computational fluid dynamics modeling of scalable stirred

105
University of Calgary PRISM: University of Calgary's Digital Repository Graduate Studies The Vault: Electronic Theses and Dissertations 2016-02-03 Computational Fluid Dynamics Modeling of Scalable Stirred Suspension Bioreactors for Pluripotent Stem Cell Expansion Le, An Le, A. (2016). Computational Fluid Dynamics Modeling of Scalable Stirred Suspension Bioreactors for Pluripotent Stem Cell Expansion (Unpublished master's thesis). University of Calgary, Calgary, AB. doi:10.11575/PRISM/25392 http://hdl.handle.net/11023/2831 master thesis University of Calgary graduate students retain copyright ownership and moral rights for their thesis. You may use this material in any way that is permitted by the Copyright Act or through licensing that has been assigned to the document. For uses that are not allowable under copyright legislation or licensing, you are required to seek permission. Downloaded from PRISM: https://prism.ucalgary.ca

Upload: others

Post on 23-Dec-2021

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Computational Fluid Dynamics Modeling of Scalable Stirred

University of Calgary

PRISM: University of Calgary's Digital Repository

Graduate Studies The Vault: Electronic Theses and Dissertations

2016-02-03

Computational Fluid Dynamics Modeling of Scalable

Stirred Suspension Bioreactors for Pluripotent Stem

Cell Expansion

Le, An

Le, A. (2016). Computational Fluid Dynamics Modeling of Scalable Stirred Suspension Bioreactors

for Pluripotent Stem Cell Expansion (Unpublished master's thesis). University of Calgary, Calgary,

AB. doi:10.11575/PRISM/25392

http://hdl.handle.net/11023/2831

master thesis

University of Calgary graduate students retain copyright ownership and moral rights for their

thesis. You may use this material in any way that is permitted by the Copyright Act or through

licensing that has been assigned to the document. For uses that are not allowable under

copyright legislation or licensing, you are required to seek permission.

Downloaded from PRISM: https://prism.ucalgary.ca

Page 2: Computational Fluid Dynamics Modeling of Scalable Stirred

UNIVERSITY OF CALGARY

Computational Fluid Dynamics Modeling of Scalable Stirred Suspension Bioreactors for

Pluripotent Stem Cell Expansion

by

An Thuy Dang Le

A THESIS

SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

IN PARTIAL FULFILMENT OF THE REQUIREMENTS FOR THE

DEGREE OF MASTER OF SCIENCE

GRADUATE PROGRAM IN BIOMEDICAL ENGINEERING

CALGARY, ALBERTA

JANUARY, 2016

© An Thuy Dang Le 2016

Page 3: Computational Fluid Dynamics Modeling of Scalable Stirred

ii

Abstract

Pluripotent stem cells (PSCs) including embryonic and induced pluripotent stem cells are known

for their potential use in cell-based therapy, disease model study, and drug screening. One of the

key challenges in pluripotent stem cell research is to establish scalable bioprocesses that reliably

produce cells with high quality at any desired quantity. Stirred suspension bioreactors (SSBs) are

known to provide a controlled and well-mixed environment for aggregate-forming cells, such as

murine and human PSCs. Hydrodynamic environment of SSBs, particularly shear stress and small

eddies, have been shown to have a significant impact on the expansion and pluripotency of

pluripotent stem cells. However, the exact mechanism has not been fully understood. In this

project, computational fluid dynamic (CFD) simulation was employed to model the hydrodynamic

environment within SSBs with various configurations and physical conditions. Understanding the

hydrodynamics is one of the first key steps in bioprocess development of PSCs using SSBs.

Page 4: Computational Fluid Dynamics Modeling of Scalable Stirred

iii

Acknowledgements

I would like to express my gratefulness to my supervisors, Dr. Michael Kallos and Dr. Ian Gates,

for their guidance and support through this project. They have provided me the best environment

for doing research, given me the opportunities to be creative, and motivated me to become a good

biomedical engineer.

I would like to thank Dr. Derrick Rancourt for giving me the opportunity to work on this exciting

project. Additionally, I would like to acknowledge Dr. Guoliang Meng for his in-depth training on

pluripotent stem cells culture and Dr. Charlie Hsu for his help with cell aggregate sizing using

Multisizer III in Krawetz’s Lab.

I would like to thank all PPRF lab members for their help and support. Every day at the lab is a

joyful day for me. Specially, I would like to acknowledge Breanna Borys for her help on the 10

ml bioreactor biological experiment and analysis.

Last but not least, I would like to thank my family for their unconditional love and care. Their trust

and understanding have given me strength to follow my dreams.

Thank you.

Page 5: Computational Fluid Dynamics Modeling of Scalable Stirred

iv

Table of Contents

Abstract ............................................................................................................................... ii Acknowledgements ............................................................................................................ iii Table of Contents ............................................................................................................... iv List of Tables ..................................................................................................................... vi List of Figures and Illustrations ........................................................................................ vii List of Symbols, Abbreviations and Nomenclature .............................................................x 

CHAPTER ONE: INTRODUCTION ..................................................................................1 Motivation ........................................................................................................................1 1.1 Research Aims ...........................................................................................................2 

CHAPTER TWO: LITERATURE REVIEW ......................................................................4 2.1 Pluripotent Stem Cells ...............................................................................................4 

2.1.1 Stem Cells and Regenerative Medicine Overview ............................................4 2.1.2 Controversial Issues Related to ESCs ...............................................................5 2.1.3 Induced-pluripotent Stem Cell Discovery and Its Significance ........................6 2.1.4 The Needs and Current Status of PSC Expansion .............................................7 2.1.5 Expansion of PSCs in Stirred Suspension Bioreactors ......................................8 2.1.6 Differentiation of PSCs in Stirred Suspension Bioreactors ...............................9 2.1.7 Challenges in PSC Bioprocess Development ..................................................10 2.1.8 The Significance of Controlling Aggregate Size .............................................12 

2.2 Computational Fluid Dynamics (CFD) ....................................................................13 2.2.1 The role of CFD Modeling in Bioprocess Development .................................13 2.2.2 Turbulent Flow ................................................................................................15 2.2.3 Governing Equations in CFD Modeling of SSBs ............................................16 2.2.4 Finite Element Method ....................................................................................17 

2.3 Conclusions ..............................................................................................................18 

CHAPTER THREE: HYDRODYNAMICS OF STANDARD LAB-SCALE 100 ML SPINNER FLASK BIOREACTOR ..........................................................................19 

3.1 Introduction ..............................................................................................................19 3.2 Materials and Methods .............................................................................................21 

3.2.1 Cell culture and Cell Aggregate Measurements ..............................................21 3.2.2 Calculations .....................................................................................................22 

3.2.2.1 Maximum Shear Stress Calculation .......................................................22 3.2.3 Computational Fluid Dynamics Modeling ......................................................23 

3.2.3.1 Geometry ...............................................................................................23 3.2.3.2 Model Physics and Grid Generation ......................................................24 

3.3 Results and Discussion ............................................................................................26 3.3.1 Grid Dependence and Model Validation .........................................................26 3.3.2 Base Case: Standard 100 ml Spinner Flask Bioreactor ...................................28 3.3.3 Hydrodynamics of 100ml Bioreactor at Different Agitation Rates .................34 3.3.4 Relationships between Hydrodynamic Parameters and Aggregate size ..........36 

3.3.4.1 Effects of shear rate on aggregate size ...................................................36 3.4 Conclusions ..............................................................................................................40 

Page 6: Computational Fluid Dynamics Modeling of Scalable Stirred

v

CHAPTER FOUR: COMPUTATIONAL FLUID DYNAMIC MODELING OF SCALED-DOWN 10 ML STIRRED SUSPENSION BIOREACTORS WITH DIFFERENT IMPELLER DESIGNS .............................................................................................42 

4.1 Introduction ..............................................................................................................42 4.2 Materials and Methods .............................................................................................44 

4.2.1 Cell Culture .....................................................................................................44 4.2.2 Computational Fluid Dynamics Modeling ......................................................45 

4.2.2.1 Geometry ...............................................................................................45 4.2.2.2 Model Physics ........................................................................................46 

4.3 Results and Discussions ...........................................................................................47 4.3.1 Base Case: Hydrodynamics of 10 ml Bioreactor at 100 rpm ..........................47 4.3.2 Comparison of 100 ml and 10 ml Bioreactor Models at 100 rpm ...................49 4.3.3 Effects of Agitation Rates ...............................................................................52 

4.3.3.1 Hydrodynamic Effects ...........................................................................52 4.3.3.2 Biological Effects ..................................................................................54 

4.3.4 Effects of Impeller Geometry ..........................................................................59 4.3.4.1 Hydrodynamic Effects ...........................................................................59 4.3.4.2 Biological Effects ..................................................................................64 

4.3.5 Effects of Shear Rate and Eddy Size on Aggregate Size ................................69 4.4 Conclusions ..............................................................................................................71 

CHAPTER FIVE: DISCUSSIONS OF LIMITATIONS, RECOMMENDATIONS, AND CONCLUSIONS.......................................................................................................73 

5.1 Discussions of Limitations and Recommendations .................................................73 5.1.1 CFD Modeling .................................................................................................73 5.1.2 Biological Experiments ...................................................................................74 

5.2 Conclusions ..............................................................................................................74 

REFERENCES ..................................................................................................................77 

APPENDIX A: CELL CULTURE ....................................................................................86 A.1. Static cell culture ....................................................................................................86 

A.1.1. Gelatin coating ...............................................................................................86 A.1.2. Static Passaging .............................................................................................86 

A.2. Standard 100 ml Spinner Flask Bioreactor ............................................................87 A.2.1. Preparation of 100 ml Bioreactor ..................................................................87 A.2.2. Bioreactor Inoculation and Feeding Regime .................................................88 A.2.3. Aggregate Size Measurement using Particle Sizer ........................................88 

APPENDIX B: HYDRODYNAMICS ..............................................................................90 B.1. Calculation Results .................................................................................................90 B.2. Quasi-steady state approximation ..........................................................................90  

Page 7: Computational Fluid Dynamics Modeling of Scalable Stirred

vi

List of Tables

Table 3–1: Dimensions of 100 ml model ...................................................................................... 24 

Table 3–2: Average velocity, shear rate, and energy dissipation rate at different fluid heights ... 34 

Table 3–3: Volume average velocity, shear rate, and energy dissipation rate at different agitation rates ........................................................................................................................ 36 

Table 4–1: Dimensions of 10 ml bioreactor models ..................................................................... 46 

Table 4–2: Comparisons of volume-averaged velocity, shear rate, and turbulent dissipation rate at different agitation rates in 10 ml bioreactor with cylinder impeller .......................... 52 

Table 4–3: Comparison of volume-averaged velocity, shear rate, and turbulent dissipation rate between bioreactor with paddle impeller and bioreactor with cylinder impeller at 150 rpm ................................................................................................................................. 60 

Table 5–1: Results from maximum shear stress calculation for 100 ml bioreactor ...................... 90 

Table 5–2: Results from Reynolds number calculation for 10 ml bioreactor ............................... 90 

Page 8: Computational Fluid Dynamics Modeling of Scalable Stirred

vii

List of Figures and Illustrations

Figure 3–1: Model geometry built in COMSOL Multiphysics. A) 3D view of fluid occupied region in 100ml bioreactor; B) Top view of model geometry .............................................. 24 

Figure 3–2: 100 ml bioreactor model grids created in COMSOL Multiphysics. A) Finer grid (948,668 tetrahedral finite elements); B) normal grid (278,951 tetrahedral elements) ........ 26 

Figure 3–3: Comparison of results of fine grid (948,668 tetrahedral finite elements) and normal grid (278,951 tetrahedral elements). ......................................................................... 27 

Figure 3–4: Maximum shear stress calculated from equation 1 and simulated using CFD ......... 28 

Figure 3–5: Results of basics hydrodynamic variables obtained from CFD simulation. A) Pressure profile; B) Velocity and streamlines; C) Shear rate; D) Energy dissipation rate ... 31 

Figure 3–6: Velocity on cut plane yz (x=0) plane. A) At t=0s, the impeller is perpendicular to the cut plane, velocity =0; B) At t=10s, the impeller is perpendicular to the cut plane, maximum velocity =0.119 m/s; B) At t=11s, the impeller is 300 from entering the cut plane, maximum velocity=0.129 m/s; D) At t=12s, the impeller is 300 leaving the cut plane, maximum velocity =0.232 m/s ................................................................................... 32 

Figure 3–7: Velocity (A), shear rate (B), and energy dissipation rate (D) at different fluid heights in 100 ml bioreactor model. Cut plane 1 (z=0.03 m) on the left side is the horizontal plane above the impeller. Cut plane 2 (z=0.02 m) in the middle is the horizontal plane through the center of the impeller. Cut plane 3 (z=0.01 m) on the right is the horizontal plane below the impeller ............................................................................ 33 

Figure 3–8: Velocity (A), shear rate (B), and energy dissipation rate (C) at different agitation rates (80 rpm, 100 rpm, and 120 rpm from left to right). All surfaces were plot at 12s of simulation .............................................................................................................................. 35 

Figure 3–9: Cell aggregate distribution (D) at 80 rpm (A), 100 rpm (B), and 120 rpm (C)......... 38 

Figure 3–10: Shear rate and mean average size. A) Volume average shear rate; B) Maximum shear rate ............................................................................................................................... 39 

Figure 3–11: Relationship between eddy size and cell aggregate size. The average eddy size was calculated from volume-average energy dissipation rate obtained by CFD simulation, and the cell aggregate size was obtained from biological experiment. .............. 40 

Figure 4–1: Model geometry built in COMSOL Multiphysics. A) 3D view of fluid occupied region in 10 ml cylindrical impeller bioreactor; B) 3D view of fluid occupied region in 10 ml paddle impeller bioreactor .......................................................................................... 45 

Figure 4–2: Basic hydrodynamics of 10 ml cylindrical impeller bioreactor at 100 rpm .............. 48 

Page 9: Computational Fluid Dynamics Modeling of Scalable Stirred

viii

Figure 4–3: Comparison of shear rate distribution in 100 ml bioreactor and 10 ml bioreactor at 100 rpm. The data was extracted from quasi-steady sate state (after 3s) of agitation. The area under the curve was normalized to unity. .............................................................. 50 

Figure 4–4: Comparison of turbulent dissipation distribution in 100 ml and 10 ml bioreactors at 100 rpm. The volume fraction of shear rate above 150 s-1 is insignificant so that the lines representing the cumulative volume of 10 ml and 100 ml bioreactor at 100 rpm no longer increase on the y-axis. The area under the curve was normalized to unity. .............. 51 

Figure 4–5: Comparison of shear rate distribution between different agitation rates in 10 ml bioreactor. The volume fraction of shear rate above 150 s-1 is insignificant so that the lines representing the cumulative volume of different agitation rates no longer increase on the y-axis. The area under the curve was normalized to unity. ........................................ 53 

Figure 4–6: Comparison of turbulent dissipation between different shear rates in 10 ml bioreactor. The volume fraction of turbulent dissipation at rates higher than 0.02 m2/s3 is insignificant so that the lines representing the cumulative volume of different agitation rates no longer increase on the y-axis. The area under the curve was normalized to unity. . 54 

Figure 4–7: Comparison of biological responses in cylindrical impeller 10 ml bioreactors at different agitation rates and in static culture. ........................................................................ 59 

Figure 4–8: Velocity vector field in 10 ml bioreactor with cylinder impeller and in 10 ml bioreactor with paddle impeller at 150 rpm. Longer and thicker arrows indicate velocity vectors with high magnitude, and the arrows point toward the direction of the velocity vectors. .................................................................................................................................. 61 

Figure 4–9: Comparison of shear rate distribution between the 10 ml bioreactor with cylinder impeller and 10 ml bioreactor with paddle impeller at 150 rpm. The area under the curve was normalized to unity. ....................................................................................................... 62 

Figure 4–10: Comparison of turbulent energy dissipation rate between 10 ml bioreactor with cylinder impeller and 10 ml bioreactor with paddle impeller at 150 rpm. The area under the curve was normalized to unity. ....................................................................................... 63 

Figure 4–11: Comparison of biological responses in 10 bioreactors with cylinder impeller and in 10 ml bioreactor with paddle impeller at 150 rpm. .................................................... 68 

Figure 4–12: Effects of shear rate on murine ESC aggregate size in 10 ml bioreactor with cylinder impeller. The error bars are standard deviations in aggregate size. ........................ 70 

Figure 4–13: Similarity in average aggregate size and eddy size at different turbulent dissipation rates in 10 ml bioreactor. .................................................................................... 71 

Figure 5–1: Shear rate distribution from 0s-12s in 100 ml Bioreactor at 100 rpm. The quasi-steady state is achieved by about 3s ...................................................................................... 91 

Page 10: Computational Fluid Dynamics Modeling of Scalable Stirred

ix

Figure 5–2: Turbulent dissipation rate distribution from 0s-12s in 100 ml Bioreactor at 100 rpm. The quasi-steady state is achieved by about 3s. ........................................................... 92 

Figure 5–3: Shear rate distribution from 0s-10s in 10 ml Bioreactor at 100 rpm. The quasi-steady state is achieved by about 4 s. .................................................................................... 93 

Figure 5–4: Turbulent dissipation rate distribution from 0s-10s in 10 ml Bioreactor at 100 rpm. The quasi-steady state is achieved by about 4 s. .......................................................... 94 

Page 11: Computational Fluid Dynamics Modeling of Scalable Stirred

x

List of Symbols, Abbreviations and Nomenclature

List of Abbreviations CFD Computational Fluid Dynamics c-Myc c-Myc proto-oncogene; Transcription factor DMEM Dulbecco's Modified Eagle's Medium EBs Embryoid Bodies ESC Embryonic Stem Cells FBS Fetal Bonvine Serum iPSC Induced Pluripotent Stem Cells Klf4 Kruppel-like factor 4; Transcription factor LIF Leukemia Inhibitory Factor Octr4 Pluripotency transcription factor PIV Particle Image Velocimetry Sox 2 Pluripotency transcription factor List of Symbols and Nomenclature Di Impeller diameter (m) Dt Vessel diameter (m) k Turbulent kinetic energy (m2/s2) k Turbulent kinetic energy (m2/s2) N Agitation rate (rpm) NP Power number P Power input (W) p Pressure (Pa) Pk Production of kinetic energy Re Reynolds number u Velocity (m/s) VL Working volume (m3) W Impeller width (m) γ Shear rate (1/s) ε Dissipation rate (m2/s3) η Eddy scale (m) μ Dynamic viscosity (Pa.s) ν Kinematic viscosity (m2/s) ρ Density (kg/m3) τ Shear stress (Pa) υT Turbulent eddy viscosity (m2/s)

Page 12: Computational Fluid Dynamics Modeling of Scalable Stirred

1

Chapter One: Introduction

Motivation

Pluripotent stem cells (PSCs) play a key role in regenerative medicine due to their ability to

differentiate into different cell types in the three germ layers (1). Recently, the new induced

pluripotent stem cell (iPSC) derivation breakthrough has made these cells even more attractive to

modern medicine, especially cell-based organ transplantation field. However, transplantation

using PSC derived cells is still remained a challenge. One of the most important reasons is that the

bioprocess for producing large quantity of PSCs has not been well defined (2, 3). It has been

estimated that each cell-based transplantation case needs at least a lot size of a billion cells per

dose to be effective (4, 5). It is nearly impossible to obtain as many cells using conventional static

culture system due to high labor and material cost, time consuming, and large incubator space

requirement (2). Stirred suspension bioreactors (SSBs) have been selected as a preferred method

for expanding PSCs largely due to its well-mixing environment, controllability, and scalability (6).

