Download - Fei Zhengzheng

Transcript
Page 1: Fei Zhengzheng

Membrane Sandwich Electroporation for In Vitro Gene Delivery

Dissertation

Presented in Partial Fulfillment of the Requirements for the Degree Doctor of Philosophy in the Graduate School of The Ohio State University

By

Zhengzheng Fei, M.S.

Graduate Program in Chemical Engineering

The Ohio State University

2009

Dissertation Committee:

Professor L. James Lee. Yang, Advisor

Professor Robert J. Lee

Professor Jessica Winter

Page 2: Fei Zhengzheng

Copyright by

Zhengzheng Fei

2009

Page 3: Fei Zhengzheng

-ii-

ABSTRACT

Gene therapy is the delivery of therapeutic genes into cells and tissues with the

aim of treating and curing a disease. As an enhanced understanding of the roles of genes

in health and disease, gene therapy is showing promise against various diseases such as

cancer, diabetes, Parkinson's disease, and several inherited physiological defects. Viral

transduction is very efficient, but safety issues, such as immune and inflammatory

responses, have hampered their clinical uses in humans. Non-viral methods, including

either chemical transfection with cationic lipids/polymers or physical transfection using

electroporation/microinjection, are becoming attractive approaches.

Electroporation is one of the most popular non-viral gene transfer methods for in

vitro cell transfection. Initial studies with electroporation experienced very low

transfection efficiencies and cell viability, severely limiting the development of this

technology. The emergence of nucleofection (a modified electroporation technology)

provided an efficient means for transfecting cells in vitro. However, nucleofection still

encounters many limitations such as the large number of cells required (>106) and high

cost involved. Moreover, cell viability is still an issue due to the high electric voltage

used and the non-uniform electric field strength distribution generated during the process.

Page 4: Fei Zhengzheng

-iii-

To address these problems, we propose to develop an electroporation system

based on an innovative micro-/nanoengineering technology for in vitro gene delivery. In

our approach, electroporation is carried out in a mild and uniform electric field, with

potential for a wider process window that can be generated to cover a wide range of cell

lines and even primary cells.

A new membrane sandwich electroporation (MSE) approach was demonstrated

using plasmids GFP and SEAP as model materials. NIH 3T3 fibroblasts were tested and a

significant improvement in transgene expression was observed compared to current

electroporation techniques. In the MSE method, the focused electric field enhances cell

permeabilization at a low electric voltage, leading to high cell viability; more important,

the sandwich membrane configuration is able to provide better gene confinement near the

cell surface, facilitating gene delivery into the cells.

Next, we demonstrated the use of femtosecond laser fabricated micro-nozzle

arrays on a gelatin-coated PET membrane for MSE. Using micro-nozzle array enhanced

MSE, we observed high and uniform gene transfection, and good cell viability of mouse

embryonic stem (ES) cells compared to the bulk electroporation. Since typically cells or

tissues from the patients are very limited and therapeutic materials such as plasmids and

oligonucleotides are very expensive, the ability to treat a small number of cells (i.e. a

Page 5: Fei Zhengzheng

-iv-

hundred) offers great potential to work with hard-to-harvest patient cells for patient-

specific ex vivo gene therapy and in vitro pharmaceutical kinetic studies.

Numerical calculation of transmembrane potential qualitatively explains the

observed differences between MSE and bulk electroporation. Since there’s a good

correlation between transfection results and transmembrane potential calculations, the

simulation process with the threshold experiments can be used to predict the transfection

results, and thus largely reduced the trial-and-error window size.

Furthermore, we successfully integrated an electrospun nanofiber scaffold as a

cell-binding substrate into MSE, called nanofiber based MSE. With a micro-well spacer,

the uniform size of mouse ES cell colonies were obtained, and plasmid transfection by

electroporation were performed during colony formation. In addition, repeated plasmid

SEAP transfection of NIH 3T3 fibroblasts was tested and better cell survival and

recovery rate was observed using the electrospun nanofiber scaffold as compared to using

micro-porous membrane. Due to its capacity of extend the exposure time with

reprogramming factors, nanofiber based MSE demonstrated the potential for efficient

induced pluripotent stem (iPS) cell generation by repeated plasmid transfection.

Page 6: Fei Zhengzheng

-v-

ACKNOWLEDGMENTS

I would like to express my sincere gratitude to my advisor Dr. L. James Lee for

his patient guidance, constructive advice and continuous support during my PhD study at

the Ohio State University. I am indebted to Dr. Yubing Xie, Dr. Shengnian Wang, Dr.

Xin Hu, Dr. Hae Woon Choi, Dr. Sadhana Sharma, Mr. Brian Henslee, Mr. Bo Yu, Dr.

Yun Wu, and Dr. Weixiong Wang for their technical support, insightful suggestions, and

encouragements. I would like to acknowledge Dr. Jingjiao Guan, Dr. Xulang Zhang, Dr.

Chee Guan Koh, Dr. Yong Yang, and all former and current group members for their

valuable discussions, and helpful comments. Thanks also go to Mr. Shi-Chiung Yu, Dr.

Chunghe Zhang, Mr. Daniel Gallego, Ms. Natalia Higuita, Mr. Yong Chae Lim, Mr. Chi

Yen, and all the students and staffs at the center for their warm help on my research

project.

The financial support and technical directions from NSF sponsored Nanoscale

Science and Engineering Center for Affordable Polymeric Biological Devices (NSEC-

CAPBD) is appreciated.

Finally, I would like to thank my parents for their love and dedications for raising,

supporting, and educating me. Great appreciations to my husband, Mr. Ziru Zhang, for

his love, support, accompany and understanding through all these years.

Page 7: Fei Zhengzheng

-vi-

VITA

June, 2001…………………………………B.S. Chemical and Biochemical Engineering,

Zhejiang University, Hangzhou, China

March, 2004……………………………… M.S. Chemical and Biochemical Engineering,

Zhejiang University, Hangzhou, China

October 2004 to present.…………………. Graduate Research Fellow,

Chemical and Biomolecular Engineering,

The Ohio State University, Columbus, OH

PUBLICATIONS

1. Fei, Zhengzheng; Wang, Shengnian; Xie, Yubing; Henslee, Brian E.; Koh, Chee

Guan; Lee, L. James. Gene transfection of mammalian cells using membrane

sandwich electroporation. Analytical Chemistry (2007), 79, 5719.

2. Guan, Yixing; Fei, Zhengzheng; Lou, Man; Yao, Shanjing. Choromatographic

refolding of recombinant human interferon gamma by an immobilized sht GroEL191-

345 column. Journal of Chouromatography A (2006), 1107, 192.

3. Guan, Yixing; Fei, Zhengzheng; Lou, Man; Yao, Shanjing. Production of

minichaperone (sht GroEL191-345) and its function in the refolding of recombinant

human interferon gamma. Protein & Peptide Letters (2005), 12, 85.

Page 8: Fei Zhengzheng

-vii-

4. Jin, Ting; Guan, Yixing; Fei, Zhengzheng; Yao, Shanjing. A combined refolding

technique for recombinant human interferon-gamma inclusion bodies by ion-

exchange chouromatography with a urea gradient. World Journal of Microbiology &

Biotechnology (2005), 21, 797.

