a complex flow geometry is performed. This analysis explains the ob-

served dynamics over a wide range of parameter values (flow rate, molecular wieght)

and illustrates the important quantitative effect of the hydrodynamic interactions on

the behavior of the system.

Page 7: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

v

Contents

Acknowledgments ii

Abstract iii

List of Figures viii

List of Tables x

1 Introduction 1

2 Problem Statement 4

2.1 Bead-spring DNA Model . . . . . . . . . . . . . . . . . . . . . . . . . 4

2.2 Kinetic theory of single polymer molecule . . . . . . . . . . . . . . . . 6

2.3 The Langevin equation and Brownian dynamics simulation . . . . . . 9

3 Intramolecular interactions and external forces 12

3.1 Spring force law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

3.2 Excluded volume interactions . . . . . . . . . . . . . . . . . . . . . . 14

3.3 Hydrodynamic interactions . . . . . . . . . . . . . . . . . . . . . . . . 15

3.4 Electrophoretic force . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

Page 8: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

vi

3.5 Polymer-wall steric interaction . . . . . . . . . . . . . . . . . . . . . . 19

4 Bistability and field-driven dynamics of confined tethered DNA 21

4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

4.2 Polymer model and simulation approach . . . . . . . . . . . . . . . . 28

4.3 Results and discussion . . . . . . . . . . . . . . . . . . . . . . . . . . 31

4.3.1 Pore-crossing geometry . . . . . . . . . . . . . . . . . . . . . . 31

4.3.2 Pore-entry geometry . . . . . . . . . . . . . . . . . . . . . . . 37

4.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41

5 Tethered DNA dynamics in shear flow 43

5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44

5.2 Discussion of Methods for Hydrodynamics of Polymer Solutions . . . 49

5.2.1 Implicit Solvent: Brownian Dynamics . . . . . . . . . . . . . . 50

5.2.2 Explicit Solvent: Continuum Methods . . . . . . . . . . . . . 52

5.2.3 Explicit Solvent: Particle Methods . . . . . . . . . . . . . . . 55

5.2.4 Coupled Methods . . . . . . . . . . . . . . . . . . . . . . . . . 56

5.3 Simulation Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57

5.3.1 Brownian Dynamics . . . . . . . . . . . . . . . . . . . . . . . 57

5.3.2 Lattice-Boltzmann . . . . . . . . . . . . . . . . . . . . . . . . 60

5.3.3 Stochastic Event-Driven Molecular Dynamics . . . . . . . . . 62

5.4 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

5.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78

6 Flow-driven DNA dynamics in complex geometries 80

6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81

Page 9: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

vii

6.2 Methods for hydrodynamics of confined polymer solutions . . . . . . 86

6.3 Polymer model and simulation method . . . . . . . . . . . . . . . . . 91

6.3.1 Model and governing equations . . . . . . . . . . . . . . . . . 91

6.3.2 Governing equations . . . . . . . . . . . . . . . . . . . . . . . 92

6.3.3 Mobility tensor and Fixman’s midpoint algorithm . . . . . . . 94

6.3.4 Chebyshev approximation . . . . . . . . . . . . . . . . . . . . 96

6.3.5 Fast Stokes solver with complex boundary conditions . . . . . 97

6.4 DNA flowing across an array of nanopits . . . . . . . . . . . . . . . . 114

6.4.1 Dynamics at low Peclet number . . . . . . . . . . . . . . . . . 117

6.4.2 Dynamics at high Peclet number . . . . . . . . . . . . . . . . 123

6.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125

7 Conclusion and future work 126

A Lattice random walk model of a tethered polymer 129

B BD/FEM algorithm for simulating DNA electrophoresis 137

C Dimer in shear flow 141

D Fixman’s midpoint algorithm 146

Page 10: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

viii

List of Figures

2.1 Schematic of a bead-spring chain in Brownian dynamics simulation. . 6

4.1 Schematic representations of soft nanomechanical bistable elements. . 26

4.2 Soft nanomechanical elements. . . . . . . . . . . . . . . . . . . . . . . 28

4.3 Top view of the simulation domains. . . . . . . . . . . . . . . . . . . 29

4.4 Time trajectories for a 3D pore-crossing system . . . . . . . . . . . . 32

4.5 Snapshots of a pore-crossing event. . . . . . . . . . . . . . . . . . . . 33

4.6 Simulation results for the quasi-2D and 3D pore crossing geometries. 35

4.7 Time trajectory and probability density for the pore entry geometry . 38

4.8 Simulation results for the quasi-2D pore-entry geometry . . . . . . . . 39

4.9 Phase transition of the “competitive bistability” system. . . . . . . . 40

5.1 Cyclic dynamics of a tethered DNA molecule . . . . . . . . . . . . . . 46

5.2 Probability distribution of the end bead of the tethered DNA molecule 68

5.3 Relaxation time of the tethered DNA . . . . . . . . . . . . . . . . . . 70

5.4 Spectral analysis of the end-bead position time series . . . . . . . . . 73

5.5 Cross correlation functions for different beads . . . . . . . . . . . . . 74

5.6 Comparison of the cross-correlation function among different methods 76

6.1 DNA flowing through nanopits . . . . . . . . . . . . . . . . . . . . . . 85

Page 11: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

ix

6.2 Schematic of the immersed boundary method . . . . . . . . . . . . . 99

6.3 Properties of the mobility tensor for the boundary particles . . . . . . 103

6.4 Screening function for the GGEM algorithm . . . . . . . . . . . . . . 105

6.5 Comparison between delta and regularized delta point forces . . . . . 106

6.6 Comparison between Hasimoto’s solution and GGEM . . . . . . . . . 110

6.7 Error analysis of GGEM . . . . . . . . . . . . . . . . . . . . . . . . . 111

6.8 Laminar channel flow . . . . . . . . . . . . . . . . . . . . . . . . . . . 113

6.9 Uniform flow around a sphere . . . . . . . . . . . . . . . . . . . . . . 115

6.10 Schematic of the immersed boundary representation of the device . . 116

6.11 Streamlines and contour plot of the streamwise velocity in a pit . . . 117

6.12 Snapshots of a hopping event . . . . . . . . . . . . . . . . . . . . . . 118

6.13 DNA dynamics at low Peclet number . . . . . . . . . . . . . . . . . . 120

6.14 Mean resident time τ v.s. Peclet number Pe. . . . . . . . . . . . . . . 121

6.15 Mean residence time τ v.s. chain length N and Peclet number Pe . . 122

6.16 DNA dynamics at high Peclet number . . . . . . . . . . . . . . . . . 124

A.1 A tethered polymer as lattice random walk. . . . . . . . . . . . . . . 133

A.2 The repulsive wall-polymer potential . . . . . . . . . . . . . . . . . . 135

B.1 Schematic of the nearest neighbor search algorithm. . . . . . . . . . . 140

C.1 Spectral analysis of a dimer in shear flow . . . . . . . . . . . . . . . . 145

Page 12: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

x

List of Tables

2.1 Properties of λ-DNA. . . . . . . . . . . . . . . . . . . . . . . . . . . 5

3.1 Spring models and the corresponding force laws. . . . . . . . . . . . 14

Page 13: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

1

Chapter 1

Introduction

Transport of polymer solutions in constricted spaces [1] is a long-standing research

topic with many applications including polymer enhanced-oil-recovery, size exclusion

chromatography [153], gel-electrophoresis [151], and recently single DNA molecule

analysis using micro- and nano-fluidic devices [140, 125]. In many fluidic devices de-

signed for separation, manipulation, and sequencing of DNA, the critical dimension

of the constriction approaches the size of the polymer molecule or smaller, and un-

derstanding the effects of spatial restrictions is essential to the rational design and

applications of those devices. In particular, capturing the interplays between polymer

dynamics, spatial confinement, and electrostatic or flow field is essential to a proper

description of these devices. It is therefore of great interest to develop validated fast

modeling and simulation tools to understand the fundamental physics of polymer

solutions in complex geometries.

There are two major types of driving forces to transport DNA molecules through

microfabricated devices: electrokinetic forces (mainly electrophoretic force) and fluid

Page 14: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

2

flow drag force. Many devices with well-defined microstructures have been proposed

for DNA manipulation, especially for separation and stretching purposes using elec-

trophoresis [45]. By contrast, much less attention has been paid on pressure (flow)

driven DNA dynamics in microfabricated devices [141, 36, 143]. There are significant

differences between the electrophoresis case and the pressure driven case. Consider

for example DNA through a slit geometry. In electrophoresis, in a uniform field, the

velocity of each DNA segment is the same everywhere in the channel; by contrast,

the unperturbed velocity profile is parabolic across the channel in the pressure driven

case at low Reynolds number, and this velocity gradient can give rise to interesting

transport phenomena, such as Taylor dispersion [141]. Furthermore, hydrodynamic

interactions (HI) between objects such as polymer segments in an unconfined domain

are long-ranged, leading to strong many-body effects, for example, Zimm scaling of the

self-diffusion coefficient of polymer. In confined geometries, the long-ranged nature

of hydrodynamic interactions changes substantially, leading to significant changes in

polymer dynamics. [1]. In a slit geometry, for example, the velocity field due to a

point force perpendicular to the walls decays exponentially. For a force parallel to

the walls, the velocity field has a parabolic form in the wall-normal direction and

decays as 1/r2. This is still a slow decay, but the symmetry of the flow leads to

cancellations upon averaging that result in screening [9, 144, 14]. Both experiments

and simulations of the diffusion of long flexible DNA molecules in slits [31, 78, 75]

are consistent with screening of hydrodynamic interactions on the scale of slit height.

Nevertheless, on smaller scales, HI are not screened and lead to changes in segment

mobilities and boundary effects such as cross-stream migration, which leads to de-

pletion layers much larger than the equilibrium chain size [77, 90, 30, 68, 29, 81]. A

Page 15: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

3

significant portion of this thesis is concerned with developing general methodologies

for Brownian dynamic simulations of DNA in complex geometries, driven by either

electrostatic field (Chapter 4 ) or flow field (Chapter 5 and Chapter 6).

This dissertation is organized as follows: In Chapter 2, we present the governing

equations for the dynamics of a single DNA molecule. Intramolecular interactions

and external potentials are introduced in Chapter 3. In Chapter 4 we present the

algorithm for DNA electrophoresis in complex geometries as applied to study the

properties of a class of soft nanomechanical control elements for microfluidics. In

Chapter 5, we discuss the Brownian dynamics simulation results of the cyclic motion

of a DNA molecule which is tethered onto a solid wall and experiences a shear flow

field. In Chapter 6, we develop and validate a method for the efficient calculation of

fluctuating hydrodynamics in complex geometries. In that chapter we also present

simulation results for the dynamics of a flow-driven DNA through a nanofluidic slit

with an embedded array of nanopits. Conclusions and future work are given in

Chapter 7.

The different sections in Chapters 4 through 6 correspond to different publica-

tions and manuscript. Those sections are therefore fairly self-contained, and some

repetition should be expected.

Page 16: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

4

Chapter 2

Problem Statement

In this dissertation, we are concerned with the dynamics of a single DNA molecule

which is confined in a fluid domain with complex boundary conditions. As we are

interested in long-time ( > 1 second) dynamics of the DNA molecule, a bead-spring

DNA model and Brownian Dynamics simulation are used in this work. A general

discussion of the wide range of techniques for modeling the hydrodynamics of polymer

chains in solution is given in Chapter 5, and a brief overview of various methodologies

for modeling hydrodynamics of polymer solutions in complex geometries is given in

Chapter 6. In the remainder of this chapter, we describe the DNA model and the

governing equations.

2.1 Bead-spring DNA Model

Double-stranded DNA is a semiflexible polymer whose physical and chemical proper-

ties have been studied extensively. The physical properties of λ-DNA, the bead-spring

model of which is used in this work, are summarized in Table 2.1. The choice of DNA

Page 17: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

5

Table 2.1: Properties of λ-DNA.

Property Symbol Value References

Number of base pares Nbp 48502 [137]

Contour length Lc 22 µm [137]

Persistence length lp 53 nm [138]

Kuhn length bk 2lp

Radius of gyration Rg 730 nm [137]

Diffusion coefficient D 0.47 µm2/s [137]

Electrophoretic mobility µ0 4.2× 10−4 cm2 /Vs [142, 102]1

model depends upon the molecule properties (chemical, mechanical, etc.) one wants

to model and the level of molecular detail one needs to retain [2]. For long time

Brownian dynamics simulation, one of the most commonly used coarse-grained mod-

els for linear DNA molecules is the bead-spring model. The molecule is modeled as

a series of Nb beads connected by Ns = Nb − 1 springs, as shown in Figure 2.1. The

total number of degree of freedom for a freely moving bead-spring chain in therefore

3Nb. Bead-spring model is the most coarse grained version of a DNA model. The

beads act as sources of fluid drag friction and the springs, which obey certain type

of spring force law, represent a collection of persistence lengths. Carefully calibrated

bead-spring chain model greatly enlarges the accessible time and length scales in nu-

merical simulation. In this work, we use the model developed by Jendrejack et al.[76],

which will be discussed in Chapter 3.

1Measured in TBE buffer (0.01M TBE, pH = 8.3, at 23 Celsius).

Page 18: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

6

O

ri

1 2

3 i

i+1

Nb-1

Nb

Figure 2.1: Schematic of a bead-spring chain in Brownian dynamics simulation.

There are two equivalent ways to describe the dynamics of a DNA or, more gen-

erally, a polymer molecule. One is based on the Fokker-Planck equation (“Diffusion

equation”) for the time evolution of the Nb beads phase space distribution function

Ψ(R, t). The other approach uses Langevin equation to describe the motion of Nb

beads. By solving the diffusion equation, one can obtain Ψ(R, t) directly. On the

other hand, with the set of coupled Langevin equations, one can directly calculate

trajectories of the beads and then obtain Ψ(R, t) by averaging. We will discuss first

the diffusion equation, and then the Langevin equation and Brownian dynamics sim-

ulation.

2.2 Kinetic theory of single polymer molecule

Kinetic theory of polymer molecule describes the dynamics of a distribution function,

i.e., evolution of the distribution function, under the action of forces. Using the

bead-spring model of polymer, the positions of the beads represent the configuration

of the polymer, and the configuration distribution function Ψ is defined so that ΨdR

Page 19: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

7

is the probability to find the position vector of the first bead within dr1 around r1 ,

second bead at dr2 around r2, and so on. The general equation for the configuration

distribution function Ψ of a bead-spring polymer is called the Fokker-Planck equation

or “Diffusion equation”. It is the derivation of this equation that we devote the

remainder of this section.

When the bead inertial relaxation times are short compared to the timescale of

interest, it is often possible to ignore inertia, that is, we can write a force balance

about bead i. The relevant forces on the bead i include: hydrodynamic force fhi ,

Brownian force f bi , and other non-hydrodynamic and non-Brownian forces fni which

is the sum of spring force, excluded volume force, and steric force between the bead

and wall. Then the force balance is written as

fhi + f bi + fni = 0, (2.1)

in which

fhi = −ζ · [ui − (u∞i + up

i )], (2.2)

f bi = −kBT∂

∂rilnΨ, (2.3)

fni = − ∂

∂riφ. (2.4)

Equation 2.2 is the Stokes drag law: the hydrodynamic force on bead i is assumed

to be linear in the slip velocity between the bead velocity ui and the solvent velocity

at the bead center (u∞i + up

i ) with the coefficient ζ . Here u∞i

is the unperturbed

velocity when the polymer is absent, and upi is the perturbed velocity due resulting

from the motion of other beads in the system. For now, we simply state that the

Page 20: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

8

perturbed velocity at ri is linearly depend on the non-hydrodynamic, non-Brownian

forces acting on all beads in the system fni as upi =

∑Nb

j=1Mij · fnj . Here Mij is a

tensor describing the hydrodynamic interaction between bead i and j.

Equation 2.3 gives the Brownian force. This equation has been derived by Bird et

al. (Ref [21]) for the case of a structureless mass point in which the chain has been

equilibrated in momentum space.

Equation 2.4 represents the force from all the non-hydrodynamic and non-Brownian

forces which stem from intra- and intermolecular potentials, and external potentials

such as electric potential and bead-wall steric repulsive potential. We will return to

a discussion of these potentials and forces in Chapter 3.

Now substitute the expressions for various forces into the force balance equation

and we obtain

− ζi · [ui − (u∞i + up

i )]− kBT∂

∂rilnΨ + fni = 0. (2.5)

Rearranging above expression, we obtain

ui = (u∞i + up

i ) +1

ζi(−kBT

∂rilnΨ + fni ) (2.6)

= u∞i +

j

(1

ζjδijI+Mij) · (−kBT

∂rjlnΨ + fnj ). (2.7)

Here, I is an identity matrix. From the first to the second line in above expression,

we use the expression for upi . According to the kinetic theory, the configuration

Page 21: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

9

distribution function satisfies the continuity equation

∂Ψ

∂t= −

i

(∂

∂ri· uiΨ). (2.8)

Defining the diffusion tensor D as

Dij =kBT

ζjδijI+ kBTMij , (2.9)

and combining 2.7 and 2.8, we obtain the diffusion equation.

∂Ψ

∂t= −

i

(∂

∂ri· (u∞

i +1

kBT

j

[Dij · (−kBT∂

∂rjlnΨ + fnj )]Ψ)) (2.10)

2.3 The Langevin equation and Brownian dynam-

ics simulation

In this section, we discuss the alternative descriptions of the dynamics of a polymer

molecule in terms of random variable dW and its defining stochastic differential

equation. This description is used in the Brownian dynamics simulation of a polymer

molecule.

The diffusion equation (Eq. 2.10) can be recast in the form of a Langevin equation

(a stochastic differential equation)[105]

dR = (U∞ +M · F+ kBT∂

∂R·M)dt +

√2B · dW, (2.11)

B ·BT = kBTM. (2.12)

Page 22: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

10

Here R is the vector containing all bead positions ri, and the vector containing

total non-Brownian, non-hydrodynamic forces acting on the beads is denoted F. The

tensor B gives the magnitude of the Brownian displacement of the polymer beads,

and is coupled to M by the Fluctuation-Dissipation theorem (Eq. 2.12). The vector

dW is a random vector composed of independent and identically distributed random

variables according to a real-valued Gaussian distribution with mean zero and variance

dt.

In Brownian dynamics simulation, applying a stochastic integration scheme to

above equation generates the trajectories of a polymer molecule under forces. For

example, using a simple forward Euler integration scheme we obtain

R(t+∆t) = R(t) + (U∞ +M · F+ kBT∂

∂R·M)∆t +

√2B ·∆W. (2.13)

In electrophoresis, DNA is a “free-draining” (FD) polymer. Therefore, we can ignore

the hydrodynamic interactions (HI) between different chain segments, and the mobil-

ity tensor M is reduced to the product of 1ζand an identity matrix. Hence, Eq. 2.13

is reduced to

R(t+∆t) = R(t) +1

ζF∆t +

2kBT

ζ·∆W. (2.14)

Above equation is relatively easy to solve compared with that in the flow-driven case.

An efficient algorithm to find the electric field at the bead position in a field with

steep gradient is essential. This issue is addressed in Appendix B.

In the flow-driven case, to capture the full fluctuating hydrodynamic in complex

geometries, we first need to turn Eq. 2.13 into a derivative-free form, as the expression

for M is lacking when the molecule is confined in complex geometries. Then, the

Page 23: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

11

stochastic differential equation can be solved based on a fast solver for Stokes flow in

complex geometries. This will be discussed in Chapter 6.

Page 24: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

12

Chapter 3

Intramolecular interactions and

external forces

The dynamics of the DNA molecule obeys Newton’s second law, and therefore, it is

determined by the total force exerted on the chain. As discussed in last Chapter,

the total force on a bead is composed of a viscous drag force fh, a Brownian force f b

due to the collisions of the solvent molecules, and all other non-hydrodynamic non-

Brownian forces fn, which includes spring forces, excluded volume forces, and any

external forces such as electrophoretic force and DNA-wall steric interaction. In this

chapter, we discuss those forces and potentials.

Page 25: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

13

3.1 Spring force law

Each spring models a fraction of the full DNA molecule which is long enough such

that its force-extension behavior obeys the Marko-Siggia spring law [26, 94],

f s =kBT

2bk[(1− |re|

Nkbk)−2 − 1 +

4|re|Nkbk

]re|re|

, (3.1)

where f s is the average force needed to get an end-to-end vector of re for each DNA

segment. The Kuhn length bk is twice the value of the persistence length, and when

a DNA molecule is divided into Nk = Lc/bk segments, each Kuhn segments can

be thought of as if they are freely jointed with each other (no bending, rotational,

or torsional potentials). This spring law was derived based on a worm-like chain

(WLC) model in polymer physics which is commonly used to describe the behavior of

semi-flexible polymers. In this model, the molecule is treated as a flexible smooth rod.

The rod’s local direction (tangential vector) decorrelates at distance s along the curve

according to exp(−s/lp), where the decay length, lp, is called the persistence length

of the chain. The stiffer the chain, the larger the persistence length. For DNA, the

persistence length is approximately 50 nm. The WLC model fits DNA force-extension

data very well up to 10 pN, or < 95% stretching ratio [26]. As long as the number of

Kuhn lengths contained in each spring is larger than 10bk, the Marko-Siggia spring

law can be used [149]. In this work, each spring represents 20bk.

There are other spring force laws proposed for simulation of polymer molecules,

which are summarized in Table 3.1, together with the WLC model.

Page 26: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

14

Table 3.1: Spring models and the corresponding force laws.

Model Force expression

Hookean f s = Hre =3kBTNkb

2k

re

FENE f s = Hre

1−(|re|/Nkbk)2

Inverse Langevin model f s = kBTbk

L−1( |re|Nkbk

) re

|re|

L(x) = cothx− 1x

WLC f s = kBT2bk

[(1− |re|Nkbk

)−2 − 1 + 4|re|Nkbk

] re

|re|

3.2 Excluded volume interactions

In polymer physics, excluded volume interaction is a type of long-ranged1 interaction

referring to the idea that in any real polymer molecule, two monomers cannot occupy

the same space. This effect plays a far more important role in polymer solution

than it does in solution of small molecules [35]. We often idealize the chain to allow

overlap of monomers, and call it an ideal chain. By contrast, a regular chain with

excluded volume interactions is called a real chain. Ideal chain does not exist in

reality, but it is used extensively because it allows rigorous mathematical treatment

of many questions. More importantly, the real chain behaves like an ideal chain in

some situations, such as when the molecule is in concentrated solutions, melts, and

in theta condition. In Appendix A, we derive an effective repulsive potential between

a polymer molecule and a hard wall using the ideal chain model.

In Brownian dynamics simulation for λ-DNA, we use an exponential form of the

1Note that excluded volume interaction is long-ranged along the chain backbone. The potentialis actually short-ranged in space (Eq. 3.2).

Page 27: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

15

repulsive excluded volume potential between bead i and bead j, when we consider a

real chain. It is obtained by considering the energy penalty due to the overlap of two

Gaussian coil [41],

Uevij =

1

2νkBTN

2k,s(

3

4πS2s

)3/2 exp(−3|ri − rj|2

4S2s

), (3.2)

where ν is the excluded volume parameter, and S2s = Nk,sb

2k/6 is the mean square

radius of gyration of an ideal chain containing Nk,s Kuhn segments of length bk.

The resulting expression describing the force acting on bead i due to the the

presence of bead j within some cut-off distance is then

f evij = νkBTN2k,sπ(

3

4πS2s

)5/2 exp(−3|ri − rj|2

4S2s

)(ri − rj), (3.3)

Although this potential is not self-consistent (any deformation of the coil caused by

the overlap has been ignored), it does provide the correct scaling relationships [76].

3.3 Hydrodynamic interactions

Beads immersed in a fluid generate flows as they move due to various forces, and

similarly they move in response to fluid motion through the Stokes drag. In this

work, we treat bead i of the bead-spring chain as a sphere of hydrodynamic radius

a. Then the relationship between the bead velocity ui and the drag force it exerts on

the fluid is given by the Stokes law fi = ζ(ui − u(ri)), where ζ is the bead friction

coefficient ζ = 6πηa, u(ri) is the fluid velocity at the bead position, and η is the fluid

viscosity. (Note that the finite size of the bead only arises in the friction coefficient.)

Page 28: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

16

Through these hydrodynamic interactions (HI), beads interact with each other and

with the walls of the confining geometry. In kinetic theory, the velocity perturbation

due to the polymer molecule, up, is generally taken to be due to a chain of point forces

acting on the fluid, and obtained by solving the incompressible Stokes flow problem,

−∇p(x) + η∇2up(x) = −∑

fnδ(x− xn), (3.4)

∇ · up(x) = 0, (3.5)

up(r) = 0, r ∈ ∂Ωb, (3.6)

where p is the pressure, up is the fluid velocity, η is the fluid viscosity, fn is the force

exerted on the fluid at point xn, δ(x) is the three-dimensional delta function, and ∂Ωb

is the boundary of the fluid domain. In the actual implementation, regularized point

forces are used, as further discussed flow. The velocity perturbation at the position

of bead i due to the bead j is represented in a Greens function form as

upi =

Nb∑

j=1

Mij · fnj (3.7)

where Mij is called hydrodynamic interaction tensor. To illustrate the components

of the mobility tensor, we consider a two-bead chain (a dumbbell) in an unbounded

domain, neglecting Brownian effects for the moment. Here the force balance shows

that the fluid velocities u(ri) experienced by bead i are given by

u1

u2

=

u∞1

u∞2

+

1ζI G∞(r1 − r2)

G∞(r2 − r1)1ζI

·

f1

f2

, (3.8)

Page 29: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

17

where G∞(r) = 18πηr

(I + rr/r2) is the Oseen-Burgers tensor, the free-space Green’s

function for Stokes’ equations, r = |r|, and I is the identity matrix. The quantity

G∞(r1 − r2) · f2 gives the velocity at position r1 generated by the force f2 exerted by

the particle at r2 on the fluid. We can write above expression succinctly as,

U = U∞ +M · F. (3.9)

We note that ∂∂r

·M = 0 for G∞. In simulation of coarse-grained polymer chain, the

singularity of the Oseen tensor must be regularized. The most common choice is the

RPY tensor [120, 157].

In a confined geometry, a wall correctionGW (r1, r2)has to be added to the mobility

tensor M, and it becomes

M =

1ζI+GW (r1, r2) G∞(r1 − r2) +GW (r1, r2)

G∞(r2 − r1) +GW (r1, r2)1ζI+GW (r1, r2)

. (3.10)

In general geometries, an analytical expression for GW (r1, r2) is not available, but

has to be calculated numerically. We note that it is not M that is needed, but rather,

the product M · f which is the velocity generated by the total non-Brownian, non-

hydrodynamic forces acting on the beads. As a result, using an accelerated Immersed

Boundary Method, we can calculateM·f for polymer in complex geometries efficiently.

We will return to a discussion on it in Chapter 6.

In Green’s function-based representation of the hydrodynamic interactions, we

assume that the beads are coupled instantaneously. In other words, the characteristic

time for the hydrodynamic interactions to propagate through the system should be

Page 30: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

18

much smaller than the time scale of other signals. These conditions are equivalent to

that the Reynolds number is small Re ≪ 1 and the March number is small Ma ≪ 1,

which are satisfied explicitly through the use of the Stokes equation.

3.4 Electrophoretic force

Electrophoretic force is commonly used to transport and manipulate DNA molecules.

In this section, we discuss DNA electrophoresis and the effective charge assigned to

each bead calculated using the electrophoretic mobility of a long DNA molecule.