The stirring mechanism in SSBs not only helps enhance mass transport, but also helps control PSC

aggregate size (7, 8). Besides mixing and aggregate size control, important environmental aspects

such as pH and dissolved oxygen can also be controlled in SSBs as in the DASGIP system (9).

SSB scaling-up bioprocess is very popular in protein production industry (10). Therefore, there is

an extensive amount of published work including simulation and experimental data on scaling

criteria of SSB in the literature (11–13).

Despite these advantages, there are still many challenges regarding the use of SSBs for

PSC expansion and differentiation. It has been shown that the hydrodynamic environment in

SSBs, particularly high shear stress and small turbulent eddies, can cause damage to the cell

membrane and alterations in gene expression (2, 14). Furthermore, reported differentiation

Page 13: Computational Fluid Dynamics Modeling of Scalable Stirred

2

efficiency in SSBs is much lower compared to differentiation efficiency in static culture system

(14, 15). PSCs tend to stay undifferentiated in fluid shear environment even in the absence of

leukemia inhibitory factor (16, 17). However, the detailed mechanisms of these effects have not

been clear.

One of the most important steps in PSC bioprocess development using SSBs is to understand

the actual hydrodynamic environment inside a particular stirring vessel (2). This can be done

efficiently using computational fluid dynamics (CFD) modeling. CFD modeling has been used

extensively in bioprocess development and design of various bioreactor configurations (18–20).

In this project, CFD is to model the hydrodynamics of standard 100 ml and scaled-down 10 ml

spinner flask bioreactors for PSC expansion. The final goal of this project was to understand the

hydrodynamic environment of SSBs at various scales and conditions as well as the effects of

impeller geometry. Biological experiments in this project focuses on cell expansion and cell

aggregate size at various conditions. It has been shown that 3D aggregate structure and size can

affect cell growth, proliferation, and cell differentiation of PSCs (21–23). Therefore, with the

information obtained from CFD simulation and biological experiment, this project will be a

valuable contribution to bioprocess development of PSCs.

1.1 Research Aims

This project was divided into the following specific aims:

Specific aim 1: “Modeling of standard lab- scale spinner flask bioreactor” is described in

Chapter 3 of this thesis. In this chapter, the hydrodynamics of 100 ml bioreactor at various agitation

rates was revealed using CFD simulation. In addition, cell aggregate size distribution data of

Page 14: Computational Fluid Dynamics Modeling of Scalable Stirred

3

murine embryonic stem cells matched with the expectations from shear stress and energy

dissipation rate profiles.

Specific aim 2: “Modeling of scaled-down 10 ml bioreactors” is described in Chapter 4 of

this thesis. The hydrodynamics environment in 10 ml bioreactors obtained from CFD modeling

was compared with standard 100 ml bioreactor. The magnitude of volume average shear stress and

energy dissipation rate can be matched to the larger scale by increasing the impeller speed in the

10 ml bioreactor. The speed increment was determined based on CFD modeling. In addition, the

effects of impeller geometry are also described in this chapter.

Specific aim 3: Based on simulation data and biological responses, the effects of shear stress

and eddy sizes on cell expansion and aggregate size were clarified. The biological data is described

in each of chapter 3 and 4 for predefined conditions. Based on the results from 100 ml and 10 ml

bioreactors, optimal expansion condition for murine PSCs and bioreactor scaling criteria were

justified.

Page 15: Computational Fluid Dynamics Modeling of Scalable Stirred

4

Chapter Two: Literature Review

2.1 Pluripotent Stem Cells

2.1.1 Stem Cells and Regenerative Medicine Overview

Discovery of stem cells by James Till and Ernest McCulloch in 1963 is one of the most

remarkable medical research achievements of the 20th century (24). The two most important

characteristics that define stem cells are cell renewal, while keeping unspecialized state, and the

ability to differentiate into different cell types (25). In stem cell biology, proliferation is the ability

of stem cells to replicate themselves multiple times, and pluripotency refers to a stem cell that has

the potential to differentiate into any of the three germ layers including ectoderm, endoderm, and

mesoderm (26). Stem cells play a significant role in wound healing and organ repair processes. In

some organs, such as the gut and bone marrow, stem cells regularly divide. However, in some

organs, such as the pancreas and the heart, stem cells only divide under certain conditions (27).

Until recently, two types of stem cells that scientists have mainly focused on are embryonic stem

cells (ESCs) and somatic or adult stem cells (ASCs).

ESCs are pluripotent stem cells derived from the inner cell mass of an early-stage embryo

called the blastocyst (28). The inner cell mass containing ESCs gives rise to the entire body of an

organism including all of the specialized cell types and organs such as the heart, lung, skin, sperm,

egg, and many other tissues. Many published research findings have shown that ESCs can be

directed to differentiate into a target cell type by controlling the microenvironment of the cells and

using suitable growth factors (27). Current advanced biomaterials and medium development,

controlling cell differentiation or maintaining their pluripotency with high efficiency are possible.

Therefore, ESCs are ideal candidates for many disease treatments as well as drug testing models.

Page 16: Computational Fluid Dynamics Modeling of Scalable Stirred

5

On the other hand, ASCs are found among specialized tissues or organs, and those cells

have the ability to differentiate into multiple but limited number of cell types (1). The main roles

of ASCs are to maintain and repair the tissue or organ in which they reside. For instance, ASCs

from the kidneys normally differentiate into the cells found in the kidneys. The use of ASCs and

their derived tissues might help minimize or eliminate rejection after transplantation. ASCs have

a great potential in specific organ repair and regeneration, but compared to ESCs, they are limited

in their differentiation potential and replicative capacity (29).

2.1.2 Controversial Issues Related to ESCs

Though offering great promise for understanding human development and hope for new

disease treatments, human ESC research has also raised numerous ethical and political

controversies. The first reason is that human ESCs involves the destruction of a human embryo.

Many people believe that an embryo is a person, and therefore, an embryo also has interests and

rights that must be respected. Based on this perspective, isolating the human ESCs from the inner

call mass of a blastocyst is undistinguishable from murder (1). Many countries including the

United States, United Kingdom, and Japan have restricted human ESC research. In 2001, President

G.W Bush allowed federal funding for stem cell research using human ESC lines that had already

existed, while prohibiting the federal funding for the derivation and use of additional ESC lines

(30).

Another important issue associated with using human ESC therapy is patient safety. Unless

the transplanted cells are derived from the patient or autologous, they will be rejected by the hostile

immune system. It has been demonstrated that ESCs up-regulate histocompatibility antigen type I

and type II when they differentiate thus triggering the immune rejection responds (29). Although

Page 17: Computational Fluid Dynamics Modeling of Scalable Stirred

6

human ESCs have been accepted for use for drug screening purposes, it is difficult to obtain these

cells from the patients. Due to these limitations, human ESC-based therapies have faced numerous

challenges in getting clinical trials and FDA approval.

2.1.3 Induced-pluripotent Stem Cell Discovery and Its Significance

Scientists have continuously tried to find alternative cell sources that have similar functions

to human ESCs, but safer and does not involve the destruction of human embryos. In 2006, Dr.

Yamanaka’s research group in Japan identified four major transcription factors that regulate

pluripotency including Klf-4, Sox-2, Oct-4, and c-Myc (31). Their study shows that by over

expressing the genes encoding those four transcriptional factors, the mouse fibroblasts can be

reprogrammed to become pluripotent stem cells. Since somatic cells become pluripotent, they can

generate any tissue in the mouse.

In 2007, the same research group showed that the same four transcription factors could also

generate induced pluripotent stem cells (iPSCs) in human (25). The generated iPSCs are

genetically and functionally similar to ESCs. This research finding has a significant impact in stem

cell research and regenerative medicine. The research trend is currently shifting toward iPSCs

generation and differentiation rather than ESCs. Recognizing the remarkable discovery of iPSCs,

the Nobel Prize in Medicine or Physiology was awarded jointly to Dr. Shinya Yamanaka and Sir

John B. Gurdon in 2012 (29).

The generation of human iPSCs has overcome major limitations in ESCs. Since iPSCs

developed from a patient’s own somatic cells, they are more accessible than ESCs and can be

easily isolated from patient skin, blood samples, and tissues of interest without the destruction of

Page 18: Computational Fluid Dynamics Modeling of Scalable Stirred

7

human embryos (28). Therefore, unlike human ESCs, human iPSCs do not raise moral and ethical

controversies.

Because somatic cells from any individual can be reprogrammed into iPSCs, it is possible

to make disease-specific cell lines from patients. Hence, besides stem cell-based therapy, one of

the most important applications of human iPSCs is specific disease model study. Additionally,

iPSCs can be used for drug screening, which is a great deal for many pharmaceutical companies

(29). Traditionally, before introduced to the market, a drug has to pass intensive animal studies

and many phases of clinical trials. Those testing processes are complicated and costly. Depending

on the patient’s age, gender, and genetic background, the drug can have different effects. One of

the best way for drug screening is probably through using patient’s specific iPSCs rather than

animal models or traditional drug testing methods. Using iPSCs can help reduce testing duration,

simplify patient selection process, and possibly reduce the pre-market cost.

2.1.4 The Needs and Current Status of PSC Expansion

One of the main goals of generating iPSCs is to use them for clinical applications such as

heart repair after myocardial infarction or B-cell transplantation in diabetic patients. The number

of cells required for each cell-based therapy ranges from few tens of millions to few billions (5).

For instance, at least 109 cardiomyocytes are needed to replace damaged cardiac tissue after

myocardial infarction (MI) (32). Furthermore, about 1.3x109 insulin producing β cells per 70-kg

patient are required for insulin independence after islet transplantation (5). Therefore, it is

necessary to design a good process for reproducible large scale production of human iPSCs before

moving toward to clinical applications.

Page 19: Computational Fluid Dynamics Modeling of Scalable Stirred

8

Currently, human iPSCs are commonly derived and expanded in static T-flasks or tissue

culture dishes with the use of fibroblast feeder cells or matrix as supporting system. Static 2-D

culture systems do not adequately represent the in vivo environment of the cells. Although matrix

and feeders offer a great support for cell growth in static cultures, they often contain animal

components and some other factors that might have negative effects on the phenotype and

genotype of the cells as well as their environment. Static culture has high variability due to the

lack of automated control system. Besides, static culture is time-consuming due to feeder cell

separation, serial cell passaging, and handling requirements (33). In order to achieve 109 cells for

transplantation, about 500 of 10-cm tissue culture dishes are need (28). Therefore, expanding

iPSCs in 2-D static culture is highly inefficient and unfeasible.

2.1.5 Expansion of PSCs in Stirred Suspension Bioreactors

Suspension bioreactors (SSBs) that have been used for animal cell cultures and protein

productions can be adapted to produce the required number of iPSCs for clinical application.

Suspension system has many advantages compared to static culture. First, this is a 3-D feeder-free

culture system which can produce comparable or even higher cell expansion but less labor

intensive than static culture system (7, 9). Second, the system can be scaled-up to meet clinical

application needs (32). Third, SSB is a well-mixed system which allows the nutrients and oxygen

to be evenly distributed in the entire culture vessel (15). This is a great system for cells grown as

aggregates, like PSCs, since while the mass transfer is much more limited in static culture when

the aggregate size gets bigger. Fourth, the system can be automatically controlled so that the

variability is very low compared to manual static culture system (33). For instance, the DASGIP

bioreactors used in many cell culture laboratories and protein production facilities come with

Page 20: Computational Fluid Dynamics Modeling of Scalable Stirred

9

different vessel sizes as well as a temperature, dissolved oxygen, and pH control system. Based on

these advantages, SSB is currently one of the most feasible and practical system for human iPSC

expansion.

Recent study of Chen at al. (32) shows that human ESCs can be grown in spinner flask

culture system for over 20 passages without loss of cell viability and growth kinetic. Based on

flow cytometry, a biomarker detection assay, the study demonstrates that those cells remain

pluripotency and normal karyotype. The resulted cumulative fold expansion after 21 passages is

about 1013. Since the culture system used is a robust scalable system for human ESC production

under GMP conditions, it has a great potential for future clinical application. Although the study

describes the production of human ESCs, the system would also be suitable for production of

human iPSC since both cell types have similar genotype, phenotype, and usually identical growth

conditions.

2.1.6 Differentiation of PSCs in Stirred Suspension Bioreactors

Differentiation is one of the most critical steps of cell production for transplantation. In

fact, the cells used for transplantation are preferably differentiated cells. For instant, iPSCs need

to be differentiated to cardiomyocytes before injecting to an MI patient. The ability to differentiate

to different cell types is a great advantage but also a problem when dealing with human ESCs and

iPSCs. For a single cell transplantation, only a single cell type is needed for a target tissue or organ.

However, it is very difficult to differentiate all the pluripotent cells into one single specific cell

type. Therefore, differentiation efficiency is normally low (34).

Many previous studies were successful in expanding human ESCs and iPSCs in suspension

bioreactors while maintaining the pluripotency of the cells. However, differentiation of the cells

Page 21: Computational Fluid Dynamics Modeling of Scalable Stirred

10

into a specific cell type in suspension culture has not been successful or very inefficient (15). One

study suggested that bioreactor culture many induce pluripotency and reduce the differentiation

efficiency through fluid shear stress (14). This can be an example of mechanotransduction

phenomenon which is the conversion of mechanical signals that the cell senses from its

environment to internal biochemical signals (35). Mechanical cures can change the cell phenotype,

proliferation, and differentiation. Therefore, in addition to biochemical factors, mechanical forces

or stresses can be used to direct the cells to differentiate into a specific cell type (8) .

Recently, due to the development of special reagents and media as well as differentiation

techniques, the efficiency of tuning cell fate in suspension bioreactor has improved. Chen et al.

(2012) also show that cardiomyocytes can be directly from pluripotent cell aggregates in SSBs.

The cell aggregates were differentiated on day 3 and the contracting differentiated cells were

observed on day 9. They did cell specific marker staining and found that about 27% percent of the

cell population was differentiated to cardiomyocytes.

2.1.7 Challenges in PSC Bioprocess Development

Human ESCs and iPSCs grow as adherent cells in static culture (34). Since these cells are

highly sensitive, they often need a support system which can be mouse fibroblast feeder cells or

extra cellular matrix components. Before transferring them to suspension culture, the cells need to

be expanded to a desirable cell quantity in static culture first. Moving the cells from static culture

with the support system to suspension culture system without the support system can be

problematic. Changes in surrounding environment can cause changes to cell growth kinetics as

well as cell phenotype (2).

Page 22: Computational Fluid Dynamics Modeling of Scalable Stirred

11

One special biological aspect that need to be taken into account while culturing human

iPSCs is that these cell survive and grow as aggregates (9). Forcing them to grow as single cells

after traditional cell passaging often result in significant cell death as well as a reduction in growth

rate (32). Due to this reason, many current techniques suggest to break the cells into small

aggregates from the big aggregates rather than into single cells. However, it is difficult to control

the new cell seeding density since the number of cells in each small aggregate can only be

estimated based on the size, and counting the number of seeding aggregates is also difficult.