5. Fei, Zhengzheng; Guan, Yixing; Yao, Shanjing. A colorimetric method to assay

biological activity of recombinant human IFN-γ. Weishengwuxue Tongbao (Chinese

Edition) (2004), 31, 65.

6. Guan, Yixing; Fei, Zhengzheng; Lou, Man; Yao, Shanjing. Minichaperone

(GroEL191-345)-mediated in vitro refolding of recombinant human interferon

gamma inclusion body. Shengwu Huaxue Yu Shengwu Wuli Jinzhan (Chinese

Edition) (2004), 31, 907.

7. Jin, Ting; Guan, Yixing; Fei, Zhengzheng; Lou, Man; Yao, Shanjing. Renaturation of

recombinant human interferon gamma inclusion body by dilution. Huagong Xuebao

(Chinese Edition) (2004), 55, 770.

FIELDS OF STUDY

Major Field: Chemical Engineering

Minor Field: Biochemical Engineering

Page 9: Fei Zhengzheng

-viii-

TABLE OF CONTENTS

Page

Abstract ............................................................................................................................... ii

Acknowledgments............................................................................................................... v

Table of contents.............................................................................................................. viii

List of Tables .................................................................................................................... xv

List of Figures .................................................................................................................. xvi

Chapter 1: Introduction................................................................................................. 1

1.1 Background ............................................................................................................. 1

1.2 Objectives ............................................................................................................... 3

1.2.1 Membrane sandwich electroporation (MSE) .................................................... 3

1.2.2 Micro-nozzle array enhanced MSE .................................................................. 3

1.2.3 Nanofiber based MSE....................................................................................... 4

Chapter 2: Literature review......................................................................................... 6

2.1 Gene delivery.......................................................................................................... 6

2.1.1 Viral versus non-viral ....................................................................................... 6

2.1.2 In vivo versus in vitro........................................................................................ 7

2.2 Electroporation........................................................................................................ 8

2.2.1 Electroporation theory ...................................................................................... 8

Page 10: Fei Zhengzheng

-ix-

2.2.1.1 Pore formation during electroporation.................................................... 9

2.2.1.2 Pore resealing after electroporation ...................................................... 11

2.2.1.3 Possible mechanisms of gene delivery by electroporation ................... 11

2.2.2 Factors to determine cell suspension electroporation ..................................... 12

2.2.2.1 Electric field strength with pulse duration and number ........................ 13

2.2.2.2 Cell density and properties.................................................................... 15

2.2.2.3 Concentration and properties of genetic materials................................ 16

2.2.2.4 Buffer composition ............................................................................... 17

2.2.2.5 Temperature .......................................................................................... 18

2.2.3 In vivo vs in vitro electroporation ................................................................... 19

2.2.4 Commercially available in vitro electroporation systems............................... 20

2.2.5 Microfluidic electroporation ........................................................................... 23

2.3 Genetically modified embryonic stem cells by electroporation ........................... 25

2.3.1 Embryonic stem cells and their properties...................................................... 25

2.3.2 Gene delivery to embryonic stem cells ........................................................... 26

2.3.3 Electroporation of embryonic stem cells ........................................................ 28

2.4 Generation of transgene-free induced pluripotent stem cells by electroporation . 29

2.4.1 Induced pluripotent stem cells ........................................................................ 29

2.4.2 Strategies to generate induced pluripotent stem cells ..................................... 31

2.4.3 Electroporation for transgene-free induced pluripotent stem cells ................. 32

Page 11: Fei Zhengzheng

-x-

Chapter 3: Membrane Sandwich Electroporation....................................................... 34

3.1 Introduction........................................................................................................... 34

3.2 Materials and methods .......................................................................................... 35

3.2.1 DNA preparation............................................................................................. 35

3.2.2 NIH 3T3 fibroblast culture and preparation.................................................... 36

3.2.3 Experimental set-up ........................................................................................ 37

3.2.4 Fabrication and assembly of microfluidic device ........................................... 39

3.2.5 Electroporation procedure............................................................................... 40

3.2.5.1 Bulk electroporation.............................................................................. 40

3.2.5.2 Localized cell electroporation and MSE............................................... 41

3.2.6 Detection of green fluorescence protein (GFP) expression ............................ 43

3.2.7 Assay for secreted alkaline phosphatase (SEAP) Activity ............................. 44

3.2.8 Cell viability.................................................................................................... 45

3.2.9 DNA distribution study by spin-disk confocal microscopy............................ 45

3.3 Results and discussions......................................................................................... 47

3.3.1 Comparison of MSE with bulk electroporation .............................................. 47

3.3.2 Comparison of MSE with localized electroporation....................................... 47

3.3.3 Mechanism analysis by a spin-disk confocal microscope .............................. 50

3.4 Conclusion ............................................................................................................ 52

Chapter 4: Micro-nozzle array enhanced membrane sandwich electroporation ........ 55

Page 12: Fei Zhengzheng

-xi-

4.1 Introduction........................................................................................................... 55

4.2 Fabrication of micro-pore arrays on gelatin-treated polyethylene terephthalate

(PET) track-etched membrane by femtosecond laser ablation ...................................... 56

4.2.1 Micro-patterning of pores by femtosecond pulsed laser ablation................... 56

4.2.2 Femtosecond laser system used in this study.................................................. 58

4.2.3 Thermal effect of femtosecond laser fabrication on gelatin-coated

polyethylene terephthalate surface.............................................................................. 61

4.2.4 Femtosecond laser drilling of gelatin-treated polyethylene terephthalate track-

etched membrane with micro-pore arrays................................................................... 63

4.3 Micro-nozzle enhanced sandwich electroporation................................................ 66

4.3.1 Experimental ................................................................................................... 66

4.3.1.1 Reporter plasmids ................................................................................. 66

4.3.1.2 Culture of mouse embryonic stem cells................................................ 68

4.3.1.3 Experimental set-up .............................................................................. 69

4.3.1.4 Electroporation procedure..................................................................... 70

4.3.1.5 Assay for transfection efficiency and cell proliferation........................ 73

4.3.1.6 Statistical analysis................................................................................. 74

4.3.2 System optimization........................................................................................ 75

4.3.4.1 Converging micro-nozzle vs straight micro-channel............................ 75

4.3.4.2 Effect of porosity and micro-pore shape of top membrane .................. 76

Page 13: Fei Zhengzheng

-xii-

4.3.4.3 Effect of top membrane location........................................................... 80

4.3.3 Comparison of MSE with bulk electroporation and nucleofection ................ 83

4.4 Simulation of transmembrane potential distribution............................................. 86

4.4.1 Three-layer model........................................................................................... 87

4.4.2 Two-dimensional (2-D) simulation process.................................................... 89

4.4.3 Simulation results............................................................................................ 91

4.4.3.1 Effect of cell shape................................................................................ 91

4.4.3.2 Effect of porosity and pore shape of top membrane ............................. 93

4.4.3.3 Converging micro-nozzle vs straight micro-channel............................ 95

4.4.3.4 Comparison of micro-nozzle enhanced sandwich electroporation with

bulk electroporation .............................................................................................. 97

