<<

Depletant induced attractions in red blood cells.

by

Austin Nehring

A thesis submitted to the School of Graduate and Postdoctoral Studies in partial fulfillment of the requirements for the degree of

Master of Science in Material Science

University of Ontario Institute of Technology

Supervisor: Dr. Hendrick de Haan

Oshawa, Ontario, Canada

October 2018

Copyright c Austin Nehring, 2018 Thesis title: Depletant induced attractions in red blood cells.

An oral defense of this thesis took place on October 19, 2018 in front of the following examining committee:

Examining Committee

Chair of Examining Committee Dr. Franco Gaspari Research Supervisor Dr. Hendrick de Haan Examining Committee Member Dr. Aaron Slepkov Examining Committee Member Dr. Olena Zenkina External Examiner Dr. Nisha Agarwal

The above committee determined that the thesis is acceptable in form and content and that a satisfactory knowledge of the field covered by the thesis was demonstrated by the candidate during an oral examination. A signed copy of the Certificate of Approval is available from the School of Graduate and Postdoctoral Studies. Abstract

Red blood cells suspended in plasma are able to aggregate into linearly stacked rouleau. The aggregations can form complex clusters and branching networks which cause complications in various pathological cases. The self assembly and biophysics behind the aggregation of red blood cells into rouleau remains under- explored. This thesis employs coarse-grained molecular simulations to model erythrocytes in a disperse limit subject to short range implicit depletion forces. This work demonstrates that the depletant interaction is sufficient to account for sudden transitions into aggregate states. Furthermore, this work demonstrates that the specific volume fraction of depletants is directly linked to the morpholo- gies of the aggregate states observed.

ii • I hereby declare that this thesis consists of original work of which I have authored. This is a true copy of the thesis, including any required final revisions, as accepted by my examiners.

• I authorize the University of Ontario Institute of Technology to lend this thesis to other institutions or individuals for the purpose of scholarly re- search. I further authorize University of Ontario Institute of Technology to reproduce this thesis by photocopying or by other means, in total or in part, at the request of other institutions or individuals for the purpose of scholarly research. I understand that my thesis will be made electronically available to the public.

Page iii of 89 Statement of contributions

• This work has been published as: Austin Nehring, Tyler N. Shendruk, and Hendrick W. de Haan. Morphology of depletant-induced erythrocyte ag- gregates. Soft Matter, 14:81608171, 2018. [1]

• The images provided by Lorena Buitrago from the Allen and Frances Adler Laboratory of Blood and Vascular Biology at the Rockefeller University for dif- ferential interference contrast microscopy images.

iv Contents

1 Introduction1

1.1 Rouleaux Formation...... 1 1.2 Depletant interaction...... 2 1.3 Introduction to Coarse-Grained Simulations...... 5 1.3.1 Molecular Dynamics...... 5 1.3.2 Coarse-Grained Molecular Dynamics...... 5 1.3.3 Langevin Dynamics...... 6 1.3.3.1 Drag in a Viscous Fluid...... 7 1.3.3.2 Thermal Fluctuations...... 8 1.3.4 Lennard Jones & Weeks-Chandler-Anderson...... 13 1.3.5 Finitely Extensible Nonlinear Elastic Bonds...... 15 1.3.5.1 Diffusion of Coupled Particles...... 16 1.3.6 Angle potentials...... 17 1.3.7 Dihedral Bonds...... 18 1.3.8 Depletant potential...... 18 1.3.9 Numerical Integration - Velocity Verlet method...... 21 1.4 Morphology of depletant-induced erythrocyte aggregates.... 23

2 Simulation Setup 24

2.1 Models...... 24 2.2 Simulation and simulation parameters...... 30 2.2.1 Aggregation of red blood cells...... 30 2.2.2 Rouleau dispersion...... 30

v 2.2.3 Interaction stability...... 31 2.2.4 Parameters...... 31 2.2.5 Observables...... 33

3 Results & Discussion 36

3.1 7 sphere cell...... 36 3.2 High aspect ratio cell...... 39 3.3 Smooth cell...... 42 3.3.1 Aggregation...... 42 3.3.2 Rouleau dispersion...... 46 3.3.3 Interaction Stability...... 47

4 Conclusion 62

Appendices 75

A Simulation time 76

B Cell diffusion 77

C GPU acceleration 79

D PhytoSpherix R code and visualization development 80

E cNAB.LAB Computing Resources 87

vi List of Figures

1.1 Schematic of Rouleaux...... 2 1.2 Schematic of Depletant Interaction...... 3 1.3 Lennard Jones and WCA...... 15 1.4 Schematic: Angle potential...... 17 1.5 Schematic: Dihedral potential...... 19 1.6 Asakura and Oosawa pair potential...... 20

2.1 7 cell model...... 25 2.2 7 cell model: bonds...... 26 2.3 19 cell model...... 26 2.4 Smoothcell model...... 27 2.5 Defining structures found within the simulation and results... 28 2.6 Rouleau-rouleau interactions within aggregated clusters..... 29

3.1 Results: 7 sphere model...... 37 3.2 Rouleau formed by 7 sphere model...... 38 3.3 Results: High aspect cell model...... 40 3.4 Results: High aspect ratio model - 20 cell...... 41 3.5 High aspect ratio cluster...... 42 3.6 High aspect ratio rouleau...... 43 3.7 Results: Smoothcell model...... 44 3.8 Smoothcell columnar rouleau...... 45 3.9 Amorphous aggregate of columnar rouleaux...... 45 3.10 Rouleau dispersion...... 47 3.11 Hysteresis comparison...... 48

Page vii of 89 3.12 Face-to-face lifetime...... 49 3.13 Face-to-face potential energy...... 50 3.14 Face-to-face normalized energy...... 51 3.15 Face-to-face lifetime, energy, normalized energy...... 53 3.16 Face-to-side life-time...... 54 3.17 Face-to-side energy...... 55 3.18 Face-to-side normalized energy...... 56 3.19 Face-to-side: Lifetime, energy, normalized energy...... 57 3.20 Comparison of 10 and 20 cell systems - number of clusters.... 58 3.21 Normalized comparison of 10 and 20 cell systems - normalized number of clusters...... 59 3.22 Comparison of 10 and 20 cell system - number of aligned..... 60 3.23 Distribution of aligned cells at φ = 0.075 ...... 60 3.24 Amorphous aggregate of 20 RBCs...... 61

B.1 Angular and rotational mean squared displacement...... 78

D.1 PhytoSphyrix R no attraction...... 85 D.2 PhytoSphyrix R low attraction...... 85 D.3 PhytoSphyrix R medium attraction...... 86 D.4 PhytoSphyrix R high attraction...... 86

List of Tables

2.2.1 Simulation Parameters in HOOMD...... 32

E.0.1 Panda - Computational Resource ...... 88 E.0.2 Computational Resources ...... 89

viii Chapter 1

Introduction

This thesis primarily focused on the formation and simulations of rouleau. The theory and methods discussed in the following chapters have also been extended to projects outside of this thesis and are presented in appendixD&E.

1.1 Rouleaux Formation

Rouleau are the organized structure of aggregate red blood cells observed in the human body. Rouleau is identified by the cells stacked on top of each other in a face to face orientation resembling a stack of coins (1.1). The observations of red blood cells forming either single rouleau or many rouleaux within microvascular flows [2,3,4] have been deeply studied. In pathological cases, such as malaria or , aggregations of red blood cells can block capillaries pre- venting blood flow [5,6,7]. Other examples arise from cardiovascular diseases and of the body [8,9, 10, 11]. Furthermore the inclusion of antibodies has been reported to be correlated with aggregates of erythrocytes near anti- gens [12, 13]. Rouleau can also arise in non-specific cases [14, 15] and can be observed in sites with inflammation [16] where blood proteins such as fibrino- gen can lead to enhanced aggregation [14, 17]. The aggregates of red blood cells can have varied lengths and also have branch points creating larger overall net- works (Fig. 1.1b) [18]. Currently there are two theories of how red blood cells

1 Figure 1.1: Images of rouleaux, columnarly ordered aggregates of red blood cells. (a) Simulation snapshot. (b) Differential interference contrast (DIC) mi- croscopy image of human erythrocytes in plasma courtesy of Lorena Buitrago (throughout). can aggregate in clusters. The first proposes plasma macromolecules bind the blood cells together. The second theory suggests the depletant induced potential can cause aggregation. This thesis aims to explore how the depletant induced potential can cause the red blood cells to collapse and explore their aggregate morphology. There is evidence to suggest that the concentration of fibrinogen or other pro- teins are correlated with the aggregation of red blood cells [19, 20]. Structural changes in prokaryotic chromosomal material is also thought to be induced by the depletion interaction [21, 22, 23]. Recently independent experimental work by Wagner et al. & Armstrong et al. concluded that the depletion interaction was sufficient for RBC aggregation [24, 25]. These observations provide a pos- sible avenue in which depletion forces are causing the adhesion of red blood cells.

1.2 Depletant interaction

Depletant interactions occur when a system has large colloids in a bath of small particles or macromolecules (depletants). The closest that a depletant can get to a colloid is the sum of the radii of the depletant and the colloid (Fig. 1.2a). This thin sheath is referred to as the exclusion layer and it represents the volume in

Page 2 of 89 Figure 1.2: (Above) Diagram illustrating the entropic basis of depletion interac- tions, in which two colloids of radius Rc are suspended in a bath of depletants of size Rd. (Below) When the two colloids make contact, their excluded volumes overlap and the depletants gain free volume, producing a net entropic gain and depletant-induced attraction which the depletant cannot access. When two colloids are brought close together the exclusion layers overlap causing an increase to the volume of the system for the depletants (Fig. 1.2b). An increase in volume is entropically favorable for the depletants while the restricted conformation is disfavorable for the colloids. If there are a sufficient number of depletants then their increase in entropy is greater than the decrease of entropy of the colloids. The result is that colloids remain close together as to maximize the systems net entropy. Generally one might think of entropy as a way for a system to disassemble and diffuse particles evenly around a volume. However, this means the deple- tant interaction can cause colloids to group together and aggregate in the right circumstances. Alternatively there is a kinetic picture of the depletant interac- tion. Imagine a sphere submerged in water, water exerts pressure on all sides of the sphere evenly. If two spheres are brought close together such that there is no fluid between them each sphere is subject to an un-even pressure. This results in a net force which keeps the spheres close together. The implementation of the

Page 3 of 89 depletant interaction is discussed further in Section 1.3.8. It is important to also recognize that specific macromolecules can interact with the surfaces of two RBCs forming short macromolecular bridges that may tether RBCs [26]. Some molecules, perhaps including the experimentally impor- tant neutral dextran, may act simultaneously as a depletant and bridging macro- molecule [27, 28, 29, 30, 31]. It is also possible that the depletant might cause the collapse of RBCs while macromolecules are additional factors in the stability of the structure over time. Nevertheless it is clear that depletant particles could be playing a role in the formation of rouleau. Currently there have been various studies have been performed at different scales of experiment and simulation. Studies using oblate spheroids to model erythrocytes have been implemented to explore blood flow [32, 33]. Studies in- cluding detailed membrane deformation and stacking of a limited number of cells have been implemented to show rouleau formation in fluid flowing through a cylinder [34]. Disk-like liquid crystals, referred to more commonly as discotic liquid crystals in the literature, show that the anisotropic shape plays a role in the alignment of the crystals [35]. Simulations of anisotropic colloids measure the change in the order parameter of a bulk phase as it transitions to different states [36, 37]. Nano-discs and vesicle-based objects have been used to build rod-like structures with long ranged order [38]. This thesis will opt to ignore many of the features of a such as the biconcavity or the details of the chemical structure. Instead, a primitive RBC model will be built using connected spheres. Using a simple structure the thesis will show that the deple- tant interaction is sufficient to explain the formation of rouleau. Furthermore, instead of moving to larger system sizes and increased densities this thesis will investigate the energetics of particular configurations to show that the results in larger systems can be informed by the stability of specific interactions between cells.

Page 4 of 89 1.3 Introduction to Coarse-Grained Simulations

1.3.1 Molecular Dynamics

Molecular dynamics is the act of solving Newtons second law by adding all the forces on a particle together and then integrating twice to find the velocity and position at some later point in time. Repeated many times over this allows the evolution of a particle’s trajectory through some environment. In the context of a system of red blood cells this would require solving for all the forces acting on the red blood cells and surrounding particles. Even a small system of just a few red blood cells in fluid could comprise of tens of billions of atoms if not more. As such calculating the forces, interactions and numerical integrations of every molecule and fluid particle would take an unreasonable amount of time. Consequently a fully detailed molecular dynamics simulation is unobtainable. However, there are three methods used to reduce the computational intensity of solving Newtons law for this system. The first method is by coarse-graining the red blood cells in the simulation. The second method will reduce the resources needed to calculate how the fluid interacts with the red blood cells. Finally, the third method will take advantage of the depletant pair potential to accelerate the calculations of interactions with the depletants.

1.3.2 Coarse-Grained Molecular Dynamics

A single red blood cell can contain millions of molecules, each comprised of indi- vidual atoms. Modeling the interactions would be extremely difficult so coarse- graining techniques are employed to reduce the computational cost of modeling the cell. The diameter of a single RBC is between [6 − 8]µm while atoms exist on the scale of 1Å. Because the forces and motion of any single atom is insignif- icant on the scale of the size and motion of the red blood cell all the atoms and molecules are grouped into a single particle. The goal is to reduce the computa-

Page 5 of 89 tional cost of the RBC without losing details like the shape, structure or how it moves. Using fewer particles makes the computation easier at the expense of in- formation about the exact structure of the RBC. The thesis will explore different coarse-grained models ranging from 7 to 19 spheres. The details of the models of the RBCs are outlined in section 2.1.

1.3.3 Langevin Dynamics

Blood contains red blood cells suspended in . Blood plasma is comprised mostly of water with proteins and molecules of various sizes [26, 39, 40, 41]. Consequently there are a significant number of particles which would need to be simulated to model the fluid [42, 43]. Furthermore, in a system of a few coarse-grained RBCs the vast majority of the computational time would be spent solving Newtons second law for all the fluid particles. A system of 10 RBCs comprised of 19 spheres (190 particles) in even a small box might still have on the order of 105 or 106 fluid molecules. Langevin Dynamics (LD) allows us to focus on red blood cells and the depletants instead of spending most of the computational effort on the fluid. If you tracked the motion of a particle in water over some time the motion appears Brownian; the particles motion is erratic and doesn’t follow any given path. LD replaces the particles interactions in the fluid with two forces that when evolved over time replicate the fluctuations and erratic movement of a particle. The proceeding section will describe how LD replaces the interactions of the fluid with two forces that are directly used to calculate the motion of the particles on the red blood cells. The end goal is to solve Newtons second law for the coarse-grained particles that make up the red blood cell. For now consider the force on just a single particle which can be written as

X ~ Fi = mi ~ai (1.3.1)

Page 6 of 89 where solving for the acceleration is relatively easy by dividing the net force by the mass. However, the net force contains the interactions with all the surround- ing fluid. The particle is being hit by many small fluid particles each transferring their own kinetic energy to the particle. Fluid as a whole drives the movement of the particle through the system by the act of many collisions between the particle and the molecules in the fluid. The particle also experiences drag via interactions with the fluid in which its own kinetic energy is transferred to the surrounding fluid. In both cases the driving and dissipation of kinetic energy on the particle are two distinct forces that can be described. Calculating the overall increase of kinetic energy from the fluctuations in the fluid and the decrease in kinetic energy from the drag caused by moving through the fluid allows us to implicitly recreate the erratic motion of the particle with only two forces.