Expanding iPSCs while maintaining their normal karyotype is challenging. It is difficult and

almost impossible to achieve homogeneous cell population since some of the cells may go under

spontaneous differentiation while some of the cells in the same culture environment try to maintain

their pluripotency (33). There are many methods to help isolate the differentiated from pluripotent

cells, but the separation process is time-consuming, especially in larger scale like suspension

bioreactor.

One of a major concern when culturing animal cells especially human ESCs and iPSCs in

suspension bioreactor is the level of shear stress. Shear stress level can be adjusted by changing

the agitation rate. It is important to run the bioreactor at an optimal speed that does not cause

damages to shear-stress sensitive cells while maintaining good mixing (15). The optimal shear

stress level is different for each suspension bioreactor system with different geometry, cell density,

and culture media viscosity (34). The most common method has been used to prevent the cells

from shear damage is by trial and error. For instance, the optimal shear stress is determined by

observing the growth kinetics of the cells at different speeds. However, this method can only give

the maximum or the minimum optimal shear stress level, but does not give the full shear stress

profile that represent the actual system. Shear stress has been found to play a role in endothelial

Page 23: Computational Fluid Dynamics Modeling of Scalable Stirred

12

cell formation and proliferation. The effects of shear stress in iPSC pluripotency and their

differentiation capability are not yet clear and need to be investigated.

It is important to have an automated system that not only monitors but also control pH,

dissolved oxygen, and agitation rate. The quantity of the cells has to come side-by-side with the

quality. Designing a good cell manufacturing process is one of the key step for human iPSCs to

get closer to cell-therapy clinical application. The quality of the iPSCs is not only dependent on

the cell source and reprogramming methods but also the expanding and differentiating process.

There have been GMP qualified suspension bioreactor system for other cell types and protein

productions, but not yet for human iPSCs (32).

2.1.8 The Significance of Controlling Aggregate Size

Aggregate size is one of the key biological parameters that needs to be controlled in

bioprocess development of PSCs (23, 36). The transport of oxygen and nutrient within the cell

aggregates is significantly influenced by the size and structure of these aggregates and the effects

of hypoxia have found to be noticeable for aggregates larger than 300 μm (37) . As the aggregate

size exceeds the tolerance limit, the diffusion process of nutrients into the core of the aggregate

becomes limited, leading to various problems such as decrease in cell proliferation and possible

necrosis toward the center of the aggregates (38, 39). For instance, study by Ungrin et al. (21) has

shown that optimal size of aggregates can enhance the expansion of human PSC in suspension

culture and the cell growth declines in excessively large aggregates.

Furthermore, aggregate size also plays an important role in PSC differentiation. One of the

most popular methods for differentiating PSCs is through the generation of heterogeneous 3-D

embryoid bodies (EBs) (40). An aggregate of PSCs in suspension culture that has the ability to

Page 24: Computational Fluid Dynamics Modeling of Scalable Stirred

13

from different cell types of all three germ lineages is considered as an EB (41). There are different

methods for generating EBs. Standard methods such as hanging drops, liquid suspension, and

methylcellulose culture have limited scalability and are difficult to control (42). On the other hand,

EB generation in SSBs is a highly scalable process and is capable of regulating the EB size (41).

It has been found that the EB size greatly influences PSC differentiation via impacting the

environmental factors that affect the stem cell differentiation such as the diffusion of soluble

molecules and cell adhesions (41). Therefore, understanding the effects of hydrodynamic

environment of SSBs on aggregate size is an essential step in bioprocess development of PSC

expansion and differentiation.

2.2 Computational Fluid Dynamics (CFD)

2.2.1 The role of CFD Modeling in Bioprocess Development

CFD has been used extensively in modeling various bioprocesses and hydrodynamics of

various bioreactor types (18). Stirring in a bioreactor is needed in order to achieve good mixing

within bioreactor vessels, but it can also cause some potential damage to the cells if the level of

shear forces caused stirring goes beyond the cells’ tolerance limit (43). The cells are dispersed

throughout the entire bioreactor system so that the level of shear stress needs to be calculated

locally and globally, not just at the tip of the impeller. Specific local hydrodynamics is difficult

and almost impossible to measure experimentally. Therefore, CFD has become an important tool

for modeling the hydrodynamic environment of bioreactor with various configurations (18, 44).

In bioreactor design, selecting a suitable impeller is one of the key steps. However,

redesigning and experimentally testing the impeller performance can be time consuming and

costly. Fortunately, CFD modeling can help in choosing the right impeller and rotating speed and

Page 25: Computational Fluid Dynamics Modeling of Scalable Stirred

14

the performance of the impeller can be directly simulated in CFD modeling within a relatively

short period of time with minimal labor cost (10).

Scaling-up and scaling-down are very important and popular in bioprocess development.

Before building production scale bioreactors, extensive rounds of testing and screening need to be

done at a small economical scale (45, 46). However, maintaining the same hydrodynamic

environment between small and large scale is challenging. There are different criteria using in

scaling of bioreactors. These include maintaining of tip speed, Reynolds number, and specific

power input (47). In order to verify the validity of these criteria, CFD modeling is needed.

The majority of work on CFD modeling of SSBs is from bioprocess development of

monoclonal antibodies (12). The main focus of these studies is to find the optimal mixing

environment to achieve highest protein production (11, 13, 18). The production scale of the

bioreactors used in protein production can go up to 25,000 L (13). Large scale bioreactors for

protein production often contain many problems such as mass transfer limitation and high shear

stress near the tip of the impeller (48). Since the final desired product is protein, the optimal

conditions are often tailored to specific production rate rather than maintaining the consistency of

the cell phenotype and genotype (49). However, this is not the case for bioprocess development of

PSCs. Since the final goal is to obtain qualified cells for transplantation, the effects of

hydrodynamics on the cell phenotype and genotype need to be carefully evaluated. Unlike CHO

cells which are grown as single cell suspension in SSBs, PSCs are grown as 3-D aggregates so that

the effects of hydrodynamics on the cells are different. Although it has been known that the size

of the cell aggregates can be altered by changing the agitation rate, the exact mechanism has not

been cleared (2). Modeling the hydrodynamic environment using CFD simulation and

Page 26: Computational Fluid Dynamics Modeling of Scalable Stirred

15

experimentally testing the growth of PSCs in small lab-scale SSBs can help clarify indicated

aggregate size control process.

2.2.2 Turbulent Flow

Flow in SSBs has often been characterized as turbulent flow. There exist different types of

flows in nature. For instant blood flow through capillaries is laminar, and air flow near the jet is

considered turbulent. There is no definite definition for turbulent flow. However, there are some

key characteristics that help define turbulent flow (50). These include:

Large Reynolds number: Reynolds number is a dimensionless number, defined as the

ratio of momentum force to viscous forces. The flow is turbulent when Reynolds number

is large and dominated by inertia forces.

Irregularity: The flow is random and chaotic, consisting of different scales of eddies. The

largest eddies are of the order of flow geometry such as boundary layer thickness, and the

smallest eddies are formed by viscous stresses dissipated into internal energy.

Dissipation: In turbulent flow, kinetic energy in the small eddies are transformed into

internal energy. The small eddies receive the kinetic energy from slightly larger eddies

and these slightly larger eddies receive energy from larger eddies and so on. This process

of multiscale eddy energy transfer is called cascade process. The largest possible eddies

are those that extend across the entire system. The smallest eddies are set by viscosity,

and smaller eddies, the stronger the velocity gradient and the more important the effect of

viscosity. Therefore, smallest eddy scale (Kolmogorov length scale) can be defined as:

η=ν3/4ε-1/4, where ν is fluid kinematic viscosity (m2/s) and ε is volume average rate of

turbulent energy dissipation (m2/s3).

Page 27: Computational Fluid Dynamics Modeling of Scalable Stirred

16

Diffusivity: The diffusivity in turbulent flow is high and mixing increases. The

turbulence escalates the exchange of momentum in boundary layers and lessens the

delays thus separation at bluff bodies such as cylinders.

Continuum: Although there are eddies in the flow, they are much larger than molecular

scales and therefore, the flow is treated as continuum.

2.2.3 Governing Equations in CFD Modeling of SSBs

Most CFD simulations are based on the governing equations of motions for a continuous

viscous fluid: Navier-Stokes equations. The key principles behind these equations are conservation

of mass, conservation of momentum, and conservation of energy (51). The hydrodynamic

properties of turbulent incompressible fluid can be described in Reynolds-Averaged Navier Stokes

equations (52) (see List of Symbols at beginning of thesis):

∂∂t

. . , . 0 2.1

Where dynamic viscosity depends only on the physical properties of fluid, and turbulent eddy

viscosity is affected by velocity fluctuation ′.

One of the most common used turbulent model in CFD is two equation k-ε model. The two

equation model relates how the energy is transferred from the larger to smaller eddies, and it can

predict the behaviors of a turbulent flow without previous knowledge of turbulent structure (53).

Equation 2.1 is complemented by two additional convection-diffusion-reaction equations in k-ε

model (54):

. k 2.2

Page 28: Computational Fluid Dynamics Modeling of Scalable Stirred

17

. 2.3

Where and the production of turbulent kinetic energy due to the mean velocity

gradient | | . The default values for model constants are: 0.09,

1.44, 1.92, 1.0, 1.3.

The k-equation (2.2) is a transport equation for turbulent kinetic energy k (m2/s2), and the ε-

equation (2.3) is an equation for the dissipation rate of turbulent kinetic energy ε (m2/s3). This

model has been proven to give reasonably good results for free-shear-layer flow with relatively

small pressure gradient (52). The k-ε model is also the most widely used in simulating the

hydrodynamic environment in SSBs (18, 19, 55), and has been validated with particle image

velocimetry (56).

2.2.4 Finite Element Method

The finite element method is one of the methods used in CFD to solve complex system of

partial differential equations from Navier-Stokes equations. Finite element method is preferred in

solving systems with complex and irregular geometries (57). Essentially, each continuous

geometry is meshed into discrete domain or grid with finite small elements, and the number of

nodes are the numbers of degrees of freedom needed to be solved. There are various finite element

shapes, and the choice of a finite element depends on the shape of original geometry, physics of

the problem, and the level of accuracy required in the final solution (58). In finite element method,

the global functional representation of a variable consist of an assembly of local functional

representations so that the global boundary conditions can be applied at local finite elements by

modification of a system of algebraic equations (58). This process of forming global representation

Page 29: Computational Fluid Dynamics Modeling of Scalable Stirred

18

from local representations is known as interpolation. It is important to carefully choose an

appropriate interpolation method conserve the continuity between adjacent elements and to

enhance the accuracy of the solution (59).

2.3 Conclusions

It has been shown that hydrodynamic environment in SSBs affects PSC expansion and

differentiation (2, 14). The first key step in PSC bioprocess development is to understand the

hydrodynamics within the bioreactor from CFD simulation. The k-ε model in CFD can be used to

predict the behaviors of a turbulent flow without previous knowledge of turbulent structure (53) .

The next step is to apply CFD results for scaling-up and scaling-down bioprocesses, for example,

selecting the right impeller design and speed to produce optimal cell expansion and aggregate size.

Page 30: Computational Fluid Dynamics Modeling of Scalable Stirred

19

Chapter Three: Hydrodynamics of Standard Lab-scale 100 ml Spinner Flask Bioreactor

3.1 Introduction

Pluripotent stem cells (PSCs) play a key role in regenerative medicine due to their self-

renewal and capability to differentiate into multiple lineages (2). Recently, with the discovery of

induced pluripotent stem cells (iPSCs), it may be possible for these cells to be used in clinical

applications such as cell-based transplantation. However, in order for cell-based transplantation to

be widely used, robust cell manufacturing processes that produce large quantity of cells with high

quality are needed. The average number of cells required per dose of transplantation is of the order

of billions (5) which is equivalent to the total number of cells harvested from more than 500

conventional static T75 tissue culture flasks. Therefore, using static culture to generate PSCs for

transplantation are impractical, costly, and labor intensive.

Stirred suspension bioreactors (SSBs) have been used to expand variety of cell types

including both anchorage dependent cells and anchorage independent cells on micro-carriers (7,

60). This system offers many advantages including scalability, a well-controlled environment,

good mixing to enhance mass transport, and less labour and material cost compared to static culture

(5). Both mouse and human PSCs have been successfully expanded in SSBs in many laboratories

without using micro-carriers (7, 61) even though the cells are anchorage dependent in static culture.

It has been shown that PSCs can be expanded in SSBs by using multiple passages while

maintaining high pluripotency (28, 62). However, differentiation in SSBs has remained a challenge

(2, 14, 15). Taiani et al. (14) showed that in bioreactor-differentiated cultures, a subpopulation of

cells in aggregates expressed the pluripotency marker Oct-4 in cell nuclei. Furthermore, Oct-4

expression still remained even after 30 days of SSB culture without leukemia inhibitory factor

Page 31: Computational Fluid Dynamics Modeling of Scalable Stirred

20

(LIF). This phenomenon raised questions about the effects of hydrodynamic environment on the

pluripotency and differentiation ability of PSCs. Study of Lara et al. (2013) demonstrated that

multiple aspects of hydrodynamic environment, such as shear stress and flow regime, affect

pluripotency marker expression. Specifically, flow exposure not only helped to maintain Oct 4

expression but also increased Oct-4 expression after 36 hours in the absence of LIF. The gene

expressions for Oct-4 and other pluripotency markers such as Nanog and Rex1 were also

significantly higher than those in static culture for the same culture duration, seeding density, and

media components (17).

The effects of hydrodynamic environment on PSCs have been shown in numerous studies

(2), but the exact mechanisms of these effects have not been fully discovered. The first step in SSB

bioprocess development is to have a good understanding of hydrodynamics within the bioreactor

and how it can affect quantifiable biological properties such as cell aggregate size and expansion.

Computational fluid dynamics (CFD) modeling is a powerful tool to investigate the

hydrodynamics inside SBBs. CFD enables one to obtain important information about the fluid

environment such as local shear stress level and energy dissipation rate that are nearly impossible

to be obtained experimentally (19). Based on many existing studies, the total shear stress acting

on the cells and aggregates comes from two main sources which are the local fluid velocity gradient

and cell interaction with turbulence eddies (19, 43). The magnitude of fluid velocity gradient is

given by:

| | (3.1)

Shear stress can be calculated from velocity gradient as:

Page 32: Computational Fluid Dynamics Modeling of Scalable Stirred

21

(3.2)

where τ (Pa) is the shear stress; γ (s-1) is shear rate or velocity gradient; u, v, and w (m/s) are the

velocity in x, y, and z direction; and μ (Pa*s) is the fluid viscosity. For Newtonian fluids such as

water, the viscosity is constant. For non-Newtonian fluids, the viscosity is a function of shear rate,

temperature, and exposure time.

Aggregates can also be sheared from the eddies with similar size or smaller (19, 43). The

size of the small eddies, η, depends on the energy dissipation rate, є, and kinematic viscosity ν,

specifically η=(ν3/є)1/4 (43). Cell aggregates have been found to be controlled by hydrodynamics

in SSBs (2), but exact mechanisms have not been cleared. Moreover, aggregate sizes are often

mentioned in the context of agitation rate which is not suitable representation of actual

hydrodynamic environment in bioreactors with different vessel and impeller geometries. The goal

of this study was to investigate the hydrodynamic environment inside a standard lab-scale 100 ml

spinner flask by using CFD simulation and determine how certain hydrodynamic variables such

as shear rate and turbulent eddies can affect aggregate forming cells such as PSCs.

3.2 Materials and Methods

3.2.1 Cell culture and Cell Aggregate Measurements

The D3 cell line of murine ESCs (provided by Dr. Rancourt, University of Calgary) were

grown on murine embryonic fibroblasts for one passage upon thawing and then were routinely

passaged twice into gelatin-coated petri dish (BD BioSciences, Bedford, MA, USA). The

expansion medium for murine ESCs consisted of 15% fetal bovine serum (Gibco, Grand Island,

NY), 1% non-essential amino acids (Gibco), 1% penicillin and streptomycin (Gibco), 0.1 mM β-

Mercaptoethanol (Sigma, St.Louis, MO), high glucose DMEM (Gibco), and 1,000U/ml leukemia

Page 33: Computational Fluid Dynamics Modeling of Scalable Stirred

22

inhibitory factor (EMD Millipore, Billerica, MA, USA). In suspension cultures, murine ESCs were

inoculated as single cells into duplicate 100ml spinner flasks (NDS Technologies, Vineland, NJ,

USA). Spinner plate speeds at 80, 100, and 120 rpm were calibrated by using a Touchless Digital

Tachometer (VWR). All cell cultures were maintained at 370C and 5% CO2. The size distribution

of murine ESCs aggregates were measured on the fifth day of suspension culture using Beckman

Coulter Multisizer III (Miami, FL, USA). For each run, murine ESC aggregates were diluted such

that the coincident counting rate was between 3-10%. Samples were counted for 20 seconds per

run, which returns about an average of 3000 events. An average of 3 runs were collected per

sample. After collection, the data were analyzed using the complementary Multisizer 3.53

software. Background signal from debris and single cells were gated out of the data range by

excluding particles smaller than 40 μm in diameter.