4.5 Conclusion ............................................................................................................ 97

Chapter 5: Nanofiber based membrane sandwich electroporation ............................. 99

5.1 Introduction........................................................................................................... 99

5.2 Materials and methods ........................................................................................ 101

5.2.1 Cell culture.................................................................................................... 101

5.2.2 Fabrication and characterization of nanofiber scaffolds with micro-well

spacers ....................................................................................................................... 101

5.2.2.1 Preparation of electrospun poly (ε-caprolactone) (PCL) /gelatin

nanofiber scaffolds.............................................................................................. 101

Page 14: Fei Zhengzheng

-xiii-

5.2.2.2 Fabrication of PCL /gelatin nanofiber scaffolds with polystyrene (PS)

micro-well arrays ................................................................................................ 103

5.2.2.3 Structure characterization by scanning electron microscopy (SEM).. 104

5.2.3 Experimental set-up for nanofiber based MSE............................................. 105

5.2.4 Electric Resistance Measurements................................................................ 106

5.2.5 Electroporation procedure............................................................................. 107

5.2.5.1 Single cell electroporation .................................................................. 107

5.2.5.2 Cell colony electroporation................................................................. 108

5.2.6 Assays for transfection efficiency and cell proliferation .............................. 109

5.2.7 Cell morphology characterization by confocal microscopy ......................... 110

5.3 Optimization of nanofiber based membrane sandwich electroporation ............. 110

5.3.1 Effect of support membrane.......................................................................... 110

5.3.2 Effect of nanofiber thickness ........................................................................ 114

5.4 Nanofiber based MSE of mouse embryonic stem (ES) cell colony.................... 116

5.4.1 Mouse ES cell colony formation with controlled size .................................. 116

5.4.2 Nanofiber based MSE with vs without micro-well spacer ........................... 118

5.4.3 Nanofiber based MSE vs Bulk electroporation of cell colony ..................... 121

5.5 Nanofiber based MSE of NIH 3T3 fibroblasts ................................................... 122

5.5.1 NIH 3T3 fibroblasts with micro-well spacer ................................................ 122

5.5.2 Repeated SEAP transfection of NIH 3T3 fibroblasts ................................... 123

Page 15: Fei Zhengzheng

-xiv-

5.6 Conclusions......................................................................................................... 126

Chapter 6: Conclusions and recommendations......................................................... 127

6.1 Conclusions......................................................................................................... 127

6.2 Recommendations............................................................................................... 129

6.2.1 Individual cell array trapping........................................................................ 129

6.2.2 Cell membrane permeability experiments .................................................... 130

6.3 Possible ways of in vivo applications.................................................................. 135

References....................................................................................................................... 137

Appendix A: Standard Curve.......................................................................................... 148

Appendix B: Optimization of bulk electroporation and nucleofection of mouse embryonic

stem (ES) cells ................................................................................................................ 149

Appendix C: Analytical solution of transmembrane potential for a two-dimentional (2-D)

cell in bulk....................................................................................................................... 155

Appendix D: G-code generation for fabricating micro-pore arrays by femtosecond laser

......................................................................................................................................... 159

Appendix E: Electroporation of mouse embryoid bodies............................................... 160

Appendix F: Membrane permeability experiment.......................................................... 162

Page 16: Fei Zhengzheng

-xv-

LIST OF TABLES

Page

Table 2.1: Commercially available electroporation systems. .......................................... 21

Table 3.1: Technical specifications of the square wave pulse generator. ........................ 39

Table 3.2: Comparison of conventional bulk electroporation (BE), localized cell electroporation (LCE), and membrane sandwich electroporation (MSE). ....................... 41

Table 3.3: Optical set-up of spin-disk confocal system. (Hemminger et al., 2007) ........ 46

Table 4.1: Technical specifications of the multi-functional pulse generator. .................. 70

Table 4.2: Comparison of nucleofection, conventional bulk electroporation by Bio-Rad Gene Pulser XCell system, and membrane sandwich electroporation (MSE).................. 71

Table 4.3: Top membranes with different pore size, pore density, and pore shape......... 77

Table 4.4: Top membranes with different distance to cell binding membrane................ 81

Table 4.5: Parameters of the three-layer model (Kotnik et al., 1997) ............................. 90

Table 5.1: Properties of three different types of membranes used as support membrane.......................................................................................................................................... 113

Table 5.2: Thickness and corresponding resistance of nanofiber layer controlled by electrospinning time........................................................................................................ 115

Page 17: Fei Zhengzheng

-xvi-

LIST OF FIGURES

Page

Figure 2.1: Theoretical drawing and cryo scanning electronen microscope (cryo-SEM; cryo =cold) images (the upper right insert) of cell membrane before electroporation (top) and pore formation by electric breakdown of the lipid bilayer after electroporation (bottom). (http://www.inovio.com)................................................................................... 10

Figure 2.2: Schematic of cell cycle. Outer ring: I = Interphase, M = Mitosis; inner ring: M = Mitosis, G1 = Gap 1, G2 = Gap 2, S = Synthesis; not in ring: G0 = Gap 0/Resting. (http://en.wikipedia.org/wiki/Cell_cycle) ......................................................................... 16

Figure 2.3: (a) Amaxa’s Nucleofection Biosystem. (www.amaxa.com); (b) Bio-Rad Gene Pulser XCell system. (www.biorad.com) ................................................................ 22

Figure 2.4: Microfluid electroporation devices: (a) single cell electroporation (Khine et

al., 2005); (b) electroporation microchip (Lin et al., 2001).............................................. 24

Figure 2.5: Three pathways to possibly reprogram multi-potent stem cells for treatment of human disorders. (Fuchs and Segre, 2000)................................................................... 27

Figure 2.6: Schematic of iPS-cell-based treatment. (Passier et al., 2008)....................... 30

Figure 2.7: Three strategies to generate induced pluripotent stem cells: (a) retroviral or lentiviral transduction, (b) adenoviral transduction, and (c) plasmid transfection. (Lowry and Plath, 2008) ................................................................................................................ 32

Figure 3.1: Plasmid maps of gWizTM green flurescence protein vector (GFP, 5757 bp) and secreted alkaline phosphatase vector (SEAP, 6569 bp). ............................................ 35

Figure 3.2: Experimental set-up (a) and fluidic device (b) of membrane sandwich electroporation (MSE). (Designed and fabricated by Mr. Shi-Chiung Yu, Dr. Weixiong Wang, and Dr. Chuhe Zhang, 2006) ................................................................................. 38

Figure 3.3: Schematic drawing of (a) cell-binding substrate in MSE disk and (b) DNA migration path during electroporation. ............................................................................. 43

Page 18: Fei Zhengzheng

-xvii-

Figure 3.4: Comparison of membrane sandwich electroporation (MSE) with conventional bulk electroporation and localized cell electroporation (LCE) using plasmid GFP. The green fluorescence indicated green fluorescence protein (GFP) expression 24 hours after bulk electroporation (a), MSE (b), and LCE with genes and cells on (c) opposite sides and (d) the same side of the support membrane. ....................................... 48

