165
Pennsylvania State University The Graduate School College of Engineering COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL INTERACTION AND DRUG DELIVERY MICROROBOTS A Dissertation in Bioengineering by Byron J. Gaskin c 2018 Byron J. Gaskin Submitted in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy August 2018

COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

  • Upload
    others

  • View
    3

  • Download
    0

Embed Size (px)

Citation preview

Page 1: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

Pennsylvania State University

The Graduate School

College of Engineering

COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL

INTERACTION AND DRUG DELIVERY MICROROBOTS

A Dissertation in

Bioengineering

by

Byron J. Gaskin

c© 2018 Byron J. Gaskin

Submitted in Partial Fulfillment

of the Requirements

for the Degree of

Doctor of Philosophy

August 2018

Page 2: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

The dissertation of Byron J. Gaskin was reviewed and approved∗ by the following:

Robert F. Kunz

Professor of Mechanical Engineering

Bioengineering Intercollege Graduate Degree Program Faculty

Dissertation Advisor, Chair of Committee

Robert L. Campbell

Associate Research Professor, Applied Research Laboratory

Cheng Dong

Professor of Bioengineering

Head of the Department of Bioengineering

William O. Hancock

Professor of Bioengineering

Professor-in-charge of Bioengineering Graduate Programs

Sean M. McIntyre

Assistant Research Professor, Applied Research Laboratory

Special Member

∗Signatures are on file in the Graduate School.

ii

Page 3: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

Abstract

A computational tool has been developed to model flowing cellular systems and hasbeen applied to direct numerical simulation of microvascular flows with a visiontowards personalized medicine. This tool couples computational fluid dynamics(CFD), computational structural mechanics (CSM), six degree-of-freedom (6DOF)motion, and surface biochemistry (SB), in the context of interface-resolved cellgeometry, to provide a detailed model of the heterogeneous blood flow microenvi-ronment. This tool can be used to study drug-mediated cellular interactions in thevasculature and design magnetically-actuated drug delivery microrobots (DDMRs)with targeting capabilities. The research hypothesis of this dissertation was thatapplying direct numerical simulation to study drug-mediated cellular interactionsand DDMR dynamics can lead to protocols for patient-specific drug treatments,including use of DDMRs.

The goal of this dissertation work was to validate the individual components ofthis tool and explore the capabilities and limitations of the tool being developed.The specific aims of this research were to develop a new coupled interface-resolvedfluid-structure-biochemistry interaction (FSBI) numerical scheme for low Reynoldsnumber vascular flows and perform direct numerical simulations of cell-cell inter-actions and DDMRs undergoing magnetic actuation.

The first specific aim is crucial as the fluid-solid interface must be handledwith care when solving fluid-structure interaction (FSI) problems. This is due tothe inconsistency that arises from solving for velocity in the fluid subdomain anddisplacement in the solid subdomain. Several coupling procedures have been devel-oped, implemented, and evaluated to determine an appropriate coupling strategyfor handling relevant problems of interacting bodies in vascular flow. The FSIformulations have been implemented into a robust, production-ready flow solvercapable of modeling interactions between cells.

iii

Page 4: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

For the second specific aim, complete FSBI problems with multiple cell typesand DDMR designs are presented to show the capabilities of the developed scheme.The input parameters for these problems include initial cell locations, biochemi-cal reaction constants, solver time step, blood cell structural properties, DDMRgeometry, and flow shear rate.

This work provides key innovations over current state-of-the-art, namely anapproach to solve the full flow-structure-biochemistry system by modifying readilyavailable vascular flow solvers, demonstrated ability of finite-volume discretizationof hyperelastic constitutive models for large motion and large strain systems, anddesign and evaluation capabilities for magnetically actuated DDMRs in microvas-cular flow.

Collectively, the developed tool can be used in future work to use patient-specific biomarker data to develop personalized drug treatments protocols, designmagnetically-actuated drug delivery microrobots for optimal tissue targeting, andeducate a reduced-order-model with decreased runtime for greater feasibility ofclinical deployment.

iv

Page 5: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

Table of Contents

List of Figures ix

List of Symbols xiii

Acknowledgments xxi

Chapter 1Introduction 11.1 Background and Significance . . . . . . . . . . . . . . . . . . . . . . 11.2 Innovation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41.3 Literature Review . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

1.3.1 Cell Computational Fluid Dynamics Simulations . . . . . . . 61.3.2 Structural Mechanics Modeling . . . . . . . . . . . . . . . . 81.3.3 Fluid Structure Interaction . . . . . . . . . . . . . . . . . . . 101.3.4 Adhesion Biochemistry . . . . . . . . . . . . . . . . . . . . . 111.3.5 Drug Delivery Microrobots . . . . . . . . . . . . . . . . . . . 13

1.4 Research Goals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15

Chapter 2Theoretical Formulation 172.1 Governing Equations . . . . . . . . . . . . . . . . . . . . . . . . . . 17

2.1.1 Cauchy Momentum Equation . . . . . . . . . . . . . . . . . 172.1.2 Fluid Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . 182.1.3 Structural Mechanics . . . . . . . . . . . . . . . . . . . . . . 19

2.1.3.1 Linear Elasticity . . . . . . . . . . . . . . . . . . . 202.1.3.2 Nonlinear Elasticity . . . . . . . . . . . . . . . . . 212.1.3.3 General Updated Lagrangian Formulation . . . . . 26

v

Page 6: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

2.1.3.4 Slow-Process Updated Lagrangian Formulation . . 292.1.3.5 Slow-Process Updated Lagrangian Boundary Con-

ditions . . . . . . . . . . . . . . . . . . . . . . . . . 312.1.4 Surface Biochemistry . . . . . . . . . . . . . . . . . . . . . . 322.1.5 Rigid Body Motion . . . . . . . . . . . . . . . . . . . . . . . 332.1.6 DDMR Magnetic Actuation . . . . . . . . . . . . . . . . . . 35

2.2 Coupling of Dynamic Systems . . . . . . . . . . . . . . . . . . . . . 372.2.1 Hydrodynamics / Fluid-Structure Interaction . . . . . . . . 37

2.2.1.1 Boundary Conditions on Rigid Bodies . . . . . . . 372.2.1.2 Boundary Conditions on Hyperelastic Bodies . . . 38

2.2.2 Biochemistry . . . . . . . . . . . . . . . . . . . . . . . . . . 382.2.2.1 Boundary Conditions on Rigid Bodies . . . . . . . 382.2.2.2 Boundary Conditions on Hyperelastic Bodies . . . 39

2.2.3 Modeling of Surface Roughness . . . . . . . . . . . . . . . . 392.2.4 Unified Governing Equation . . . . . . . . . . . . . . . . . . 40

Chapter 3Computational Implementation 413.1 Finite Volume Discretization . . . . . . . . . . . . . . . . . . . . . . 41

3.1.1 Spatial interpolation of a scalar φ . . . . . . . . . . . . . . . 423.1.2 Explicit spatial gradient of a scalar φ . . . . . . . . . . . . . 433.1.3 Implicit spatial gradient of a scalar φ . . . . . . . . . . . . . 433.1.4 Implicit temporal gradient of a scalar φ . . . . . . . . . . . . 44

3.2 Semi-implicit Six-Degrees-of-Freedom Coupling . . . . . . . . . . . 443.3 Structural Mechanics . . . . . . . . . . . . . . . . . . . . . . . . . . 46

3.3.1 Fixed-shape Leukocyte Rolling . . . . . . . . . . . . . . . . 463.3.2 Finite-Volume Structural Mechanics Formulation . . . . . . 503.3.3 Finite Volume Linear Elastostatics . . . . . . . . . . . . . . 50

3.3.3.1 Discrete Governing Equations . . . . . . . . . . . . 503.3.3.2 Implementation of Boundary Conditions . . . . . . 51

3.3.4 Finite Volume Nonlinear Elastodynamics . . . . . . . . . . . 513.3.4.1 Discrete Governing Equations . . . . . . . . . . . . 513.3.4.2 Accommodation of Inertial Contributions . . . . . 523.3.4.3 Lagged Correction of Non-linear Terms . . . . . . . 533.3.4.4 Quantification of Constitutive Model Selection . . . 53

3.4 Proximity-based adaptive timestepping . . . . . . . . . . . . . . . . 553.5 Fourier Stability Analysis of Computational Implementations . . . . 56

3.5.1 Analysis of 1D Laplace Equations . . . . . . . . . . . . . . . 573.5.1.1 Jacobi Method . . . . . . . . . . . . . . . . . . . . 583.5.1.2 Relaxed Jacobi Method . . . . . . . . . . . . . . . 60

vi

Page 7: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

3.5.1.3 Gauss-Seidel method . . . . . . . . . . . . . . . . . 613.5.1.4 Relaxed Gauss-Seidel Method . . . . . . . . . . . . 61

3.5.2 Analysis of 3D Steady Linear Elasticity Equations . . . . . . 653.5.3 Fourier Stability of Analysis of Multi-Step Solution Procedures 74

3.6 Miscellaneous High Performance Computing Improvements . . . . . 753.6.1 Linear Solver Performance Optimization . . . . . . . . . . . 75

3.6.1.1 Empirically-based Performance Optimization . . . 753.6.1.2 Fourier-based Performance Optimization . . . . . . 77

Chapter 4Verification, Validation, and Results 824.1 Fixed Shape Leukocyte Rolling . . . . . . . . . . . . . . . . . . . . 824.2 Finite-Volume Structural Mechanics . . . . . . . . . . . . . . . . . . 86

4.2.1 3D Linear Elastostatics . . . . . . . . . . . . . . . . . . . . . 864.2.2 Stretching of Prismatic Beam under Self-Weight . . . . . . . 864.2.3 Flexure of Prismatic Beam Due to End Loading . . . . . . . 894.2.4 Saint-Venant Kirchhoff Hyperelasticity . . . . . . . . . . . . 94

4.2.4.1 Hyperelastic Sphere undergoing Rigid Body Rota-tion . . . . . . . . . . . . . . . . . . . . . . . . . . 94

4.3 Single Body Simulations . . . . . . . . . . . . . . . . . . . . . . . . 954.3.1 Impulsively Started Rigid TC in Uniform Flow . . . . . . . . 954.3.2 Rigid TC in Linear Shear Flow . . . . . . . . . . . . . . . . 984.3.3 Rigid Helical Microswimmer . . . . . . . . . . . . . . . . . . 103

4.3.3.1 Constant Torque . . . . . . . . . . . . . . . . . . . 1074.3.3.2 Constant Angular Velocity . . . . . . . . . . . . . . 110

4.3.4 Magnetically Actuated Hyperelastic Microbead in CouetteFlow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111

4.4 Multiple Body Simulations . . . . . . . . . . . . . . . . . . . . . . . 1144.4.1 Free-Flowing Rigid TC and Wall-Adhered Hyperelastic PMN

in Couette Flow . . . . . . . . . . . . . . . . . . . . . . . . . 1144.4.2 Rigid Two-cell Aggregate Formation Simulations . . . . . . 118

Chapter 5Conclusions and Future Work 1215.1 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1215.2 Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122

Appendix A124

A.1 SI Scaling of Microvascular System . . . . . . . . . . . . . . . . . . 124

vii

Page 8: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

Appendix B127

B.1 Gradient Reconstruction using Least Square Optimization . . . . . 127B.1.1 Derivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127B.1.2 Boundary Conditions . . . . . . . . . . . . . . . . . . . . . . 131

B.1.2.1 Dirichlet BC . . . . . . . . . . . . . . . . . . . . . 131B.1.2.2 Extrapolation BC . . . . . . . . . . . . . . . . . . 131

Bibliography 132

viii

Page 9: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

List of Figures

1.1 Simulation of heterogeneous biological cell flow. Tumor cells andpolymorphonuclear leukocytes are near the endothelial wall. Redblood cells are near the flow centerline. . . . . . . . . . . . . . . . . 2

1.2 Flowchart of CellCFD-PSU computational tool. . . . . . . . . . . . 41.3 Pictorial of ligand and receptor distribution of interest in the Melanoma-

PMN-Endothelium system. . . . . . . . . . . . . . . . . . . . . . . . 121.4 Types of magnetic microrobots and their actuation methods. Re-

produced from [88] . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

2.1 Microscopy image of PMN attached to endothelial wall in shearflow. The flow profile around the cell is obtained using particleimage velocimetry (PIV) [53] . . . . . . . . . . . . . . . . . . . . . . 19

2.2 Simplified representation of deformation gradient. . . . . . . . . . . 222.3 Simplified representation of deformation gradient decomposition us-

ing updated Lagrangian approach. Total Lagrangian considers ini-tial and deformed configurations only. Updated Lagrangian alsoconsiders a time-dependent intermediary configuration. . . . . . . . 27

2.4 Methods of magnetic actuation. Reproduced from [88] . . . . . . . 352.5 Cell surfaces contain many complex structures which are modeled

using a sub-grid fictitious repulsion force. [28,103] . . . . . . . . . . 39

3.1 Arbitrarily shaped polyhedra with shared face. . . . . . . . . . . . 423.2 Analytic transformation from a sphere to deformed PMN. . . . . . 483.3 Flowchart of fixed-shape PMN rolling algorithm . . . . . . . . . . . 483.4 Spectral radii of relaxed Gauss-Seidel applied to 1D Laplace’s equa-

tion as function of wavenumber . . . . . . . . . . . . . . . . . . . . 633.5 Maximum spectral radius of relaxed Gauss-Seidel applied to 1D

Laplace’s equation as function of relaxation factor. Spectral radiiare obtained by sampling ρk at wavenumber π/64. . . . . . . . . . . 64

3.6 Spectral Radius of relaxed Jacobi method applied to linear elasticityas function of φy and φz when φx ≈ 0.0 . . . . . . . . . . . . . . . . 71

ix

Page 10: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

3.7 Spectral Radius of relaxed Jacobi method applied to linear elasticityas function of φy and φz when φx = π . . . . . . . . . . . . . . . . . 72

3.8 Spectral radius of relaxed Jacobi method applied to linear elasticityas function of wavenumber when φx = φy = φz for various relax-ation factors. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73

3.9 Sample residual profile showing solution root mean square (RMS)error as a function of solver iteration number . . . . . . . . . . . . . 76

3.10 Estimated performance increase of two-step relaxed Jacobi methodapplied to 3D linear elasticity equations. Two-step methods showthe potential of 78% increase in performance over the single steprelaxed Jacobi method. . . . . . . . . . . . . . . . . . . . . . . . . . 81

4.1 Trajectory of point on the surface of rolling PMN. . . . . . . . . . . 834.2 X coordinate of surface point vs time during PMN rolling. . . . . . 844.3 Y coordinate of surface point vs time during PMN rolling. . . . . . 854.4 Pictorial representation of prismatic beam stretching under self-

weight. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 874.5 Deformation of bar due to self weight. . . . . . . . . . . . . . . . . . 884.6 Grid convergence plot for stretching of beam due to self weight. . . 894.7 Pictorial representation of prismatic beam experiencing flexure due

to end loading T . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 904.8 Deformation of bar due to flexure traction loading at bar end. . . . 924.9 Grid convergence plot for flexure of beam due to end loading. . . . 934.10 Surface mesh of sphere. . . . . . . . . . . . . . . . . . . . . . . . . . 944.11 Pictorial representation of computational domain and boundary

conditions for rigid TC in uniform flow. . . . . . . . . . . . . . . . . 964.12 Angled view of TC and slice of flow field along the centerline of

the computational domain at t = 10µs. Slice is colored by velocitymagnitude and flow is in the +x direction. Gradients in the velocityfield indicate the TC has not yet reached Uinf . . . . . . . . . . . . 97

4.13 Analytic and computed velocity profiles of rigid TC in uniform flowas function of time. . . . . . . . . . . . . . . . . . . . . . . . . . . . 98

4.14 Pictorial representation of computational domain and boundaryconditions for rigid TC in linear shear flow. . . . . . . . . . . . . . . 99

4.15 Angled view of TC and slice of flow field along the centerline ofthe computational domain at t = 1ms. Slice is colored by velocitymagnitude and flow is in the +x direction. Gradients in the velocityfield are due to the prescribed flow shear rate. . . . . . . . . . . . . 100

x

Page 11: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

4.16 Trajectory of TC centroid in linear shear flow. Dashed vertical linesrepresents the end of the modeled computational domain. Cyclicconditions are used to increase the effective flow length. . . . . . . . 101

4.17 Velocity profile of TC centroid in linear shear flow. Simulation endsas TC collides into the top wall of the computational domain. Walleffects appear in the profile as signal noise towards the end of thesimulation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102

4.18 Schematic of helical microswimmer with labeled geometric entities.Helical geometries are defined by helix angle (θ), helix radius (R),filament radius (r), helix pitch (λ), and number of turns (n) . . . . 104

4.19 Mesh of the helical microswimmer used in the simulations per-formed in this work. Geometric properties of this helix are foundin Table 4.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105

4.20 Mesh of the helical microswimmer and slice of the meshed fluiddomain used in the simulations performed in this work. Pointson the interface of the fluid and helix domains match exactly asconformal meshing is used at the interface. . . . . . . . . . . . . . . 106

4.21 Helix in flow and slice of flow field along the centerline of the compu-tational domain at t = 97µs. Slice is colored by the z-componentsof velocity. Red indicates flow coming out of the page and blueindicated flow going into the page. . . . . . . . . . . . . . . . . . . 108

4.22 Comparison of computed axial velocity profile and predictions basedon the Abbott model for a helical microswimmer undergoing con-stant torque. These models are in agreement with a slight differencein the profile slope. This difference is likely caused by the idealizedviscous drag approximation in the Abbott model. . . . . . . . . . . 109

4.23 Comparison of computed normalized axial velocity profile for a he-lical microswimmer undergoing constant angular velocity. Velocityis normalized by the Abbott model prediction. The steady-statevelocity asymptotes to approximately 0.86. . . . . . . . . . . . . . . 111

4.24 Pictorial representation of computational domain for wall-adjacenthyperelastic microbead in Couette flow. . . . . . . . . . . . . . . . . 112

4.25 Side view of the hyperelastic microbead and slice of flow field alongthe centerline of the computational domain. The microbead hasreached its equilibrium shape. Flow is in the +x direction . . . . . . 113

4.26 Comparison of initial and final shape of wall-adjacent magnetic mi-crobead in linear shear flow. The final shape has a flat bottomsurface and has evolved to an equilibrium shape best suited for thisflow region. Flow is from left to right. . . . . . . . . . . . . . . . . . 114

xi

Page 12: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

4.27 Pictorial representation of computational domain for free-flowingrigid TC and wall-adhered hyperelastic PMN in Couette flow. . . . 115

4.28 Rigid TC and wall-adhered hyperelastic PMN at various times dur-ing the simulation. The bodies are colored by magnitude of dis-placement. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116

4.29 Rigid TC, wall-adhered hyperelastic PMN, and slice of flow fieldalong centerline of computational domain at t = 9, 320µs. Theslice is colored by magnitude of flow velocity. . . . . . . . . . . . . . 117

4.30 TC velocity and number of bonds as a function of time. Initialdrop in velocity indicates collision between TC and PMN. TC-PMNaggregate forms when TC velocity reaches zero. . . . . . . . . . . . 119

B.1 Uniform rectilinear computational compact stencil. . . . . . . . . . 128

xii

Page 13: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

List of Symbols

Abbreviations

2PK Second Piola-Kirchhoff

6DOF Six degree-of-freedom

ALE Arbitrary Lagrangian-Eulerian

CFD Computational Fluid Dynamics

CSM Computational Structural Mechanics

DDMR Drug Delivery Microrobot

FBM Flowing Blood Models

FEM Finite Element Method

FSA Fourier Stability Analysis

FSBI Fluid-Structure-Biochemistry Interaction

FSI Fluid-Structure Interaction

FVM Finite Volume Method

GS Gauss-Seidel Method

ICAM Intercellular Adhesion Molecule

J Jacobi Method

LE Linear Elasticity

Pa Pascals

xiii

Page 14: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

PDE Partial Differential Equation

PMN Polymorphonuclear Leukocytes

RBC Red Blood Cell

RBF Radial Basis Function

Re Reynolds Number

RGS Relaxed Gauss-Seidel Method

RJ Relaxed Jacobi Method

RMS Root-mean Squared

SPUL Slow-process updated Lagrangian

SVK Saint Venant-Kirchhoff

TC Tumor Cell

UL Updated Lagrangian

Greek Characters

α Wavenumber per unit length (Fourier stability analysis)Constant speedup (solver performance optimization)

αa Constant for material domain a

βa Maximum spectral radius for solver configuration a

γ Flow shear rate

γn Lighthill drag approximation on thin rigid helical bodies

γp Lighthill drag approximation on thin rigid helical bodies

δij ij-th component of Kronecker delta

ε Constant (radial basis function)

εnx Solution roundoff error of scalar φ at location x and iteration n assuminginfinite mathematical precision

xiv

Page 15: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

εnx Fourier transform of solution roundoff error of scalar φ at location x anditeration n assuming infinite mathematical precision

εijk ijk-th component of Levi-Civita symbol (i.e., permutation symbol)

θ Wavenumber (Fourier stability analysis)Helix angle (DDMR)

θi i-th component of angular displacement vector

λ Lame Constant (solid domain)Equilibrium distance (biochemistry)Helix pitch (DDMR)

µ Fluid viscosity (fluid domain)Shear modulus (solid domain)

µ Material magnetic moment per atom

ν Poisson’s ratio

ρ DensitySpectral radius (Fourier stability analysis)

ρk Spectral radius at wave mode k (Fourier stability analysis)

ρak Spectral radius at wave mode k for method a (Fourier stability analysis)

σij ij-th component of Cauchy stress tensor

τij ij-th component of shear tensor

φ Arbitrary scalar

φx Arbitrary scalar at location x

φnx Arbitrary scalar at location x and iteration n assuming infinite mathemat-ical precision

φnx Arbitrary scalar at location x and iteration n assuming finite mathematicalprecision

χ Deformation vector from reference to deformed configurations

Ψ Hyperelastic strain energy function

xv

Page 16: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

ω Relaxation factorAngular velocity

ωimp Relaxation factor of semi-implicit 6DOF motion solver

Roman Characters

A Arbitrary constant

At Arbitrary constant at time t

Ax Prescribed shape constant (PMN shape transform)Stability Constant (Fourier stability analysis)

Ay Prescribed shape constant (PMN shape transform)Stability Constant (Fourier stability analysis)

Az Stability Constant (Fourier stability analysis)

A Arbitrary second-order tensor

b Nonlinear spring constant

bi i-th component of body force vector

B Arbitrary constant

Bt Arbitrary constant at time t

Bi i-th component of magnetic field vector

Bx Stability Constant (Fourier stability analysis)

By Prescribed shape constant (PMN shape transform)Stability Constant (Fourier stability analysis)

Bz Stability Constant (Fourier stability analysis)

c Constant

Ca Body configuration a

Cx Stability Constant (Fourier stability analysis)

Cy Stability Constant (Fourier stability analysis)

xvi

Page 17: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

Cz Stability Constant (Fourier stability analysis)

C Right Cauchy-Green deformation tensor

d Distance between points

dc Distance between centroids of two cells

Dx Stability Constant (Fourier stability analysis)

Dy Stability Constant (Fourier stability analysis)

Dz Stability Constant (Fourier stability analysis)

ej j-th component of unit normal along bond line of action

E Modulus of elasticity

E Green-Lagrangian strain tensor

E′′ij ij-th component of Green-Lagrangian strain tensor from meshed to de-

formed (unknown) configurations

f(x) Arbitrary function of x

f cost Optimization routine cost function

f ref Reference value of optimization routine cost function

f bondj j-th bond formed

fnk Finite-difference coefficient for k-th order derivative at the n-th grid point

Fi i-th component of force vector

f′(x) First spatial derivative of arbitrary function of x

F′ij ij-th component of deformation gradient from referenced to meshed con-

figurations

F′′ij ij-th component of deformation gradient from meshed to deformed (un-

known) configurations

Fij ij-th component of deformation gradient

F Deformation gradient

xvii

Page 18: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

g Gravity

Gk Amplification factor of wave mode k

h Grid spacing

H′′ij ij-th component of displacement gradient from meshed to deformed (un-

known) configurations

Ii i-th tensor invariant

I Identity tensor

J Determinant of deformation gradient

k Wave mode

kb Boltzmann constant

kon Affinity of a molecule to form a bond

k0on Affinity of a molecule to form a bond under equilibrium conditions

koff Affinity of a molecule to break a bond

k0off Affinity of a molecule to break a bond under equilibrium conditions

lp(a, b) lp norm between vectors a and b

L Length of domain

m MassMeter (scale)

msat Saturation magnetization

Mi i-th component of magnetization vector

n Number of turns in helix (DDMR)

nL Number of molecules per surface area

ni i-th component of outward unit normal vector

nfi i-th component of fluid domain outward unit normal vector

nsi i-th component of solid domain outward unit normal vector

xviii

Page 19: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

~n Unit normal vector of shared face of polyhedra

N Number of points in domain (Fourier stability analysis)

NA Avogadro’s number

p Fluid pressureProbability (biochemistry)Arbitrary point (adaptive timestepping)

P First Piola-Kirchhoff stress tensor

Qaij ij-th component of explicit surface stress tensor for material a

r RadiusHelix filament radius (DDMR)

ri i-th component of radius

rn Error residual at n-th iteration of linear solver

~r12 Vector from the centroid of polyhedron 1 to the centroid of polyhedron 2

δr1 Vector from the centroid of polyhedron 1 to shared polyhedra face

R Helix radius (DDMR)

s Spring constantSecond (scale)Estimated speedup (solver performance optimization)

S Surface

sts Transition spring constant

Si i-th component of surface outward area vector

S′ij ij-th component of second Piola-Kirchhoff stress tensor relating refer-

enced and meshed configurations

S′′ij ij-th component of second Piola-Kirchhoff stress tensor relating meshed

and deformed (unknown) configurations

Sij ij-th component of second Piola-Kirchhoff stress tensor

~S Surface area vector pointing from polyhedron 1 to polyhedron 2

xix

Page 20: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

S Second Piola-Kirchhoff stress tensor

t Time

∆t Timestep

∆t∗ Estimated time scale (adaptive timestepping)

T Temperature

Ti i-th component of traction vector

u Linear velocity (DDMR)

ui i-th component of velocity (fluid domain)i-th component of displacement (solid domain)

uai i-th component of displacement on configuration a

v Deformed volume (solid domain)

vai i-th component of velocity of point P a (adaptive timestepping)

va′i i-th component of velocity of point P a relative to point P b (adaptive

timestepping)

V Volume (fluid domain)Reference volume (solid domain)

wi Weight associated with i-th radial basis function

xi i-th spatial componenti-th component of linear displacement vector (rigid body motion) i-thapproximation of value x

xai i-th spatial component in configuration a

x Point on deformed configuration

∆x Increment in x direction

X Point on reference configuration

xx

Page 21: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

Acknowledgments

I would like to thank Dr. Kunz for having provided me with such a great oppor-tunity. My Penn State journey began with a summer internship in Dr. Kunz’sresearch group and has led to immense growth not only as a person but also as aresearcher.