The velocity of a DNA molecule (> 100 bp) in free electrophoresis has little depen-

dence on length [142]. When an electric field is applied across an electrolyte solution

containing DNA molecules, the counter-ion cloud around the DNA experiences a force

in the opposite direction to that of the DNA. This force generates ab electroosmotic

flow which balances the flow perturbation generated by the DNA molecule. Hence,

we can say that hydrodynamic interactions are screened and DNA is “free-draining”

during free electrophoresis. Every segment feels the same friction from the fluid, so

DNA molecules of different sizes will have the same mobility.

We consider a DNA molecule subject to an external electric force E. The free

electrophoretic mobility of DNA µ0 is defined as the ratio between the electrophoretic

velocity v and electric field strength E. For a “free-draining” DNA bead-spring chain,

the total friction coefficient of the chain is the sum of all the friction coefficients ζ of

the beads. Likewise, the total electric charge of the chain is Q = Nbqb where qb is

the effective charge per bead. Balancing the electric force Fe = QE = NbqbE and the

Page 31: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

19

friction force Fh = Nbζv, results in the following center of mass velocity

v =NbqbE

Nbζ=

qbE

ζ. (3.11)

Hence, the electrophoretic mobility of DNA in free-solution (µ0) is given by

µ0 ≡v

E=

qbζ. (3.12)

This expression indicates that the electrophoretic mobility of DNA in free solution is

independent of the length or the molecular weight of the chain. For a “free-draining”

chain with diffusivity D, the friction coefficient is ζ = kBT/NbD. Therefore, given

the electrophoretic mobility, the effective charge per bead is

qb = ζµ0 =kBT

NbDµ0. (3.13)

Using the diffusion coefficient of λ-DNA during free solution electrophoresis D = 0.55

µm2/s, and the electrophoresis mobility obtained under the same condition µ0 =

4.2×10−4cm2/Vs [102], qb is determined to be 178 electrons per bead, for the Nb = 11

bead-spring model used in this work.

3.5 Polymer-wall steric interaction

A potential needs to be devised to prevent penetration of beads through solid walls.

However, if the potential is steep near a wall, such as that of a Lennard-Jones po-

tential, a very small time step should be used. On the other hand, in the case of

Page 32: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

20

soft potential, a bead can sometimes penetrate the wall. Jendrejack et al. devised a

potential barrier for bead-wall interactions which is harder than the Gaussian soft po-

tential used for bead-bead interaction and softer than typical Lennard-Jones potential

[77],

Uwalli =

Awallb−1k δ−2

wall(hi − δwall)3, hi < δwall

0, hi > δwall.

(3.14)

Here hi represents the perpendicular distance of bead i from the wall, δwall is the

cut-off distance.

The choice of the repulsive potential is arbitrary as long as it can prevent the

bead from penetrating the wall with reasonably small time step. In Appendix A, we

derive an effective repulsive potential between a polymer molecule and a hard wall

using the ideal chain model.

Page 33: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

21

Chapter 4

Bistability and field-driven

dynamics of confined tethered

DNA

DNA is “free-draining” during electrophoresis as discussed in Chapter 3. As a con-

sequence, it cannot be separated by free electrophoresis. A sieving medium, such

as polymer gel or microfabricated structure, that induces a size-dependent mobil-

ity is necessary for DNA separations. Many microfluidic devices with well-defined

structures have been proposed for DNA manipulation, especially for separation and

stretching purposes using electrophoresis [45]. These devices are typically composed of

periodic obstacle arrays. When an electrostatic field is introduced, a nonuniform field

with steep gradient is generated in the device. Thus, understanding DNA dynamics in

nonuniform electric fields is an important issue for the design of high performance mi-

Page 34: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

22

crofluidic devices. The motivation of the work presented in this chapter1 is to build a

general framework to analyze DNA dynamics in nonhomogeneous electrostatic fields

in complex microfluidic geometries. A Brownian dynamics/Finite element method

algorithm is presented in Appendix B. In the remainder of this chapter, we focus on

its application to study the properties a novel class of soft nanomechanical control

elements we proposed for microfluidic devices.

4.1 Introduction

A fundamental unit in many engineering systems is an element with two distinct

states – on/off, open/closed, left/right etc. Of particular interest and utility in many

cases is the property of bistability: each state is robust in the sense that a finite

perturbation must be temporarily introduced to switch the system from one state to

the other. Many classical macroscopic mechanical systems display bistability, and ef-

forts have recently been made on many fronts to design, construct and analyze “soft”

bistable systems on smaller and smaller scales, down to the molecular level. Steen

and coworkers [70, 152] have constructed systems of pairs of droplets connected by

a flow channel. This system exhibits bistability via the existence of situations where

two configurations with minimal surface energy can arise, with one drop large and the

other small, or vice versa. Switching between states can be achieved, for example, via

pressure fluctuations or electroosmotic pumping of fluid between droplets. Groisman

et al. [63] and Arratia et al. [10] have demonstrated microfluidic systems in which

elastic liquids (solutions of long flexible polymer molecules) in complex geometries

1This chapter is based on the publication: Y. Zhang, J. J. de Pablo, and M. D. Graham, SoftMatter, 5, 3694.

Page 35: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

23

can exhibit bistability via symmetry-breaking flow instabilities. At the level of single

flexible molecules in solution, bistability arises through coil-stretch hysteresis in suffi-

ciently long DNA molecules in solution during extensional flow [126]. (We will use the

term “bistable” loosely here, to denote situations where the energy barrier between

two metastable states is greater that about kBT , where kb is Boltzmann’s constant

and T is absolute temperature.) A similar phenomenon is predicted for a chain teth-

ered on a no-slip wall, in the case where there is a stagnation point flow and the chain

tether point is at the stagnation point [18]. Banavar et al. [15] used simulations of

model chain molecules to predict bistability in the conformations of the molecules.

In particular, they predict for a certain case the coexistence of a single helix and a

dual helix in which the chain folds in half and self-interacts to form a double helix.

Our interest here is in bistability of polymer chain conformations associated with the

combination of tethering and confinement within a micro- or nanoscale device.

Partially confined configurations of polymer chains, where different parts of a

chain experience different degrees of confinement, give rise to interesting phenomena of

entropic origin. The entropy difference between a portion of polymer chain in confined

space and in open space gives rise to an entropically induced recoil force, which tries to

pull the whole chain into the open space. Using a nanochannel/microchannel device,

Mannion et al. estimated this entropically induced force of double-stranded (ds-)

DNA to be about 102 − 103fN [92]. According to the force versus extension curve of

ds-DNA, a force of this magnitude will generate a fractional extension of the molecule

of about 0.5 [94]. In a related calculation, Bickel et al. determined the force required

to tether an end of an ideal chain onto a hard wall, which is also a force due to the

decrease of available number of configurations [20]. This tethering force converges to

Page 36: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

24

a constant value kT/lk as chain length diverges. For DNA with a Kuhn length of

lk ≈ 100nm, this force is about 40 fN at room temperature. Partial confinement

has been exploited to manipulate a DNA molecule using a nanoslit device (50 - 100

nm across), on which one wall had square nanopits (100 nm deep and 100 - 500 nm

wide)[116]. Because the confinement is weaker in the regions of the slit where there

is a pit, DNA segments are preferentially found there. This entropic well tends to pin

sections of the chain in these regions. When multiple pits are present (and not too

far apart), stretched sections of DNA will span the more confined regions between

pairs of pits. Binder and coworkers have used Monte Carlo simulations and theory

to study a tethered polymer in good solvent compressed by a circular disk centered a

distance H above the tether point [95, 71]. For moderate compression, the polymer is

“imprisoned” underneath the disk adopting a quasi-2D random walk configuration by

forming blobs with diameter equivalent to the disk height H . When a certain height

Himp, part of the chain “escapes” this imprisonment and forms a “stem-and-flower”

configuration where the blobs underneath the disk are stretched into an extended

configuration by the entropic force exerted by the escaped parts. Using Monte Carlo

simulations, they found that there is a regime of heights Hesc ≤ H ≤ Himp where the

chain exhibits “two-phase coexistence” (bistability) of the escaped and the imprisoned

configurations. Both simulation results and analytical theory indicate that the escape

transition is a first-order phase transition. This provides an example of bistability

created by a combination of tethering and partial confinement, specifically by the

competition between the entropy gain of the escaped segments and the entropy loss

and energy penalty of extending the portion of the chain that remains trapped under

the disk.

Page 37: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

25

Partially confined configurations also lead to interesting dynamics [103, 99]. Trapped

within a two-dimensional array of spherical cavities interconnected by circular holes,

short linear DNA strongly localize in cavities and only sporadically “jump” through

holes [103]. To jump out of a cavity, small DNA segments penetrate into the neighbor-

ing cavity first and form a partial confined configuration. The fluctuation of the prob-

ing segments beyond a threshold leads to an abrupt jump of the entire molecule. This

is in accordance with previous theoretical studies of entropic barriers by Muthukumar

[100, 99].

In the present work, we examine, using Brownian dynamics simulations of a coarse-

grained model for long flexible DNA molecules, the possibility of generating bistable

behavior using long flexible linear polymer molecules end-tethered in a confined ge-

ometry. Fig. 4.1(a) shows a very simple example of what we will call an “entropically

bistable” system. (The system studied by Binder and coworkers [95] is also in this

class.) Here a polymer molecule is tethered within a small pore of width 2d connect-

ing two open regions. For a sufficiently long chain (Rg ≫ d, where Rg is the radius

of gyration of the chain in the absence of confinement), one clearly expects that the

most entropically favorable situations will be those where the entire chain is in one

open region or the other. Thermal fluctuations or transient application of an electric

field (if the chain is a polyelectrolyte) or flow field might drive the chain from one

region to the other over an entropic barrier. One realization of this type of system

would be the fully three-dimensional situation where the pore is a hole connecting

two half-spaces, as shown in Fig. 4.1(d). A more experimentally tractable version,

fabricated using nanolithography techniques, is the “quasi-two-dimensional” (Q2D)

situation shown in Fig. 4.1(c), where top and bottom bounding walls exist. Both

Page 38: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

26

z y

xh=2d

r

z

y x

(c) (d)

l

2d

(a)

thermalnoise

transientf ie ld

z

y

x

(b)

mediumfield

z

y

x 2d

nofield

highfield

or

Figure 4.1: Schematic representations of soft nanomechanical bistable elements. (a) Pore-crossing geometry:an entropically bistable system. (b) Pore-entry geometry: a competitively bistable system. (c) Quasi-2D pore-crossing geometry. (d) 3D pore-crossing geometry.

Page 39: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

27

situations will be studied here.

We also consider the possibility of systems where the origin of bistability is some-

what more subtle. Consider the system shown in Fig. 4.1(b), comprising a polyelec-

trolyte end-tethered at the mouth of a long pore. In the absence of an electric field,

the most likely conformations of the chain lie outside the pore (left). If a sufficiently

large field is applied pulling the chain into the pore, entropy will be overcome by

electrostatic energy and the chain will reside entirely in the pore (right). We hypoth-

esize that at intermediate field strengths, a nontrivial competition between entropy

and electrostatic energy will allow the possibility of two stable states (center), one

where the chain is mostly outside the pore, in a high entropy state, and the other

where the chain is mostly inside the pore, in a low entropy state. This system will

be called “competitively bistable”. We will study the quasi-two-dimensional version

of this geometry.

Many fundamental issues arise when considering these systems. What is the

nature of the transition states in these systems? How high are the free energy barriers

between states? How do these depend on the length of the chain and, in the Q2D

case, the degree of lateral confinement? What are the dynamics of transitions between

the states? How will those dynamics differ in the electrostatically driven and flow

driven cases? To what extent do the dynamics of these systems follow Kramers-type

kinetics? This initial report will only begin to touch on some of these issues.

Potential applications of these systems can also be considered. A chain tethered in

a micro/nanochannel under sufficiently confining conditions would behave somewhat

like a porous medium, resisting fluid flow and blocking the transport of sufficiently

large solutes. This property in combination with entropic bistability might be ex-

Page 40: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

28

(a) (b)

TransientFieldTransientField

TransientFieldTransientField

z

yx

z

yx

Figure 4.2: Possible applications of tethered polymer molecules as control elementsin micro/nanofluidic devices. (a) Switch. (b) Gate.

ploited as shown in Fig. 4.2. In the left image, a polyelectrolyte chain is configured so

that it blocks transport of the round solute particles in the left channel, while allowing

transport of the square ones, or vice versa. Switching would occur through transient

application of an electric field in the z direction. In the right image, an entropically

bistable geometry serves as a gate that can be opened or closed by a transient electric

field.

4.2 Polymer model and simulation approach

We use Brownian dynamics simulation of a model of long ds-DNA molecules to study

the statics and dynamics of a tethered chains in confined systems. Following previous

work [76, 74], a double-stranded DNA molecule is described by a bead-spring chain

model composed of Nb beads of hydrodynamic radius a = 77 nm connected by Ns =

Nb − 1 entropic springs. Each bead represents a DNA segment of 4850 base pairs,

i.e., Nb = 11 corresponds to a stained λ-DNA, which has a contour length of 22 µm

Page 41: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

29

2d=14 a

l=2 a

(a) (b)

z

y

x

ze

ze

2d=14 a

Pore crossing Pore entry

Figure 4.3: Top view of the simulation domains. Black areas are impenetrable walls;the free end of the chain is a filled circle. In quasi-2D cases, the tethered bead-springchains are bounded in x direction by walls at x = −d and x = +d. (a) Pore-crossinggeometry. (b) Pore-entry geometry.

and radius of gyration of 730 nm. An effective charge of Q = 178 e is assigned to

each bead, which is calculated based on the free solution mobility of DNA [142, 102].

The length unit in this work is a = 77 nm. The springs connecting the beads obey a

worm-like chain force law [94]

Fsij =

kBT

2bk

[

(1− |rj − ri|Nk,sbk

)−2 − 1 +4|rj − ri|Nk,sbk

]

]

rj − ri|rj − ri|

. (4.1)

Here, bk is the Kuhn length for DNA and Nk,s = 20 is the number of Kuhn length

per spring. The physical confinement is taken into account through an empirical

bead-wall repulsive potential of the form

Uwalli =

Awallb−1k δ−2

wall(hi − δwall)3, hi < δwall

0 hi > δwall.

(4.2)

Page 42: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

30

Here hi represents the perpendicular distance of bead i from the wall, δwall is the

cut-off distance. In this work, we choose Awall = 50kBT/3 and δwall = bkN1/2k,s /2 =

3.01a = 0.24 µm. In this work, we ignore hydrodynamic interactions (HI) between

chain segments and assume that the chain is ideal. The force balance on the beads

leads to a stochastic differential equation[105]:

dR =1

ζFdt+

2kBT

ζdW, (4.3)

where R is the vector containing bead positions ri, ζ is the friction coefficient of

the bead and F is the vector of non-hydrodynamic and non-Brownian forces. The

components of dW are obtained from a real-valued Gaussian distribution with mean

zero and variance dt. We use a standard (stochastic) Euler scheme[105] for time-

integration.

In the pore crossing system (Fig. 4.3 (a)), we study the dependence of the free

energy barrier on chain length and confinement in both a quasi-2D and 3D geometries.

The wall thickness is 2 a and the pore size is 14 a. For the quasi-2D case (Fig. 4.1 (c)),

the chain is confined in the x direction, and the distance between the two impenetrable

planes are h = 14 a. For the 3D case (Fig. 4.1 (d)), the chain is unbounded in both

x and y directions.

In the pore entry system (Fig. 4.3 (b)), we consider only the quasi-2D case. The

chain is tethered to the center of the entrance to a square channel which runs along

the +z axis. An electrostatic field is applied along the −z direction. In the present

model electrostatic interactions between beads are neglected, consistent with behavior

in a high ionic strength solvent, where electrostatic interactions are screened. Elec-

Page 43: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

31

troosmosis of counterions screens hydrodynamic interactions, as does confinement,

justifying their neglect in the present simplified model. The electric field strength E

at z = −∞ is E = E∞ez. The relative field strength is denoted as ER = αE∞/E0,

where E0 = kBT/aQ = 18.74 V/cm is the field strength unit and α is a constant

to make ER order O(1). When ER = 1, the corresponding E∞ = 1.23 V/m, and

the electric force fE = E∞Q = 3.51 × 10−2 fN. Even inside the channel where the

electric field gradient is around ten times larger than E∞, the electric force is very

small compared with thermal agitation kBT/a = 53.4 fN. The electric field is obtained

by solving the governing Laplace equation with no flux boundary conditions on the

channel walls and Dirichlet boundary conditions at the left and the right edges of the

domain. The commercial PDE solver COMSOL, based on the finite element method

(FEM), is used. When d = 7a, the simulation boxes are y × z = 200a × 100a and

14a× 50a for the outer chamber and the channel, respectively. As the field gradient

becomes uniform and constant going along the channel, we can use a short simula-

tion box for the channel. We use the Lagrange-quadratic square elements (1a × 1a)

provided by COMSOL. As the polymer model is coarse-grained to a length scale of

a, this resolution of the field is adequate.

4.3 Results and discussion

4.3.1 Pore-crossing geometry

We consider first the pore-crossing geometry of Fig. 4.3(a). Only the equilibrium

behavior in this system will be studied. Fig. 4.4(a) is a typical time series plot of the

z component of the free end ze of a chain with Nb = 15 in the 3D case. This plot

Page 44: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

32

0 500 1000 1500 2000-4

-2

0

2

4

t (s)

z e (m

)

tw

(a)

1070 1080 1090 1100-4

-2

0

2

4

t (s)

z e (m

)

(b)

0 1000 2000 3000 4000 50000.0

0.1

0.2

0.3

(t w)

tw(s)

tw / exp(-tw/

(c)

Figure 4.4: Simulation results for an ideal chain tethered to the center of the porein a 3D pore-crossing geometry (Fig. 4.3(a)). (a) Typical time series plot of the zcomponent of the free end of a tethered chain with Nb = 15 and d = 7a. (b) Blowupof a segment of the time series plotted in (a). (c) Waiting time distribution and bestfit to an exponential distribution.

Page 45: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

33

(a) (b) (c)

(d) (e) (f)

z( m)m

y(

)m

m

−3 −2 −1 0 1 2 3−2

−1

0

1

2

−3 −2 −1 0 1 2 3−2

−1

0

1

2

−3 −2 −1 0 1 2 3−2

−1

0

1

2

−3 −2 −1 0 1 2 3−2

−1

0

1

2

−3 −2 −1 0 1 2 3−2

−1

0

1

2

−3 −2 −1 0 1 2 3−2

−1

0

1

2

Figure 4.5: Snapshots of a crossing event in the quasi-2D pore-crossing geometry.Time interval between frames is 0.5s.

clearly shows that ze fluctuates around two metastable states. Fig. 4.4(b) is a blowup

of the time interval 1070-1100 s. The chain end stays on one side of the wall for a

time interval tw, which we will call the waiting time, and then crosses to the other

side. Fig. 4.5 shows the dynamics of a typical crossing event. The free end of the

tethered chain first diffuses to the pore and then through to the other side of the wall.

We notice that this step of end bead crossing does not necessarily cause a crossing

of the whole chain. A successful crossing event occurs when the number of beads on

one side of the wall changes from Nb − 1 to 0 or vice versa. The time points tic when

a crossing completes are recorded. We define the waiting time as the time between

two adjacent crossing events tiw = ti+1c − tic. As seen in Fig. 4.4(a), the crossing time

is very small compared to the waiting time, which is well-fit to an exponential with

chain-length-dependent decay rate 1/τ , as shown in Fig. 4.4(c). This result indicates

Page 46: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

34

that, to a good approximation, the crossing events are independent of one another –

crossing is a Poisson process.

To gain further information about the crossing process and the transition state,

we consider the joint behavior of the z-component of the chain end, ze and the z-

component of the center bead, zm (where m = Nb/2 and (Nb + 1)/2 when Nb is

even and odd respectively). The joint probability density for these two variables is

denoted ρ(ze, zm). Fig. 4.6(a) shows − ln ρ(ze, zm) for a chain with Nb = 13 chain in

a 3D pore-crossing geometry. It clearly demonstrates that there are two metastable

states (maxima in ρ(ze, zm)), in which both ze and zm are on the same side of the

wall, and two transition states when the end bead and the center bead are located

on different sides of the wall. The transition states are in accordance with what we

observed in the snapshots shown in Fig. 4.5.

A simpler representation of the free energy surface explored by this system is

given simply by the probability density for the z-component of the end bead ρ(ze).

Unlike the joint distribution ρ(ze, zm), this representation is too low-dimensional to

provide insight into the structure of the transition state. On the other hand, it is

simple to construct and visualize, and a simple potential of mean force (PMF) can

be constructed as

F (ze) = F ∗ − kBT ln

[

ρ(ze)

ρ∗

]

, (4.4)

where F ∗ and ρ∗ are arbitrary reference values. In Fig. 4.6(b) we show ρ(ze) (dashed

circle) and corresponding F (ze) (solid line) when F ∗ = Fwell = 0 (Nb = 20) in a

quasi-2D system . As expected, the PMF is composed of two energy wells which

are separated by a energy barrier – the difference in energy between these is denoted

∆Fmax. Fig. 4.6(c) shows a plot of the energy barrier ∆Fmax/kT as a function of

Page 47: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

35

-0.50

-1.0

-1.5

-2.0

-2.5

-3.0

-3.0

-3.5

-3.5

-4.0

-4.0

-4 -2 0 2 4-3

-2

-1

0

1

2

3

ze( m)

zm(

m)

-5.0-4.5-4.0-3.5-3.0-2.5-2.0-1.5-1.0-0.500

(a)

-3 -2 -1 0 1 2 30

1

2

3(b)

ze( m)F(

z e)/kT

Fmax

10 15 20 25 300

1

2

3

4

5

6

7

Fmax kT a ln(Nb)

a=2.58 0.12

a=1.41 0.03

(c)

F max

/kT

Nb

3D Q2D

10 15 20 25 30

10

100

1000(d)

4.16 0.14

2.45 0.05

3D Q2D

Nb

(s)

b

Figure 4.6: Simulation results for an ideal chain tethered to the center of the pore inquasi-2D and 3D pore crossing geometries (Fig. 4.3(a)). (a) The negative of logarithmof the joint probability density function of the z-component of the end bead ze and thez-component of the center bead zm for a Nb = 13 chain in 3D. (b) Reduced potentialof mean force ∆F(ze)/kBT for a Nb = 20 chain in Q2D. (c) Semilog plot of reducedenergy barrier vs. chain length when d = 7a (d) Log-log plot of mean first passagetime τ v.s. chain length when d = 7a. The error bars are about the same size as thesymbols.

Page 48: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

36

Nb for the 3D and quasi-2D cases. For the range of Nb considered it appears that

∆Fmax/kT = a lnNb.

Fig. 4.6(c) shows the dependence of the mean waiting time τ on molecular weight.

Over the limited range of molecular weights studied, the results appear to follow a

power law: τ = τmNαb . If the transition process follows Kramers kinetics [160], one

would expect that τ = τ0 exp(∆Fmax/kbT ), which in the present case would imply

that τ0 = τmNα−ab . If hydrodynamic interactions were included in the model, one

might expect α to decrease slightly, based on simulation results for translocation of

a free polymer chain through a pore ([8, 73, 56, 54]). Since a is determined from the

free energy landscape of the system, it is unaffected by the presence or absence of

hydrodynamic interactions in the model. On the other hand, the pore geometry (size,

shape) might affect both these parameters; in particular changing the pore size will

change the free energy landscape and thus the exponent a. The nature and origin of

the observed exponents will be the subject of future work; we note that extension of

Kramers’ theory of barrier crossing of a point particle to a polymer system is a topic

of some current research [106, 129, 128].

Finally, we observe that the simulation results indicate that confinement from 3D

to quasi-2D greatly reduces the energy barrier and leads to a larger transition rate

(Fig. 4.6(c) and (d)). As the crossing rate is inversely proportional to the free energy

barrier, which is closely related to the number of accessible chain configurations, this

is not surprising by considering that the number of accessible chain configurations in

the quasi-2D case is greatly reduced compared to that in the 3D case.

Page 49: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

37

4.3.2 Pore-entry geometry

The previous section reported a system where geometry alone determines the free

energy landscape of the tethered polymer. The present section turns to a case where

entropy and electrostatic energy compete, and address the possibility that bistability

of a tethered chain can arise from this competition. The situation under consideration

is the pore-entry geometry of Fig. 4.3(b). Only the quasi-2D case will be considered.

In the absence of a field, a chain with Rg ≥ d would not be expected to sample the

pore, while if the field is sufficiently large it is expected that the entire chain will

reside in the pore. The regime of interest is chain lengths such that Rg ≥ d and

intermediate values of the field strength.

Fig. 4.7(a) shows a time series of ze for the case Nb = 10 and ER = 5. As in the

pore-crossing geometry, bistability is observed, which can be seen more clearly from

the joint probability density function ρ(ze, zm) illustrated in Fig. 4.7(b). The path

connecting the two metastable states indicates, again, the transition state. Fig. 4.8(a)

shows PMF curves based on ρ(ze) as functions of ER when Nb = 10. As the electric

force is small in the outer chamber, it does not significantly contribute to the energy

of the chain when ER . 10.0 and the energy profile outside the channel is insensitive

to the change of ER. The energy well inside the channel is created by a balance

between the entropic force and the electric force. We denote the critical field strength

above which this energy well appears as Elow (≈ 2.2 for Nb = 10). As ER increases

further, the free energy barrier for entry to the pore weakens, and eventually vanishes

once ER exceeds a field strength denoted Ehigh When Elow < ER < Ehigh, the system

shows competitive bistability that is created by the interplay between entropy and

electrostatic energy. Fig. 4.8(b) shows the positions of the minima in Fig. 4.8(a) as

Page 50: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

38

0 20 40 60 80 100

-2

-1

0

1

2

3

4(a)

t(s)

ze(

m)

-3 -2 -1 0 1 2 3 4-3

-2

-1

0

1

2

3

4

zm(

m)

ze( m)

(b) -4.00-3.50-3.00-2.50-2.00-1.50-1.00-0.5000

Figure 4.7: Simulation results for a tethered ideal chain in the quasi-2D pore-entrygeometry (Fig. 4.3(b)) for the case Nb = 10, ER = 5, d = 7a. (a) A typical timeseries plot of the z component of the free end of a tethered chain. (b) The negativelogarithm of the joint probability density function ρ(ze, zm).

Page 51: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

39(c) (d)

(e) (f)

(g) (h)

-2 0 2 4 6 8 10 12 14-1

0

1

2

3

g

h

f

e

d

c

(b)

Elow

z e0(

m)

ER

OUT IN

Ehigh

z( m)m

y(

m)

m

−2 −1 0 1 2 3−2

−1

0

1

2

−2 −1 0 1 2 3−2

−1

0

1

2

−2 −1 0 1 2 3−2

−1

0

1

2

−2 −1 0 1 2 3−2

−1

0

1

2−2 −1 0 1 2 3

−2

−1

0

1

2

−2 −1 0 1 2 3−2

−1

0

1

2

-3 -2 -1 0 1 2 3 4-6-5-4-3-2-101234

E0 E1 E2.2 E3 E4 E5 E7 E9

ze0

(a)

ze( m)

F(z e)/k

T

Figure 4.8: Simulation results of an ideal chain tethered to the center of the entrance in a quasi-2D pore entrygeometry, with Nb = 10. (a) Potential of mean force as function of last bead position ze and reduced electric fieldgradient ER. Top to bottom curves correspond to ER = 0, 1.0, 2.2, 3.0, 4.0, 5.0, 7.0, and 9.0, respectively. (b)Most probable end bead positions ze0 (i.e., maxima of ρ(ze) or minima of ∆F (ze)) as a function of ER for state”IN” and state ”OUT”. The error bars are almost the same size as the symbols. (c)(d),(e)(f),(g)(h) are snapshotsat ER = 2.3, 5, and 9, respectively.