Figure 3.5: Comparison of membrane sandwich electroporation (MSE) with localized cell electroporation (LCE) using plasmids SEAP. The bars indicated the activity levels of secreted alkaline phosphatase (SEAP) expressed by NIH 3T3 cells 48 hours after electroporation. Data were plotted with the standard deviation from the mean (n=3). .... 49

Figure 3.6: (a) DNA distribution in the gap between two membranes in the observed domain. The zero position is set at the surface of the top membrane. 3 sets of consecutive images were analyzed at each z slice. (b, c) Confocal images of the slices near the top membrane (x = 0) and in the middle of the two membranes (x = 5.2µm)........................ 53

Figure 4.1: Physical phenomena that are present when machining with a long laser pulse (a) and ultrafast laser pulses (b). (http://www.cmxr.com/Industrial/Handbook.htm)....... 57

Figure 4.2: Femtosecond laser CPA system (Model 2161, Clark-MXR) with micro-station. The arrow indicates the laser pathway. ................................................................ 58

Figure 4.3: Schematic drawing of regenerative amplifier, including High Reflective (HR) mirror, Faraday Rotator (FR), Pockels cell (PC), Dichouroic mirror (DM), and Radiofrequency (RF) unit.(Clark-MXR, CPA 2110 User manual. 2nd Edition, 2004) ... 59

Figure 4.4: Block diagram of the beam delivery system set-up. ..................................... 60

Figure 4.5: Heat effect of various laser beam power on the surrounding gelatin-coated polyethylene terephthalate (PET) surface. ........................................................................ 63

Figure 4.6: SEM image of PET track-etched membrane with average pore size of 400 nm after coating with gelatin. White arrows point out the pores blocked with gelatin. ... 64

Figure 4.7: (a) shape and size of the micro-pores produced under various laser beam power up to 4 mW; (b) SEM images of micro-pores on the gelatin coating side produced at the average laser beam power of 2.5 (upper) and 3.5 mW (lower)............................... 65

Page 19: Fei Zhengzheng

-xviii-

Figure 4.8: Plasmid map of pmaxGFP, encoding the new green fluorescent protein from Pontellina sp. (http://www.lonzabio.com/uploads/tx_mwaxmarketingmaterial/ amaxa_Newsletter_amaxa-news-03.pdf).......................................................................... 67

Figure 4.9: The second generation of MSE system, including a multi-functional pulse generator, and a platform, which is able to handle three fluidic devices in parallel. (Designed and fabricated by Mr. Mr. Shi-Chiung Yu and Dr. Shengnian Wang, 2008) . 69

Figure 4.10: (a) Schematic of MSE disk set-up; (b) Schematic of DNA migration path in the MSE device. DNA molecules migrate from cathode to anode. .................................. 73

Figure 4.11: Effect of different pore shapes, micro-channel () and micro-nozzle (), on mouse ES cell transfection by MSE. The bars indicate total activity of SEAP expression 24 hours after MSE under the optimized electrical field (Appendix B). ......................... 76

Figure 4.12: Comparison of top membrane with different micro-pore size and micro-pore density in MSE: (a) transfection efficiency and (b) cell viability of mouse ES cells 24 hours after MSE. ............................................................................................................... 79

Figure 4.13: Effect of top membrane location in MSE on cell transfection: (a) Transfection efficiency and (b) cell viability of mouse ES cells 24 hours after electroporation. ................................................................................................................. 82

Figure 4.14: Comparison of mouse ES cell transfection by micro-nozzle array enhanced MSE, bulk electroporation by Bio-Rad Gene Pulser, and nucleofection. (a) Transfection efficiency and (b) cell viability 24 hours after electroporation. From left to right, bulk

electroporation with initial input cell number of 6101× and 5101× ; micro-nozzle

enhanced MSE with initial input cell number of 4101× ; and nucleofection with initial

input cell number of 6101× . ............................................................................................. 84

Figure 4.15: GFP transfection of mouse ES cell by micro-nozzle enhanced MSE. A

hundred of cells were trapped on a 1010 × micro-nozzle array, and (a) phase contrast and (b) fluorescent images were taken 24 hours after electroporation. ................................... 86

Figure 4.16: Simulation comparison of top membranes with different pore size, pore density, and pore shape. (a) Schematic diagram of Cases I to IV, from left to right, with electric field lines across/around a single cell; (b) calculated transmembrane potential distribution. θ is the angle around cell surface. ................................................................ 92

Page 20: Fei Zhengzheng

-xix-

Figure 4.17: Simulation comparison of transmembrane potential distribution of top membranes with different pore size, pore density, and pore shape. ................................. 94

Figure 4.18: Simulation comparison between support membranes with micro-nozzles and micro-channels (a) electric potential distribution and electric field lines across/around a single cell near a micro-nozzle (left) and micro-channel (right); (b) calculated transmembrane potential distribution. θ is the angle around cell surface. ........................ 96

Figure 5.1: Schematic diagram of fiber formation by electrospining process where a droplet of a polymer solution is elongated by a high electrical field. (http://nano.mtu.edu/Electrospinning_start.html)........................................................... 102

Figure 5.2: Schematic of fabricating electrospun nanofiber scaffold with polystyrene (PS) micro-well arrays. (1) PDMS stamp with micro-pillar arrays; (2) Drop-cast PS solution; (3) PS solution is spin-coated and it de-wets on the surface of the PDMS stamp; (4) PS in-between the features is removed; (5) PS micro-well arrays were bonded to electrospun nanofiber scaffold by thermal bonding. (Gallego et al., In preparation)..... 104

Figure 5.3: SEM image of PCL/gelatin nanofiber scaffolds with 300 µm PS micro-wells.......................................................................................................................................... 105

Figure 5.4: Schematic drawing of electrospun nanofiber based MSE........................... 106

Figure 5.5: Comparison of different cell-binding substrates used for membrane sandwich electroporation. The transfection efficiency (a) and cell viability (b) of mouse embryonic stem cells, and the resistance (R) of MSE disk (c) were presented using PET membrane only, aluminum oxide membrane only, PET membrane with nanofibers, and aluminum oxide membrane with nanofibers.................................................................................... 111

Figure 5.6: Effect of electrospun nanofiber thickness on the transfection efficiency (a) and cell viability (b) of mouse embryonic stem cells. The thickness of nanofiber layer corresponds to electrospinning time as shown in Table 5.2........................................... 115

Figure 5.7: Confocal images of mouse ES cell colonies after cultured 24 (a, c) and 48 (b, d) hours on randomly distributed PCL/gelatin nanofiber scaffolds without (a, b) and with (c, d) 100 µm PS micro-wells. The cell seeding density was 5,000 / mm2. Cells were fixed with 70% ethanol and stained with PI dye. The length of the standard bars is 100 µm. .................................................................................................................................. 117