1.3.3.1 Drag in a Viscous Fluid

Drag is a complex and non trivial attribute of an object. Drag behaves very dif- ferently depending on the object’s orientation and velocity relative to the fluid itself. This makes it difficult to determine exactly how the fluid and object are interacting. The Reynolds number quantifies the relation of the objects size and velocity in the fluid µ−1 Re = νL (1.3.2) ρ

ν represents the characteristic velocity of the particle in the fluid. L represents the characteristic length scale of the object, µ is the dynamic viscosity of the fluid and ρ represents the density of the fluid. The Reynolds number itself is a dimensionless quantity but its value can give a good indication of the type of environment the object is in. For a human treading in water our characteristic length scale exists at meters and the velocity of our arms and legs in meters per second. Using that and the density of water

Page 7 of 89 !−1 m 8.9 · 10−4[P a s] Re = 1[m] · 1[ ] ≈ 1.1 · 106 (1.3.3) Human in Water s Kg 1000[ m3 ] a Reynolds number at that magnitude corresponds to turbulent fluid. However, the length scale of a red blood cell is µm. The resulting Reynolds number is about 1.1, which is significantly different; a blood cell exists in fluid close to laminar flow. For systems of very small Reynolds numbers such as Re < 1 a spherical object experiences drag that obeys stokes law:

~ Fdrag = −6πµrs~v (1.3.4)

The drag force is related to the radius of the sphere rs and the velocity relative to the fluid. It is more common to see the force written in a condensed form

~ Fdrag = −γ~v (1.3.5)

where −γ = 6πµrs is named the friction coefficient. The drag term encompasses all the interactions that would otherwise slow down the particle as it moves through the fluid.

1.3.3.2 Thermal Fluctuations

Thermal fluctuations arise from the collisions of the particle with the fluid. In thermal equilibrium the fluid collides randomly in all directions with the parti- cle. While these collisions can impart enough kinetic energy to cause the parti- cle to move because they come from all directions the particle is not directed towards any location. However, there can be situations where enough fluid molecules hit the particle on the same side or an energetic particle can transfer a significant amount momentum relative to the rest of the fluid. Both situations cause a non-negligible increase in the kinetic energy of the particle at some in-

Page 8 of 89 stantaneous moment in time. This is represented by a random term

~   R(t) = Rx(t),Ry(t),Rz(t) (1.3.6)

which is a vector comprised of three orthogonal components (Cartesian) at some point in time. It is important to note that being random means that the vector at any point in time is not correlated to itself at any other point in time. As a consequence the average of the random term over a long time of each com- ponent is zero. This is mathematically represented

~   hR(t)it = hRx(t)it, hRy(t)it, hRz(t)it (1.3.7) ~   hR(t)it = 0, 0, 0 (1.3.8) where each component has a zero average over a long time. The random term can be modeled by a stationary Gaussian process with a zero mean. Although, collectively a series of non-zero terms acting on an object still results in motion over time. A function R~(t) can be used to model a particle which is being subject to some instantaneous non-zero force. The function R~(t) has a range between the values [−1, 1] and outputs a random number for any input of time t. The form of the functions is represented by

~ ~ Frandom = KrR(t) (1.3.9)

where additionally the multiplied constant Kr represents the magnitude of the random term. A fluid at a high temperature or low viscosity would have larger

fluctuations which can be represented by a large Kr. Alternatively a fluid at a low temperature would have smaller less frequent collisions which would be represented by a small Kr. The particles’ interactions with temperature and vis- cosity are also coupled to the diffusion as described by the Einstein—Smoluchowski

Page 9 of 89 relation

kBT D = (1.3.10) γ which depends on the temperature the friction coefficient and the boltzmann constant kB. Kr should be something that relates the temperature and drag to the fluid in an appropriate manner to the model. To determine Kr we write the equation of motion of the particle considering drag (1.3.5) and the random term (1.3.9)

X ~ m~ai = F ~ ~ m~ai = KrR(t) − γv(t) (1.3.11) in the overdamped limit. When the Reynolds number is very small objects are dominated by the viscous effects of the fluid. In this limit any inertia that a parti- cle has immediately decays away. Consequently any acceleration that a particle was subject to also quickly decays. This is represented by setting the accelera- tion term ~ai ≈ 0. With no inertia the motion of our particle acts as if it is given a burst of kinetic energy, which then is transferred into the fluid by drag as its kinetic energy is damped out by the drag term. With acceleration set to zero and equating the two terms on the right hand side we come to a relation

~ ~ KrR(t) = γv(t) (1.3.12) between the random force term and drag. Similar to the random term we can take the time average to explore the equation of motion over time. Averaging

Page 10 of 89 over a long time of each side reveals that the average velocity is zero

~ hKrithR(t)it = hγith~v(t)it, Average of a constant is the constant itself  Kr 0 = γh~v(t)it, Average of the random term is zero 1.3.8

0 = h~v(t)it (1.3.13) which indicates that our particle does not experience drift. In thermal equilib- rium with a non-flowing fluid our particle does not have a motion in any par- ticular direction. However, the particle still has a mean square displacement. Multiplying Eqn.1.3.12 with itself at another point in time t0

~ ~ 0 0 KrR(t) · KrR(t ) = γ~v(t) · γ~v(t ) (1.3.14)

2 ~ ~ 0 2 0 Kr R(t)R(t ) = γ ~v(t)~v(t ) (1.3.15)

The term R~(t)R~(t0) is the relationship of each random vector at any two points in time. When averaged the two terms should only be correlated if t = t0, other- wise they are uncorrelated; this is a delta function hR~(t)R~(t0)i = δ(t0 − t). When integrated the delta function collapses on itself being equal to 1 when t = t0. Ad- ditionally a dimensionality factor of 3 comes from the R~ term being comprised of three dimensions

Z t Z t Z t Z t 2 ~ ~ 0 0 2 0 0 Kr hR(t)R(t )idtdt = γ h~v(t)~v(t )idtdt 0 0 0 0 Z t Z t Z t 2 0 0 2 0 0 Kr δ(t − t)dt = γ h~v(t)~v(t )idtdt 0 0 0 Z t Z t Z t 2 0 2 0 0 Kr (3)dt = γ h~v(t)~v(t )idtdt 0 0 0 Z t Z t 2 2 0 0 Kr (3t) = γ h~v(t)~v(t )idtdt (1.3.16) 0 0 furthermore the Green-Kubo theorem [44] states that right hand side represents the mean squared displacement, which in 3 dimensions is 6Dt. Einstein connects the MSD to the diffusion through his diffusion equation (Eq. 1.3.10) which when

Page 11 of 89 substituted k T  K2(3t) = γ26 B t (1.3.17) r γ allows us to solve for Kr. When substituted back into 1.3.9 the full term can be written as ~ p ~ Frandom = 2kBT γR(t) (1.3.18)

which is the second force acting on the particle at any instant. This force im- parts kinetic energy onto the particle in a random direction and in opposition the increased kinetic energy from each successive kick is damped out by the drag term. The balance between the two terms acts as a thermostat. Increasing the drag decreases the particles kinetic energy over time while increasing the mag- nitude of the random force increases the kinetic energy on the particle. Adding the drag (1.3.5) and the random term (1.3.18) to P F~ = m~a yields the second order Langevin equation

~ ~ Frandom + Fdrag = m~a p ~ 2kBT γR(t) − γ~v = m~a (1.3.19)

which represents the effects of a fluid on a particle. In summary, MD can model and solve how a red blood cell will be effected by depletants. A red blood cell is very large and comprised of many atoms and molecules, so we will coarse-grain the model to remove the high computational task of computing the interactions within a red blood cell. With a model on the order of tens of particles even a small simulation box would still have potentially hundreds of thousands of water molecules, thus most of our time would be spent calculat- ing water-water interactions. Langevin dynamics replaces the fluid with a drag term and a random term drastically reducing the computational impact of the fluid. We can now solve Newtons second law for a particle in our MD simula- tions very quickly and all of the computational time is spent on the motion of the red blood cells.

Page 12 of 89 1.3.4 Lennard Jones & Weeks-Chandler-Anderson

When considering how red blood cells will be interacting one can imagine the RBCs bumping off each other or sliding along the surface of another cell. In each case the red blood cells have excluded volume which prevent the other RBCs from occupying the same space. When choosing to model how the RBCs interact the easiest method is to approach collisions like billiard balls; there is no force when they are not touching and a repulsive force when they are. Un- fortunately this causes a dire issue in simulations as these are discontinuous forces which create large numerical errors every time two particles interact. It is therefore advantageous to have a smooth transition between a zero force and a high repulsive force. A modified version of the Lennard Jones potential (LJ) is adopted to accomplish this task (Fig.1.3, left). The LJ potential has zero potential when the particles are far away. As two particles start to approach each other there is an attractive well which brings them closer together. If two particles get even closer the potential then rises to infinity (Fig.1.3, left). The Lennard Jones potential considers the interaction between any ith particle to its jth neighbor with some separation (r). The nominal diameter of the particle sigma (σ) and energy epsilon () represent the distance and energy scale. The LJ can be written as the sum of two terms

" #  σ 12  σ 6 ULJ = 4 − (1.3.20) rij rij

where there is an attractive term of 6 modeled from dipole-dipole interac- tions. The second term is repulsive to the power of 12, the number 12 is chosen as it is the square of the attractive term. The minimum of the function reaches −

1 at exactly r = 2 6 σ. The LJ potential has the repulsive behaviour which satisfies the requirements of having continuous forces. However, the long attractive re- gion is not of interest to this study as we are testing only the effects of depletants. The attractive tail models long range electrostatic interactions while depletants

Page 13 of 89 have a specific short ranged potential which also has a specified cut-off. The de- pletant interaction is discussed in greater detail in section 1.3.8. Therefore the LJ potential is modified in two ways, the first being that the function is cut off after

1 the global minimum (Fig.1.3, center). This is chosen at a critical value rc = 2 6 σ where the derivative of the potential is exactly zero. This is accomplished by changing to a piece-wise function

 " # 12 6   σ   σ  4 − , if r < rc  rij rij UT-LJ =  0, r ≥ rc with the aforementioned cut-off. To keep the energy in the simulation consistent the potential is shifted by  which is the value of the function at the minimum (Fig.1.3, right). This ensures when no particles are interacting the system has no potential energy due to its interaction terms.

 " # 12 6   σ   σ  4 − + , if r < rc  rij rij UTS-LJ =  0, r ≥ rc

This ensures that the function continuously approaches zero energy as the separation of two particles increases. It sufficiently approximates a hard sphere model without causing any discrete jumps or regions where numerical errors can easily be created. The shift is not important as the force is the gradient of the potential, however if the energy between two particles was of interest the shift ensures correct measurements are recorded. This particular modification of the LJ potential is called the Weeks-Chandler-Anderson potential (WCA) [45] (Fig.1.3, right). The WCA potential will be used to maintain excluded volume between the red blood cells in the simulation. Using just the equation of motion (Eq.1.3.19) and adding in the WCA allows for one to assemble a box of repulsively interact- ing particles. The following sections detail how to couple particles together into

Page 14 of 89 Figure 1.3: The Lennard Jones interaction, truncated LJ and shifted & truncated LJ with  = σ = 1.0. The vertical dotted line indicates the cut off (rc) where the slope of the function is equal to zero. The rightmost panel indicates the addition of  = 1.0 to increase the minimum of the function from −1.0 to 0.0. This partic- ular modification is known as the Weeks-Chandler-Anderson potential (WCA). the structure of the red blood cell.

1.3.5 Finitely Extensible Nonlinear Elastic Bonds.

Building a red blood cell out of particles requires a way to keep the particles to- gether. It is common to use a spring to bond particles so that they remain close.

1 2 The best place to start is the Hookean spring F = 2 kx which increases the force on the particles as their separation increases. Unfortunately a Hookean spring has a unintended side effect; there is no way to limit how far two particles can separate. In the context of our RBC this means that its possible a large thermal fluctuation or a collision could push particles between the bonds of their neigh- bors; This is known as bond crossing. To prevent this from occurring the Finitely Extensible Nonlinear Elastic bond (FENE) is chosen instead. The FENE bond is similar in form to a Hookean spring but allows one to choose the maximum pos- sible stretch a bond can have. The FENE potential is a function of the separation of two particles by r:

2 ! 1 2 r UFENE(r) = krmax ln 1 − 2 (1.3.21) 2 rmax

where rmax is the largest distance that two objects can separate. The choice of

Page 15 of 89 rmax is usually set to 1.5 diameter of the particle. Meaning if two particles were to separate they only create an opening which is half the diameter of the parti- cles. If another particle was to cross the bond it would require a large amount of energy to create a opening that it could fit through. The spring constant k mag- nifies the potential and increasing k directly increases the restoring force of the FENE bond. The constants chosen reflect those of the polymer model of Kremer and Grest [46] where the spring constant is set to 30/σ2 with a maximum bond distance 1.5σ.

1.3.5.1 Diffusion of Coupled Particles

When two particles are coupled together via FENE bonds the diffusion of the pair will be lower then that of a single particle. This is due to the random ther- mal kick that is applied on each particle being independent. Thermal kicks that are pointed in different directions are reflected by the movement of the center of mass of the pair. As N many particles are joined together the sum of the random kicks quickly approaches zero which would prevent any diffusion from occur- ring at all. To alleviate the effect of the reduced diffusion the drag coefficient can be reduced by a factor related to the number of particles in an object. This al- lows each thermal kick to have a greater effect over some distance which allows the larger paired particle to diffuse at a greater rate. In practice the diffusion constant is decreased by a factor of N −1 so the drag term can be multiplied by a factor of N −1 to remain consistent. However, it should be noted that real objects do not behave this way. Diffusion is also related to the size, shape and surface area. The choice to decrease the drag allows the RBCs to counter the effects of reduced diffusion. There isnt any large asymmetric effects as all RBCs have the same number of particles and thus are all affected in the same manner.

Page 16 of 89 1.3.6 Angle potentials

Particles which are bonded together, and have excluded volumes are considered “freely jointed”. As long as the center to center distance of each particle satisfies the FENE and the WCA potentials there is nothing to stop the particles from ro- tating around each other. In chains of particles this allows the objects to contort and twist without penalty. To ensure that the RBC does not freely twist or move unphysically a harmonic angle potential is added. This harmonic potential is evaluated between three particles across the center of the cell (Fig. 1.4). The potential considers the angle generated by the triplet, and then applies a corre- sponding force to the center monomer. In this case the angle θ0 where θ0 = π

1  2 V (θ) = k θ − θ0 (1.3.22) 2 prevents the middle of the red blood cell from moving out of the plane of the disk. Similar to the coefficient in the FENE bond k can be increased to reduce the central monomers’ motion out of the plane of the RBC. This is chosen to maintain a disk-like shape while allowing some flexibility in the cell.

Figure 1.4: Schematic of the angle potential between 3 spheres i,j,k. The angle θ is evaluated between pairs i,j and j,k. Pictured in solid blue is the angle bond between the particles

Page 17 of 89 1.3.7 Dihedral Bonds

The final bond used to generate the red blood cell is called a dihedral bond. Mo- tion of the exterior particles is controlled by the dihedral bond. It is a harmonic potential which is placed across four particles in consecutive order (Fig. 1.5). Two sets of three particles (i,j,k & j,k,l) each create a plane. The intersection of the planes occurs at some angle θ. The dihedral bond controls the alignment of the intersecting planes by keeping them at a specified angle. In the case of a disk, setting the rest angle between the planes to be 0◦increases the rigidity of the red blood cell by forcing the particles to all lay flat.