I would like to thank my committee members, Dr. Cheng Dong, Dr. WilliamHancock, Dr. Robert Campbell, and Dr. Sean McIntyre for their helpful commentsand insights throughout this process. They have each had a great impact on howI view and approach my research.

I would like to thank the Applied Research Laboratory and everyone at theGarfield Thomas Water Tunnel. The Laboratory’s Eric Walker Assistantship pro-vided most of my funding throughout graduate school, for which I am extremelygrateful.

I would like to thank the Penn State Office of Graduate Educational EquityPrograms (OGEEP) and all the college multicultural offices at Penn State. Theseoffices play a huge role in the lives of underrepresented minority students at PennState by ensuring we have the resources needed to excel.

I would like to thank the Alfred P. Sloan Foundation’s Minority Ph.D. Pro-gram for providing me with financial support and great professional developmentopportunities.

I would like to thank all the friends I have made during my time at Penn State.They have all made this a wonderful experience.

Finally, I would like to thank my family for always supporting my dreams.They have always held me to my word and pushed me to achieve all that I set outto accomplish.

xxi

Page 22: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

“I have no special talent. I am only passionately curious.”

— Albert Einstein

xxii

Page 23: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

Chapter 1Introduction

In medicine, the treatment options used for a specific condition can have sig-

nificant impacts on a patient’s health outcomes and ongoing quality-of-life with

each option having its own set of advantages and disadvantages. A data-driven

simulation-based tool that accounts for patient-specific parameters can help in

creating personalized treatment strategies. In the past several decades, there has

been significant interest in patient-specific medical treatment strategies [1–3] with

approaches including (but not limited to) genomics-based profiling [4–7], clinical

endpoint based profiling (using information such as patient’s history of health and

disease) [8–10], and biomarker based profiling [11–13]; this study takes a biomarker-

based approach to personalized medicine.

1.1 Background and Significance

The scope of this research is focused on conditions that involve the interactions

of cells, drugs, and drug delivery microrobots (DDMRs) in microvascular flow.

Developing a better understanding of cellular interaction in heterogeneous flow

environments, like the system depicted in Figure 1.1, is critical to developing new

medical treatments as these interactions may cause complex changes in the overall

system dynamics. Any change to the cellular microenvironment may alter variables

such as flow properties, body geometries, or biochemical reaction rates which, in

Page 24: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

2

turn, affect the rates of cellular interaction.

Figure 1.1: Simulation of heterogeneous biological cell flow. Tumor cells and poly-morphonuclear leukocytes are near the endothelial wall. Red blood cells are nearthe flow centerline.

Understanding the coupling between the cellular flow microenvironment and

rates of cellular interactions is also useful in the design and evaluation of DDMRs.

Many DDMR designs have been proposed in the literature but there is a lack

of computational platforms available to evaluate and compare arbitrary designs.

Some of the proposed designs include bio-inspired synthetic microstructures, re-

programmed bacteria [14–17], and microbiorobots (i.e., aggregates of synthetic mi-

crostructures and flagellated bacteria) [16], with each design being quite different

from the others. Several groups have developed preliminary models to determine

flow characteristics of DDMRs. However, these models are usually design specific,

often approximate blood as being a monophasic Newtonian fluid without account-

ing for the presence of discrete cells, and neglect cellular interactions that occur

in the cellular flow microenvironment. There is also work to be done on optimiz-

ing magnetic actuation fields for specific DDMR designs. The computational tool

developed through this research can be used to evaluate the preliminary DDMR

Page 25: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

3

models presented in the literature and develop models from novel DDMR designs.

Another major thrust of this research is the development of interface resolved

fluid-structure interaction (FSI) capabilities to better capture the influences of the

cell membrane, cell molecular scale structure, and DDMR design on the overall

system dynamics. This work focused on developing FSI formulations and test

cases for physiologically-relevant flows and structures. For initial feasibility stud-

ies, these FSI models were developed for Newtonian fluids and linear elastic solids.

There has been some promising work done on the feasibility of applying finite vol-

ume formulations to linear elasticity equations [18–24] with some work focusing

specifically on FSI applications [25–27] that helped in this effort. The FSI mod-

els were subsequently enriched with more sophisticated treatments of structural

mechanics.

A general workflow for the computational tool is shown in Figure 1.2. There

have been many changes to the workflow compared to earlier versions of this tool

[28, 29]. One large area of modification is in the “Solve coupled CFD / CSM /

Biochem / 6DOF Problem” block. Within this block, numerics of each physically-

and biologically-relevant subsystem have been implemented and accommodated in

a way that efficiently solves the complete biological cell flow problem; namely, this

work focused on the coupling of the CFD and CSM through the FSI and 6DOF

formulations.

Solving the CSM problem required additional modification to the meshing

strategy being used. The computational tool implemented here recomputes the

fluid domain mesh at every timestep and retains the the solid body meshes through-

out the simulation. Retention of the solid body meshes allows for easier accumu-

lation of deformation over time. The computed displacement field within the solid

body is then used to update the solid mesh at the end of each timestep. This mesh-

ing strategy allows for the accomodation of large deformation structural mechanics

models which typically rely on stress history in the body.

Page 26: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

4

Figure 1.2: Flowchart of CellCFD-PSU computational tool.

1.2 Innovation

The tool developed here resolves the physics and biology necessary to study the

effects of interactions among micrometer scale bodies in vascular flow. One impor-

tant feature of this tool is interface-resolved modeling of arbitrarily-shaped bodies

through the use of adaptive conformal meshing; this gives the tool an advantage

not only in analyzing novel DDMR designs but also in studying diseases that cause

irregular cell morphologies (e.g., sickle-cell anemia [30]).

In the analysis of DDMRs, the tool’s capablility of modeling microrobots that

are fully synthetic (e.g., microfabricated structures), fully biological (e.g., repro-

grammed bacteria), or synthetic-biological aggregates (e.g., microfabricated struc-

tures with monolayer coating on bacterial swarmer cells) is demonstrated.

In the development of an interface-resolving FSI formulation for low Reynolds

number flow, this research has explored the feasibility of implementing these FSI

Page 27: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

5

formulations into existing, production-ready vascular flow solvers. This research

has also explored techniques to increase computational efficiency of these imple-

mented FSI formulations.

Key innovations over current state-of-the-art:

• Developed an approach to solve the full flow-structure-biochemistry system

by modifying readily available vascular flow solvers.

• Demonstrated the ability of finite-volume discretization of hyperelastic con-

stitutive models for large motion and large strain systems.

• Designed and evaluated capabilities for magnetically actuated DDMRs in

microvascular flow.

1.3 Literature Review

In the last several decades, there have been many attempts at modeling flowing

blood systems. Computational modeling of flowing blood systems provide valuable

insight into a variety of disease states and mechanisms. These models can advance

the understanding of disease by providing data typically unavailable through clin-

ical or experimental approaches. Example of conditions that have benefited from

the use of flowing blood models (FBMs) include sickle cell anemia [30–37], throm-

bus formation, growth, and dislodging [38–42], malaria [33, 43], hemolysis [44],

and cancer metastasis [28,45,46]. These models can additionally lead to potential

improvements in drug and therapy development for these diseases.

There are, however, significant scientific challenges associated with the devel-

opment of FBMs. Complex internal cell structures, wide-ranging fluid regimes,

and rich biochemistry must all be coupled to achieve a physically correct model.

The coupling of these nonlinear systems often leads to large computational re-

quirements that can make large scale simulation intractable. Each of these physics

Page 28: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

6

may have effects acting across various space and time scales leading to issues of

multiscale resolution.

This work began as an effort to model cancer metastasis [28,45,46] and has since

evolved to a more general FBM which incorporates the key innovations previously

stated. The implementation of a finite volume interface-resolved large deformation

fluid-structure-biochemistry (FSBI) coupling is particularly noteworthy. Literature

reviews for each component of this FBM are presented next.

1.3.1 Cell Computational Fluid Dynamics Simulations

Literature covering cell fluid mechanics often assumes blood to be a non-Newtonian

fluid [47–49]. The non-Newtonian assumption is valid for capturing the aggre-

gate behavior of blood plasma along with all of the blood constituents (e.g., cells,

platelets, dissolved proteins). These blood constituents are explicitly modeled in

this work and their effects are captured by their corresponding models. As such,

the flow model will only be responsible for capturing the behavior of the Newtonian

blood plasma [28,29,45,47–54].

While Newtonian models are more straightforward than non-Newtonian mod-

els, there is still significant computational cost associated with solving the full

Navier-Stokes equations which contributes to that of the overall FBM. There are

a number of methods available to solve the Navier-Stokes equations for blood flow

systems including particle methods, Lattice-Boltzmann methods, interface cap-

turing methods, hydrodynamic approximation functions, and interface resolved

methods.

Particle methods are popular in FBMs due to their ability to handle systems

with various fluids or materials of differing physical properties [55–57]. This class

of method is Lagrangian in nature. Discrete particles are tracked in time as they

interact with each other and the domain boundaries. Particle differential opera-

tors are applied to the equations governing conservation of mass, momentum, and

energy producing the necessary particle interaction terms. These operators can be

derived to account only for a given particle and its neighbors within a specified

Page 29: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

7

radius [58]. An advantage to using this type of method is that material interfaces

need not be resolved. Setting the correct material properties for each particle dur-

ing case initialization allows for interface dynamics to be captured through particle

interactions. A primary disadvantage of these methods is their inability to handle

fine, complex structures.

Lattice-Boltzmann methods are similar to particle methods but solve the dis-

crete Boltzmann equations instead of the Navier-Stokes equations [59]. Many of

the advantages of using particle methods still apply to this class of methods. An

additional advantage of using Lattice-Boltzmann methods is the ability to run

on massively parallel computing systems by design. Lattice-Boltzmann methods,

unlike particle methods, uses a fixed lattice (i.e., mesh) to create a number of

volumes within the domain. These volumes are filled with a finite number of parti-

cles that evolve in time similarly to the aforementioned particle methods. Each of

these volumes can be solved in parallel with the use of coupling terms that share

information across volume faces.

Interface capturing methods are another widely used approach to model mul-

tiphase flows where all interfacial scales are fully resolved. This class of methods

uses an Eulerian approach, where the Navier-Stokes equations are solved on a fixed

mesh. Effects of interfaces and membranes within the domain are incorporated into

the flow solution through the use of source terms representing the sub-grid scale

dynamics (e.g., surface tension). Examples of interface capturing methods are

volume-of-fluid (VOF) [60] and immersed boundary (IB) [61–69].

Hydrodynamic approximation functions are an approach to fluid dynamics

modeling that captures flow effects for a desired flow regime. Using these functions

allows for quick computation of the approximate hydrodynamic force applied to

the blood cells in the flow. This approach has been used extensively [38,44,70–72]

due to its low computational cost despite its inaccuracies as the desired flow moves

away from the flow regime for which the function was developed.

While interface-resolved methods are the most accurate approach to modeling

fluid dynamics in cell-resolving FBMs, they are also the most computationally

Page 30: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

8

expensive. Conformal meshing techniques are used to explicitly resolve an interface

through time. The appropriate interface condition can be enforced at the correct

locations in this class of methods. This class of method is well suited for complex

geometries as comformal meshing will be used to resolve all of the desired interfaces;

although, care must be taken to ensure to balance between mesh resolution and

computational cost.

In this work, an interface-conformal finite volume approach will be used [29,45,

52, 73, 74]. This approach uses a discretization of the integral form of the Stokes

equations shown in Equation 2.7. The equations are integrated over each volume

in the domain and transformed to face integrals through the divergence theorem as

described in Section 3.1. This approach has also been implemented into a number

of commercially available CFD packages (e.g., STAR-CCM+ and OpenFOAM).

The procedures and formulations described in the work can be implemented into

a code using the interface-resolved finite volume approach.

1.3.2 Structural Mechanics Modeling

A single fluid dynamics model can be used in this work without regard for the

number of blood cells or blood cell types in the flow domain. A similar approach

cannot be used with the modeling of structural mechanics as each blood cell type

has unique structural properties due to varying molecular and internal structures.

For example, red blood cells (RBCs) are observed to be highly deformable [68,75]

and must be modeled with care, while cancerous tumor cells (TCs) experience

small deformations and can be modelled as rigid bodies [46, 51, 53, 76]. As such,

any approach used for structural mechanics modeling must be able to accommodate

a range of blood cell constitutive models.

Modeling of RBCs is particularly challanging due to their high rates of deforma-

tion. An overview of RBC structural mechanics and modeling has been presented

by Fruend [75]. RBC deformations are important as they play a large role in TC

and polymorphonuclear leukocyte (PMN) margination (i.e., migration to the vessel

wall during blood flow).

Page 31: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

9

This work focuses on modeling the near-wall region and assumes the cell mar-

gination has already taken place. This assumption allows for priority to be placed

on capturing the structural deformations of PMNs and TCs near vessel walls.

PMNs are drastically less deformable than RBCs [75]. Previous work on PMN

modeling was able to obtain great understanding of the adhesion biochemistry

mechanisms, even when treating the PMNs as rigid bodies [29, 52]. However,

wall-adherent PMNs experience large deformations due to high shear conditions

with elongations of nearly 150% at wall shear rates of 800s−1 [77]. As such, the

structural mechanics of PMNs must be adequetly captured for modeling cellular

interactions in the near-wall region. One approach to model PMNs is using a pre-

stressed membrane with a viscoelastic Maxwell fluid interior [28,45]. This approach

works well for small deformations but requires unphysically large interior viscosities

to capture large deformations; a phenomenon likely due to the nearly rigid PMN

nucleus. An improvement to this model is a 3-layer system which places a rigid

core within the interior Maxwell fluid [28,45].

This work sought to couple FSBI systems through unified governing equations

and identical discretization operators everywhere in the computational domain.

Many CFD solvers use the finite volume method (FVM) to discretize and solve

fluid systems of interest. This work used structural mechanics models discretized

using FVM to develop FSBI formulations that could be implemented into existing

FVM flow solvers.

Early attempts to develop FVM structural mechanics models used linear elas-

ticity to compute stresses and deformations due to specified loadings on prismatic

beams [18, 23–27]. A substantial amount of work was then done to accomodate

multi-material interfaces [22] using continuity of traction and displacement at the

interface. Much of this work had been done in the context of linear elasticity and

was then extended to large deformation hyperelastic models [78]. The Saint-Venant

Kirchoff hyperelasticity model was discretized for FVM-based solvers using both

total Lagrangian and updated Lagrangian approaches. The difference between the

total and updated Lagrangian approaches are anecdotally shown in Figure 2.3. The

Page 32: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

10

updated Lagrangian approach is of interest as it allows the structural mechanics

computations to be performed on a known geometry that can change in time. The

reference geometry in this approach is always set to the meshed deformed body

geometry used in the solver rather than an arbitrary stress-free configuration. De-

tailed description of the total and updated Lagrangian approaches, albeit in the

context of FEM, was presented by Sussman and Bathe [79].

In this work, an updated Lagrangian approach to hyperelasticity is derived

using FVM and coupled to the fluid system using continuity of traction and no-

slip conditions at the fluid-structure interface.

1.3.3 Fluid Structure Interaction

Interface-conformal methods allow for the most general description of multicom-

ponent systems by not only meshing each of the system components but also the

interfaces of each component. This description can be naturally extended to FSI

problems by meshing the flow domain, solid interior, and fluid-solid interface with

a single mesh. This class of FSI methods require mesh motion techniques and are

sometimes referred to as Arbitrary Lagrangian-Eulerian (ALE) methods [80, 81].

The advantage of using ALE methods is the ability to model interface geometry

and solve governing equations in each material exactly as needed. In the context

of cellular interaction simulation, ALE methods also allow for the modeling of

molecular biochemistry at the cell surface in a way that accounts for distribution

and individual interactions between adhesion molecules.

These methods solve the appropriate governing equations in material with nec-

essary boundary conditions at the interface. Practically, this approach could be

implemented in a number of ways.

The first, and most conceptually simple, is to use two seperate solvers (one

for the fluid and one for the solid) each using the subset of the conformal mesh

corresponding to its material domain [45]. Information from either solver can be

passed to the other for use at the interface boundaries. This is an iterative process

that continues until the two solvers have reached a consistent solution; a process

Page 33: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

11

that may be time-consuming and computationally expensive.

An alternative approach is to use a single solver to solve all relevant governing

equations in the domain. This approach reduces communication costs as all infor-

mation transfer can be done using shared memory communication and provides a

straightforward framework in which to model highly complicated multi-material

systems. This approach was used in this work and led to the development of a

Finite Volume FSBI solver.

1.3.4 Adhesion Biochemistry

Protein expression in the vasculature causes activation of ligands and receptors

on the surface of circulating cells. Compatible ligand-receptor pairs may form

biochemical bonds creating adhesion phenomena among the cells and allowing

the formation of cellular aggregates. In the context of PMN-TC-Endothelium

adhesion, the surface molecules of interest would be ICAM-1 (intercellular adhesion

molecule) on the surfaces of the TC and endothelial cells, β-2 integrins on the PMN

surface, and E- and P-selectins on the endothelial cell surfaces [46,50,51,76,77]. A

pictorial representation of the surface molecule distribution can be seen in Figure

1.3.

This system allows for PMN-TC interaction through bonds formed between

ICAM-1 and β-2 while PMN-Endothelial interactions use strong bonds formed

between ICAM-1 and β-2 and weak bonds formed by selectin molecules. TC-

Endothelium adhesion cannot happen directly but rather must be mediated by

PMN adhesion. Simulation of such a system requires adequate modeling of the

relevant adhesion kinetics.

The adhesion kinetics model used in this work uses a probabilistic approach to

bond formation and breakage. The model used is based upon the early work of

Bell, Dembo, and Hammer [70,71,82,83]. Springs are used to model the adhesion

molecules where the bond formation and breakage rates are based on the molecule

separation distance and several empirically obtained coefficients. A uniform dis-

tribution is prescribed for each of the adhesion molecules on the surface of every

Page 34: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

12

Figure 1.3: Pictorial of ligand and receptor distribution of interest in theMelanoma-PMN-Endothelium system.

cell in the simulation. Cell surface based models of biochemical adhesion were first

introduced by Hammer [70,71,83].

The work by Hammer and coworkers introduced surface receptor distributions

on idealized PMNs based on estimated geometric parameters and empirically com-

puted surface densities. The general approach introduced by Hammer is used in

this work with some increased complexity. This work does not comformally rep-

resent the microvilli making it necessary to modify the model by appropriately

adjusting the prescribed receptor surface distribution and incorporating a ficti-

tious repulsion force to ensure the cells do not get non-physically close [52, 54].

The adhesion kinetics model must also be consistent with both the cell surface

discretization and the local probabilistic nature of the kinetics as bonds can form

or break at any time during the simulation.

Page 35: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

13

1.3.5 Drug Delivery Microrobots

Drug delivery microrobots (DDMRs) are an exciting technological advancement

showing great promise for the future of personalized medicine. Progress has been

made in developing biologically-inspired microscale systems capable of targeted

drug delivery. These microsystems either reproduce desired traits of microorgan-

isms or reprogram microorganisms to perform desired actions [84].

All modern DDMR research begins with the understanding of low Reynolds

number fluid mechanics. The highly viscous nature of these flows require non-

reciprocating motion for propulsion. Unlike free flowing cells, DDMRs must be

capable of more than simply travel along with the fluid flow. Biological motors such

as flagella and cilia have been used as inspiration for creating efficient propulsion

for microrobots. Flagella are typically helical and rotate to achieve propulsion.

Cilia produce locomotion through a non-reciprocal swimming motion consisting of

a powerstroke and recovery stroke.

Design of biologically-inspired motors must consider actuation of motor com-

ponents. Size limitations of these microsystems limit the amount of feasible ac-

tuation methods. A promising mode of actuation is through use of magnetic

fields. Magnetic acuation allows for the development non-invasive targeting ca-

pabities that do not require fuel storage within the microsystem [16]. Figure 1.4

shows various proposed magnetically actuated DDMR design types. Two designs

of magnetically actuated microsystems frequently discussed in the literature are

magnetic microbeads [16,85–88] and flagella-inspired magnetic helical microswim-

mers [16, 85, 86, 88, 89]. Both of these system types are explored in this work and

simulated using the developed computational tool.

Page 36: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

14

Figure 1.4: Types of magnetic microrobots and their actuation methods. Repro-duced from [88]

Individual DDMR designs have been well-characterized for cases when the bod-

ies are rigid. Introduction of deformable microsystems presents an additional chal-

lenge to modeling of these systems due to the added fluid-structure interaction

complexity. Attempts have been made to model deformable DDMRs, specifically

elastic-tailed helical microswimmers [90], showing the need for more robust mod-

eling capabilities.

A large obstacle in the clinical use of DDMRs is the localized control of multi-

ple microrobots using a global inpul signal [91]. Clinical applications of targeting

DDMRs may require many microrobots approaching multiple target sites simul-

taneously. Using a single magnetic source to provide localized instruction to mi-

crorobots can be accomplished through the use of complex magnetic fields and

active control algorithms. The computational tool developed in this work provides

a framework in which to develop and enhance active controller systems for DDMR

swarms beyond the traditional experimental approach.