Page 52: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

40

0.00 0.02 0.04 0.06 0.08 0.10 0.120

2

4

6

8

10

12

14

State "IN"+"OUT"

State "OUT"

State "IN"

Elow

Ehigh

(a)

ER

1/Nb

0.0 0.5 1.0 1.5 2.0 2.5 3.00.00.10.20.30.40.50.60.70.80.91.0(b)

Pin

ER/E1/2

Nb=10 Nb=15 Nb=20 Nb=25

Figure 4.9: (a) Phase diagram showing the region in parameter space where the chainshows bistability in the pore-entry geometry. (b) The probability to find the chaininside the channel Pin as function of ER normalized by E1/2.

functions of ER – the regime of bistability is clearly seen. Fig. 4.8(c) shows typical

conformations of the chain at various points on Fig. 4.8(b).

Finally, we address the phase transition behavior of the tethered polymer molecule

in the competitively bistable system. In Fig. 4.9(a), the shaded area shows the param-

eter regime in ER and 1/NB where bistability occurs. Fig. 4.9(b) shows the probability

Pin =∫∞0

dzeρ(ze) of the chain end residing inside of the channel, as a function of ER

for various Nb. We have normalized ER with E1/2, the (molecular-weight dependent)

field strength where Pin = 0.5. From these plots it appears that as N → ∞, the

slope diverges and the region of bistability vanishes, indicating a first-order phase

transition. Similar behavior was noted in the escape transition system studied by

Binder and coworkers [95].

Page 53: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

41

4.4 Conclusions

Simulations predict that flexible polymers end-tethered in a pore within a confined

geometry can display multiple free energy minima separated by substantial barriers

– bistability. For a tether point in a pore between two large chambers, bistability is

expected based on simple considerations of entropy and geometry -we denote this as

“entropic bistability”. In this situation dynamic simulations suggest that the crossing

of the chain end through the pore is the dominant mode of transition between the two

states, and consideration of the joint probability density function of the position of the

chain end and the chain midpoint indicates the presence of a transition state where

the chain end is on one side of the pore, while the chain midpoint is on the other. For

a chain end-tethered at the mouth of a long pore and subjects to a field that tends to

draw the chain into the pore, we predict the existence of a second kind of bistability,

arising from the interplay of conformational entropy and electrostatic energy; we

call this “competitive bistability”. In particular, simulations show the presence of a

regime of field strength in which there are two free-energy minima, one corresponding

to a chain that is largely outside the pore in an entropically favorable state, and one

where the chain is largely inside the pore, in an electrostatically favorable state.

Both fundamental and technological question are raised by these results. The

systems under consideration here comprise model systems for activated rate processes

in systems with many degrees of freedom and provide an opportunity to compare

experiment, theory and simulations for this class of systems. Of particular interest

may be the difference in dynamics when transition from one state to another is driven

by fluid flow rather than an electrostatic field. These systems or variations of them

might also have applications as electrically or fluidically actuated nanomechanical or

Page 54: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

42

nanoelectromechanical elements, such as valves or switches.

Page 55: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

43

Chapter 5

Tethered DNA dynamics in shear

flow

In Chapter 5 and Chapter 6, we discuss flow-driven DNA dynamics. To study the

effects of solid impenetrable walls on the dynamics of a nearby DNA molecule, we

start by re-examining a problem with simple geometry: the cyclic dynamics of a

tethered DNA molecule in shear flow where the chain is grafted to a flat wall. We

compare three simulation methods: Brownian Dynamics (BD), the Lattice Boltzmann

Method (LBM), and a recent Stochastic Event-Driven Molecular Dynamics (SEDMD)

algorithm.1 We focus on the dynamics of the free end (last bead) of the tethered chain

and we examine the cross-correlation function (CCF) and power spectral density

(PSD) of the chain extensions in the flow and gradient directions as a function of

chain length N and dimensionless shear rate Wi. Extensive simulation results suggest

1This chapter is based on the publication: Y. Zhang, A. Donev, T. Weisgraber, B. J. Alder, M.D. Graham, and J. J. de Pablo, J. Chem. Phys. 130, 234902. LBM and SEDMD calculations wereperformed by our collaborators at the Lawrence Livermore National Lab.

Page 56: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

44

a classical fluctuation-dissipation stochastic process and question the existence of

periodicity of the cyclic dynamics, as previously claimed. We support our numerical

findings with a simple analytical calculation for a harmonic dimer in shear flow given

in Appendix C.

5.1 Introduction

The interaction of polymer molecules with fluid flow has been studied both theoret-

ically [130, 21] and experimentally [109, 108, 136, 135, 57] for several decades. The

behavior of polymer chains in flow is determined by an intricate interplay between

the flow gradients, chain elasticity, thermal fluctuations, and the physical confinement

[78, 75]. The dynamics of tethered polymer molecules (“polymer brushes”) in shear

flow has received considerable attention due to its relevance to diverse important

applications, such as colloidal stabilization, surface adhesion, and lubrication [96].

In contrast to previous work on the collective motion of polymer brushes [96, 55],

Doyle et al. studied the dynamics of a single tethered DNA in uniform shear flow

using fluorescence videomicroscopy [46]. Enhanced temporal fluctuations in the chain

extension were observed, and were attributed to the coupling of advection in the

flow direction and diffusion in the gradient direction. A cyclic dynamics mechanism

(Fig.5.1), closely related to the tumbling dynamics of a free polymer molecule in

shear flow [115, 57, 127], was proposed based on results from Brownian dynamics

simulations. No peaks were observed in the calculated power spectral density (PSD)

of the DNA extension in the flow direction, and the authors therefore suggested that

the cyclic dynamics of a tethered chain in shear flow is aperiodic. An important

Page 57: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

45

physical question is whether there is a characteristic timescale associated with the

cyclic motion that is distinct from the internal relaxation time of the chain.

Several computational studies revisited the problem of a tethered chain in shear

flow by looking at different variables relating to both the flow and gradient directions,

such as extensions along both flow and gradient directions [37], polymer orientation

angle defined through the gyration tensor [127], and angle between the wall and

the vector joining the tethering point to the center-of-mass of the chain [62]. The

cross-correlation functions for such variables exhibit signatures of the proposed cyclic

motion in the form of peaks at non-zero delay time. Because of the particular choice

of variables in Ref. [62], the lack of such peaks at small Weissenberg numbers was

attributed to the existence of a critical Weissenberg number; as demonstrated in Ref.

[42], choosing a different sets of variables shows that the signature peaks exist even at

small shear rates. In Ref. [62], the position of the peak in the studied cross-correlation

functions was interpreted as a characteristic cycling time, and it was found to be a

fraction of the relaxation time of the polymer chain. In Ref. [127] the tumbling motion

of a free polymer chain in shear flow was studied experimentally and computationally,

and wide peaks were found in the Fourier spectra of the time series of the angle of

the chain relative to the flow direction. These peaks were identified as evidence of

periodic motion of the tumbling molecule. The characteristic tumbling time (period)

was extracted from the position of the peak in the spectrum and was found to be

in good agreement with the experimentally-measured tumbling frequency. These

finding for a free chain in shear flow inspired similar studies of a tethered chain,

and similar observations of periodic motion with a characteristic period about an

order of magnitude larger than the relaxation time were reported [127, 37, 38]. In

Page 58: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

46

12

11

1)

2)

3)

4)

z

x

y

Figure 5.1: Snapshots taken from a simulation run withWi = 5 to show tethered DNAdynamics. The beads are labeled from 1 to 11 as shown. Cyclic motion mechanismproposed by Doyle et al. is composed of four stages: 1) (Re)coiling; 2) initiating; 3)stretching; 4) rotating [46]

several later studies of the tumbling emotion of a free polymer chain in shear flow,

experimental results [57], numerical simulation [115], and theory [34, 156] all suggest

that the intervals between successive tumbling events are exponentially-distributed

with a decay constant equal to the relaxation time of the chain. Such an exponential

tail implies that the tumbling events occur as a Poisson-like process, which is aperiodic

despite the existence of a characteristic timescale (frequency of repetition). If the

tumbling events of a free polymer in shear flow is aperiodic, intuitively, adding of a

wall to break the symmetry of the motion should leave the dynamics of a tethered

chain aperiodic.

Different authors use the terms “cyclic” (repetitive) and “periodic” with differ-

Page 59: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

47

ent meanings, and it is therefore important to give our definitions. Periodicity is

the quality of occurring in regular time intervals (periods). Periodic motion has

correlation functions that are (possibly damped) oscillatory functions, and spectra

that have sharp peaks. Noise (fluctuations) and the associated dissipation will al-

ways broaden any peaks that are related to underlying deterministic periodic motion

(and consequently, exponentially dampen the oscillations in the real-space correla-

tion functions). As an example, for a rigid spheroid in shear flow, there is indeed

periodic motion (Jeffery’s orbits) in pure shear flow. Adding fluctuations, when they

are small, is expected to preserve that but introduce some broadening of the spectral

peaks [88]. In contrast, a cycle usually means a process that eventually returns to its

beginning and then repeats itself in the same sequence. The end-to-end tumbling of

a single polymer molecule in shear flow provides a relevant example. In this paper

we analytically calculate the power spectrum for a tethered dimer in shear flow, and

find an exponentially-decaying cross-correlation function that has the relaxation time

as the only characteristic timescale. More importantly, this analytical example shows

that the power spectrum can exhibit a wide peak at small frequencies without any

underlying periodic motion, and that the location of a maximum in the PSD is not

necessarily an indication of a new timescale. The analytical results for a tethered

dimer are consistent with our numerical observations for longer tethered chains in

shear flow. Therefore, our investigations do not confirm the existence of periodic

motion with a period distinct from the relaxation time of the chain, as previously

suggested in the literature [127, 37, 38].

We apply three different solvent representations to the same problem of a teth-

ered chain in shear flow. The models are an implicit solvent (Brownian Dynamics

Page 60: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

48

[74], BD), a continuum solvent (Lattice-Boltzmann [5], LB), and a particle solvent

(Direct Simulation Monte Carlo [42], DSMC). Such comparison between widely dif-

fering methods on the same problem is important as a validation of their range of

applicability. It is also important to compare the computational performance of the

different methods. In this work, different chain representations and boundary condi-

tions make a direct quantitative comparison impossible. Specifically, the BD polymer

is a worm-like chain representative of semi-flexible DNA, in the LB simulations it is

a flexible chain of repulsive spheres, and in the DSMC solvent it is a flexible chain

of hard spheres. However, we can access the importance of the details of the chain

model, and in particular, of chain elasticity, and thus test the widely used assumption

that the dynamics scales with the Weissenberg number Wi = γτ , where γ is the shear

rate and τ the chain relaxation time, independent of the details of the model. For this

particular problem of a tethered chain in shear flow, we find good agreement between

the different methods.

A general discussion of the wide range of techniques for modeling the hydrody-

namics of polymer chains in solution is given in Section 5.2. Further details about the

three specific techniques we use in this paper to study the tethered polymer problem

are given in Section 5.3. In Section 5.4 we present our results, and finally, in Section

5.5 we give some concluding remarks.

Page 61: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

49

5.2 Discussion of Methods for Hydrodynamics of

Polymer Solutions

In this Section we give a brief overview of various methodologies for modeling hy-

drodynamics of soft matter systems, notably, polymer solutions (see also review by

Duenweg and Ladd [47]). The various methods for computational hydrodynamics

of polymer solutions can be divided in two major categories. The first are purely

continuum methods that use constitutive equations for the polymer solution. These

models only apply at macroscopic scales, when the number of polymer chains in an

elementary fluid flow volume is large, so that statistical averages of the chain con-

formations can be used as parameters in constitutive models of the time-dependent

stress as a function of the strain rate history. The construction of such constitu-

tive models is ad hoc and rather difficult in situations where conformations of the

chains couple to an unsteady flow, as, for example, in the problem of turbulent drag

reduction. Additionally, such continuum methods do not apply to situations where

the dynamics of individual chains are of interest, such as a DNA molecule flowing

through a micro-channel or DNA translocation through a pore. The second major

category of methods explicitly simulates the motion of each polymer chain using some

form of molecular dynamics. The simplest chain model is a dumbbell. Multi-bead

representations of the chains are capable of complex chain conformations but require

models for the bead-bead interactions. Such details of the polymer model are impor-

tant for both physical fidelity and computational efficiency. For example, preventing

chain-chain crossing can require stiff interactions for excluded volume terms, which

in turn can lead to small time steps. There are also two major types of algorithms

Page 62: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

50

for dealing with the solvent. One represents the solvent implicitly, and the others use

an explicit solvent. An implicit solvent is most efficient computationally, however, it

can only be used if the fluid flow in the absence of the polymer is known analytically

or can easily be pre-computed numerically (e.g., stationary flow), and if the polymer

chains themselves do not alter the background flow.

5.2.1 Implicit Solvent: Brownian Dynamics

The most widely used implicit-solvent algorithm is Brownian dynamics [74], described

in more detail in Section 5.3.1. The method involves solving first-order differential

equations of motion for the positions of the beads with additional forces due to the

presence of the solvent. These solvent forces can be separated into a deterministic

portion, for which a (linear) analytical approximation is used, and a stochastic por-

tion, which is assumed to be white noise. The fluctuation-dissipation theorem is used

to set the magnitude of the stochastic forcing. Brownian dynamics relies on several

assumptions usually valid in microfluidic applications. The first assumption is that

of small Reynolds number laminar (usually stationary) flow adequately described by

a linearized Navier-Stokes equation. The second assumption is that hydrodynamic

fields develop infinitely quickly relative to the rate at which the polymer conforma-

tion changes, so that a quasi-stationary approximation can be used to describe the

perturbation of the flow field induced by the motion of the beads. This approxi-

mation leads to Stokes friction on single beads, as well as hydrodynamic interaction

pairwise terms approximated with a long-range Oseen tensor as derived through an

asymptotic (t → ∞) analysis for point particles. The free-draining approximation

of Brownian dynamics neglects these pairwise hydrodynamic interactions. The in-

Page 63: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

51

clusion of pairwise hydrodynamic interactions leads to a matrix formulation of the

fluctuation-dissipation theorem and therefore factorization of a matrix of the size of

the number of beads is required at every time-step. Various numerical tools have

been devised to avoid dense factorization [76, 132, 74, 67], thereby reducing the cost

of a single time step in Brownian dynamics with hydrodynamic interactions.

Brownian Dynamics should be distinguished from Langevin dynamics, in which

second-order (Newton’s) equations of motion are used for the beads, that is, both

the bead velocities and positions are included as explicit degrees of freedom (but the

solvent is still implicit) [122]. This assumes that there is a large separation of time-

scales between the fluid degrees of freedom and the velocities of the beads, which

is in fact only true if the beads are much denser than the solvent. Furthermore, a

much smaller timestep necessary to resolve the faster dynamics (relaxation) of the

bead velocities. Therefore, Langevin dynamics finds its use only when the solvent is

represented explicitly, so that calculating the friction and stochastic forces no longer

requires factorization of the mobility tensor.

An important advantage of Brownian dynamics is that it simulates the limit of

zero Reynolds number exactly. It can also often exactly account for simple boundary

conditions (e.g., flow in an infinite half plane) without resorting to approximations

that truncate the flow field to a finite domain, such as the commonly-used periodic

boundary conditions. Brownian dynamics is relatively easy to implement, however,

complex boundary conditions, such as indentations or bumps on walls, requires care

so that analytical approximations to the Oseen tensor that preserve the positive-

definiteness of the diffusion tensor [69]. While the computational cost can rise rapidly

as the number of beads is increased when direct implementations are used, novel

Page 64: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

52

schemes can be used to truncate the long-range hydrodynamic interactions and yield a

linear dependence on system size, similarly to the handling of electrostatic interactions

in spectral [132] and multipole methods [67].

5.2.2 Explicit Solvent: Continuum Methods

In order to capture the bi-directional coupling between the motion of the polymer

and the flow around it, it is necessary to explicitly represent the solvent. The first

level of approximation is to use a continuum description of the solvent assuming the

applicability of the Navier-Stokes (NS) PDEs at small length scales. Typically an

incompressible assumption is made, which is appropriate at sufficiently low Mach

numbers if acoustic waves are not of interest. Additional approximations such as lin-

earization or an iso-thermal approximation may be appropriate. The time-dependent

(unsteady) NS equations can be solved by any of the numerous existing CFD al-

gorithms, including explicit, implicit, or semi-implicit algorithms of varying level of

complexity [131, 146, 12]. One of the advantages of the PDE formulation over par-

ticle methods is the ability to use powerful adaptive mesh resolution techniques that

allow coarsening of the mesh away from the region of interest, here polymer chains.

However, the case of complex boundary conditions such as needed, for example, in

the handling of moving beads or flow through porous media, presents difficulties. An

alternative to solving the Navier-Stokes PDEs is to use the Lattice-Boltzmann (LB)

method [150], as discussed in Section 5.3.2. It requires small time steps limited by

CFL-type conditions, however, each of the time steps is efficient. Recently, so-called

entropic LB schemes have been developed that posses a discrete H-function, resulting

in unconditional numerical stability even at high Reynolds numbers [23]. LB has been

Page 65: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

53

found competitive with NS solvers in many situations and has the further advantage

that it is based on kinetic theory and allows a more detailed level of description than

NS. An important advantage of LB solvers is also their ability to handle complex

boundary conditions. Recently, Chen et al. have provided a detailed comparison

between BD and LB simulations on a DNA model that shows that the LB method

provides a reasonable description of the results of more precise BD simulations at low

Reynolds numbers [32].

Thermal Fluctuations

Most continuum fluid dynamics methods are deterministic and thus do not include

internal fluctuations of the hydrodynamic fields. Fluctuations become more important

the smaller the length scale of interest, and are crucial for polymer flows. Including

thermal fluctuations in a continuum formulation has been carried out for both CFD

and LB algorithms. The Landau-Lifshitz Navier Stokes (LLNS) equations include

thermal fluctuations in the stress tensor but numerical schemes to solve them are

not nearly as advanced as are the standard CFD solvers [131, 51, 19]. Fluctuations

have been included in LB and do not pose any particular numerical problems [3].

Fluctuations have also been included in incompressible solvers in conjunction with

the Immersed Boundary Method [12, 83]. The ability to turn fluctuations on or off

is an important advantage of continuum-based methods over particle methods.

Coupling with the Polymer Chains

Regardless of what continuum method is employed, it is necessary to couple that

method to the MD description of the polymer chains. The simplest and most com-

Page 66: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

54

monly used coupling scheme is to approximate the beads as points and assume

for the solvent-induced force on the polymer beads the Stokes-Langevin form F =

−6πRHηvf+FS, where vf is an estimate of the local fluid velocity and FS is an uncor-

related stochastic force whose magnitude obeys the fluctuation-dissipation theorem

[146, 58]. This approximation is similar to that in Brownian dynamics, namely, Stokes

law is only valid in quasi-static continuum situations, relying on the separations of

time and length scales which are usually only marginally separated in realistic situa-

tions. Typically the strength of the coupling, RH , is empirically tuned to reproduce

experimental measurements. The coupling can also be dealt with when the beads

occupy an actual volume, free of fluid. Then stick or slip boundary condition at the

surface of the beads are employed, as in both NS [131] and LB [150] simulations of

colloidal dispersions. However, these methods are rarely used in polymer simulations

due to the complexity when many moving particles are involved, because, the grid

size needs to be smaller than the bead size and may need to be adaptively changed

when the bead moves.

A different alternative is provided by the Immersed Boundary method [12], where

the fluid occupies the whole space and the particles, represented as immersed struc-

tures, move together with the fluid with a velocity that is a localized average of the

fluid velocity. This eliminates the bead inertia from the problem and the need to ex-

plicitly enforce boundary conditions on the surface of the beads. The method can be

seen as an alternative to Brownian dynamics that correctly captures time-dependent

momentum transport in the fluid by explicitly representing the fluid flow, and also

includes thermodynamically-consistent thermal fluctuations [83].

Page 67: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

55

5.2.3 Explicit Solvent: Particle Methods

An alternative to continuum methods is to use a particle representation of the fluid.

The most detailed description is a MD simulation of both the fluid and the solvent.

Unlike the classical NS equations, MD automatically and correctly includes fluctua-

tions, internal fluid structure, diffusion, and non-linear transport. Particle methods

are also typically simple to implement and can easily accommodate complex bound-

ary conditions.Typically a truncated repulsive Lenard-Jones potential is used for the

solvent-solvent interactions. However, even with massive parallelization such MD

simulations are limited to short total times and therefore efforts have been made to

coarse-grain the solvent to a mesoscopic representation. There, the fluid particles are

no longer representative of solvent molecules, but are larger having different dynamics

and interactions with each other. However, the viscosity and the stress fluctuations

in the solvent must be reproduced correctly. There are mesoscopic particle solvents

of progressively decreasing level of microscopic fidelity, and thus increasing efficiency.

The handling of the coupling between the solvent and the beads is a separate issue,

like for continuum solvents. A particle solvent may be coupled to a polymer chain

by including explicit short-ranged solvent-bead continuous [89] or hard-spheres [42]

interaction potentials. Efficiency can further be gained by coarse graining the bead-

solvent interactions as well, typically using the same ideas as used to coarse grain

the solvent-solvent interactions [82, 98]. Dissipative Particle Dynamics (DPD) [111]

further coarsens the solvent molecules to obtain a system of weakly-repulsive spheres

interacting with a mixture of conservative, stochastic, and dissipative forces. The

conservative forces can be used to reproduce the solvent equation of state, while the

dissipative forces model viscous friction. The stochastic forces act as a thermostat

Page 68: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

56

that ensures detailed balance and correct thermal fluctuations in the DPD fluid. The

method has great flexibility and requires significantly less solvent particles and larger

time-steps than classical MD, however, it still requires costly integration of differen-

tial equations of motion for each of the solvent particles. Such integration of ODEs

can be avoided by using a kinetic Monte Carlo method, such as Direct Simulation

Monte Carlo (DSMC), to represent the solvent-solvent interactions. The idea is to

use stochastic conservative collisions between nearby solvent particles to represent

the exchange of momentum and energy. Both multi-particle collisions [82, 98] and bi-

nary collisions [42] have been used, as described in Section 5.3.3. The computational

efficiency comes at the cost of neglecting the structure of the solvent, as in contin-

uum methods. Recently a new Stochastic Hard-Sphere Dynamics method has been

proposed that also uses uncorrelated stochastic binary collisions but still produces a

non-trivial fluid structure and a thermodynamically-consistent non-ideal equation of

state, similar to those of a DPD fluid [43].

5.2.4 Coupled Methods

Methods that combine several of the techniques described above into a single con-

currently coupled simulation can take advantage of their region of validity. Such a

simulation may involve several levels each with a different level of microscopic de-

tail. For example, molecular dynamics with complete atomistic detail and realistic

potentials may be used for the polymer chain(s) and nearby solvent. The solvent can

then be coarse grained to a mesoscopic particle fluid sufficiently far from any chains.

The particle method can then be coupled to an explicit fluctuating hydrodynamic

description with a fine grid, for example, LB or a fluctuating NS solver. Finally, the

Page 69: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

57

hydro grid can be adaptively coarsened in regions even farther from the chain, and a

non-fluctuating continuum solver used. This last macroscopic level can use a different

method from the fluctuating hydrodynamics level, for example, it could be an incom-

pressible NS solver. Much remains to be done to enable a truly multiscale simulation

capable of bridging from microscopic to macroscopic length and time-scales [155, 39].

5.3 Simulation Methods

In this Section we describe in further technical detail the three different techniques

we apply to the tethered polymer problem. The majority of the methodology has

been previously published so here we only summarize the essential points and cite

the relevant works.

5.3.1 Brownian Dynamics

Details of the DNA model and Brownian dynamics simulation method that we use

can be found in Refs. [76, 74]. We discretize a double-stranded DNA molecule into

a bead-spring chain composed of Nb beads of radius Rb = 77nm (the unit of length,

1 l.u. = 77nm) connected by Ns = Nb − 1 entropic springs. Each spring represents a

DNA segment of 4850 base pairs, so that Nb = 11 corresponds to a stained λ-DNA,

which has a contour length of 21 µm. In Brownian dynamics, a force balance on this

chain leads to a stochastic differential equation for the dynamics of the chain [105],

∆R = [U+D · FkBT

+∂

∂R·D]∆t +

√2B ·∆W (5.1)

Page 70: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

58

where R is the vector containing bead positions, R = r1, ..., rN, U is the unper-

turbed velocity field at the bead centers, kB is Boltzmann constant, T is absolute

temperature, F is the non-hydrodynamic and non-Brownian forces, and D = B ·BT

is the diffusion tensor. The components of ∆W are obtained from a real-valued

Gaussian distribution with mean zero and variance dt. In a unbounded space, the

hydrodynamic interactions (HI) enter the chain dynamics through the diffusion ten-

sor,

Dij = kBT [(6πηa)−1Iδij +Ωij ] (5.2)

where η is the viscosity of the solvent, a is the bead hydrodynamic radius, I is the

unit tensor, δij is the Kronecker delta, and Ω is the HI (Stokeslet or Oseen) tensor.

Recent work has provided evidence of hydrodynamic coupling to the wall and ex-

perimental validation of the use of point-particle (Stokeslet) hydrodynamic interac-

tions (HI) to describe the motion of Brownian particles near a surface [48]. Therefore,

it is essential to have wall corrected HI in the simulation to capture the dynamics of

a tethered chain correctly. In a bounded space, like near a solid wall, the HI tensor

is modified to,

Ωij = (1− δij)ΩOB(ri − rj) +ΩW (ri − rj) (5.3)

where ΩOB is the free-space diffusion tensor, and ΩW is the correction which accounts

for the no-slip constraint on the wall. The solution for a Stokeslet above a flat plate

given by Blake allows us to calculate ΩW exactly [22]. In a square channel or complex

geometries, we need to solve this problem numerically with a finite element method

to determine ΩW at a grid of points [78]. Based on this description of near-wall HI,

Jendrejack et al. [77] predicted that the DNA molecules migrate away from the wall

Page 71: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

59

in shear flow, leading to the formation of depletion layers in the near wall region. This

prediction has been verified in recent experiments of dilute DNA solutions undergoing

pressure-driven flow in microchannels [4, 30]. In different works, Delgado-Buscalioni

used a hybrid particle-continuum model method to describe HI [37] and Schroeder et

al. used unbounded space HI [127] to study the motion of a tethered chain.

We further assume that the chain is ideal (no self-excluded volume interactions

between different beads). The entropic springs connecting the beads obey a worm-like

chain law

Fsij =

kBT

2bk[(1− |rj − ri|

Nk,sbk)−2 − 1 +

4|rj − ri|Nk,sbk

]rj − ri|rj − ri|

, (5.4)

where bk is the Kuhn length for DNA and Nk,s is the number of Kuhn lengths per

spring. The physical confinement is taken into account through an empirical bead-

wall repulsive potential of the form

Uwalli = Awallb

−1k δ−2

wall(hi − δwall)3, (5.5)

when hi < δwall, where hi represents the perpendicular distance of bead i from the

wall, δwall is the cut-off distance. In this work, we choose Awall = 25kBT and δwall =

bkN1/2k,s /2 = 0.24 µm. All of the parameters a, bk, ν are the same as used in previous

work, where it has been shown to successfully reproduce the static and dynamic

properties of DNA with contour length 10µm− 126µm [76, 78]. For each parameter

set, the sample size is 30 chains unless otherwise specified. All results are presented

for DNA at room temperature in a solvent with a viscosity of 1 cP .