1   V (r) = k 1 + d cos n · φ(r) (1.3.23) 2

By overlapping dihedral bonds around the outside of the cell the bonds prevents any individual part of the disk from twisting out of plane. Increasing k increases the rigidity of the exterior ring of the RBC. Similar to the angle bonds if the dihedral bonds are removed the exterior is free to fold into itself or fluctuate wildly. These bonds are chosen to eliminate extreme deformation while still trying to remain true to the general shape of a disk.

1.3.8 Depletant potential

The potentials discussed so far have set up the foundations to create a simulation of RBCs that have excluded volume. LD provides the motion and drag from the fluid and the WCA enforces that cells cannot pass through each other. The three bond types (FENE, angle, dihedral) control the structure of the cell. The last interaction is the attractive potential between cells. Depletant interactions will be modeled after using the Asakura and Oosawa (AO) model between spheres [47, 48, 49]. These interactions are short range attractions which depends only on the size of the colloid (Rc) and the depletant (Rd). For spherical colloids of

Page 18 of 89 Figure 1.5: The dihedral potential is generated between 4 particles i,j,k,l. The angle between the planes i,j,k (green) and j,k,l (red) is the dihedral angle. In the RBC this angle is set to θ = 0◦ which keeps the planes directly on top of each other.

radius Rc, the AO pair potential [47] is

 3  UAO R 3 r 1 r 3 = − 1 − + φd (1.3.24) kT Rd 4 R 16 R

where R = Rc +Rd and r is the center-to-center separation. The advantage of the depletant potential is that it doesn’t require information about the biology of the environment. It only requires information about the geometry of the particles and depletants. Because the size of the particles and depletants are kept con- stant the only variable is how many depletants are in the system. Increasing the volume fraction of depletants increases the attractive potential of the AO poten- tial (1.6). The AO potential also scales with the energy kT . Another advantage is that the potential replaces the need for the explicit depletants. Knowing the volume fraction of depletants in the simulation allows for the only variable to be separation distance, which can be easily calculated. Sections 1.3.3.1& 1.3.4 described the motion of a single particle in the simu- lations. The FENE bond, WCA, angle and dihedral potentials describe the forces of neighboring connected particles. The depletant potential describes the attrac-

Page 19 of 89 Figure 1.6: Energy plot of the AO pair potential between two colloids of radius Rc = 2.5 in a bath of depletants with radius Rc = 0.1. Shown are three plots with different volume fractions φ = [0.05, 0.10, 0.15]. tive interaction between the particles of different cells. All of these can be written in full in Newtons second law

X F~ = m~a ~ ~ ∇(UFENE + UWCA + UAngle + UDihedral + UDepletant) + Frandom + Fdrag = m~a (1.3.25) for each particle. The gradients are evaluated at each timestep and summed together as part of the net force on a particle at any given time. Once all the forces are determined and the acceleration has been solved for the particles velocity can be determined. If the velocity of the particle is known then the position of the particle at some later point in time can be calculated. Once the position of a particle is found all the forces can be recalculated and the process repeated. However, this can be a difficult task with some nuances which arise in how you reach the final position.

Page 20 of 89 1.3.9 Numerical Integration - Velocity Verlet method

As the framework of the model has been set up it is now useful to describe how the dynamics of the system are determined. The equation of motion needs to be numerically integrated as the set of N coupled differential equations represent a far too complex system to solve by hand. By solving for the acceleration, and then integrating twice the particles’ position can be determined at some later step in time. The timestep ∆t represents the increment in time. The first method to do this is by Euler’s method. The algorithm would follow these steps:

1. At time t0 calculate the force on a particle i as governed by its potentials.

2. The force on the particle divided by its mass is then the acceleration ~ai = ~ Fi(t0) Mi

3. Calculate the velocity: ~vi(t0 + ∆t) = ~vi(t0) + ~ai∆t

1 2 4. Calculate the position: ~xi(t0 + ∆t) = ~xi(t0) + ~vi∆t + 2~ai(t)(∆t)

5. Return to step 1

An issue arises with using this naive method in that Euler’s integration method is known as a non-symplectic integrator. Simply put this method does not con- serve energy. The error arises in the final step where the particle is placed in a new position according to the velocity. If a particle is moved from outside an energy barrier to inside an energy barrier (such as the WCA) it’s kinetic energy isnt modified. On the subsequent timestep the particle receives a huge force pushing it out of the field vastly increasing its own kinetic energy even further. This can cause systems of particles to begin to boil, or freeze depending on the potential as the energy in the system drifts away from the initial value. In order to keep track of the position and the velocity of each particle though time and conserve energy the Velocity Verlet halfstep algorithm is used. At each point in time it integrates all the forces on each particle and progresses them

Page 21 of 89 forward by the timestep noted as ∆t just as in Euler’s method. However, the

1 velocity and position are first calculated halfway through the timestep 2 ∆t; then all the forces and recalculated and the velocity is updated before finding the final position. This small change alters the kinetic energy of a particle as it enters a potential field so that the total energy (kinetic and potential) are constant. The modified algorithm follows these steps:

1. At time t0 calculate the force on a particle i as governed by its potentials.

~ 2. The Force on the particle divided by its mass is then the acceleration Fi(t0)·

Mi = ~ai

1 1 3. Calculate the halfstep velocity: ~vi(t0 + 2 ∆t) = ~vi(t0) + 2~ai(t0)∆t

4. Calculate the new position using the halfstep velocity: ~xi(t0+∆t) = ~xi(t0)+

1 ~vi(t0 + 2 ∆t)∆t

5. Calculate the forces on each monomer at the new position.

~ −1 6. Calculate the updated accelerations on the particle Fi(t0 + ∆t) · Mi = ~ai

1 1 7. Update the velocity at the final position: ~vi(t0 +∆t) = ~vi(t0 + 2 ∆t)+ 2~ai(t0 + 1 2 ∆t)∆t

8. Return to step 1

Note that the Verlet halfstep is not the same as using Eulers method with a smaller timestep. Reducing the timestep in Eulers method reduces the magni- tude of the error but it still does not conserve energy; while the Verlet integration modifies the kinetic energy of the particle at its original position and at its final position. This process is important as our models must conserve energy through the simulation.

Page 22 of 89 1.4 Morphology of depletant-induced erythrocyte ag-

gregates

The following sections of the simulation setup, results, discussion and conclu- sion have been consolidated and published in the paper “Morphology of depletant- induced erythrocyte aggregates” [1]. In addition to those results this thesis also includes an extended discussion on simulation details and 2 additional RBC models tested.

Page 23 of 89 Chapter 2

Simulation Setup

To model the red blood cell a coarse-grained method is employed. As men- tioned previously a full simulation including all of the particles of a single RBC would contain millions of atoms. Because the forces and motion of the atoms and molecules that make up a red blood cell are insignificant to the total object we group sections of the cell together in an "average" particle. Taking an extreme view and pretending the entire RBC is just a single particle has a lot of negative consequences which include a complete loss of structure. Thus we need to cre- ate a model which is coarse-grained enough to still maintain the asymmetry of the RBC but not to over simplify the system. Three models of red blood cells are tested within this thesis.

2.1 Models

The initial model has a center sphere connected to six adjacent spheres creating a rudimentary disk. This represents moving from a single sphere and towards modeling a red blood cell. This model will be called the 7-sphere model (Fig. 2.1) and it will also be the basis for the following models. Each sphere has a diameter of σ, where the magnitude of σ is equal to 5.0. As a result the total diameter of the RBC is 3σ and the thickness of the cell is 1σ. The aspect ratio is the ratio of the diameter to the thickness which for the 7-sphere model is equal

24 to 3. This aspect ratio is consistent with the average dimensions of a erythrocyte being ≈ 8.0mm in diameter and 2.0mm in thickness.

Figure 2.1: Structure of the 7-sphere model. The blue skeleton represents the attractive FENE bonds between the spheres while the red shells represent the repulsive WCA potential.

Each sphere is also subject to the WCA potential to prevent them from over- lapping. All neighbors in the 7-sphere model are given FENE bonds between them which generates the overall skeleton of the model(Fig. 2.1 blue). As dis- cussed previously there are two more types of bonds are used. Angle bonds are added between opposing spheres and dihedral bonds are placed around the exterior ring (Fig. 2.2 a,b). These bonds are used to ensure that the RBC has enough rigidity to maintain its disk-like shape while still allowing it to flex and fluctuate thermally. The result is the simplest possible model that has a disk-like shape. In the next model the 7-sphere model is extended upon and one additional ring of particles is added. This second ring is comprised of 12 spheres bringing the total number of spheres to 19 (Fig. 2.3). The 19-sphere model still has all nearest neighbors bonded with a FENE interaction. Both angle and dihedral bonds are used to ensure that the 19-sphere RBC is relatively rigid. The largest difference is in the aspect ratio. The diameter of the 19-sphere model extends to 5σ therefore, the aspect ratio is increased to 5. This model is also referred to as

Page 25 of 89 Figure 2.2: Bonds within the 7-sphere model. The small spheres represent the locations of the center of each particle similar to Fig.2.1. A: The angle bond is placed between opposing spheres crossing the center. The three angle bonds are placed: j,i,m - k,i,n - l,i,o. B: Dihedral bonds are consecutively overlapped around outside ring. The first two dihedral bonds in the exterior ring are placed at locations j,k,l & k,l,m. Not pictured are the remaining rotations around the exterior ring, a total of 6 dihedral bonds are used. the high aspect ratio model. This difference in aspect ratio will become relevant during the discussion of the maximal size of rouleaux in the results section.

Figure 2.3: Structure of the 19-sphere model. The skeleton represents the attrac- tive FENE bonds between the spheres while the red shells represent the repul- sive WCA potential.

Both the 7-sphere and 19-sphere model have non-overlapping components creating grooves between the spheres. A physical red blood cell does not have grooves along the surface and instead it is smooth. To improve upon the pre- vious iterations the third model uses overlapping to spheres address this issue.

Page 26 of 89 1 Additional spheres between the center and the outside ring placed at 2 σ create a smoother surface. This model is referred to as the smooth cell model (2.4). The smooth cell model still has a diameter of only 3σ meaning the aspect ratio is equal to the 7-sphere model.

Figure 2.4: Left: Surface of the model, comprised of 19 spheres: 1 central sphere, 6 for the interior ring, and 12 spheres on the exterior ring. Right: Skeleton of the model. FENE bonds between the spheres are represented by the blue lines. These bonds act as the skeleton of the cell maintaining its rigidity and preventing the spheres from separating.

Angle bonds are placed between opposing spheres in each ring through the center. And dihedral bonds are placed around the external ring. Because the in- ner ring is bound to both the central monomer and the neighboring exterior ring dihedral bonds are not placed on the inner ring. This is due to the surrounding neighbors already being constrained to a disk shape prevent the inner ring from fluctuating. In the discussion of the smoothness of the model it may be tempting to con- tinuously add on more spheres. Increasing the number of spheres in the model provides more a "realistic" (although still coarse grained) cell while simultane- ously negatively impacting performance of the simulations. The idea is not to simulate perfect models but instead show that the behaviour holds for all mod- els. The different models will be used to ensure that the results are robust as you move into more complicated models or different environments. These three models test both how the "smoothness" of the cell and the aspect ratio affects the results.

Page 27 of 89 Figure 2.5: Typical structures found in aggregates. Above: Simulation snapshots. Below: DIC microscopy images of erythrocytes. A: Two RBCs in a face to face configuration. B: The fourth cell which sits perpendicular to the three cells in a stack is defined to be in a face to side configuration. C: Multiple cells aligned into a single rouleau, all cells exist in a face to face configuration. D: Multiple cells in both face to face and also in face to side configurations this is referred to as an amorphous structure.

In the simulations 4 types of structures are observed (Fig 2.5). These four structures will be referenced throughout the results and the following sections. These structures do not represent all possible structures but instead are struc- tures that consistently appear at the scale of our simulations. The first structure is a doublet of two RBCs in the rouleau formation; this doublet is referred to as the face-to-face configuration (2.5 a). The second structure is a single cell which has aggregated perpendicularly on a rouleau which will be called the face-to- side configuration (2.5 b). Both of these smaller structures make up a subset of larger clusters. The face- to-face configuration generates rouleau while the face-to-side configuration is very commonly found in the formation of disordered clusters (2.5 c, d). Amor- phous, or disordered clusters themselves are not created from large globs of red blood cells as the word amorphous might suggest. The clusters are best de- scribed as an aggregate of many small interacting rouleau. Each cluster might

Page 28 of 89 have smaller rouleau on the order from 3 to 5 cells interacting in different con- figurations. To be clear in describing structures this thesis will reserve the usage of "rouleau" to refer to only the singular column of red blood cells. And the term amorphous to describe clusters that are not a single rouleau. Amorphous structures have varied morphologies which can be difficult to describe. There are three more specific assemblages which can be used to iden- tify parts of an amorphous cluster. Parallel, bisecting and the perpendicular (face-to-side) interactions (2.6) can all be observed. An amorphous cluster may have combinations of these types of interactions which comprise the overall structure.

Figure 2.6: Common rouleau-rouleau interactions. Above: Simulation snap- shots. Below: DIC microscopy images of erythrocytes. A: Two rouleau laying side-to-side in a parallel configuration. B: Perpendicular rouleau in bi-secting contact. C: Rouleaux interacting in a face-to-side configuration.

These different types interactions will be referenced later in the results sec- tion.

Page 29 of 89 2.2 Simulation and simulation parameters

There are three different types of simulations performed in this thesis.

2.2.1 Aggregation of red blood cells

The first simulation explores how red blood cells initialized in a disperse volume aggregate together under various static volume fractions of depletants. The cells are built in a grid-like pattern and then equilibrated with 2·106 integration steps. This ensures that the cells are no longer correlated to their initial configurations. Following this the depletant interaction is activated and the cells are again al- lowed to equilibrate for 5 · 107 integration steps. Cells are allowed to continue moving while data is periodically taken for 2 · 107 integration steps.

2.2.2 Rouleau dispersion

The second type of simulation examines the stability of a pre-existing rouleau (2.5 c). The simulations begin with cells aligned into a single coloumn with the depletant interaction turned on. During a warm-up phase the thermal energy is slowly increased from kT = 0 to 1 over 104 integration steps. kT is a measure of the energy in our simulations and the energy in a simulation is reported in units of kT . Because the potentials also scale with kT the equilibrium states do not rely on the value of kT itself. We choose kT = 1.0 as the energy due to convenience. This allows the column of rouleau to fluctuate into the lowest energy state before any data is collected. After the warm-up is completed the simulation continues with data being taken at a constant temperature of kT = 1 for 2 · 107 integration steps.

Page 30 of 89 2.2.3 Interaction stability

The final type of simulations provides data on the interaction energy of spe- cific contact configurations and their lifetimes. In these simulations two cells are initialized in face-to-face contact (2.5 a), or a triplet rouleau and flanking fourth RBC with its face on the side of the column (2.5 b). The warm-up phase is similar to dispersed rouleau where the initial structure is subject to a warm-up phase for 104 integration steps. Data is collected for the entirety of the simula- tion. These starting conditions illustrate the stability and lifetime of particular aggregate structures.