Page 21: Fei Zhengzheng

-xx-

Figure 5.8: SEAP transfection of mouse embryonic stem cells by nanofiber based MSE without and with micro-well spacer, bulk electroporation by Bio-Rad Gene Pulser X-Cell system, and nucleofection: (a) transfection efficiency (b) cell viability 24 hours after electroporation. ............................................................................................................... 119

Figure 5.9: Confocal images of mouse ES cells 6 (a, c) and 30 (b, d) hours after nanofiber based MSE without (a, b) and with (c, d) 100 µm micro-well spacer. Mouse ES cells were fixed with 70% ethanol and stained with PI dye. The length of the standard bars is 100 µm................................................................................................................. 120

Figure 5.10: Morphology of NIH 3T3 fibroblasts with (a) and without (b) 300 µm micro-well spacer after 48 hours. .............................................................................................. 123

Figure 5.11: Repeated SEAP transfection of NIH 3T3 fibroblasts using PET micro-porous membrane () and electrospun PCL/gelatin nanofiber scaffolds (): (a) transfection efficiency (b) cell viability at day 1 and 2 post-electroporation. Both cell-binding substrates had the micro-well spacer. ................................................................ 125

Figure 6.1: Schematic illustration of individual cell array trapping by an optical tweezer array created by focused laser beam through a micro-lens array.................................... 130

Figure 6.2: Electroporation of a cell. The electroporation mediated gene transfection process includes two parts: (1) cell membrane break-down and reseal, (2) genes bounding to the cell membrane during the electroporation, and entering cell plasma by endocytosis. (http://www.inovo.com).................................................................................................. 131

Figure 6.3: The third generation of MSE system, including (a) a MSE stage with two MSE disks and (b) an electroporation box, which is able to connect with one AC pulse generator and one DC power supply. Each MSE disk consists of a bottom and a top piece. (Designed and fabricated by Mr. Shi-Chiung Yu and Dr. Weixiong Wang, 2009) ....... 133

Page 22: Fei Zhengzheng

-1-

CHAPTER 1: INTRODUCTION

1.1 Background

Gene therapy is the delivery of therapeutic genes into cells and tissues with the

aim of treating and curing a disease. (Mulligan, 1993) As an enhanced understanding of

the roles of genes in health and disease, gene therapy is showing promise against various

diseases such as cancer, diabetes, Parkinson's disease, and several inherited physiological

defects. (Izumikawa et al., 2005; Le Meur et al., 2007) For instance, gene therapy has

already successfully restored the health of two children with severe combined

immunodeficiency in 1990. (Blaese et al., 1995) Since then, many clinical trials in human

are paving their ways towards treating human diseases.

Over the past decades, the use of genetically modified primary embryonic stem

(ES) cells is becoming an attractive tool for fundamental studies as well as clinical

applications. (Ben-Nuna and Benvenisty, 2006; O’Connor and Crystal1, 2006;

Strulovici1 et al., 2007) ES cells are pluripotent cells derived from the inner cell mass of

an in vitro fertilized embryo grown to the early stage, know as a blastocyst. ES cells are

good candidates for cell-based therapies due to their unlimited self-renewal capacity and

differentiation potential into various cell types that can function as neurons, muscles,

bone, or blood.

Page 23: Fei Zhengzheng

-2-

Genetic manipulation of cultured ES cells provides great potential for ex vivo

gene therapy. (Schindhelm and Nordon, 1999) However, little progress has been made

due to the difficulties involved in successful transfection. Viral transduction of ES cells is

very efficient, but safety issues, such as immune and inflammatory responses, have

hampered their clinical uses in humans. (Niidome and Huang, 2002; Thomas et al., 2003)

Non-viral methods, including either chemical transfection with cationic lipids/polymers

or physical transfection using electroporation/microinjection, are becoming attractive

approaches. (Mehier-Humbert and Guy, 2005; Wells, 2004)

Electroporation is one of the most popular non-viral gene transfer methods for ES

cell transfection. (Tompers and Labosky, 2004) Initial studies with electroporation of ES

cells experienced very low transfection efficiencies (<20%) and cell viability (<50%),

(Mohour et al., 2006) severely limiting the development of this technology. The

emergence of nucleofection (a modified electroporation technology) provided an efficient

means for transfecting ES cells in vitro. (Lorenz and Harnack, 2004; Siemen et al., 2005)

However, nucleofection of ES cells still encounters many limitations such as the large

number of cells required (>106) and high cost involved. Moreover, cell viability is still an

issue due to the high electric voltage used and the non-uniform electric field strength

distribution generated during the process.

Page 24: Fei Zhengzheng

-3-

1.2 Objectives

To address these problems, we propose to develop an electroporation system

based on an innovative micro-/nanoengineering technology for in vitro gene delivery. In

our approach, electroporation is carried out in a mild and uniform electric field, with

potential for a wider process window that can be generated to cover a wide range of cell

lines and even primary cells.

1.2.1 Membrane sandwich electroporation (MSE)

A novel electro-transfection method, called membrane sandwich electroporation

(MSE) was developed. NIH 3T3 fibroblasts were used as cell models and were tested

using reporter genes (plasmid pGFP and pSEAP). In the MSE set-up, the focused electric

field enhances cell permeabilization at a low electric voltage, leading to high cell

viability; at the same time, the sandwich membrane configuration is able to provide better

gene confinement near the cell surface, facilitating gene delivery into the cells. Compared

to current bulk electroporation techniques, the MSE method increased transfection

efficiency and cell viability.

1.2.2 Micro-nozzle array enhanced MSE

The MSE design could not provide a uniform electric field distribution to each

cell because of randomly distributed pores on the track-etched membrane. To address this

Page 25: Fei Zhengzheng

-4-

limitation, we recently fabricated polyethylene terephthalate (PET) membranes with well-

defined micro-hole arrays by using femtosecond pulsed laser ablation. By adjusting the

laser output powers and laser beam focus points, we were able to produce both

converging micro-nozzle and straight micro-channel arrays on the membrane. This new

design was used for gene transfection of mouse embryonic stem (ES) cells. The observed

good transfection results are explained by numerical calculations of the transmembrane

potential distribution on the cell surface.

1.2.3 Nanofiber based MSE

We further integrated the MSE method with a three-dimensional (3-D)

electrospun nanofiber cell culture system for genetic modification of ES cells in a gentle

manner without breaking the colonies. Electrospun fibers are known to provide an in

vivo-like environment required for healthy and viable cells. A combination of MSE and

fibers allows for the cells to be cultured on the same substrate before and after

transfection, and reduces the number of harsh instances, such as repeated cell

trypsinization for getting single cell suspension, and longer time duration outside the

incubator, which cells have to undergo during conventional electroporation. Furthermore,

this method provides the flexibility of using appropriate type of support membrane

substrates specific for an intended application and cell type. The applicability of this

method was demonstrated using mouse ES cells and NIH 3T3 fibroblasts as a model

system.

Page 26: Fei Zhengzheng

-5-

It should be stated that the results and the descriptions of membrane sandwich

electroporation presented in Chapter 3 have already been included in a published journal

paper: Fei et al., 2007. The experimental procedure and many results presented in the

Chapter 4 and 5 are being included in two manuscripts under preparation.