Page 37: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

15

1.4 Research Goals

The goal of this research is to develop a computational tool to study drug-mediated

cellular interactions in the vasculature and to design drug delivery microrobots

(DDMRs) with targeting capabilities. Cellular interactions play an important role

in disease management in nearly all medical interventions [92–96]. The research

hypothesis is that applying direct numerical simulation to study drug-mediated cel-

lular interactions and DDMR dynamics can lead to protocols for patient-specific

drug treatments, including use of DDMRs. Development of this tool is an evolution

of previous work to determine the probability of circulating tumor cell vascular ex-

travasation [28,29,45,52,54,73]. This tool advances the current state of technology

by discretely resolving the interfaces of all bodies of interest to better capture the

relevant physics and biochemistry.

The specific aims of this research were to develop a new coupled interface-

resolved fluid-surface-biochemistry interaction (FSBI) numerical scheme for low

Reynolds number vascular flows and perform direct numerical simulations of cell-

cell interactions and DDMRs undergoing magnetic actuation.

The first specific aim is crucial as the fluid-solid interface must be handled

with care when solving FSI problems. This is due to the inconsistency that arises

from solving for velocity in the fluid subdomain and displacement in the solid

subdomain. Several coupling procedures have been developed, implemented, and

evaluated to determine an optimal coupling strategy for handling relevant problems

of interacting bodies in vascular flow. The FSI formulations have been implemented

into a robust, production-ready flow solver [74] capable of modeling interactions

between cells [29, 52,54].

For the second specific aim, complete FSBI problems with multiple cell types

and DDMR designs are presented to show the capabilities of the developed scheme.

The input parameters for these problems include initial cell locations, biochemi-

cal reaction constants, solver time step, blood cell structural properties, DDMR

geometry, and flow shear rate.

Collectively, the developed tool can be used in future work to:

Page 38: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

16

• Use patient-specific biomarker data to develop personalized drug treatments

protocols.

• Design magnetically-actuated drug delivery microrobots for optimal tissue

targeting.

• Educate a reduced-order-model with decreased runtime for greater feasibility

of clinical deployment.

Page 39: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

Chapter 2Theoretical Formulation

2.1 Governing Equations

2.1.1 Cauchy Momentum Equation

An underlying approximation throughout this work is the use of a continuum de-

scription of the fluid and solid materials. At the heart of the continuum description

is Cauchy’s Theorem which states the existence of a spatial tensor σij such that

Ti = σijnj (2.1)

where ni is the i-th component of unit normal vector at a given spatial location, Ti

is the i-th component of the traction applied at that spatial location. The spatial

tensor σij is referred to as the Cauchy stress tensor.

From this theorem it is possible to derive a description of the local balance of

linear momentum in the form

ρDuiDt

=∂σij∂xj

+ bi (2.2)

where ρ is the material density, ui is the i-th component of material velocity,

and bi is the i-th component of the body force experienced by the material.

Furthermore, it is possible to show that

Page 40: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

18

σij = σji (2.3)

to satisfy the balance of angular momentum. A complete derivation of the Cauchy

momentum equation can be found in Chapter 19 of [97]. Starting with Equation

2.2 allows for both fluid and solid models to be obtained by substituting the model

for σij that best describes each material.

2.1.2 Fluid Dynamics

The flow system is being modeled as a highly viscous, incompressible Newtonian

fluid. The Cauchy stress, σij, for such a material can be expressed as:

σfluidij = −pδij + µ

(∂ui∂xj

+∂uj∂xi

). (2.4)

Due to the low Reynolds number (Re � 1), the flow is in the Stoke’s regime

and with negligible momentum. As such, the flow field at any time is the solution

of an elliptic boundary value problem with no time derivative term. Therefore,

it is appropriate that the flow be governed by the steady Stokes and continuity

equations [29]:

µ∂2ui∂xj∂xj

=∂p

∂xi− bi, (2.5)

∂ui∂xi

= 0. (2.6)

In Equations 2.5 and 2.6, µ is the fluid molecular viscosity, ui is the i-th com-

ponent of the velocity vector, xi is the i-th component of the spatial coordinate

vector, bi is the i-th component of the body force vector (per unit volume), and p

is the fluid pressure. Performing a volume integral on Equation 2.5 and applying

the divergence theorem gives:

∫S

µ∂ui∂xj

dSj =

∫S

pfacedSi −∫V

bidV. (2.7)

Page 41: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

19

In finite-volume fluid dynamics solvers, Equation 2.7 is solved on every control

volume in the domain. The flow solver used in this tool is valid for this case as

show in Refs [28,74].

2.1.3 Structural Mechanics

As early as the 1950s, scientists have known that biological cells are not perfectly

rigid and have sought to develop models for cellular mechanical properties [98].

Figure 2.1: Microscopy image of PMN attached to endothelial wall in shear flow.The flow profile around the cell is obtained using particle image velocimetry (PIV)[53]

In this work, the first attempt at capturing the effects of structural mechan-

ics was fixed-shape cellular rolling. This approach is described in more detail in

Section 3.3.1. Previous studies have shown that polymorphonuclear leukocytes

(PMNs) adjacent to the vascular wall can form a weak adhesion and roll along the

wall surface [28, 45, 51, 53, 76]. During this rolling, a consistent shear-dependent

shape was observed across a range of experiments; an example of one such shape

can be seen in Figure 2.1.

The structural mechanics can be modeled using steady linear elasticity coupled

with the inter-cellular biochemistry in an effort to add more physical richness to

the computational tool. This linear elasticity approach requires those governing

equations be expressed in a finite-volume formulation for straightforward imple-

mentation into an existing fluid dynamics solver.

Page 42: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

20

2.1.3.1 Linear Elasticity

For a compressible linear elastic (Hookean) material, the stress tensor is given by

σij = λ∂uk∂xk

δij + µ

(∂ui∂xj

+∂uj∂xi

), (2.8)

where λ and µ are the Lame constants and ui is the i-th component of the dis-

placement vector. In the case of elastostatics, the governing equation is

∂σij∂xj

+ bi = 0. (2.9)

Substituting σij into Equation 2.9, performing a volume integral, applying the

divergence theorem, and rearranging gives

∫S

µ∂ui∂xj

dSj = −(∫

S

λ∂uk∂xk

dSi +

∫S

µ∂uj∂xi

dSj +

∫V

bidV

). (2.10)

A quick observation shows that Equations 2.7 and 2.10 are of similar form. This

is one indication that it may be possible to add linear elastic capabilities into an

existing finite volume fluid solver with minimal modification. The linear elasticity

formulation will be validated in a modified fluid solver using canonical beam cases.

Implementation of linear elasticity is important in developing a general struc-

tural mechanics approach, although quick analysis shows that linear elasticity itself

is ill-suited for microvascular flow problems. Bodies in microvascular flow environ-

ments may experience large force loadings and large rotations. Large force loadings

become problematic since the resulting deformations may violate the small defor-

mation assumption used to derive the governing equations of linear elastic bodies.

Issues also arise when bodies undergoing large rotations are described using linear

elasticity. Rigid body rotation should not affect the stress field of a body, as it

does not affect the strain field. However, linear elasticity may predict non-zero

stresses resulting from rigid body rotations. For example, consider a body un-

dergoing rotation θ. The displacements associated with this motion are expressed

as,

Page 43: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

21

u

v

w

=

(cos θ − 1) − sin θ 0

sin θ (cos θ − 1) 0

0 0 0

x− cx

y − cy

z − cz

. (2.11)

Using this expression of the body displacements, the spatial gradients of the dis-

placement field is expressed as,

∂u

∂x= (cos θ − 1), (2.12)

∂u

∂y= − sin θ, (2.13)

∂v

∂x= sin θ, (2.14)

∂v

∂y= (cos θ − 1), (2.15)

∂u

∂z=∂v

∂z=∂w

∂x=∂w

∂y=∂w

∂z= 0. (2.16)

Subsituting these values into Equation 2.8 produce non-zero stress values when

θ 6= 0. Nonetheless, solving linear elastic problems using a computational fluid

dynamics solver serves as a proof-of-concept that fluid and solid systems can be

solved simultaneously using a single unified solver.

Nonlinear elasticity is discussed in the following section and introduces concepts

necessary to capture the richness of loading application and removal on a moving

deformable body.

2.1.3.2 Nonlinear Elasticity

To explore the concepts involved in much of nonlinear elasticity, it is first necessary

to delve into the notion of body configurations and the resulting definitions. Con-

Page 44: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

22

figurations are critical in describing a class of nonlinear elastic materials known as

Green, or hyperelastic, materials.

Figure 2.2: Simplified representation of deformation gradient.

Consider a point X on an body denoted as the reference, or undeformed, con-

figuration. Now consider that reference body is moved through space to become

a deformed body at time t. Following point X through the body motion, that

material particle will be at point x(X, t) on the deformed body; this motion can

be described using a deformation vector, χ(X, t), which maps any point X in the

reference body to its position in the deformed body. This relationship is formally

expressed as

x = χ(X, t). (2.17)

Moving one step forward, it is possible to use the descriptions of point mapping

to explore how curves and surfaces deform during body motion. Consider a vector

Page 45: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

23

dX at point X in the reference body. Observing the vector dX through the body

motion, the vector will transform into dx at point x in the deformed body. The

deformation of this vector can be described using the deformation gradient tensor,

F(X, t); this relationship is expressed as

dx = F(X, t)dX, (2.18)

where F(X, t) is defined as

F(X, t) =∂χ(X, t)

∂X=∂x

∂X. (2.19)

Another important transformation to explore is how volume elements change

during body motion. Consider two volume elements dV and dv defined in the ref-

erence and deformed configurations, respectively. The relationship between these

volume elements is

dv = JdV, (2.20)

J(X, t) := detF(X, t) > 0, (2.21)

where J must be greater than zero due to the physical constraint of material

impenetrability. More in-depth coverage of these preliminary definitions can be

found in references [97,99,100].

Armed with the definitions of χ(X, t), F(X, t), and J it is possible to present

the hyperelastic material theory which is suitable for large-strain deformations.

Hyperelastic materials are defined by a strain-energy function, Ψ = Ψ(F), that

is solely a function of the deformation gradient, F. Taking the derivative of Ψ with

respect to F produces the first Piola-Kirchhoff stress tensor, P.

P =∂Ψ(F)

∂F(2.22)

It is important to note that Ψ = Ψ(F) = Ψ(C) = Ψ(E), that is, the strain energy

Page 46: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

24

depends on the strain and not rotation. Here, C and E are the right Cauchy-Green

deformation tensor and Green-Lagrange strain tensor, respectively. Both of these

tensors are solely functions of F, expressed as:

C = FTF, (2.23)

E =1

2(FTF− I). (2.24)

As shown in Chapter 6 of [99], P can also be expressed as:

P =∂Ψ(F)

∂F= 2F

∂Ψ(C)

∂C= F

∂Ψ(E)

∂E. (2.25)

For isotropic hyperelastic materials, those whose strain energy function does

not change due to rigid-body motion, the strain energy function can be expressed

in terms of the tensor invariants:

Ψ = Ψ[I1(F), I2(F), I3(F)],

= Ψ[I1(C), I2(C), I3(C)],

= Ψ[I1(E), I2(E), I3(E)].

(2.26)

For an arbitrary second-order tensor over a three-dimensional vector space, A, the

three tensor invariants and derivatives of these invariants can be calculated from

the components of A expressed with respect to an orthonormal basis as:

I1(A) = tr(A) = A11 + A22 + A33, (2.27)

I2(A) =1

2[tr(A)2 + tr(A2)]

= A11A22 + A22A33 + A11A33 − A12A21 − A23A32 − A13A31,

(2.28)

I3(A) = det(A), (2.29)

Page 47: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

25

∂I1

∂A= I, (2.30)

∂I2

∂A= I1I−AT , (2.31)

∂I3

∂A= I3A

−T . (2.32)

Having a strain-energy function as a function of the tensor invariants allows for

the derivative of the strain energy function to be written as

∂Ψ(C)

∂C=∂Ψ

∂I1

∂I1

∂C+∂Ψ

∂I2

∂I2

∂C+∂Ψ

∂I3

∂I3

∂C. (2.33)

Knowing C is symmetric, this derivative can be rewritten as

∂Ψ(C)

∂C=∂Ψ

∂I1

I +∂Ψ

∂I2

[I1(C)I−C] +∂Ψ

∂I3

[I3(C)C−1]. (2.34)

Once obtained, the first Piola-Kirchhoff tensor, P, can be related to other stress

tensors as

σ = J−1PFT , (2.35)

S = F−1P, (2.36)

where σ is the Cauchy stress tensor and S is the second Piola-Kirchhoff stress

tensor.

A model known as the Saint-Venant Kirchhoff (SVK) model will be used for

much of the nonlinear elasticity in this work. The SVK model is characterized by

the strain-energy function

Ψ(E)SV K =λ

2tr(E)2 + µtr(E2), (2.37)

where λ is the first Lame parameter of the hyperelastic material and µ is the shear

Page 48: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

26

modulus of the hyperelastic material. This strain-energy function allows us to

obtain P as

PSV K = F∂Ψ(E)SV K

∂E,

= F[λtr(E)I + 2µE],

= λtr(E)F + 2µFE.

(2.38)

This leads to a Cauchy stress tensor of the form

σSV K =λtr(E)

JFFT +

JFEFT , (2.39)

which can be expressed solely in term of F as:

σSV K =λ[tr(FFT )− 3]

2JFFT +

µ

JFFTFFT − µ

JFFT (2.40)

Equation 2.40 shows that the Cauchy stress tensor, σ, can be expressed solely

as a function of the deformation gradient, F, for an isotropic hyperelastic mate-

rial. The logical next step is to present a description of the deformation gradient

consistent with the finite volume computational discretization. An appropriate

description of the deformation gradient can be obtained using the Updated La-

grangian approach.

2.1.3.3 General Updated Lagrangian Formulation

Using an Updated Lagrangian (UL) kinematic description of an elastic body, three

configurations are used to describe the body’s motion. The three configurations

used are the base configuration (CB) defined as the body at t = 0 (where stress is

known and often assumed to be zero), the reference configuration (Ct) defined as

the geometry of the body stored in the solver at time t, and a current configuration

(C∗) defined as the future body being solved for in the solver. At each instant,

Ct is the converged solution from the previous timestep. At t = 0, Ct and CB are

identical.

Proceeding in index notation, deformation gradients are defined as

Page 49: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

27

Figure 2.3: Simplified representation of deformation gradient decomposition usingupdated Lagrangian approach. Total Lagrangian considers initial and deformedconfigurations only. Updated Lagrangian also considers a time-dependent inter-mediary configuration.

dxti = F′

ijdxBj , (2.41)

dx∗i = F′′

ijdxtj, (2.42)

where xai is the i-th spatial location of an arbitrary point in configuration Ca.

Expressing the deformation of the C∗ in terms of CB gives

dx∗i = F′′

ijdxtj = F

′′

ikF′

kjdxBj = Fijdx

Bj , (2.43)

Fij = F′′

ikF′

kj, (2.44)

Page 50: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

28

where Fij is the ij-th component of the deformation gradient from CB to C∗

through Ct. Fij are the terms to be used in the constitutive model chosen for a

given material.

Computing Fij requires computing both F′ij and F

′′ij then taking the product of

the two tensors. For convenience, these terms are calculaed in the solver using the

available gradient solvers. It is important to note that spatial gradient operators

in the solver are with respect to xt. Using the built-in gradient operator, it is

straightforward to compute (F′ij)−1 as

(F′

ij)−1 =

∂xBi∂xtj

. (2.45)

The first step in computing (F′ij)−1 is to ensure the topology of the discretized

body is identical at every time t; doing this eliminates the need to perform mesh-

mesh interpolation. Next, the spatial locations at every quadrature point in CB

must be saved (in this case, the spatial locations of the volume centroids are saved).

These spatial locations are loaded into solver and stored for their corresponding

mesh element. The gradient of the spatial locations gives (F′ij)−1. Lastly, (F

′ij)−1

is inverted to give F′ij. Physical constraints require that det[F

′] > 0, ensuring F

is always invertible. Moreover, F′ij is a mxm square matrix when working in m

dimensional space. In this case, F′ij is represented as square, invertible, and 3x3.

Some additional care is taken when computing F′′ij to allow for future implicit

implementations of constitutive models. Ideally F′′ij should be a function of the

state variable since it is dependent on future values. Fortunately, it is possible to

express F′′ij as a function of displacement, u∗i .

uti = xti − xBi , (2.46)

u∗i = x∗i − xti, (2.47)

Page 51: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

29

H′′

ij =∂u∗i∂xtj

=∂(x∗i − xti)

∂xtj=∂x∗i∂xtj− ∂xti∂xtj

= F′′

ij − δij, (2.48)

giving

F′′

ij =∂u∗i∂xtj

+ δij. (2.49)

This formulation of F′′ij allows for hyperelastic constitutive models to be ex-

pressed as a function of the state variable u∗i . Using this expression to compute

Fij gives

Fij = F′′

ikF′

kj =

(∂u∗i∂xtk

+ δik

)F′

kj = F′

kj

∂u∗i∂xtk

+ F′

ij (2.50)

where F′ij is constant during each timestep. This expression of Fij is then substi-

tuted into the following definition of the second Piola-Kirchhoff stress tensor for

the SVK model to obtain the stress in terms of the unknown displacement

SSV Kij = µ (FkiFkj − δij) +λ

2(FkmFkm − 3) δij (2.51)

2.1.3.4 Slow-Process Updated Lagrangian Formulation

The general updated Lagrangian formulation can be simplified in cases when ei-

ther deformation happens slowly or there is adequate temporal resolution. This

simplification of the general updated Lagrangian formulation will be referred to as

the slow-process updated Lagrangian formulation (SPUL).

The balance of linear momentum for a hyperelastic material with respect second

Piola-Kirchhoff stress is:

ρ∂2ui∂t∂t

=∂(FikSkj)

∂xj+ bi. (2.52)

The change in stress and deformation is assumed to be small when the defor-

mation process occurs slowly. As such, each of these variables can be expressed as

their value at a given point in time plus some small change.

Page 52: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

30

ρ∂2(ui + δui)

∂t∂t=∂[(Fik + δFik)(Skj + δSkj)]

∂xj+ bi, (2.53)

where δSkj is the second Piola-Kirchhoff stress computed using the values of the

displacement increment, δui, instead of the displacement itself. δFij is defined as

δFij = F′′

ij =∂δu∗i∂xtj

+ δij. (2.54)

Subtracting Equation 2.52 from Equation 2.53 gives

ρ∂2δui∂t∂t

=∂(δFikSkj + FikδSkj + δFikδSkj)

∂xj+ bi. (2.55)

The slow-process assumption allows for the approximation that (Fij ≈ δij)

as the deformation change is captured in the δFij term. This further simplifies

Equation 2.55 to

ρ∂2δui∂t∂t

=∂δSij∂xj

+∂[δFik(Skj + δSkj)]

∂xj+ bi. (2.56)

Equation 2.56 is identical to the updated Lagrangian governing equations ob-

tained by Cardiff [78]. The Second Piola-Kirchhoff (2PK) Stress Tensor, Sij, and

its increment, δSkj, are defined for a SVK material as

Sij = µ(F′

kiF′

kj − δij)

2

(F′

kmF′

km − 3)δij

= µ

(∂uti∂xBj

+∂utj∂xBi

+∂utk∂xBi

∂utk∂xBj

)+λ

2

(∂utm∂xBm

+∂utm∂xBm

+∂utk∂xBm

∂utk∂xBm

)δij,

(2.57)

δSij = µ(F′′

kiF′′

kj − δij)

2

(F′′

kmF′′

km − 3)δij

= µ

(∂u∗i∂xtj

+∂u∗j∂xti

+∂u∗k∂xti

∂u∗k∂xtj

)+λ

2

(∂u∗m∂xtm

+∂u∗m∂xtm

+∂u∗k∂xtm

∂u∗k∂xtm

)δij.

(2.58)

These equations show the gradients of the reference stress are taken in a co-

Page 53: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

31

ordinate system different from that of the increment stress. This need not be the

case. All gradients may be taken with respect to the xti coordinates. It is bene-

ficial to have all gradients taken with respect to the xti coordinates, as this is the

geometry represented in the flow solver, allowing gradients to be computed using

simple interpolation and the divergence theorem. Obtaining Sij using gradients

with respect to xti is done by first computing

(F′

ij)−1 =

∂xBi∂xtj

, (2.59)

at each spatial point of interest. (F′ij)−1 is the gradient of the initial geometry with

respect to the geometry at time t. The deformation gradient yields a 3x3 square

matrix when operating in three-dimensional space. As such, the values of F′ij can

be obtained by inverting a 3x3 matrix at each spatial point of interest.

The slow-process updated Lagrangian formulation was used for modeling all

hyperelastic materials in this work.

2.1.3.5 Slow-Process Updated Lagrangian Boundary Conditions

Dirichlet boundary conditions are straightforward to implement, as displacement is

specified on the boundary. The values of [δSij+δFik(Skj+δSkj)] are computed using

the prescribed displacements for the slow-process updated Lagrangian formulation.

In practice, it is also necessary to compute the displacement gradients at the

boundary to fully implement the Dirichlet boundary condition. This can be done

in a number of ways including least-square approximation methods or distance-

weighted averaging.

Traction (Neumann) boundary conditions tend to be of more importance in

FSI problems. Care must be taken when applying traction boundary conditions

at the surface of hyperelastic bodies as the effective traction is a function of the

applied traction and the internal body stresses. The traction at the surface of a

hyperelastic body using an updated Lagrangian formulation can be expressed as

Page 54: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

32

T effectivei = T surfacei − Sijnsj . (2.60)

The traction boundary condition is introduced into the governing equations at

the appropriate faces as

T effectivei = [δSij + δFik(Skj + δSkj)]nfj , (2.61)

where T effectivei can be used at the faces without the need to compute [δSij +

δFik(Skj + δSkj)]nfj .

2.1.4 Surface Biochemistry

The surface biochemistry formulation used is nearly identical to the work of Behr

and Gaskin [29,52].

The second law of thermodynamics and equilibrium conditions are coupled to

govern the bond formation and breakage in the domain. Modifications were then

made to the biochemistry formulation to allow for localized modeling of individual

bonds.

Localized bond formation and breakage probabilities are calculated for each

compatible molecule pair in the system. The association rate, kon, and dissociation

rate, koff , are calculated as,

kon = k0onnLALexp

(−sts(d− λ)2

2kbT

), (2.62)

koff = k0offexp

((s− sts)(d− λ)2

2kbT

), (2.63)

where k0on is the association rate at equilibrium, k0

off is the dissociation rate at

equilibrium, AL is the surface area of the discrete mesh face, nL is the molecule

surface density, s is the bond spring constant, sts is the bond spring constant

during the transition state, d is the separation distance of the two molecules, λ is

the equilibrium spring length, T is the local temperature, and kb is Boltzmann’s

Page 55: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

33

constant.

The value of kon is then corrected to allow for localized bond modeling,

kon,g = [∑faces

(nLAL)]k0onexp

(−sts(dc − λ)2

2kbT

), (2.64)

kon,ave =

∑faces

kon

number of faces, (2.65)

kon,corr = konkon,gkon,ave

, (2.66)

where dc is the distance between the centroids of the two cells.

For each molecule pair, the probability of bond formation is calculated as,

P = 1− exp(−kon,corr∆t), (2.67)

where ∆t is the elapsed simulation time since the previous calculation. A random

number is then generated and a bond is formed if the calculated probability is

greater than the random number. Similarly, bond breakage is calculated as,

P = 1− exp(−koff∆t). (2.68)

The force due to an individual bond is computed as

f bondj = s(d− λ)ej, (2.69)

where ej is a unit normal at the bond site along the line of action of the bond.