2.2.4 Parameters

In order to setup and accomplish this process HOOMD-Blue is used [50, 51]. HOOMD is a general purpose simulation package which is used for particle sim- ulations. The simulation set-up, parameters and coefficients are defined within a python environment. HOOMD itself then generates the simulation environment and takes advantage of algorithms written in languages such as C to accomplish repeated computational tasks. The simulation package also takes advantage of additional computational hardware to accelerate large scale tasks, such as inte- grating the velocity Verlet algorithm (1.3.9), or updating the neighbor list of ad- jacent particles. A full list of the simulation parameters tabulated in Table 2.2.1 and a small discussion on how we take advantage of the graphical processing units (GPUs) is mentioned in AppendixC The depletant interaction is not built into HOOMD. Thus as such a custom potential and force is added via a force table. HOOMD computes the force val- ues along a grid during initialization and then refers to the values when adding the force before integrating. Because the force table is a discrete grid, interpo- lation is used for any distances which are between any two points. This means that if a point is between two known values the average is evaluated. The force

Page 31 of 89 Table 2.2.1: Simulation Parameters in HOOMD

Parameter Coefficients Sigma σ = 5 1 Radius of colloid Rc = 2 σ 1 Radius of depletant Rd = 50 σ Energy kT = 1.0 1 Integration Step dt = 100 Friction ζ = 1.0 FENE k FENE = 30.0 r0 = 1.5 shift = 4.0 ◦ Angle k angle = 45.0 θ0 = 180.0 Dihedral k Dihedral = 45.0 d = −1.0 n = 3.0 1 WCA  = 1.0 σ = 1.0 rcut-off = 2 6 , shift = 4.0 3 3 3 Box Dimensions V10 Cell = (20.0σ) VHigh density = (20.0σ) VEqual Density = (25.2σ) of the depletant potential is computed as the negative gradient of the potential. The depletant potential only depends on the separation distance r

!  R 3 3 r 1  r 3 Vdepletant = −1.0 1.0 − + φd Rd 4 R 16 R

!  R 3 3 1 3  r2  Fdepletant = 1.0 0.0 − + 3 φd Rd 4 R 16 R

The grid is calculated from 0.01 to 2R−0.01 where R = (Rc +Rd). The grid is incremented in spacings of 0.01. The cutoff at 2R − 0.01 is because the depletant interaction is manually set to 0 at r = 2R. In order to acquire strong statistics about the variance and error about each aggregate structure every simulation is repeated 100 times. In a case such as aggregation, simulation data is taken every 100 integration steps. With 2 · 105 integration steps this results in 2000 snapshots of the final structure over time. Thus a single observable for one volume fraction is averaged over 200,000 snap- shots of data when considering 100 total systems. This process is repeated for different volume fractions. In HOOMD a simulation timestep does not represent a 1 to 1 correspondence to a unit of such as a second. There are two ways time can be calculated or related to the simulation. The first is to use the units provided by HOOMD to acquire

Page 32 of 89 how long each timestep is in terms of time, this is accomplished in appendixA. The second method is to relate the blood cell’s diffusion to its own geometry and is discussed in appendixB.

2.2.5 Observables

When discussing the collection of data we have devised three metrics which are used to describe our system. The first metric is the number of clusters. The num- ber of clusters is high for systems that have not aggregated and low for systems which have aggregated into a single cluster. The criteria for two cells being in the same cluster is based on the sphere to sphere distance of the particles that make up each cell. Because σ represents the overall diameter of the spheres that comprise the cell it is also a good choice to determine if two cells are touching. Any two spheres that have a separation less than (σ+0.05σ) are considered close enough to be touching. The small buffer zone (+0.05σ) allows small fluctuations to occur without causing large variations in the data. This also gives the added benefit that if a cell does separate it must diffuse at least (+0.05σ) away before being considered on its own. The number of clusters is a metric which gives information on clustered cells. However, it does not tell us anything about the cluster itself. To gain further in- sight into the exact shape, the number of aligned cells is calculated. The condi- tions for being aligned are more strict. The center to center distance of each cell must also pass an additional distance criteria. Any two cells that are in a face √ to face configuration must have a center to center separation of less than 2σ. One might consider comparing the normal of the face of each cell to determine if they are aligned, and indeed that criteria does work. However, the normal doesn’t consider translational separation of the disks. Two disks with aligned normals can still be offset. This means that an additional calculation of compar- ing the center to center distance must be done in order to ensure the cells are aligned on top of each other. Thus checking the normals of each cell is a redun-

Page 33 of 89 dant calculation. The only possible configuration two disks can have when the centers are in contact is to be aligned. Furthermore, in a situation where our simulation reports multiple unique clusters the largest number of aligned is re- ported. For example if two clusters or a single cluster exists with 7 and 3 aligned cells the metric will report 7 aligned cells. The final metric that is tracked is the size of the largest cluster. This is im- portant during the transition between many independent clusters and a single cluster. The maximum size can help identify if a system of 2 clusters is com- prised of two equal sized clusters or a single dominant cluster with a lone cell still diffusing in the environment. Because simulations are repeated many times, the statistical average of these quantities is reported in the results section. In short the metrics are: • The average number of clusters. • The average largest number of aligned cells. • The average size of the largest cluster. These metrics are calculated by generating an adjacency graph for the cells in the simulation. An adjacency graph is a representation of nodes in a network that are connected together. This is accomplished by creating a matrix which contains information on the relation between cells in the simulation. Once the adjacency graph is created the number of connected components in that graph directly calculates how many cells are in a cluster. As an example, if a simulation consisted of three separated cells the connected components would be 3, this represents 3 independent clusters. If all three cells were in contact with each other the connected components would be 1. The python package NetworkX is used to quickly compute these quantities [52]. To calculate the number of connected components an adjacency matrix is computed. The adjacency matrix has the dimensions (Number of cells)2 where each ij component represents the contact of two cells. When two cells are touching their respective ij elements are set to 1.0, otherwise their elements are set to 0.0. The quantity M = (A +

Page 34 of 89 I)n where A is the adjacency matrix and I is the identity matrix determines the number of connected components. M is only positive when the exponent n is less than the number of connected components in the matrix. With only 10 and 20 cell systems the number of connected components can be found quickly by starting at n = 0 and incrementing by 1 to n = 10 or n = 20. As n is incremented the value of M will decrease. When M is returned as zero or a negative number the number of connected components is n − 1 (the last value which returned a positive output). For computational sake there are some small efficiency gains that can be con- sidered. There are certain situations where the orientation or separation of par- ticles flags a known result. For example rather than calculating the distance between every sphere in two RBCs the first calculation is the center to center distance. If the center to center distance is greater than 4σ the cells cannot be touching in any possible orientation. Therefore, there is no reason to continue checking the rest of the particles within the model RBC. If the center to center distance is within (σ + 0.05σ) then not only are the cells touching they are also aligned. It would be impossible to orient the centers so close to each other in any way other than in a face to face configuration. In this case the two cells are known to be touching and aligned. If the result is between the two criteria (greater than σ + 0.05σ & less than 4σ ) then the distance and alignment criteria are checked for each of the corresponding particles in a cell. Furthermore, at any point in time if the two RBCs are resolved as touching then the remaining exterior spheres are not checked as that becomes a redundant calculation. These checks serve to increase the throughput of the simulations.

Page 35 of 89 Chapter 3

Results & Discussion

In order of appearance the results from the 7 cell model, the high aspect ratio model and the smooth cell model are presented. The models, set-up and pa- rameters of the simulation are described in detail within section2. Within each model the results of the aggregation and separation simulations are presented first. The results from the initialized cell configurations are presented second and the final results pertain to changes in density and system size.

3.1 7 sphere cell

The first model tested is the 7 sphere model as discussed in section 2.1. The av- erages of the number of clusters, maximum cluster size and number of aligned are plotted as a function of depletant volume fraction in Fig. 3.1. At low volume fractions φ = [0.00, 0.08] there is no observed aggregation of the cells. The cells are dispersed and each "cluster" is a single cell thus the number of clusters is the number of cells, 10. This is due to the thermal energy being greater than the attractive potential. Cells cannot aggregate into stable clusters without sep- arating shortly thereafter from the random force kicking them out of any paired configurations. Shortly after φ = 0.08 there is a sudden transition downwards in the number of clusters. Cells are now able to form clusters which are stable for long periods of time. Beyond this transition the number of clusters remains

36 at ≈ 1.0 This transition happens very quickly over the scale of volume fractions tested. Transition to a single cluster marks the point where the depletion forces dominate over thermal forces. It should be noted that while the transition occurs quickly over φ the system is required to be equilibrated for a sufficient period. The closer the system is to the transition the longer it takes to equilibrate into a single cluster.

10

8

6

4

2 1 0.06 0.08 0.1 0.12 0.14 0.16 0.18

Figure 3.1: 3 quantities are tracked versus the volume fraction φ. Red circle: The number of clusters represents unique groups within the simulation, collapse into a single cluster occurs at φ ≥ 0.08. Error bars represent the standard devia- tion. Blue hollow circles: The size of the largest cluster. Purple dots: Mean number of aligned cells. The standard error is represented by the size of the data point.

At the same time as the collapse, the number of aligned cells and the maxi- mum size of the cluster increases. At its peak the number of aligned cells reaches about 5 cells with a variance between 4 to 7. Increasing the volume fraction φ changes the morphology in the structure. The number of aligned cells drops to about 3 cells and at very high volume fractions is reduced further. The peak in the number of aligned cells shows the possibility that not only can the depletant interaction cause aggregation, but also the formation of rouleau. However, the

Page 37 of 89 peak does not reach the maximum number of cells. Furthermore, the alignment of cells is non-monotonic in φ. Increasing φ initially has no effect on the num- ber of aligned cells. However, the number of aligned cells peaks immediately after the collapse of RBCs into a single cluster. As φ is increased the number of aligned cells decreases until it plateaus near 2-3 aligned cells. Viewing what a typical structure looks like near the maximum number of aligned shows interac- tions between pairs of rouleau (Fig. 3.2). The rouleau does not form in a straight line but instead a tilted stack. Furthermore the non overlapping spheres allow for other cells to aggregate into the corrugations created by the tilted column. As a consequence a cell which diffuses onto the side of a rouleau may find itself in a stable configuration. The single cell on the side provides a new face that another cell can come into contact with and aggregate onto. This cascades as rouleau begin to form branches off the original structure.

Figure 3.2: Simulation snapshot of the 7 sphere model at φ = 0.085, each cell is provided with its own colour to help identify the structure. (a) Image showing 6 vertically aligned cells on the left of the structure with a cluster of 3 cells sit- uated on the side in a bisecting position. (b) The snapshot has been rotated by 90◦showing the corrugated positions of the cells along the side.

This model provides a great starting point. The depletant interaction is caus- ing cells to cluster together and directly after the collapse [φ ≈ 0.085] a peak in alignment is observed. However the non overlapping spheres give rise to offset

Page 38 of 89 positioning. Furthermore, the corrugations maximize the well depth of aggre- gating cells which allow long lived stable aggregated clusters. When four cells are aligned there is 3 grooves along the side of the structure for a cell to aggregate onto (Fig. 3.2 a). These 3 grooves are where flanking RBCs are commonly found in stable positions. The alignment of 3-4 cells corresponds with the plateau seen in Fig. 3.1 and is also related to the aspect ratio of the cell (3). To test if the as- pect ratio is the limiting factor of both the peak in alignment and the cause of the plateau of aligned cells, the same simulation is repeated with a higher aspect ratio model.

3.2 High aspect ratio cell

The high aspect ratio model increases the diameter of the cell to 5σ while the thickness remains the same. The previous results suggested that the smaller aspect ratio (3σ) prevented long rouleaux to form. As seen in Fig. 3.3 the average number of clusters behaves very similar to the 7-sphere model. At low volume fractions the cells do not aggregate. There is a transition from free cells to a singular cluster which also corresponds to a peak in alignment. Cell collapse into a cluster occurs slightly earlier than the previous model; differences in the RBC model are responsible for the shift. As the volume fraction is increased the number of aligned cells decays until it reaches a minimum around 3 cells (Fig. 3.3). Compared to the 7 sphere model the increased aspect ratio changes the be- haviour in the number of aligned cells. In particular the maximum number of aligned in the peak is much higher with 8 cells aligned rather than 4-5. Further- more, while the 7-sphere model quickly plateaued around 3 aligned cells the high aspect ratio model smoothly decays towards 3-4 aligned cells. In the region of peak alignment single columns of rouleau can form although it is not very often and this is represented by the standard error barely reaching above 9 aligned cells (Fig. 3.3 purple). A number of aligned cells near 10 encour-

Page 39 of 89 10 Number of Max Cluster size Clusters Aligned Cells 8

6

Quantity 4

2

0.06 0.07 0.08 0.09 0.1 0.11 0.12 0.13 0.14 0.15 Volume fraction - φ

Figure 3.3: 3 quantities are tracked versus the increase in volume fraction. Red diamond: The number of clusters represents unique groups within the simula- tion, collapse into a single cluster occurs at φ = 0.075. Blue circles: The size of the largest cluster. Purple dots: Mean number of aligned cells. Outer error bars represent standard deviation while the inner error bars represent the standard error in the measurement. aged simulations of larger systems to explore the maximum number of aligned cells. This was accomplished by increasing the number of cells to 20 while also increasing the system volume to (25.2σ)3 to maintain equal density to the 10 cell system. Simulating a system with 20 cells at equal density also shows interesting be- haviour, the results are plotted in figure 3.4. The peak number of aligned cells averages around 16. Similar to the 10 cell system the RBCs see a peak in alignment but do not see full alignment at any one particular volume fraction. Visualizing a typical structure found near the peak reveals that many of the clusters are formed of one main stack and another rouleau interacting in either a parallel, bi-secting or perpendicular configuration (Fig. 3.5). As the rouleau begin to form other cells are likely to interact along the sides of the column. The cells which are not aligned provide a face or branch which other cells may aggregate onto; sim- ilar to the to the 7 sphere model. It is still within possible to find completely

Page 40 of 89 20 Number of Max Cluster size Clusters Aligned Cells 15

10 Quantity

5

0.06 0.07 0.08 0.09 0.1 0.11 0.12 0.13 0.14 0.15 Volume fraction - φ

Figure 3.4: 20 RBCs at equal density to 10 cells. Collapse occurs at the same volume fraction with significant peak in number of aligned. Aligned cells decay to 4 cells with variation around 3 to 5. aligned rouleau within a set of simulations, although it does require many at- tempts (Fig. 3.6). Furthermore, the corrugated surface of the RBCs results in the off-set stacking within any single rouleau but also provides grooves in which other cells or rouleau can fit into. This issue is the exact same as the 7 sphere model. Hence, while increasing the number of cells does allow longer rouleau to form, the same mechanisms that prevent the 10 cell system from consistently forming a single rouleau also exist in the 20 cell case. Increasing the aspect ratio directly increases the length that a rouleau can be before the face-to-side and rouleau-rouleau interactions start to interfere in the formation of a single rouleau. However, the corrugated surface in both the 7- sphere model and the high aspect ratio limit the consistency of rouleau length. Longer rouleau provide many small grooves along the side in which a cell can situate itself into. Increasing the aspect ratio causes the face-to-face interaction to become stronger. The corrugations also enhance the stability of the face-to- side interaction albeit less than the face-to-face. An increased aspect ratio could allow simulations to result in single column rouleau in a 10 or 20 cell system. In- creasing the aspect ratio beyond 5 is no longer relevant to modeling a red blood

Page 41 of 89 Figure 3.5: The high aspect ratio has preferential face to face stacking of the RBCs. Long columns of rouleau are formed just after the initial collapse. How- ever, due to the non overlapping spheres the cells offset to situate within the corrugations in the model. This offset also allows cells to aggregate on the side of a forming rouleau. In this picture a rouleau of 15 cells (vertical) and a per- pendicular rouleau of 5 cells (right, horizontal) interact. cell. It should be noted that cylindrical structures can also arise in liquid crys- tals [53]. In systems of nano-discs [38, 54, 55], disk-like-micelles [56, 57] or clay platelets [58] can stack into each other in columnar structures. The self assembly of discotic liquid crystals [59, 60, 61] shows that increasing the aspect ratio will not change the resulting behaviour.