To study the dynamics of a tethered chain, beads are labeled from 1 to Nb + 1,

starting from the tethered point, as illustrated in Fig. 1. The fluid velocity in the

Page 72: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

60

flow direction z is a linear function of distance from the wall in the gradient direction

x, vz = γx, where γ is the shear rate, and vx = 0 and vy = 0. Following common

experimental practice, the longest relaxation time is calculated by allowing a chain

that is initially stretched using a large shear rate to relax to equilibrium. Near

equilibrium, the relaxation time is determined by an exponential decay fit the chain

extension along the stretch direction,

〈X2〉 = (X2(0)− 〈X2〉eq) exp(−t

τ) + 〈X2〉eq. (5.6)

An exponential fit to the autocorrelation of the chain extension (relative to equilib-

rium) parallel to the wall gives similar results. The relaxation time for our λ-DNA

is estimated to be 0.59s at room temperature, which is in good agreement with the

experimental result of 0.51s [46] after extrapolating the viscosity to 1 cP .

5.3.2 Lattice-Boltzmann

In addition to Brownian Dynamics, we examine the short time correlations of a teth-

ered polymer in a uniform shear flow using a hybrid Lattice Boltzmann (LB) and

Molecular Dynamics (MD) code based on the method by Ahlrichs and Dunweg [5].

The Lattice Boltzmann method is a mesoscopic approach to fluid flow calculation

and is based on a discrete version of the Boltzmann equation with enough detail to

recover hydrodynamic behavior. The LB equation describes the evolution of a single-

particle distribution function, fi (x, t), which is the mass density of particles moving

Page 73: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

61

with velocity ei at a time t and position x on a cubic lattice,

fi (x+ ei∆t, t +∆t) = fi (x, t) +∑

j

Aij

[

fj (x, t)− f eqj (x, t)

]

. (5.7)

The set of velocities ei is discrete and chosen such that x + ei∆t always remains a

lattice site. The last term describes the collision process in which the distribution

function relaxes to a local equilibrium, for which we utilize the BGK (Bhatnagar-

Gross-Krook) approximation to the collision operator, Aij = −τ−1δij , where τ is a

relaxation time. The macroscopic hydrodynamic quantities, density ρ, momentum j =

ρu, and momentum flux Π, are computed from moments of the particle distribution

function,

ρ =∑

i

fi, j =∑

i

fiei, and Π =∑

i

fiei ⊗ ei. (5.8)

The equilibrium distribution depends on the macroscopic variables and its form is

given by

f eqi (x, t) = wiρ

[

1 +ei · uc2s

+(ei · u)22c4s

− u2

2c2s

]

, (5.9)

where the weights wi depend on the particle velocity discretization and are determined

by mass and momentum conservation. The lattice sound speed is cs = ∆x/√3∆t,

where ∆x is the lattice spacing. In this work we solved the distribution function on

the standard D3Q19 lattice [150] where the 19 particle velocity components consist

of one rest particle, the 6 nearest neighbors in a simple cubic lattice, and the 12

next nearest neighbors in the [110] directions. The corresponding weights are 1/3,

1/18, and 1/36. The LB method avoids the additional mathematical complexities of

Navier-Stokes PDE solvers and is straightforward to parallelize efficiently. Using a

Page 74: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

62

Chapman-Enskog expansion, the lattice-Boltzmann equation can recover the Navier-

Stokes equations for small Mach and Knudsen numbers, and, within these limits it

is second-order accurate in space and time. Compared to the other two methods we

apply to the tethered polymer problem, BD and SEDMD, LB is less efficient in this

case since it solves for the solvent in the entire domain, even relatively far from the

polymer chain.

In the LB calculations, the polymer is represented by 25 point particles joined by

finitely extendable nonlinear elastic (FENE) springs and interact through a repulsive

Lennard-Jones potential among each other and with the walls. Solvent fluctuations

are incorporated by adding a stochastic term to the right hand side of the LB equation.

This term introduces fluctuations into the momentum flux in a manner that satisfies

the fluctuation-dissipation theorem [150]. Coupling between the LB for the solvent

and the MD for the solute is achieved through Stokes drag forces and white-noise

stochastic forces acting on the monomers. The first monomer in the chain is tethered

to the stationary lower wall in a domain having 36, 22, and 24 lattice sites in the

streamwise, spanwise, and wall normal directions. The streamwise and spanwise

directions are periodic and the bounding upper wall moves with constant velocity,

providing the uniform shear.

5.3.3 Stochastic Event-Driven Molecular Dynamics

In addition to Brownian dynamics and Lattice-Boltzmann, we have also applied a

purely particle-based method to the tethered polymer problem. The Stochastic Event-

Driven Molecular Dynamics (SEDMD) algorithm introduced in Ref. [42] combines

Event-Driven Molecular Dynamics (EDMD) for the polymer particles with Direct

Page 75: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

63

Simulation Monte Carlo (DSMC) [7] for the solvent particles. In SEDMD, the poly-

mers are represented as chains of hard spheres tethered by square wells. The solvent

particles are realistically smaller than the beads and are considered as hard spheres

that interact with the polymer beads with the usual hard-core repulsion. The al-

gorithm processes true (deterministic, exact) binary collisions between the solvent

particles and the beads, without any approximate coupling or stochastic forcing.

However, the solvent particles themselves do not directly interact with each other,

that is, they can freely pass through each other as for an ideal gas. Deterministic

collisions between the solvent particles are replaced with momentum- and energy-

conserving stochastic collisions between nearby solvent particles. This gives realistic

hydrodynamic behavior and fluctuations in the solvent, with tunable viscosity and

thermal conductivity, but without internal fluid structure. A recent modification of

the DSMC algorithm can be used to achieve a non-ideal equation of state for the

stochastic solvent that is thermodynamically-consistent with the density fluctuations

[43].

Hard-sphere models of polymer chains have been used in EDMD simulations for

some time [139, 107, 101]. These models typically involve, in addition to the usual

hard-core exclusion, additional square well interactions to model chain connectivity.

Recent studies have used square well attraction to model the effect of solvent quality

[104]. Even more complex square well models have been developed for polymers

with chemical structure and it has been demonstrated that such models, despite their

apparent simplicity, can successfully reproduce the complex packing structures found

in polymer aggregation [101, 107]. Here we use the simplest model of a polymer

chain, namely, a linear chain of Nb particles tethered by unbreakable bonds. This is

Page 76: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

64

similar to the commonly-used freely jointed bead-spring FENE model model used in

time-driven MD. The length of the tethers has been chosen to be 1.1Db, where Db is

the diameter of the beads.

Several particle methods for hydrodynamics have been described in the literature,

such as MD [13], dissipative particle dynamics (DPD) [111], and multi-particle col-

lision dynamics (MPCD) [119, 89]. Molecular dynamics is the most accurate model

of the fluid structure and dynamics, however, it is very computationally demanding

due to the need to integrate equations of motion with small time steps ∆t and calcu-

late interparticle forces at every time step. The key idea behind DSMC is to replace

deterministic interactions between the particles with stochastic momentum exchange

(collisions) between nearby particles. The standard DSMC [7] algorithm starts with

a time step where particles are propagated advectively, r′

i = ri + vi∆t, and sorted

into a grid of cells. Then, a certain number Ncoll ∼ ΓcNc(Nc − 1)∆t of stochastic

conservative collisions are executed between pairs of particles randomly chosen from

the Nc particles inside the cell. For mean free paths comparable to the cell size, the

grid of cells should be shifted randomly before each collision step to ensure Galilean

invariance. The collision rate Γc and the pairwise probability distributions are chosen

based on kinetic theory.

In SEDMD the polymer chains and the bead-solvent interactions are handled us-

ing hard-sphere event-driven molecular dynamics (EDMD) [6, 44, 139, 104] instead

of the time-driven MD (TDMD) widely used for continuous potentials. The essential

difference between EDMD and TDMD is that EDMD is asynchronous and there is no

time step, instead, collisions between hard particles are explicitly predicted and pro-

cessed at their exact (to numerical precision) time of occurrence. Since particles move

Page 77: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

65

along simple trajectories (straight lines) between collisions, the algorithm does not

waste any time simulating motion in between events (collisions). SEDMD combines

time-driven DSMC with EDMD by splitting the particles between ED particles and

TD particles. Roughly speaking, only the polymer beads and the DSMC particles

surrounding them are treated asynchronously as in EDMD. The rest of the DSMC

particles that are not even inserted into the event queue. Instead, they are handled

using a time-driven (TD) algorithm very similar to that used in traditional DSMC.

In three dimensions, a very large number of solvent particles is required to fill the

simulation domain. The majority of these particles are far from the polymer chain

and they are unlikely to significantly impact or be impacted by the motion of the

polymer chain. We therefore approximate the behavior of the solvent particles suffi-

ciently far away from any polymer beads with that of a quasi-equilibrium ensemble.

In this ensemble the positions of the particles are as in equilibrium and the velocities

follow a local Maxwellian distribution whose mean is the macroscopic local veloc-

ity. These particles are not simulated explicitly, rather, we can think of the polymer

chain and the surrounding DSMC fluid as being embedded into an infinite reservoir

of DSMC particles which enter and leave the simulation domain following the appro-

priate distributions. Using such open or stochastic boundary conditions dramatically

improves the speed, at the cost of small errors due to truncation of hydrodynamic

fields. This truncation can be avoided by coupling DSMC to a continuum fluctuating

hydrodynamic solver [155].

We have made several runs for different polymer lengths and also bead sizes. One

set of runs used either Nb = 25 or 50 large beads each about 10 times larger than

a solvent particle. Another set of runs used either Nb = 30 or 60 small beads each

Page 78: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

66

identical to a solvent particle, with faster execution but nearly identical results. In the

simulations reported here we have used rough wall BCs for collisions between DSMC

and non-DSMC particles [42]. This emulates a non-stick boundary condition at the

surface of the polymer beads. Using specular (slip) conditions lowers the friction

coefficient, but does not qualitatively affect the behavior of tethered polymers. All of

the runs used open boundary conditions, where about 153 DSMC cells around each

bead were explicitly simulated. Note that for (partially) collapsed polymer chains the

total number of explicitly simulated cells is much smaller than 153Nb. The Nb = 30

runs were run for about 6000τ0 relaxation times, and such a run takes about 6 days

on a single 2.4GHz Dual-Core AMD Opteron processor. Even for such long runs

the statistical errors due to the strong fluctuations in the polymer conformations are

large, especially for correlation functions at long time lags t > τ .

5.4 Results

The main goal of our paper is to reinvestigate the tethered chain problem through

extensive long time simulations (thousands of longest relaxation time of the teth-

ered polymer, τ) involving different representations of polymer and solvent, including

Brownian dynamics [74] (BD), the Lattice Boltzmann method [5] (LBM), and a re-

cent Stochastic Event-Driven Molecular Dynamics [42] (SEDMD) algorithm. In this

section we present comparison results from our simulations. More extensive results

for the tethered polymer problem obtained using the SEDMD algorithm are presented

in Ref. [42]. Since the three different methods that we use give similar results and

Brownian Dynamics is the fastest methodology, the majority of the results we present

Page 79: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

67

will be from BD simulations with Nb = 11 (Ns = 10), unless otherwise indicated. Of

the three methods used here, LB is the slowest and thus the LB results are of more

limited duration. We emphasize that direct computational comparison between the

methods is unfair. Most significantly, the LB runs use periodic boundary conditions

and have to fill the whole simulation domain with explicit solvent (lattice points).

By contrast, the SEDMD runs use open boundaries and thus use much less explicit

solvent, whereas the Brownian dynamics does not use an explicit solvent at all.

Doyle et al. proposed a cyclic dynamics mechanism for a tethered polymer chain in

shear flow (Fig. 5.1) based on Brownian dynamics simulation results [46]. According

to this scenario, when thermal fluctuations cause motion in the gradient direction

x (from state 1 to state 2), the chain is driven away from the wall and experiences

higher hydrodynamic drag. This leads to further stretching and an increase of the

extension in the flow direction z (state 3). Due to the finite extensibility of the chain,

the extension in the z direction is finite and depends on the shear rate and chain

properties. After stretching, the coupled torque of the hydrodynamic drag and spring

forces will rotate the chain towards the wall (state 4). As the chain get closer to the

wall, the flow velocity decreases and entropic recoiling becomes dominant, resulting in

a decrease of the z extension (state 1). The tethered chain could take other dynamical

paths than following the one described above, such as restretching or recoiling after

state 2 by random motion in −x and −z direction, respectively.

In Fig. 5.2 we show the probability distribution function (pdf) ρ(z, x) of the end

bead in the z − x plane at Wi = 0 and Wi = 2 for the three different methods. The

results are presented in dimensionless units by normalizing the unit of length by the

average radius of gyration in the x direction at Wi = 0. Here, again, we want to

Page 80: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

68

-1 0 1 2 3 40.0

0.5

1.0

1.5

2.0

2.5

3.0

(a) Wi = 0

( )b Wi = 2

-3 -2 -1 0 1 2 30.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

-3 -2 -1 0 1 2 30.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

-3 -2 -1 0 1 2 30.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.000.0330.0650.0980.130.160.200.230.26

-1 0 1 2 3 40.0

0.5

1.0

1.5

2.0

2.5

3.0

-1 0 1 2 3 40.0

0.5

1.0

1.5

2.0

2.5

3.000.0810.160.240.330.410.490.570.65

Figure 5.2: Probability distribution of the end bead of the tethered DNA moleculein a dimensionless x − z plane at Wi = 0 and Wi = 2. The visible differences canlikely be attributed to the differences in the boundary conditions between the differentmethods, as well as the different elasticity of the chains. (Left) Brownian dynamics.(Middle) Stochastic Event-Driven Molecular Dynamics. (Right) Lattice BoltzmannMethod.

Page 81: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

69

emphasize that we are not expecting perfect match between methods. In particular,

the different methods implement different effective boundary conditions at the wall

surface. In Brownian dynamics, an essentially reflective boundary condition appears,

while in the case of a hard-sphere chain a perfectly reflective boundary condition is

appropriate. For the LB runs an intermediate case appears, where the repulsion from

the wall is stronger than hard spheres but still finite-ranged. These boundary effects

are clearly visible in the results in Fig. 5.2, where the BD results show a depletion

layer near the wall where as the SEDMD and LB results show the bead spending

more time near the wall.

In Fig. 5.3 we compare the dependence of the relaxation times τx/y/z along the

three different axes on the flow rate among the three different methods. The figure

shows reasonable agreement between the different techniques, especially considering

the large errors inherent in determining relaxation times. We calculate the relaxation

times by fitting an exponential decay to the intermediate portion of the autocorre-

lation function 0.2 < C(t) < 0.8 of the position of the end bead along the three

coordinate axes. The LB calculations use periodic boundaries with a narrower box

in the spanwise (y) direction than in the streamwise (x) direction, which makes the

relaxation times τx(Wi = 0) and τy(Wi = 0) unequal, as they must be by symmetry.

We have scaled τy(Wi) (the shorter axes) in the LB results by a constant factor so as

to correct this strong boundary effect at Wi = 0. Among the three relaxation times,

the relaxation in the direction perpendicular to the wall τz is the shortest, even for

no flow. Note than in this work, following Ref. [42], the relaxation time along the

flow direction τx is used to define the internal relaxation time and thus Wi when

comparing among the different methods. Note also that it is τx that seems to get

Page 82: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

70

0 1 2 3 4 5 6

Wi

0

0.5

1

1.5

2

τ /

τ 0

xzy

circles=BD, squares=SEDMD, diamonds=LB

0 1 2 3 4 5 60

0.5

1

1.5

2

τz / τ

x

τy / τ

x

Figure 5.3: Dependence of the dimensionless relaxation time τ(Wi)/τ(Wi = 0) of thetethered chain along the three coordinate axes as a function of dimensionless flowrate Wi. The inset shows the ratios of the different relaxation times as a function ofWi.

Page 83: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

71

most strongly reduced as Wi increases.

To study the time scale associated with the fluctuating process (cycle) quantita-

tively and to find the correlation between different chain segments, we calculated the

cross-correlation functions (CCF) of beads’ positions. We also calculated the power

spectral density (PSD) in search of periodicity. The CCF and PSD are the natural

tools for examining the relationship between two time dependent random variables

in the time and frequency domain respectively. The mean-removed CCF of two time

series α(t) and β(t) is defined as

Cαβ(T ) =E[(α(t + T )− α)(β(t)− β)]

σασβ

(5.10)

where α = E(α) is the mean, σ2α = E(α2)−[E(α)]2 is the standard deviation, and T is

the time lag. A significant peak in the CCF at lag T indicates that α(t) is correlated

to β(t) when delayed by time T . In the frequency domain, the PSD is the norm of

the Fourier transform of the CCF,

Sαβ(ν) =

∫ ∞

−∞Cαβ(T )exp(−2iπνT )dT

(5.11)

Note that this is the standard definition used in the engineering literature, and here

the frequency ν = 1/T is actual frequency (inverse period) rather than angular fre-

quency ω = 2πν. To produce a PSD with accurate sampling around interesting

frequencies, long simulation times and a high sampling frequency are essential. We

examined various choices of variables to represent the motion of the chain and have

found little qualitative difference between them. We have chosen the position of the

end bead rNb= (x, y, z) as be the best option [42]. Extensive computational efforts

Page 84: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

72

have been undertaken to determine the CCF and PSD of the end bead coordinates

as function of chain length N and shear flow parameter Wi.

The CCF Czx(t) of the end bead at various Wi is shown in Fig. 5.4(a). The shape

of the CCF is consistent with the cyclic dynamics mechanism proposed by Doyle.

Clearly, in the absence of flow, Wi = 0, the movements in the x and z directions are

uncorrelated on all time scales. When shear flow is introduced, the movements in flow

direction and gradient direction are coupled together due to the nature of the flow

and the finite extensibility of the chain, as reflected in the rise of a prominent peak in

the CCF. When thermal motion in +x direction occurs, the chain will be stretched

with an increase in +z, which leads to a positive correlation. Similarly, when motion

in −x direction is introduced, the chain will recoil in the −z direction as the drag

decreases, which also leads to a positive correlation. As expected, the larger the

shear rate, the greater the correlation. There’s only one significant peak in the long

time correlation function, shown in the inlet of Fig. 5.4(a), which suggests that all

correlations are short-lived and not periodic. Turning attention to the correlation

between different chain segments, Fig. 5.5 shows the CCFs of several beads along

the chain at Wi = 2. One striking feature is that for beads sufficiently far from the

tether all curves pass the time axis at the same time lag. The fact that all CCFs

have the same shape indicates that a common movement pattern exists for the whole

chain. The inset in Fig. 5.5(a) shows the CCFs of the x coordinates of different

beads. Although the correlation decays as the distance along the chain increases, it

confirms that all beads move in a cooperative manner, indicated by the fact that the

peak positions are all at zero time lag. The CCF for the end bead for various chain

lengths are compared in Fig. 5.5(b), to show that there is no fundamental difference

Page 85: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

73

Figure 5.4: (a) Normalized cross-correlation functions (CCF) Czx(t) of the endbead’s coordinates in flow direction z and gradient direction x as a function of non-dimensional time, at various Wi for Ns = 10. The inset shows longer time lags.(b) Power spectral density (PSD) Szx(ν) of the end bead’s coordinates as a func-tion of non-dimensional frequency. The results are averaged over 30 runs for a totalsimulation time is 103τ , and 104τ for Wi = 5.

Page 86: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

74

Figure 5.5: (a). CCFs of end beads’ coordinates at Wi = 2 for a chain with Ns = 10and simulation time is 1000 τ . The number in the legend is the bead label as shownin Fig. 5.1. The inset shows the CCFs for x coordinates of different beads to studythe correlation of the dynamics between different beads (similar results are obtainedfor the y axes). (b) CCFs of end bead as function of chain length at Wi = 5. Thenumber in the legend is the number of springs Ns in the chain. The inset shows longertime lags.

Page 87: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

75

between different chain lengths, ranging from 20 µm (Ns = 10) to 80 µm (Ns = 40).

We have also established that these results are insensitive to the cut-off distance and

the magnitude of the repulsive potential between the wall and chain segments.

In Fig. 5.6 we compare the cross-correlation function Czx(t) at Wi = 2 among

the three different methods: Brownian Dynamics, Lattice-Boltzmann, and Stochastic

Event-Driven Molecular Dynamics. In particular, our goal is to verify the pervasive

assumption that the dynamics of polymer chains in shear flow is essentially universally

quantitatively determined by Wi for a wide range of flexible chains. Furthermore, it is

important to cross-validate the different methods against each other, given that each

of them makes certain assumptions and has somewhat different range of applicabil-

ity. The results in Fig. 5.6 indeed show reasonable agreement between the different

methods. Perfect agreement is not expected because the polymer models are different

among the different methods. The cross-correlations we measure are not consistent

with periodic motion. The PSD calculation does not show discernible peaks either,

as shown in Fig. 5.4(b). All that we can reliably extract from the results is that

the response of the chain to a large thermal fluctuation (the “cycle”) is reproducible

for short times, and we find no evidence of sustained correlations (oscillations) at

times longer than the internal relaxation time of the chain. For a free chain in shear

flow, where rotations of the chain are possible, one can count the number of tumbling

events per unit time and define that as a cycling time. The distribution of the delays

between successive tumbling events is itself important. If this distribution is sharply

peaked, that would be consistent with a periodic motion with a well-defined period.

If the distribution is exponential, this would indicate a Poisson-like tumbling process.

Several recent works have proposed an exponential distribution for the delay between

Page 88: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

76

-2.5 -2 -1.5 -1 -0.5 0 0.5 1 1.5 2

t / τ

-0.1

0

0.1

Czx

BD Wi=2.0 (N=20, error~0.005)

SEDMD Wi=1.8 (N=30, error~0.01)

LB Wi=2.0 (N=25, error~0.05)

Figure 5.6: Comparison of the cross-correlation function Czx(t) at Wi = 2 among thethree different methods: Brownian Dynamics (30 runs about ∼ 1000τ long), Lattice-Boltzmann (run is ∼ 600τ long), and Stochastic Event-Driven Molecular Dynamics(10 runs about ∼ 1000τ long).

Page 89: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

77

successive tumblings [57, 34, 156]. Furthermore, the tumbling time was found to be

related to the internal relaxation time of the chain [34, 156]. For tethered chains, we

cannot even identify and count a unique event such as tumbling and thus we cannot

extract a repetition frequency for the “cycle”.

In Appendix C, we analytically calculate the CCF for a Brownian particle tethered

to the origin with a harmonic spring and subjected to shear flow. This simple dimer

model qualitatively reproduces the features we see in the CCF for the tethered chains,

namely, a single peak at t ∼ τ of width ∼ τ and height ∼ Wi. Better quantitative

agreement is obtained when a nonlinear spring and a hard wall surface are also in-

cluded (without hydrodynamics). The PSD for the dimer model shows no peaks and

there is only a single time-scale in the dynamics, namely, the intrinsic relaxation time

τ . Furthermore, the analytical form of the CCF shows that by a slight modification of

a tunable parameter one can obtain a CCF fully-consistent with our numerical results

for longer chains. This analytical CCF has an analytical PSD that does show a broad

peak at small frequencies ντ ∼ 0.1, very similar to the previously reported peaks

used to justify the claims to periodicity in the chain motion [127, 37, 38]. This peak

is weak and broad even when plotted on a logarithmic axes and its exact shape and

maximum will vary depending on the particular model, variables used in calculating

the PSD, Wi, the definition used for calculating τ and Wi, etc. We therefore believe

that its interpretation as evidence of periodic motion is not justified.

The calculations in Appendix C for a dimer in shear flow also demonstrate that

a qualitatively similar behavior is observed even without hydrodynamic interactions.

Our results from Brownian Dynamics simulations in the free-draining limit confirm

this and show that the HI do not affect the results significantly, so long as the relax-

Page 90: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

78

ation time is recalculated when computing Wi. In the tethered case, we believe that

the competition between frictional and elastic restoring forcing dominates and the

hydrodynamic interactions are a weak perturbation. Therefore, it is not surprising

that the proper inclusion of hydrodynamic interactions is not essential for the teth-

ered polymer problem, as reasoned theoretically for a free chain in shear flow in Ref.

[156].

5.5 Conclusions

We studied the dynamics of a polymer molecule tethered to a hard wall and subjected

to a shear flow. We found consistent results among three methods utilizing different

representations of the solvent, Brownian Dynamics (BD), Lattice-Boltzmann (LB),

and Stochastic Event-Driven Molecular Dynamics (SEDMD). Specifically, BD im-

plicitly represents the solvent, LB explicitly represents the solvent flow on a discrete

lattice, and SEDMD utilizes a particle-based solvent. The three methods also utilized

different polymer chains, namely, the BD simulations used a worm-like chain, the LB

simulations used a FENE-LJ chain, and for SEDMD we used a tethered chain of hard

spheres.

The correlation functions of the position of the end bead question the existence of

periodic motion, as previously suggested. The cross-correlation function between the

bead positions along the flow and gradient directions shows a single peak indicative

of a fluctuation-dissipation cycle of duration comparable to the relaxation time of

the polymer. The corresponding Fourier representation, the power-spectral density,

shows no peaks. We find that neither the chain length of the polymer N , nor the

Page 91: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

79

dimensionless shear rate Wi, qualitatively alter the results, and in the Appendix C we

give some calculations for a very simple model of a dimer in shear flow that reproduces

the essential features of the observed peak in the cross-correlation function.

While our conclusions are rather different from other authors, our results are sta-

tistically consistent with those presented in the literature. Specifically, the shape and

position of the peaks in the cross-correlation functions are very similar to reported re-

sults, however, we did not observe large oscillations in the CCFs previously identified

as signatures of periodic motion [37]. We believe that this is due to the requirement of

very long simulation times to obtain good statistics for the time-correlation functions

at long time lags, as necessary to establish periodicity. Not all previous studies have

been able to reach sufficiently long simulation times. Another important point we

clarified is that maxima in the power-spectral density does not necessarily indicates a

periodic motion, which we demonstrate in Section C using an analytic dimer model.

Namely, an analytical shape is suggested by the dimer calculations that can exhibit

peaks very similar to those reported in the literature through small adjustments of

a tunable parameter, whose appropriate value likely depends on details of the model

used and the exact variables used in the calculations of the power spectrum. Fur-

thermore, different if not conflicting ways have been used to define and calculate the

“cycling time”, without properly distinguishing between the duration of a cycle and

the interval between cycles. Even more importantly, the very concept of a cycle in the

chain motion as a well-defined countable event, analogous to the case of a free chain

in shear flow, should be questioned. Our results are consistent with a simple tradi-

tional picture of continuous thermal fluctuations dissipated by deterministic friction,

leading to exponentially-decaying correlation functions.