2.1.5 Rigid Body Motion

It is necessary to capture the trajectory of rigid bodies to fully understand the

cellular interactions happening during the cancer extravasation process. A common

approach is to describe the trajectories as an extension of Newton’s second law of

motion,

Page 56: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

34

d2xidt2

=Fim, (2.70)

ICijd2θjdt2

= Ti, (2.71)

where xi is the i-th component of the linear displacement vector, θi is the i-th

component of the vector of Euler angles, Fi is the i-th component of the sum of

all forces applied at the centroid, Ti is the i-th component of the torque applied

about the centroid, m is the mass of the body, and ICij is the moment of inertia

corresponding to the ij-th axis.

To accommodate bodies of arbitrary geometry, the moment of inertia is calcu-

lated as,

ICij =

∫V

ρrirjdV, (2.72)

for each axis where ri is the i-th component of the radius vector to a point in the

body from the axis passing through the body centroid C, and ρ is the mass density

at each point ri.

Past work [29, 45, 52, 73] used an explicit Euler six degree-of-freedom (6DOF)

solver to compute trajectories of rigid bodies despite the strict limitations of such

an approach [29]. Once the right-hand sides of Equations 2.70 and 2.71 were

computed, each of the equations were marched forward in time using the explicit

Euler method.

Building upon the computational tool developed in [29], the rigid body 6DOF

solver was incorporated into the flow solver and semi-implicitly coupled to the

flow field solution procedure; an approach described in greater detail in Section

3.2. This implicit coupling of the 6DOF solver relaxed the stability restrictions of

the computational system and allows for much better runtime performance.

Another approach explored in this work is to include the contributions of inertia

into the momentum conservation equations and remove the 6DOF solver entirely.

Such an approach can be formulated to allow for body deformability. The Up-

Page 57: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

35

dated Lagrangian technique described in Section 2.1.3.2 not only computed the

deformation in a hyperelastic material but also the rigid body contributions of

the material’s trajectory. The computational implementation of this approach is

described in detail in Section 3.3.4.2.

2.1.6 DDMR Magnetic Actuation

Many DDMRs are controlled via magnetic actuation. This conveniently allows for

the use of MRI machines to simultaneously actuate and image the device [17, 87].

Moving forward, there is much work to be done in optimization of the applied

magnetic fields for desired targeting. Figure 1.4 shows several proposed DDMR

designs, and the optimal method of magnetic actuation for each design. When

looking solely at propulsion, torque-driven helical microswimmer DDMRs are most

efficiently driven by rotating magnetic fields, while spherical DDMRs are best

driven by field gradients [88]. However, these analyses do not account for the

desired interactions of the DDMR with the surrounding microenvironment.

Figure 2.4: Methods of magnetic actuation. Reproduced from [88]

Previous work has shown that fluid shear rate affects rates of cellular adhesion

in microvascular flow [50]; where relative translational and rotational velocities of

the bodies of interest, due to flow shear, play important roles in cellular interaction.

Extending these findings to DDMR applications, it is necessary to design DDMR

Page 58: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

36

geometry and actuation with an eye towards both efficient propulsion and desired

biochemical interactions.

The force and torque on an arbitrary body for a given magnetic field Bi are

expressed, respectively, as

Fmi = VMj

∂Bi

∂xj, (2.73)

Tmi = V εijkMjBk, (2.74)

where Mi is the i-th component of body magnetization, V is the volume of the

magnetic object, and εijk is the Levi-Civita symbol (commonly referred to as the

permutation symbol) [88]. Equations 2.73 and 2.74 can be used to determine

the components of motion due to the magnetic actuation. Therefore, DDMR

designs can be actuated in the computational tool using prescribed magnetic fields

to explore the effect of actuation on DDMR targeting capabilities (e.g., efficient

propulsion and desired biochemical interactions).

This work assumes DDMRs are ferromagnetic permanent magnets at satura-

tion magnetization. Ferromagnetic materials are made up of atoms with large

dipole moments due primarily to electron spins [101]. These dipoles in permanent

magnets align themselves due to interactions with one another in the absence of

an externally applied field. In the presence of an external field, these dipoles align

based on the strength of the applied field until all dipoles are in alignment; at which

point, the magnet has reached saturation. Saturation magnetization is defined as

the maximum magnetic dipole per unit volume [101,102].

For this class of magnets, saturation magnetization can be computed as

msat = µatom

volume= µ

mass/volume

mass/atom,

= µρatom

mass= µρ

(nucleon

mass

)(atom

nucleons

),

= µρNA

(atom

nucleons

),

(2.75)

Page 59: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

37

where µ is the material’s magnetic moment per atom, ρ is material density, NA is

Avogadro’s number, and the last term is the reciprocal of the number of nucleons

per atom. For a pure element, the last term is simply the reciprocal of the element’s

atomic number.

At saturation, magnetization, Mi, of the body is defined as

Mi = msatBi. (2.76)

Using this description alone, it may seem that (Tmi = 0) for any possible configu-

ration of the magnetic field. However, (Tmi = 0) only holds for external magnetic

fields that are constant in time. Instantaneous changes in Bi can produce torque-

inducing misalignment of the material’s magnetic dipoles.

2.2 Coupling of Dynamic Systems

2.2.1 Hydrodynamics / Fluid-Structure Interaction

2.2.1.1 Boundary Conditions on Rigid Bodies

Given a flow field around a rigid body, the resultant forces and torques of a New-

tonian fluid acting on a rigid body are

F fluidi =

∫A

(µ∂ui∂xm

nfm + µ∂um∂xi

nfm − pnfi

)dA, (2.77)

T fluidi =

∫A

εijkrj

(µ∂uk∂xm

nfm + µ∂um∂xk

nfm − pnfk

)dA, (2.78)

where p is the fluid pressure at the surface, rj is the j-th index of the vector

pointing from the surface location to the centroid of the rigid body, and εijk is the

Levi-Civita symbol (commonly referred to as the permutation symbol).

Page 60: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

38

2.2.1.2 Boundary Conditions on Hyperelastic Bodies

Implementation of FSI boundary conditions on hyperelastic bodies follows the Neu-

mann boundary condition approach described in Section 2.1.3.5. The traction at

the surface of a hyperelastic body in flow using an updated Lagrangian formulation

can be expressed as

T fluidi = σfluidij nfj = −pnfi + µ

(∂ui∂xj

+∂uj∂xi

)nfj , (2.79)

T FSIi = T fluidi − Sijnsj ,

= T fluidi + Sijnfj ,

=

(∂ui∂xj

+∂uj∂xi

)+ Sij

]nfj − pn

fi ,

(2.80)

The traction boundary condition is introduced into the governing equations at

the appropriate faces as

T FSIi = [δSij + δFik(Skj + δSkj)]nfj , (2.81)

where T FSIi can be used at the faces without the need to compute [δSij+δFik(Skj+

δSkj)]nfj .

2.2.2 Biochemistry

2.2.2.1 Boundary Conditions on Rigid Bodies

Forces applied at the surface of rigid bodies are transformed into equivalent forces

and torques applied at the centroid of the rigid body. The resultant force and

torque at the centroid of the rigid body due to biochemistry is

F bondsi =

∑bond

f bondi , (2.82)

Page 61: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

39

T bondsi =∑bond

(εijkrjfbondk ). (2.83)

2.2.2.2 Boundary Conditions on Hyperelastic Bodies

Forces applied at the surface of hyperelastic bodies are added to the T surface term

in Equation 2.60.

For example, the value of T surface for an FSI face with a biochemical bond

would be

T surfacei = T fluidi + f bondi . (2.84)

This value can then be used to determine T effectivei at the boundary surface.

2.2.3 Modeling of Surface Roughness

Sub-grid scale modeling is used in this work to capture the effects of microstruc-

tures on the surface of cells. Figure 2.5 shows an image of these microstructures

which can be fairly complex. All cells must maintain a minimum separation dis-

tance to account for the presence of the surface microstructures. The specified

minimum separation distance is enforced using a fictitious localized repulsion force.

Figure 2.5: Cell surfaces contain many complex structures which are modeled usinga sub-grid fictitious repulsion force. [28, 103]

The repulsion force used in this work is

fi,rep = [b/d3]ei, (2.85)

Page 62: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

40

where b is a constant repulsion parameter and d is the separation distance between

two points on adjacent cells. The force due to repulsion can be applied to bodies

similarly to the treatment of the effects due to biochemistry.

2.2.4 Unified Governing Equation

Re-arranging of the equations governing conservation of momentum in the fluid

and solid materials gives a single unified equation to represent all material in the

domain. The unified governing equation validates the concept of simultaneously

solving the entire computational domain in an FSI problem and is expressed as

αk∫V

ρ∂2ui∂t∂t

dV −∫S

µ∂ui∂xj

dSj =

∫V

bidV +

∫S

QkijdSj, (2.86)

αstokes = αLinElas = 0, (2.87)

αSV K = 1, (2.88)

Qstokesij = −pfaceδij, (2.89)

QLinElasij = µ

∂uj∂xi

+ λ∂uk∂xk

δij, (2.90)

QSV Kij = µ

(∂uj∂xi

+∂uk∂xi

∂uk∂xj

)+λ

2

(∂um∂xm

+∂um∂xm

+∂uk∂xm

∂uk∂xm

)δij+F

ik

(S′

kj + S′′

kj

),

(2.91)

where αk is a constant determined by material k. Care must be taken in the

physical meaning of state variable ui; the state variable represents flow velocity

in the fluid domain and material displacement in the solid domain. The coupling

conditions described above ensure state variable on either side of an FSI interface

are appropriately treated.

Page 63: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

Chapter 3Computational Implementation

3.1 Finite Volume Discretization

As shown in Chapter 2, the governing equations for the fluid, linear elastic, and

hyperelastic systems all have the same form. The similarity of the equations allows

for the development of a unified solver to simultaneously compute the solutions in

all of the material regions. This work uses a finite volume discretization approach

to solve the fluid and structural domains of the system.

Figure 3.1 shows two adjacent arbitrarily shaped polyhedra with a shared face.

The vector ~r12 points from the centroid of polyhedron 1 to the centroid of poly-

hedron 2. The distance from the centroid of polyhedron 1 to the centroid of the

shared face is denoted as δr1. The area of the shared face is denoted by the vector

~S with the area vector pointing in the direction from polyhedron 1 to polyhedron

2. The unit normal vector to the shared face, ~n, is obtained by normalizing ~S.

All variable data is stored at the centroid of the polyhedron. By convention,

all terms on the left side of the equation will be treated implicitly and all terms of

the right side of the equation will be treated explicitly using lagged values. There

are several discrete operators of importance in this work:

• Spatial interpolation of a scalar φ,

• Explicit spatial gradient of a scalar φ,

Page 64: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

42

Figure 3.1: Arbitrarily shaped polyhedra with shared face.

• Implicit spatial gradient of a scalar φ,

• Implicit temporal gradient of a scalar φ.

Each operator discretization will be individually addressed. More detailed infor-

mation about the discretization used in the baseline solver can be found in [74].

3.1.1 Spatial interpolation of a scalar φ

Spatial interpolation of a scalar φ is crucial as information is stored at the poly-

hedron centroid while the governing equations require values at the faces of the

polyhedra. The value φface is approximated as

φface = (1− k)φ1 + kφ2, (3.1)

Page 65: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

43

k ≡ δr1

δr1 + δr2, (3.2)

where φ1 is the value of φ at the centroid of polyhedron 1. This operator returns

an exact linear interpolation if vectors ~r12 and ~S are parallel.

3.1.2 Explicit spatial gradient of a scalar φ

Gradients of φ can be computed at the centroid of an arbitrary polyhedron using

the divergence theorem

∂φ

∂xi=

1

V

faces∑f

(Sfi φ

f), (3.3)

where V is the volume of the polyhedron, Sfi is the i-th component of the outward

area vector of face f , and φf is the value of φ at face f . The value of this gradient

can then be interpolated to any face using the interpolation operator introduced

above.

3.1.3 Implicit spatial gradient of a scalar φ

The gradient of scalar φ can be arranged to allow for implicit treatment. The

gradient on the shared face shown in Figure 3.1 can be expressed as

∂φ

∂xi

∣∣∣∣face =∂φ

∂xi

∣∣∣∣face −[(

∂φ

∂xj

)face r12j

|r12|

]r12i

|r12|+

[(∂φ

∂xj

)face r12j

|r12|

]r12i

|r12|, (3.4)

where is |r12| the magnitude of ~r12. The third term in Equation 3.4 represents

the contributions of the spatial gradient parallel to ~r12 while the first two terms

represent the orthogonal contributions.

The parallel contributions can be discretely expressed as

Page 66: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

44

[(∂φ

∂xj

)face r12j

|r12|

]r12i

|r12|=

(φ2 − φ1

|r12||r12|

)r12i , (3.5)

while the orthogonal contributions are computed and treated explicitly.

3.1.4 Implicit temporal gradient of a scalar φ

The inertial term in many systems can be expressed as the temporal gradient of a

scalar φ

∂2φ

∂t∂t. (3.6)

Applying a finite difference approach, the second time derivative of a scalar φ

can be discretely expressed as the summation,

∂2φ

∂t∂t≈

np∑n=1

(fn2 φn) , (3.7)

where np is the number of grid points in the temporal stencil, fn2 is the finite

difference coefficient for second derivative at the n-th grid point, and φn is the

value of φ at the n-th grid point. Since the term is being discretized in time, the

grid point n corresponds to an instance in time. The coefficients fn in Equation 3.7

can be computed to arbitrary orders of accuracy using Fornberg’s method [104].

3.2 Semi-implicit Six-Degrees-of-Freedom Cou-

pling

For the rigid body and linear elastostatic structural models, a semi-implicit six

degree-of-freedom (6DOF) solver was used to relax the restrictions of the explicit

off-board 6DOF solver.

The implementation of the semi-implicit 6DOF solver involved modifications to

the explicit implementation described in [29]. The governing equations for 6DOF

Page 67: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

45

motion, as described in Section 2.1.5, can be represented as a first order ordinary

differential equation of the form

dA

dt= B. (3.8)

Beginning with the explicit approach, a first-order differential equation can be

marched forward in time as

At+∆t = At + ∆t

(dA

dt

)t= At +Bt∆t, (3.9)

where the updated value At+∆t depends solely on values known from previous

points in time. This approach is typically the simplest time marching strategy to

implement, but comes with a restrictive stability limitation. Previous work found

the maximum stable timestep for this type of flow system to be around 30µs [29].

Such a limit forces the timestep value to be much smaller than needed to adequately

resolve the system dynamics. In fact, it would take nearly 300 timesteps for an

8µm radius spherical body in a typical microvascular flow to travel the length of

one radius using a timestep of 30µs. The strict stability limit imposed by the

explicit time marching led to the exploration of implicit marching schemes that

could be implemented in a manner consistent with the FSBI solver.

The semi-implicit 6DOF coupling strategy implemented was motivated by the

iterative solution strategy used in the underlying FSBI solver. The underlying

FSBI uses the SIMPLEC algorithm [74] where the state variable being solved is

predicted then corrected iteratively until the solution converges. Each corrector

step in the predictor-corrector loop provides a potential solution to that attempts

to enforce the problems constraints. The idea for the semi-implicit solver is to use

the FSBI corrector solution to compute a potential solution for the 6DOF motion

that is coupled back into the fluid system through the no-slip condition at the

body surface. A semi-implicit discretization of Equation 3.9 can be written as

At+∆t,∗ = At +Bt+∆t,∗∆t, (3.10)

Page 68: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

46

where the updated value At+∆t,∗ is an estimated value that depends on not only

A from the previous timestep but also the most recent potential solution of B.

However, due to the very nature of the iterative procedure, precautions must be

taken to prevent any solution error from growing uncontrollably with each iteration.

The relaxation factor, ωimp, motivated by the under-relaxation approach often

used in SIMPLEC is used to help with stability of the routine. This relaxation

factor is prescribed as a value between zero and one and implemented as

At+∆t,∗ = ωimp(At +Bt+∆t,∗∆t

)+ (1− ωimp)At+∆t,∗∗, (3.11)

where At+∆t,∗∗ is the value of At+∆t,∗ computed at the previous solver iteration.

The relaxed semi-implicit 6DOF motion solver implementation has greatly im-

proved the stability limitations. In practice, the solver remains stable even with

timestep values on the order of 1000µs.

3.3 Structural Mechanics

Rigid body motion solver performance has been greatly improved by use of the

implicit iterative discretization compared to the explicit discretization. Advances

have also been made in the structural modeling of deformable bodies in microvas-

cular flow environments.

In this work, three approaches were explored in the context of structural mod-

eling for coupling to fluid systems with biochemical interactions: 1) fixed-shape

leukocyte rolling discussed in Section 3.3.1, 2) Linear elastic solid modeling dis-

cussed in Section 3.3.3, and 3) Non-linear hyperelastic solid modeling discussed in

Section 3.3.4.

3.3.1 Fixed-shape Leukocyte Rolling

In the near-wall region, deformable leukocytes undergo complex flow-structure-

wall-biochemistry interactions. These interactions lead to PMN rolling along the

Page 69: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

47

endothelial wall with a nominal shape and rolling velocity. Experimental observa-

tions have found the rolling shapes and velocities to be functions of flow parameters

and local biochemical expression [53]. An approximation was sought to model the

complex flow-structure-wall-biochemistry interactions of leukocytes in the near-

wall region in the manner that was computationally efficient and retained the

leukocyte surface mesh topology, to allow for modeling of long-term biochemical

interaction. A viable approximation for the leukocyte near-wall interactions is the

fixed-shape leukocyte rolling strategy.

Analytic transform equations were created to transform a sphere to match

the nominal rolling shape from experimental observations. Beginning with a unit

sphere centered at origin, the nominal rolling shape is obtained using the following

equations:

xnew = xold + Axcos

(πxold

2r

), (3.12)

ynew = yold −max(0, yold)(Ay)−min(0, yold)

[Bycos

(πxold

8r

)], (3.13)

znew = zold, (3.14)

where xold is the x-coordinate of an arbitrary point belonging to the undeformed

sphere, r is the radius of the undeformed sphere, and Ax, Ay, and By are pre-

scribed shape constants chosen based on the fluid shear rate and local biochemical

expression. An example of a shape obtained through this transform is shown in

Figure 3.2.

The obtained leukocyte shape is sent to the FSBI solver along with the pre-

scribed rolling velocities. The velocity at the surface of the leukocyte is computed

using the prescribed values of leukocyte translational and rotational velocities.

A crucial step in the leukocyte rolling strategy is transforming the deformed

shape back into a sphere to rotate the surface mesh. While the sphere-to-leukocyte

Page 70: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

48

Figure 3.2: Analytic transformation from a sphere to deformed PMN.

Figure 3.3: Flowchart of fixed-shape PMN rolling algorithm

transform described by Equations 3.12, 3.13, and 3.14 does not have an analytic

inverse, root-finding methods can be used to approximate the inverse transform.

The Newton-Raphson method is well-suited to approximating the roots of a

Page 71: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

49

real-valued function by using an iterative approach to find a value x that returns

the root of a given function f(x) [105]. The value of x is approximated by

xn+1 = xn −f(xn)

f ′(xn), (3.15)

where xn is the n-th approximation of the value x. This method requires as initial

guess of x, the function f(x), and the derivatives of the function f(x).

Fortunately, the derivatives of the sphere-to-leukocyte analytic transform are

easily obtained by taking the derivatives of Equations 3.12, 3.13, and 3.14. These

derivatives are found to be:

∂xnew

∂xold= 1− πAx

8sin

(πxold

8

), (3.16)

∂ynew

∂xold= 0 ∀ yold ≥ 0, (3.17)

∂ynew

∂yold= 1− Ay ∀ yold ≥ 0, (3.18)

∂ynew

∂xold=

[πByy

old

32

]cos

(πxold

32

)∀ yold < 0, (3.19)

∂ynew

∂yold= 1−Bycos

(πxold

32

)∀ yold < 0, (3.20)

∂znew

∂zold= 1, (3.21)

∂xnew

∂yold=∂xnew

∂zold=∂ynew

∂zold=∂znew

∂xold=∂znew

∂yold= 0. (3.22)

When solving for the inverse tranform using known values of xnew and ynew, the

values of xold and yold can be computed separately for ease of computational im-

plementation. First, xold should be computed using the Newton-Raphson method

since xnew is solely a function of xold. Then, ynew can be computed since ynew is a

Page 72: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

50

function of xold and yold, where yold is the only unknown since xold was solved for

in the previous step. This approach only requires the derivatives ∂xnew

∂xoldand ∂ynew

∂yold

be used by the Newton-Raphson root-finding routine.

3.3.2 Finite-Volume Structural Mechanics Formulation

3.3.3 Finite Volume Linear Elastostatics

3.3.3.1 Discrete Governing Equations

The governing equation for conservation of momentum in linear elastostatic ma-

terials is provided in Section 2.2.4. The discrete form of this equations for an

arbitrary polyhedron in the computational domain is written as

−faces∑f

(µ∂ui∂xj

Sj

)= biV +

faces∑f

(µ∂uj∂xi

Sj + λ∂uk∂xk

Si

), (3.23)

where the terms on the left side of the equation are treated implicitly and terms

on the right side of the equation are treated explicitly. The difficulty of solving

this governing equation using fixed-point iterative schemes has been discussed [18]

where the following modification was proposed to improve solver issues

−faces∑f

[(2µ+ λ)

∂ui∂xj

Sj

]= biV +

faces∑f

[µ∂uj∂xi

Sj + λ∂uk∂xk

Si − (µ+ λ)∂ui∂xj

Sj

].

(3.24)

This modification serves to increase the diagonal dominance of the linear algebraic

system leading to the better solver performance.

Inertial contributions of linear elastic materials are accommodated by using the

rigid body 6DOF solver to compute the trajectory of the body geometry obtained

from the solution to the linear elastostatic problem.

Page 73: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

51

3.3.3.2 Implementation of Boundary Conditions

Boundary conditions for the linear elastic problems of interest are prescribed trac-

tion and prescribed displacement.

In the case of prescribed displacement on a boundary face, the gradient terms

at that face are approximated by one-sided gradients using the prescribed face

displacement and the value at the polyhedron centroid.

In the case of prescribed traction on a boundary face, the following relation is

used:

Ti = µ∂ui∂xj

nj + µ∂uj∂xi

nj + λ∂uk∂xk

ni. (3.25)

Prescribed traction conditions can be enforced by substituting Equation 3.25 into

the discrete governing equations at the appropriate boundary faces.

In practice, for steady simulations, there should always exist at least one point

in the domain where displacement is prescribed to ensure the displacement solution

is not off by an arbitrary constant of integration. No such challenge arises for

unsteady simulations as the solid body will simply accelerate due to the applied

force imbalance.

3.3.4 Finite Volume Nonlinear Elastodynamics

3.3.4.1 Discrete Governing Equations

The governing equation for conservation of momentum in hyperelastic materials

using the Saint-Venant Kirchhoff constituent model is provided in Section 2.2.4.

The discrete form of these equations for an arbitrary polyhedron in the computa-

tional domain is written as

ρ∂2ui∂t∂t

V −faces∑f

(µ∂ui∂xj

Sj

)= biV +

faces∑f

(QSV Kij Sj

), (3.26)

where the terms on the left side of the equation are treated implicitly and terms

on the right side of the equation are treated explicitly.

Page 74: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

52

Similar to the linear elastostatic equations, the discrete form of the governing

equations for SVK materials are modified for increased diagonal dominance as

ρ∂2ui∂t∂t

V −faces∑f

[(2µ+ λ)

∂ui∂xj

Sj

]= biV +

faces∑f

[QSV Kij Sj − (µ+ λ)

∂ui∂xj

Sj

]. (3.27)

Special attention must be given to the computation of the temporal displacement

gradient for the accommodation of the inertial contributions, as well as for the

transformation of the deformation gradient from the stress-reference configuration

to the meshed configuration at a given instance in time.