3.3 Smooth cell

3.3.1 Aggregation

The final model presented returns to the 3:1 aspect ratio while diminishing the corrugated surface of the cell as discussed in section. 2.1. This will allow us to test the effect of the surface independent of the change in aspect ratio. In general at low volume fractions, no aggregation is observed. There is a transition to a

Page 42 of 89 Figure 3.6: 19 cells aligned in rouleau configuration. The corrugated surface causes tilted aggregation in the rouleau. The high aspect ratio increases the sta- bility of the face-to-face interaction. Rouleau must form larger columns (≈ 5 cells) before interactions on the side are stable for long times. single cluster which marks the point where the depletion forces are greater than the thermal fluctuations in the system. Corresponding with this transition is an increase in the number of aligned cells and maximum cluster size. This point

∗ ∗ will be referred to as φff where φff = 0.07. This marks the point at which face-to- face aggregation is observed. Increasing the volume fraction continues to yield a single cluster with different morphologies (Fig. 3.7). Compared to the 7-sphere model and the high aspect models there is to- tal alignment in the rouleau past the collapse transition. Along 3 points (φ = [0.072, 0.074, 0.076]) all 10 cells consistently align into a single column of rouleau. Beyond φ = 0.078 the number of aligned cells begins to slowly decay. It is not until φ ≈ 0.0875 that the deviation around the mean drops below 10 cells. This value is notable as interactions like the face-to-side start to cause a drop in the number of aligned cells. These configurations are long-lived are indicated to be-

∗ ∗ ∗ gin at φfs = 0.0875. Both φff = 0.07 & φfs = 0.0875 are indicated in Figure 3.7 as

Page 43 of 89 Figure 3.7: 10 cell system. At low volume fractions (φ), the attraction from the de- pletant interaction is less than the thermal fluctuations. As such the cells do not aggregate. At intermediate volume fractions the attraction of the cells is greater then the thermal fluctuations causing collapse in the number of clusters. The number of aligned quickly rises to 10. When all 10 cells are aligned the RBCs are in a single rouleau. At high volume fractions the number of aligned cells drops as amorphous structures are observed. Inset are images of both a rouleau and an amorphous structure. Two dotted vertical lines correspond to the tran- sitions where the system collapses into rouleau, and then decays to amorphous structures as more meta-stable states become long lived. vertical black dotted lines. Visualizing an example of rouleau found in the smooth cell model shows that the overlapping spheres do not allow large grooves that diffusing cells can wedge themselves into. Furthermore, the staggered stacking of face-to-face cells is greatly reduced which is closer to the actual alignment of RBCs in rouleau (Fig. 3.8). Around φ = 0.09 all 10 cells in a single rouleau is a relatively rare occurrence; a value of 10 is barely within a standard deviation. Instead, there is continual decay in the number of aligned cells until at φ = 0.15 less than 4 cells are aligned on average. At these large volume fractions, the depletant forces are very strong and thus cells tend to stick together in whatever configuration they first contact each other. Morphologies such as the inset in Fig.3.7 or Fig.3.9 are commonly found. The smooth cell model retains the preferential face-to-face aggregation

Page 44 of 89 Figure 3.8: A single column of rouleau. The smooth cell model retains the same aspect ratio as the 7-sphere model (2.1) but has high preferential stacking similar to the 19-sphere model (3.7). By using overlapping spheres the offset which was noticeable in both previous models is greatly reduced. This rouleau is typical of the clusters found between φ = 0.071 and φ = 0.09. similar to the high aspect ratio but also maintains the correct aspect ratio like the 7 sphere model. The rest of the analysis is accomplished using the smoothcell model as it presents the best aspects of both the 7 sphere and high aspect ratio models.

Figure 3.9: At high volume fractions small stacks of rouleaux aggregate together. The entire structure is comprised of small groups of rouleaux which have found stable positions against each other, forming a amorphous aggregate of rouleaux.

When cells align in a face-to-face configuration they maximize the overlap of the excluded volume. The largest depletant interaction occurs when two cells are in a face-to-face configuration. Thus in any aggregate the face-to-face con- figuration should be the most stable configuration. This raises the question “Are

Page 45 of 89 amorphous shapes more stable at high volume fractions than rouleau?”.

3.3.2 Rouleau dispersion

The aggregation simulations show that at high volume fractions the cells find themselves in amorphous clusters. Rouleau dispersion tests at which volume fractions a single rouleau will break apart and return to a state of separate cells. Cells are initialized in a column of rouleau at the very beginning of the simu- lation. (2.5(c),Fig. 3.8). Because the face-to-face interaction is predicted to be greater than the face-to-side it should correspond to the global minimum in en- ergy. As expected at high volume fractions the RBCs remain aligned in a single rouleau until the volume fraction is sufficiently low as to allow cells to diffuse

∗ (Fig. 3.10). Once the volume fraction is less than φff rouleau disperses suddenly into separate clusters and remain independent for the rest of the simulation time. The distribution of the collapse is narrower as it is unlikely that cells escape the

∗ global minimum above φff. Stating the answer to the posed question, amorphous shapes are not more stable instead they are trapped meta-stable states that are long lived. At sufficient volume fractions cells can aggregate in unfavorable lo- cations but remain stuck for very long times. Comparing the aggregation of red blood cells (Fig. 3.7) to the dispersion (Fig. 3.10) the aggregate-dispersion transition occurs at slightly different volume fractions, suggesting hysteresis. The dispersion of RBCs exists at a slightly re- duced volume fraction; aggregation seems to occur at φ = 0.07 while cells dis- perse around φ = 0.067 (Fig. 3.11). Cells themselves do not experience hysteresis due to memory or any intrinsic property. Instead hysteresis occurs due to the history of the state. That is to say, cells that are separated take time to aggregate together and cells that are aligned take time to disperse. Increasing the simula- tion time would reduce the coercivity from ≈ 0.003 to a lower value and align the transitions. However, rouleau formation is a time dependent process in the body which means that it should not be unreasonable to expect some hysteresis

Page 46 of 89 Figure 3.10: Starting in a single rouleau with all 10 cells aligned. At low vol- ume fractions the cluster falls apart into 10 separate cells (Number of clusters equals 10). At an intermediate volume fraction the attraction from the depletant interaction is less then thermal fluctuations and cells separate over a small win- dow of volume fractions. Higher volume fractions increase the attractive force between the cells causing the structure to be stable for very long time periods. Unlike systems where cells aggregate together, at high volume fractions if the initial configuration is rouleau then the structure remains in that conformation. in both simulation and experiment.

3.3.3 Interaction Stability

Studies which include large numbers of discotic particles to explore liquid crys- tals described how a dense system of RBCs would act in a confined space [62, 63, 59, 64]. Furthermore, high density blood flow has been explored using prin- cipals from discotic liquid crystals [65, 66]. However, rather than increasing the number of red blood cells to try and mimic the large scale systems it is valuable to understand why the transition to rouleau occurs. To explore the transition to a cluster of RBCs simulations of only particular configurations are employed. A column of rouleau should form when the depth of the energy well in the face-to-face configuration is greater than the thermal fluctuations (Fig. 2.5,a). Similarly one should expect to see the transition from rouleaux into more disordered aggregates when the well-depth of face-to-side is larger than thermal energies (Fig. 2.5,b). The first situation is the transition

Page 47 of 89 Figure 3.11: . Hysteresis in the aggregate-dispersion transition. Symbols and colours corre- spond to Figs. 3.7 and 3.10. (a) The time averaged maximum number of aligned RBCs near the aggregation point for aggregation (filled) and rouleau dispersion (open). The aggregation path decays at greater volume fractions than shown here. The coercivity is approximately 0.003. The dashed vertical line demarcates ∗ φff = 0.07, the volume fraction for which face-to-face contacts are long lived. (b) The same as panel (a) but for the number of clusters. The two curves are mirror images of each other in the vicinity of the dispersion-aggregate transition point. from free cells to a cluster, two cells are placed face-to-face and released at a given depletant fraction. Life-time of the doublet conformation is tracked over the course of the simulation. Cells are considered separated when they are no longer part of the cluster, because our system is only testing 2 cells this is simply when the number of clusters transitions from 1 to 2 (3.12). At very low volume fractions the cells separate immediately within just a

Page 48 of 89 Figure 3.12: The average value of the lifetime of a structure of just two cells in a face to face configuration. At low volume fractions the cells immediately sepa- rate. The lifetime rises with the volume fraction and reaches the midpoint of the total simulation time around φ = 0.07, in the aggregation simulations. Beyond this point the cells are likely to survive as a pair for more then half our simula- tion time. At higher volume fractions the cells are stable in that configuration. The standard deviation is plotted as the outer error bars and the error is plotted as the inner error bars. The dotted line at the bottom represents the mean time to escape at a volume fraction of φ = 0.0. few timesteps. Increasing the volume fraction increases the average lifetime of the doublet, at first the increase is small. However after φ = 0.065 the lifetime rapidly increases to the maximum time limit corresponding to cells that never separated. Interestingly the point φ = 0.07 corresponds to when cells reach half the total simulation time. This point marks the same location that collapse was initially observed in the aggregation plots (Fig. 3.7). Originally the critical

∗ volume fraction of face-to-face interaction was determined to be φff = 0.07. Now ∗ ∗ φff := 0.07, where φff defines this interaction to be at the particular value. Beyond this volume fraction cells are likely to stay in a collapsed state for at least half the simulation. By φ = 0.075 all simulations reach the end without any separation, indicated by the vast reduction in variance. The critical volume fraction for the face-to-face interaction is reasonably close to physiological systems. In vitro experiments of red blood cell aggregation uti- lized a volume fraction of φ = 7.8% using various macromolecules [25]. Whole

Page 49 of 89 blood is approximately 55% plasma and only 10% of plasma is protein, the re- maining 90% is water. These proteins are comprised of serum albumin (∼ 55%), fibrinogen (∼ 5%) and globulins (∼ 40%). Serum albumin is too small to induce aggregation [25] leaving only fibrinogen and globulins as the principal compo- nents that could act as depletants. Using the molecular mass and hydrodynamic radius of both fibrinogen (340kDa, and a hydrodynamic radius of ∼ 11nm) and a globulin such as the α-2-macroglobulin (725kDa, ∼ 9nm [25]) results in a vol- ume fraction of ∼ 7.5%. In the context of the simulations the size ratio of the

1 depletant to σ is 25 while the size ratio of fibrinogen to roughly a third of the 1 diameter of an RBC (≈ 2µm) is 180 . The size ratio of albumin would be roughly 1 500 .

0

-10

-20

-30

-40 Total Energy Energy Total -50

-60 * φff -70 Cells are unstable Cells are stable -80 0.06 0.065 0.07 0.075 0.08 0.085 Volume fraction - φ

Figure 3.13: The Total energy in the face to face configuration of two cells. This is the sum of both the repulsive LJ interaction and the attractive depletant interac- ∗ tion. The critical volume fraction φff = 0.07 is indicated with the dotted vertical line. Volume fractions lower then that correspond with systems that tend to be unstable, and volume fractions higher tend to be stable.

The total interaction energy is also recorded in the simulations (Fig. 3.13).

∗ The interaction energy increases linearly. Increasing φ beyond the φff drastically decreases the variance around the mean which indicates there are less fluctua-

Page 50 of 89 tions in the data. This occurs because the energy calculated at each time step is similar to the previous step. Indicating that the doublet is more stable, resulting

∗ in long lived face-to-face interactions. At φff the total energy is recorded as -37.8 kT, which indicates adhesion many times the thermal energy (kT = 1.0). This seems out of place and much higher than one might guess where collapse would occur. -0.6 Cells are unstable Cells are stable -0.8 * φff -1

-1.2 # participating -1.4 40

-1.6 30

20 -1.8 10 Energy per participating particle * φff -2 0 0.06 0.07 0.08 -2.2 0.06 0.065 0.07 0.075 0.08 0.085 Volume fraction - φ

Figure 3.14: The total energy in a cell normalized by the amount of participating particles. Each cell has 19 particles, thus the highest possible number of particles that can participate in an interaction is 38. The inset graph shows in red (trian- gles) the mean number of participations. The line which deviates upwards is the number of total contacts, as two cells are squished together at high volume fractions particles can increase the number of total contacts they have with the opposing cell. Near and below the critical phi value the number of contacts is 1 to 1 with the number of participating particles.

However, the model RBCs are not rigid disks, but rather composed of 19 in- dividual particles joined together by many bonds with many internal degrees of freedom. To account for this, the total number of particles that are participating in the depletion interactions is calculated (Fig. 3.14, inset). The average number of particles participating in the doublet contact increases with φ. Again, there

∗ is a sudden decrease in the variance, indicating a stable value. Beyond φff the

Page 51 of 89 number of participating particles saturates to 38 indicating that every particle is in contact. The energy normalized by the number of participating particles

∗ is around −1.1kT at φff . This energy agrees with the intuition that aggregation occurs when the depletion interactions are stronger than the thermal energy. As all three metrics share the same axis it can be useful to visualize them to- gether (Fig. 3.15). Plotting the original lifetime as a semi-log plot shows a change

∗ in behaviour past φff which is represented by the change in slope (Fig. 3.15, a). The total energy decreases nearly linearly with φ but the standard deviation sud-

∗ denly decreases at φff (Fig. 3.15, b). This also corresponds to when the normal- ized energy per interacting particle is greater than the thermal energy (Fig. 3.15, c). Previously the peak in aligned cells was qualitatively measured in figures 3.1 and 3.7. However, now the peak in alignment can be attributed to the stability

∗ of the face-to-face configuration. In particular below φff there are no configu-

∗ rations where two cells are stable for a significant period of time while after φff the only stable configuration is the face-to-face orientation. These calculations are repeated for the face-to-side configurations. Results for the time to separate are shown in Fig. 3.16. The results are similar to the face-to-face case, but the critical volume fraction is now at a larger value where

∗ φfs := 00.0875. This is the point at which the cells take on average half of the total

∗ simulation to separate, the same critera used to define φff . Note that the transi- tion region with large error bars is now broader indicating that the behaviour in the face-to-side case is more varied than the face-to-face case. The total energy is again calculated and is shown in Fig. 3.17.

∗ The total depletant energy is around -30 kT at φfs, similar to the face-to-face case. However, the calculation has much more variance than the face-to-face configuration. When cells are sticking together in a face-to-face configuration they are less mobile with respect to one another and thus little variation is ob- served. Conversely, a cell that is stuck to the side of a roll is free to rotate around an axis parallel to the axis of the roll. This type of rocking motion is observed

Page 52 of 89 Figure 3.15: The stability of face-to-face doublets. (a) The average lifetime of a doublet. At lower volume fractions than shown, the lifetime is constant. Near ∗ the dispersion transition, the lifetime grows exponentially until φff = 0.07 (ver- tical dashed line) at which point doublets are likely to survive for more than half the simulation time. This point marks the critical volume fraction beyond which face-to-face contacts are stable on the timescale of the simulations. (b) The to- tal interaction energy in simulation units of the face-to-face configuration. (c) The interaction energy normalized by the number of participating particles. (in- set) The mean number of particles participating in the interaction (red triangles) rises with volume fraction. Particles can have multiple pair interactions so the total number pair interactions (green line) differs at strong adhesion.