3.2.2 Calculations

3.2.2.1 Maximum Shear Stress Calculation

The maximum shear stress in SSB can be calculated from (63):

τMax 5.33ρ єν (3.3 )

where τMax (Pa) is the maximum shear stress, ρ (kg/m3) is the fluid density, ν (m2/s) is the kinematic

viscosity, and ε (m2/s3) is energy dissipation rate which relates to vessel geometry, impeller

geometry and properties of fluid via (63):

(3.3 a)

where VL (m3) is the fluid volume and P (W) is the power input which is estimated from:

Page 34: Computational Fluid Dynamics Modeling of Scalable Stirred

23

P NP N3 Di5 ρ (3.3 b)

where N (rotation per second) is the agitation rate, Di (m) is the impeller diameter, and NP is the

power number that is defined by:

(3.3 c)

NP can be estimated according to Nagata (63) s follows:

. .

. . (3.3 d)

. (3.3 e)

(3.3 f)

. . . (3.3 g)

. . . . (3.3 k)

3.2.3 Computational Fluid Dynamics Modeling

3.2.3.1 Geometry

Standard lab-scale 100 ml bioreactor have been used extensively in expansion of both

murine and human PSCs (7, 61, 62). The bioreactor consists of a glass vessel and a top mounted

magnetic based impeller. In addition, there is a small indentation (9 mm) at the center bottom of

the flask which helps reduce the dead space right below the impeller. The geometry built in

COMSOL Multiphysics is only up to the level of fluid height (Figure 3-1).

Page 35: Computational Fluid Dynamics Modeling of Scalable Stirred

24

Figure 3–1: Model geometry built in COMSOL Multiphysics. A) 3D view of fluid occupied region in 100ml bioreactor; B) Top view of model geometry

Table 3–1: Dimensions of 100 ml model

Dimension Scale and unit Vessel inner diameter 0.064 (m) Fluid height 0.034 (m) Impeller shaft diameter 0.0159 (m) Impeller shaft length (height) 0.019 (m) Impeller diameter (horizontal length) 0.051 (m) Impeller width (thickness) 0.012 (m) Indentation height from the bottom vessel 0.009 (m)

3.2.3.2 Model Physics and Grid Generation

The flow is considered to be fully turbulent when the Reynolds’ number, defined by

Re=N(Di2)/υ, is greater than 20,000 and fully laminar when it is less than 10 (55). The type of flow

in between fully laminar and fully turbulent is transition flow which is difficult to characterize. In

this study, Re ranges from 4.2x103 to 6.4x103 which is closer to the turbulent regime. As a

consequence, in the CFD model, the Reynolds Averaged Navier-Stokes (RANS) k-ε technique

was chosen to model the flow in 100 ml at operated agitation rates from 80 rpm to 120 rpm. The

Page 36: Computational Fluid Dynamics Modeling of Scalable Stirred

25

k-ε model is very popular for industrial applications due to its good convergence rate and relatively

low memory requirements compared to other turbulent models (64). All the solid walls are

presented by non-slip conditions and the free surface is specified with an open boundary

(atmospheric pressure) condition; the effect of capillary effects are not taken into account. The

model assumed that the fluid properties are as same as water as the majority of cell culture medium

is made up of water.

The commercial finite element multiphysics software package COMSOL Multiphysics

Version 4.4 (COMSOL, Inc. California, USA) was used to model the hydrodynamics of 100 ml

bioreactor. The simulations were run on Intel dual core Xeon 3.30 GHz with 24 GB RAM. All

cores were under load during each simulation and about 8 Gb out of 24 Gb of memory was

allocated for COMSOL Multiphysics.

The finite element grid was created by using the physics-controlled mesh generation option

in COMSOL. The finer grid (Figure 3-2A) consists of 1,117,436 finite elements whereas the

normal grid (Figure 3-2B) consists of 350,161 finite elements. All simulations at different agitation

rates were run using fine grid with 1,629,259 degrees of freedom. Each speed was simulated for

12s duration with average CPU run time of 96 hours. Data from the simulation were then exported

into Excel spreadsheet for further analysis. Most flow visualizations were performed by using

built-in post-processing tools within the COMSOL Multiphysics package.

Page 37: Computational Fluid Dynamics Modeling of Scalable Stirred

26

Figure 3–2: 100 ml bioreactor model grids created in COMSOL Multiphysics. A) Finer grid (948,668 tetrahedral finite elements); B) normal grid (278,951 tetrahedral elements)

3.3 Results and Discussion

3.3.1 Grid Dependence and Model Validation

In CFD modeling, grid generation is a crucial step that affects the convergence, solution

accuracy, and CPU time (53). Due to the strong interaction of the mean flow and turbulence,

turbulent flow numerical solutions and results are more susceptible to grid dependency than

laminar flow (65, 66). In this study, 100 ml model at 100 rpm were solved using two different grid

sizes: normal grid consists of 278,951 tetrahedral finite elements and fine grid consists of 948,668

tetrahedral finite elements. The results from two grids sizes were compared and plotted on Figure

3–3 . The average pressure was the same in both, and the difference in average velocity and shear

stress are less than 2.5%. The difference in average energy dissipation rate was 8%, but the

difference in average eddy scale was only 2%. Since the differences in solutions between two grid

sizes were relatively small, no further grid refinements were carried out. Figure 3–4 shows

A B

Page 38: Computational Fluid Dynamics Modeling of Scalable Stirred

27

maximum shear stress results obtained from equation 1 and CFD simulation. The difference was

2% at 120 rpm, 4.5% at 100 rpm, and 9.5% at 80 rpm. As expected, higher agitation rate results in

larger Reynolds number and greater maximum shear stress. In this study, the flow at 120 rpm is

closest to fully turbulent which the CFD model and equation 1 calculation precisely described.

Figure 3–3: Comparison of results of fine grid (948,668 tetrahedral finite elements) and normal grid (278,951 tetrahedral elements).

0

0.2

0.4

0.6

0.8

1

1.2

Avg Pressure Avg Velocity Avg Shear Rate Avg EnergyDissipation rate

Avg Eddy Size

Rel

ativ

e R

atio

Model Results

Relative Results from Fine Grid and Normal Grid

Fine Grid Normal Grid

Page 39: Computational Fluid Dynamics Modeling of Scalable Stirred

28

Figure 3–4: Maximum shear stress calculated from equation 1 and simulated using CFD

3.3.2 Base Case: Standard 100 ml Spinner Flask Bioreactor

Previous studies in the same bioreactor system have shown that murine and human ESCs

grew best at 100 rpm (7, 61) which was chosen as the base case for this study. In this study, the

CFD simulations give constant profiles of velocity, shear rate, and energy dissipation rate after

0.000

0.200

0.400

0.600

0.800

80 rpm 100 rpm 120 rpm

Max

imu

m S

hea

r S

tres

s (P

a)

Agitation rate in 100 ml Bioreactor (rpm)

Maximum Shear Stress

Calculated Max Shear Simulated Max Shear

Page 40: Computational Fluid Dynamics Modeling of Scalable Stirred

29

about 3s (see Appendix B for detailed profiles). All results in this study were extracted from 10s

or beyond

Figure 3-5 shows distributions of the pressure, velocity, shear stress, and energy dissipation

rate. Both pressure and shear stress can exert forces to the cells but with different magnitude and

direction. The pressure in the bioreactor was essentially hydrostatic pressure which depended on

fluid height and gravity. Many studies showed that the effects due to hydrostatic pressure are

minimal compared to the effects due to shear stress (67). Therefore, in a small biological system

like 100 ml spinner flask, the effects of hydrostatic pressure on the cells are often neglected.

The velocity magnitude of velocity is highest at the tip of the impeller as expected. Below

the impeller, the flow is less turbulent and the mixing is not as active as in the region near the tip

of the impeller. The dimple-like indentation on the center bottom of the flask was designed to

prevent the stagnation of the fluid in that area. Without this indentation, the cells in this region

would grow into large cell aggregates and cell death would occur due to mass transfer limitation.

Figure 3-5B showed strong flow separation near the tip of the impeller. This suggests not only

good mixing around the tip of the impeller but also high shear stress (Figure 3–5C) and energy

dissipation rate (Figure 3–5D). High shear stress and small eddy due to high energy dissipation

rates at this region have been reported as the sources of cell damage and cell behavior changes in

SSBs (63). However, the significantly high shear stress region only occupies a small fraction of

the bioreactor volume. More than 95% of bioreactor volume produced shear rate level less than 50

s-1 which was too low to cause cell damage. However, small level shear stress levels might cause

some changes in genetic construction of the cells which needs further investigation from

mechanotransduction studies.

Page 41: Computational Fluid Dynamics Modeling of Scalable Stirred

30

Figure 3–6 showed how the velocity profile on yz (x=0) plane changed over time. Initially,

the impeller is perpendicular to the cut plane with 0 velocity. The impeller rotates counter

clockwise at the rate of 100 rotations per minute, and the velocity profile is plotted at 10s, 11s, and

12s. The results show that the velocity magnitude is highest and flow separation is strongest at 12s

when the impellers has just passed the plane. When the tip of the impeller moves further away

from the cut plane as in 10s, the velocity decreases and weaker radial flow is observed. At 11s, the

impeller is moving closer to the cut plane, the velocity started to increase to slightly higher than

the velocity obtained at 10s, but much lower than on obtained at 12s.

Figure 3-7 shows the changes in velocity, shear rate, and energy dissipation rate at different

horizontal or z- planes. At the middle of the impeller, z=0.02m, the velocity was highest. It is

consistent with Figure 3-6 that the velocity is higher in the wake region side of the impeller. Shear

stress and energy dissipation rate are also highest near the tip of the impeller. Velocity, shear rate,

and energy dissipation rate reduced significantly as the cut plane is further away from the impellers

(Table 3-2).

These results suggested that impeller speed and position play an important role in

magnitude and distribution of velocity, shear rate, and energy dissipation rate in the bioreactor.

Furthermore, strongest mixing happens near the tip of the impeller, but potential damages to the

cells are also highest around this region. In actual practice of bioreactor culture, the height of the

impeller was adjusted to a consistent level to obtain good mixing and low variability between runs.

Page 42: Computational Fluid Dynamics Modeling of Scalable Stirred

31

Figure 3–5: Results of basics hydrodynamic variables obtained from CFD simulation. A) Pressure profile; B) Velocity and streamlines; C) Shear rate; D) Energy dissipation rate

A) Pressure B) Velocity and Streamline

C) Shear Rate D) Energy Dissipation

Page 43: Computational Fluid Dynamics Modeling of Scalable Stirred

32

Figure 3–6: Velocity on cut plane yz (x=0) plane. A) At t=0s, the impeller is perpendicular to the cut plane, velocity =0; B) At t=10s, the impeller is perpendicular to the cut plane, maximum velocity =0.119 m/s; B) At t=11s, the impeller is 300 from entering the cut plane, maximum velocity=0.129 m/s; D) At t=12s, the impeller is 300 leaving the cut plane, maximum velocity =0.232 m/s

A B

C D

Page 44: Computational Fluid Dynamics Modeling of Scalable Stirred

33

Figure 3–7: Velocity (A), shear rate (B), and energy dissipation rate (D) at different fluid heights in 100 ml bioreactor model. Cut plane 1 (z=0.03 m) on the left side is the horizontal plane above the impeller. Cut plane 2 (z=0.02 m) in the middle is the horizontal plane through the center of the impeller. Cut plane 3 (z=0.01 m) on the right is the horizontal plane below the impeller

A) Velocity (m/s)

C) Shear Rate (s-1)

B) Energy Dissipation rate (m2/s3)

Page 45: Computational Fluid Dynamics Modeling of Scalable Stirred

34

Table 3–2: Average velocity, shear rate, and energy dissipation rate at different fluid heights

Above the impeller (z=0.03 m)

Middle of Impeller (z=0.2 m)

Below the Impeller (z=0.01 m)

Velocity 0.047 (m/s) 0.096 (m/s) 0.051 (m/s)

Shear rate 12.04 (s-1) 30.6(s-1) 7.47 (s-1)

Energy Dissipation rate 1.19E-3(m2/s3) 1.09E-2(m2/s3) 7.48E-4 (m2/s3)

3.3.3 Hydrodynamics of 100ml Bioreactor at Different Agitation Rates

Agitation rate is meaningless alone without considering vessel dimensions, impeller

geometry, and fluid properties. Instead of agitation rate, other characteristic flow parameters such

as velocity, shear rate, and eddy scale should be used for each particular bioreactor system.

Figure 3-8 shows distributions of the velocity, shear rate, and energy dissipation rate on

middle horizontal cut plane (z=0.02m) at 12s. On this plane, the effects of changing agitation rate

are strong compared to other horizontal distances in the entire bioreactor volume. At 12s, the

impeller position is at different locations with different velocity magnitudes depending on the

agitation rate. As expected, increase in agitation rate leads to increase in maximum and volume

averaged velocity magnitude, shear rate, and energy dissipation rate (Figure 3-8 and Table 3-3). If

20% increase in agitation rate from 100 rpm to 120 rpm gives a rise to about 20% increase in shear

rate, the increase in energy dissipations rate went up by 60%. Significant increase in energy

dissipation rate is expected based on the relationship between dissipation rate є and agitation rate

N described in equation 3.3. Besides, 25% increase in agitation rate from 80 rpm to 100 rpm led

to 25% increase in shear rate and 76% increase in energy dissipation rate. Therefore, it was

Page 46: Computational Fluid Dynamics Modeling of Scalable Stirred

35

expected that the difference in cell aggregate size between 80 rpm and 100 rpm would be larger

than between 100 rpm and 120 rpm.

Figure 3–8: Velocity (A), shear rate (B), and energy dissipation rate (C) at different agitation rates (80 rpm, 100 rpm, and 120 rpm from left to right). All surfaces were plot at 12s of simulation

A) Velocity

B) Shear Rate

C) Energy Dissipation Rate

Page 47: Computational Fluid Dynamics Modeling of Scalable Stirred

36

Table 3–3: Volume average velocity, shear rate, and energy dissipation rate at different agitation rates

80 rpm 100 rpm 120 rpm

Velocity 0.048 (m/s) 0.0586 (m/s) 0.0690 (m/s)

Shear rate 11.27 (s-1) 14.1 (s-1) 16.9 (s-1)

Energy Dissipation rate 1.79E-3 (m2/s3) 3.15E-3 (m2/s3) 5.04E-3 (m2/s3)

3.3.4 Relationships between Hydrodynamic Parameters and Aggregate size

3.3.4.1 Effects of shear rate on aggregate size

The biological results of cell aggregate size distribution shown in Figure 3-9 agreeably

reflected the CFD results above. The average aggregate diameter from Multisize measurement at

100 rpm and 80 rpm were similar to the average aggregate diameters at the same agitation rates in

previous study by Comier et al. (7) demonstrating the consistency of cell aggregate generation in

used bioreactor system. The difference in cell aggregate size between 80 rpm and 100 rpm (33%)

was indeed larger than between 100 rpm and 120 rpm (13.7%). When mode of aggregate size from

100 rpm sample (152 μm) and 120 rpm sample (130 μm) almost overlapped, the mode size

distribution from 80 rpm sample distinctly shifted to the right (236 μm) Therefore, larger change

in shear rate and energy dissipation rate indeed lead to larger change in cell aggregate size.

Figure 3-10 shows the relationship between shear rate and cell aggregate size. Higher

maximum shear rate and average shear rate leads to smaller average cell aggregate size. The error

bars in the graphs are the standard deviations which indicated how wide the aggregate size spreads

throughout the bioreactor volume. The number of small aggregates in 120 rpm sample is highest

Page 48: Computational Fluid Dynamics Modeling of Scalable Stirred

37

due to highest shear and energy dissipation rate. Although volume-averaged shear rate and energy

dissipation rate at 80 rpm are low and most of the aggregates were large, some of the aggregates

might be broken up near the tip of the impeller. Hence, the 80 rpm sample has largest standard

deviation among three tested agitation rates. The aggregate size at 100 rpm is most uniformly

distributed with less small aggregates (less than 70 μm) compared to other two agitation rates.

Therefore, 100 rpm has been reported to be the most optimal agitation rate for growing murine

ESCs in the same bioreactor system (7).

The relationship between eddy size and cell damage have been a topic of debate in

bioreactor bioprocess development for many years. It was found that the relationship between cell

aggregate size and energy dissipation rate was similar to the relationship between eddy size and

energy dissipation rate (63). Therefore, cell aggregates would have similar size as eddy size.

Before CFD simulation was used popularly in modeling hydrodynamics of SBBs, only minimum

Komogorov length scale was calculated based on maximum energy dissipation rate near the tip of

the impeller and known kinematic viscosity. It was hypothesized that potential cell damages occur

when eddy scale was smaller or equal to the size of single cell or cell aggregate (43). However,

the details of this relationship has not been clear. In addition, the hydrodynamic environment of

SSB is not uniform, and there is a distribution of velocity, shear rate, and energy dissipation rate.

Therefore, one single calculated maximum shear or Komogorov scale are not enough to explain

the potential damages as well as variation in aggregate size within the bioreactor.