3.3.4.2 Accommodation of Inertial Contributions

Given uni = (xni − xn−1i ) in the solid domain, where xni is the i-th coordinate of the

spatial location at time n, the discrete form of the second derivative of uni can be

expressed as

∂2ui∂t∂t

≈np∑n=1

(fn2 xni )−

np∑n=1

(fn2 x

n−1i

). (3.28)

The following expression is then obtained through algebraic manipulation:

∂2ui∂t∂t

≈ fnp2 unpi + fnp2 xnp−1i +

np−1∑n=1

{(fn2 − fn+1

2 )xni}− f 1

2x0i . (3.29)

Equation 3.29 is useful in that it allows for an implicit treatment of inertial

contributions. The first term on the right side of the equations contains the un-

known state variable unpi and can be treated implicitly by adding the coefficient

fnp2 to the corresponding point coefficient. Every other term relies solely on the

history of mesh motion and can be added to the source term in the solid mechanics

solver.

Page 75: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

53

3.3.4.3 Lagged Correction of Non-linear Terms

The definition of QSV Kij shows the SVK governing equations are non-linear in dis-

placement ui. Linearization is needed to numerically solve this system using any

linear solver.

While there are many ways to handle the non-linear terms in the model, this

work treats all non-linear terms explicitly. Non-linear terms are computed using

the displacement field provided by the most recent solution of the iterative non-

linear solver.

One interesting area of future work is an exploration of non-linear treatments

on numerical stability of the solution procedure. Implicit handling of some terms in

QSV Kij could provide increased stability and result in overall increases in solver per-

formance. The procedure described in Section 3.5 gives a framework for quantifying

the effects of linearization treatments on iterative solver stability and performance.

3.3.4.4 Quantification of Constitutive Model Selection

Continuity of traction dictates that (Sijnfluidj = T fluidi ) at any point on a fluid-solid

interface. This condition can be used not only to estimate the material properties

best suited for a particular constitutive model but also to quantify the error as-

sociated with selection of given constitutive models based on known observations.

Use of this condition requires knowledge of the solid’s stress-free geometry, the

loading applied to the solid surface, and the solid’s equilibrium geometry based on

the applied loading.

At each point on the surface of the material:

1. Compute deformation gradient between the reference and equilibrium con-

figurations Fij given

(Fij =

∂xequlibriumi

∂xstressFreej

).

2. Construct appropriate stress tensor Sij(Fij, φ1, φ2, ..., φk) as a function of Fij

and the k properties describing the material, (φ1, φ2, ..., φk).

Page 76: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

54

3. Compute lp norm between vectors [S(F, φ1, φ2, ..., φk)nfluid] and T fluid.

Once the lp norm has been computed at each point, the following cost function

can be used in an optimization routine:

f cost(φ1, φ2, ..., φk) =∑

points on surface

{lp[T fluid, S(F, φ1, φ2, ..., φk)nfluid

]}. (3.30)

An optimization routine can be used to find values of (φ1, φ2, ..., φk) that min-

imize equation 3.30.

For leukocytes in this work, the cells are described using the Saint-Venant

Kirchhoff (SVK) constitutive model, assumed to be spheres of radius 4µm when

unstressed, and describe the equilibrium shape of a wall-adhered PMN using the

transformation given in Section 3.3.1.

The SVK stress tensor expressed in terms of F and material properties are

shown in Equation 2.51. In Equation 2.51, material properties µ and λ are func-

tions of the modulus of elasticity, E, and Poisson’s ratio, ν, as follows:

µ =E

2(1 + ν), (3.31)

λ =Eν

(1− 2ν)(1 + ν). (3.32)

The values of T fluid at each surface point is obtained by running a flow calcu-

lation of the deformed leukocyte body fixed at a specified location on the vascular

wall. The shear rate of the flow is chosen to match experimental data [53] and

the leukocyte is assumed to have no further deformation or velocity at the surface.

The obtained flow field is used to compute the traction at the surface of the cell.

Quantification of constitutive model selection error can be done by comparing

a given set of constitutive model and material properties with a reference cost

function value. The following reference cost function value is chosen due to its

indifference towards the constitutive model selection

Page 77: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

55

f ref =∑

points on surface

[lp(T fluid, 0

)], (3.33)

where f ref > 0 if, and only if, there exists at least one point on the material surface

where T fluid 6= 0.

3.4 Proximity-based adaptive timestepping

A proximity-based adaptive timestepping approach was implemented to maintain

high temporal resolution throughout the simulations without sacrificing computa-

tional efficiency. This approach uses body trajectories to compute time-to-collision

for resolved bodies.

Consider two points, P a and P b, each of having an associated spatial location,

xi, and velocity, vi. The line-of-action, di, is then defined as

di = xbi − xai . (3.34)

Next, the frame of reference is shifted such that the velocity of P b, vbi , is identically

zero. The velocities in this frame are expressed as:

vbi = 0, (3.35)

va′

i = vai − vbi . (3.36)

The relative speed of the P a along the line of action is obtained as the magnitude

of the projection of va′i onto a unit vector along d. A characteristic time length

can be computed using the relative speed along d and the length of d:

vrel =1

|d|va′

i di, (3.37)

Page 78: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

56

∆t∗ =|d|vrel

=|d|2

vai di. (3.38)

This relation can be used when vrel and |d| are both greater than zero. Non-

positive values of vrel may occur when the points are moving away from each other.

In such cases, ∆t∗ would not be considered while determining the new timestep

value. Values of |d| will only be zero in the event of collision, at which point the

finite-volume approximation used to solve the governing equations of the physical

systems may no longer be valid.

The characteristic timescale, ∆t∗, can be used to set a timestep that adequately

resolves the physics temporally. This work used a timestep of ∆t∗/N where N is an

integer. This value allows for N time steps before collision, assuming the velocities

remain unchanged. By analyzing all point-pairs in the domain, ∆t∗ is set to be

the minimum computed characteristic timescale. Bounds on the timescale will be

set, however, to ensure values of ∆t∗ are appropriate for the simulation.

3.5 Fourier Stability Analysis of Computational

Implementations

In the development of new computational methods, it is important to understand

the limits within which the method can be confidently relied upon. Moreover, it is

desirable to quickly estimate performance of a method without allocating the full

amount of computational resources required by the method.

Fourier stability analysis (FSA) of discrete computational operators provides

a framework to determine both stability and expected performance of a compu-

tational method. One drawback of the FSA approach, however, is the under-

lying assumption that all boundaries of the computational domain are periodic.

Nonetheless, FSA remains a powerful tool to explore method feasibility, despite its

inability to accommodate Dirichlet or Neumann boundary conditions.

Computational methods will rarely, if ever, outperform the estimates provided

Page 79: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

57

by FSA in practice. This holds especially true when comparing performance esti-

mates to systems with Dirichlet or Neumann boundary conditions. As such, FSA

provides the best-case scenario stability estimate for a given method.

This section will demonstrate FSA using various governing equations represen-

tative of those often found in computational mechanics.

3.5.1 Analysis of 1D Laplace Equations

FSA will first be performed on the scalar Laplace’s equation in one dimension (1D)

expressed as

∂2φ

∂x2= 0. (3.39)

Laplace’s equation is the simplest example of an eliptic partial differential equation,

and provides great insight into many physical systems, including fluid dynamics

and heat conduction [106,107].

The finite difference approximation of second derivative centered in space using

uniform grid spacing gives

∂2φ

∂x2≈ φx+∆x − 2φx + φx−∆x

∆x2= 0. (3.40)

Solving for φx gives

φx =φx+∆x + φx−∆x

2. (3.41)

Four iterative computational solution strategies will be applied to the 1D scalar

Laplace’s equations and analyzed using FSA:

1. Jacobi method,

2. Relaxed Jacobi method,

3. Gauss-Seidel method,

4. Relaxed Gauss-Seidel method.

Page 80: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

58

The Jacobi and Gauss-Seidel iterative methods are widely used in numerical

analysis. Details of these methods in general can be found in nearly all numerical

analysis textbooks.

3.5.1.1 Jacobi Method

Using the Jacobi method, the solution to any point is computed using lagged

values of the solution at every other point in the domain. Applying this approach

to solve for φx, the value of terms on the right side of the equation are set to the

solution from the previous iteration. The discrete iterative equation describing φx

is expressed as

φn+1x =

φnx+∆x + φnx−∆x

2, (3.42)

where n is the iteration number of the Jacobi procedure. Initial guess values are

used for φ0x to start the iterative process.

A solution roundoff error can be defined as

εnx = φnx − φnx, (3.43)

where φnx is the solution assuming finite mathematical precision and φnx is the

solution assuming infinite mathematical precision.

The roundoff error at each iteration can be expressed as

εn+1x =

εnx+∆x + εnx−∆x

2. (3.44)

Uniformly distributing N + 1 points in the domain of length L, such that

N∆x = L, the discrete Fourier series of εnx can be expressed as

εnx =N∑n=1

(εnke

iαx), (3.45)

α =πk

N, (3.46)

Page 81: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

59

where α is the wavenumber per unit length.

Substituting Equation 3.45 into Equation 3.44 gives

N∑n=1

{eiαx

[εn+1k − εnk

2

(eiα∆x − e−iα∆x

)]}= 0. (3.47)

Since eiαx 6= 0, Equation 3.47 can only be satisfied, in general, if

εn+1k =

εnk2

(eiα∆x − e−iα∆x

), (3.48)

which can be further simplified as

εn+1k = εnk cos(α∆x). (3.49)

The amplification factor of solution error, Gk, of the iterative process for mode

k is defined as the ratio:

Gk :=εn+1k

εnk(3.50)

The spectral radius for each mode k, ρk, is defined as

ρk = |Gk| =∣∣∣∣ εn+1k

εnk

∣∣∣∣ . (3.51)

An iterative process will only converge if ρ < 1 for all k in the range (0 < α∆x ≤ π).

Substituting Equation 3.49 into Equation 3.51 shows ρk for the Jacobi method,

ρJk , to be

ρJk =

∣∣∣∣cos

(πk∆x

N

)∣∣∣∣ . (3.52)

Upon inspection, it is found that ρJk = 1 when k = 0 and k = N/∆x. Therefore,

Fourier stability analysis predicts the Jacobi iterative method will not converge

when attempting to solve the 1D Laplace’s equation due to the spectral radius

at k = N/∆x. The condition of k = 0 is never fully realized, in practice, as the

minumum value of k is determined by the discrete mesh spacing.

Page 82: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

60

3.5.1.2 Relaxed Jacobi Method

Although the classic Jacobi method is found inadequate to solve the 1D Laplace’s

equation, the method benefits from relaxation strategies to help convergence to-

wards an error-free solution.

The relaxed Jacobi method is implemented by redefining φn+1x to be

φn+1x = ω

(φnx+∆x + φnx−∆x

2

)+ (1− ω)φnx, (3.53)

where the relaxation factor, ω, is a positive non-zero real-valued number that serves

to weight the updated value of φx. The system is considered underrelaxed when

ω < 1 and overrelaxed when ω > 1. It can be seen that choosing ω = 1 reduces

the iterative process to the classic Jacobi method.

Applying Fourier stability analysis to this system, the spectral radius of the

relaxed Jacobi method, ρRJk , is found to be

ρRJk =

∣∣∣∣εn+1k

εnk

∣∣∣∣ =

∣∣∣∣ωcos(πk∆x

N

)− ω + 1

∣∣∣∣ . (3.54)

The spectral radius, ρRJk , has two values of interest. The first is k = 0 where

ρ = |ω − ω + 1| = 1, (3.55)

which shows there is no damping of the zero wavenumber. The other value of

interest is k = N/∆x where

ρ = |−ω − ω + 1| = |1− 2ω| . (3.56)

Fourier stability analysis predicts the relaxed Jacobi iterative method, when

attempting to solve the 1D Laplace’s equation, will converge towards an error-free

solution when underrelaxed but diverge when overrelaxed.

Page 83: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

61

3.5.1.3 Gauss-Seidel method

The Gauss-Seidel iteration procedure is similar to Jacobi, except that updated

values of φ are used as the solution progresses. This leads to φn+1x expressed as

φn+1x =

(φnx+∆x + φn+1

x−∆x

2

). (3.57)

Applying Fourier stability analysis to the Gauss-Seidel expression of φn+1x gives

εn+1k = εnk

(eiπk∆xN

2− e− iπk∆xN

), (3.58)

ρGSk =

∣∣∣∣εn+1k

εnk

∣∣∣∣ =

∣∣∣∣∣ eiiπk∆xN

2− e−i iπk∆xN

∣∣∣∣∣ =1√

5− 4 cos(iπk∆xN

) . (3.59)

which indicates the classic Gauss-Seidel method will converge when solving the 1D

Laplace’s equation since ρGSk < 1 for all values of necessary values of k.

3.5.1.4 Relaxed Gauss-Seidel Method

While the classic Gauss-Seidel method is shown to converge the 1D Laplace’s

equation, it is possible to increase the performance of the iterative process.

The spectral radius of an iterative procedure gives insight into not only solver

stability but also expected solver performance. Solver performance increases as

max(ρk) decreases for appropriate values of k. Section 3.6.1.2 explains the rela-

tionship between spectral radius and solver performance in greater detail.

As with the relaxed Jacobi method, φn+1x corresponding to the relaxed Gauss-

Seidel procedure is expressed as

φn+1x = ω

(φnx+∆x + φn+1

x−∆x

2

)+ (1− ω)φnx. (3.60)

Applying Fourier stability analysis to the relaxed Gauss-Seidel expression of φn+1x

gives

Page 84: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

62

εn+1k = εnk

(ωe

iπk∆xN + 2(1− ω)

2− ωe− iπk∆xN

), (3.61)

ρRGSk =

∣∣∣∣εn+1k

εnk

∣∣∣∣ =

∣∣∣∣∣ωeiπk∆xN + 2(1− ω)

2− ωe− iπk∆xN

∣∣∣∣∣=

√ω4 [sin(2θ)− 2 sin(θ)]2 + [ω2(cos(2θ)− 2 cos(θ)) + 4(ω − 1)]2√

[−4ω cos(θ) + ω2 + 4]2,

(3.62)

θ = α∆x =πk∆x

N. (3.63)

Equation 3.62 can be used to sweep through values of ω and minimize max(ρk).

Figure 3.4 shows the distribution of ρk as a function of θ for various values of ω

while Figure 3.5 shows max(ρk) as a function of ω.

Figure 3.5 shows that the minimum value of max(ρk), corresponding with the

best estimated solver performance, for the relaxed Gauss-Seidel method is obtained

near ω = 2. In practice, the optimal value of ω also depends on the smallest

wavenumber present in the system. On an evenly spaced grid with N + 1 grid

points, the smallest realized wavenumber corresponds to θ = π/N .

Page 85: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

63

Figure 3.4: Spectral radii of relaxed Gauss-Seidel applied to 1D Laplace’s equationas function of wavenumber

Page 86: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

64

Figure 3.5: Maximum spectral radius of relaxed Gauss-Seidel applied to 1DLaplace’s equation as function of relaxation factor. Spectral radii are obtainedby sampling ρk at wavenumber π/64.

Page 87: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

65

3.5.2 Analysis of 3D Steady Linear Elasticity Equations

The steady linear elastic equations in index notation are

∂σij∂xj

+ bi = 0, (3.64)

σij = µ

(∂ui∂xj

+∂uj∂xi

)+ λ

∂uk∂xk

δij. (3.65)

Assuming no body forces,

∂σij∂xj

= µ

(∂2ui∂xj∂xj

+∂2uj∂xi∂xj

)+ λ

∂2uk∂xi∂xk

= 0. (3.66)

Writing Equation 3.66 in x, y, and z dimensions, respectively, produces

µ

(∂2u

∂x2+∂2u

∂y2+∂2u

∂z2

)+ (µ+ λ)

(∂2u

∂x2+

∂2v

∂x∂y+

∂2w

∂x∂z

)= 0, (3.67)

µ

(∂2v

∂x2+∂2v

∂y2+∂2v

∂z2

)+ (µ+ λ)

(∂2u

∂x∂y+∂2v

∂y2+

∂2w

∂y∂z

)= 0, (3.68)

µ

(∂2w

∂x2+∂2w

∂y2+∂2w

∂z2

)+ (µ+ λ)

(∂2u

∂x∂z+

∂2v

∂y∂z+∂2w

∂z2

)= 0. (3.69)

Reorganizing the equations to combine all normal diffusion contributions shows

(2∂2u

∂x2+∂2u

∂y2+∂2u

∂z2

)+ λ

∂2u

∂x2

]+ (µ+ λ)

(∂2v

∂x∂y+

∂2w

∂x∂z

)= 0, (3.70)

(∂2v

∂x2+ 2

∂2v

∂y2+∂2v

∂z2

)+ λ

∂2v

∂y2

]+ (µ+ λ)

(∂2u

∂x∂y+

∂2w

∂y∂z

)= 0, (3.71)

Page 88: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

66

(∂2w

∂x2+∂2w

∂y2+ 2

∂2w

∂z2

)+ λ

∂2w

∂z2

]+ (µ+ λ)

(∂2u

∂x∂z+

∂2v

∂y∂z

)= 0. (3.72)

These equations can now be discretized using the relaxed Jacobi iterative

method then transformed into Fourier space. Equation 3.70 will be used to show

this process.

A first-order and second-order central finite difference can each be written as

∂φ

∂x≈ φx+∆x,y,z − φx−∆x,y,z

2∆x, (3.73)

∂2φ

∂x2≈ φx+∆x,y,z − 2φx,y,z + φx−∆x,y,z

∆x2. (3.74)

In the n-th iteration of the point Jacobi iterative method, φ at the point of

interest will be that of step n while all other values of φ will be obtained from the

previous Jacobi step. As such,

∂φ

∂x≈φnx+∆x,y,z − φnx−∆x,y,z

2∆x, (3.75)

∂2φ

∂x2≈φnx+∆x,y,z − 2φn+1

x,y,z + φnx−∆x,y,z

∆x2. (3.76)

For the cross diffusion terms, finite difference approximations can be written

as

∂2φ

∂x∂y≈φnx+∆x,y+∆y,z − φnx+∆x,y−∆y,z − φnx−∆x,y+∆y,z + φnx−∆x,y−∆y,z

4∆x∆y. (3.77)

Substituting the finite differences into Equation 3.70 produces

Page 89: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

67

[(2µ+ λ)

(unx+∆x,y,z − 2un+1

x,y,z + unx−∆x,y,z

∆x2

)+ µ

(unx,y+∆y,z − 2un+1

x,y,z + unx,y−∆y,z

∆y2

)+ µ

(unx,y,z+∆z − 2un+1

x,y,z + unx,y,z−∆z

∆z2

)]+ (µ+ λ)

(vnx+∆x,y+∆y,z − vnx+∆x,y−∆y,z − vnx−∆x,y+∆y,z + vnx−∆x,y−∆y,z

4∆x∆y

+wnx+∆x,y,z+∆z − wnx+∆x,y,z−∆z − wnx−∆x,y,z+∆z + wnx−∆x,y,z−∆z

4∆x∆z

)= 0.

(3.78)

Solving for un+1x,y,z and applying relaxation yields

un+1x,y,z = ω

{[(2µ+ λ)

(unx+∆x,y,z + unx−∆x,y,z

Ax∆x2

)+ µ

(unx,y+∆y,z + unx,y−∆y,z

Ax∆y2

)+ µ

(unx,y,z+∆z + unx,y,z−∆z

Ax∆z2

)]+ (µ+ λ)

(vnx+∆x,y+∆y,z − vnx+∆x,y−∆y,z − vnx−∆x,y+∆y,z + vnx−∆x,y−∆y,z

4Ax∆x∆y

+wnx+∆x,y,z+∆z − wnx+∆x,y,z−∆z − wnx−∆x,y,z+∆z + wnx−∆x,y,z−∆z

4Ax∆x∆z

)}+ (1− ω)unx,y,z,

(3.79)

Ax = 2

(2µ+ λ

∆x2+

µ

∆y2+

µ

∆z2

). (3.80)

The discrete Fourier transform for φ can be expressed as,

φnx,y,z → φnq,r,seiφqxeiφryeiφsz. (3.81)

For the sake of brevity while working with a multidimensional vector equation,

φnx,y,z will be denoted as the solution round-off error of φnx,y,z.

Applying the Fourier transform to Equation 3.78 and dividing through by

(eiφqxeiφryeiφsz) gives

Page 90: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

68

un+1q,r,s = ω

{unq,r,s

[(2µ+ λ)

(eiφq∆x + e−iφq∆x

Ax∆x2

)+ µ

(eiφr∆y + e−iφr∆y

Ax∆y2

)+ µ

(eiφs∆z + e−iφs∆z

Ax∆z2

)]+ (µ+ λ)

(vnq,r,s

eiφq∆xeiφr∆y − eiφq∆xe−iφr∆y − e−iφq∆xeiφr∆y + e−iφq∆xe−iφr∆y

4Ax∆x∆y

+ wnq,r,seiφq∆xeiφs∆z − eiφq∆xe−iφs∆z − e−iφq∆xeiφs∆z + e−iφq∆xe−iφs∆z

4Ax∆x∆z

)}+ (1− ω)unq,r,s.

(3.82)

Which can be cleaned up as

un+1q,r,s = Bxu

nq,r,s + Cxv

nq,r,s +Dxw

nq,r,s, (3.83)

Ax = 2

(2µ+ λ

∆x2+

µ

∆y2+

µ

∆z2

), (3.84)

Bx =ω

Ax

[2µ+ λ

∆x2

(eiφq∆x + e−iφq∆x

)+

µ

∆y2

(eiφr∆y + e−iφr∆y

)+

µ

∆z2

(eiφs∆z + e−iφs∆z

)]+ (1− ω),

(3.85)

Cx =ω(µ+ λ)

4Ax∆x∆y

[ei(φq∆x+φr∆y) − ei(φq∆x−φr∆y)

− e−i(φq∆x−φr∆y) + e−i(φq∆x+φr∆y)

],

(3.86)

Dx =ω(µ+ λ)

4Ax∆x∆z

[ei(φq∆x+φs∆z) − ei(φq∆x−φs∆z)

− e−i(φq∆x−φs∆z) + e−i(φq∆x+φs∆z)].

(3.87)

Page 91: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

69

Repeating this procedure for the y- and z-dimension produces

vn+1q,r,s = Byu

nq,r,s + Cyv

nq,r,s +Dyw

nq,r,s, (3.88)

wn+1q,r,s = Bzu

nq,r,s + Czv

nq,r,s +Dzw

nq,r,s, (3.89)

Ay = 2

∆x2+

2µ+ λ

∆y2+

µ

∆z2

), (3.90)

Az = 2

∆x2+

µ

∆y2+

2µ+ λ

∆z2

), (3.91)

By =AxCxAy

, (3.92)

Bz =AxDx

Az, (3.93)

Cy =ω

Ay

∆x2

(eiφq∆x + e−iφq∆x

)+

2µ+ λ

∆y2

(eiφr∆y + e−iφr∆y

)+

µ

∆z2

(eiφs∆z + e−iφs∆z

)]+ (1− ω),

(3.94)

Cz =AyDy

Az, (3.95)

Dy =ω(µ+ λ)

4Ay∆y∆z

[ei(φr∆y+φs∆z) − ei(φr∆y−φs∆z)

− e−i(φr∆y−φs∆z) + e−i(φr∆y+φs∆z)

],

(3.96)

Page 92: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

70

Dz =ω

Az

∆x2

(eiφq∆x + e−iφq∆x

)+

µ

∆y2

(eiφr∆y + e−iφr∆y

)+

2µ+ λ

∆z2

(eiφs∆z + e−iφs∆z

)]+ (1− ω).

(3.97)

These three equations can be expressed simultaneously as

Bx Cx Dx

By Cy Dy

Bz Cz Dz

unq,r,s

vnq,r,s

wnq,r,s

=

un+1q,r,s

vn+1q,r,s

wn+1q,r,s

.The stability of this scheme is determined by the spectral radius of matrix G.

G(φq, φr, φs) =

Bx Cx Dx

By Cy Dy

Bz Cz Dz

. (3.98)

The spectral radius is defined as the maximum eigenvalue of the linear system.

Figures 3.6 and 3.7 show contour plots of the spectral radius as a function of φy

and φz when φx = 0 and φx = π, respectively. Figure 3.8 shows spectral radius

as a function of wavenumber when φx = φy = φz. The spectral radii profiles show

that for 3D linear elasticity the maximum spectral radius value may not occur at

the smallest wavenumber when using the relaxed Jacobi method. As such, care

must be taken when searching for the maximum value of spectral radius for a given

relaxation factor as the maximum may occur at any wavenumber present in the

discretized system.