Page 53 of 89 Simulation Time 5000

4000

3000

2000

Time before cells separate 1000 Mean life time 0 0.07 0.075 0.08 0.085 0.09 0.095 0.1 Volume fraction - φ

Figure 3.16: The average lifetime of a structure of a cell positioned on the side of a rouleaux. At low volume fractions the cell diffuses away quickly, at φ = 0.0875 the critical volume fraction is labeled. The standard deviation is plotted as the outer error bars and the variance is plotted as the inner error bars. The lifetime saturates completely at φ = 0.1. in the simulations. As the cell rocks back-and-forth, the total depletant energy changes given larger standard deviation. This is is confirmed in the inset to Fig. 3.18 where the number of participating particles shows very large variations, even at high volume fractions. Note also that the number of participating par- ticles saturates around 20 indicating that at most ≈ 10 particles from the cell on the side can interact in this configuration. This is significantly less than the 19 of the face-to-face configuration. The main figure of Fig. 3.18 shows the energy normalized by the number of participating particles. This normalization again brings the values closer to

∗ the expected value of kT . However, the normalized energy at φfs is around -1.7

∗ kT which is deeper than the value of -1.1 kT found at φff . Recall that the fluc- tuations in the energy are much larger for face-to-side than face-to-face. Hence, although the average is deeper at -1.7 kT , particular configurations can fluctuate out of the potential well which could allow separation. That is, as the cell rocks back-and-forth, it can experience reduced depletant attraction and thus escape the stack. Each result of the face-to-side configuration is plotted on a shared φ axis in

Page 54 of 89 0

-10

-20

-30 Total Energy Total

-40

-50 φc,fs

Cells are unstable Cells are stable -60 0.08 0.085 0.09 0.095 0.1 Volume fraction - φ

Figure 3.17: The Total energy in the face to face configuration of two cells. This is the sum of both the repulsive LJ interaction and the attractive depletant in- ∗ teraction. The critical volume fraction φfs = 0.0875 is indicated with the dotted vertical line. Volume fractions lower then that correspond with systems that tend to be unstable, and volume fractions higher tend to be stable.

Fig. 3.19. A semi-log scale of the lifetime shows a distinct change in the be- haviour lifetime of the face-to-side configuration (Fig. 3.19, a). Nearing the sta-

∗ ∗ bility point φfs the rate of the lifetime is continuously increasing. Once past φfs the rate of increase in lifetime tapers off approaching the full simulation time. Similar to the key results shown in the face-to-face configuration the stability of the face-to-side gives a deeper understanding of the behaviour in the number of aligned cells shown in Fig 3.7. The first stable configuration as φ is increased is rouleau and the second stable configuration is the face-to-side interaction. Using

∗ ∗ φff and φfs we can try and indicate results for a different system. We should first

∗ observe collapse with only rouleau at φff , and then see a decrease in the number

∗ of aligned cells past φfs. To test the robustness of the results presented for the 10 cell system, simu- lations were also performed with 20 cells. Two 20 cells systems were designed. In the first, the volume of the simulation box is kept constant but the number of cells is increased from 10 to 20. Hence, these simulations are performed at a

Page 55 of 89 -1 Participating particles 25 -1.2 20

-1.4 15

10 -1.6 5 φc,fs -1.8 0 0.08 0.085 0.09 0.095 0.1 -2 φ

-2.2 φ Energy per participating particle c,fs -2.4 Cells are unstable Cells are stable -2.6 0.08 0.085 0.09 0.095 0.1 Volume fraction - φ

Figure 3.18: The interaction energy in Fig 3.17 normalized by the number of par- ticipating particles. (inset) The mean number of particles participating in the interaction (red triangles) rises with volume fraction. Particles can have multi- ple pair interactions so the total number pair interactions (green line) differs at strong adhesion. The total number of particle interactions saturates around 20. higher density. In the second both the number of cells and the volume of the system was increased to maintain the equal density to the 10 cell case. The col- lapse of cells into a cluster is examined in Fig. 3.20 where all three system setups are plotted. Comparing between the curves, differences are only apparent in the transi- tion region as all cells are free at low φ and all cells are in one cluster at large φ. There is consistency across all three curves in the sense that aggregation is observed to first occur at φ > 0.067 in all three cases. At φ = 0.068 the same degree of aggregation is observed for setups with the same box size, v = (20σ)3 (red & blue), while a smaller degree of aggregation is observed for the larger box, v = (25.2σ)3 (green). This result indicates that the average time for cells or clusters of cells to collide with one another may influence the degree of aggrega- tion in the transition region, which is to say that a larger box simply means that cells take a longer time to find each other. It is also possible that the equilibration

Page 56 of 89 Figure 3.19: A combination plot of (a) Fig. 3.16 plotted on a log scale Aligned ∗ with (b) Fig. 3.17 and c Fig. 3.18. The dashed vertical line now marks the φfs = 0.0875, above which it is probable that face-to-side aggregates endure for more than half the simulation time.

Page 57 of 89 time is sufficient for the smaller box but not so for the larger one.

Figure 3.20: The collapse of all curves begins near φ = 0.067, and all curves con- ∗ verge at φff = 0.07. This indicates that small changes in density or the number of cells does not effect the collapse point. Because the collapse depends on the energetics of the face-to-face interaction rather then system volume or number of cells in a disperse system.

Because two of the systems exist at 20 cells the graphs do not truly overlap each other. Another way to view the collapse is by normalizing the results based on the total number of cells (Fig. 3.21). When normalized the systems reinforce that they all start to collapse at the same point φ > 0.067. Interestingly while the systems of similar volume matched each other in Fig. 3.20 the systems of similar density follow more closely to each other in Fig 3.21. This seems to indicate that during the start of collapse the systems of similar density can find cells at nearly the same rate (higher density finds cells quicker and thus collapses faster). At the end of the collapse systems match each other at full collapse. The collapse trend continues beyond φ = 0.069 where the results are in good agreement with the number of clusters. At large enough φ values, all setups converge to a single cluster indicating that the equilibration time is long enough for complete aggregation. However, noting that the scale of Fig. 3.20 & Fig.3.21 only spans from φ = 0.067 to φ = 0.07 the deviations overall are small compared to previous results such as Fig. 3.7. In fact, the most important result is that the

∗ predicted φff works for all three cases since nearly complete collapse is observed

Page 58 of 89 1 N = 10, V = (20 )3 3 0.8 N = 20, V = (25.2 ) N = 20, V = (20 )3

0.6

0.4

Number of Clusters of Number 0.2 * φ ff

0 0.065 0.066 0.067 0.068 0.069 0.07 0.071 0.072 Volume fraction - φ

Figure 3.21: The normalized values of the 10 and 20 cell system, The same plot as Fig. 3.20. the value of 1 corresponds to no clusters in the simulation and a value of 0 represents a complete collapse. at this φ value in each curve. The number of aligned cells as a function of φ is shown in Fig. 3.22 for all three setups. For each case, the average number of aligned cells is very close to

∗ the total number of cells just after φff . That is, almost all the cells are lining up in a single roll. The data from the large box does not quite get to 20, but again this is likely an equilibration time issue. The number of aligned cells stays remains near the maximum for a larger range of φ for the 10 cell case: rolls of 10 are observed up to φ ≈ 0.09 while the 20 cell rolls are observed only until φ ≈ 0.08 for both box geometries. In both the 20 cell cases the formation of two rolls of significant length is observed. The larger number of cells provides additional configurations that two rouleau may interact in. Parallel or bisecting configurations (2.6 a,b) can form at these volume fractions which are stable for the simulation time. As the number of cells is increased and the number of possible configura- tions increases plotting the maximum number of aligned cells hides useful in- formation. Figure 3.23 shows the distributions of aligned cells at φ = 0.075 for figure 3.20 at the end of the simulations. Most notably the average number of

Page 59 of 89 Figure 3.22: Plot of the number of aligned 10 and two 20 cell systems versus the volume fraction. Systems collapse to the same values and at similar points. In this plot the number of aligned cells increases to a maximum at the same volume fractions and decays. All systems come to a full collapse at the same depletant fraction.

N = 20, V = (25.2 3 0.6 0.5 0.4

0.2

0 Probability of being in state of Probability 5 10 15 20 Magnitude of alignment

Figure 3.23: 20 RBCs at φ = 0.075 at a system volume of (25, 2σ)3. This corre- sponds to the green square in Fig. 3.22 at that φ value. The probability of being in an aligned group plotted against the magnitude of the number of aligned cells. Most of the cells (≈ 50%) are in an aligned cluster of length 20 while situ- ations where two smaller groups of can cluster. Note that at φ = 0.075 all cells are within the same cluster and the number of clusters is equal to 1 as shown in Fig. 3.20.

Page 60 of 89 aligned cells (≈ 17) is not the most common rouleau size. Instead the most com- mon rouleau size is of length 20 with a second distribution of rouleau around length 7 to 13. This corresponds with groups of interacting rouleaux that are in long lived meta-stable states. These shorter rouleaux are not to be confused with groups that have not yet found each other due to an equilibration issue. As shown previously in Fig.3.20 at φ = 0.075 the number of clusters is equal to 1. Beyond φ = 0.09, there is a steady decrease in the number of aligned cells. All systems continue to decay until they reach about 4-5 aligned cells. This value makes sense when considering the face-to-side interaction is well beyond its threshold allowing stable non-rouleau configurations. Futhermore, a short roll of cells provides enough surface area that either another cell, or another neigh- boring rouleau can interact and aggregate onto. At large volume fractions, these states are long lived yielding amorphous clusters which comprise many small rouleau trapped in deep meta stable wells. An example of an amorphous cluster is shown in Fig. 3.24.

Figure 3.24: 20 RBCs at φ = 0.10, the cells commonly cluster together into amor- phous conformations at these volume fractions. This particular structure has groups of 2,3,4 and the middle node with 5 aligned cells that have aggregated together. This aggregate contains parallel, bisecting and perpendicular (face-to- side) interactions. This structure would report a number of aligned cells of 5 which corresponds to the right hand side of 3.22.

Page 61 of 89 Chapter 4

Conclusion

When red blood cells aggregate into rouleau and other amorphous clusters their morphology appears to be very complex. This thesis studies the formation of ag- gregates due to a well-defined short-ranged depletion attraction using coarse- grained numerical simulations. Viewed in this manner the biophysical phe- nomenon of rouleau can be viewed as a consequence of entropic and geometrical interactions. Employing a simple model of an RBC paired with the AO potential and ignoring biochemically specific bridging mechanisms, our simulations re- produced much of the variety of clinically observed states, including dispersed cells, rouleaux and amorphous aggregations. At low volume fractions the cells remain free and able to diffuse. At a critical volume fraction the attractive forces are greater than the thermal fluctuations which cause a collapse into a single cluster. The morphology of the cluster depends on the specific volume fraction. There is a narrow window of volume fractions directly after the collapse where single columnar rouleau is consistently observed. Higher volume fractions re- sult in varied aggregate clusters which are trapped in long lived meta stable states. Rouleau formation is a consequence of the disk-like shape of the red blood cell. Understanding how geometry and shape affects cellular behaviour is an exciting research direction [67, 68], and this thesis has sought to bridge self-assembly in inanimate anisotropic colloids to . The three mod- els tested each demonstrate the effects of geometry on the resulting aggregate

62 morphologies. The initial 7 sphere model provides a small region of enhanced rouleau formation which is ultimately hindered by the corrugated surface. The high aspect ratio model provides a greater interaction strength in the rouleau configuration which increases the number of aligned cells after the collapse. However, the corrugated surface continues to allow stable configurations along the side of the rouleau which are long lived. While increasing the aspect ratio increases the number of aligned it begins to more closely represent the aggrega- tion of disks rather than red blood cells. Finally, the smooth cell model removes the corrugated surface while keeping a similar aspect ratio to the 7 sphere model. This model has a region of consistent single rouleau formation before other con- figurations become available at increased depletion interactions. Recent work in the field of sticky-spheres has shown the self assembly pro- cesses in colloids [69, 70, 71]. While the geometry of disks provides the colum- nar ordering found in liquid crystals and discotic crystals [72]. This thesis pro- vides a simplistic model that replicates the clustering processes and the struc- tural order found in more complex molecules and systems such as lyotropic columnar liquid cyrstals [73, 74]. Furthermore, the energetics between specific conformations independently validate the observed values of collapse in larger systems. The particular life- times and energetics of states are used to predict the transitions from a system of free red blood cells to either rouleau or amorphous structures. Finally the predictions are tested for robustness by changing the parameters of the system. Systems of moderately higher density and larger systems of similar density be- have in the same predictable manner. RBCs could serve as predictable colloid based on the volume fraction of depletants in the system. Exploiting rouleau for- mation at controlled depletant volume fractions could be used to synthesize rod- like configurations to generate structures or networks [75, 76, 77, 78]. Larger net- works of rouleau aggregation could be analyzed by the distributions of aligned cells and clusters. Which could then be a useful comparison to determine the

Page 63 of 89 type of environment that an experimental system is in. Image processing of ag- gregated erythrocytes could be an exciting avenue for bringing the analysis of self-assembly to diagnosing biological conditions where systems of interacting rouleaux are formed.

Page 64 of 89 Bibliography

[1] Austin Nehring, Tyler N. Shendruk, and Hendrick W. de Haan. Morphol- ogy of depletant-induced erythrocyte aggregates. Soft Matter, 14:8160– 8171, 2018.

[2] J. Liam McWhirter, Hiroshi Noguchi, and Gerhard Gompper. Deforma- tion and clustering of red blood cells in microcapillary flows. Soft Matter, 7:10967–10977, 2011.

[3] Dinar Katanov, Gerhard Gompper, and Dmitry A. Fedosov. Microvascu- lar blood flow resistance: Role of red blood cell migration and dispersion. Microvascular Research, 99:57 – 66, 2015.

[4] Luca Lanotte, Johannes Mauer, Simon Mendez, Dmitry A. Fedosov, Jean- Marc Fromental, Viviana Claveria, Franck Nicoud, Gerhard Gompper, and Manouk Abkarian. Red cells’ dynamic morphologies govern blood shear thinning under microcirculatory flow conditions. Proceedings of the National Academy of Sciences, 113(47):13289–13294, 2016.

[5] J. Patrick Shelby, John White, Karthikeyan Ganesan, Pradipsinh K. Rathod, and Daniel T. Chiu. A microfluidic model for single-cell capillary obstruc- tion by plasmodium falciparum-infected erythrocytes. Proceedings of the National Academy of Sciences, 100(25):14618–14622, 2003.

[6] P. C. Obiefuna and D. P. Photiades. Sickle discocytes form more rouleaux in vitro than normal erythrocytes. Journal of Tropical Medicine and Hygiene, 93(3):210–214, 1990.

65 [7] P. C. Obiefuna. Rouleaux formation in sickle cell traits. Journal of Tropical Medicine and Hygiene, 94(1):42–44, 1991.

[8] C. Le Devehat, T. Khodabandehlou, and M. Vimeux. Impaired hemorhe- ological properties in diabetic patients with lower limb arterial ischaemia. Clinical and Microcirculation, 25(2):43–48, 2001.

[9] Byoung Kwon Lee, Azhil Durairaj, Anilkumar Mehra, Rosalinda B. Wenby, Herbert J. Meiselman, and Tamas Alexy. Microcirculatory dysfunction in cardiac syndrome X: Role of abnormal blood rheology. Microcirculation, 15(5):451–459, 2010.

[10] Elena I. Ryabchikova, Larisa V. Kolesnikova, and Sergei V. Luchko. An anal- ysis of features of pathogenesis in two animal models of ebola virus infec- tion. Journal of Infectious Diseases, 179:S199–S202, 1999.