In this study, with the help of CFD modeling, not only distribution of shear rate, but also

distribution of energy dissipation rate and eddy scale can be obtained. Figure 3-11 shows that the

volume average eddy size from CFD simulation is actually similar to average aggregate size. This

finding helped explain how the aggregate size is possibly controlled in the bioreactor. Because of

Page 49: Computational Fluid Dynamics Modeling of Scalable Stirred

38

the forces exerted by turbulent eddies on the aggregates varies with size, the number of cells on an

aggregate being sheared off from cell aggregate-eddy interaction would also differ. Therefore,

resulted aggregates would have different sizes. Unlike cell damage which was difficult to predict

and quantify, cell aggregate size can be accurately measured using large sample size as in this

study. Additional CFD simulations at different agitation rates and corresponding biological

experiments can be done to reconfirm the similarity between cell aggregate distribution and eddy

size distribution in 100 ml spinner flask.

Figure 3–9: Cell aggregate distribution (D) at 80 rpm (A), 100 rpm (B), and 120 rpm (C)

Page 50: Computational Fluid Dynamics Modeling of Scalable Stirred

39

Figure 3–10: Shear rate and mean average size. A) Volume average shear rate; B) Maximum shear rate

40

80

120

160

200

240

280

400 450 500 550 600 650 700 750

Ave

rage

Agg

rega

te S

ize

(µm

)

Maximum Shear Rate (s-1)

A) Maximum Shear Rate and Mean Aggregate Size

40

90

140

190

240

290

12 13 14 15 16 17

Ave

rage

Agg

rega

te S

ize

(µm

)

Shear Rate (s-1)

B) Volume Average Shear Rate and Aggregate Size

Page 51: Computational Fluid Dynamics Modeling of Scalable Stirred

40

Figure 3–11: Relationship between eddy size and cell aggregate size. The average eddy size was calculated from volume-average energy dissipation rate obtained by CFD simulation, and the cell aggregate size was obtained from biological experiment.

3.4 Conclusions

In this study, the hydrodynamics of standard lab-scale 100ml bioreactor spinner flask was

characterized. The maximum shear stress levels obtained from CFD simulations were similar to

the results obtained from calculation described by Nagata (63). The impeller plays a significant

role on shear stress and energy dissipation rate distribution inside the bioreactors. Specifically,

closer to the tip of the impeller, the flow was more turbulent with higher velocity, shear rate, and

energy dissipation rate. Significant high shear and small eddy region only occupied a small volume

fraction within the bioreactor. Therefore, the maximum shear rate was about 40 times greater than

volume average shear rate. Near the impeller, the shear stress applied on the cells is larger than at

other region so that the cell membrane can get damaged and the aggregates can be broken up

0

50

100

150

200

0 0.001 0.002 0.003 0.004 0.005 0.006

Siz

e (µ

m)

Volume -averaged є (m/s3)

CFD Average Eddy Size and Average Aggregate Size at Different Turbulent Dissipation Rates

CFD Average Eddy Size Average Aggregate Size

Page 52: Computational Fluid Dynamics Modeling of Scalable Stirred

41

easier. Although the volume fraction of high shear region is low, there are still chances that the

cell and aggregates travel through this region. Therefore, there are always small aggregates in the

bioreactor despite of low agitation rates.

Cell aggregate size was measured by using an automatic particle sizer Multisizer III. The

aggregate size agreed with CFD results in many aspects. Larger increase in shear rate and energy

dissipation rate led to larger decrease in cell aggregate size. In addition, average aggregate size

was also similar to volume-averaged eddy size obtained from CFD simulation. The results

suggested that aggregate size distribution was possibly controlled by shear rate and eddy size

distribution inside the bioreactor. Increase in agitation rate with a consistent increment of 20 rpm

did not bring the same level increment in shear aggregate size. This is important information for

bioprocess development of aggregate forming cells like PSCs. For bioreactor scaling up, it would

be ideal to keep shear rate and energy dissipation rate consistent. The effect of changes certain

physical parameters such as impeller geometry, vessel dimensions, fluid viscosity, and agitation

rate on hydrodynamic environment also need to be carefully evaluated in order to design a good

bioprocess for PSC expansion. It is important to keep the cell aggregate size consistent since large

aggregate might increase mass transfer limitation and cell loss (34, 63).

Page 53: Computational Fluid Dynamics Modeling of Scalable Stirred

42

Chapter Four: Computational Fluid Dynamic Modeling of Scaled-Down 10 mL Stirred Suspension Bioreactors with Different Impeller Designs

4.1 Introduction

Stirred suspension bioreactors (SSBs) have been used widely for various purposes including

protein production, cell expansion, and cell differentiation. Recently, it has been reported that

induced pluripotent stem cells (PSCs) can also be derived in SSBs (15). There are many great

advantages that make SSBs an ideal choice for bioprocess development of mammalian cells in

general and stem cells in particular. These include creating well-mixing environment (7, 19),

reducing labor cost from conventional static culture (5), and having great scalability (9, 32, 46).

Before executing the main production plan in large scale bioreactors, extensive rounds of screening

and testing at smaller scale bioreactor needs to be done to select the optimal condition for cell

growth or protein production (68). The smallest scale that represents the bigger system would be

the most ideal choice for screening purposes.

Pluripotent stem cells have been commonly cultured in the laboratory using 100 ml spinner

flask bioreactors (5, 7, 69) . In many cases, 100 ml bioreactors are used for screening purposes to

determine optimal seeding density, feeding regime, effects of growth factors, effects of agitation

rates, and expansion of different cell lines (7, 61, 62). Each screening might need multiple

bioreactors in order to have a good sample size for final analysis. Using 100 ml bioreactor for

screening and factorial design experiments can be time consuming and costly due to large number

of cells required for seeding and amount of culture media needed for initial inoculation and

feeding. Therefore, scale-down version of 100 ml bioreactor is needed for high throughput

screening.

Page 54: Computational Fluid Dynamics Modeling of Scalable Stirred

43

Developing small scaled SSBs has been a new trend in bioprocess development (55). One

of the newest and most advanced small scale SSB system is the ambr from TAP Biosystem, UK.

This system uses 24 disposable bioreactors controlled by an automated work station. Each

bioreactor has 10 to 15 ml of working volume and its contents are stirred by an internal impeller.

Hydrodynamic environment of single ambr vessel was characterized using computational fluid

dynamics (CFD) modeling and mixing time measurement (13). Although a large part of

hydrodynamics in ambr system was revealed in this study, some aspects of bioreactor scaling-

down has not been clear. For instance, the effects of rectangular shape of the ambr vessel was not

investigated. The similarity between ambr and 5 L bioreactor was just based on the tip speed and

dissipation rate with no information on shear stress. The cells chosen for this study were CHO

cells which respond to shear stress differently than stem cells. Therefore, ambr system might not

be an ideal choice for growing shear sensitive cells such as PSCs until the effects of hydrodynamic

environment in amber system on the cells have been fully characterized.

Here, we use 10 ml bioreactors (HexaScreen, Barcelona, Spain) for PSC culture. These

bioreactors closely resemble the standard lab-scale 100 ml spinner flasks with cylindrical vessels

and top mounted rotating system. Unlike the ambr vessels, these 10 ml HexaScreen bioreactors

are reusable that allows the researchers to run multiple experiments with relatively low cost. Each

small vessel can be treated as a single T-25 flask or 10 cm2 petri dish for daily sampling and

multipurpose screening without concerning about contamination or hydrodynamic changes as in

100 ml bioreactor way of sampling.

In this study, we used computational fluid dynamic (CFD) to model the hydrodynamic

environment of 10 ml bioreactor to test whether it can be used as a scaled-down model of standard

100 ml spinner flask. Besides, biological experiments with pluripotent stem cells were conducted

Page 55: Computational Fluid Dynamics Modeling of Scalable Stirred

44

to verify that the hydrodynamic environment in 10 ml bioreactor is suitable for shear-sensitive and

aggregate forming cell culture.

4.2 Materials and Methods

4.2.1 Cell Culture

The D3 cell line of murine ESCs (provided by Dr. Rancourt, University of Calgary) were

grown on murine embryonic fibroblasts for one passage upon thawing and then were routinely

passaged twice into gelatin-coated petri dish (BD BioSciences, Bedford, MA, USA). The

expansion medium for murine ESCs consisted of 15% fetal bovine serum (Gibco, Grand Island,

NY), 1% non-essential amino acids (Gibco), 1% penicillin and streptomycin (Gibco), 0.1 mM β-

Mercaptoethanol (Sigma, St.Louis, MO), high glucose DMEM (Gibco), and 1,000U/ml leukemia

inhibitory factor (EMD Millipore, Billerica, MA, USA). In suspension cultures, murine ESCs were

inoculated as single cells into duplicate 10 ml spinner flasks (HexaScreen, Barcelona, Spain) at

density of 33, 000 cells/ml. BioWiggler (Global Cell Solutions, Charlottesville, VA, USA) stir

plate with 8 programmed stirring positions was used to agitate 10 ml bioreactors. The accuracy

and precision of bioreactor agitation rates at 100, 120, and 150 rpm were tested by using a

Touchless Digital Tachometer (VWR). All cell cultures were maintained at 370 and 5% CO2.

Duplicate of the whole 10 ml bioreactor was taken for daily examining of cell aggregate size,

growth, and metabolites. Cell aggregate size was measured based on photographs taken on Zeiss

microscope and built in size measurement tool (AnxioVision 4). Cell count and cell viability

assessment were performed using standard hemocytometer, and cell metabolites were measured

using a Bio-profile 100 Plus (NOVA Biomedical, Waltham, MA, USA).

Page 56: Computational Fluid Dynamics Modeling of Scalable Stirred

45

4.2.2 Computational Fluid Dynamics Modeling

4.2.2.1 Geometry

In this study, 10 ml bioreactors with two different impeller design were modelled. The

vessel diameter for both models stayed the same. One model has simple cylinder impeller (Figure

4–1A) and another one with overall paddle-shape impeller (Figure 4–1B). Similar to 100 ml

bioreactors, 10 ml bioreactors also had a small indentation on the center bottom of the vessel to

reduce the dead space. The dimensions of the 10 ml bioreactor models were listed on Table 4–1.

Figure 4–1: Model geometry built in COMSOL Multiphysics. A) 3D view of fluid occupied region in 10 ml cylindrical impeller bioreactor; B) 3D view of fluid occupied region in 10 ml paddle impeller bioreactor

Page 57: Computational Fluid Dynamics Modeling of Scalable Stirred

46

Table 4–1: Dimensions of 10 ml bioreactor models

Dimension Cylinder Impeller Paddle Impeller

Vessel inner diameter 0.032 (m) 0.032 (m)

Fluid height 0.0145(m) 0.015 (m)

Impeller diameter (horizontal length) 0.0254 (m) 0.0274 (m)

Impeller width (vertical length) 0.0054 (m) 0.0074 (m)

Indentation height from the bottom vessel 0.005 (m) 0.005 (m)

4.2.2.2 Model Physics

The flow in stirred vessel is considered to be fully laminar when the Reynolds number is

less than 10 and fully turbulent number is greater than 20,000 (13). The transition region is where

the Reynolds number falls between this range. Transition flow is difficult to characterize and there

has been yet a commercially available CFD program that can fully solve this flow behavior. In this

study, Reynolds number from 1,031 to 1,858 which is closer to the turbulent regime. Therefore, in

the CFD model, Reynolds Averaged Navier-Stokes (RANS) k-ε technique was chosen to model

the flow in 10 ml bioreactor at 100, 120, and 150 rpm. The k-є model is very popular for industrial

applications due to its good convergence rate and relatively low memory requirements compared

to other turbulent models (18). All the solid walls are presented by non-slip conditions and the free

surface is specified with an open boundary (atmospheric pressure) condition; the effects of

capillary effects are not taken into account. The model assumed that the fluid properties are as

same as water.

Page 58: Computational Fluid Dynamics Modeling of Scalable Stirred

47

The commercial finite element Multiphysics software package COMSOL Multiphysics

Version 4.4 (COMSOL, Inc. California, USA) was used to model the hydrodynamics of 10 ml

bioreactor. The simulations were run on Intel dual core Xeon 3.30 GHz with 24 GB RAM. All

cores were under load during each simulation and about 8 Gb out of 24 Gb of memory was

allocated for COMSOL Multiphysics.

The finite element grid was created using physics-controlled mesh generation option in

COMSOL. The grid consisted of 534,346 finite elements which was comparable with the fine grid

generated for 100 ml bioreactor in previous study. Each agitation rate was simulated for 12s

duration and 826,993 degree of freedom was solved with average CPU run time of 70 hours. Data

from the simulations were then exported into Excel spreadsheet for further analysis. Flow

visualizations were performed by using built-in port-processing tools within COMSOL

Multiphysics package.

4.3 Results and Discussions

4.3.1 Base Case: Hydrodynamics of 10 ml Bioreactor at 100 rpm

Figure 4–2 shows the distributions of the pressure, velocity, shear stress, and energy

dissipation rate at 100 rpm. Similar to 100 ml model (refer to Chapter 3), the pressure in 10 ml

model is basically hydrostatic pressure that has no significant effect on the cells at resulted level

(67). The velocity is highest at the tip of the impeller, toward the wake region. Due to non-slip

condition, the velocity magnitude increases in region that is closer to the rotating impeller. Shear

rate and turbulent dissipation rate are also highest near the tip of the impeller. The region with

relatively high shear rate (greater than 100 s-1) and high turbulent dissipation rate (greater than

0.03 m2/s-3) only occupies a small volume fraction of the entire bioreactor.

Page 59: Computational Fluid Dynamics Modeling of Scalable Stirred

48

Figure 4–2: Basic hydrodynamics of 10 ml cylindrical impeller bioreactor at 100 rpm

Page 60: Computational Fluid Dynamics Modeling of Scalable Stirred

49

4.3.2 Comparison of 100 ml and 10 ml Bioreactor Models at 100 rpm

As previously discussed in 100 ml models (refer to Chapter 3), higher volume averaged

shear rates and energy dissipation rates led to smaller aggregate sizes. Figure 4–3 shows that shear

rate distribution in 100 ml and 10 ml bioreactors at 100 rpm are similar. Cumulative volume

fraction is the sum of volume fractions for all shear rate levels up to the current one, and it equals

to 1 for the entire graph from 0 s-1 to 150 s-1. Al though the maximum shear rate is much higher

than 150 s-1, the volume fraction for shear stress level higher than 150 s-1 is very small so that the

graphs no longer increases vertically. There is steep increase from lowest shear rate of about 5 s-1

up to shear rate of 60 s-1, indicating that shear rate of this range occupies the most volume fraction

of both 10 ml and 100 ml bioreactors. The line graph representing10 ml bioreactor is above the

line graph representing 100 ml bioreactor for shear rate under 60 s-1, but the order is reversed for

shear rate above 60s-1. Therefore, the overall volume- averaged mean shear rate in 100 ml

bioreactor is higher than in 10 ml bioreactor. In fact, the different in volume- averaged shear rate

of the two bioreactor scales is about 14% (Table 4–2).

Figure 4–4 shows the comparison of turbulent dissipation rate between two bioreactor

scales. The volume fraction in 10 ml bioreactor stops increasing after its turbulent dissipation rate

reaches 0.01m2/s3. However, the volume fraction of higher dissipation rate (above 0.01 m2/s3) is

still continuing to increase. In fact, the volume averaged turbulent dissipation rate in 100 ml model

is significantly higher than in 10 ml model (about 158% higher, Table 4–2). In addition, the volume

average velocity in 100 ml bioreactor is also higher than in 10 ml bioreactor (Table 4–2). In order

to match up the shear rate and energy dissipation profiles between the two bioreactor scales,

agitation rate increase in 10 ml bioreactor is needed. The level of increase is predicted by CFD

simulation and verified by biological experiments.

Page 61: Computational Fluid Dynamics Modeling of Scalable Stirred

50

Figure 4–3: Comparison of shear rate distribution in 100 ml bioreactor and 10 ml bioreactor at 100 rpm. The data was extracted from quasi-steady sate state (after 3s) of agitation. The area under the curve was normalized to unity.

0

0.01

0.02

0.03

0.04

0.05

0.06

0.07

0.08

0 20 40 60 80 100 120 140 160

Cu

mu

lati

ve V

olu

me

Fra

ctio

n

Shear Rate (1/s)

Shear Rate Distribution in 10 ml and 100 ml Bioreactors at 100 RPM

10 ml-100 rpm 100 ml-100 rpm

Page 62: Computational Fluid Dynamics Modeling of Scalable Stirred

51

Figure 4–4: Comparison of turbulent dissipation distribution in 100 ml and 10 ml bioreactors at 100 rpm. The volume fraction of shear rate above 150 s-1 is insignificant so that the lines representing the cumulative volume of 10 ml and 100 ml bioreactor at 100 rpm no longer increase on the y-axis. The area under the curve was normalized to unity.