Page 27: Fei Zhengzheng

-6-

CHAPTER 2: LITERATURE REVIEW

2.1 Gene delivery

A major challenge in gene therapy is to deliver therapeutic genes into the

designated target cells with a high transfection efficiency and minimal cell damage.

There are two basic criteria used to distinct the field of gene delivery: viral versus non-

viral and in vivo versus in vitro.

2.1.1 Viral versus non-viral

Viral gene delivery uses genetically modified viruses to introduce genes into cells

or tissues by infection. These recombinant viruses are highly efficient, and have been

used in clinical trials since 1990. (Blaese et al., 1995) However, safety concerns, such as

immune and inflammatory responses, are limiting their clinical applications, (Thomas et

al., 2003) especially after the first gene therapy death caused by the adenovirus-based

treatment reported in 1999. (Hollon, 2002) Another challenge relates to virus-based gene

delivery is the high cost of viral vector production. In addition, there is always the fear

that viral vectors, once inside the patient, may recover their ability to cause disease.

Page 28: Fei Zhengzheng

-7-

Non-viral systems have become widespread for gene delivery because of their

relative safety and low manufacturing cost. Non-viral methods can be divided into two

groups: chemical and physical (or mechanical). (Niidome and Huang, 2002) Chemical

methods use reagents such as cationic lipids and polymers, or proteins that will complex

with DNA or RNA, condensing it into particles and directing it to the cells. Physical

methods involve introducing plasmids into cells by mechanical ways, including

microinjection, particle bombardment (gene gun), electroporation, sonoporation, and

laser irradiation. As compared to chemical methods, physical methods can transfer naked

DNA into cells directly and avoid the harmful side effects associated with synthetic

vectors, such as lipoplexes and polyplexes. (Mehier-Humbert and Guy, 2005)

2.1.2 In vivo versus in vitro

In vivo gene delivery refers to introducing the genes directly into the affected

tissue inside a living organism, and requires that the vector be targeted specifically and at

sufficiently high frequencies to the desired cell types. Viruses and liposomes have been

widely investigated as in vivo carriers, but safety issues such as immune response and

cytotoxicity have limited their clinical applications. (Niidome and Huang, 2002; Thomas

et al., 2003) Physical methods are more benign, because they can directly transfer naked

DNA into cells and avoid the risks associated with introducing a secondary agent.

(Mehier-Humbert and Guy, 2005; Wells, 2004)

Page 29: Fei Zhengzheng

-8-

In vitro gene delivery involves performing the experiments in a controlled

environment outside a living organism. This type of research aims at describing the

effects of an experimental variable on a subset of an organism's constituent parts.

Currently, electroporation is one of the most popular research tools used for gene transfer

into mammalian cells in vitro.

2.2 Electroporation

2.2.1 Electroporation theory

Electroporation (or electropermeabilization) is a process that can induce transient

openings in cell plasma membrane by executing external electric field on cells, and thus

increased the permeability of cell membrane. (Chernomordik et al., 1987; Chang et al.,

1992; Weaver and Chizmadzhev, 1996; Gabriel and Teissie, 1997 and 1999) Since it was

developed in the early 1980’s (Neumann et al., 1982), electroporaiton has been widely

used to deliver exogenous macromolecules into cytoplasm, ranging from ions, drugs,

dyes, tracers, and antibodies, to DNA, RNA and oligonucleotides. (Coulberson et al.,

2003)

Normally, one or multiple short and high-voltage pulses are imposed on cells, and

cell membranes are permeable at the locations of highest transmembrane potential

gradient, typically the areas closest to the electrodes (the anode side opens first, followed

Page 30: Fei Zhengzheng

-9-

by the cathode side). (Gehl, 2003) Since the applied pulses are short, the lipid bilayer

structure can be restored and, therefore, the cell survives.

2.2.1.1 Pore formation during electroporation

The experimental observations of electroporating planar lipid bilayers

(Zimmermann and Vienken, 1982; Chernomordik et al., 1983; Glaser et al., 1988) and

the molecular dynamic (MD) simulations (Mounir, 2005) demonstrated that the kinetics

of pore formation by electric breakdown of the lipid bilayer (as shown in Figure 2.1)

includes three steps:

(1) Induction: Water fingers penetrate the hydrophobic core of the lipid bilayer

from both sides at the beginning of applying the electric field, and extend to form the

water wires.

(2) Stabilization: Hydrophobic polar heads migrate toward the interior of the

bilayer surrounding with hydrophilic polar heads, and thus stable large water pores are

formed.

(3) Resealing: water channels disappear and polar lipid head-groups go back to

the lipid-water interface after the electric field is turned off.

Chang and Reese (1990) observed pore-like crater structures or volcano funnels

of 20 to 120 nm diameters in electroporated red blood cell membrane by rapid-freezing

electron microscopy. The analytical methods indicated that the initial electropores should

Page 31: Fei Zhengzheng

-10-

be much smaller. (Neumann et al., 1999) The observation of Chang and Reese (1990)

most probably results from the enlargement of smaller primary pores by osmotic or

hydrostatic pressure due to Maxwell stress.

Figure 2.1: Theoretical drawing and cryo scanning electronen microscope (cryo-SEM;

cryo =cold) images (the upper right insert) of cell membrane before electroporation (top)

and pore formation by electric breakdown of the lipid bilayer after electroporation

(bottom). (http://www.inovio.com)

Page 32: Fei Zhengzheng

-11-

Besides leading to electric breakdown of the lipid bilayer, the transmembrane

potential gradient can also cause the opening of many protein channels in cell membrane.

(Tsong, 1991) Therefore, electric field induced pore formation occurs in both lipid

domain and protein channels.

2.2.1.2 Pore resealing after electroporation

The kinetics of pore resealing after electroporation is mainly measured by

studying the time course of the decrease in membrane conductivity (Chen and Lee, 1994;

Chernomordik et al., 1987), membrane permeability to small inorganic ions (Bier et al.,

2002), and the fraction of cells permeable to certain membrane-impermeant compounds

(Gabriel and Teissie, 1995; Rols and Teissie, 1989; Zimmermann et al., 1980). The

resealing process after cell electroporation was also analyzed theoretically, and compared

with experimental data available in the literature. (Saulis, 1997; Bier et al., 2002) It is

generally agreed that the resealing of the membrane requires seconds to minutes, and it is

a random process.