Page 93: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

71

Figure 3.6: Spectral Radius of relaxed Jacobi method applied to linear elasticityas function of φy and φz when φx ≈ 0.0

Page 94: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

72

Figure 3.7: Spectral Radius of relaxed Jacobi method applied to linear elasticityas function of φy and φz when φx = π

Page 95: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

73

Figure 3.8: Spectral radius of relaxed Jacobi method applied to linear elasticity asfunction of wavenumber when φx = φy = φz for various relaxation factors.

Page 96: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

74

3.5.3 Fourier Stability of Analysis of Multi-Step Solution

Procedures

Fourier stability analysis typically assumes a single-step iterative method, meaning

a single, constant relaxation factor is used throughout the entire iterative proce-

dure. Work has been done showing the benefit of using multi-step solution pro-

cedures. This section briefly discusses the extension of a typical Fourier stability

analysis to such a procedure.

Assume an iterative procedure with stability matrix being a function of the

relaxation parameter, g(ω).

Given a two-step procedure, the analysis would provide

g(ω1)unq,r,s = u′

q,r,s, (3.99)

g(ω2)u′

q,r,s = un+1q,r,s, (3.100)

g(ω2)g(ω1)unq,r,s = un+1q,r,s. (3.101)

The effects of the stability matrices are multiplicative and the generalized form

for an m-step is [m−1∏i=0

g(ωm−i)

]unq,r,s = un+1

q,r,s. (3.102)

This relation allows the term∏m−1

i=0 g(ωm−i) to be used in analyzing multi-step

stability. Furthermore, the relaxation factors used in the multi-step procedure can

be optimized to improve code performance (as discussed in Section 3.6.1.2).

Page 97: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

75

3.6 Miscellaneous High Performance Computing

Improvements

3.6.1 Linear Solver Performance Optimization

The possibility of multi-step Jacobi and Gauss-Seidel linear solvers opens up the

opportunity to explore the performance benefits associated with various sets of

relaxation factors for arbitrary numbers of steps. Given the nonlinear nature of

the physical system being solved, incremental improvements in linear solver per-

formance can have compounding effects on the overall performance of the compu-

tational tool developed through this work.

Two approaches to optimizing the linear solver performance are as follows:

1) an empirical-based approach that evaluates performance using the nonlinear

residual history and 2) a Fourier stability based approach that uses the spectral

radius of the numerical system to predict performance.

3.6.1.1 Empirically-based Performance Optimization

An initial attempt at optimizing linear solver performance is to find a combina-

tion of relaxation factors that minimize the slope of the solver error residual as a

function of iteration number. Figure 3.9 shows an example of such a residual plot.

As seen in the plot, the relationship becomes linear after many iterations. The

slope of this linear region can be computed and used as the cost function for the

optimization routine.

To help improve the performance of the optimization routine, radial basis func-

tions (RBFs) are used to develop a response surface of the function space. The

optimization routine is then used to find the global minimum of the response sur-

face. The optimal solution returned by the routine is evaluated using the linear

solver. The solution obtained from the linear solver is used to update the re-

sponse surface and re-run the optimization routine. This cycle is repeated until

the optimization routine converges on an optimal solution.

Page 98: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

76

Figure 3.9: Sample residual profile showing solution root mean square (RMS) erroras a function of solver iteration number

RBF response surfaces provide an approximation to the unknown function gov-

erning system stability as a function of relaxation factors. The resulting function

approximation will pass exactly through the points used to compute the curve pa-

rameters since RBFs assume observe points to be exact [108]. While there many

forms of RBFs, this work uses the multiquadric radial basis function expressed as

φ(r) =√

1 + (εr)2, (3.103)

Page 99: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

77

r = ||x− xi||, (3.104)

where ε is an adjustable constant set as the average norm between all observed

points, x is an arbitrary vector corresponding to a point in parameter space, xi is

the i-th observation point used to create the curve fit, and r is the norm computed

between x and xi.

Using the definition of φ, the curve-fit approximation is computed as

y(x)RBF, approx =N∑i=1

wiφi(||x− xi||) (3.105)

where N is the total number of observed points, φi is the RBF associated with the

i-th observation point, and wi is the weight associated with the i-th RBF.

The final curve-fit is a linear combination of the RBFs, φ, placed at each of the

observation points. This linearity allows the weights to be obtained using a linear

least-squares approach similar to that shown in Section B.1.

The linear least-squares approach allows the weights to be solved as

w = (XTX)−1XTyobserved, (3.106)

Xij =∂y(x)RBF, approxi

∂wj= φ(||xi − xj||). (3.107)

X will be a square matrix of rank m, where m is the number of observation points.

The performance benefit of using an RBF response surface as an approximation

of the response surface is in the ability to quickly evaluate function calls and the

possibility to analytically compute the gradient for use as input into the optimiza-

tion routine.

3.6.1.2 Fourier-based Performance Optimization

Let β0 be the maximum spectral radius of the a given solver configuration 0. It is

possible to compute the speedup, s, of any other solver configuration i as follows

Page 100: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

78

s =ln(βi)

ln(β0). (3.108)

This equation is obtained from the relationship between spectral radius of sta-

bility matrix and solution error.

Beginning with the relationship

∣∣∣∣rn+1

rn

∣∣∣∣ = β, (3.109)

where rn is the error residual at n-th iteration of the linear solver. Given the

iterative nature of this equation, a multiplicative relationship can be shown to be

∣∣∣∣rn+α

rn

∣∣∣∣ = βα. (3.110)

Now assume that in reducing the residuals by a factor of 10 takes α/c and α

iterations for models 0 and a, respectively, the corresponding spectral radii are

related as

0.1 = βα/c0 = βαa . (3.111)

Taking the natural logarithm of each side produces

α

cln(β0) = αln(βa), (3.112)

which is simplified as

s =1

c=

ln(βa)

ln(β0), (3.113)

This analysis does assume, however, that each iteration requires the same amount

of computational work.

The ability to compute code speedup using spectral radii allows for the possibil-

ity of devising an optimization routine to maximize code performance, particularly

for multi-step solution procedures. It is possible to find a set of relaxation factors,

Page 101: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

79

for a given multi-step procedure, that minimizes the spectral radius and, in turn,

maximizes code performance of the solver. However, developing such an optimiza-

tion scheme is no trivial task.

The first step in developing an optimization routine for multi-step solution pro-

cedures is obtaining the spectral radius for an arbitrary set of relaxation factors.

There has been some work done in finding spectral radii and optimal relaxation

factors for linear solvers of elliptic PDE systems [109, 110]. While these studies

provided analytic approaches to the system optimization, the analysis was per-

formed for scalar PDEs in 1D and 2D. The work being presented deals with 3D

vector PDEs of elliptic and parabolic nature.

Section 3.5.2 shows the Fourier stability analysis of the static 3D linear elasticity

equations and presents an expression for the stability matrix G as a function of the

wavenumber in each spatial dimension. Any arbitrary combination of wavenumbers

will produce three eigenvalues for the 3x3 matrix G. The spectral radius of the

system is the maximum eigenvalue of all possible values for matrix G with φ in

the range [φmin, π].

One approach to obtaining the spectral radius of G is to sweep through possible

values of φ to find the maximum system eigenvalue; this approach, however, is time

consuming and computationally inefficient.

A second approach is to analytically derive the maximum eigenvalues by finding

values of φ that satisfy the condition

∂G

∂ei= 0, (3.114)

where ei is the i-th system eigenvalue.

A third approach is to use a numerical optimization routine to find the maxi-

mum system eigenvalue. Using a metaheuristic optimization approach (e.g., tabu

search, simulated annealing, particle swarm) it may be possible to quickly locate

the global minimum. This third approach will be used in this work due to its ease

of implementation.

The second, and final, step in developing an optimization routine for multi-

Page 102: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

80

step solution procedures is finding the set of relaxation factors that minimize the

spectral radius of matrix G, thus increasing speedup. This step of the routine

will use the subroutine developed in step one to compute the spectral radii, then

compute speedup by using Equation 3.108 and the spectral radius of a reference

solver.

For illustrative purposes, this optimization routine was performed on a two-step

solver using the model system described in Section 3.5.2.

Figure 3.10 shows a plot of the computed performance increase for a given pair

of relaxation factors; any combination of relaxation factors predicted to diverge

is prescribed a speedup value of zero. All remaining speedup values are made

negative so that global minimum of the system corresponds to maximum predicted

speedup. The speedup values in this plot are obtained using a simulated annealing

optimization routine.

Page 103: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

81

Figure 3.10: Estimated performance increase of two-step relaxed Jacobi methodapplied to 3D linear elasticity equations. Two-step methods show the potential of78% increase in performance over the single step relaxed Jacobi method.

Page 104: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

Chapter 4Verification, Validation, and Results

4.1 Fixed Shape Leukocyte Rolling

The PMN rolling algorithm presented in Section 3.3.1 allows the model PMN to

move along the near-wall region in a manner that captures complex flow-structure-

biochemistry interactions.

Figures 4.1, 4.2, and 4.3 show the trajectory of a point on the PMN surface. The

PMN translational and rotational velocities are set to 120 µm/µs and 120 rad/µs,

respectively. Assuming a perfect sphere in pure rolling leads to both velocities

having the same magnitude equal to that of the shear rate. The trajectory plots

do, however, show some slipping at wall-adjacent points. The slippage is caused

by the analytic transform.

The oscillation of the surface point is indicative of the PMN surface rotating

about the body centroid.

For this case, the centroid height on the undeformed sphere is set to zero and

the PMN centroid is moving in the positive-x direction. Figure 4.1 shows the path

of the point throughout the entire simulation. The point path is smooth and much

longer on the top of the PMN than the bottom. Figure 4.2 shows the point does

spend more time in forward motion than backwards motion. Figure 4.3 shows

the point does spends equal time in upward motion and downward motion. The

smooth trajectory and sustained time along the top of the PMN allows for stable

Page 105: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

83

Figure 4.1: Trajectory of point on the surface of rolling PMN.

localized modeling of biochemistry.

Applying the rolling algorithm to every point belonging to the PMN compu-

tational mesh enables mesh motion while retaining the mesh topology needed to

accommodate localized modeling of biochemical bonds.

Page 106: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

84

Figure 4.2: X coordinate of surface point vs time during PMN rolling.

Page 107: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

85

Figure 4.3: Y coordinate of surface point vs time during PMN rolling.

Page 108: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

86

4.2 Finite-Volume Structural Mechanics

4.2.1 3D Linear Elastostatics

Linear elastostatics, while ill-suited for cellular FSBI problems in the microvascula-

ture, provides great insight into the ability to solve structural mechanics problems

using a baseline solver designed for fluid mechanics. Testing of the linear elastic-

ity implementation is done using classical beam problems for which linear elastic

analytic solutions exist.

4.2.2 Stretching of Prismatic Beam under Self-Weight

Consider a 3D rectangular beam as seen in Figure 4.4. With gravity acting in the

positive-z direction, the beam will experience the body force b = [0, 0, ρg]; where

ρ is the material mass density and g is the acceleration due to gravity.

For the prismatic beam test cases, notional properties of steel are used in the

simulations; modulus of Elasticity, E, set to 200 GPa and Poisson’s ratio, ν, set to

0.3. The beam dimensions are set to 0.3 meters, 0.3 meters, and 2.4 meters in the

x, y, and z directions, respectively. Grid spacing is chosen to be uniform in every

direction.

In this configuration, the boundary conditions are

σz(x, y, L) = 0,

σx = σy = 0,

τxy = τyz = τzx = 0.

(4.1)

Using the direct integration approach, the stress field in the rectangular beam

is found to be

Page 109: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

87

Figure 4.4: Pictorial representation of prismatic beam stretching under self-weight.

σz(x, y, z) = ρg(L− z),

σx = σy = 0,

τxy = τyz = τzx = 0.

(4.2)

Assuming symmetry about the surface at (z = 0) and no rigid-body motion (i.e.,

center of beam does not move and axes does not rotate) at point C (x = y = z = 0)

gives the solution

u(x, y, z) =νρgx

E(z − L),

v(x, y, z) =νρgy

E(z − L),

w(x, y, z) = − ρg2E

[z2 − 2Lz + ν(x2 + y2)

].

(4.3)

The results of this test case are shown in Figure 4.5. Necking of the beam does

Page 110: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

88

occur at z = 0 which is consistent with the symmetry about the z = 0 surface.

Grid convergence results using the l2-norm of σzz for this case are shown in

Figure 4.6. The lp-norm on an arbitrary function a is described in Equation 4.4 for

an arbitrary value of p. The order of convergence for the linear elasticity solver was

found to be 1.45. While the discrete operators described in Section 3.1 are second-

order accurate on a block mesh with uniform grid spacing, the boundary condition

implementation is less than second-order. Proper implementation of second-order

accurate boundary conditions will then return a second-order accuracy for the

entire solver.

lp =(∑

(|aobserved − aexact|))1/p

(4.4)

Figure 4.5: Deformation of bar due to self weight.

Page 111: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

89

Figure 4.6: Grid convergence plot for stretching of beam due to self weight.

4.2.3 Flexure of Prismatic Beam Due to End Loading

Given the same geometry and material properties as the self-weight case, consider

a 3D rectangular beam subject to end loading T as shown in Figure 4.7. T = 90N

is applied at the shear center of the (z = L) cross-section to ensure the body is

not subject to any torsion.

In this configuration, the boundary conditions are

Page 112: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

90

Figure 4.7: Pictorial representation of prismatic beam experiencing flexure due toend loading T .

u(x, y, 0) = v(x, y, 0) = w(x, y, 0) = 0,

by(0, 0, L) =T

dV=

T

∆x∆y∆z,

σz(x, y, L) = 0,

σx(x, y, z) = σy(x, y, z) = 0,

τxy(x, y, z) = 0.

(4.5)

Assuming a harmonic stress function, as presented by Sadd [111], the stress

field in the rectangular beam is found to be

Page 113: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

91

τxz(x, y, z) =νA2T

2(1 + ν)π2Ix

∞∑n=1

(−1)n

n2

sin2nπxA

sinh2nπyA

coshnπBA

,

τyz(x, y, z) =T

2Ix

(B2

4− y2

)+

νT

6(1 + ν)Ix

[3x2 − A2

4− 12A2

4π2

∞∑n=1

(−1)n

n2

cos2nπxA

cosh2nπyA

coshnπBA

],

σz(x, y, z) = − TIxy(L− z),

Ix =AB3

12.

(4.6)

where Ix is the area moment of inertia of the cross-sections normal to the z-direction

and T is the magnitude of the end loading. More detailed description of the analytic

solution technique can be found in Sadd [111].

The results of this test case are shown in Figure 4.8. The prismatic beam flexes

as expected. Grid convergence results using the l2-norm of σzz for this case are

shown in Figure 4.9 and reproduce the 1.45 order of convergence shown by the

self-weight case.

Page 114: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

92

Figure 4.8: Deformation of bar due to flexure traction loading at bar end.

Page 115: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

93

Figure 4.9: Grid convergence plot for flexure of beam due to end loading.

Page 116: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

94

4.2.4 Saint-Venant Kirchhoff Hyperelasticity

The implementation of hyperelastic structural mechanics is built upon the linear

elastic functionality. Addition of the explicitly-treated nonlinear terms arising from

the SVK constitutive model can be tested using steady rigid body rotation cases.

4.2.4.1 Hyperelastic Sphere undergoing Rigid Body Rotation

Consider a 3D sphere centered at c=[cx, cy, cz] as shown in Figure 4.10.

Figure 4.10: Surface mesh of sphere.

A rotation θ is applied at the sphere center about an axis pointing the positive

z-direction passing through the sphere center. The displacement of any point on

the body due to the rotation θ is

Page 117: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

95

u

v

w

=

(cos θ − 1) − sin θ 0

sin θ (cos θ − 1) 0

0 0 0

x− cx

y − cy

z − cz

. (4.7)

Equation 4.7 can be used to show E′′ij to be

E′′

ij = 0.5(F′′

kiF′′

kj − δij) = 0. (4.8)

It can also be shown that any hyperelastic strain measure will be identically

zero for any steady rigid body rotation. As such, this case is used to ensure the

correct implementation of hyperelastic constitutive models.

The displacements computed using Equation 4.7 are set as the boundary con-

ditions at the surface of the sphere. The l2-norm of the stress field computed in

the body was zero for arbitrary values of θ.

4.3 Single Body Simulations

A number of single body cases are presented using scales relevant to the microvas-

cular environment.

4.3.1 Impulsively Started Rigid TC in Uniform Flow

An important case to test the semi-implicit rigid body 6DOF motion solver is an

impulsively started rigid body in uniform low Reynolds number flow, as shown in

Figure 4.11.

The flow-induced drag on a rigid TC in low Reynolds number uniform flow is

computed using Stokes’s law

F (t)stokes = 6πµr[uflow − u(t)TC ], (4.9)

where µ is the dynamic viscosity of the fluid, r is the sphere radius, and uflow is

Page 118: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

96

Figure 4.11: Pictorial representation of computational domain and boundary con-ditions for rigid TC in uniform flow.

the velocity of the uniform flow field. Solving Newton’s law of motion gives the

expected velocity profile for the rigid sphere

du(t)TC

dt=

6πµr

ρV[uflow − u(t)TC ], (4.10)

∴ u(t)TC = uflow(

1− e−6πµrtρV

). (4.11)

where ρ and V are the density and volume of the TC, respectively.

The time to steady-state can be found using Equation 4.11 solving for t0.99

where [u(t0.99)TC/uflow] = 0.99

t0.99 =−ρV6πµr

ln(0.01), (4.12)

Setting µ = 1 pgµm·µs , ρ = 1 pg

µm3 , and r = 8µm, time to steady-state is found

to be t0.99 ≈ 65.5µs. The value of t0.99 is less than the stability limit of the

Page 119: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

97

semi-implicit 6DOF solver allowing timestep to be chosen based solely on desired

temporal resolution.

This case was run using the presented FSBI solver with UInf = 1.4× 10−4 µmµs

.

Figure 4.12 shows the flow domain at t = 10µs. Figure 4.13 shows the 6DOF

solver results alongside the profile given by Equation 4.11. The two trajectories

are seen to be in agreement.

Figure 4.12: Angled view of TC and slice of flow field along the centerline of thecomputational domain at t = 10µs. Slice is colored by velocity magnitude andflow is in the +x direction. Gradients in the velocity field indicate the TC has notyet reached Uinf

Page 120: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

98

Figure 4.13: Analytic and computed velocity profiles of rigid TC in uniform flowas function of time.

4.3.2 Rigid TC in Linear Shear Flow

Another case to explore the ability of the semi-implicit 6DOF solver is a rigid body

in linear shear flow, as shown in Figure 4.14.

The linear shear flow will induce translation and rotation on the TC. In fact,

the TC will generate lift in the direction of increasing velocity due to the shear

[112,113].

Setting µ = 1 pgµm·µs , ρ = 1 pg

µm3 , r = 8µm, and γ = 50s−1, this case was simulated

using the presented FSBI solver. The flow field computed by the simulation is

Page 121: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

99

Figure 4.14: Pictorial representation of computational domain and boundary con-ditions for rigid TC in linear shear flow.

shown in Figure 4.15. The results of this case are shown in Figures 4.16 and

4.17. This case makes use of the body cyclic conditions proposed in [29]. The

vertical dashed lines in Figure 4.16 denote the end of the discretized computational

domain. The cyclic conditions allow for an effective flow length of 480µm despite

using a computational domain of length 120µm. The case ended when the rigid

TC collided with the top wall of the computational domain. The effects of the top

wall can be seen in Figure 4.17 by the increase in signal noise towards the end of

the simulation.

Page 122: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

100

Figure 4.15: Angled view of TC and slice of flow field along the centerline of thecomputational domain at t = 1ms. Slice is colored by velocity magnitude and flowis in the +x direction. Gradients in the velocity field are due to the prescribedflow shear rate.

Page 123: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

101

Figure 4.16: Trajectory of TC centroid in linear shear flow. Dashed vertical linesrepresents the end of the modeled computational domain. Cyclic conditions areused to increase the effective flow length.

Page 124: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

102

Figure 4.17: Velocity profile of TC centroid in linear shear flow. Simulation endsas TC collides into the top wall of the computational domain. Wall effects appearin the profile as signal noise towards the end of the simulation.

Page 125: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

103

4.3.3 Rigid Helical Microswimmer

One exciting application of this work is the simulation of rigid helical microswim-

mers. Much experimental work has been done on the design, fabrication, and

actuation of helical microswimmers [88,90] which can be enhanced by appropriate

modeling techniques.

The presented FSBI solver is well-suited to simulate helical microswimmers

and demonstrate this ability. Two simulation are performed for a rigid helical

microswimmer was placed in a quiescent fluid. In the first simulation, a constant

torque due to the rotation of a magnetic field aligned with the helical axis is applied

to the microswimmer. In the second simulation, the microswimmer is subject to a

constant angular velocity. Flow is induced by the microswimmer due to the no-slip

condition at the fluid-solid interface. The parameters used are listed in Table 4.1

and Figure 4.19 shows the resulting microswimmer. Effects of frequency step-out,

loss of propulsion due to the microswimmers inability to remain in sync with the

input signal, are ignored in these simulations.

Page 126: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

104

Figure 4.18: Schematic of helical microswimmer with labeled geometric entities.Helical geometries are defined by helix angle (θ), helix radius (R), filament radius(r), helix pitch (λ), and number of turns (n)

Page 127: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

105

Figure 4.19: Mesh of the helical microswimmer used in the simulations performedin this work. Geometric properties of this helix are found in Table 4.1

Page 128: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

106

Figure 4.20: Mesh of the helical microswimmer and slice of the meshed fluid domainused in the simulations performed in this work. Points on the interface of the fluidand helix domains match exactly as conformal meshing is used at the interface.

Page 129: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

107

4.3.3.1 Constant Torque

In this simulation, a constant torque per unit volume of 1.95× 10−2pg/µm/µs2 is

applied to a rigid helical microswimmer. The model presented by Abbott [90] pro-

vides an expected relationship between angular velocity and translational velocity,

which is expressed in Equation 4.13. This expected relationship is obtained using

Purcell’s helical motion propulsion matrix [114] and Lighthill’s approximation of

drag on thin rigid helical bodies [115],

u =−baω, (4.13)

a = 2πnR

(γp cos2(θ) + γn sin2(θ)

sin(θ)

), (4.14)

b = 2πnR2(γp − γn) cos(θ), (4.15)

γp =2πµ

ln(

0.36πRr sin θ

) , (4.16)

γn =4πµ

ln(

0.36πRr sin θ

)+ 0.5

. (4.17)

With constant torque acting on the body, the expected angular and transla-

tional velocity profiles are

dω(t)

dt=TaxialIaxial

(4.18)

Table 4.1: Properties of Helical Microswimmer

Pitch, λ [µm] 2Filament Radius, r [µm] 0.1Helix Angle, θ [rad] tan−1(π)Helix Radius, R [µm] 1Number of Turns, n [-] 5

Page 130: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

108

Figure 4.21: Helix in flow and slice of flow field along the centerline of the compu-tational domain at t = 97µs. Slice is colored by the z-components of velocity. Redindicates flow coming out of the page and blue indicated flow going into the page.

ω(t) =TaxialIaxial

t (4.19)

Setting µfluid = 1 pgµm·µs , ρfluid = 1 pg

µm3 , the computed trajectory of the mi-

croswimmer centroid is shown in Figure 4.22 and the flow field computed by the

simulation is shown in Figure 4.21. The velocity profile computed using the pre-

sented FSBI solver gives a linear profile which favorably agrees with the Abbott

model expectations. The profile slopes describe the acceleration of the body. The

ratio of the computed acceleration and the Abbott model expectation is approxi-

mately 1.45. This deviation can likely be attributed to the Lighthill drag approx-

imation.

Page 131: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

109

Lighthill approximated the viscous drag on a helix assuming stokeslets along

the helical path. This approach may not accurately capture the effect a helical

segment may have on the effective drag of a near-by segment. More work can be

done to explore potential relationships between the magnitude of deviation from

theory and helical geometric parameters.