Page 92: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

80

Chapter 6

Flow-driven DNA dynamics in

complex geometries

In this chapter, we consider the dynamics of a flow-driven DNAmolecule in micro/nano-

fluidic devices. In this case, the configuration-dependence of the mobility tensor can-

not be ignored and the solvent velocity field is in general non-linear on the length scale

of the molecule. We present an immersed boundary method that allows fast Brown-

ian dynamics simulation of polymer chains and other particles in complex geometries

with fluctuating hydrodynamics.1 This approach is applied to study the dynamics

of a flow-driven DNA molecule through a nanofluidic slit with an embedded array

of nanopits. We investigate the dynamics of the DNA molecule as a function of the

Peclet number and chain length, as well as the influence of hydrodynamic interactions

by comparing with free draining simulation results.

1This chapter is based on the manuscript: Y. Zhang, J. J. de Pablo, and M. D. Graham, submittedto J. Chem. Phys. (April 12, 2011)

Page 93: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

81

6.1 Introduction

Transport of polymer solutions in constricted spaces [1] is a long-standing research

topic with many applications including polymer enhanced-oil-recovery, size exclusion

chromatography [153] and gel-electrophoresis [151] and, recently, single DNAmolecule

analysis using micro- and nano-fluidic devices [140, 125]. In many fluidic devices

designed for separation, manipulation, and sequencing of DNA, the critical dimension

of the constriction approaches the radius of gyration of the polymer molecule or

smaller. In particular, capturing the interaction between polymer dynamics and fluid

motion in a confined geometry is essential to a proper description of these devices.[1]

The purpose of the present work is to introduce a computational approach to the

dynamics of polymers in complex confined geometries, and to apply the approach to

an interesting recent set of experiments on the flow of DNA solutions over arrays of

nanopits [36].

Many devices with well-defined microstructure have been proposed for DNA ma-

nipulation, especially for separation and stretching purposes using electrophoresis[45].

By contrast, much less attention has been paid on pressure (flow) driven DNA dy-

namics in microfabricated devices [141, 36, 143]. There are significant differences

between the electrophoresis case and the pressure driven case. Consider for example

DNA through a slit geometry. In electrophoresis, in a uniform field, the velocity of

each DNA segment is the same everywhere in the channel; by contrast, the unper-

turbed velocity profile is parabolic across the channel in the pressure driven case at

low Reynolds number, and this velocity gradient can give rise to interesting transport

phenomena, such as Taylor dispersion [141]. Furthermore, hydrodynamic interactions

(HI) between objects such as polymer segments in an unconfined domain are long-

Page 94: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

82

ranged, leading to strong many-body effects, for example, Zimm scaling of the self-

diffusion coefficient of a polymer chain. In confined geometries, the long-ranged nature

of hydrodynamic interactions changes substantially, leading to significant changes in

polymer dynamics. [1]. In a slit geometry, for example, the velocity field due to a

point force perpendicular to the walls decays exponentially. For a force parallel to the

walls, the velocity field has a parabolic form in the wall-normal direction and decays

as 1/r2. This is still a slow decay, but the symmetry of the flow leads to cancellations

upon averaging that result in screening [9, 144, 14]. Both experiments and simula-

tions of the diffusion of long flexible DNA molecules in slits [31, 78, 75] are consistent

with screening of hydrodynamic interactions on the scale of slit height. Nevertheless,

on smaller scales, HI are not screened and lead to changes in segment mobilities and

boundary effects such as cross-stream migration, which leads to depletion layers much

larger than the equilibrium chain size [77, 90, 30, 68, 29, 81].

Several recent computational studies investigated the effect of HI on dynamics

of polymer molecules in complex geometries.[54, 73, 93, 66, 154, 29, 56] Among the

problems studied in these works, forced translocation of a polymer through a nanopore

has received substantial attention because of its potential applications to rapid DNA

sequencing. In this problem of translocation, one end of a polymer molecule is placed

at the entrance to a nanopore embedded in a membrane, and then the polymer

molecule is pulled through the pore by a constant force exerted on that leading end

segment or on the chain segments in the pore. Using different simulation methods

and/or different polymer models, Izmitli et al. [73], Hernandez-Ortiz et al.[66], and

Fyta et al.[54] all found that HI shortens the translocation time and alters the scaling

exponents for the power-law dependence of the translocation time on polymer length

Page 95: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

83

[54, 73]. Using the general geometry Ewald-like method (GGEM)[67], Hernandez-

Ortiz et al. studied the effect of HI on polymer solutions with finite concentration

flowing over a channel with grooves oriented perpendicular to the flow direction was

considered.[69] At low concentration and high Weissenberg number Wi, which is

defined as the polymer relaxation time times fluid deformation rate, the groove was

almost completely depleted of polymer chains. At concentration approaching overlap,

the concentration difference between the bulk and groove is substantially reduced, but

only if hydrodynamic interactions were included in the simulation. All these studies

suggest that HI is not only important but essential in the simulation of polymer

solutions in complex geometries.

The specific system of interest in the present work is a nanoslit device that exploits

partial confinement to manipulate a DNA molecule. [116, 36] In the experiments, the

height of the channel is H = 100 nm, and the bottom wall of the slit contains square

nanopits (100 nm deep and 500–1000 nm wide). When a solution of λ−DNA, which

has a radius of gyration Rg of 770 nm, is introduced into such a system, the DNA

is highly confined as reflected by the confinement ratio Rg/H = 7.7. Because the

confinement is weaker in the regions of the slit where there is a pit, DNA segments

are preferentially found there. This entropic well tends to pin sections of the chain in

these regions (Fig. 6.1(b)). When multiple pits are present (and not too far apart, 1-2

µm in experiments), stretched sections of DNA will span the more confined regions

between pairs of pits. When a pressure drop is applied across the channel, the DNA

molecule is transported downstream by the flow, hopping between neighboring pits

(Fig. 6.1(c)). When a molecule gets trapped in a pit, a free end of the molecule

first diffuses out of the pit, then is dragged downstream by the flow. When the free

Page 96: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

84

end hits the next pit, the rest of the chain is dragged along into the second pit.

In Ref. [36], an equilibrium model was proposed, based on the hypothesis that the

dynamics of DNA at low Peclet number, which is defined as the ratio between the

diffusive and convective time scale, is thermally activated, and thus the Kramer’s

theory of barrier crossing could be applied. There are two important issues that

were not addressed in Ref. [36]. First, no direct evidence was shown to support the

hypothesis that the dynamics is thermally activated. Specifically, no analysis about

the distribution of the residence time in each pit as a function of Peclet number

(fluid velocity) was performed. Second, how the chain dynamics change with DNA

size was not systematically examined. We address these issues using an accelerated

Brownian dynamics method for modeling the hydrodynamics of polymer solutions

in complex geometries. This method is based on Green’s function solutions of the

Stokes equations, combined with a fast Stokes solver and a variant of the immersed

boundary method.

This paper is organized as follows: A general discussion of the wide range of tech-

niques for modeling the hydrodynamics of polymer solutions in complex geometries is

given in Sec. 6.2. In Sec. 6.3, we present the DNA model, governing equations, and

the new immersed boundary method that we have developed to calculate fluctuating

hydrodynamics in complex geometries. Sec. 6.4 gives results from the application of

the algorithm to the problem of flow-driven DNA dynamics in a nanofluidic device

with embedded nanopits array, and some concluding remarks are given in Sec. 6.5.

Page 97: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

85

Figure 6.1: DNA flowing through a nanoslit with embedded nanopit arrays (repro-duced with permission from Ref. [36]. Copyright 2009 by the Institute of Physics) (a)Scanning electron micrograph of typical 1× 1 µm square pits embedded in a nanoslitat 1 µm intervals. (b) Epifluorescent images of stained λ-DNA molecules confined ina 107 nm deep slit embedded with 1× 1 µm pits spaced 2 µm apart. (c) Fluorescentimages of λ-DNA travelling across a nanopit array under an applied pressure of 40mbar in the same device as shown in (b).

Page 98: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

86

6.2 Methods for hydrodynamics of confined poly-

mer solutions

In this section we give a brief overview of various methodologies for modeling hydro-

dynamics of polymer solutions in complex geometries. For theoretical analysis, the

reader is referred to Graham.[1]

The various methods for computational hydrodynamics of polymer solutions can

be divided into two major categories. The first category is purely continuum methods

that use constitutive equations for the polymer solution. These models only apply at

macroscopic scales, when the number of polymer chains in an elementary fluid flow

volume is large, so that statistical averages of the chain conformations can be used

as parameters in constitutive models of the time-dependent stress as a function of

the strain rate history. The second major category of methods explicitly simulates

the motion of each polymer chain using a coarse-grained (bead-spring) model of the

chains. Multibead representations of the chains are capable of complex chain confor-

mations but require models for the bead-bead interactions. There are also two major

types of approaches for dealing with the fluid dynamics of the solvent and thus the

hydrodynamic interactions between chain segments. One approach, exemplified by

Brownian dynamics (BD) and Lattice Boltzmann (LB) methods, uses a continuum

model (Stokes, Navier-Stokes or Boltzmann equation) to represent the solvent. The

other uses a discrete, particle-level, model to represent the solvent; examples include

full atomistic molecular dynamics (MD), Stochastic Rotational Dynamics (SRD), and

Dissipative Particle Dynamics (DPD). As we are interested in long-time (¿ 1 second)

dynamics of a single polymer molecule in complex geometries, this brief review focuses

Page 99: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

87

on mesoscopic methods, namely, BD, LB, SRD, and DPD, especially their different

treatments of hydrodynamics, boundary conditions, and Brownian fluctuations.

In Brownian dynamics simulations of polymer molecules in unbounded or peri-

odic domain, the HI are usually incorporated using analytical Green’s functions for

Stokes’ equations: I.e. the beads exert point forces on the fluid and Green’s functions

are used to determine the fluid velocity at any other bead position. For a system with

N beads, these interactions are incorporated in theN×N mobility matrix[105]. While

for true point forces, the Green’s function for the Stokes equations – the Oseen tensor

– is appropriate, in Brownian dynamics simulation it is necessary to use regularized

point forces. The most common regularization is the Rotne-Prager-Yamakawa tensor

[120, 157], while in the present work a different, computationally convenient regu-

larization is used. Corresponding to the mobility matrix representation of the HI, a

matrix formulation of the fluctuation-dissipation theorem – a factorization of the mo-

bility tensor – must be implemented. Ermak and McCammon suggested a Cholesky

factorization which requires O(N3) operations[49], while Fixman [53] observed that

the square root matrix of the mobility matrix can be accurately approximated via

Chebyshev polynomials, and this algorithm brings down the computation complexity

to roughly O(N2) or O(N2.25).[76] Another advantage of Fixman’s algorithm, one

that is essential to this study, is that it works even when an analytical form of the

Green’s function is not available, as is the case for flow in complex geometries. For

periodic domains, the scaling can be dramatically improved, to O(N logN), by using

particle-mesh Ewald (PME) methods, which are based on Hasimoto’s solution for the

Green’s function for Stokes flow driven by a periodic array of point forces [65, 16].

BD can also be efficiently applied to nonperiodic geometries. In simple cases

Page 100: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

88

like flow in a half-plane bounded by a no-slip wall, the exact Green’s function for

that geometry can be used. To handle more complex boundary conditions, such as

indentations or bumps on walls, it requires incorporating other methods such as finite

element methods [78, 69], which can be done in the framework of the general geometry

Ewald-like method (GGEM) [67], which also has O(N logN) or O(N) scaling. In

the present work, we combine GGEM with a variant of the Immersed Boundary

Method.[110, 97, 114]

The lattice Boltzmann method is a mesoscopic approach to fluid flow calculation

and is based on a discrete version of the Boltzmann equation with enough detail to

recover hydrodynamic behavior [47]. The LBM avoids the some of the computational

complexities of Navier-Stokes (NS) solvers and is straightforward to parallelize effi-

ciently. Using a Chapman–Enskog expansion, the LB equation can recover the NS

equations for small finite Mach and Knudsen numbers, and, within these limits it is

second-order accurate in space and time. Coupling between the LB for the solvent

and the molecular dynamics for the polymer is achieved through Stokes drag forces.

The Brownian fluctuation is introduced via random fluctuations added to the stress

tensor and the beads dynamics directly. [47] Usually, the no-slip boundary is intro-

duced by a bounce-back collision rule in which incoming fluid particles are reflected

back towards the nodes from which they originated, and this results in no-slip bound-

ary conditions. For complex boundaries, however, the boundary implementation may

not be trivial [85, 33, 28]. The Lattice Boltzmann-molecular dynamics scheme has

been successfully applied to the problems of polymer translocation through a pore

[54, 73, 93], polymer cyclic dynamics near a solid wall [158], and polymer migration

in a square channel [81].

Page 101: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

89

Stochastic rotation dynamics is a relatively recent particle-based mesoscopic sim-

ulation method. In the SRD algorithm, the fluid is modelled as a system of point

particles. The algorithm consists of two steps, the streaming step and the collision

step.[80, 59] In the streaming step, the particles move ballistically during a certain

amount of time. In the collision step, the particles are sorted into cubic cells of a cer-

tain size. In each cell, the velocity vector of every particle is rotated an angle α about

a randomly chosen axis, relative to the center-of-mass velocity of all the particles

within that particular cell. In addition, a random shift is added in order to ensure

Galilean invariance. The polymer segments are coupled to the fluid by including

them in the collision step. The streaming and collision steps are designed to conserve

mass, momentum, and energy. In the long time limit, correct hydrodynamics are

recovered. SRD also naturally contains Brownian fluctuations. To implement no-slip

boundary conditions in SRD, there are two approaches: the bounce-back scheme and

the stochastic scheme. The bounce-back scheme was first proposed by Malevanets

and Kapral [91], and is a direct analog of the no-slip boundary condition commonly

applied in LBM. Lamura and Gompper found that the simple bounce-back rule is

not sufficient to guarantee no-slip without corrections. [87] In the stochastic scheme,

proposed by Inoue et al., the velocity of the SRD particle that comes into a wall

cell is changed to a random velocity obeying a Maxwell-Boltzmann distribution[72].

This boundary condition, however, allows flow to penetrate small objects.[60] Watari

et al.[154] and Chelakkot et al.[29] have used this method to study the dynamics

of polymer solutions between two parallel planes, and qualitatively reproduced the

hydrodynamic migration phenomenon.

Dissipative particle dynamics coarse-grains the solvent molecules to obtain a sys-

Page 102: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

90

tem of weakly repulsive (soft) spheres interacting with a mixture of conservative,

stochastic, and dissipative forces.[50, 64, 118] The conservative forces can be used to

reproduce the solvent equation of state, while the dissipative forces model viscous

friction. The stochastic forces act as a thermostat that ensures detailed balance and

correct thermal fluctuations in the DPD fluid. The method has great flexibility and

requires significantly fewer solvent particles and larger time steps than classical MD.

The influence of solid boundaries is typically very strong in DPD simulations. The

modeling of no-slip boundaries in the context of DPD has been addressed in the past

by freezing the particles that represent the solid boundaries[117], and by modifying

the repulsive forces[79]. While these efforts introduce minimal additional algorithmic

complexity to the standard DPD formulation, care has to be taken to avoid numer-

ical artifacts such as depletion of particles near the wall or artificial ordering of the

near-wall particles leading to large density fluctuations[112].

To study the dynamics of polymer solutions in complex geometries, the ability to

treat no-slip boundary conditions efficiently and accurately becomes important. In all

three mesoscopic models (LBM, SRD, and DPD), it is not a trivial task to correctly

implement the no-slip boundary conditions. Another difference between BD and the

three mesoscopic models is that those methods describe compressible inertial fluids

in which hydrodynamics need time to propagate through the solution; the Reynolds

and Mach numbers Re and Ma are finite. In Green’s function-based methods, on the

other hand, the condition Re ≪ 1, Ma ≪ 1 are satisfied explicitly through the use

of the Stokes equation.

In the present study, we propose an immersed boundary method-based Green’s

function method, where the boundary is represented by many regularized points and

Page 103: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

91

the Stokes equation is solved on a Cartesian grid. The validity and the accuracy of

the new method are scrutinized by simulating laminar flow in a slit, uniform flow

around a solid sphere, and flow generated by a Stokeslet near a single wall.

6.3 Polymer model and simulation method

6.3.1 Model and governing equations

In this section, we present the coarse-grained model used in this work to study the

dynamics of polymer in complex geometries. We focus attention on a model of λ-

DNA[158, 76, 74], described by a bead-spring chain model composed of Nb beads

of hydrodynamic radius a = 77 nm connected by Ns = Nb − 1 entropic springs. a

is also the length unit used through out this work for nondimensionlization. Each

bead represents a DNA segment of 4850 base pairs, i.e., Nb = 11 corresponds to

a fluorescently stained λ-DNA, which has a contour length of 21 µm and radius of

gyration of 770 nm.[136, 137] The length unit in this work is a = 77 nm. The spring

connecting the beads i, j obeys a worm-like chain force law [94]

f sij =kBT

2bk[(1− |rj − ri|

Nk,sbk)−2 − 1 +

4|rj − ri|Nk,sbk

]rj − ri|rj − ri|

. (6.1)

Here, bk is the Kuhn length for DNA and Nk,s is the number of Kuhn length per

spring. The physical confinement, or the steric interaction between DNA and wall, is

Page 104: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

92

taken into account through an empirical bead-wall repulsive potential of the form

Uwalli =

Awallb−1k δ−2

wall(hi − δwall)3, hi < δwall

0, hi > δwall.

(6.2)

Here hi represents the perpendicular distance of bead i from the wall, δwall is the cut-

off distance. In this work, we choose Awall = 50kBT/3 and δwall = 1a. The excluded

volume potential between two distinct beads is

Uevij =

1

2νkBTN

2k,s(

3

4πS2s

)3/2 exp(−3|ri − rj|2

4S2s

), (6.3)

where ν is the excluded volume parameter, and S2s = Nk,sb

2k/6.

All of the parameters a, bk, ν are the same as used in previous works[76, 74],

where it has been shown to successfully reproduce the static (radius of gyration, Rg)

and dynamic properties (diffusivity D and longest relaxation time τ0) of DNA with

contour length of 10− 126µm.

6.3.2 Governing equations

We are concerned with the numerical simulation of the Brownian motion of a single

DNA molecule immersed in an isothermal incompressible Newtonian solvent in the

limit of zero Reynolds number. When the DNA molecule is represented by the bead-

spring model as described above, the equation of motion for each bead is determined

by the balance of Stokes drag on the bead due to the surrounding fluid, the Brownian

forces exerted by the fluid on the bead, and other non-hydrodynamic, non-Brownian

intra/inter molecular forces. The force balance on the beads leads to a stochastic

Page 105: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

93

differential equation which governs the dynamics of the chain and can be solved

numerically[105]:

dR = (U∞ +M · F+ kBT∂

∂R·M)dt +

√2B · dW, (6.4)

B ·BT = kBTM. (6.5)

Here R is the vector containing all bead positions ri, kB is the Boltzmann constant

and T is the absolute temperature. The vector containing total non-Brownian, non-

hydrodynamic forces acting on the beads is denoted F. From Eq. 6.4 we can see that

there are four contributions to the change of chain configuration after a time step

dt: unperturbed flow velocity U∞ in complex geometries, perturbed velocity M · F

due to the polymer forces, drift due to the configuration dependence of the mobility

tensor ∂∂R

·M, and Brownian displacement√2B ·dW. The unperturbed flow velocity

U∞ at R is the velocity when DNA is absent in the system.

As noted, the second term in Eq. 6.4, M · F, is the velocity generated by the

motion of the polymer beads. It arises from the fact that beads immersed in a fluid

generate flows as they move due to various forces, and similarly they move in response

to fluid motion through the Stokes drag. As described in last section, we treat bead i

of the bead-spring chain as a sphere of hydrodynamic radius a. Then the relationship

between the bead velocity ui and the drag force it exerts on the fluid is given by

the Stokes law fi = ζ(ui − u(ri)), where ζ is the bead friction coefficient ζ = 6πηa,

u(ri) is the fluid velocity at the bead position, and η is the fluid viscosity. (Note

that the finite size of the bead only arises in the friction coefficient.) Through these

hydrodynamic interactions (HI), beads interact with each other and with the walls

Page 106: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

94

of the confining geometry. The mobility tensor M, will be discussed in detail in Sec.

6.3.3.

For systems with configuration-dependent mobility tensors, the third term in Eq.

6.4 has to be included to obtain the correct dynamics. This is because the configura-

tion dependence of the mobility tensor results in a mean drift with a mean velocity

∂∂R

·M.[61] In complex geometries, the expression for M is lacking. Thus to solve Eq.

6.4, we need to turn it into a derivative-free form, and this will also be discussed in

Sec. 6.3.3.

The tensor B gives the magnitude of the Brownian displacement of the polymer

beads, and is coupled to M by the Fluctuation-Dissipation theorem (Eq. 6.5). The

vector dW is a random vector composed of independent and identically distributed

random variables according to a real-valued Gaussian distribution with mean zero

and variance dt. We will use the Chebyshev approximation given in Sec. 6.3.4 to

calculate B · dW.

6.3.3 Mobility tensor and Fixman’s midpoint algorithm

To illustrate the components of the mobility tensor, we consider a two-bead chain

(a dumbbell) in an unbounded domain, neglecting Brownian effects for the moment.

Here the force balance shows that the fluid velocities u(ri) experienced by bead i are

given by

u1

u2

=

u∞1

u∞2

+

1ζI G∞(r1 − r2)

G∞(r2 − r1)1ζI

·

f1

f2

, (6.6)

Page 107: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

95

where G∞(r) = 18πηr

(I + rr/r2) is the Oseen-Burgers tensor, the free-space Green’s

function for Stokes’ equations, r = |r|, and I is the identity matrix. The quantity

G∞(r1 − r2) · f2 gives the velocity at position r1 generated by the force f2 exerted by

the particle at r2 on the fluid. We can write above expression succinctly as,

U = U∞ +M · F. (6.7)

In a confined geometry, a wall correction GW (r1, r2) has to be added to the mobility

tensor M, and it becomes

M =

1ζI+GW (r1, r2) G∞(r1 − r2) +GW (r1, r2)

G∞(r2 − r1) +GW (r1, r2)1ζI+GW (r1, r2)

. (6.8)

In general geometries, an analytical expression for GW (r1, r2) is not available, but

has to be calculated numerically.[78] In particular, we will see that it is not M that

is needed, but rather, the product M · F, which is the velocity generated by the

total non-Brownian, non-hydrodynamic forces acting on the beads. Using an accel-

erated Immersed Boundary Method, we can calculate M · F for polymer in complex

geometries efficiently, as discussed in Sec. 6.3.5.

Fixman proposed a method to numerically integrate Eq. 6.4 without calculating

the derivative term ∂∂R

·M.[52] The idea is to approximate the first-order derivative

Page 108: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

96

with a finite difference approximation, and transform Eq. 6.4 into

R∗ = R+√2B(R) ·∆W,

∆R = [U∞ +M(R) · F]∆t +

√2

2kBTM(R∗) · [(B−1(R))

T ·∆W]

+

√2

2B(R) ·∆W, (6.9)

where ∆W is a vector of independent Gaussian random variables with variance ∆t.

A derivation of this algorithm is given in Appendix D.

6.3.4 Chebyshev approximation

In Eq. 6.9, there are three matrix-vector multiplications that need to be evaluated:

M · x, (B−1)T · x, and B · x, where x may be any real vector. The Brownian ten-

sor B is coupled to the mobility tensor M by the Fluctuation-Dissipation theorem

kBTM = B ·BT. A common choice for B is the square root matrix S, satisfying

kBTM = S · ST with S = ST. With that, we only need to evaluate M · x and S−1 · x,

because (S−1)T = (ST)−1 = S−1 and S · x = S · (S · S−1) · x = kBTM · (S−1) · x. Given

an algorithm to evaluate M · x, Fixman noted that one can evaluate S−1 · x using

Chebyshev polynomial approximation over the range [λmin, λmax], where λmin and

λmax are the minimum and maximum eigenvalues of M respectively [76, 53]. The

Page 109: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

97

approximation yL to S−1 · x is a series of matrix-vector multiplications,

yL =L∑

0

alxl, (6.10)

x0 = dw, (6.11)

x1 = [daM+ dbI] · dw, (6.12)

xl+1 = 2[daM+ dbI] · xl − xl−1. (6.13)

Here al are the Chebyshev coefficients of the scalar function 1/√x over the domain

[λmin, λmax], and da = 2/(λmax−λmin)and db = −(λmax+λmin)/(λmax−λmin). From

the above equations, we see that rather than M, all that we need is the product

of M and a vector, which can be calculated efficiently by an Immersed Boundary

Method as will be described in the next section. To calculate the eigenvalues λ of

M, we use power iteration to calculate λmax and Arnold iteration (ARPACK) to

calculate λmin.[147] Both algorithms require only dot product calculations between

M and some vector x.

In summary, with a fast algorithm to evaluate M · x for arbitrary x, we can

simulate the Brownian movement of a bead-spring polymer in complex geometries.

This fast algorithm is discussed in next section.

6.3.5 Fast Stokes solver with complex boundary conditions

Note that M · F is the fluid velocity generated by the the vector of point forces, F.

Thus, to calculate M · F, we need to solve the Stokes flow equations with distributed

Page 110: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

98

point forces with no-slip boundary conditions:

−∇p(x) + η∇2u(x) = −∑

fnδ(x− xn), (6.14)

∇ · u(x) = 0, (6.15)

u(r) = 0, r ∈ ∂Ωb, (6.16)

where p is the pressure, u is the fluid velocity, η is the fluid viscosity, fn is the force

exerted on the fluid at point xn, δ(x) is the three-dimensional delta function, and

∂Ωb is the boundary of the fluid domain. In the actual implementation, regularized

point forces are used, as further discussed below.

Provided a fast Stokes solver for distributed point forces within a periodic do-

main (GGEM), we developed an accelerated Immersed Boundary Method (IBM) for

the Stokes equations with complex boundary conditions by representing the no-slip

boundary as forcing term (momentum source) in the Stokes equation. To employ

the fast Fourier transform (FFT) technique to speed up the calculation, we chose a

periodic domain. We present first the IBM algorithm, then the GGEM algorithm.

Immersed boundary method

A generic three-dimensional domain is shown schematically in Fig. 6.2. Polymer

beads and solid boundaries are embedded in a rectangular periodic domain Ω =

[−Lx/2, Lx/2] × [−Ly/2, Lx/2 × [−Lz/2, Lz/2]. The beads of the polymer molecule

are represented by the filled symbols, and their position vector is R. The beads

exert forces obeying Stokes law on the surrounding fluid. A no-slip boundary ∂Ωb

is represented by regularized point forces located at boundary nodes Rb, which are

Page 111: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

99

Figure 6.2: Schematic of the immersed boundary method. The system is a threedimensional periodic domain, in this study, a rectangular parallelepiped of Lx×Ly ×Lz. The beads of the polymer molecule is represented by the filled symbols, and theirposition vector is R. The no-slip boundary ∂Ωb is represented by regularized pointforces located at boundary nodes Rb, which are uniformly spaced with a boundarymesh size of h.

Page 112: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

100

uniformly spaced with a boundary mesh size of h. These boundary nodes do not

move relative to the fluid, but rather exert forces on the fluid such that the velocity

of fluid at Rb is zero. Hence, the no-slip boundary condition is satisfied at each nodes.

The advantages of IBM include its ease of programming due to use of a Cartesian

grid, and its potential for parallel processing as the boundary nodes are treated in a

similar fashion as the polymer beads.