2.2.1.3 Possible mechanisms of gene delivery by electroporation

Two possible mechanisms for gene delivery by electroporation have been

proposed and verified by both in vitro and in vivo studies (Golzio et al., 1998; Mir et al.,

1999): diffusion-controlled and electrophoresis-driven uptake, both generally occurring

Page 33: Fei Zhengzheng

-12-

near the temporary openings facing the cathode. Diffusion-controlled uptake involves

DNA molecules binding to the cell surface and diffusing into the cytoplasm leading to the

cell transfection. (Sukharev et al., 1992) Electrophoresis-driven uptake follows two steps:

(1) Induced by an applied electric field, cell membrane breaks down and results in

formation of lipid vesicles containing DNA; and (2) DNA molecules are loaded into cells

by endocytosis after electroporation. (Klenchin et al., 1991; Sukharev et al., 1992;

Glogauer et al., 1993)

2.2.2 Factors to determine cell suspension electroporation

The degree of transfection by electroporation is highly cell-dependent, and

normally assessed by two aspects: the transfection rate of therapeutic materials

(transfection efficiency) and the survival rate of electropermeabilized cells (cell

viability). (Chu et al., 1987)

The transfection efficiency of electroporation depends on various factors,

including the status of cells (e.g. growth phase, density, size, orientation) (Tsong, 1991),

the physical and chemical properties of therapeutic materials (e.g. DNA size,

configuration and concentration) (Tsong and Xie, 1997), and applied electric conditions

(e.g. pulse amplitude, duration and number) (Kotnik et al., 2003; Gehl, 2003). Other

issues that should be considered when performing electroporation include temperature,

post-pulse manipulation and composition of electrode and pulsing medium.

Page 34: Fei Zhengzheng

-13-

Generally, cell survival is not a concern when working with bacterial cells, but

can be a major issue with mammalian cells. The viability of mammalian cells depends on

the resealing of electric-field-induced pores in cell membrane after electroporation and

the extent of excess leakage of intracellular molecules during electroporation. (Rols and

Teissie, 1990)

2.2.2.1 Electric field strength with pulse duration and number

For a certain cell line, extent of cell permeabilization is dependent on the

amplitude of electric pulses, and degree of molecule transportation is dependent on the

duration and number of electric pulses. (Gehl, 2003)

The transmembrane potential difference can be described by the following

equation (Teissie and Rols, 1993):

)]exp(1[cos τθ textm rfgEV −−=∆

(2.1)

where:

∆Vm: transmembrane potential difference, V/cm;

f : shape factor of impacted cells under an external field;

g: relative electric permeability;

Eext: external electric field strength, V/cm;

r: radius of the cell, µm;

θ : angle between Eext and the point on the cell membrane;

Page 35: Fei Zhengzheng

-14-

t: elapsed time after the field is turned on, s;

τ : charging or relaxation time of membrane, s.

Typically, for a pure dielectric pulse, g =1. Transient and reversible breakdown of

the membrane can be achieved only when the transmembrane potential surpasses the cell

membrane capacitance, expressed as a threshold potential, ∆Vs. Considering the same

lipid bilayer feature of cellular membrane for eukaryotic cells, ∆Vs is reported to be 200

mV. (Gehl, 2003) As shown in Eq. (2.1), transmembrane potential difference is

dependent on the external field and cell properties. (Kotnik et al., 1997) When ∆Vm >∆Vs,

it is believed that transient hydrophilic pores are formed.

It is generally agreed that the formation of electropores takes place on the order of

micro- to miliseconds, whereas resealing of the membrane requires seconds to minutes.

(Rols and Teissie, 1990) If extensive electroporation is applied, there will be a high

probability for slow recovery of cells and the loss of intracellular components, leading to

irreversible permeabilization and eventually cell death. In addition, cell lysis occurs as

the phenomena of the electroporation-induced apoptosis, if the electric field strength is

too high or the pulse duration is too long. Typical electric field strengths used in cell

suspension electroporation are 0.5 ~ 1.0 kV/cm for mammalian cells. (Rols and Teissie,

1990)

Pulse shape has also played different roles in cell electroporation. (Kotnik et al.,

2003) There are two popular types of pulse shapes: square wave (rectangular) or

Page 36: Fei Zhengzheng

-15-

exponentially decay. (Chang et al., 1992) An exponentially decay pulse is generated by

discharging a capacitor, which is relatively simple but cannot provide precisely control of

pulse parameters. A square wave pulse is generated by a more sophisticated instrument to

accurately control both the intensity and the duration of each pulse.

2.2.2.2 Cell density and properties

One essential factor for achieving good transfection and survival rate of the cells

is cell density in the suspension during the electroporation. Usually if the cell density is

lower than half million per milliliter ( 5105× /mL), the survival rate of cells is extremely

low. However, if the cell density is too high, the transfection efficiency largely decreases

as the result of the shield effect among the cells and electrofusion between cells in

contact (Tsong, 1991).

The capacity to undergo and recover from electroporation is highly cell-

dependent. Two significant factors are the size and growth phase of the recipient cells.

Larger the cell diameter, lower the recovery rate of cell membrane after electroporation.

(Potter, 1993) Electroporation of mammalian cells in logarithmic growth phase is much

more efficient than treatment in early or stationary growth phase. (Anderson et al, 1991)

Additionally, plasmids can access the nucleus more easily during the mitotic phase (M

phase) of the cell cycle (Figure 2.2) than entering post-mitotic state (G0 phase) outside

of the cell cycle. (Golzio et al., 2002a)

Page 37: Fei Zhengzheng

-16-

Figure 2.2: Schematic of cell cycle. Outer ring: I = Interphase, M = Mitosis; inner ring:

M = Mitosis, G1 = Gap 1, G2 = Gap 2, S = Synthesis; not in ring: G0 = Gap 0/Resting.

(http://en.wikipedia.org/wiki/Cell_cycle)

2.2.2.3 Concentration and properties of genetic materials

Tsong and Xie (1997) have reported that the amount of DNA bound to the cell

membrane during electroporation is proportional to the amount of DNA transferred into

the cells post-electroporation. It is essential to facilitate binding of DNA to the cell

surface during electroporation. The increase of DNA concentration can arise the

possibility of DNA bound to cell membrane, especially around the effective area facing

the cathode, and thus enhance the transfection efficiency.

Page 38: Fei Zhengzheng

-17-

The size (molecular weight) affects uptake pathway: Smaller genes, such as

oligonucleofide, siRNA, and microRNA, are more easily transferred through the cell

membrane by diffusion of surface bound genes and electroosmosis of genes in the bulk

solution, while large plasmid delivery into cytosol is dominated by the electrophoresis of

surface bound DNA. The integration and expression of the loaded DNA is strongly

dependent on the DNA configuration (supercoiled, circular-relaxed, or linearized). Linear

DNA can be more effectively integrated into the host genome, but relative unstable in the

cytoplasm compared to supercoiled and circular DNA. (Tsong and Xie, 1997)

2.2.2.4 Buffer composition

Another key factor to influence the cell transfection is the recipe of

electroporation buffer. In many cases, the presence of high ionic strength in

electroporation buffer causes electric arching, especially during the application of longer

high-voltage pulses, and may kill most of cells. Therefore buffer solutions having a low

ionic strength and thus low conductivity were used in order to avoid cell damage as a

result of high currents when using high conductivity buffers. (Rols and Teissie, 1990)

Divalent cations, such as calcium (Ca2+) and magnesium (Mg2+) ions, are frequently

added to the electroporation buffer. Free Ca2+ and Mg2+ stimulates the pore resealing,

leading to an increase in cell survival rate. (Klenchin et al., 1991) Mg2+ (up to 10mM)

facilitates the binding of DNA to the cell membrane, resulting in an elevated transfection

rate. (Tsong, 1991) In addition, it has been reported that low concentration of glucose or

Page 39: Fei Zhengzheng

-18-

sucrose has certain effect on transfection efficiency and cell viability in some type of

cells. (Myers and Tisa, 2003)

Change of the microenvironment of cells is believed to be the cause of cell death.