[11] Y. Goldin, T. Tulshinski, Y. Arbel, O. Rogowski, R. Ben Ami, J. Serov, P. Halperin, I. Shapira, and S. Berliner. Rheological consequences of acute infections: The rheodifference between viral and bacterial infections. Clin- ical Hemorheology and Microcirculation, 36(2):111–119, 2007.

[12] Mary Lea Killian. Hemagglutination Assay for Influenza Virus. Springer New York, New York, NY, 2014.

[13] Amy C. Valenciano, Rick L. Cowell, Theresa E. Rizzi, and Ronald D. Tyler. Section 2 - Red blood cells. In Amy C. Valenciano, Rick L. Cowell, Theresa E. Rizzi, and Ronald D. Tyler, editors, Atlas of Canine and Feline Peripheral Blood Smears, pages 23 – 109.e2. Mosby, 2014.

[14] Oguz K. Baskurt and Herbert J. Meiselman. : Basic aspects and clinical importance. Clinical Hemorheology and Microcirculation, 53(1-2):23–37, 2013.

Page 66 of 89 [15] Walter H. Reinhart, Nathaniel Z. Piety, and Sergey S. Shevkoplyas. Influ- ence of red blood cell aggregation on perfusion of an artificial microvascu- lar network. Microcirculation, 24(5):e12317, 2016.

[16] R. Ben Ami, G. Barshtein, D. Zeltser, Y. Goldberg, I. Shapira, A. Roth, G. Keren, H. Miller, V. Prochorov, A. Eldor, S. Berliner, and S. Yedgar. Parameters of red blood cell aggregation as correlates of the inflamma- tory state. American Journal of Physiology-Heart and Circulatory Physiology, 280(5):H1982–H1988, 2001. PMID: 11299197.

[17] M. W. Rampling. Red cell aggregation and yield stress. Clinical Blood Rhe- ology, 1:45–64, 1988.

[18] Shu Chien and Kung-Ming Jan. Ultrastructural basis of the mechanism of rouleaux formation. Microvascular Research, 5(2):155–166, 1973.

[19] Björn Neu and Herbert J. Meiselman. Depletion-mediated red blood cell aggregation in polymer solutions. Biophysical Journal, 83(5):2482–2490, 2002.

[20] Björn Neu, Rosalinda Wenby, and Herbert J. Meiselman. Effects of dex- tran molecular weight on red blood cell aggregation. Biophysical Journal, 95(6):3059 – 3065, 2008.

[21] Suckjoon Jun and Andrew Wright. Entropy as the driver of chromosome segregation. Nature Reviews Microbiology, 8(8):600, 2010.

[22] James Pelletier, Ken Halvorsen, Bae-Yeun Ha, Raffaella Paparcone, Steven J. Sandler, Conrad L. Woldringh, Wesley P. Wong, and Suckjoon Jun. Physi- cal manipulation of the Escherichia coli chromosome reveals its soft nature. Proceedings of the National Academy of Sciences, 109(40):E2649–E2656, 2012.

[23] Tyler N. Shendruk, Martin Bertrand, Hendrick W. de Haan, James L. Harden, and Gary W. Slater. Simulating the entropic collapse of coarse- grained chromosomes. Biophysical Journal, 108(4):810–820, 2015.

Page 67 of 89 [24] Christian Wagner, Patrick Steffen, and Saša Svetina. Aggregation of red blood cells: From rouleaux to clot formation. Comptes Rendus Physique, 14(6):459–469, 2013.

[25] Jonathan K. Armstrong, Rosalinda B. Wenby, Herbert J. Meiselman, and Timothy C. Fisher. The hydrodynamic radii of macromolecules and their effect on red blood cell aggregation. Biophysical Journal, 87(6):4259–4270, 2004.

[26] Edward W. Merrill, Edwin R. Gilliland, T. S. Lee, and Edwin W. Salzman. Blood rheology: Effect of fibrinogen deduced by addition. Circulation Re- search, 18(4):437–446, 1966.

[27] Shu Chien and Kung-Ming Jan. Red cell aggregation by macromolecules: Roles of surface adsorption and electrostatic repulsion. Journal of Cellular Biochemistry, 1(4-5):385–409, 1973.

[28] D. E. Brooks. Mechanism of red cell aggregation. In Blood Cells, Rheology, and Aging, pages 158–162. Springer, 1988.

[29] H. Bäumler, B. Neu, E. Donath, and H. Kiesewetter. Basic phenomena of red blood cell rouleaux formation. Biorheology, 36(5, 6):439–442, 1999.

[30] A. Pribush, D. Zilberman-Kravits, and N. Meyerstein. The mechanism of the dextran-induced red blood cell aggregation. European Biophysics Journal, 36(2):85–94, 2007.

[31] Lian Zhao, Waraporn Kaewprayoon, Hong Zhou, Radostina Georgieva, and Hans Baeumler. RBC aggregation in dextran solutions can be measured by flow cytometry. Clinical Hemorheology and Microcirculation, 65(1):93–101, 2017.

[32] W. Pan, B. Caswell, and G. E. Karniadakis. A low-dimensional model for the red blood cell. Soft Matter, 6:4366–4376, 2010.

Page 68 of 89 [33] D. A. Fedosov, W. Pan, B. Caswell, G. Gompper, and G. E. Karniadakis. Pre- dicting human blood viscosity in silico. Proceedings of the National Academy of Sciences, 108(29):11772–11777, 2011.

[34] J. Liam McWhirter, Hiroshi Noguchi, and Gerhard Gompper. Flow-induced clustering and alignment of vesicles and red blood cells in microcapillaries. Proceedings of the National Academy of Sciences, 106(15):6039–6043, 2009.

[35] S. Chandrasekhar, B. K. Sadashiva, and K. A. Suresh. Liquid crystals of disc-like molecules. Pramana, 9(5):471–480, 1977.

[36] Zhan-Wei Li, Zhao-Yan Sun, and Zhong-Yuan Lu. Simulation model for hierarchical self-assembly of soft disklike particles. The Journal of Physical Chemistry B, 114(7):2353–2358, 2010.

[37] Matthieu Marechal, Alejandro Cuetos, Bruno MartÃŋnez-Haya, and Mar- jolein Dijkstra. Phase behavior of hard colloidal platelets using free energy calculations. The Journal of Chemical Physics, 134(9):094501, 2011.

[38] Lilian C. Hsiao, Benjamin A. Schultz, Jens Glaser, Michael Engel, Megan E Szakasits, Sharon C. Glotzer, and Michael J. Solomon. Metastable orienta- tional order of colloidal discoids. 2015.

[39] George B. Thurston. Rheological parameters for the viscosity viscoelasticity and thixotropy of blood. Biorheology, 16(3):149–162, 1979.

[40] M. R. Hardeman, P. T. Goedhart, J. G. G. Dobbe, and K. P. Lettinga. Laser- assisted optical rotational cell analyser (LORCA); I. A new instrument for measurement of various structural hemorheological parameters. Clinical Hemorheology and Microcirculation, 14(4):605–618, 1994.

[41] Barbara M. Johnston, Peter R. Johnston, Stuart Corney, and David Kil- patrick. Non-Newtonian blood flow in human right coronary arteries: Steady state simulations. Journal of Biomechanics, 37(5):709–720, 2004.

Page 69 of 89 [42] Yaling Liu and Wing Kam Liu. Rheology of red blood cell aggregation by computer simulation. Journal of Computational Physics, 220(1):139 – 154, 2006.

[43] L. Goubergrits, E. Wellnhofer, and U. Kertzscher. Choice and impact of a non-Newtonian blood model for wall shear stress profiling of coronary arteries. In Alexei Katashev, Yuri Dekhtyar, and Janis Spigulis, editors, 14th Nordic-Baltic Conference on Biomedical Engineering and Medical Physics, pages 111–114, Berlin, Heidelberg, 2008. Springer Berlin Heidelberg.

[44] Rep Kubo. The fluctuation-dissipation theorem. Reports on progress in physics, 29(1):255, 1966.

[45] John D. Weeks, David Chandler, and Hans C. Andersen. Role of repulsive forces in determining the equilibrium structure of simple liquids. Journal of Chemical Physics, 54(12):5237–5247, 1971.

[46] Gary S. Grest and Kurt Kremer. Molecular dynamics simulation for poly- mers in the presence of a heat bath. Physical Review A, 33(5):3628, 1986.

[47] Sho Asakura and Fumio Oosawa. On interaction between two bodies immersed in a solution of macromolecules. Journal of Chemical Physics, 22(7):1255–1256, 1954.

[48] Henk N. W. Lekkerkerker and Remco Tuinier. Colloids and the Depletion In- teraction. Lecture Notes in Physics.

[49] Tyler N. Shendruk, Martin Bertrand, James L. Harden, Gary W. Slater, and Hendrick W. de Haan. Coarse-grained molecular dynamics simulations of depletion-induced interactions for soft matter systems. Journal of Chemical Physics, 141(24):244910, 2014.

[50] Joshua A. Anderson, Chris D. Lorenz, and Alex Travesset. General purpose molecular dynamics simulations fully implemented on graphics processing units. Journal of Computational Physics, 227(10):5342–5359, 2008.

Page 70 of 89 [51] Jens Glaser, Trung Dac Nguyen, Joshua A. Anderson, Pak Lui, Filippo Spiga, Jaime A. Millan, David C. Morse, and Sharon C. Glotzer. Strong scaling of general-purpose molecular dynamics simulations on GPUs. Com- puter Physics Communications, 192:97–107, 2015.

[52] Aric Hagberg, Pieter Swart, and Daniel S. Chult. Exploring network structure, dynamics, and function using NetworkX. Technical report, Los Alamos National Lab.(LANL), Los Alamos, NM. (United States), 2008.

[53] Liquid crystals of disc-like molecules. Philosophical Transactions of the Royal Society of London A: Mathematical, Physical and Engineering Sciences, 309(1507):93–103, 1983.

[54] Alejandro Cuetos and Bruno Martínez-Haya. Columnar phases of discotic spherocylinders. The Journal of chemical physics, 129(21):214706, 2008.

[55] Aaron E. Saunders, Ali Ghezelbash, Detlef-M. Smilgies, Michael B. Sigman, and Brian A. Korgel. Columnar self-assembly of colloidal nanodisks. Nano Letters, 6(12):2959–2963, 2006. PMID: 17163739.

[56] Honggang Cui, Zhiyun Chen, Sheng Zhong, Karen L Wooley, and Dar- rin J Pochan. Block copolymer assembly via kinetic control. Science, 317(5838):647–650, 2007.

[57] Giuseppe Colafemmina, Raffaella Recchia, Andrea S Ferrante, Samiul Amin, and Gerardo Palazzo. Lauric acid-induced formation of a lyotropic nematic phase of disk-shaped micelles. The Journal of Physical Chemistry B, 114(21):7250–7260, 2010.

[58] Sara Jabbari-Farouji, Jean-Jacques Weis, Patrick Davidson, Pierre Levitz, and Emmanuel Trizac. Interplay of anisotropy in shape and interac- tions in charged platelet suspensions. The Journal of Chemical Physics, 141(22):224510, 2014.

Page 71 of 89 [59] Richard J Bushby and Owen R Lozman. Discotic liquid crystals 25 years on. Current opinion in colloid & interface science, 7(5-6):343–354, 2002.

[60] Sandeep Kumar. Self-organization of disc-like molecules: Chemical as- pects. Chem. Soc. Rev., 35:83–109, 2006.

[61] Bilal R Kaafarani. Discotic liquid crystals for opto-electronic applications. Chemistry of Materials, 23(3):378–396, 2010.

[62] Zhan-Wei Li, Li-Jun Chen, Ying Zhao, and Zhong-Yuan Lu. Ordered packing of soft discoidal system. The Journal of Physical Chemistry B, 112(44):13842–13848, 2008.

[63] Michael J Solomon. Directions for targeted self-assembly of anisotropic col- loids from statistical thermodynamics. Current Opinion in Colloid & Interface Science, 16(2):158–167, 2011.

[64] Paul AC O’Brien, Michael P Allen, David L Cheung, Matthew Dennison, and Andrew Masters. Elastic constants of hard thick platelets by monte carlo simulation and virial expansion. Soft Matter, 7(1):153–162, 2011.

[65] Roland G Winkler, Dmitry A Fedosov, and Gerhard Gompper. Dynamical and rheological properties of soft colloid suspensions. Current opinion in colloid & interface science, 19(6):594–610, 2014.

[66] Dmitry A Fedosov, Hiroshi Noguchi, and Gerhard Gompper. Multiscale modeling of blood flow: from single cells to blood rheology. Biomechanics and modeling in mechanobiology, 13(2):239–258, 2014.

[67] Nathan D. Bade, Randall D. Kamien, Richard K. Assoian, and Kathleen J. Stebe. Curvature and rho activation differentially control the alignment of cells and stress fibers. Science Advances, 3(9), 2017.

Page 72 of 89 [68] Nathan D. Bade, Tina Xu, Randall D. Kamien, Richard K. Assoian, and Kathleen J. Stebe. Gaussian curvature directs stress fiber orientation and cell migration. Biophysical Journal, 114(6):1467 – 1476, 2018.

[69] Zorana Zeravcic, Vinothan N. Manoharan, and Michael P. Brenner. Col- loquium: Toward living matter with colloidal particles. Review of Modern Physics, 89:031001, 2017.

[70] Yoav Kallus and Miranda Holmes-Cerfon. Free energy of singular sticky- sphere clusters. Physical Review E, 95:022130, 2017.

[71] Miranda Holmes-Cerfon. Sticky-sphere clusters. Annual Review of Con- densed Matter Physics, 8(1):77–98, 2017.

[72] L. Schmidt-Mende, A. Fechtenkötter, K. Müllen, E. Moons, R. H. Friend, and J. D. MacKenzie. Self-organized discotic liquid crystals for high- efficiency organic photovoltaics. Science, 293(5532):1119–1122, 2001.

[73] Matthew A. Lohr, Marcello Cavallaro, Daniel A. Beller, Kathleen J. Stebe, Randall D. Kamien, Peter J. Collings, and Arjun G. Yodh. Elasticity- dependent self-assembly of micro-templated chromonic liquid crystal films. Soft Matter, 10:3477–3484, 2014.

[74] Erwan Paineau, Marie-Eve M. Krapf, Mohamed-Salah Amara, Natalia V. Matskova, Ivan Dozov, Stéphan Rouzière, Antoine Thill, Pascale Launois, and Patrick Davidson. A liquid-crystalline hexagonal columnar phase in highly-dilute suspensions of imogolite nanotubes. Nature Communications, 7:10271, 2016.

[75] Lilian C. Hsiao, Benjamin A. Schultz, Jens Glaser, Michael Engel, Megan E Szakasits, Sharon C. Glotzer, and Michael J. Solomon. Metastable orienta- tional order of colloidal discoids. Nature Communications, 6:8507, 2015.

Page 73 of 89 [76] Mahesh Ganesan and Michael J. Solomon. High-density equilibrium phases of colloidal ellipsoids by application of optically enhanced, direct current electric fields. Soft Matter, 13:3768–3776, 2017.

[77] Daniel Morphew and Dwaipayan Chakrabarti. Clusters of anisotropic col- loidal particles: From colloidal molecules to supracolloidal structures. Cur- rent Opinion in Colloid & Interface Science, 30:70 – 80, 2017.

[78] Bing Liu, Yuanyuan Wu, and Shuping Zhao. Anisotropic colloids: From non-templated synthesis to patchy templated synthesis. Chemistry–A Euro- pean Journal, 2018.