Figure 4.22: Comparison of computed axial velocity profile and predictions basedon the Abbott model for a helical microswimmer undergoing constant torque.These models are in agreement with a slight difference in the profile slope. Thisdifference is likely caused by the idealized viscous drag approximation in the Abbottmodel.

Page 132: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

110

4.3.3.2 Constant Angular Velocity

In this simulation, a rigid helical microswimmer is subjected to a constant angular

velocity of 10 revolutions per second about the microswimmer’s helical axis. Using

the same fluid and geometric properties as the constant torque simulation, the

computed normalized velocity profile of the microswimmer is shown in Figure

4.23. The velocity profile shown is normalized by the steady-state axial velocity

predicted by the Abbott model. The FSBI solver found the normalized steady-

state velocity to be approximately 0.86. As with the constate torque case, the

velocity profile’s deviation from the value predicted by the Abbott model is likely

attributed to the drag approximation used by Abbott.

Page 133: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

111

Figure 4.23: Comparison of computed normalized axial velocity profile for a helicalmicroswimmer undergoing constant angular velocity. Velocity is normalized by theAbbott model prediction. The steady-state velocity asymptotes to approximately0.86.

4.3.4 Magnetically Actuated Hyperelastic Microbead in Cou-

ette Flow

Another area of interest is the design and modeling of magnetically actuated de-

formable DDMRs.

For this case, an elastic body was placed near a wall with the nominal PMN

shape presented in Section 3.3.1, as show in Figure 4.24.

Page 134: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

112

Figure 4.24: Pictorial representation of computational domain for wall-adjacenthyperelastic microbead in Couette flow.

The FSI traction interface condition was enforced on the entire body surface.

The body was tethered in place using a volumetric body force representative of

magnetic forces. The tethering body force per unit volume is described in Equation

4.20 and applied to each polyhedron inside the deformable body.

bmagi =1

V body,tot

∫S

(µ∂ui∂xj− pfaceδij

)dSj, (4.20)

where the integral is computed at the body surface, ui is the i-th component of

the flow velocity, pface is flow pressure at the surface, µ is the fluid viscosity, and

V body,tot is the total body volume.

This case was run with a timestep of ∆t = 10µs, modulus of Elasticity of

E = 0.504Pa, Poisson’s ratio of ν = 0.133, flow shear rate of 50s−1, the un-

deformed microbead is sphere of radius 4µm, the microbead shape is initialized

using the PMN transform described in Section 3.3.1, and the centroid of the trans-

formed microbead is set to a height of 2.25µm. Figure 4.25 shows the equilibrium

Page 135: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

113

microbead shape and a slice of the flow field over the body along the center of the

computational domain. Figure 4.26 shows a comparison of the body shapes and

location at various points in time during the simulation. The final shape has a flat

bottom surface and has evolved to an equilibrium shape best suited for this flow

region.

In addition to obtaining the equilibrium shape, this case provides a glimpse

into the development of active controllers for microbead design. The body force

obtained from Equation 4.20 can be used with Equations 2.73 and 2.74 to esti-

mate the required magnetic field needed to hold the bead in place. Furthermore,

Equations 4.20, 2.70, 2.71, 2.73, and 2.74 can be used together to determine the

field needed to move the bead along a prescribed trajectory.

Figure 4.25: Side view of the hyperelastic microbead and slice of flow field along thecenterline of the computational domain. The microbead has reached its equilibriumshape. Flow is in the +x direction

Page 136: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

114

Figure 4.26: Comparison of initial and final shape of wall-adjacent magnetic mi-crobead in linear shear flow. The final shape has a flat bottom surface and hasevolved to an equilibrium shape best suited for this flow region. Flow is from leftto right.

4.4 Multiple Body Simulations

4.4.1 Free-Flowing Rigid TC and Wall-Adhered Hypere-

lastic PMN in Couette Flow

The FSBI solver presented has been developed to handle an arbitrary number of

bodies with differing structural mechanics treatments. This capability is demon-

strated by using the unified solver to simultaneously model a rigid TC and hyper-

elastic PMN in Couette flow in the same computational domain.

Page 137: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

115

A wall-adhered PMN is placed in the near-wall region and modeled using the

SVK hyperelastic constitutive model. A rigid TC is placed upstream of the PMN.

Figure 4.27 shows a pictorial representation of the computational domain.

Figure 4.27: Pictorial representation of computational domain for free-flowing rigidTC and wall-adhered hyperelastic PMN in Couette flow.

For simplicity, zero-displacement Dirichlet conditions are enforced on PMN

faces near the wall and the FSI traction matching interface condition is enforced

on all other PMN faces. The TC trajectory is solved using the 6DOF solver.

The zero-displacement condition on the near-wall PMN faces, while non-physical,

allows for an exploration of the hyperelasticity solver’s ability without requiring

the fine mesh resolution necessary to resolve the flow between the wall and fixed

body.

Figure 4.28 shows the cell shapes and positions at various timepoints through-

out the simulation. The rigid TC translates and rotates as expected in shear flow.

The PMN deforms in a shearing motion and begins to exhibit signs of rolling along

the endothelial surface. Figure 4.29 gives a side view of the computational domain

Page 138: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

116

at t = 9, 320µs to better show the computed shapes.

Figure 4.28: Rigid TC and wall-adhered hyperelastic PMN at various times duringthe simulation. The bodies are colored by magnitude of displacement.

More work is needed to fully capture the PMN rolling motion using this FSBI

formulation. One approach is to use the localized biochemistry model presented

in Chapter 2 to solver for all interactions between wall-adjacent PMN and the

Page 139: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

117

Figure 4.29: Rigid TC, wall-adhered hyperelastic PMN, and slice of flow fieldalong centerline of computational domain at t = 9, 320µs. The slice is colored bymagnitude of flow velocity.

endothelial wall. This approach, however, may significantly increase the compu-

tational cost of these simulations as the mesh on the surface of the endothelial

wall would need to be refined for adequate spatial resolution of the biochemical

interactions. Another approach is to develop a different biochemistry model that

is not dependent on the resolution of the computational mesh at the wall (e.g., a

biochemistry model based primarily on distance from the wall). Either approach

may allow wall-adjacent PMN to roll along the endothelial wall. Nonetheless, this

dissertation provides much of the framework needed to model that phenomenon.

Page 140: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

118

4.4.2 Rigid Two-cell Aggregate Formation Simulations

After demonstrating all of the new functionality implemented into the CellCFD-

PSU framework, the rigid two-cell ADH1 case from [29] was re-run to ensure the

new FSBI solver could reproduce the results.

In this case, a free-flowing rigid TC interacts with a wall-adhered rigid PMN.

An initial collision occurs between the bodies, followed by a period of transient

biochemical interactions, and finally aggregate formation. An aggregate is declared

to be formed once the TC has no motion relative to the fixed PMN.

The parameters used to initialize the system are found in Table 4.2. The PMN

rolling model was used with translational and rotational velocities set to 0 µm/µs

and 0 rad/µs, respectively, to fix the PMN at the wall. A simple moving average

using 75 sample points was used to smooth the velocity signal in post-processing

to better show when aggregate formation occurs.

Table 4.2: ADH1 Case Parameters

Parameter Value

Equilibrium Dissociation Rate, k0off [s−1] 0.3

Equilibrium Association Rate, k0on [s−1] 3× 10−3

Equilibrium Spring Length, λ [µm] 0.05

Flow Shear Rate, γ [s−1] 150

Spring Constant, s [N/m]: 2× 10−3

Transition Spring Constant, sts [N/m]: 1× 10−3

TC Molecular Surface Density (ICAM-1), nICAML [N/m2]: 13× 10−12

PMN Molecular Surface Density (LFA-1), nLFAL [N/m2]: 13× 10−12

Domain Size [µm]:X 60Y 32Z 42

Tumor Cell Initial Centroid [µm]:X 18Y 10Z 21

PMN Initial Centroid [µm]:X 30Y 2.5Z 21

Page 141: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

119

Figure 4.30 shows the results of the simulation. The results first show a sharp

decrease in TC velocity indicative of a collision event. During this collision, there

is some biochemical interacting shown by small amounts of bond formation and

breakage. TC velocity begins to increase as the cell rolls over the PMN surface

due to fluid forcing.

Figure 4.30: TC velocity and number of bonds as a function of time. Initial dropin velocity indicates collision between TC and PMN. TC-PMN aggregate formswhen TC velocity reaches zero.

As the TC rolls over the PMN, significant biochemical interaction takes place

and formation of long lasting bonds occur. An interaction threshold is eventually

surpassed and the TC velocity once again decreases until the TC has no motion

Page 142: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

120

relative to the PMN. The TC-PMN aggregate is formed once the TC velocity

reaches zero. The TC velocity does slightly oscillate about zero once the aggregate

has been formed. This velocity oscillation can be attributed to the continued

formation and breakage of individual bonds and fluid forcing acting on the TC.

Page 143: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

Chapter 5Conclusions and Future Work

5.1 Conclusion

A computational tool has been developed to model flowing cellular systems and

has been applied to direct numerical simulation of microvascular flows with a vision

towards personalized medicine. This tool couples CFD, computational structural

mechanics, 6DOF motion, and surface biochemistry, in the context of interface-

resolved cell geometry, to provide a detailed model of the heterogeneous blood

flow microenvironment. This tool can be used to study drug-mediated cellular

interactions in the vasculature and to design magnetically actuated DDMRs with

targeting capabilities.

Structural mechanics models were introduced, along with accompanying FSBI

coupling conditions for each of the models. The semi-implicit 6DOF motion solver

uses the iterative nature of many flow solvers to increase stability of the motion

solver. The iterative implicit procedure developed allowed for orders-of-magnitude

speedup over explicit 6DOF motion solvers. The PMN rolling model provides

a computationally straightforward approach to capture complex flow-structure-

biochemistry interactions of wall-adhered PMNs. The rolling model can be tuned

to match experimentally observed phenomena. The finite-volume discretization of

the hyperelastic Saint-Venant Kirchhoff constitutive model allows for large defor-

mation modeling of arbitrary bodies in microvascular flow. This hyperelasticity

Page 144: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

122

implementation shows it is possible to model systems including both large strains

and large rigid body motions. Both structural mechanics models were implemented

into an existing flow-solver showing the ability to backfit other existing codes with

structural modeling capability.

DDMR designs were simulated using this tool, showing great promise for the

future of DDMR computational analysis. A rigid helical microswimmer was placed

in a quiescent fluid and actuated with both a prescribed torque and a prescribed

angular velocity, reproducing the effects of a rotating magnetic field. The results

of the helical microswimmer simulation were in agreement with theoretical predic-

tions. A hyperelastic microbead was placed in the near-wall region of a Couette

flow and held in place with a volumetric body force reproducing the effects of an

applied magnetic field. The microbead evolved to an equilibrium shape due to the

applied fluid forcing and volumetric body force.

Multiple linear solver analysis methods were presented. Fourier stability analy-

sis (FSA) allows for performance evaluations of computational systems with apriori

knowledge of the underlying governing equations and discretization strategy. An

empirically-based performance optimization routine was also presented requiring

no knowledge of the underlying governing equations or discretization strategy.

5.2 Future Work

Various aspects of this research can benefit from further development.

One area of potential benefit is exploring ways to fully capture the PMN rolling

motion using the hyperelasticity structural modeling implementation. Figure 4.28

shows the onset of PMN rolling, however, more work is needed to better model

the biochemical interactions between the PMN and the endothelium wall. Use of

the localized biochemistry model would require greatly increased spatial resolution

along the endothelial wall and smaller timesteps to ensure the bond physics are

adequately resolved. Such an increase would significantly impact runtime of the

computational tool and likely render runtimes intractable.

Page 145: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

123

A second area for further exploration is the implementation of hyperelastic

models that do not depend on the slow-process assumption. More general hypere-

lastic models would have the potential advantage of better stability of the physical

system and may allow for arbitrarily large timesteps. However, rearranging the

general hyperelastic governing equations into the unified equation form presented

in Section 2.2.4 is not immediately obvious.

A third area of exploration is using this tool to better understand swarm con-

trol of DDMRs in microvascular environments. It is possible to determine the

magnetic actuation force and torque needed to move a DDMR along a prescribed

trajectory using Equations 4.20, 2.70, and 2.71. Equations 2.73 and 2.74 can

then be used to solve the inverse problem of determining the necessary magnetic

field. Furthermore, solving the inverse magnetic field problem provides a path to-

wards the development of active controllers capable of providing localized DDMR

instructions using a single global signal.

Page 146: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

Appendix A

A.1 SI Scaling of Microvascular System

Mass: picogram [pg]

Length: micrometer [µm]

Time: microsecond [µs]

1 picogram = 10−12 grams = 10−15 kilograms

1 micrometer = 10−6 meters

1 microsecond = 10−6 seconds

Velocity:

1[ms

]=

(1m · 106µm

1m

)(1

1s· 1s

106µs

)= 1

[µm

µs

](A.1)

Density:

1

[kg

m3

]=

(1kg · 103g

1kg· 1012pg

1g

)(1

1m· 1m

106µm

)3

=1015

1018

[pg

µm3

]= 10−3

[pg

µm3

] (A.2)

Page 147: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

125

Viscosity:

1

[kg

m · s

]=

(1kg · 103g

1kg· 1012pg

1g

)(1

1m· 1m

106µm

)(1

1s· 1s

106µs

)=

1015

1012

[pg

µm · µs

]= 103

[pg

µm · µs

] (A.3)

Force (N):

1

[kg ·ms2

]=

(1kg · 103g

1kg· 1012pg

1g

)(1m · 106µm

1m

)(1

1s· 1s

106µs

)2

=1021

1012

[pg · µmµs2

]= 109

[pg · µmµs2

] (A.4)

Pressure (Pa):

1

[kg

m · s2

]=

(1kg · 103g

1kg· 1012pg

1g

)(1

1m· 1m

106µm

)(1

1s· 1s

106µs

)2

=1015

1018

[pg

µm · µs2

]= 10−3

[pg

µm · µs2

] (A.5)

Surface Density:

1

[molecule

m2

]= (1 molecule)

(1

1m· 1m

106µm

)2

= 10−12

[molecule

µm2

](A.6)

Association Constant:

1

[1

M · s

]= 1

[liter

mole · s

]= 10−3

[m3

mole · s

]= 10−3

(1m · 106µm

1m

)3(1

1s· 1s

106µs

)(1

1 mole

)=

1015

106

[µm3

mole · µs

]= 109

[µm3

mole · µs

] (A.7)

Dissociation Constant:

1

[1

s

]=

(1

1s· 1s

106µs

)= 10−6

[1

µs

](A.8)

Page 148: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

126

Spring Constant (N/m):

1

[kg

s2

]=

(1kg · 103g

1kg· 1012pg

1g

)(1

1s· 1s

106µs

)2

=1015

1012

[pg

µs2

]= 103

[pg

µs2

] (A.9)

Base SI units of Boltzmann Constant (J/K):

1

[m2 · kgK · s2

]=

(1kg · 103g

1kg· 1012pg

1g

)(1m · 106µm

1m

)2(1

1s· 1s

106µs

)2(1

K

)=

1027

1012

[µm2 · pgK · µs2

]= 1015

[µm2 · pgK · µs2

](A.10)

Therefore, Boltzmann Constant:

1.38064852× 10−23

[m2 · kgK · s2

]= 1.38064852× 10−8

[µm2 · pgK · µs2

](A.11)

Page 149: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

Appendix B

B.1 Gradient Reconstruction using Least Square

Optimization

B.1.1 Derivation

In finite-volume formulations, it is often necessary to reconstruct the gradient

field of a scalar. One approach to doing this is using least-squared optimization.

Referring to the computational stencil shown in Figure B.1, a Taylor expansion of

variable Φ at point N taken about point P yields:

ΦN = ΦP + δxNPδΦ

δx

∣∣∣∣P

+ δyNPδΦ

δy

∣∣∣∣P

+ δzNPδΦ

δz

∣∣∣∣P

+H.O.T, (B.1)

where δγNP = γN − γP and higher-order terms (H.O.T) are neglected. As with

every Taylor series expansion, the approximation error is the sum of the higher

order terms neglected higher-order terms. Therefore, the error eN can be computed

as

eN = ΦP − ΦN + δxNPδΦ

δx

∣∣∣∣P

+ δyNPδΦ

δy

∣∣∣∣P

+ δzNPδΦ

δz

∣∣∣∣P

. (B.2)

For numerical stability, it is common practice to use a weighted error by multi-

plying ek with some weighting function wk (where k is the index of the neighboring

cell). For this work, an inverse distance weighting function will be used. This

weighting function is written as

Page 150: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

128

Figure B.1: Uniform rectilinear computational compact stencil.

wk =1√

(δxkP )2 + (δykP )2 + (δzkP )2. (B.3)

Next, the square of the weighted error between the cell-center P and each of

it neighbors are summed and this value is used as the objective function to be

minimized through the optimization process.

Fobj =k∑

(wkek)2, (B.4)

where k represents the index of the neighboring cells. In the 2D stencil shown in

B.1, k = N,S,W,E.

The goal is to compute the gradient values at point P that minimize Fobj.

Assuming Fobj is convex, we can find the global minimum with respect to the

gradient where

δ∑

(wkek)2

δ δΦδx

∣∣P

=δ∑

(wkek)2

δ δΦδy

∣∣P

=δ∑

(wkek)2

δ δΦδz

∣∣P

= 0. (B.5)

Page 151: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

129

Beginning with the derivative with respect to δΦδx

∣∣P

gives

δ∑

(wkek)2

δ δΦδx

∣∣P

=∑(

w2k

δe2k

δ δΦδx

∣∣P

)= 0, (B.6)

where the derivative of e2k can be easily computed as

δe2k

δ δΦδx

∣∣P

= 2ekδxkP . (B.7)

Substituting this derivative back into the minimized objective function gives

∑(w2kekδxkP

)=∑[

w2k

(ΦP − Φk + δxkP

δΦ

δx

∣∣∣∣P

+ δykPδΦ

δy

∣∣∣∣P

+ δzkPδΦ

δz

∣∣∣∣P

)δxkP

]= 0.

(B.8)

Some algebra allows this summation to be expressed as

δΦ

δx

∣∣∣∣P

∑(w2kδx

2kP

)+δΦ

δy

∣∣∣∣P

∑(w2kδxkP δykP

)+δΦ

δz

∣∣∣∣P

∑(w2kδxkP δzkP

)=∑[

w2kδxkP (Φk − ΦP )

].

(B.9)

This new form can be rewritten as

axδΦ

δx

∣∣∣∣P

+ bxδΦ

δy

∣∣∣∣P

+ cxδΦ

δz

∣∣∣∣P

= fx, (B.10a)

ax =∑(

w2kδx

2kP

), (B.10b)

bx =∑(

w2kδxkP δykP

), (B.10c)

cx =∑(

w2kδxkP δzkP

), (B.10d)

fx =∑[

w2kδxkP (Φk − ΦP )

]. (B.10e)

Repeating this technique for derivatives of Fobj with respect to δΦδy

∣∣P

and δΦδz

∣∣P

Page 152: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

130

yields the following equations, respectively:

ayδΦ

δx

∣∣∣∣P

+ byδΦ

δy

∣∣∣∣P

+ cyδΦ

δz

∣∣∣∣P

= fy, (B.11a)

ay =∑(

w2kδxkP δykP

), (B.11b)

by =∑(

w2kδy

2kP

), (B.11c)

cy =∑(

w2kδykP δzkP

), (B.11d)

fy =∑[

w2kδykP (Φk − ΦP )

], (B.11e)

azδΦ

δx

∣∣∣∣P

+ bzδΦ

δy

∣∣∣∣P

+ czδΦ

δz

∣∣∣∣P

= fz, (B.12a)

az =∑(

w2kδxkP δzkP

), (B.12b)

bz =∑(

w2kδykP δzkP

), (B.12c)

cz =∑(

w2kδz

2kP

), (B.12d)

fz =∑[

w2kδzkP (Φk − ΦP )

]. (B.12e)

This system of equations can be expressed in matrix form as:

ax bx cx

ay by cy

az bz cz

δΦδx

δΦδy

δΦδz

P

=

fx

fy

fz

. (B.13)

Now, inverting the 3x3 matrix on the LHS gives the three components of the

∇Φ at cell-center P.

Page 153: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

131

B.1.2 Boundary Conditions

B.1.2.1 Dirichlet BC

Implementation of the Dirichlet BC is straight forward. The distances used are

computed from the cell-center P to the boundary face and ΦBC is used on the

RHS.

B.1.2.2 Extrapolation BC

If Φ is not specified on the boundary face, the value of Φ may be extrapolated to

the face using a Taylor series expansion. Therefore, we can write Φ on boundary

face b as

Φb = ΦP + δxbPδΦ

δx

∣∣∣∣P

+ δybPδΦ

δy

∣∣∣∣P

+ δzbPδΦ

δz

∣∣∣∣P

. (B.14)

Then we can compute (Φb − ΦP ) as

Φb − ΦP = δxbPδΦ

δx

∣∣∣∣P

+ δybPδΦ

δy

∣∣∣∣P

+ δzbPδΦ

δz

∣∣∣∣P

. (B.15)

In this method, the RHS will be treated explicitly. For the extrapolation BC,

the values of ∇Φ used to compute (Φb − ΦP ) should be the gradient values from

the previous iteration.

Page 154: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

Bibliography

[1] Sander, C. (2000) “Genomic medicine and the future of health care,” Sci-ence, 287(5460), pp. 1977–1978.

[2] Taylor, C. A., M. T. Draney, J. P. Ku, D. Parker, B. N. Steele,K. Wang, and C. K. Zarins (1999) “Predictive medicine: computationaltechniques in therapeutic decision-making,” Computer aided surgery, 4(5),pp. 231–247.

[3] Hamburg, M. A. and F. S. Collins (2010) “The path to personalizedmedicine,” New England Journal of Medicine, 363(4), pp. 301–304.

[4] Chin, L., J. N. Andersen, and P. A. Futreal (2011) “Cancer genomics:from discovery science to personalized medicine,” Nature medicine, 17(3),pp. 297–303.

[5] Clark-Langone, K. M., C. Sangli, J. Krishnakumar, and D. Wat-son (2010) “Translating tumor biology into personalized treatment planning:analytical performance characteristics of the Oncotype DX R©Colon CancerAssay,” BMC Cancer, 10(1), pp. 1–11.

[6] Project, T. C. L. C. G. and N. G. M. NGM (2013) “A genomics-basedclassification of human lung tumors,” Science translational medicine, 5(209),p. 209ra153.

[7] Roukos, D. H. (2008) “Innovative genomic-based model for personalizedtreatment of gastric cancer: integrating current standards and new technolo-gies,” Expert review of molecular diagnostics, 8(1), pp. 29–39.

[8] Salgia, R., T. Hensing, N. Campbell, A. K. Salama, M. Mait-land, P. Hoffman, V. Villaflor, and E. E. Vokes (2011) “Personal-ized Treatment of Lung Cancer,” Seminars in Oncology, 38(2), pp. 274–283.

Page 155: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

133

[9] Sandborn, W. J., B. G. Feagan, S. B. Hanauer, H. Lochs,R. Lofberg, R. Modigliani, D. H. Present, P. Rutgeerts,J. Scholmerich, E. F. Stange, and Others (2002) “A review of ac-tivity indices and efficacy endpoints for clinical trials of medical therapy inadults with Crohn’s disease,” Gastroenterology, 122(2), pp. 512–530.

[10] Oberthuer, A., D. Juraeva, L. Li, Y. Kahlert, F. Westermann,R. Eils, F. Berthold, L. Shi, R. D. Wolfinger, M. Fischer, andOthers (2010) “Comparison of performance of one-color and two-colorgene-expression analyses in predicting clinical endpoints of neuroblastomapatients,” The pharmacogenomics journal, 10(4), pp. 258–266.

[11] Strimbu, K. and J. A. Tavel (2010) “What are biomarkers?” CurrentOpinion in HIV and AIDS, 5(6), p. 463.