0

0.0005

0.001

0.0015

0.002

0.0025

0.003

0 0.005 0.01 0.015 0.02

Cu

mu

lati

ve V

olu

me

Fra

ctio

n

Turbulent Dissipation Rate (m2/s3)

Turbulent Dissipation Rate in 10 ml and 100 ml Bioreactors at 100 RPM

10 ml-100 rpm 100 ml-100 rpm

Page 63: Computational Fluid Dynamics Modeling of Scalable Stirred

52

4.3.3 Effects of Agitation Rates

4.3.3.1 Hydrodynamic Effects

At lower range of shear rate (less than 50 s-1), the volume fraction for this range is highest

at 100 rpm and lowest at 150 rpm (Figure 4–5). The order is reversed for shear rate greater than

50 s-1. Overall, it shows that the shear rate is higher at higher agitation rate and greater increase in

agitation rate leads to greater increase in shear rate. For instance, 20% increase in agitation rate

from 100 rpm to 120 rpm gives 21% increase in volume averaged shear rate, and 25% increase in

agitation rate from 120 rpm to 150 rpm gives 25.5% increase in volume averaged shear rate in 10

ml bioreactor (Table 4–2). These results are consistent with previous 100 ml model.

The turbulent dissipation rate in Figure 4–6 shows similar trend to Figure 4-5. Overall, the

turbulent dissipation rate is highest at 150 rpm and lowest at 100 rpm in 10 ml bioreactor. However,

the level of difference in turbulent dissipation rate is more dramatic than in shear rate. Specifically,

20% increase in agitation rate leads to 56% increase in volume averaged turbulent dissipation rate,

and 25% increase in agitation rate leads to 66% increase in volume averaged turbulent dissipation

rate. Interestingly, volume averaged turbulent dissipation rate (0.003 m2/s3) in 10 ml bioreactor at

150 rpm matches with 100 ml bioreactor at 100 rpm.

Table 4–2: Comparisons of volume-averaged velocity, shear rate, and turbulent dissipation rate at different agitation rates in 10 ml bioreactor with cylinder impeller

100 rpm 120 rpm 150 rpm Velocity (m/s) 0.028 0.033 0.040 Shear Rate (s-1) 12.34 14.95 18.76 Energy Dissipation Rate (m2/s3) 1.16E-3 1.81E-3 3.0E-3

Page 64: Computational Fluid Dynamics Modeling of Scalable Stirred

53

Figure 4–5: Comparison of shear rate distribution between different agitation rates in 10 ml bioreactor. The volume fraction of shear rate above 150 s-1 is insignificant so that the lines representing the cumulative volume of different agitation rates no longer increase on the y-axis. The area under the curve was normalized to unity.

0

0.01

0.02

0.03

0.04

0.05

0.06

0.07

0.08

0 30 60 90 120 150

Cu

mu

lati

ve V

olu

me

Fra

ctio

n

Shear Rate (s-1)

Shear Rate at Different Agitation Rates

100 rpm 120 rpm 150 rpm

Page 65: Computational Fluid Dynamics Modeling of Scalable Stirred

54

Figure 4–6: Comparison of turbulent dissipation between different shear rates in 10 ml bioreactor. The volume fraction of turbulent dissipation at rates higher than 0.02 m2/s3 is insignificant so that the lines representing the cumulative volume of different agitation rates no longer increase on the y-axis. The area under the curve was normalized to unity.

4.3.3.2 Biological Effects

As mentioned previously, the volume averaged turbulent dissipation rate in 10 ml at 150

rpm matches with the optimal value in 100 ml bioreactor at 100 rpm. Therefore, it is expected that

150 rpm in 10 ml bioreactor would give the best cell expansion out of three tested agitation rates.

Indeed, the results show that the 10 ml bioreactor culture achieves highest cell expansion at 150

0

0.0005

0.001

0.0015

0.002

0.0025

0.003

0 0.005 0.01 0.015 0.02

Cu

mu

lati

ve V

olu

me

Fra

ctio

n

Dissipation Rate (m2/s3)

Turbulent Dissipation Rate at Different Agitation Rates

100 rpm 120 rpm 150 rpm

Page 66: Computational Fluid Dynamics Modeling of Scalable Stirred

55

rpm (Figure 4-7A). Due to lower shear rate and low energy dissipation rate, the aggregate size in

bioreactors agitated at 100 rpm are large (Figure 4-7B) leading to possible mass transfer limitation

that affects cell proliferation (2, 23, 70). Therefore, the cells expansion at 100 rpm in 10 ml

bioreactor is very low and seems to have decreased after the third day. The cell aggregate size at

150 rpm in 10 ml bioreactor is smallest among tested agitation rates and is similar to the aggregate

size at 100 rpm in 100 ml bioreactor. The viability of the cells in 10 ml bioreactors operated at 150

rpm remained above 95% even at day 4. The viability at other agitation rates was slightly lower

but was still higher than 90%. Based on hydrodynamic and biological response similarities, 10 ml

bioreactor at 150 rpm can be considered to be the scaled-down model of 100 ml bioreactor at 100

rpm.

The media properties of murine ESCs were examined by measuring the pH, nutrient

concentration (glucose), and waste products (lactate and ammonia) daily for both experiments

throughout the culture period. The glucose concentration in the media for the mESCs that were

expanded at different agitation rates in the stirred suspension and in the static cultures decreased

steadily over the course of the culture periods (Figure 4-7D). However, the glucose concentration

after four days remained above zero for all conditions, indicating that nutrient concentration is not

limiting cell proliferation after four days.

The pH of the media dropped over the culture period for all conditions (Figure 4-7C). By

the fourth day of culture, the pH of the static condition was significantly lower than all bioreactor

conditions. Although the final cell multiplication ratio was greater in the 150 rpm mini bioreactor

compared to static culture, it is possible that the pH in the static condition is lower due to a

difference in growth conditions of the suspended spheres and adherent monolayers. There is a

limited growth surface area in the static culture compared to the stirred suspension bioreactors

Page 67: Computational Fluid Dynamics Modeling of Scalable Stirred

56

causing increased cell death and release of lactic acid into the culture media. Low dissolved oxygen

concentration and the absence of agitation could attribute to the decrease in expansion of mESC

cells in static condition. It has been reported that a pH below 7.1 reduces cell proliferation in stem

cells (32). Figure 4-7G shows that pH remains in healthy range for all bioreactor conditions, but

high acidic for static condition as a result of high lactic acid concentration. Although ammonia

level increased (Figure 4-7F), the concentration of ammonia on day 4 for all culture conditions

was below 1.5 mml/L which is not considered to be inhibitory to mammalian cell growth (7). Since

the difference in metabolites is small and the concentrations of glucose, lactate, and ammonia are

within the safe limit, the differences in cell expansion and aggregate size between three agitation

rates are mainly due to hydrodynamic environment.

0

5

10

15

20

25

30

35

40

45

0 1 2 3 4

Mu

ltip

lica

tion

Rat

io

Time (days)

A) Multiplication Ratios at Different Agitation Rates

100rpm 120rpm 150rpm Static

Page 68: Computational Fluid Dynamics Modeling of Scalable Stirred

57

0

50

100

150

200

250

300

350

400

100 120 150

Agg

rega

te D

iam

eter

m)

Agitation Rate (rpm)

B) Mean Aggregate Diameters at Different Agitation Rates

6.5

6.7

6.9

7.1

7.3

7.5

7.7

7.9

0 1 2 3 4

pH

Time (days)

C) Change in pH of Culture Media

100rpm 120rpm 150rpm Static

Page 69: Computational Fluid Dynamics Modeling of Scalable Stirred

58

0.0

1.0

2.0

3.0

4.0

1 2 3 4

g/L

Time (days)

D) Glucose Concentration in the Media

100rpm 120rpm 150rpm Static

0.0

1.0

2.0

3.0

4.0

1 2 3 4

g/L

Time (days)

E) Lactate Concentration in the Media

100rpm 120rpm 150rpm Static

Page 70: Computational Fluid Dynamics Modeling of Scalable Stirred

59

Figure 4–7: Comparison of biological responses in cylindrical impeller 10 ml bioreactors at different agitation rates and in static culture.

4.3.4 Effects of Impeller Geometry

4.3.4.1 Hydrodynamic Effects

Figure 4–8 shows the velocity field of 10 ml bioreactor with paddle impeller and 10 ml

bioreactor with cylinder impeller at 150 rpm. The impellers rotate counter clockwise as shown by

the directions of small arrows. Near the bottom of the bioreactor vessel, the velocities are more

uniform with regular axial direction. Closer to the impeller at the middle of the bioreactor height,

the velocity increases and radial flow is very strong. At this region, the flow is highly turbulent

with strong flow separation shown by the arrows. Visually, the radial flow is stronger and velocity

magnitude is higher in the bioreactor with paddle impeller. In fact, the volume averaged velocity

0

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

1 2 3 4

mm

ol/L

Time (days)

F) Ammonia Concentration in the Media

100rpm 120rpm 150rpm Static

Page 71: Computational Fluid Dynamics Modeling of Scalable Stirred

60

in the paddle impeller model is about 85% greater than in cylinder impeller model at the same

agitation rate (Table 4–3).

Figure 4–9 shows that shear rate in paddle impeller model is higher than in cylinder

impeller. Due to the rectangular shape with relatively small corners, shear rate near the tip of

paddle impeller is much higher than maximum shear rate in cylinder impeller model. In the entire

bioreactor volume, the average shear rate in paddle impeller model is 74% greater than in cylinder

impeller model (Table 4–3). The impeller geometry also strongly affects the turbulent dissipation

rate in 10 ml bioreactors (Figure 4–10). The volume average turbulent dissipation rate in paddle

impeller model is about 3.2 times greater than in cylinder impeller model (Table 4–3). Hence, with

higher shear rate and turbulent dissipation rate, the 10 ml bioreactor with paddle impeller is

expected to produce smaller cell aggregates.

Table 4–3: Comparison of volume-averaged velocity, shear rate, and turbulent dissipation rate between bioreactor with paddle impeller and bioreactor with cylinder impeller at 150 rpm

Cylinder Impeller (150 rpm) Paddle Impeller (150 rpm)

Velocity (m/s) 0.040 0.075

Shear Rate (s-1) 18.76 32.8

Energy Dissipation Rate (m2/s3) 3.0E-3 9.7E-3

Page 72: Computational Fluid Dynamics Modeling of Scalable Stirred

61

Figure 4–8: Velocity vector field in 10 ml bioreactor with cylinder impeller and in 10 ml bioreactor with paddle impeller at 150 rpm. Longer and thicker arrows indicate velocity vectors with high magnitude, and the arrows point toward the direction of the velocity vectors.

Paddle Impeller

Cylinder Impeller

Page 73: Computational Fluid Dynamics Modeling of Scalable Stirred

62

Figure 4–9: Comparison of shear rate distribution between the 10 ml bioreactor with cylinder impeller and 10 ml bioreactor with paddle impeller at 150 rpm. The area under the curve was normalized to unity.

0

0.01

0.02

0.03

0.04

0.05

0.06

0.07

0.08

0.09

0 30 60 90 120 150

Cu

mu

lati

ve V

olu

me

Fra

ctio

n

Shear Rate (s-1)

Effects of Impeller Geometry on Shear Rate

Cylinder Impeller (150 rpm) Paddle Impeller (150 rpm)

Page 74: Computational Fluid Dynamics Modeling of Scalable Stirred

63

Figure 4–10: Comparison of turbulent energy dissipation rate between 10 ml bioreactor with cylinder impeller and 10 ml bioreactor with paddle impeller at 150 rpm. The area under the curve was normalized to unity.

0

0.0005

0.001

0.0015

0.002

0.0025

0.003

0.0035

0 0.005 0.01 0.015 0.02

Cu

mu

lati

ve V

olu

me

Fra

ctio

n

Turbulent Dissipation Rate (m2/s3)

Effects of Impeller Geometry on Turbulent Dissipation Rate

Cylinder Impeller (150rpm) Paddle Impeller (150rpm)

Page 75: Computational Fluid Dynamics Modeling of Scalable Stirred

64

4.3.4.2 Biological Effects

Cell growth in 10 ml bioreactor with paddle impeller and in 10 ml bioreactor with cylinder

impeller at 150 rpm are shown on Figure 4-11A. The cell density in bioreactor with cylinder

impeller is higher than in bioreactor with paddle impeller. However, the difference is less

significant compared to the cell growth at lower agitation rates. Changes in pH (Figure 4-11B),

glucose concentration (Figure 4-11E), lactate concentration (Figure 4-11F), and ammonia

concentration (Figure 4-11G) are also similar for both bioreactor systems with different impeller

geometries. As expected, the cell aggregate size in 10 ml bioreactor with paddle impeller (Figure

4-11C) is smaller than the aggregate size in 10 ml bioreactor with cylinder impeller (Figure 4-11D)

at the same agitation rate. Due to stronger shear rate and turbulent dissipation rate, there are more

small aggregates in bioreactor with paddle impeller giving the mean aggregate diameter of 114 μm

which is smaller than aggregate dimeter in 10 ml bioreactor with cylinder impeller (150 μm).

Relatively smaller aggregate below 250 μm are not affected by mass transfer limitation and are

easier to split into single cells than bigger aggregates. Therefore, resulted cell density in bioreactor

at 150 rpm with both impeller geometries was much higher than the viable cell density produced

at lower agitation rates. Furthermore, the viability of both culture conditions were always above

95% in 4 days of culture duration. Although high shear rate and energy dissipation rate might cause

some damages to the cells, these hydrodynamic ingredients are necessary to maintain the healthy

cell aggregate size as well as cell growth in suspension culture.

Page 76: Computational Fluid Dynamics Modeling of Scalable Stirred

65

0.0E+00

2.0E+05

4.0E+05

6.0E+05

8.0E+05

1.0E+06

1.2E+06

1.4E+06

1 2 3 4

Via

ble

Cel

l Con

cen

trat

ion

(ce

lls/

mL

)

Time (days)

A) Cell growth

150 rpm, Cylinder 150rpm, Paddle

Page 77: Computational Fluid Dynamics Modeling of Scalable Stirred

66

6.0

6.5

7.0

7.5

8.0

1 2 3 4

pH

Time (days)

B) Change in pH of Media

150 rpm, Cylinder 150 rpm, Paddle

0.E+00

5.E-02

1.E-01

2.E-01

2.E-01

60 80 100 120 140 160 180 200 210 220 240

Tot

al N

um

ber

Fra

ctio

n

Aggregate Diameter (µm)

C) Cylinder Impeller at 150 RPMAggregate Size Distribution

Page 78: Computational Fluid Dynamics Modeling of Scalable Stirred

67

0.E+00

5.E-02

1.E-01

2.E-01

2.E-01

3.E-01

60 80 100 120 140 160 180 200 220

Tot

al N

um

ber

Fra

ctio

n

Aggregate Diameter (μm)

D) Paddle Impeller at 150 RPMAggregate Size Distribution

0.0

1.0

2.0

3.0

4.0

1 2 3 4

g/L

Time (days)

E) Glucose Concentration in Media

150 rpm, Cylinder 150 rpm, Paddle

Page 79: Computational Fluid Dynamics Modeling of Scalable Stirred

68

Figure 4–11: Comparison of biological responses in 10 bioreactors with cylinder impeller and in 10 ml bioreactor with paddle impeller at 150 rpm.

0.0

1.0

2.0

3.0

4.0

1 2 3 4

g/L

Time (days)

F) Lactate Concentration in Media

150 rpm, Cylinder 150 rpm, Paddle

0

0.5

1

1.5

2

1 2 3 4

mm

ol/L

Time (days)

G) Ammonia Concentration in Media

150 rpm, Cylinder 150 rpm, Paddle

Page 80: Computational Fluid Dynamics Modeling of Scalable Stirred

69

4.3.5 Effects of Shear Rate and Eddy Size on Aggregate Size

Figure 4–12 shows the relationship between shear rate and cell aggregate size in 10 ml

bioreactor. The results are very similar to 100 ml model described in previous chapter. Higher

shear leads to smaller aggregate, and the level of difference in aggregate size is dependent on the

level of difference in shear rate. For instance, 25% increase in shear rate from 100 rpm to 120 rpm

leads to 13% decrease in aggregate size, and 74% increase in shear rate from 120 rpm to 150 rpm

leads to 31% decrease in aggregate size. At low shear level as in 10 bioreactors being agitated at

100 rpm, majority of the cell aggregates stay in low shear region so that the overall aggregate size

is very large. However, despite of the agitation rates, aggregate breakup near the tip of rotating

impeller is always possible. Therefore, there are always small aggregates present in stirred

suspension culture, and higher shear would produce higher number of small aggregates.

Figure 4–13 shows the similarity of average eddy size and average cell aggregate size in

10 ml bioreactor at different turbulent dissipation rates. As turbulent dissipation rate increases, the

eddy size decreases (71). Results from previous 100 ml model in previous chapter showed that

average cell aggregate have similar size as average eddy size at different agitation rates. In fact,

the average aggregate size and average eddy size in 10 ml bioreactor are also similar. The

difference is only 15 µm or less which is about size of a single murine ESC. Therefore, once again,

the results show that CFD modeling can be used to predict hydrodynamics and cell aggregate size

in stirred suspension bioreactor.

Page 81: Computational Fluid Dynamics Modeling of Scalable Stirred

70

Figure 4–12: Effects of shear rate on murine ESC aggregate size in 10 ml bioreactor with cylinder impeller. The error bars are standard deviations in aggregate size.