[79] Thomas Williams, Colin Kelley, and many others. Gnuplot 4.6: An interac- tive plotting program. http://gnuplot.sourceforge.net/, 2013.

[80] Kevin G Phillips, Steven L Jacques, and Owen JT McCarty. Measurement of single cell refractive index, dry mass, volume, and density using a tran- sillumination microscope. Physical review letters, 109(11):118105, 2012.

Page 74 of 89 Appendices

75 Appendix A

Simulation time

In the HOOMD-Blue simulation package all values are fully self-consistent. There are no conversion factors between units. There are three fundamental units in HOOMD-Blue:

• Distance - D

• Energy - E

• Mass - M

All other units are values are derived from the combination of these units. In our simulations the derivation for time follows as

r MD2 time - τ = (A.0.1) E

This means that there is a free choice in units to represent time. Using some estimated values for the mass and energy of our system [80]

s 10−12 kg(10−6 m)2 time - τ = = 10−9s (A.0.2) 10−6 kJ/mol

puts our simulation timestep on the order of nanoseconds. This estimate is not intended to replicate real time but rather indicates that the simulations exist at a reasonable timescale.

76 Appendix B

Cell diffusion

An alternative way to measure simulation time is to relate the motion of the object to its own geometry. Rather than stating the simulation timesteps one can choose to use a physical timescale. To accomplish this measurements of the mean square displacement (MSD) and the mean square angular displacement (MSAD) are measured (Fig B.1). The mean square displacement tracks the mo- tion of the center of mass over time. The MSAD is calculated by tracking the rotation of the normal of face of the red blood cell. As the cell rotates away from its initial position the change in angle is acquired via the dot product between the normal vectors along time. Using the MSD and the MSAD the diffusion coefficient can be solved via the Einstein—Smoluchowski relation (1.3.10). The diffusion of the center of mass is D = 0.01033 ± 0.004%σ2/τ and the

−1 rotational diffusion is DR = 0.06858 ± 0.06%τ . This means that the rotational

4 diffusion coefficient corresponds to an average time of about τR ≈ 3 × 10 τ to accumulate π/2 radians of rotation. In short, this gives a good idea of how long it takes for a cell to become uncorrelated with itself with respect to rotation. In terms of simulation time the warm up phase can be represented by ≈ 70τR, the simulation time as 1700τR and the data collection phase as 600τR.

77 1e5 1e4

1e4 1e3

1e3 MSD MSAD

1e2 1e2

10 10 1000 10000 100000 100 1000 10000 100000 Simulation time Simulation time Figure B.1: (Left) The translational mean squared displacement (MSD) of the red blood cells center of mass. Measured diffusion coefficient of the RBC is shown as a red dotted line, with bars indicating the fitting range used. The exponent of 1.0 corresponds to a diffusive motion (2.0 corresponds to ballistic motion). The 1 green dashed line shows the theoretical expectation of 19 t for comparison where 1 19 corresponds to 19 particles used in the smoothcell model. (Right) The MSAD of the RBC’s normal over time.

Page 78 of 89 Appendix C

GPU acceleration

As the complexity and size of a system is increased the required time to ac- complish a simulation also increases. By taking advantage of additional hard- ware found in a computer the calculations can be accomplished more efficiently. HOOMD accomplishes this by using graphical processing units (GPUs) to work alongside the CPU to rapidly iterate through repeated calculations. To acquire strong statistics about the equilibrium position at any given vol- ume fraction our simulations use multiple layered systems. Copies of the same system can be layered on top each other where each is in its own unique uni- verse. Each system is initialized with a different trajectory to prevent any sys- tem from replicating the trajectory of another. This ensures that each system is independent from all others. Forces like the depletant interaction or WCA and the corresponding particles only exist within their own layer and there is no interactions between the layers. By using GPU accelerated software packages simulations are able to scale up in size and layers without a significant penalty in time as compared to non-accelerated systems.

79 Appendix D

PhytoSpherix R code and visualization development

Phytospherix R are nanoparticles developed by Mirexus Inc. These are com- prised of phytoglycogen which is a naturally occurring molecule that is safe for human consumption and readily available from low cost sources. Phytospherix R have many attributes which make them suitable for development in products like skin care or medical applications. However, due to the nature of the nanopar- ticle the exact structure has been difficult to experimentally resolve. The nanopar- ticle is comprised of a central protein which has 3-4 branches of glycogen at- tached to it. As each individual glycogen branch grows outwards once it reaches ≈ 10 monomers it splits into two new separate branches. The new branches proceed to grow until they split again once they reach ≈ 10 monomers. Conse- quently this results in a structure which quickly grows from 3-4 branches to 6-8, then to 12-16 and so forth. This branching and exponential growth is referred to as a dendritic structure. However, because of this exponential growth the nanoparticle has an incred- ibly complex structure. Early theory and estimations theorized a structure with a hollow core and a very dense shell. Experimentally the structure’s radius and density was inconsistently reported by different methods such as X-ray diffrac-

80 tion, AFM measurements or neutron scattering. The radius as reported by neu- tron scattering suggested a radius of ≈ 20nm while X-ray diffraction indicated a radius of ≈ 35nm. The initial simulations used beads joined by the FENE bond (1.3.5) to model the glycogen branches. These branches had excluded volume which was im- plemented by the WCA (1.3.4). Atomistic simulations of glycogen branches had suggested that there exists hydrophobic interactions between the branches in the molecule. A consequence of these hydrophobic interactions resulted in short range attractions between the branches causing the molecule to become more dense. I modified the original code to include the effects of the hydrophobic in- teractions between the glycogen branches. In a coarse grained simulation this is represented by an attractive force between particles. This was implemented by using a force-shifted lennard jones interaction. This is a modified LJ potential which is shifted such that the derivative is equal to 0 at a particular separation distance. The shifted-LJ potential has a free parameter eplison () which repre- sents the strength of the attraction. A zero  effectively eliminates any attraction between particles while a large epsilon ( ' 1.0) represents strong interactions that tightly bind neighboring monomers together in attractive wells. The original code was written in such a way that particles would place them- selves one at a time to build the branches. The advantage of this allows the structure to equilibrate as it grows; it models a realistic process of the glycogen branches growing. However, building the particle like this allows particles to grow in confined areas without regard for nearby branches which might con- flict in 3D space. As such this method requires a force-cap to be implemented. A force-cap prevents the resultant force between any two particles being unreal- istically high. In simulation if two particles are placed in the same location or if a particle is placed inside a neighboring branch the force returned by the WCA po- tential would cause the simulations to stop. The force cap prevents the sudden ejection of particles from hyper-elastic collisions and instead lets them slowly

Page 81 of 89 phase through each other until the repulsive force is much more manageable. The segment of code for the force cap follows as such:

### Pair Table Force Function ###

def FSLJ_ForceCap(r,rmin,rmax,rcut,sigma,epsilon,force_cap,C1,V):

# Take derivative ofV(r)= −4∗epsilon∗(−12∗(sigma∗∗12)∗(1.0/rcut)−13+

6∗sigma∗∗6∗(1.0/rcut)∗∗7)

F = −4∗epsilon∗(−12∗(sigma∗∗12)∗((1.0/r)∗∗13) + (6∗(sigma∗∗6))∗((1.0/r)∗∗7)) −

C1

# enforce force cap to prevent particle ejection during equilibration

if (F > force_cap):

F = force_cap

# Return force to the langevin equation,V ==0 as potential isnt needed.

return (V,F)

# Constants for force calculation

table = pair. table(width=2000)

rcut = 1.5

sigma = 1.0

force_cap = 15.0

V = 0

# This isa constant for the force asa parameter to ensure it goes to zero. It

is the derivative ofV evaluated atr=rcut

C1 = −4∗epsilon∗(−12∗(sigma∗∗12)∗((1.0/rcut)∗∗13) + 6∗sigma∗∗6∗((1.0/rcut)∗∗7))

### Tables for each particle type ###

table . pair_coeff . set(’A’,’A’, func=FSLJ_ForceCap, rmin=0.01, rmax=rcut,

coeff=dict(rcut=rcut, sigma=1.0, epsilon=epsilon, force_cap=force_cap,C1=C1))

table . pair_coeff . set(’A’,’Z’, func=FSLJ_ForceCap, rmin=0.01, rmax=rcut,

Page 82 of 89 coeff=dict(rcut=rcut, sigma=1.0, epsilon=epsilon, force_cap=force_cap,C1=C1))

table . pair_coeff . set(’Z’,’Z’, func=FSLJ_ForceCap, rmin=0.01, rmax=rcut,

coeff=dict(rcut=rcut, sigma=1.0, epsilon=epsilon, force_cap=force_cap,C1=C1))

## End of pair table, You can call pair table witha changing force cap via

changing force cap and re−running func=FSLJ_ForceCap

Using the force cap the simulation would then build the nanoparticle over time. To get the final structure the force-cap would need to be released to allow regular thermal fluctuations to equilibrate the molecule. This was accomplished by incrementally removing the force-cap until the numerical value was much greater than any two particle interactions. The second segment of the modified code is as follows:

#Equilibrate structure before any analysis

fori in xrange(15,150,(1500/15)):

printi

force_cap = i

table.pair_coeff.set(’A’,’A’, func=FSLJ_ForceCap, rmin=0.01,

rmax=rcut, coeff=dict(rcut=rcut, sigma=1.0, epsilon=epsilon,

force_cap=force_cap,C1=C1))

table.pair_coeff.set(’A’,’Z’, func=FSLJ_ForceCap, rmin=0.01,

rmax=rcut, coeff=dict(rcut=rcut, sigma=1.0, epsilon=epsilon,

force_cap=force_cap,C1=C1))

table.pair_coeff.set(’Z’,’Z’, func=FSLJ_ForceCap, rmin=0.01,

rmax=rcut, coeff=dict(rcut=rcut, sigma=1.0, epsilon=epsilon,

force_cap=force_cap,C1=C1))

run(1e3)

#At the end of the loop turn force cap"off" by making it

unrealistically high

force_cap = 1e10

table.pair_coeff.set(’A’,’A’, func=FSLJ_ForceCap, rmin=0.01,

Page 83 of 89 rmax=rcut, coeff=dict(rcut=rcut, sigma=1.0, epsilon=epsilon,

force_cap=force_cap,C1=C1))

table.pair_coeff.set(’A’,’Z’, func=FSLJ_ForceCap, rmin=0.01,

rmax=rcut, coeff=dict(rcut=rcut, sigma=1.0, epsilon=epsilon,

force_cap=force_cap,C1=C1))

table.pair_coeff.set(’Z’,’Z’, func=FSLJ_ForceCap, rmin=0.01,

rmax=rcut, coeff=dict(rcut=rcut, sigma=1.0, epsilon=epsilon,

force_cap=force_cap,C1=C1))

The current literature suggested the structure of the dendritic nanoparticle had the most of its mass on the exterior of the particle. The dense exterior shell would give way to a porous core which potentially had openings to fit bio-compatible molecules inside. Unfortunately a dense shell also requires spe- cialized methods to insert any bio-compatible molecules. Simulations resulted immediately in the reversal of the dense shell theory. The final structures had a dense core with branches which were free to float and explore the exterior en- vironment (Fig. D.1.) This can be advantageous for both functionalizing and getting a molecule to diffuse into the nanoparticle. The images were generated in VMD using Nvidias Opti-X graphics process- ing. The images D.2, D.3 and D.4 depict the Phytospherix R nanoparticle at in- creasing attraction strengths. As the interaction increases the particle loses its hairy colloid shape and collapses into a contained dense ball. Furthermore, the existence of exterior hairs could catch water which gives the molecule a large hydrodynamic radius compared to its physical radius. These images presented to Mirexus generated a drastic shift away from the previous literature results of a dense shell with a porous center. The simulations and the visualizations of the structure have since been the foundation for more complex simulations. The type of structures have also helped direct experimental collaborators which have since provided evidence to support the hairy colloid structure.

Page 84 of 89 Figure D.1: A completed structure of the PhytoSpherix particle with no attrac- tion between branches. The branches near the exterior are dispersed around the outside resulting in a hairy colloid. The exact radius is non specific due to the discrepancy between where the "outside" truly ends. However branches could extend as far as ≈ 60nm to 70nm.

Figure D.2: A half slice of a PhytoSpherix R nanoparticle in a low attraction regime. The center slightly closes into itself while free ends are left to explore the surrounding environment. Of interest is the back folding which has occurred in the image represented by the dark blue the last branching glycogen monomers. colors of almost every generation can be seen folding back into the center. The backfolding reveals that the outer layer is not comprised of only the last gener- ation, and the inner core is not generated from only the initial branches. The estimated diameter of a particle like this is between 30-45nm, compared to a PhytoSpherix R particle which is estimated to be 35nm.

Page 85 of 89 Figure D.3: A half slice of a PhytoSpherix R nanoparticle. Increasing the attrac- tive potential collapses the center of the particle. However, free ends are still able to fluctuate outwards. The estimated radius is about 25-35nm which corre- sponds closely to experiential results.

Figure D.4: A half slice of a PhytoSpherix R nanoparticle. At extremely high attraction the particle completely collapses into itself. This particle has an overall diameter of ≈ 20nm. Small compared to experimental results but might hold insight into the nanoparticle in a different solution.

Page 86 of 89 Appendix E cNAB.LAB Computing Resources

During the development of simulations of Phytospherix R the computational re- sources required to develop and test code increased dramatically. At that point in time the laboratory resources utilized local development on a laptop, which could not handle the increasing requirements of the simulations. I built desk- tops focused on being able to make effective use of graphical processing units (GPU) computation while simultaneously being able to handle both high data input/output and large data manipulation. Initially a consumer level desktop GPU was purchased and added into local machine to test scaling and compati- bility. After successful trials the first purposed built desktop was configured and utilized to generate the results for the Phytospherix R simulations. The compo- nents are outlined in Table. E.0.1. With the initial success granted by switching to local development on GPU’s other projects immediately transitioned to a GPU accelerated desktop develop- ment. I was responsible for the planning and building of another six machines: Polar, Ailurus, Teddy, Arktos, Cinnamon & Grizzly. These machines had to allow a user to interface with them for both code development, visualization and be connected such that anyone in the lab had remote access to all machines. Fur- thermore, I was also tasked with outfitting three MCS faculty machines with new components such as hard drives, RAM and GPU’s to allow them to add to local resources. A shorter list describing some of the main components in the

87 Table E.0.1: Panda - Computational Resource

Component ID CPU Intel i7 6700k CPU Cooler Corsair H110i GTX AIO Motherboard ASUS Z170-A Pro GPU GTX Titan-X (Maxwell) GTX Titan-X (Maxwell) PSU EVGA 1000 P2 RAM 32GB Corsair LPX DDR4 Storage Samsung 850 evo 1TB Western Digital Black 4TB Western Digital Black 4TB Case Corsair 780T Full Tower

six desktops are outlined in table E.0.2.

Page 88 of 89 Table E.0.2: Computational Resources

Polar & Ailurus Teddy Arktos & Cinna- Grizzly mon

CPU Intel i7 6700k Intel i7 7700k Intel i7 8700k AMD Threadrip- per 1950x Motherboard ASUS Z170-A ASUS Z270-A ASUS Z370-A ASUS TR4 Zenith GPU EVGA 1080 EVGA 1080 Ti PNY 1080 Ti PNY 1080 Ti EVGA 1080 EVGA 1080 Ti PNY 1080 Ti PNY 1080 Ti Storage ADATA 1TB SSD Crucial 1TB SSD Crucial 1TB SSD Crucial 1TB SSD WD Black 4TB WD Black 4TB RAM 32GB 32GB 32GB 64GB

Page 89 of 89