[12] Duffy, M. J. and J. Crown (2008) “A Personalized Approach to CancerTreatment: How Biomarkers Can Help,” Clinical Chemistry, 54(11), pp.1770–1779.

[13] Ziegler, A., A. Koch, K. Krockenberger, and A. Großhennig(2012) “Personalized medicine using DNA biomarkers: a review,” Humangenetics, 131(10), pp. 1627–1638.

[14] Felfoul, O., M. Mohammadi, S. Taherkhani, D. de Lanauze, Y. Z.Xu, D. Loghin, S. Essa, S. Jancik, D. Houle, M. Lafleur, and Oth-ers (2016) “Magneto-aerotactic bacteria deliver drug-containing nanolipo-somes to tumour hypoxic regions,” Nature Nanotechnology.

[15] Khalil, I. S. M., V. Magdanz, S. Sanchez, O. G. Schmidt, L. Abel-mann, and S. Misra (2013) “Magnetic control of potential microroboticdrug delivery systems: nanoparticles, magnetotactic bacteria and self-propelled microjets,” in 2013 35th Annual International Conference of theIEEE Engineering in Medicine and Biology Society (EMBC), IEEE, pp.5299–5302.

[16] Martel, S. (2016) “Swimming microorganisms acting as nanorobots versusartificial nanorobotic agents: A perspective view from an historical retro-spective on the future of medical nanorobotics in the largest known three-dimensional biomicrofluidic networks,” Biomicrofluidics, 10(2), p. 21301.

[17] Martel, S., M. Mohammadi, O. Felfoul, Z. Lu, and P. Pouponneau(2009) “Flagellated magnetotactic bacteria as controlled MRI-trackablepropulsion and steering systems for medical nanorobots operating in the hu-man microvasculature,” The International journal of robotics research, 28(4),pp. 571–582.

Page 156: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

134

[18] Jasak, H. and H. G. Weller (2000) “Application of the finite volumemethod and unstructured meshes to linear elasticity,” International journalfor numerical methods in engineering, 48(2), pp. 267–287.

[19] Jasak, H. (2007) “Updated Lagrangian finite volume solver for large defor-mation dynamic response of elastic body,” Transactions of FAMENA, 31(1),p. 55.

[20] Demirdzic, I. and S. Muzaferija (1994) “Finite volume method for stressanalysis in complex domains,” International Journal for Numerical Methodsin Engineering, 37(21), pp. 3751–3766.

[21] Demirdzic, I. (2016) “A fourth-order finite volume method for structuralanalysis,” Applied Mathematical Modelling, 40(4), pp. 3104–3114.

[22] Tukovic, Z., A. Ivankovic, and A. Karac (2013) “Finite-volume stressanalysis in multi-material linear elastic body,” International journal for nu-merical methods in engineering, 93(4), pp. 400–419.

[23] Bijelonja, I., I. Demirdzic, and S. Muzaferija (2006) “A finite volumemethod for incompressible linear elasticity,” Computer methods in appliedmechanics and engineering, 195(44), pp. 6378–6390.

[24] Slone, A. K., C. Bailey, and M. Cross (2003) “Dynamic solid mechanicsusing finite volume methods,” Applied mathematical modelling, 27(2), pp.69–87.

[25] Greenshields, C. J. and H. G. Weller (2005) “A unified formulationfor continuum mechanics applied to fluid–structure interaction in flexibletubes,” International Journal for Numerical Methods in Engineering, 64(12),pp. 1575–1593.

[26] Slone, A. K., K. Pericleous, C. Bailey, and M. Cross (2002) “Dy-namic fluid–structure interaction using finite volume unstructured mesh pro-cedures,” Computers & structures, 80(5), pp. 371–390.

[27] Slone, A. K., K. Pericleous, C. Bailey, M. Cross, and C. Ben-nett (2004) “A finite volume unstructured mesh approach to dynamic fluid–structure interaction: an assessment of the challenge of predicting the onsetof flutter,” Applied mathematical modelling, 28(2), pp. 211–239.

[28] Hoskins, M. (2008) Development of a computational fluid dynamics toolto explore the interactions between cancer cells and leukocytes, Ph.D. thesis,The Pennsylvania State University.

Page 157: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

135

[29] Gaskin, B. J. (2014) Coupled Flow-Biochemistry Simulations of DynamicSystems of Blood Cells, Ph.D. thesis, The Pennsylvania State University.

[30] Lei, H. and G. E. Karniadakis (2015) “Multiscale Modeling of Sickle CellAnemia,” in Modeling the Heart and the Circulatory System, Springer, pp.119–156.

[31] ——— (2012) “Quantifying the Rheological and Hemodynamic Character-istics of Sickle Cell Anemia,” Biophysical Journal, 102(2), pp. 185–194.

[32] ——— (2013) “Probing vasoocclusion phenomena in sickle cell anemia viamesoscopic simulations,” Proceedings of the National Academy of Sciences,110(28), pp. 11326–11330.

[33] Fedosov, D. A., M. Dao, G. E. Karniadakis, and S. Suresh (2014)“Computational Biorheology of Human Blood Flow in Health and Disease,”Annals of Biomedical Engineering, 42(2), pp. 368–387.

[34] Epstein, F. H. and H. F. Bunn (1997) “Pathogenesis and treatment ofsickle cell disease,” New England Journal of Medicine, 337(11), pp. 762–769.

[35] Ingram, V. M. (1957) “Gene mutations in human haemoglobin: thechemical difference between normal and sickle cell haemoglobin,” Nature,180(4581), pp. 326–328.

[36] Pauling, L., H. A. Itano, S. J. Singer, and I. C. Wells (2004) “SickleCell Anemia, a Molecular Disease,” Landmarks in Medical Genetics: ClassicPapers with Commentaries, 51, p. 200.

[37] Platt, O. S., D. J. Brambilla, W. F. Rosse, P. F. Milner, O. Cas-tro, M. H. Steinberg, and P. P. Klug (1994) “Mortality in sickle celldisease–life expectancy and risk factors for early death,” New England Jour-nal of Medicine, 330(23), pp. 1639–1644.

[38] Mori, D., K. Yano, K.-i. Tsubota, T. Ishikawa, S. Wada, and T. Ya-maguchi (2008) “Computational study on effect of red blood cells on pri-mary thrombus formation,” Thrombosis Research, 123(1), pp. 114–121.

[39] Burt, V. L., P. Whelton, E. J. Roccella, C. Brown, J. A. Cut-ler, M. Higgins, M. J. Horan, and D. Labarthe (1995) “Prevalenceof hypertension in the US adult population results from the Third NationalHealth and Nutrition Examination Survey, 1988-1991,” Hypertension, 25(3),pp. 305–313.

[40] Chaouat, A., E. Weitzenblum, and T. Higenbottam (1996) “Therole of thrombosis in severe pulmonary hypertension,” European RespiratoryJournal, 9(2), pp. 356–363.

Page 158: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

136

[41] Fedullo, P. F., W. R. Auger, K. M. Kerr, and L. J. Rubin (2001)“Chronic thromboembolic pulmonary hypertension,” New England Journalof Medicine, 345(20), pp. 1465–1472.

[42] Sowers, J. R., M. Epstein, and E. D. Frohlich (2001) “Diabetes,hypertension, and cardiovascular disease an update,” Hypertension, 37(4),pp. 1053–1059.

[43] Ye, T., N. Phan-Thien, B. C. Khoo, and C. T. Lim (2013) “Stretchingand Relaxation of Malaria-Infected Red Blood Cells,” Biophysical Journal,105(5), pp. 1103–1109.

[44] Nakamura, M., S. Bessho, and S. Wada (2014) “Analysis of Red BloodCell Deformation under Fast Shear Flow for Better Estimation of Hemoly-sis,” International Journal for Numerical Methods in Biomedical Engineer-ing, 30(1), pp. 42–54.

[45] Hoskins, M. H., R. F. Kunz, J. E. Bistline, and C. Dong (2009) “Cou-pled flow-structure-biochemistry simulations of dynamic systems of bloodcells using an adaptive surface tracking method,” Journal of Fluids andStructures, 25, pp. 936–953.

[46] Liang, S., M. Hoskins, P. Khanna, R. F. Kunz, and C. Dong (2008)“Effects of the Tumor-Leukocyte Microenvironment on MelanomaNeutrophilAdhesion to the Endothelium in a Shear Flow,” Cellular and Molecular Bio-engineering, 1(2-3), pp. 189–200.

[47] Rodkiewicz, C. M., P. Sinha, and J. S. Kennedy (1990) “On the Ap-plication of a Constitutive Equation for Whole Human Blood,” Journal ofBiomechanical Engineering, 112(2), p. 198.

[48] Johnston, B. M., P. R. Johnston, S. Corney, and D. Kilpatrick(2004) “Non-Newtonian blood flow in human right coronary arteries: steadystate simulations,” Journal of Biomechanics, 37(5), pp. 709–720.

[49] Ballyk, P., D. Steinman, and C. Ethier (2017) “Simulation of non-Newtonian blood flow in an end-to-side anastomosis,” Biorheology, 31(5),pp. 565–586.

[50] Slattery, M. J. and C. Dong (2003) “Neutrophils influence melanomaadhesion and migration under flow conditions,” International journal of can-cer, 106(5), pp. 713–722.

[51] Dong, C., M. J. Slattery, S. Liang, and H.-H. Peng (2005)“Melanoma Cell Extravasation under Flow Conditions is Modulated by

Page 159: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

137

Leukocytes and Endogenously Produced Interleukin 8,” Mol Cell Biomech,2(3), pp. 145–159.

[52] Behr, J., B. Gaskin, C. Fu, C. Dong, and R. Kunz “Localized Modelingof Biochemical and Flow Interactions During Cancer Cell Adhesion,” .

[53] Leyton-Mange, J., S. Yang, M. Henty, R. Kunz, J. Zahn, andC. Dong (2006) “Design of a Side-View Particle Image Velocimetry Sys-tem for Cellular Adhesion Analysis,” Biomechanical Engineering, 128, pp.271–278.

[54] Behr, J. (2013) Multiscale modeling of cancer cell adhesion, Ph.D. thesis,The Pennsylvania State University.

[55] Tsubota, K.-i., S. Wada, and T. Yamaguchi (2006) “Particle methodfor computer simulation of red blood cell motion in blood flow,” ComputerMethods and Programs in Biomedicine, 83(2), pp. 139–146.

[56] Muller, Matthias and Schirm, Simon and Teschner, M. (2004)“Interactive blood simulation for virtual surgery based on smoothed particlehydrodynamics,” Technology and Health Care, 12(1), pp. 25—-31.

[57] Yamaguchi, T., T. Ishikawa, Y. Imai, N. Matsuki, M. Xenos,Y. Deng, and D. Bluestein (2010) “Particle-Based Methods for Multi-scale Modeling of Blood Flow in the Circulation and in Devices: Challengesand Future Directions,” Annals of Biomedical Engineering, 38(3), pp. 1225–1235.

[58] Eldredge, J. D., A. Leonard, and T. Colonius (2002) “A General De-terministic Treatment of Derivatives in Particle Methods,” Journal of Com-putational Physics, 180(2), pp. 686–709.

[59] Chen, S. and G. D. Doolen (1998) “LATTICE BOLTZMANN METHODFOR FLUID FLOWS,” Annual Review of Fluid Mechanics, 30(1), pp. 329–364.

[60] Khismatullin, D. B. and G. A. Truskey (2005) “Three-dimensionalnumerical simulation of receptor-mediated leukocyte adhesion to surfaces:Effects of cell deformability and viscoelasticity,” Physics of Fluids, 17(3), p.031505.

[61] N’Dri, N., W. Shyy, and R. Tran-Son-Tay (2003) “ComputationalModeling of Cell Adhesion and Movement Using a Continuum-Kinetics Ap-proach,” Biophysical Journal, 85(4), pp. 2273–2286.

Page 160: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

138

[62] Doddi, S. K. and P. Bagchi (2009) “Three-dimensional computationalmodeling of multiple deformable cells flowing in microvessels,” Physical Re-view E, 79(4), p. 046318.

[63] Bagchi, P. (2007) “Mesoscale Simulation of Blood Flow in Small Vessels,”Biophysical Journal, 92(6), pp. 1858–1877.

[64] Pan, T.-W., L. Shi, and R. Glowinski (2010) “A DLM/FD/IB methodfor simulating cell/cell and cell/particle interaction in microchannels,” Chi-nese Annals of Mathematics, Series B, 31(6), pp. 975–990.

[65] Kan, H.-C., H. S. Udaykumar, W. Shyy, and R. Tran-Son-Tay(1998) “Hydrodynamics of a compound drop with application to leukocytemodeling,” Physics of Fluids, 10(4), pp. 760–774.

[66] Pawar, P., S. Jadhav, C. D. Eggleton, and K. Konstantopoulos(2008) “Roles of cell and microvillus deformation and receptor-ligand bindingkinetics in cell rolling,” American Journal of Physiology-Heart and Circula-tory Physiology, 295(4), pp. H1439–H1450.

[67] Crowl, L. M. and A. L. Fogelson (2010) “Computational model ofwhole blood exhibiting lateral platelet motion induced by red blood cells,” In-ternational Journal for Numerical Methods in Biomedical Engineering, 26(3-4), pp. 471–487.

[68] II, S., K. SUGIYAMA, S. TAKAGI, and Y. MATSUMOTO (2012) “AComputational Blood Flow Analysis in a Capillary Vessel including Multi-ple Red Blood Cells and Platelets,” Journal of Biomechanical Science andEngineering, 7(1), pp. 72–83.

[69] Jadhav, S., C. D. Eggleton, and K. Konstantopoulos (2005) “A 3-Dcomputational model predicts that cell deformation affects selectin-mediatedleukocyte rolling,” Biophysical journal, 88(1), pp. 96–104.

[70] Hammer, D. A. and S. M. Apte (1992) “Simulation of cell rolling andadhesion on surfaces in shear flow: general results and analysis of selectin-mediated neutrophil adhesion,” Biophysical journal, 63(1), pp. 35–57.

[71] Chang, K.-C. and D. A. Hammer (1996) “Influence of Direction and Typeof Applied Force on the Detachment of Macromolecularly-Bound Particlesfrom Surfaces,” Langmuir, 12(9), pp. 2271–2282.

[72] Caputo, K. E., D. Lee, M. R. King, and D. A. Hammer (2007) “Ad-hesive Dynamics Simulations of the Shear Threshold Effect for Leukocytes,”Biophysical Journal, 92(3), pp. 787–797.

Page 161: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

139

[73] Kunz, R. F., B. J. Gaskin, Q. Li, S. Davanloo-Tajbakhsh, andC. Dong (2015) “Multi-scale biological and physical modelling of the tu-mour micro-environment,” Drug Discovery Today: Disease Models, 16, pp.7–15.

[74] Kunz, R. F., W. S. Yu, S. P. Antal, and S. M. Ettorre (2001) “Anunstructured two-fluid method based on the coupled phasic exchange algo-rithm,” in AIAA Computational Fluid Dynamics Conference. AIAA Paper,vol. 2672.

[75] Freund, J. B. (2014) “Numerical Simulation of Flowing Blood Cells,” An-nual Review of Fluid Mechanics, 46(1), pp. 67–95.

[76] Cao, J., B. Donell, D. R. Deaver, M. B. Lawrence, and C. Dong(1998) “In Vitro Side-View Imaging Technique and Analysis of Human T-Leukemic Cell Adhesion to ICAM-1 in Shear Flow,” Microvascular research,55(2), pp. 124–137.

[77] Dong, C., J. Cao, E. J. Struble, and H. H. Lipowsky (1999) “Me-chanics of Leukocyte Deformation and Adhesion to Endothelium in ShearFlow,” Annals of Biomedical Engineering, 27(3), pp. 298–312.

[78] Cardiff, P. (2012) Development of the finite volume method for hip jointstress analysis, Ph.D. thesis, University College Dublin.

[79] Sussman, T. and K.-J. Bathe (1987) “A finite element formulation fornonlinear incompressible elastic and inelastic analysis,” Computers & Struc-tures, 26(1-2), pp. 357–409.

[80] Hirt, C., A. Amsden, and J. Cook (1974) “An arbitrary Lagrangian-Eulerian computing method for all flow speeds,” Journal of ComputationalPhysics, 14(3), pp. 227–253.

[81] Hu, H. H., N. Patankar, and M. Zhu (2001) “Direct Numerical Simula-tions of FluidSolid Systems Using the Arbitrary LagrangianEulerian Tech-nique,” Journal of Computational Physics, 169(2), pp. 427–462.

[82] Bell, G. I. and Others (1978) “Models for the specific adhesion of cellsto cells,” Science, 200(4342), pp. 618–627.

[83] Dembo, M., D. C. Torney, K. Saxman, and D. Hammer (1988) “Thereaction-limited kinetics of membrane-to-surface adhesion and detachment,”Proceedings of the Royal Society of London. Series B. Biological Sciences,234(1274), pp. 55–83.

Page 162: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

140

[84] Kim, M. J., A. A. Julius, and E. Steager (2012) “Historical Overview,”in Microbiorobotics, Elsevier, pp. xix–xxiv.

[85] Cheang, U. K., F. Meshkati, H. Kim, K. Lee, H. C. Fu, and M. J.Kim (2016) “Versatile microrobotics using simple modular subunits,” Scien-tific Reports, 6.

[86] Fischer, P. and A. Ghosh (2011) “Magnetically actuated propulsion atlow Reynolds numbers: towards nanoscale control,” Nanoscale, 3(2), pp.557–563.

[87] Mathieu, J.-B. and S. Martel (2007) “In vivo validation of a propul-sion method for untethered medical microrobots using a clinical magneticresonance imaging system,” in 2007 IEEE/RSJ International Conference onIntelligent Robots and Systems, IEEE, pp. 502–508.

[88] Peyer, K. E., L. Zhang, and B. J. Nelson (2013) “Bio-inspired magneticswimming microrobots for biomedical applications,” Nanoscale, 5(4), pp.1259–1272.

[89] Xu, T., G. Hwang, N. Andreff, and S. Regnier “Influence of geometryon swimming performance of helical swimmers using DoE,” Journal of Micro-Bio Robotics, pp. 1–10.

[90] Abbott, J. J., K. E. Peyer, M. C. Lagomarsino, L. Zhang, L. Dong,I. K. Kaliakatsos, and B. J. Nelson (2009) “How Should MicrorobotsSwim?” The International Journal of Robotics Research, 28(11-12), pp.1434–1447.

[91] Becker, A. T. (2017) “Controlling swarms of robots with global inputs:Breaking symmetry,” in Microbiorobotics, Elsevier, pp. 3–20.

[92] Barrere, F., C. A. van Blitterswijk, and K. de Groot (2006) “Boneregeneration: molecular and cellular interactions with calcium phosphateceramics,” International journal of nanomedicine, 1(3), p. 317.

[93] Wurmser, A. E., T. D. Palmer, and F. H. Gage (2004) “CellularInteractions in the Stem Cell Niche,” Science, 304(5675), pp. 1253–1255.

[94] Agrawal, B., S. J. Gendler, and B. Longenecker (1998) “The biolog-ical role of mucins in cellular interactions and immune regulation: prospectsfor cancer immunotherapy,” Molecular Medicine Today, 4(9), pp. 397–403.

[95] Sims, N. A. and J. H. Gooi (2008) “Bone remodeling: Multiple cellu-lar interactions required for coupling of bone formation and resorption,” inSeminars in cell & developmental biology, vol. 19, Elsevier, pp. 444–451.

Page 163: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

141

[96] Wang, D. and S. J. Lippard (2005) “Cellular processing of platinum an-ticancer drugs,” Nature reviews Drug discovery, 4(4), pp. 307–320.

[97] Gurtin, M. E., E. Fried, and L. Anand (2010) The mechanics andthermodynamics of continua, Cambridge University Press.

[98] Mitchison, J. M. and M. M. Swann (1954) “The mechanical propertiesof the cell surface,” Journal of Experimental Biology, 31(3), pp. 461–472.

[99] Holzapfel, G. A. (2000) Nonlinear solid mechanics, vol. 24, Wiley Chich-ester.

[100] Ogden, R. W. (1997) Non-linear elastic deformations, Courier Corpora-tion.

[101] Hayt Jr WH; Buck, J. A. (2006), “Engineering Electromagnetics. Vol7,” .

[102] Nagy, Z. G. (2011) Modeling and analysis of the magnetization, torqueand dynamics of untethered soft-magnetic microrobots, Ph.D. thesis, ETHZURICH.

[103] Park, E. Y. H., M. J. Smith, E. S. Stropp, K. R. Snapp, J. A.DiVietro, W. F. Walker, D. W. Schmidtke, S. L. Diamond, andM. B. Lawrence (2002) “Comparison of PSGL-1 Microbead and Neu-trophil Rolling: Microvillus Elongation Stabilizes P-Selectin Bond Clusters,”Biophysical Journal, 82(4), pp. 1835–1847.

[104] Fornberg, B. (1998) “Calculation of Weights in Finite Difference Formu-las,” SIAM Review, 40(3), pp. 685–691.

[105] Burden, R. and J. D. Faires (2001) Numerical Analysis, Brooks/Cole.

[106] Kundu, PK and Cohen, I. (2008) Fluid Mechanics, Elsevier AcademicPress.

[107] Bergman, T. L., F. P. Incropera, D. P. DeWitt, and A. S. Lavine(2011) Fundamentals of Heat and Mass Transfer, John Wiley & Sons.

[108] Buhmann, M. D. (2003) Radial Basis Functions: Theory and Implementa-tions, Cambridge University Press.

[109] Kunz, R. F., B. W. Siebert, W. Cope, N. F. Foster, S. P. An-tal, and S. M. Ettorre (1998) “A coupled phasic exchange algorithm forthree-dimensional multi-field analysis of heated flows with mass transfer,”Computers & Fluids, 27(7), pp. 741–768.

Page 164: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

142

[110] Yang, X. I. and R. Mittal (2014) “Acceleration of the Jacobi iterativemethod by factors exceeding 100 using scheduled relaxation,” Journal ofComputational Physics, 274, pp. 695–708.

[111] Sadd, M. H. (2009) Elasticity: Theory, Applications, and Numerics, ACA-DEMIC PRESS INC.

[112] Saffman, P. G. (1965) “The lift on a small sphere in a slow shear flow,”Journal of Fluid Mechanics, 22(02), p. 385.

[113] Cherukat, P., J. McLaughlin, and A. Graham (1994) “The inertiallift on a rigid sphere translating in a linear shear flow field,” InternationalJournal of Multiphase Flow, 20(2), pp. 339–353.

[114] Purcell, E. M. (1977) “Life at low Reynolds number,” American Journalof Physics, 45(1), pp. 3–11.

[115] Lighthill, J. (1976) “Flagellar Hydrodynamics,” SIAM Review, 18(2), pp.161–230.

Page 165: COUPLED BIOPHYSICAL SIMULATIONS OF CELL-CELL …

Vita

Byron J. Gaskin

Byron Gaskin grew up in Miami, Florida where he graduated from MonsignorEdward Pace High School in 2007. It was during high school that Byron was firstintroduced to computer programming. Upon completion of high school, Byronattended Florida International University in Miami where he obtained a Bache-lor’s Degree in Mechanical Engineering. During his undergraduate career, Byronconducted undergraduate research in computational fluid dynamics and worked asan intern at the Penn State Applied Research Laboratory.

Byron began his Master’s degree at Penn State in Mechanical Engineering inAugust 2012 and completed the degree in August 2014. He completed a Mas-ter’s thesis titled “Coupled Flow-Biochemistry Simulations of Dynamic Systems ofBlood Cells”.

After completion of his Master’s degree, Byron started work on his doctor-ate in Bioengineering in Fall 2014. Byron was, in Spring 2018, selected for theEmerging Leaders in Data Science Fellowship at the National Institute of Allergiesand Infectious Diseases (NIAID) at the National Institutes of Health (NIH). Uponcompletion of the doctorate, Byron will begin his 12-month data science fellowshipat NIAID in Rockville, MD.

Byron is a member of Sigma Phi Epsilon Fraternity and a founder of the OmegaGamma chapter of Theta Tau Fraternity.