Denoting Mibm as the mobility tensor for all the regularized point forces in the

periodic domain, i.e., polymer beads and boundary nodes, the velocities generated by

those point forces are,

Mibm · Fibm = Uibm. (6.17)

Partitioning Mibm into smaller tensors for polymer-polymer, polymer-boundary, and

boundary-boundary interactions, the relationship between forces and velocities at the

polymer and boundary points is given by

Uibm =

Up

Ub

=

Uuni

Uuni

+

Mpp Mbp

Mpb Mbb

·

Fp

Fb

, (6.18)

where p and b denote polymer and boundary respectively, and Uuni is the uniform

background velocity which is generated by a constant pressure drop. Here Mpp and

Mbb are symmetric and positive definite (and thus invertible) matrices. The quantities

Mpb · Fp and Mbp · Fb are the velocities at boundary points generated by polymer

forces Fp and the velocities at polymer points generated by boundary forces Fb,

respectively, and we are able to calculate them using periodic GGEM, which will be

discussed in the next section. Note that the unperturbed Stokes flow velocity U∞ in

Page 113: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

101

a complex geometry is the sum of the uniform background flow Uuni and a correcting

flow generated by the boundary points Ubr, U∞ = Uuni + Ubr. From Eq. 6.18 we

obtain:

Up = Uuni +Mpp · Fp +Mbp · Fb, (6.19)

Ub = Uuni +Mpb · Fp +Mbb · Fb = 0. (6.20)

Noting that Mbb is invertible, we obtain:

Fb = (Mbb)−1 · (−Uuni −Mpb · Fp), (6.21)

Up = Uuni +Mbp(Mbb)−1 · (−Uuni) +Mpp · Fp −Mbp(Mbb)−1Mpb · Fp

= Uuni +Mbp(Mbb)−1 · (−Uuni) + [Mpp −Mbp(Mbb)−1Mpb] · Fp. (6.22)

Comparing the above equation to U∞ = Uuni +Ubr, we obtain

Ubr = Mbp(Mbb)−1 · (−Uuni). (6.23)

By construction, Eq. 6.20 and Eq. 6.21 also suggest an algorithm to calculate Up:

(1) Given the polymer configuration R(tn), evaluate polymer forces Fp(tn) includ-

ing spring forces, excluded volume interactions, and repulsive wall-bead interactions;

(2) Calculate the perturbed velocities Upb = Mpb ·Fp generated by polymer forces

at the boundary points;

(3) Determine the boundary forces Fb(tn) that satisfy Mbb · Fb(tn) = −Uuni −

Upb(tn) (Eq. 6.21);

(4) Evaluate the polymer bead velocities Up = Uuni +Mpp · Fp +Mbp · Fb.

Page 114: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

102

Step (3) in this algorithm requires an efficient iterative scheme to calculate the

boundary forces that specify the velocity field in the domain we are interested in. In

other words, our goal is to solve Mbb ·X = U, where U is given and X is unknown.

The number of iterations required to reduce the relative error to meet some specified

tolerance is a function of the condition number ofMbb (Figure 6.3(a)), which is defined

as the ratio of the largest over the smallest eigenvalues ofMbb. We found that a simple

conjugate gradient method [123] is sufficient. First, in the algorithm of conjugate

gradient method, only matrix-vector multiplications are calculated. Second, it can

reduce the relative residual ||Mbb · Fb − U||/||U|| (Eq. 6.21) to less than 10−4 in

tens of iterations as shown in Fig. 6.3(b), even when the condition number of Mbb is

very large. It should be noted that a similar approach, in a non-Brownian boundary

element context, has been proposed by Zhao et al. to study the flow of blood cells in

complex geometries.[159]

Periodic GGEM

The core of our algorithm is a fast solver for the Stokes equations with distributed

point forces. In this section, we present the general geometry Ewald-like method

(GGEM), which can serve as a “blackbox” to evaluate M · x for a given point force

distribution in periodic domain.[67, 68, 66, 114] For periodic boundary condition,

Ewald sum and particle-mesh Ewald (PME) methods are available, which are based

on the Hasimoto’s solution for Stokes flow driven by a periodic array of point forces.

These methods reduce the operation to O(N logN), compared with O(N2) opera-

tions of direct evaluation of all the pairwise interactions.[65, 132, 124] Fast multipole

method is another alternative, and the complexity is O(N).[145] GGEM is also an

Page 115: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

103

Figure 6.3: Properties of the Mbb tensor for the case where a spherical boundary (R =3) is represented by uniformly distributed points on the surface and the parameters forthe calculation are Lx = Ly = Lz = 10, dx = dy = dz = 0.25, ξ = 4.0. (a) Conditionnumbers as a function of number of points on the sphere shell. (b) Relative residualas a function of number of iterations.

Page 116: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

104

O(N logN) method when fast Fourier transforms are implemented.

The central idea that makes GGEM a fast algorithm is separating the point force

density ρ = Σfnδ(x− xn) into a local part ρl = Σfn(δ(x− xn) − g(x− xn)) and a

global part ρg = Σfng(x− xn) by introducing a screening function g(x − xn). In

this work, the screening function g(r) is a modified Gaussian with α as the screening

parameter

g(r) = (α3/π3/2)e(−α2r2)(5/2− α2r2), (6.24)

where r is the radial position vector and r = |r| (Fig. 6.4). This smooth function

satisfies

g(r)dr = 1 for any value of α, and when α → ∞ it becomes a three-

dimensional Dirac delta function. Correspondingly, the velocity field generated by ρ is

separated to a local contribution and a global contribution u = ul+ug. We chose g(x−

xn) so that an analytical expression for ul is available and ul decays exponentially

fast. Thus, the interactions due to ρl are short ranged, and only interactions within

some cut-off distance Rc need to be considered in the calculation of ul. In the present

case, the global velocity ug is going to be calculated by solving the Stokes equations

on a structured mesh (grid) using three-dimensional fast Fourier transform. Note

in contrast to conventional Ewald-based methods, FFT is not necessary to achieve

computational efficiency in GGEM, so the GGEM approach is not limited to periodic

domains.[67, 69, 148, 114]

Local velocity field Consider first the local velocity at x, ul(x), which results

from the local force density ρl,

ul(x) =∑

n

Gl(x− xn) · fn, (6.25)

Page 117: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

105

Figure 6.4: Screening function g(αr) (Eq. 6.24) for the GGEM algorithm.

where the “local Green’s function” is

Gl(x) =1

8πη(I+

xx

r2)erfc(αr)

r− 1

8πη(I− xx

r2)2α

π1/2e(−α2r2). (6.26)

Here, r = |x| and I is the identity matrix. In this work, to avoid singularity associated

with delta point forces, we use regularized point forces by replacing δ(r) with g(r)

(Eq. 6.24) with a regularizing parameter ξ and the corresponding regularized local

Green’s function is

GRl (x) =

1

8πη(I+

xx

r2)(erf(ξr)

r− erf(αr)

r)+

1

8πη(I− xx

r2)(

π1/2e(−ξ2r2)− 2α

π1/2e(−α2r2)).

(6.27)

When x → 0,

GR0

l =1

8πη(4ξ√π− 4α√

π)I, (6.28)

and the velocity at x = 0 stays finite. From Eq. 6.27, we see that because of the

presence of the screening function, velocity decays exponentially over a distance pro-

Page 118: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

106

Figure 6.5: (Free space) Comparison of the x component of the velocity field drivenby a delta point force (lines) acting along +x-axis and by a regularized delta pointforce (symbols) with a form of modified Gaussian (Eq. 6.24, α = 2.0). (Top) alongx-axis. (Bottom) along y-axis.

portional to 1/α, provided that α ≪ ξ. Calculation of the local velocity field begins

by identifying point forces located within the cut-off distance Rc = 4/α for a given

x. Any interactions beyond Rc is ignored. The value of α is empirically determined,

and is a constant for certain simulation. To make sure the local velocity satisfies the

periodic boundary conditions, we apply the minimum image convention which states

that, in the periodic system, the cutoff distance must be smaller than half the shortest

box length. In Fig. 6.5, we compare the velocity generated by a delta point force and

a regularized point force.

The regularizing parameter ξ is different for the polymer beads and the boundary

nodes. Consider first the boundary nodes. As with the conventional IBM [110, 97],

the boundary force density is distributed onto the grid, which is uniformly spaced

with a mesh size of h, through a regularized delta function. In this work, we choose

the modified Gaussian as in Eq. 6.24, and we choose ξ such that ξh = O(1). This

Page 119: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

107

ensures that the force density associated with each grid node is spread over the length

scale of the associated elements, thereby preventing fluid from penetrating the solid

boundaries and also preventing unphysically large fluid. For the polymer beads, the

choice for ξ is straightforward and is inversely proportional to the hydrodynamic

radius of the polymer bead, ξa ≈ 1. In the present work, we choose ξ so that

ξa = 3/√π, with which the maximum fluid velocity is equal to that of a particle with

radius a and the pair mobility remains positive-definite.[69]

For any point x in the fluid that is not on the boundary or a polymer bead, the

local contribution to the velocity is given by

ul(x) =

n∑

i

GRl (x− xi) · f bi +

m∑

j

GRl (x− xj) · fpj . (6.29)

Here, n and m are the number of boundary points and polymer beads, respectively,

which lie within the cut-off distance centered at x. For a boundary point, the local

contribution is

ul(xk) =

n∑

i

GRl (x− xi) · f bi +

m∑

j

GRl (x− xj) · fpj . (6.30)

Note that in the first term, the i = k term requires evaluation of GRl (0), which is

given by Eq. 6.28.

For a polymer bead, the local contribution is

ul(xk) =n

i

GRl (x− xi) · f bi +

m∑

j,j 6=k

GRl (x− xj) · fpj . (6.31)

The exclusion of the “self-term”, j = k, in the polymer local velocity calculation is

Page 120: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

108

because in the Brownian dynamics simulation polymer beads do not move with the

fluid velocity. Thus the velocity “seen” by a polymer bead should not include the

velocity generated by the bead itself as it moves through the fluid.

Global velocity field The velocity field driven by the global part of force density

ρg is obtained by solving the Stokes equations with periodic boundary condition on

a grid using fast Fourier transform (FFT).

The continuous force density g(r) has to be replaced by a grid based force density.

In this work, we use the modified Gaussian function to distribute the force density

onto the grid points which are within the cut-off distance Rc. This is simply a Fourier

collocation approach to the global problem.[27] Again, we use periodic boundary

condition with minimum image convention to make sure the forcing term is periodic

in all three spatial directions.

The linearity of Stokes equation allows us to exploit the efficiency of FFT. We

assume the velocity, forcing term, and pressure gradient are periodic in all three

spatial directions. The velocity vector U(x) and pressure gradient ∇p(x) at the grid

point x = (x1, x2, x3) are

u(x) =∑

kuke

−ik·x,

∇p(x) =∑

kpke

−k·x. (6.32)

Here, k = (k1, k2, k3) is the wavenumber vector and k2 = k21 + k2

2 + k23. Now we apply

Fourier transform to the Stokes equations Eq. 6.15 and obtain

pk − ηk2uk = −f , ik · uk = 0. (6.33)

Page 121: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

109

When k 6= 0, taking the scalar product of the Fourier transformed momentum equa-

tion with ik yields

pk = −(k · f)k2

k, (6.34)

uk =f

ηk2− (k · f)

ηk4k.

When k = 0, p0 = −f , which means that the force exerted on the fluid is balanced

by the mean pressure gradient of the field. Then the global velocity field is obtained

by inverse FFT. To get the velocity at the polymer points from the velocity on the

mesh, we perform back-interpolation using quadratic Lagrange polynomials. It is this

interpolation step that ultimately controls the order of accuracy of the solution.

Finally, note that in the global velocity calculation for polymer beads, we also

have to exclude the self-interaction term. This velocity field is determined by the

free-space regularized Stokeslet (Eq. 6.26). Therefore, the total velocity seen by a

polymer bead is

u(xpj ) = ul(x

pj ) + ug(x

pj )− lim

x→0

Gl(x) · fpj . (6.35)

Validation: GGEM We investigated the accuracy of GGEM as function of screen-

ing parameter α and mesh size ∆x, against Hasimoto’s solutions of spatially periodic

Stokes equation[65, 113]. The simulation box is a cube of side length L = 10. A

delta point force acting along x direction with unit strength is placed at the center

of the box. Fig. 6.6 shows Hasimoto’s solution and the numerical solution using

GGEM for the x component of velocity at line x = 0, y = 0 and line y = 0, z = 0.

Fig. 6.7(a) shows the error as a function of screening parameter α for three different

Page 122: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

110

Figure 6.6: Comparison of Hasimoto’s result[65] (symbols) and numerical solution(lines) for the x component of velocity due to point force acting along +x directionin a periodic domain.

mesh sizes ∆x. Notice that, empirically, when 20/L < α < 0.8/∆x the error does

not depend on α. The lower bound is related to the restriction that cut-off radius

Rc = 4/α should be smaller than half of the box size, and the upper bound is about

the number of grid points which need to be included to resolve the length scale α−1.

Fig. 6.7(b) shows the error as a function of mesh size ∆x together with least squares

fitting to estimate the order of accuracy. The error is defined as the L2 norm of the

difference between GGEM results and Hasimoto’s ||E||2 = ||uGGEM − uHasimoto||2 at

200 sampling points which are randomly distributed within the simulation box. With

Lagrange back-interpolation, GGEM has order 3 of accuracy. This result does not

change with the positions of the sampling points. Unless otherwise noted, throughout

the rest of this study, we choose the screening parameter which satisfies α∆x = 0.8,

and ∆x is determined by the desired error tolerance.

Page 123: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

111

Figure 6.7: Error ||E||2 as a function of (a) screening parameter α and (b) mesh size∆x for a point force at the center of a cubic periodic domain.

Page 124: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

112

Validation: IBM We now show numerical results of three test cases using the

IBM algorithm proposed above: pressure driven flow through a slit geometry (Fig.

6.8), uniform flow around a sphere (Fig. 6.9(a)), and flow generated by a Stokeslet

near a solid plane (Fig. 6.9(b)). We use the slit case to investigate the numerical

error of IBM. For laminar flow in a slit geometry, the velocity only depends on the

perpendicular distance away from the wall, and it is parabolic Ux(z) = 3/2Uavg(1 −

(z/H)2). In the numerical calculations, the periodic domain is a cube with side of

L = 40, and the wall is represented by a square lattice with grid size of h. In all the

calculations, we used a mesh size of ∆x = 0.05L for the global velocity calculations.

Fig. 6.8(a) shows excellent agreement between the analytic solution and the numerical

solution obtained by IBM. The absolute error ||Unumerical −Uexact||2 is shown as a

function of h in Fig. 6.8(b); decays roughly as h1.2, consistent with the general result

that IBM is first-order accurate.[86] The local error is the largest near the boundary

and decays very rapidly into the flow (< 1× 10−5 on the center line).

The uniform flow around a sphere case demonstrates that our algorithm works well

for curved boundary as well. The exact solution to the flow field in an unbounded

domain is:

Ur = U∞[1− 3

2(a

r) +

1

2(a

r)3]cos(θ),

Uθ = −U∞[1− 3

4(a

r)− 1

4(a

r)3]sin(θ).

Here, θ is the angle away background flow direction, a = 1 is the radius of the sphere,

and r is the radial distance from the center of the sphere. We use a large simulation

box (L = 500) to minimize the effect of periodic boundary. When the background flow

Page 125: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

113

Figure 6.8: (a) Velocity profile of laminar flow through a slit. The line represents theanalytic profile and the symbols represent the numerical result by IBM. (b) Error ofIBM as a function of boundary grid size h and the slope is 1.22.

Page 126: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

114

is U∞ =(1, 0, 0), we use N = 200 regularized points, which are uniformly distributed

on a spherical shell of radius 1, to represent an unit sphere placed at the center of

the simulation box. The mesh size is ∆x = 1 and α∆x = 0.1 in this case. From Fig.

6.9(a) we see that IBM gives a satisfactory result.

In the case of a Stokeslet above a solid wall, numerical results are tested against

the exact results from the analytical expression of the mobility tensor for a point

force near a single solid wall, obtained by Blake[22]. The calculation is similar to that

in the slit case except that a point force is placed near a wall, and the separation

between the walls is large to minimize the effect from the other wall. The parameters

for this calculation are L = 40,∆x = 0.5, and h = 0.5. The wall is the x − y plane,

and the Stokeslet is situated at (0, 0, 2), pointing in the x direction. From Fig. 6.9(b)

we see that IBM gives excellent results for this case, which involves both force points

and boundary points.

6.4 DNA flowing across an array of nanopits

In this section, we apply the immersed boundary method described in last section to

study the dynamics of a DNA molecule flowing through a nanoslit with embedded

nanopit arrays (Fig. 6.10). The schematic of the problem is given in Fig. 6.10. The

device is periodic in the streamwise (x) and spanwise (y) directions. One periodic or

one unit cell is a rectangular parallelepiped of Lx ×Ly × 2H = 40a× 40a× 8a. Here,

a = 77 nm is the bead radius in the bead-spring model representing the DNA. One

pit is at the center of the unit cell and has a size of Ld × Ld ×H = 16a× 16a× 4a.

The pit-to-pit distance is 2Lc = 24a.

Page 127: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

115

Figure 6.9: Validation of IBM algorithm. Open symbols are numerical results andlines are analytical calculations. (a) Velocity profile Ux around a unit sphere withno-slip boundary condition. Open circles represent points along the x-axis and opensquares are for points along y-axis. (b) Stokeslet near a single wall. Velocity profileof Ux along +z-direction for various (x, y) pair. Open squares are for (x, y) = (3, 3)and open circles are for (x, y) = (1, 1).

Page 128: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

116

Figure 6.10: Schematic and the immersed boundary representation of the nanopitproblem. Gray spheres indicate points at which no-slip boundary conditions aresatisfied.

In our simulation, there are two unit cells in a three dimensional periodic simula-

tion box of 2Lx × Ly × 2H . The boundary is represented by boundary points which

are regularly spaced with a mesh size of 1a, as shown in Fig. 6.10. Correspondingly,

we chose α = 0.8a and ξ = 2.5a as the input parameters for the immersed boundary

method. A Peclet number, the ratio between the diffusive and convective time scale,

is defined as Pe = UpHD

, and the Weissenberg number is Wi = τ0γ = τ0Up

H, where Up

is the maximum x component of the unperturbed velocity in the y−z plane, D is the

chain diffusivity, τ0 is the longest relaxation time of the chain in bulk solution, and

γ is shear rate in the pit. Note that Pe and Wi are linked by the confinement ratio

Rg/H , as Wi ∼ Pe(Rg

H)2. For λ−DNA, which has a radius of gyration about 10a,

the effective confinement ratio is about 5, considering that the cut-off distance for

the excluded bead-wall interaction is 1a. For λ−DNA we use the τ0 value obtained

by Jendrejack et al.[74]. Figure 6.11 shows several streamlines of the Stokes flow and

the contour plot of the streamwise velocity in the absence of the DNA molecule.

Page 129: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

117

Figure 6.11: (Top) Streamlines in the nanopit and (Bottom) contour plot of thestreamwise velocity on the x− z plane.

6.4.1 Dynamics at low Peclet number

In the introduction we described the experimental observation of the chain hopping

between pits, and the mechanism of a free end diffusing and being dragged down-

stream from one pit to the next. This behavior is reproduced in our simulations, as

shown by the sequence of snapshots in Fig. 6.12. To study the dynamics quantita-

tively, we tracked the center-of-mass of the DNA molecule and extracted dynamic

properties of DNA from its trajectories. Fig. 6.13(a) is a typical time series plot of

the x-component of the center-of-mass xc of a λ-DNA molecule. This figure clearly

demonstrates that DNA hops from pit to pit at low Peclet number: the molecule

stays in one pit for a time interval td, which we will call the residence time, and then

jumps to the next pit. As seen in Fig. 6.13(a) at low Peclet number, the time for the

Page 130: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

118

Figure 6.12: Snapshots of a hopping event (from (a) to (f)) (top-down view, Pe = 3.5).

Page 131: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

119

molecule to get to the next pit is very small compared to the residence time. The

probability distribution of td is well-fit to an exponential ρ(td) = 1/τ exp(−td/τ),

as shown in Fig. 6.13(b). This result indicates that, to a good approximation, the

hopping events are independent of one another—the hopping dynamics is a Poisson

process at low Pe, and the rate parameter which characterizes the Poisson process is

1/τ , where τ is the mean residence time.

Intuitively, one expects that the larger the applied pressure gradient, the faster the

chain hops, or the smaller τ becomes. Fig. 6.14 shows the Peclet-number dependence

of the mean residence time τ , nondimensionalized by the longest relaxation time τ0

of λ−DNA. Over the range of Peclet number studied, the results appear to follow

a exponential distribution τ ∼ τh exp(−αPe), where 1/τh is basically the hopping

frequency as Pe → 0. The physical mechanism of hopping becomes clear when we

write the Peclet number as Pe = (Upζc)HkBT

, where ζc is the friction coefficient of the

chain and we have written the chain diffusivity D as kBTζc

. The term Upζc is a measure

of the hydrodynamic drag force on a chain in the pit. Therefore, the product in the

numerator is the work done by the fluid to drag the molecule out of the pit, and Pe

is the ratio of that work to thermal energy. Hence, our simulation demonstrates that

at low Pe, the dynamics of the hopping is indeed thermally activated.

To examine the effect of HI on the dynamics, we also performed simulation for the

“free draining” (FD) case, where HI were ignored ( in Fig. 6.14). In FD simulation,

the beads are not coupled through hydrodynamical interactions. The mobility tensor

is reduced to the product of 1ζand the identity matrix. Comparing the results obtained

for HI and FD, we can see that the mean residence time is indeed affected by the

hydrodynamic interactions: it takes a longer time for the molecule to hop to the

Page 132: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

120

Figure 6.13: (a) Time series plot of the x-component of the center of mass of a λ−DNA driven through a nanopit array for two low values of Pe. (b) Residence timedistribution and best fit to an exponential distribution for conditions shown in (a).

Page 133: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

121

Figure 6.14: Mean resident time τ v.s. Peclet number Pe.

next pit in the FD case. This indicates that although HI is screened on a length

scale larger than the slit height, it plays an important role in dynamics happening

on smaller length scales. In this case, the “cooperative” motion of DNA segments

due to HI in the complex geometry gives rise to an increase of the hopping frequency.

We also noticed that the difference between HI and FD gets smaller as the Peclet

number increases. As Peclet number increases, the transport process is dominated by

DNA-wall steric interaction while hydrodynamic interactions play a relatively minor

role. This observation is in line with the observations of Hernandez-Ortiz et al. in

the problem of hydrodynamic effects on polymer translocation through a pore.[66]

We also studied the relationship between mean residence time and chain length. In

general, separation of a mixture of chains of different sizes can only occur if the mean

polymer velocity through the channel strongly depends on the chain length. Fig. 6.15

shows the dependence of the mean residence time on chain length for two different

Page 134: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

122

Figure 6.15: Mean residence time τ v.s. chain length N and Peclet number Pe

Peclet numbers. First, as we increase the Peclet number from 3.5 to 5.4, τ becomes

smaller. Second, at a given Pe, we observe a strong molecular weight dependence of τ

at low Pe, while τ saturates in the long chain limit. The first observation is consistent

with the results shown in Fig. 6.14 that increasing Pe lowers the energy barrier

height. The mechanism responsible for the second observation is not clear. Based on

the barrier crossing theory, we have the following hypothesis. For a given pit size,

there is a corresponding DNA size Lc, which fills the pit. For a DNA molecule longer

than Lc, the energy barrier is determined only by the chain segments with the same

length as Lc, i.e. it is determined by the part of the chain residing in the pit. Since

the transport velocity of DNA in the slit is independent of chain length [141], we can

assume the reaction rate prefactor in the barrier crossing theory is also independent

of chain length, thus, in the long chain limit, the mean residence time saturates. The

effect of HI gets more pronounced as the chain length gets longer, as shown by the

Page 135: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

123

empty square symbols in Fig. 6.15. This is consistent with the observation of Izmitli

et al.[73] and Hernandez-Ortiz et al.[66] in the problem of hydrodynamic effects on

polymer translocation through a pore.

6.4.2 Dynamics at high Peclet number

When we increase the pressure drop, the transport characteristics of DNA in the

device change, as indicated by the change of the shape of the probability density

function of residence time (Fig. 6.16(a)). In contrast with ρ(td) at low Peclet num-

ber, the distribution at high Peclet number can be fitted to a Gaussian ρ(td) =

1w√2π

exp(− (td−Tc)2

2w2 ), with mean Tc and variance w2. This shift from exponential to

Gaussian reflects the change of the dominant physics as Pe increases, from a pri-

marily stochastic process at low Peclet number to a primarily deterministic one at

high Peclet number. Qualitatively, our simulation results are consistent with exper-

imental observations [36]. We do not seek a quantitative comparison, as data about

the mean residence time were not reported in the experimental study. Fig. 6.16(b)

shows the mean residence time as function of Pe, which is also quite different from

that at low Pe. It appears to follow a power law at high Pe, Tc ∼ 1/Pe. This is

not surprising as at high Pe, as the deterministic drifting process dominates. Hence

the residence time is well-approximated by the time it takes to travel across one pit,

which is approximately Tc ≈ Lx/Up ∼ 1/Pe. The variance of the residence time,

w2, approaches a scaling of Pe−2 in the high Peclet number limit. This is consistent

with the macrotransport theory prediction of force-driven transport of a point-size

Brownian particle in a slowly varying periodic channel.[25, 84]

Page 136: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

124

Figure 6.16: (a) Residence time distribution at high Peclet number and best fit to aGaussian distribution. (b) Fitting parameters for Gaussian distribution at differentPeclet numbers.

Page 137: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

125

6.5 Conclusions

In this work, we presented an immersed boundary method (IBM) based on the gen-

eral geometry Ewald-like method (GGEM), aimed at simulating Brownian motion of

polymer in complex geometries with full description of hydrodynamic interactions.

The core idea is to replace the complex boundary with regularized point forces, cho-

sen to satisfy the appropriate boundary conditions. The IBM-GGEM methodology

correctly reproduces the velocity field in several test cases including cases involving

curved boundary and point momentum sources.

As an application of this methodology we have considered the simulation of a single

DNA molecule driven through a nanoslit with embedded nanopits array, by a pressure

gradient. We studied the dynamics of the DNA as a function of the Peclet number

and chain length, as well as the influence of hydrodynamic interactions by comparing

with free draining simulation results. We found that the transport characteristics of

the hopping dynamics in this device differ at low and high Peclet number, and for

long DNA, relative to the pit size, the dynamics is governed by the segments residing

in the pit. We also found that HI plays an important quantitative role in the hopping

dynamics even in such a highly confined system.

We expect that our algorithms will find many applications in micro- and nano-

fluidics.

Page 138: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

126

Chapter 7

Conclusion and future work

In conclusion, we have developed general numerical frameworks to study electrophoretic-

driven and flow-driven DNA dynamics in complex geometries, and we presented a

systematic investigation of three different problems using Brownian dynamics simu-

lations.

In both the problem of soft nanomechanical elements and the problem of flowing

DNA through nanopits, the concepts of entropic trapping and barrier crossing give

satisfying explanations to the observed dynamics. It seems that we have a good

understanding of the basic physics in those microfluidic devices, where DNA dynamics

is determined by the combined effects of major driving forces and chain configuration

entropy. Now the goal is shifted to apply the known physics to the design of devices.

This echoes the motivation of this work, which is to build validated numerical tools

to help the design and optimization of novel processes and devices in a very broad

area.

Considering the future of both the simulation methods and their applications

Page 139: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

127

to the problem of DNA dynamics in complex geometries, harnessing the increasing

computational power to conduct more detailed simulations of larger system for longer

time is an inevitable trend.