0

50

100

150

200

250

0 5 10 15 20 25 30 35

Agg

rega

te S

ize

(µm

)

Volume Averaged Shear Rate (s-1)

Effects of Shear Rate on Aggregate Size in 10 ml Bioreactors

Page 82: Computational Fluid Dynamics Modeling of Scalable Stirred

71

Figure 4–13: Similarity in average aggregate size and eddy size at different turbulent dissipation rates in 10 ml bioreactor.

4.4 Conclusions

Hydrodynamics of 10 ml bioreactor at different agitation and with different impeller

geometries were modeled using COMSOL Multiphysics. Velocity, pressure, shear rate, and energy

dissipation rate profiles in 10 ml bioreactor were similar to 100 ml bioreactor. However, the

magnitude of these hydrodynamic values were smaller in 10 ml bioreactor than in 100 ml

bioreactor at the same agitation rate. This can be adjusted by increase the agitation rate in 10 ml

bioreactor. It was found in this study that, 150 rpm in 10 ml bioreactor has similar hydrodynamics

and biological responses as 100 rpm in 100 ml bioreactor. Therefore, 10 ml bioreactor can be used

as a scaled down model of 100 ml bioreactor with a predefined adjustment in impeller speed. CFD

0

20

40

60

80

100

120

140

160

180

0 0.002 0.004 0.006 0.008 0.01 0.012

Siz

e (µ

m)

Turbulent Dissipation Rate (m2/s3)

Aggregate Size and Eddy Size at Different Turbulent Dissipation Rates

Mean Eddy Size Mean Aggregate Size

Page 83: Computational Fluid Dynamics Modeling of Scalable Stirred

72

modeling is a great tool to help determine the impeller speed and predict the hydrodynamic change

in scaling down bioprocess.

In this study, over 4-day culture period, the cell aggregate size and cell expansion were found

to be highly dependent on hydrodynamic environment. The metabolites of the cells in 10 ml

bioreactors were found to be in a healthy range. Extended culture duration can be done in the future

to further investigate the effects of metabolites on the cells.

The effects of shear rate and eddy size on cell aggregate size was reconfirmed in this study.

Similar to 100 ml bioreactor, the level of effects was dependent on the level of changes in

hydrodynamics, mainly shear rate and energy dissipation rate. Once again, it was shown that the

average aggregate sizes are similar to average eddy sizes at corresponding turbulent dissipation

rates. Therefore, CFD can be used to not only predict the hydrodynamics, but also the performance

of the bioreactor on producing the optimal cell expansion and cell aggregate size.

Page 84: Computational Fluid Dynamics Modeling of Scalable Stirred

73

Chapter Five: Discussions of Limitations, Recommendations, and Conclusions

5.1 Discussions of Limitations and Recommendations

5.1.1 CFD Modeling

Although CFD simulation used in this project has captured many important information on

hydrodynamic environment inside the bioreactors with various configurations, there are several

limitations that need further investigations. In order to accurately apply k-є model, the flow has to

be fully turbulent. However, based on the calculated Reynolds number, the flow is actually in the

transition zone which is not yet fully turbulent (13). The turbulent flow assumption is needed not

only in this studies but other similar studies because the transition flow is difficult to characterize

and current CFD packages are not fully capable of modeling this type of flow (18). The k-є has

shown good agreement with experimental results for small pressure gradient, but less accurate for

adverse pressure gradients (52). Moving forward from this project, experimental validation of CFD

simulation can be done using PIV (72). The velocity field obtained from PIV can be used to derive

shear stress and energy dissipation rate as in CFD modeling.

The fluid used in all simulations was assumed to be as same as water. Although this

assumption is widely used in many CFD modeling studies (13, 18, 19, 73), the exact fluid property,

particularly cell culture medium with presence of cells, is needed in order to obtain the accurate

model outputs. The viscosity of fluid can be measured experimentally using rheometer. However,

depending on protein and cell concentration, the fluid mixture can behave differently (74).

Therefore, measuring the viscosity of cell culture medium at various shear rate and cell density is

necessary in order to accurately determine fluid properties of cell culture medium. This can be a

constant number for Newtonian fluid or a function of shear rate for non-Newtonian fluid (43, 75).

Page 85: Computational Fluid Dynamics Modeling of Scalable Stirred

74

The grid generations were performed without taking into account the size of the aggregates.

The model assumed one-phase turbulent flow and the cell aggregate size and mass were neglected

to simplify the numerical complexity and to save computational cost. However, as the exact level

of forces acting on the cells of the aggregates is needed for other studies such as

mechanotransduction study, the size and the mass of the aggregates should be incorporated into

the CFD model and the mesh size should be adaptive to the size of the aggregates. In this case,

two-phase flow can be employed to model the fluid and solid cell aggregate interaction which may

involve more complex numerical solutions and higher computation cost (76).

5.1.2 Biological Experiments

The main focus of this project is to model the fluid environment and correlate the

simulation data with cell expansion and aggregate size. The next step would be staining for

pluripotency markers and testing differentiation capability of resulted cell aggregates to clarify

the different effects of hydrodynamics on cell behaviors.

Although the results show good agreements between 100 ml and 10 ml bioreactors on the

relationship between small eddy size and aggregate size, additional data at different agitation

rates are needed to further support this finding. In addition, it is worth testing similar conditions

using different aggregate-forming cells such as human ESCs and iPSCs.

5.2 Conclusions

CFD simulation is a powerful tool for modeling the hydrodynamics of SSBs. Using CFD

simulation, important hydrodynamic variables such as fluid velocity, shear rate, and energy

dissipation rates can be obtain locally and globally. These are difficult to get experimentally, even

Page 86: Computational Fluid Dynamics Modeling of Scalable Stirred

75

with the most used velocity imaging technique like PIV (77). In Chapter 3, standard lab-scale 100

ml spinner flask was simulated using RANS k-ε model in COMSOL Multiphysics. As expected

the shear rate and energy dissipation rate were highest near the tip of the impeller. The significant

difference in volume-averaged and maximum values of shear rate (~40 times) and energy

dissipation rates (~100 times) suggesting that the “possible cell damage” region only occupied a

small volume fraction of the bioreactor. However, more or less, the cell and aggregates could travel

through this region and experience high shear. Even though the aggregate size increased as shear

rate end energy dissipation rate decreased, small aggregate sizes always presented in all stirring

conditions.

In Chapter 4, the same method was used to model the hydrodynamics of 10 ml bioreactor.

The velocity, shear rate, and energy dissipation rate profiles were similar to 100 ml model in

Chapter 3. The magnitude of the two bioreactor scales could be matched up by increasing the

agitation rate in 10 ml bioreactor. It was shown in Chapter 4 that, 150 rpm in 10 ml bioreactors

gave similar volume-averaged turbulent dissipation rate as 100 rpm in 100 ml bioreactor described

in Chapter 4. Therefore, based on hydrodynamic profile, 10 ml bioreactor at 150 rpm could be

considered as a scaled-down model of 100 ml bioreactor at 100 rpm.

In addition to effects of scaling-down, the effects of impeller geometry were also described

in Chapter 4. The 10 ml bioreactor with paddle impeller produced higher shear than 10 ml

bioreactor with cylinder impeller at the same agitation rate. The paddle impeller provided more

sweeping area than cylinder impeller due to 7.8% difference in length and width. The relatively

thin and small corner led to high local shear rate and high energy dissipation rate near the tip of

the impeller. Overall volume-average shear rate in 10 ml bioreactor with paddle impeller was 75%

higher than the 10 ml bioreactor with cylinder impeller, at 150 rpm.

Page 87: Computational Fluid Dynamics Modeling of Scalable Stirred

76

Biological results in Chapters 3 and 4 agreed with the expectations based on hydrodynamic

profiles. In higher shear environment, the cells aggregates were less likely to be broken up so that

the mean aggregate size was higher at lower shear rate. The same observations in aggregate size

were obtained in 10 ml bioreactors. Furthermore, biological results also showed that the average

aggregate size obtained from 10 ml bioreactor at 150 rpm (150 μm) was similar to the average

aggregate size in 100 ml bioreactor at 100 rpm (135 μm). The difference was very small and is

about the size of a single ES cell (78).Therefore, high through put screenings and experimental

testing could be done using 10 ml scaled-down bioreactors instead of using 100 ml bioreactors.

Another useful of information for cell aggregate control and bioreactor scaling was the

relationship between average eddy size and average aggregate diameter. The average small eddy

sizes in Chapters 3 and 4 were calculated from volume-averaged turbulent dissipation rate (12).

The results in 100 ml model (chapter 3) and 10 ml model (chapter 4) showed that the average cell

aggregate size was similar to the average eddy size obtained from CFD modeling. Interestingly,

this result also imply that the volume-averaged turbulent dissipation rate can be considered as a

scale-up and scale-down criterion in bioreactor design for expanding aggregate-forming cells such

as PSCs. Using turbulent energy dissipation rate as a scaling criterion is in fact very popular in

bioprocess development (12, 79). However, the volume-averaged energy dissipation rate here can

only be obtained from CFD simulation.

Overall, CFD simulation results and biological results showed good agreements. The results

from this work can be applied into bioprocess design and development of aggregate-forming cells.

Furthermore, the information on hydrodynamics of 100 ml and 10 ml bioreactors can be used in

mechanotransduction studies to investigate the effects of shear stress on gene expressions of the

cells.

Page 88: Computational Fluid Dynamics Modeling of Scalable Stirred

77

References

1. Walia B, Satija N, Tripathi RP, Gangenahalli GU. 2012. Induced pluripotent stem cells:

fundamentals and applications of the reprogramming process and its ramifications on

regenerative medicine. Stem Cell Rev. Reports. 8(1):100–115

2. Fridley KM, Kinney M a, McDevitt TC. 2012. Hydrodynamic modulation of pluripotent

stem cells. Stem Cell Res. Ther. 3(6):45

3. Loring JF, McDevitt TC, Palecek SP, Schaffer D V, Zandstra PW, Nerem RM. 2014. A

global assessment of stem cell engineering. Tissue Eng. Part A. 00(00):1–15

4. Nienow AW, Rafiq Q a., Coopman K, Hewitt CJ. 2014. A potentially scalable method for

the harvesting of hmscs from microcarriers. Biochem. Eng. J. 85:79–88

5. Kehoe DE, Jing D, Lock LT, Tzanakakis ES. 2010. Scalable stirred-suspension bioreactor

culture of human pluripotent stem cells. Tissue Eng. Part A. 16(2):405–21

6. Odeleye a. OO, Marsh DTJ, Osborne MD, Lye GJ, Micheletti M. 2014. On the fluid

dynamics of a laboratory scale single-use stirred bioreactor. Chem. Eng. Sci.

111(100):299–312

7. Cormier JT, zur Nieden NI, Rancourt DE, Kallos MS. 2006. Expansion of undifferentiated

murine embryonic stem cells as aggregates in suspension culture bioreactors. Tissue Eng.

12(11):3233–45

8. Fok EYL, Zandstra PW. 2005. Shear-controlled single-step mouse embryonic stem cell

expansion and embryoid body-based differentiation. Stem Cells. 23(9):1333–42

9. Olmer R, Lange A, Selzer S, Kasper C, Haverich A, et al. 2012. Suspension culture of

human pluripotent stem cells in controlled, stirred bioreactors. Tissue Eng. Part C

Methods. 18(10):772–84

Page 89: Computational Fluid Dynamics Modeling of Scalable Stirred

78

10. Dhanasekharan K. 2006. Design and scale-up of bioreactors using computer simulations.

Bioprocess Int. 4(3):34–42

11. Bonvillani P, Ferrari MP, Ducrós EM, Orejas J a. 2006. Theoretical and experimental

study of the effects of scale-up on mixing time for a stirred-tank bioreactor. Brazilian J.

Chem. Eng. 23(1):1–7

12. Johnson C, Natarajan V, Antoniou C. 2014. Verification of energy dissipation rate

scalability in pilot and production scale bioreactors using cfd. Biotechnol. Prog., pp. 1–5

13. Nienow AW, Rielly CD, Brosnan K, Bargh N, Lee K, et al. 2013. The physical

characterisation of a microscale parallel bioreactor platform with an industrial cho cell line

expressing an igg4. Biochem. Eng. J. 76:25–36

14. Taiani JT, Krawetz RJ, Zur Nieden NI, Elizabeth Wu Y, Kallos MS, et al. 2010. Reduced

differentiation efficiency of murine embryonic stem cells in stirred suspension bioreactors.

Stem Cells Dev. 19(7):989–98

15. Shafa M, Day B, Yamashita A, Meng G, Liu S, et al. 2012. Derivation of ipscs in stirred

suspension bioreactors. Nat. Methods. 9:465–66

16. Gareau T, Lara GG, Shepherd RD, Krawetz R, Rancourt DE, et al. 2014. Shear stress

influences the pluripotency of murine embryonic stem cells in stirred suspension

bioreactors. J. Tissue Eng. Regen. Med. 8(4):268–78

17. Lara GG, Hazenbiller O, Gareau T, Shepherd RD, Kallos MS, et al. 2013. Fluid flow

modulation of murine embryonic stem cell pluripotency gene expression in the absence of

lif. Cell. Mol. Bioeng. 6(3):335–45

18. Kelly WJ. 2008. Using computational fluid dynamics to characterize and improve

bioreactor performance. Biotechnol. Appl. Biochem. 49(Pt 4):225–38

Page 90: Computational Fluid Dynamics Modeling of Scalable Stirred

79

19. Kaiser S, Löffelholz C. 2011. Cfd for characterizing standard and single-use stirred cell

culture bioreactors. Comput. Fluid Dyn. Technol. Appl., pp. 97–122

20. Bhole M, Ford C, Bennington CPJ. 2009. Characterization of axial flow impellers in pulp

fibre suspensions. Chem. Eng. Res. Des. 87(4):648–53

21. Ungrin MD, Clarke G, Yin T, Niebrugge S, Nostro MC, et al. 2012. Rational bioprocess

design for human pluripotent stem cell expansion and endoderm differentiation based on

cellular dynamics. Biotechnol. Bioeng. 109(4):853–66

22. Bauwens CL, Peerani R, Niebruegge S, Woodhouse KA, Kumacheva E, et al. 2008.

Control of human embryonic stem cell colony and aggregate size heterogeneity influences

differentiation trajectories. Stem Cells. 26(9):2300–2310

23. Fluri D a, Tonge PD, Song H, Baptista RP, Shakiba N, et al. 2012. Derivation, expansion

and differentiation of induced pluripotent stem cells in continuous suspension cultures.

Nat. Methods. 9(5):509–16

24. Maehle A. 2013. Europe pmc funders group ambiguous cells : the emergence of the stem

cell concept in the nineteenth and twentieth centuries. . 65(4):

25. Lian Q, Chow Y, Esteban MA, Pei D, Tse HF. 2010. Future perspective of induced

pluripotent stem cells for diagnosis, drug screening and treatment of human diseases.

Thromb. Haemost. 104(1):39–44

26. Chidgey AP, Layton D, Trounson A, Boyd RL. 2008. Tolerance strategies for stem-cell-

based therapies. Nature. 453(May):330–37

27. Hanley J, Rastegarlari G, Nathwani AC. 2010. An introduction to induced pluripotent

stem cells. Br. J. Haematol. 151(1):16–24

28. Olmer R, Haase A, Merkert S, Cui W, Palecek J, et al. 2010. Long term expansion of

Page 91: Computational Fluid Dynamics Modeling of Scalable Stirred

80

undifferentiated human ips and es cells in suspension culture using a defined medium.

Stem Cell Res. 5(1):51–64

29. Wong WT, Sayed N, Cooke JP. 2013. Induced pluripotent stem cells: how they will

change the practice of cardiovascular medicine. Methodist Debakey Cardiovasc. J.

9(4):206–9

30. Lo B, Parham L. 2009. Ethical issues in stem cell research. Endocr Rev. 30(3):204–13

31. Lengner CJ. 2010. Ips cell technology in regenerative medicine. Ann. N. Y. Acad. Sci.

1192:38–44

32. Chen VC, Couture SM, Ye J, Lin Z, Hua G, et al. 2012. Scalable gmp compliant

suspension culture system for human es cells. Stem Cell Res. 8(3):388–402

33. O’Brien C, Laslett AL. 2012. Suspended in culture--human pluripotent cells for scalable

technologies. Stem Cell Res. 9(2):167–70

34. Abbasalizadeh S, Larijani MR, Samadian A, Baharvand H. 2012. Bioprocess development

for mass production of size-controlled human pluripotent stem cell aggregates in stirred

suspension bioreactor. Tissue Eng. Part C Methods. 18(11):831–51

35. Ingber DE. 2006. Cellular mechanotransduction: putting all the pieces together again.

FASEB J. 20(7):27

36. Carpenedo RL, Sargent CY, McDevitt TC. 2007. Rotary suspension culture enhances the

efficiency, yield, and homogeneity of embryoid body differentiation. Stem Cells.