Future work may include:

• New mechanical model for DNA under nanometer scale confinement. When

a DNA chain is confined to a space smaller than its persistence length, which

is already achieved in experiments, the bead-spring model breaks down. We

thus need to build a finer-scale model of the DNA to study these systems, for

example, a bead-rod model with a rod length of 10 nm. For a λ-DNA molecule,

this means thousands of beads, which is out of the reach of traditional Brownian

dynamics simulation technique when long-ranged hydrodynamic interactions are

included. But it might be feasible with our new accelerated immersed boundary

method using graphic processing unit.

• Graphic processing unit (GPU) acceleration. The parallel architecture of GPU

makes it very effective for algorithms where processing of large blocks of data is

done in parallel. It has been applied to accelerate molecular dynamics simula-

tion and computational fluid dynamics calculation. As recently reported, GPU

calculations achieved tens of fold, even hundreds of fold in some case, speedup

over an optimized CPU implementation. The advantages of our accelerated

Brownian dynamics algorithm is that it divides the long-ranged interactions

into short-ranged pair-wise interactions, which can be calculated using analyti-

cal expressions, and long-ranged interactions, which change smoothly with small

gradient and can be calculated on a regular structured Cartesian grid. Both

calculations are straightforward to parallelize efficiently. Also, in our algorithm,

Page 140: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

128

the boundary nodes are treated in a similar fashion as the polymer beads, which

gives its potential for parallel processing.

Page 141: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

129

Appendix A

Lattice random walk model of a

tethered polymer

In this appendix, we use a random walk lattice model to calculate the thermodynamic

properties of an ideal polymer molecule tethered to a hard nonadsorbing wall.

For an ideal chain, the entropy S(r) associated with all chain conformations start-

ing from an origin and ending at r is simply related to the number of distinct walks

Z(r) connecting the origin and r in n steps.[35]

S(r) = kB ln[Zn(r)] (A.1)

Therefore, the free energy of an ideal chain is F(r)/kBT = − ln[Zn(r)]. It is useful to

know how does the free energy change as a function of a single reaction coordinate,

such as the perpendicular distance of the free end to the wall ze for a tethered molecule.

It could reduce the problem to one dimensional and allow us to apply Kramer’s theory

of barrier crossing to study the dynamic properties of the walk. Here we introduce

Page 142: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

130

the potential of mean force of a polymer chain as a function of ze

F(ze) = F∗(ze)− kBT ln[〈ρ(ze)〉〈ρ∗(ze)〉

] (A.2)

where F∗(ze) and 〈ρ∗(ze)〉 are arbitrary reference values.[121] The average distribution

function 〈ρ(ze)〉 along the reaction coordinate is obtained from a Boltzmann weighted

average

〈ρ(ze)〉 =∫

drδ(z′

e(r)− ze)e−U(r)/kBT

dre−U(r)/kBT, (A.3)

where U(R) represents the total energy of the chain as a function of the beads’

positions R and δ(z′

e(R)− ze) picks out the configurations with the same ze. For an

ideal chain, every path has the same weight, and above expression reduces to

〈ρ(ze)〉 =Zn(ze)

Ztotn

, (A.4)

and therefor, the potential of mean force is

F(ze) = F∗(ze)− kBT ln[Zn(ze)

Z∗n(ze)

]. (A.5)

Above expressions indicate one can calculate the free energy profile along the chosen

reaction coordinate (ze), by calculate the ratio of the number of configurations of a

specific state to that of the reference state.

The property of an ideal chain is often studied using a lattice random walk model

[35]. For a n-step random walk on an integer lattice, we can directly count the

number of walks starting origin and ending at (x, y, z). When the walk is near a wall,

absorbing boundary condition should be used, i.e., we should discard those walks

Page 143: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

131

touching the wall at least once.[40, 134] Once the number of walks satisfying both

absorbing boundary and various constraints are obtained, the free energy is readily

obtainable using thermodynamic relationships presented above. Next, we present the

results of combinatorics calculations for counting the chain configurations.

According to Bertrand’s “ballot theorem”, the number of 1-D random walks on

the integers of n steps, from the origin to the point x > 0 and never return to the

origin, is

Z1dt (x,N) =

x

N

(

NN+x2

)

, (A.6)

assuming N and x have the same parity mod(N − x, 2) = 0. Thus, the total number

of walks is

Z1dt =

x

W 1dt (x,N) =

x

x

N

(

NN+x2

)

. (A.7)

Assuming N = 2k + 1(k = 0, 1, ...),

Z1dt =

x=1

x

N

(

NN+x2

)

=1

N

x=1

x

(

NN+x2

)

=1

N[

(

N

k + 1

)

+ 3

(

N

k + 2

)

+ ...+ (2k − 1)

(

N

k + k

)

+ (2k + 1)

(

N

k + k + 1

)

].

(A.8)

Note that (2i − 1) = (k + i) − (k − i − 1),(

Nk+i

)

=(

Nk+1−i

)

and (k + i)(

Nk+i

)

=

Page 144: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

132

(N − k + 1)(

NN−k+1

)

. Last expression is

Z1dt =

1

N[

(

N

k + 1

)

+ 3

(

N

k + 2

)

+ ...+ (2k − 1)

(

N

k + k

)

+ (2k + 1)

(

N

k + k + 1

)

].

=1

N[(k + 1)

(

N

k + 1

)

− k

(

N

k

)

+ ...+ (2k)

(

N

2k

)

−(

N

1

)

+N ]

=1

N(k + 1)

(

N

k + 1

)

=

(

2k

k

)

. (A.9)

Similarly, we can prove that when N = 2k(k = 1, 2...), Z1dt =

(

2k−1k

)

. Therefore,

Z1dt =

(

N−1[N/2]

)

, where [x] is the largest integer smaller than x.

In 2-D case, assuming there are i steps in the wall normal direction and N − i

steps in the wall parallel direction, the total number of paths are

Z2dt =

N∑

i=1

(

N − 1

i− 1

)(

i− 1

[ i2]

)

2N−i =

(

2N − 1

N − 1

)

. (A.10)

Similarly, in 3-D, the total number of paths is

Z3dt =

N∑

i=1

(

N − 1

i− 1

)(

i− 1

[ i2]

)

4N−i. (A.11)

In summary, we obtain the analytical expressions of the number of configurations of

Page 145: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

133

1

2N steps

x z

y

2

0

Figure A.1: A tethered polymer as lattice random walk.

a tethered polymer molecule, and the total numbers of paths are

Z1dt (N) =

N∑

x=1

x

N

(

NN+x2

)

=

(

N − 1

[N2]

)

≈ 2N√2πN

, (A.12)

Z2dt (N) =

N∑

m=1

(

N − 1

m− 1

)(

m− 1

[m2]

)

2N−m =

(

2N − 1

N − 1

)

≈ 4N√4πN

, (A.13)

Z3dt (N) =

N∑

m=1

(

N − 1

m− 1

)(

m− 1

[m2]

)

4N−m ≈ 6N√6πN

, (A.14)

where m is the number of steps in the confined(wall normal) direction. There is

no closed expression for the sum in 3-D. However, the approximation 6N/√6πN

converges to the exact result as N increases (data not shown), following the same

scaling as those for 1-D and 2-D cases. In this work, the total number of N -step tail

is called the fundamental solution of a tethered chain, and it is denote ZN0 .

When a polymer is grafted to a bilayer, it can induce gelation or other phase

Page 146: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

134

changes in lamellar phases, which might be related to the interactions of locally

deformed membrane patches which are created by the entropic forces exerted by the

grafted polymer molecules[20]. When an ideal polymer chain is grafted to a hard wall,

the number of accessible configuration is greatly reduced which leads to an entropy

penalty. Phenomenologically, we observe a repulsive force which drives the polymer

away from the wall, i.e., we need to apply a tethering force to fix the end of polymer.

We are interested in the dependence of this tethering force on polymer chain length.

For a lattice polymer, the required force f0 is related to the number of configurations

Z as

f0 = −∂F(l)

∂l|l=0 ≈

F0 − F1

1a=

kBT

a(lnZN+1

0 − lnZN0 ), (A.15)

where l is the perpendicular distance of the tethered end from the wall, the lattice

grid size is a = 1 which has a physical meaning of persistence length of the polymer

molecule. Using the expression of the number of tail configuration (Eq. A.14), we

know that ZN0 ≈ 6N√

6πNand ZN+1

0 ≈ 6N+1√6π(N+1)

. Plug these into last equation and we

obtain

f0 =kBT

aln

ZN+10

ZN0

≈ kBT

aln 6

N

N + 1=

kBT

a(ln 6− 1

2ln

N + 1

N)

≈ kBT

a(ln 6− 1

2N). (A.16)

Clearly, the force is repulsive, and the longer the chain, the larger the force.

Interestingly, the force becomes independent of the chain length in the long chain

limit. The expression we obtained is similar to that of Bickel et al. [20] using the

Page 147: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

135

0 1 2 310

−5

10−4

10−3

10−2

10−1

100

101

l/Rg

∆F(l/

Rg)/

k BT

N=6

N=8

N=10

N=12

Figure A.2: The wall-polymer potential calculated using the fundamental solution ofa tethered chain for various chain length (symbols). Dashed line is a fit to Eq. A.23

diffusion equation[20], f0 = kBTa

exp(− a2

4R2g) ≈ kBT

a(1 − a2

4R2g), and it has the same

dependence on persistence length a and on chain length n. The difference of the

constant ln 6 is due to the nature of the lattice. To our best knowledge, there is no

experimental results for us to compare with. For a semi-flexible fluorescently stained

λ-DNA with a = 50nm and Rg = 770 nm, the tethering force is about 0.82 pN. For a

flexible molecule of poly(methacrylic acid) (PMAA) with a=0.3nm and Rg = 700nm,

however, the tethering force is 11.2 pN.

Above analysis can be extended to calculate the effective repulsive potential be-

tween a free polymer molecule and a hard wall using the fundamental solution of a

tethered molecule. The total number of N -step paths in the half space starting l away

Page 148: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

136

from the wall is denoted WNl , and it can be obtained recursively as,

WN0 = ZN

0 , (A.17)

WN1 = ZN+1

0 = 4WN−11 +WN−1

2 , (A.18)

WN2 = ZN+12

0 − 4ZN+10 = 4WN−1

2 +WN−11 +WN−1

3 (A.19)

WN3 = ZN+2

0 − 8ZN+10 + 15ZN+1

0 = 4WN−13 +WN−1

2 +WN−14 (A.20)

WNi+1 = WN+1

i − 4WNi +WN

i−1. (A.21)

We can see that WNl is a linear combination of the fundamental solutions ZN+i

0 , i =

1, 2, ..., l − 1, and the coefficients can be constructed recursively. Once all WNl are

calculated, taking the free chain as the reference state, the wall-polymer potential

can be calculated according to Eq. A.5 as

∆F(l, N)

kBT= − ln

WNl

6N. (A.22)

Results for various chain lengths are plotted in Fig. A.2. First, when l is normal-

ized by the size or the molecule Rg =√Na, where a is the lattice size, all results

collapse onto a master curve. Second, the interaction potential decreases exponen-

tially fast when l < Rg and it can be fitted to the following equation

∆F id(l, N)

kBT= − ln(erf(

l

Rg)), (A.23)

which is obtained from a continuous description of the random walk using the diffu-

sion equation[24]. This repulsive potential can be used in mesoscale simulation of a

polymer molecule near a single wall.

Page 149: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

137

Appendix B

BD/FEM algorithm for simulating

DNA electrophoresis

The objective of this appendix is to present a numerical framework, Brownian dynam-

ics (BD)/Finite element method (FEM), to simulate DNA electrophoresis in complex

geometries. The proposed numerical scheme is composed of three parts:

• a bead-spring DNA model,

• a finite element scheme for the non-homogeneous electric field,

• and a coupling algorithm to find the values of the electric field at the DNA

beads location.

The bead-spring model has been described in Chapter 3. In the reminder of this

appendix, we discuss the finite element scheme and the coupling algorithm.

For a linear DNA molecule, the Debye length (κ−1), which is the length scale

over the charges on the DNA backbone are screened out by the counter-ions, is much

Page 150: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

138

smaller than the persistence length of DNA in sufficiently concentrated salt solution

(typically κ−1 is approximately 2 nm in thickness under physiological conditions).

Thus, DNA globally behaves like a neutral polymer and the electric field is governed

by Laplaces equation, since local electric field disturbance due to a DNA molecule is

screened over the Debye length. Also, we consider the system where electroosmotic

flow is eliminated using a polymer layer grafted on the walls. In terms of geometry, we

consider a microfluidic device with a constant channel height, and the channel height

is usually much smaller than the other length scales of the device. Therefore, the

problem is reduced to two dimensional. Still, with increasing complexity of geometry,

a numerical method is required to calculate the electric field in a microfluidics with

complex geometries.

The electric field is computed from the electric potential with the relationship

E = −∇φ in an arbitrary domain, where φ denotes electric potential. The governing

equations for the electrostatic potential are,

∇φ2 = 0 in Ω, (B.1)

φ = φ0 or∂φ

∂n= 0 on ∂Ω, (B.2)

where n is the normal vector at the boundary. We consider two types of boundary

conditions. One is Dirichlet boundary condition which specifies the values of poten-

tial on the boundary of the domain, such as the electric potentials at the inlet and

outlet. The other is Neumann boundary condition which specifies the values that

the derivative of potential is to take at the insulating channel walls. In experiments,

the device is usually made of poly(dimethylsiloxane) (PDMS) which is an insulator.

Page 151: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

139

Hence, a zero gradient boundary condition is assigned along the channel walls.

In this work, we use a commercial partial differential equation (PDE) solver,

COMSOL, which is based on the finite element method (FEM), to solve the governing

Laplace’s equation for the electrostatic potential (Eq. B.2). For detailed information

on FEM, the readers are referred to [17] and COMSOL manual. In a microfluidic

device, steep field gradient exists around obstacles or sharp corners, which means that

the space around an immersed structure like a post, and around a sharp corner like

near a contraction/expansion region, should be more finely discretized. Once electric

potential is obtained, E is simply computed from E = ∇φ and stored for lookup. To

calculate the field strength at the bead location, we need a fast algorithm to find the

triangular element surrounding the bead.

For an irregular triangular mesh, finding the electric field at a specified point in

the domain is not a trivial problem. As we just discussed above, to resolve the steep

field gradient around the obstacles and corners, very fine mesh needs to be used. This

means that we have a large number of elements to look through. We address this

task as follows. We first assume that the target point is closer to the centroid of the

element enclosing it, than to the centroids of all the other elements. Of course this is

not always true as shown in Fig. B.1. So we need to find the four nearest centroids

to the target, and then find the one enclosing it using a confirming algorithm. Now

the problem is reduced to find the 4-nearest centroids to the target. We can solve

it using the nearest neighbor search algorithm, which is a well-known algorithm in

computational geometry [133]. In this work, we use a variant of this algorithm called

Approximate Nearest Neighbor Searching (ANN) developed by David M. Mount and

Sunil Arya [11]. Finding the nearest point needs O(logN) operation in the case of

Page 152: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

140

Figure B.1: Schematic of the nearest neighbor search algorithm (NNS). To find theelement enclosing the target point, we use NNS to find the k-nearest centroids of theelements.

randomly distributed points. The worst case search time, when a 2-dimensional KD-

tree containing N nodes is used, is O(N1/2). This allows us to use very fine mesh in

the calculation.

Page 153: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

141

Appendix C

Dimer in shear flow

In this appendix we analytically and numerically consider the case of a Brownian

particle attached to the origin with a harmonic spring with stiffness k = κζ , where ζ

is the friction coefficient, subject to shear flow, vx = γy. If additionally we include a

hard wall at y = 0 such that the random walk is restricted to the upper half-space,

y > 0, we are essentially considering a tethered polymer chain composed of two beads.

Note that this problem is essentially two-dimensional.

The overdamped Langevin equations for the particle coordinates are

x = γy − κx+ Fx

y = −κy + Fy,

where F denotes the random forcing. After performing a Fourier transform in time,

Page 154: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

142

we get the solution in Fourier space

x =(iν + κ)Fx + γFy

(iν + κ)2

y =Fy

(iν + κ),

from which we can obtain all cross-correlation functions using the identities⟨

Fx⋆Fx

=⟨

Fy⋆Fy

= α and⟨

Fx⋆Fy

= 0. In particular, we obtain the monotonically-

decreasing non-normalized PSD

Sxy(ν) =∥

∥Cxy

∥= ‖〈x⋆y〉‖ =

αγ

(ν2 + κ2)3/2,

and, after an inverse Fourier transform, the non-normalized CCF

Cxy(t) =

αγe−κt(2κt + 1)/(4κ2) for t ≥ 0

αγeκt/(4κ2) for t < 0.

In the case of no shear flow, γ = 0, we obtain that Cxx(t) = αe−κ|t|/(2κ), showing that

the relaxation time is τ = κ−1 and thus Wi = γ/κ. For the harmonic spring dimer

the relaxation time does not depend on Wi. The cross-correlation function shows

a single peak at tmax = (2κ)−1 = τ/2, and after proper normalization, Cxy(t) =

Cxy(t)/√

Cxx(t = 0)Cyy(t = 0), the height of the peak in the CCF is found to be

Cmaxxy = Cxy(τ/2) =

√2e−1/2Wi

2 +Wi2. (C.1)

This analytically-solvable dimer model, even without a hard wall, reproduces the

Page 155: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

143

characteristics of the CCF that we observe for tethered polymer chains in shear flow.

Specifically, Cxy(t) has an asymmetric peak of width ∼ τ centered at t = τ/2 and

height ∼ Wi. There is no periodicity in the motion of the dimer and no “cycling”

time-scale other than the intrinsic relaxation time τ .

The dimer problem can no longer be solved analytically if a hard wall is present

or if the spring is non-linear (e.g., FENE or worm-like). We can, however, study the

dimer with a non-linear spring and/or in the presence of a hard wall numerically using

Brownian Dynamics (without hydrodynamics). Some results for Wi = 2 are given in

Fig. C.1, where we also show the analytical solution for the harmonic dimer and the

results for longer tethered chains. When a hard wall is present, the numerical results

show that the position of the peak in the CCF shifts to smaller times and reduces in

height. For the non-linear springs, the position of the peak moves to smaller times as

Wi increases, exactly as we observe for the tethered chains. The height of the peak

is several times larger for a dimer than for a chain with N ≫ 1 beads, which is not

unexpected.

Even after including non-linearity and the hard wall, the dimer model fails to

reproduce the smaller but still substantial negative peak at t < 0 that we observe in

the CCFs for the longer tethered chains at small Wi. An analytical calculation for a

harmonic chain tethered to a point and subjected to shear flow might reproduce that

feature as well. We can mimic such a peak by constructing an artificial CCF,

Cxy(t) = Cxy(t)− αCxy(−t), (C.2)

where 0 < α < 1 controls the depth of the negative peak, and Cxy is the analytical

Page 156: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

144

CCF for the harmonic dimer . As illustrated in Fig. C.1, such an empirical fit

matches the numerical results quite well. The Fourier transform of Eq. (C.2) gives

an empirical PSD of the form

S(2πν = Ω/τ) ∼ Wi

(1 + α2)(1 + Ω2)− 2α(1− Ω2)

1 + 2Ω2 + Ω4,

which for α > 1/3 exhibits a wide maximum at frequencies Ω = 2πτ/T ∼ 0.5, i.e.,

at a period T ∼ 10τ . As illustrated in Fig. C.1, the maximum in this PSD is very

reminiscent of the “peaks” in the PSD observed in Refs. [127, 37, 38], where they

were attributed to the existence of a periodic motion with period of about 10τ . The

analytical shape of the PSD only involves τ as a relevant timescale, and the cross-

correlation function has an exponential decay at large times ∼ exp(−t/τ), just like

the autocorrelation function for the end-to-end vector used to define relaxation times.

Such an exponential decay is inconsistent with periodic motion, but is consistent with

some recent theoretical models that suggest similar correlations for a free chain in

shear flow [34, 156]. In summary, as seen from this simple analytical example of a

dimer in a flow, a maximum in the PSD does not imply any periodic motion and the

claim of an existence of a new physical timescale other than the internal relaxation

time of the polymer is not justified.

Page 157: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

145

0.01 0.1 1

Dimensionless frequency f=ντ

0.01

0.1

1

PS

D

α=1/2α=2/3α=3/4

-2 -1 0 1 2 3

t / τ

-0.2

-0.1

0

0.1

0.2

0.3

0.4

Cxy

Rescaled theory Cxy

(2τ)/2

Harmonic dumbbellFENE dumbbellWormlike dumbbellRescaled chain 3C

xz(τ) (BD, N=20)

Empirical !t ( α=2/3)

PS

D

Cxy

0.3

0.2

0.4

0.1

0.0

-0.1

-0.2-2 -1 0 1 2 3

t/t

1

0.1

0.01

0.01 0.1 1

Dimensionless frequency f=nt

Figure C.1: The left panel shows the cross-correlation function for a dimer (dumbbell)tethered to a hard wall and subjected to shear flow, for a harmonic, FENE anda worm-like spring. We also show a rescaled form of the analytical solution for aharmonic dimer in shear flow (without a hard wall). The height of the peak diminishesby a factor of about 2 when a hard wall is present, so we have rescaled the analyticalsolution for the harmonic dumbbell accordingly. The position of the peak shiftsto smaller times when a hard wall is present as well, and we have thus rescaledthe time for the analytical solution. The CCF for a wormlike chain of N = 20beads, as obtained from Brownian Dynamics simulations, is also shown for qualitativecomparison after scaling by a factor of 3 to bring its height in agreement with thedimer case. We also show an empirical fit to the Brownian Dynamics simulations ofthe form proposed in Eq. (C.2), for which the PSD can be analytically calculatedand shows a maximum at period T ∼ 10τ , depending on the value of the tunableparameter α, as illustrated in the right panel.

Page 158: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

146

Appendix D

Fixman’s midpoint algorithm

Fixman[52] proposed a method that avoids computing the derivative of diffusion

tensor, which is generally unknown for complex geometries. As a simple illustration

of how this method works, we start with the one dimensional case with only the drift

caused by the position-dependence of D, and a forward Euler step is simply

∆x = x(t+∆t)− x(t) = (d

dxD(x))∆t +

√2B(x(t))∆w. (D.1)

We can approximate the derivative using a finite difference, in which case this equation

becomes

∆x =D(x(t) + ∆x)−D(x(t))

∆x∆t +

√2B(x(t))∆w. (D.2)

This is an implicit method as information at t +∆t is needed. Fixman proposed an

intermediate step defined as

∆x∗ = x∗(t +∆t)− x(t) =√2B(x(t))∆w. (D.3)

Page 159: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

147

This is a valid approximation, as in Eq. D.2 the first term is proportional to ∆t and

the second term is proportional to ∆w ∝ ∆t1/2. When ∆t → 0, the first term is van-

ishingly small compared to the second term. Then the finite difference approximation

of the derivative term becomes

D(x(t) + ∆x)−D(x(t))

∆x≈ D(x∗)−D(x(t))

∆x∗ =D(x∗)−D(x(t))√

2B(x(t))∆w. (D.4)

Inserting this into Eq. D.2 and notice that ∆t = (∆w)2 when ∆t → 0, we finally

have

∆x =D(x∗)−D(x(t))√

2B(x(t))∆w +

√2B(x(t))∆w. (D.5)

The vector version of this, with D = B ·B, is given by

∆r =√2/2(D(r∗)−D(r(t)))(B−1(r(t)))T ·∆w +

√2B(r(t)) ·∆w (D.6)

=√2/2D(r∗)(B−1(r(t)))T ·∆w +

√2/2B(r(t)) ·∆w. (D.7)

We now show that the first term on the RHS of Eq. D.6 equals (∂/∂r ·D)dt when

∆t → 0. The ith component of the first term is√2/2(Dij(r

∗)−Dij(r))(B−1)Tjk∆wk.

Expand Dij around r = r(t)

Dij(r∗) = Dij(r) +

∂Dij

∂rl∆r∗l +O(∆r∗2l )). (D.8)

Plug in Eq. D.3 and ignore the higher order terms:

Dij(r∗)−Dij(r) =

∂Dij

∂rl

√2Blm∆wm. (D.9)

Page 160: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

148

Plug this into

√2/2(Dij(r

∗)−Dij(r))(B−1)Tjk∆wk =

√2/2

∂Dij

∂rl

√2Blm∆wm(B

−1)Tjk∆wk

=∂Dij

∂rlBlm(B

−1)Tjk∆tδmk

=∂Dij

∂rlBlkB

−1kj ∆t

=∂Dij

∂rl∆tδlj

=∂Dij

∂rj∆t. (D.10)

In the second step, we use the property of Wiener process dwidwj = δijdt.

Page 161: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

149

Bibliography

[1] M. D. Graham. Fluid Dynamics of Dissolved Polymer Molecules in Confined

Geometries. Annu. Rev. Fluid Mech., 43:273, 2011.

[2] J. J. de Pablo. Coarse-grained simulations of macromolecules: From DNA to

nanocomposites. Annu. Rev. Phys. Chem., 62:555, 2011.

[3] R. Adhikari, K. Stratford, M. E. Cates, and A. J. Wagner. Fluctuating Lattice

Boltzmann. Europhys. Lett., 71:473, 2005.

[4] U. S. Agarwal, A. Dutta, and R. A. Mashelkar. Effect of flexibility on the

shear-induced migration of short-chain polymers in parabolic channel flow. J.

Fluid Mech., 557:297, 2006.

[5] P. Ahlrichs and B. Duenweg. Simulation of a single polymer chain in solution

by combining lattice Boltzmann and molecular dynamics. J. Chem. Phys.,

111:8225, 1999.

[6] B. J. Alder and T. E. Wainwright. Studies in molecular dynamics. I. General

method. J. Chem. Phys., 31:459, 1959.

Page 162: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

150

[7] F. J. Alexander and A. L. Garcia. The Direct Simulation Monte Carlo Method.

Comput. Phys., 11:588, 1997.

[8] I. Ali and J. M. Yeomans. Polymer translocation: The effect of backflow. J.

Chem. Phys., 123:234903, 2005.

[9] A. Alvarez and R. Soto. Dynamics of a suspension confined in a thin cell. Phys.

Fluids, 17:093103, 2005.

[10] P. E Arratia, C. C Thomas, J Diorio, and J. P Gollub. Elastic instabilities of

polymer solutions in cross-channel flow. Phys. Rev. Lett., 96:4, 2006.

[11] S. Arya and D. M. Mount. Approximate nearest neighbor queries in fixed dimen-

sions. In Proceedings of the fourth annual ACM-SIAM Symposium on Discrete

algorithms, SODA ’93, pages 271, Philadelphia, PA, USA, 1993. Society for

Industrial and Applied Mathematics.

[12] P. J. Atzberger, P. R. Kramer, and C. S. Peskin. A stochastic immersed bound-

ary method for fluid-structure dynamics at microscopic length scales. J. Com-

put. Phys, 224:1255, 2007.

[13] C. Aust, M. Kroger, and S. Hess. Structure and dynamics of dilute polymer solu-

tions under shear flow via nonequilibrium molecular dynamics. Macromolecules,

32:5660, 1999.

[14] A. Balducci, P. Mao, J. Han, and P. S. Doyle. Double-stranded DNA diffusion

in slitlike nanochannels. Macromolecules, 39:6273, 2006.

Page 163: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

151

[15] J. R. Banavar, M. Cieplak, T. X. Hoang, and A. Maritan. First-principles design

of nanomachines. Proc. Natl. Acad. Sci. U. S. A., 106:6900, 2009.