25(9):2224–34

37. Wu J, Rostami MR, Cadavid Olaya DP, Tzanakakis ES. 2014. Oxygen transport and stem

cell aggregation in stirred-suspension bioreactor cultures. PLoS One. 9(7):1–12

38. Koay EJ, Athanasiou K a. 2008. Hypoxic chondrogenic differentiation of human

Page 92: Computational Fluid Dynamics Modeling of Scalable Stirred

81

embryonic stem cells enhances cartilage protein synthesis and biomechanical

functionality. Osteoarthr. Cartil. 16(12):1450–56

39. Abaci HE, Truitt R, Luong E, Drazer G, Gerecht S. 2010. Adaptation to oxygen

deprivation in cultures of human pluripotent stem cells, endothelial progenitor cells, and

umbilical vein endothelial cells. Am J Physiol Cell Physiol. 298(11):1527–37

40. Niebruegge S, Bauwens CL, Peerani R, Thavandiran N, Masse S, et al. 2009. Generation

of human embryonic stem cell-derived mesoderm and cardiac cells using size-specified

aggregates in an oxygen-controlled bioreactor. Biotechnol. Bioeng. 102(2):493–507

41. Bratt-Leal AM, Carpenedo RL, McDevitt TC. 2009. Engineering the embryoid body

microenvironment to direct embryonic stem cell differentiation. Biotechnol. Prog.

25(1):43–51

42. Dang SM, Gerecht-Nir S, Chen J, Itskovitz-Eldor J, Zandstra PW. 2004. Controlled,

scalable embryonic stem cell differentiation culture. Stem Cells. 22(3):275–82

43. Cherry RS, Papoutsakis ET. 1986. Hydrodynamic effects on cells in agitated tissue culture

reactors. Bioprocess Eng. 1(1):29–41

44. Kaiser SC, Jossen V, Schirmaier C, Eibl D, Brill S, et al. 2013. Fluid flow and cell

proliferation of mesenchymal adipose-derived stem cells in small-scale, stirred, single-use

bioreactors. Chemie-Ingenieur-Technik. 85(1-2):95–102

45. Lamping SR, Zhang H, Allen B, Ayazi Shamlou P. 2003. Design of a prototype miniature

bioreactor for high throughput automated bioprocessing. Chem. Eng. Sci. 58(3-6):747–58

46. Alvarez MM, Bulnes-Abundis D, Carrillo-Cocoom LM, Aguilar-Yañez JM, Johnson KE.

2008. A scalable 30 ml bioreactor for biopharmaceutical screening. AIChE Conf. Proc.

47. Chisti Y. 1993. Animal cell culture in stirred bioreactors: observations on scale-up.

Page 93: Computational Fluid Dynamics Modeling of Scalable Stirred

82

Bioprocess Eng. 9(5):191–96

48. Spier RE. 1991. Large-scale mammalian cell culture: methods, applications and products.

Curr. Opin. Biotechnol. 2(3):375–79

49. Jayapal K, Wlaschin K, Hu W, Yap G. 2007. Recombinant protein therapeutics from cho

cells-20 years and counting. Chem. Eng. Prog. 103(10):40–47

50. Davidson L. 2003. An introduction to turbulence models

51. Chung. 2009. Computational fluid dynamics., Vol. 82 Spec No. S1–2 pp.

52. Bardina J, Huang PG, Coakley TJ. 1997. Turbulence modeling validation, testing, and

development. NASA Tech. Memo.

53. Wilcox DC. 2006. Turbulence Modeling for CFD (Third Edition). DCW Industries

54. Kuzmin D, Mierka O, Turek S. 2007. On the implementation of the κ-ε turbulence model

in incompressible flow solvers based on a finite element discretisation. Int. J. Comput. Sci.

Math. 1(2/3/4):193

55. Nienow AW, Rielly CD, Brosnan K, Bargh N, Lee K, et al. 2013. The physical

characterisation of a microscale parallel bioreactor platform with an industrial cho cell line

expressing an igg4. Biochem. Eng. J. 76:25–36

56. Sucosky P, Osorio DF, Brown JB, Neitzel GP. 2004. Fluid mechanics of a spinner-flask

bioreactor. Biotechnol. Bioeng. 85(1):34–46

57. Fletcher C a. J a J. 1987. Computational techniques for fluid dynamics 1

58. Zienkiewicz OC, Taylor RL. 2000. The finite element method (fluid dynamics). Methods.

3:347

59. Pironneau O. 1988. Finite element method for fluid. . 1988:176

60. Liovic P, Šutalo ID, Stewart R, Glattauer V, Meagher L. 2012. Fluid flow and stresses on

Page 94: Computational Fluid Dynamics Modeling of Scalable Stirred

83

microcarriers in spinner flask bioreactors. , pp. 1–6

61. Hunt MM, Meng G, Rancourt DE, Gates ID, Kallos MS. 2014. Factorial experimental

design for the culture of human embryonic stem cells as aggregates in stirred suspension

bioreactors reveals the potential for interaction effects between bioprocess parameters.

Tissue Eng. Part C Methods. 20(1):76–89

62. zur Nieden NI, Cormier JT, Rancourt DE, Kallos MS. 2007. Embryonic stem cells remain

highly pluripotent following long term expansion as aggregates in suspension bioreactors.

J. Biotechnol. 129(3):421–32

63. Sen A, Kallos MS, Behie L a. 2002. Expansion of mammalian neural stem cells in

bioreactors: effect of power input and medium viscosity. Dev. Brain Res. 134(1-2):103–13

64. Matsubara C, Kuo T, Wu H. 2013. Comparison of the effects of k - ϵ , k - ω , and zero

equation models on characterization of turbulent permeability of porous media. .

2(June):43–50

65. Dimotakis PE. 2005. Turbulent mixing. Annu. Rev. Fluid Mech. 37(1):329–56

66. Wilcox DC. 1993. Turbulence modeling for cfd

67. Vozzi F, Bianchi F, Ahluwalia A, Domenici C. 2014. Hydrostatic pressure and shear

stress affect endothelin-1 and nitric oxide release by endothelial cells in bioreactors.

Biotechnol. J. 9(1):146–54

68. Yu P, Lee TS, Zeng Y, Low HT. 2005. Fluid dynamics of a micro-bioreactor for tissue

engineering. Fluid Dyn. Mater. Process. 1(3):235–46

69. Wang Y, Chou BK, Dowey S, He C, Gerecht S, Cheng L. 2013. Scalable expansion of

human induced pluripotent stem cells in the defined xeno-free e8 medium under adherent

and suspension culture conditions. Stem Cell Res. 11(3):1103–16

Page 95: Computational Fluid Dynamics Modeling of Scalable Stirred

84

70. Ungrin MD, Joshi C, Nica A, Bauwens C, Zandstra PW. 2008. Reproducible, ultra high-

throughput formation of multicellular organization from single cell suspension-derived

human embryonic stem cell aggregates. PLoS One. 3(2):e1565

71. Sen A, Kallos MS, Behie L a. 2002. Expansion of mammalian neural stem cells in

bioreactors: effect of power input and medium viscosity. Dev. Brain Res. 134:103–13

72. Ismadi M-Z, Gupta P, Fouras A, Verma P, Jadhav S, et al. 2014. Flow characterization of

a spinner flask for induced pluripotent stem cell culture application. PLoS One.

9(10):e106493

73. Annabi N, Tsang K, Mithieux SM, Nikkhah M, Ameri A, et al. 2013. Highly elastic

micropatterned hydrogel for engineering functional cardiac tissue. Adv. Funct. Mater.

23(39):4950–59

74. Almeida PL, Portela R, Sobral RG, Grilo IR, Cidade T, Leal CR. 2014. Living bacteria

rheology: population growth, aggregation patterns and cooperative behaviour under

different shear flows. Phys. Rev. E. 90:022720

75. Sánchez Pérez J a., Rodríguez Porcel EM, Casas López JL, Fernández Sevilla JM, Chisti

Y. 2006. Shear rate in stirred tank and bubble column bioreactors. Chem. Eng. J. 124(1-

3):1–5

76. Hu HH, Joseph DD, Crochet MJ. 1992. Direct simulation of fluid particle motions

77. Nobach H, Bodenschatz E. 2009. Limitations of accuracy in piv due to individual

variations of particle image intensities. Exp. Fluids. 47(1):27–38

78. Zwaka TP, Thomson J a. 2003. Homologous recombination in human embryonic stem

cells. Nat. Biotechnol. 21(3):319–21

79. Scott WH, Thomas CR, Hewitt CJ, Lewis G, Meier SJ, et al. 2012. Scale-down studies for

Page 96: Computational Fluid Dynamics Modeling of Scalable Stirred

85

assessing the impact of different stress parameters on growth and product quality during

mammalian cell culture. 14th Eur. Conf. Mix. 431–36

Page 97: Computational Fluid Dynamics Modeling of Scalable Stirred

86

APPENDIX A: CELL CULTURE

A.1. Static cell culture

A.1.1. Gelatin coating

Each 100 ml of 0.1% gelatin solution was made by adding 0.1 gram of gelatin type B into

100ml sterile double distilled water. The solution was autoclaved, cooled down, and then stored at

40C for future use.

About 3ml of gelatin solution was added to fully cover the surface of a 10 cm2 tissue culture

dish. The tissue culture dish was let sit in the sterile laminar flow hood at room temperature for 20

minutes. The gelatin solution was then removed from the tissue culture dish and discarded. The

dish was let dry in the hood with the lid slightly opened for about 15 to 20 minutes. The gelatin-

coated tissue culture dish could be used instantly, or be wrapped with aluminum foil and stored in

4C fridge, for up to three days.

A.1.2. Static Passaging

Before passaging the cells, 10 ml of murine ESC culture medium was added to each gelatin-

coated tissue culture dish and then pre-incubated at 37C and 5% CO2.

Murine ESCs of D3 cell line were passaged every two days when the cultured reached

70% confluency. At this state the cells were in exponential growth phase and were healthy.

Overgrown culture could promote cell differentiation where the cells lost the ability to form well

defined colonies.

In the beginning of passaging, the consumed culture media was straightforwardly removed

from the tissue culture since murine ESCs adhered well to the bottom surface of the dish. The dish

was rinsed with 8ml DPBS 1X buffer solution. For cell detachment, 3ml of 0.25% trypsin-EDTA

Page 98: Computational Fluid Dynamics Modeling of Scalable Stirred

87

was added to the entire bottom surface of each 10 cm2 tissue culture dish and then incubated at

37C for 3-5 minutes until the cells have detached and lift off from the cell culture surface. After

the incubation, the cells were not completely dissociated, but remained as suspended colonies.

These colonies were broken up into single cells by pipetting the cell suspension up and down for

about 10 times until the single cells suspension had achieved. At least 3 ml of murine ESC culture

medium was added to the cell suspension in each culture dish to block the Trypsin from damaging

the cell membrane. The whole solution was then transferred to 15 ml sterile conical tube. The tube

was then centrifuged at 300 *g for 5 minutes. The supernatant was removed and 2 ml murine ESC

culture medium was added to each conical to suspend the single cells contained in the pellet. Two

50µm of homogenous cell suspension solution were taken for cell counting. Cell count was done

by using hemocytometer. Based on the information from the cell count, appropriate cell suspension

volume was added to each of the new 10cm^2 tissue culture dish to reach the inoculation density

of 40,000 cells/ml. The culture dishes were then incubated at 37C and 5% CO2.

A.2. Standard 100 ml Spinner Flask Bioreactor

A.2.1. Preparation of 100 ml Bioreactor

Six 100 ml spinner flask bioreactors were siliconized with Sigmacoat in order to prevent

the cells from sticking to the wall of the bioreactors. After siliconization process, the bioreactors

were then soaked with DPBS 1X solution for one day and after that, they were washed and soaked

with double distilled water for another day. The bioreactors were then let dry to be autoclaved for

cell culture use.

Agitation rate of three spinner plates was calibrated using Touchless Digital Tachometer.

The speed of the impeller corresponded with the frequency of the beam of light reflected off self-

Page 99: Computational Fluid Dynamics Modeling of Scalable Stirred

88

adhesive tabs. The speed was measured twice to ensure its accuracy and precision. The agitation

rates used in this experiment were 80, 100, and 120 rpm.

A.2.2. Bioreactor Inoculation and Feeding Regime

The cells were first expanded in static culture to in order to have a total of 3.5 x106 cells

for each 100 ml bioreactor. Before bioreactor inoculation, the required amount of murine ESC

culture media was calculated and pre-warmed in the incubator for about 2 hours. The first part of

bioreactor inoculation is basically static culture passing. After counting the cells, appropriate

volume of cell suspension solution was added to each bioreactor to achieve an inoculation density

of 3.5 x104 cells/ml. Three pairs of bioreactors (80, 100, and 120rpm) were agitated using

calibrated spinner plates inside the incubators at 37C and 5% CO2.

Media change took place on day 3 and day 4 of bioreactor cell culture. A pair of bioreactors

was taken out of the incubator and let sit in the laminar flow hood for one minute. Because of

gravity, the cell aggregates will settle at the bottom of the bioreactor, leaving the top part of the

bioreactor clear. Top 50 ml of spent media was removed from each bioreactor and replaced by 50

ml fresh media that had had been pre-warmed in the incubator.

A.2.3. Aggregate Size Measurement using Particle Sizer

The size distribution of murine ESC aggregates was measured using a Beckman Coulter

Multisizer III (Miami, FL, USA) with a 1000-um aperture, which has an effective particle-size

range of 20 µm to 600 µm. Murine ESC aggregate were collected from the bioreactor by gravity

sedimentation; supernatant was aspirated and aggregates were then re-suspended in 60:40 isoton

II (Beckman Coulter, Mississauga, ON) and glycerol (Fischer Scientific, Ottawa, ON). The

Page 100: Computational Fluid Dynamics Modeling of Scalable Stirred

89

instrument was set to apply a current of 800 uA, with a preamp gain of 4, and a maximum cell

volume of 25 µm3. Diluent resistivity and baseline noise were re-calculated prior to each session

by measuring noise level to determine the lower size threshold. The average noise level range

between 2.9 - 3.1% of the aperture diameter, which equate to a lower size limit of 29.8 µm to 34

µm. The amplitude of the voltage pulse associated with each particle was converted to a particle

size based on calibration coefficient using a L90 Latex Particle Standard with a nominal size of 90

µm with the bin size set to 128 and bin spacing on log diameter. For each run, murine ESC

aggregates were diluted such that the coincident counting rate was between 3-10%. Samples were

counted for 20 seconds per run, which returns about an average of 3000 events. An average of 3

runs were collected per sample. After collection, the data were analyzed using the Multisizer 3.53

software with counts plotted against diameters as a histogram using linear bins and a linear scale

for the x-axis. Background signal from debris and single cells were gated out of the data range by

excluding particles smaller than 40 μm in diameter.

Page 101: Computational Fluid Dynamics Modeling of Scalable Stirred

90

APPENDIX B: HYDRODYNAMICS

B.1. Calculation Results

Reynolds number can be calculated as: Re=N(Di)2/υ

The maximum shear stress in SSB was calculated from (63) described in Chapter 3.

Table 5–1: Results from maximum shear stress calculation for 100 ml bioreactor

Agitation rate (rpm)

Reynolds number Power number

Maximum Shear stress (Pa)

80 3570 0.611 0.408

100 4463 0.579 0.556

120 5356 0.554 0.715

Table 5–2: Results from Reynolds number calculation for 10 ml bioreactor

Agitation rate (rpm) Reynolds number

100 1064

120 1277

150 1597

B.2. Quasi-steady state approximation

Below graphs were taken from CFD solutions. Quasi-steady state solution is achieved

when the color lines representing the volumetric shear rate profile at each time point are aligned.

In this case, the quasi-steady state is about 3 s. The cumulative relative volume refers to the sum

of volume fraction up to the current bin. For example, in figure 5-1, above 3 s, the volume fraction

Page 102: Computational Fluid Dynamics Modeling of Scalable Stirred

91

for shear rate values that are below 300 s-1 is large (steep increase in the color lines). After this

range (0-250 s-1), the increase in cumulative volume fraction is insignificant so that the color lines

stay at the same y value. In other words, the volume fraction of shear rate below 400 s-1 is

essentially the same as the volume fraction of shear rate below 300 s-1 because the volume fraction

between 300 s-1 and 400 s-1 is very small and almost equals to zero.

Figure 5–1: Shear rate distribution from 0s-12s in 100 ml Bioreactor at 100 rpm. The quasi-steady state is achieved by about 3s

Shear Rate Distribution in 100 ml Bioreactor at 100 rpm

Page 103: Computational Fluid Dynamics Modeling of Scalable Stirred

92

Figure 5–2: Turbulent dissipation rate distribution from 0s-12s in 100 ml Bioreactor at 100 rpm. The quasi-steady state is achieved by about 3s.

Turbulent Dissipation Rate Distribution in

100 ml Bioreactor at 100 rpm

Page 104: Computational Fluid Dynamics Modeling of Scalable Stirred

93

Figure 5–3: Shear rate distribution from 0s-10s in 10 ml Bioreactor at 100 rpm. The quasi-steady state is achieved by about 4 s.

Shear Rate Distribution in 10 ml Bioreactor at 100 rpm

Page 105: Computational Fluid Dynamics Modeling of Scalable Stirred

94

Figure 5–4: Turbulent dissipation rate distribution from 0s-10s in 10 ml Bioreactor at 100 rpm. The quasi-steady state is achieved by about 4 s.

Turbulent Dissipation Rate Distribution in

10 ml Bioreactor at 100 rpm