[16] A. J. Banchio and J. F. Brady. Accelerated Stokesian dynamics: Brownian

motion. J. Chem. Phys, 118:10323, 2003.

[17] K. J. Bathe. Finite Element Procedures. Prentice-Hall, Englewood Cliffs, NJ,

1996.

[18] V. A. Beck and E. S. G. Shaqfeh. Ergodicity breaking and conformational

hysteresis in the dynamics of a polymer tethered at a surface stagnation point.

J. Chem. Phys., 124:094902, 2006.

[19] J. B. Bell, A. Garcia, and S. A. Williams. Numerical Methods for the Stochastic

Landau-Lifshitz Navier-Stokes Equations. Phys. Rev. E, 76:016708, 2007.

[20] T. Bickel, C. Jeppesen, and C. M. Marques. Local entropic effects of polymers

grafted to soft interfaces. Eur. Phys. J. E, 4:33, 2001.

[21] R. B. Bird, C. F. Curtiss, R. C. Amstrong, and O. Hasager. Dynamics of

Polymer Liquids Vol. 2. Wiley-InterScience; 2 edition, 2001.

[22] J. R. Blake. Image system for a Stokeslet in a no-slip boundary. Proc. Camb.

Philos. S-M, 70:303, 1971.

[23] B. M. Boghosian, P. J. Love, P. V. Coveney, I. V. Karlin, S. Succi, and J. Yepez.

Galilean-invariant lattice-Boltzmann models with H theorem. Phys. Rev. E,

68:025103, 2003.

Page 164: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

152

[24] P. G. Bolhuis, A. A. Louis, J. P. Hansen, and E. J. Meijer. Accurate effective

pair potentials for polymer solutions. J. Chem. Phys., 114:4296, 2001.

[25] H. Brenner and D. A. Edwards. Macrotransport Processes. Butterworth-

Heinemann, 1993.

[26] C. Bustamante, J. F. Marko, E. D. Siggia, and S. Smith. Entropic elasticity of

lambda-phage DNA. Science, 265:1599, 1994.

[27] C. G. Canuto, M.Y. Hussaini, A. Quarteroni, and T.A. Zang. Spectral Methods:

Fundamentals in Single Domains. Springer-Verlag, 2006.

[28] C. Chang, C. Liu, and C. Lin. Boundary conditions for lattice Boltzmann

simulations with complex geometry flows. Comput. Math. Appl., 58:940, 2009.

[29] R. Chelakkot, R. G. Winkler, and G. Gompper. Migration of semiflexible poly-

mers in microcapillary flow. EPL, 91:14001, 2010.

[30] Y. L. Chen, M. D. Graham, J. J. de Pablo, K. Jo, and D. C. Schwartz. DNA

molecules in microfluidic oscillatory flow. Macromolecules, 38:6680, 2005.

[31] Y. L. Chen, M. D. Graham, J. J. de Pablo, G. C. Randall, M. Gupta, and P. S.

Doyle. Conformation and dynamics of single DNA molecules in parallel-plate

slit microchannels. Phys. Rev. E, 70:060901(R), 2004.

[32] Y. L. Chen, H. Ma, M. D. Graham, and J. J. de Pablo. Modeling DNA in con-

finement: A comparison between the brownian dynamics and lattice Boltzmann

method. Macromolecules, 40:5978, 2007.

Page 165: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

153

[33] B. Chun and A. J. C. Ladd. Interpolated boundary condition for lattice Boltz-

mann simulations of flows in narrow gaps. Phys. Rev. E, 75:066705, 2007.

[34] D. Das and S. Sabhapandit. Accurate Statistics of a Flexible Polymer Chain in

Shear Flow. Phys. Rev. Lett., 101:188301, 2008.

[35] P. G. de Gennes. Scaling Concepts in Polymer Physics. Cornell University

Press, 1979.

[36] J. T. Del Bonis-O’Donnell, W. Reisner, and D. Stein. Pressure-driven DNA

transport across an artificial nanotopography. New J. Phys., 11:075032, 2009.

[37] R. Delgado-Buscalioni. Cyclic motion of a grafted polymer under shear flow.

Phys. Rev. Lett, 96:088303, 2006.

[38] R. Delgado-Buscalioni. Dynamics of a Single Tethered Polymer under Shear

Flow. AIP Conference Proceedings, 913:114, 2007.

[39] R. Delgado-Buscalioni and G. De Fabritiis. Embedding molecular dynamics

within fluctuating hydrodynamics in multiscale simulations of liquids. Phys.

Rev. E, 76:036709, 2007.

[40] E. A. Dimarzio. Proper accounting of conformations of a polymer near a surface.

J. Chem. Phys., 42:2101, 1965.

[41] M. Doi and S. F. Edwards. The Theory of Polymer Dynamics. Oxford Press,

New York, 1986.

[42] A. Donev, A. L. Garcia, and B. J. Alder. Stochastic Event-Driven Molecular

Dynamics. J. Comp. Phys., 227:2644, 2008.

Page 166: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

154

[43] A. Donev, A. L. Garcia, and B. J. Alder. Stochastic Hard-Sphere Dynamics for

Hydrodynamics of Non-Ideal Fluids. Phys. Rev. Lett, 101:075902, 2008.

[44] A. Donev, S. Torquato, and F. H. Stillinger. Neighbor List Collision-Driven

Molecular Dynamics Simulation for Nonspherical Particles: I. Algorithmic De-

tails II. Applications to Ellipses and Ellipsoids. J. Comp. Phys., 202:737–764,

765–793, 2005.

[45] K. D. Dorfman. DNA electrophoresis in microfabricated devices. Rev. Mod.

Phys., 82:2903, 2010.

[46] P. S. Doyle, B. Ladoux, and J. L. Viovy. Dynamics of a tethered polymer in

shear flow. Phys. Rev. Lett., 84:4769, 2000.

[47] B. Duenweg and A. J. C. Ladd. Lattice Boltzmann simulations of soft matter

systems. Adv. Polym. Sci., 221:89, 2009.

[48] E. R. Dufresne, T. M. Squires, M. P. Brenner, and D. G. Grier. Hydrodynamic

coupling of two brownian spheres to a planar surface. Phys. Rev. Lett., 85:3317,

2000.

[49] D. L. Ermak and J. A. McCammon. Brownian dynamics with hydrodynamic

interactions. J. Chem. Phys., 69:1352, 1978.

[50] P. Espanol and P. Warren. Statistical mechanics of dissipative particle dynam-

ics. EPL, 30:191, 1995.

[51] G. De Fabritiis, M. Serrano, R. Delgado-Buscalioni, and P. V. Coveney. Fluc-

Page 167: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

155

tuating hydrodynamic modeling of fluids at the nanoscale. Phys. Rev. E,

75:026307, 2007.

[52] M. Fixman. Simulation of polymer dynamics 1. general theory. J. Chem. Phys,

69:1527, 1978.

[53] M. Fixman. Construction of Langevin forces in the simulation of hydrodynamic

interaction. Macromolecules, 19:1204, 1986.

[54] M. Fyta, S. Melchionna, S. Succi, and E. Kaxiras. Hydrodynamic correlations

in the translocation of a biopolymer through a nanopore: Theory and multiscale

simulations. Phys. Rev. E, 78:036704, 2008.

[55] G. Fytas, S. H. Anastasiadis, R. Seghrouchni, D. Vlassopoulos, J. Li, B. J.

Factor, W. Theobald, and C. Toprakcioglu. Probing collective motions of ter-

minally anchored polymers. Science, 274:2041, 1996.

[56] M. Gauthier and G. Slater. Molecular dynamics simulation of a polymer chain

translocating through a nanoscopic pore. Eur. Phys. J. E, 25:17, 2008.

[57] S. Gerashchenko and V. Steinberg. Statistics of tumbling of a single polymer

molecule in shear flow. Phys. Rev. Lett., 96:038304, 2006.

[58] G. Giupponi, G. De Fabritiis, and P. V. Coveney. Hybrid method coupling fluc-

tuating hydrodynamics and molecular dynamics for the simulation of macro-

molecules. J. Chem. Phys., 126:154903, 2007.

[59] G. Gompper, T. Ihle, D. Kroll, and R. Winkler. Multi-particle collision dynam-

Page 168: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

156

ics: A particle-based mesoscale simulation approach to the hydrodynamics of

complex fluids. Adv. Polym. Sci., 221:1, 2009.

[60] I. O. Gotze, H. Noguchi, and G. Gompper. Mesoscale simulations of hydrody-

namic squirmer interactions. Phys. Rev. E, 76:046705, 2007.

[61] P. S. Grassia, E. J. Hinch, and L. C. Nitsche. Computer simulations of Brownian

motion of complex systems. J. Fluid Mech., 282:373, 1995.

[62] Y. Gratton and G. W. Slater. Molecular dynamics study of tethered polymers

in shear flow. Eur. Phys. J. E, 17:455, 2005.

[63] A. Groisman, M. Enzelberger, and S. R. Quake. Microfluidic memory and

control devices. Science, 300:955, 2003.

[64] R. D. Groot and P. B. Warren. Dissipative particle dynamics: Bridging the gap

between atomistic and mesoscopic simulation. J. Chem. Phys., 107:4423, 1997.

[65] H. Hasimoto. On the periodic fundamental solutions of the Stokes equations

and their application to viscous flow past array of spheres. J. Fluid Mech.,

5:317, 1959.

[66] J. P. Hernandez-Ortiz, M. Chopra, S. Geier, and J. J. de Pablo. Hydrodynamic

effects on the translocation rate of a polymer through a pore. J. Chem. Phys,

131:044904, 2009.

[67] J. P. Hernandez-Ortiz, J. J. de Pablo, and M. D. Graham. Fast computa-

tion of many-particle hydrodynamic and electrostatic interactions in a confined

geometry. Phys. Rev. Lett., 98:140602, 2007.

Page 169: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

157

[68] J. P. Hernandez-Ortiz, H. Ma, J. J. de Pablo, and M. D. Graham. Cross-stream-

line migration in confined flowing polymer solutions: Theory and simulation.

Phys. Fluids, 18:123101, 2006.

[69] J. P. Hernandez-Ortiz, H. Ma, J. J. de Pablo, and M. D. Graham. Concen-

tration distribution during flow of confined flowing polymer solutions at finite

concentration: slit and grooved channel. Korea-Aust. Rheol. J., 20:143, 2008.

[70] A. Hirsa, C. A. Lopez, M. A. Laytin, M. J. Vogel, and P. H. Steen. Low-

dissipation capillary switches at small scales. Appl. Phys. Lett., 86(1):014106,

2005.

[71] H. P. Hsu, K. Binder, L. I. Klushin, and A. M Skvortsov. What is the order

of the two-dimensional polymer escape transition? Phys. Rev. E., 76:021108,

2007.

[72] Y. Inoue, Y. Chen, and H. Ohashi. A mesoscopic simulation model for immis-

cible multiphase fluids. J. Comput. Phys., 201:191, 2004.

[73] A. Izmitli, D. C. Schwartz, M. D. Graham, and J. J. de Pablo. The effect

of hydrodynamic interactions on the dynamics of DNA translocation through

pores. J. Chem. Phys, 128:085102, 2008.

[74] R. M. Jendrejack, J. J. de Pablo, and M. D. Graham. Stochastic simulations

of DNA in flow: Dynamics and the effects of hydrodynamic interactions. J.

Chem. Phys, 116:7752, 2002.

[75] R. M. Jendrejack, E. T. Dimalanta, D. C. Schwartz, M. D. Graham, and J. J.

de Pablo. DNA dynamics in a microchannel. Phys. Rev. Lett., 91:038102, 2003.

Page 170: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

158

[76] R. M. Jendrejack, M. D. Graham, and J. J. de Pablo. Hydrodynamic interac-

tions in long chain polymers: Application of the chebyshev polynomial approx-

imation in stochastic simulations. J. Chem. Phys, 113:2894, 2000.

[77] R. M. Jendrejack, D. C. Schwartz, J. J. de Pablo, and M. D. Graham. Shear-

induced migration in flowing polymer solutions: Simulation of long-chain de-

oxyribose nucleic acid in microchannels. J. Chem. Phys, 120:2513, 2004.

[78] R. M. Jendrejack, D. C. Schwartz, M. D. Graham, and J. J. de Pablo. Ef-

fect of confinement on DNA dynamics in microfluidic devices. J. Chem. Phys,

119:1165, 2003.

[79] J. L. Jones, M. Lal, J. N. Ruddock, and N. A. Spenley. Dynamics of a drop at

a liquid/solid interface in simple shear fields: A mesoscopic simulation study.

Faraday Discuss., 112:129, 1999.

[80] R. Kapral. Multiparticle collision dynamics: Simulation of complex systems on

mesoscales. Adv. Chem. Phys., 140:89, 2008.

[81] R. Kekre, J. E. Butler, and A. J. C. Ladd. Comparison of lattice-Boltzmann and

Brownian-dynamics simulations of polymer migration in confined flows. Phys.

Rev. E, 82:011802, 2010.

[82] N. Kikuchi, J. F. Ryder, C. M. Pooley, and J. M. Yeomans. Kinetics of the poly-

mer collapse transition: The role of hydrodynamics. Phys. Rev. E, 71:061804,

2005.

[83] P. R. Kramer, C. S. Peskin, and P. J. Atzberger. On the foundations of the

Page 171: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

159

stochastic immersed boundary method. Comput. Method. Appl. M., 197:2232,

2008.

[84] N. Laachi, M. Kenward, E. Yariv, and K. D. Dorfman. Force-driven transport

through periodic entropy barriers. EPL, 80:50009, 2007.

[85] A. J. C. Ladd and R. Verberg. Lattice-Boltzmann simulations of particle-fluid

suspensions. J. Stat. Phys., 104:1191, 2001.

[86] M. C. Lai and C. S. Peskin. An immersed boundary method with formal second-

order accuracy and reduced numerical viscosity. J. Comput. Phys, 160:705,

2000.

[87] A. Lamura and G. Gompper. Numerical study of the flow around a cylinder

using multi-particle collision dynamics. Eur. Phys. J. E, 9:477, 2002.

[88] L. G. Leal and E. J. Hinch. The effect of weak Brownian rotations on particles

in shear flow. J. Fluid Mech., 46:685, 1971.

[89] S. H. Lee and R. Kapral. Mesoscopic description of solvent effects on polymer

dynamics. J. Chem. Phys., 124:214901, 2006.

[90] H. Ma and M. D. Graham. Theory of shear-induced migration in dilute polymer

solutions near solid boundaries. Phys. Fluids, 17:083103, 2005.

[91] A. Malevanets and R. Kapral. Mesoscopic model for solvent dynamics. J. Chem.

Phys., 110:8605, 1999.

[92] J. T. Mannion, C. H. Reccius, J. D. Cross, and H. G. Craighead. Conformational

Page 172: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

160

analysis of single DNA molecules undergoing entropically induced motion in

nanochannels. Biophys. J., 90:4538, 2006.

[93] A. P. Markesteijn, O. B. Usta, I. Ali, A. C. Balazs, and J. M. Yeomans. Flow

injection of polymers into nanopores. Soft Matter, 5:4575, 2009.

[94] J. F. Marko and E. D. Siggia. Stretching DNA. Macromolecules, 28:8759, 1995.

[95] A. Milchev, V. Yamakov, and K. Binder. Escape transition of a polymer chain:

Phenomenological theory and monte carlo simulations. Phys. Chem. Chem.

Phys., 1:2083, 1999.

[96] S. T. Milner. Polymer brushes. Science, 251:905, 1991.

[97] R. Mittal and G. Iaccarino. Immersed boundary methods. Annu. Rev. Fluid

Mech., 37:239, 2005.

[98] K. Mussawisade, M. Ripoll, R. G. Winkler, and G. Gompper. Dynamics of

polymers in a particle-based mesoscopic solvent. J. Chem. Phys., 123:144905,

2005.

[99] M. Muthukumar. Translocation of a confined polymer through a hole. Phys.

Rev. Lett., 86:3188, 2001.

[100] M. Muthukumar and A. Baumgartner. Effects of entropic barriers on polymer

dynamics. Macromolecules, 22:1937, 1989.

[101] H. D. Nguyen and C. K. Hall. Molecular Dynamics Simulations of Fibril For-

mation by Random Coil Peptides. Proc. Natl. Acad. Sci., USA 101:16180, 2004.

Page 173: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

161

[102] A. E. Nkodo, J. M. Garnier, B. Tinland, H. J. Ren, C. Desruisseaux, L. C. Mc-

Cormick, G. Drouin, and G. W. Slater. Diffusion coefficient of DNA molecules

during free solution electrophoresis. Electrophoresis, 22:2424, 2001.

[103] D. Nykypanchuk, H. H. Strey, and D. A. Hoagland. Brownian motion of DNA

confined within a two-dimensional array. Science, 297:987, 2002.

[104] S. B. Opps, J. M. Polson, and N. A. Risk. Discontinuous molecular dynamics

simulation study of polymer collapse. J. Chem. Phys., 125:194904, 2006.

[105] H. C. Ottinger. Stochastic Processes in Polymeric Fluids: Tools and Examples

for Developing Simulation Algorithms. Springer, 1995.

[106] P. J. Park and W. Sung. Dynamics of a polymer surmounting a potential

barrier: The kramers problem for polymers. J. Chem. Phys., 111:5259, 1999.

[107] S. Peng, F. Ding, B. Urbanc, S. V. Buldyrev, L. Cruz, H. E. Stanley, and N. V.

Dokholyan. Discrete molecular dynamics simulations of peptide aggregation.

Phys. Rev. E, 69:041908, 2004.

[108] T. T. Perkins, D. E. Smith, and S. Chu. Single polymer dynamics in an elon-

gational flow. Science, 276:2016, 1997.

[109] T. T. Perkins, D. E. Smith, R. G. Larson, and S. Chu. Stretching of a single

tethered polymer in a uniform-flow. Science, 268:83, 1995.

[110] C. S. Peskin. Flow patterns around hear valves - numerical method. J. Comput.

Phys, 10:252, 1972.

Page 174: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

162

[111] F. Xijunand N. Phan-Thien, S. Chen, X. Wu, and T. Y. Ng. Simulating flow of

DNA suspension using dissipative particle dynamics. Phys. Fluids, 18:063102,

2006.

[112] I. V. Pivkin and G. E. Karniadakis. Controlling density fluctuations in wall-

bounded dissipative particle dynamics systems. Phys. Rev. Lett., 96:206001,

2006.

[113] C. Pozrikidis. Computation of periodic Green’s functions of Stokes flow. J.

Eng. M., 30:79, 1996.

[114] P. Pranay, S.G. Anekal, J.P. Hernandez-Ortiz, and M.D. Graham. Pair collisions

of fluid-filled elastic capsules in shear flow: Effects of membrane properties and

polymer additives. Phys. Fluids, 22:123103, 2010.

[115] A. Puliafitoa and K. Turitsyn. Numerical study of polymer tumbling in linear

shear flows. Physica D, 211:9, 2005.

[116] W. Reisner, N. B. Larsen, H. Flyvbjerg, J. O. Tegenfeldt, and A. Kristensen.

Directed self-organization of single DNA molecules in a nanoslit via embedded

nanopit arrays. Proc. Natl. Acad. Sci. U.S.A., 106:79, 2009.

[117] M. Revenga, I. Zuniga, and P. Espanol. Boundary conditions in dissipative

particle dynamics. Comput. Phys. Commun., 121-122:309, 1999.

[118] M. Ripoll, M. H. Ernst, and P. Espanol. Large scale and mesoscopic hydrody-

namics for dissipative particle dynamics. J. Chem. Phys., 115:7271, 2001.

Page 175: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

163

[119] M. Ripoll, K. Mussawisade, R. G. Winkler, and G. Gompper. Low-Reynolds-

number hydrodynamics of complex fluids by multi-particle-collision dynamics.

Europhys. Lett., 68:106, 2004.

[120] J. Rotne and S. Prager. Variational treatment of hydrodynamic interaction in

polymers. J. Chem. Phys., 50:4831, 1969.

[121] B. Roux. The calculation of the potential of mean force using computer simu-

lations. Comput. Phys. Commun., 91:275, 1995.

[122] J. N. Roux. Brownian particles at different times scales: a new derivation of

the Smoluchowski equation. Phys. A, 188:526, 1992.

[123] Y. Saad. Iterative Methods for Sparse Linear Systems, Second Edition. Society

for Industrial and Applied Mathematics, 2 edition, April 2003.

[124] D. Saintillan, E. Darve, and E. S. G. Shaqfeh. A smooth particle-mesh Ewald

algorithm for Stokes suspension simulations: The sedimentation of fibers. Phys.

Fluids, 17:033301, 2005.

[125] R. B. Schoch, J. Han, and P. Renaud. Transport phenomena in nanofluidics.

Rev. Mod. Phys., 80:839, 2008.

[126] C. M. Schroeder, H. P. Babcock, E. S. G. Shaqfeh, and S. Chu. Observation of

polymer conformation hysteresis in extensional flow. Science, 301:1515, 2003.

[127] C. M. Schroeder, R. E. Teixeira, E. S. G. Shaqfeh, and S. Chu. Characteristic

periodic motion of polymers in shear flow. Phys. Rev. Lett., 95, 2005. 018301.

Page 176: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

164

[128] K. L. Sebastian and A. Debnath. Polymer in a double well: dynamics of translo-

cation of short chains over a barrier. J. Phys.-Condens. Mat., 18:S283, 2006.

[129] K. L. Sebastian and A. K. R. Paul. Kramers problem for a polymer in a double

well. Phys. Rev. E, 62:927, 2000.

[130] E. S. G. Shaqfeh. The dynamics of single-molecule DNA in flow. J. Non-

Newtonian Fluid Mech., 130:1, 2005.

[131] N. Sharma and N. A. Patankar. Direct numerical simulation of the Brownian

motion of particles by using fluctuating hydrodynamic equations. J. Comput.

Phys., 201:466, 2004.

[132] A. Sierou and J. F. Brady. Accelerated Stokesian Dynamics simulations. J.

Fluid Mech., 448:115, 2001.

[133] S. S. Skiena. The Algorithm Design Manual. Springer-Verlag, New York, 1997.

[134] M. Slutsky. Diffusion in a half-space: From lord kelvin to path integrals. Am.

J. of Phys., 73:308, 2005.

[135] D. E. Smith, H. P. Babcock, and S. Chu. Single-polymer dynamics in steady

shear flow. Science, 283:1724, 1999.

[136] D. E. Smith and S. Chu. Response of flexible polymers to a sudden elongational

flow. Science, 281:1335, 1998.

[137] D. E. Smith, T. T. Perkins, and S. Chu. Dynamical Scaling of DNA Diffusion

Coefficients. Macromolecules, 29:1372, 1996.

Page 177: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

165

[138] S. B. Smith, L. Finzi, and C. Bustamante. Direct mechanical measurements

of the elasticity of single DNA-molecules by using magnetic beads. Science,

258:1122, 1992.

[139] S. W. Smith, C. K. Hall, and B. D. Freeman. Molecular Dynamics for Polymeric

Fluids Using Discontinuous Potentials. J. Comp. Phys., 134:16, 1997.

[140] T. M. Squires and S. R. Quake. Microfluidics: Fluid physics at the nanoliter

scale. Rev. Mod. Phys., 77:977, 2005.

[141] D. Stein, F. H. J. van der Heyden, W. J. A. Koopmans, and C. Dekker. Pressure-

driven transport of confined DNA polymers in fluidic channels. Proc. Natl. Acad.

Sci. U.S.A., 103:15853, 2006.

[142] N. C. Stellwagen, C. Gelfi, and P. G. Righetti. The free solution mobility of

DNA. Biopolymers, 42:687, 1997.

[143] N. P. Teclemariam, V. A. Beck, E. S. G. Shaqfeh, and S. J. Muller. Dynamics

of DNA polymers in post arrays: Comparison of single molecule experiments

and simulations. Macromolecules, 40:3848, 2007.

[144] T. Tlusty. Screening by symmetry of long-range hydrodynamic interactions of

polymers confined in sheets. Macromolecules, 39:3927, 2006.

[145] A. K. Tornberg and L. Greengard. A fast multipole method for the three-

dimensional stokes equations. J. Comput. Phys., 227:1613, 2008.

[146] D. Trebotich, G. H. Miller, P. Colella, D. T. Graves, D. F. Martin, and P. O.

Page 178: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

166

Schwartz. A Tightly Coupled Particle-Fluid Model for DNA-Laden Flows in

Complex Microscale Geometries. Comp. Fluid Solid Mech., pages 1018, 2005.

[147] L. N. Trefethen and D. Bau. Numerical Linear Algebra. SIAM: Society for

Industrial and Applied Mathematics, 1997.

[148] P. T. Underhill, J. P. Hernandez-Ortiz, and M. D. Graham. Diffusion and spatial

correlations in suspensions of swimming particles. Phys. Rev. Lett., 100:248101,

2008.

[149] P. T. Underhill and P. S. Doyle. On the coarse-graining of polymers into bead-

spring chains. J. Non-Newtonian Fluid Mech., 122:31, 2004.

[150] O. B. Usta, A. J. .C Ladd, and J. E. Butler. Lattice-Boltzmann simulations

of the dynamics of polymer solutions in periodic and confined geometries. J.

Chem. Phys, 122:094902, 2005.

[151] J. L. Viovy. Electrophoresis of DNA and other polyelectrolytes: Physical mech-

anisms. Rev. Mod. Phys., 72:813, 2000.

[152] M. J. Vogel, P. Ehrhard, and P. H. Steen. The electroosmotic droplet switch:

Countering capillarity with electrokinetics. Proc. Natl. Acad. Sci. U.S.A.,

102:11974, 2005.

[153] Y. Wang, I. Teraoka, F. Y. Hansen, . H. Peters, and O. Hassager. A Theo-

retical Study of the Separation Principle in Size Exclusion Chromatography.

Macromolecules, 43:1651, 2010.

Page 179: BROWNIAN DYNAMICS SIMULATION OF DNA IN COMPLEX …Li, Li Xi, Samartha Anekal, Matthias Rink, Mauricio Lopez, Aslin Izmitli, Pratik Pranay, Kushal Sinha, Rafael Henriquez Rivera, and

167

[154] N. Watari, M. Makino, N. Kikuchi, R.G. Larson, and M. Doi. Simulation of

DNA motion in a microchannel using stochastic rotation dynamics. J. Chem.

Phys, 126:094902, 2007.

[155] S. A. Williams, J. B. Bell, and A. L. Garcia. Algorithm Refinement for Fluc-

tuating Hydrodynamics. SIAM Multiscale Modeling and Simulation, 6:1256,

2008.

[156] R. G. Winkler. Semiflexible Polymers in Shear Flow. Phys. Rev. Lett.,

97:128301, 2006.

[157] H. Yamakawa. Transport properties of polymer chains in dilute solution: Hy-

drodynamic interaction. J. Chem. Phys., 53:436, 1970.

[158] Y. Zhang, A. Donev, T. Weisgraber, B. J. Alder, M. D. Graham, and J. J.

de Pablo. Tethered DNA dynamics in shear flow. J. Chem. Phys, 130:234902,

2009.

[159] H. Zhao, A. H. G. Isfahani, L. N. Olson, and J. B. Freund. A spectral boundary

integral method for flowing blood cells. J. Comput. Phys., 229:3726, 2010.

[160] R. Zwanzig. Nonequilibrium Statistical Mechanics. Oxford University Press,

2001.


Recommended