Although the role of loss of intracellular components is unclear, electroporation can

abolish the osmotic balance of cells with the external environment, especially the

equilibrium of individual ions, like the equilibrium of Na-K-ATPase at the plasma

membrane. More importantly, the colloidal-osmotic effect is lethal as it causes the

swelling of cells. The excess large molecules inside cells lead to the osmotic imbalance

inside. Ions are transported towards the outside to attain ionic equilibrium and water

flow inside, resulting in swelling. If the resealing of the cell membrane is ineffective,

continuous swelling may lead to cell bursting and death. Several surfactants, such as

dextran, PEG and poloxamer 188, were used in the external medium to balance the

colloid osmotic pressure changed by DNA or macromolecule delivery. (Kanduser et al.,

2003)

2.2.2.5 Temperature

Besides reversible breakdown of the lipid domain induced during the

electroporation, many voltage-sensitive protein channels in the cell membrane may open

when they experience an applied external electric field. Local temperature increase as a

result of Joule heating (or ohmic / resistive heating) generated with electric current, and

Page 40: Fei Zhengzheng

-19-

high temperature around cells may denature these protein channels. To avoid the Joule

heating effect, it is commonly suggested to pre-incubate the cell suspension on ice before

electroporation.

On the other hand, a better cell survival rate was achieved when culturing the

cells in 37°C incubator right after electroporation rather than leaving them at room

temperature. It is possible that an increase of post-pulse temperature facilitates the

membrane recovery, and thus improved the cell viability.

2.2.3 In vivo vs in vitro electroporation

In vivo electroporation has been attempted in the past two decades to deliver

therapeutic materials into targeted tissues and organs. (Jaroszeski et al., 1999) Generally

it can be divided into two major groups: delivery of anticancer drugs (such as bleomycin

and cisplatin) for cancer therapy, and delivery of DNA, RNA or DNA vaccines for gene

therapy and DNA vaccination. For cancer therapy, successful examples of electroporation

trials have been done on animal or human patients with basal cell carcinoma, melanoma,

and head and neck cancers. In gene therapy or DNA vaccination, any type of cell or

tissue can be a target, and successful targeting locations have been reported in skin, liver,

muscle, brain and tumors; with muscle and skin as the two most popular targeted tissues.

For in vivo electroporation, the electric field distribution is a key factor. Different shapes

of electrode probes have been tested, such as plate electrodes and needle electrodes.

Page 41: Fei Zhengzheng

-20-

Needle electrodes are preferred due to good electric contact as well as stability. However,

the electric field distribution must be given particular consideration. Because of the lack

of a fundamental understanding of the biophysics of electroporation, current in vivo

process must be performed by trial-and-error.

To understand the fundamentals of the delivery mechanism, in vitro

electroporation has been widely used by researchers as a routine method for transient cell

transfection by plasmids or short strands of interfering RNA (eg. siRNA and microRNA).

(Coulberson et al., 2003) Because of its simplicity and reproducibility, in vitro

electroporation has also been used for introducing exogenous macromolecules, such as

antibodies and enzymes, into cells, and thus has been identified as a rapid and affordable

screening method to assess the behavior of fusion proteins within cells before using more

costly in vivo transgenic animal models.

Experimental results indicate that cell death is more likely in vitro than in vivo.

The most likely reason is that there is a large loss of intracellular molecules due to

abundant extracellular space for in vitro studies, while this space is much more limited

for in vivo studies inside tissues. (Chang et al., 1992)

2.2.4 Commercially available in vitro electroporation systems

In vitro electroporation systems are commercially available and parallel plate

electrode probes are widely used. The in vitro electroporation systems available in the

Page 42: Fei Zhengzheng

-21-

U.S. market are listed in the Table 2.1. The leading systems, Amaxa Nucleofection

Biosystem and Bio-Rad Gene Pulser XCell system, are shown in Figure 2.3.

Table 2.1: Commercially available electroporation systems.

Company Generators Cost Advantage Disadvantage

Nucleofector Single cuvette:

$10K Amaxa

96-well Shuttle $20K

High

Efficiency

Unknown electric

condition;

Expensive cell-

specific buffer

Gene Pulser XCell Single cuvette:

$6K Bio-Rad

Gene Pulser

MXCell

96-well Shuttle:

$ 10K

BTX ECM 830 Single cuvette or

25-/96-well: $6K

Low cost; Limited efficiency

Page 43: Fei Zhengzheng

-22-

Figure 2.3: (a) Amaxa’s Nucleofection Biosystem. (www.amaxa.com); (b) Bio-Rad

Gene Pulser XCell system. (www.biorad.com)

Although commercial electroporation systems have been reasonably successful,

the variation in transfection efficiency is large. Furthermore, the outcome of an

electroporation protocol is cell-type specific and varies among cells in a given

population. The quick development of proper protocols for different cells (in vitro) or

tissues (in vivo) is difficult due to the lack of detailed understanding of the

Page 44: Fei Zhengzheng

-23-

electroporation mechanism and many variations presenting in the bulk systems. It is

necessary to develop new electroporation methods that can replace the current

approaches.

2.2.5 Microfluidic electroporation

Microfluidic devices have been used for electroporation since 1999 (Huang and

Rubinsky). Microfluidic electroporation offers several advantages compared to bulk

electroporation: (1) various configurations of electrodes can be patterned and placed in

close proximity so that very low potential differences are sufficient for electroporation of

the cell membrane (can be as low as 1 V/cm); (Khine et al., 2005) (2) cell handling and

manipulation are much easier and uniform electroporation is possible; (3) less reagents

are needed for transfection; and (4) the process can be monitored at the single cell level

for intracellular content transport.

Current microfluidic electroporation devices focus mostly on cell lysis. (Lee and

Tai, 1999; Lu et al., 2005) In general, solid electrodes are patterned in close proximity

with different shapes (e.g., saw-tooth structure) to concentrate electric field strength, and

the electric current is focused at a small constriction zone. In recent years, more and more

designs were being used for drug and gene delivery to mammalian cells. (Huang and

Rubinsky, 2003; Lin et al., 2001; Khine et al., 2005) Figure 2.4 shows two microfluidic

electroporation designs for mammalian cell transfection.

Page 45: Fei Zhengzheng

-24-

Figure 2.4: Microfluid electroporation devices: (a) single cell electroporation (Khine et

al., 2005); (b) electroporation microchip (Lin et al., 2001).

Page 46: Fei Zhengzheng

-25-

2.3 Genetically modified embryonic stem cells by electroporation

2.3.1 Embryonic stem cells and their properties

Embryonic stem (ES) cells are derived from the inner cell mass of a blastocyst.

ES cells were first isolated from mouse embryos in 1981 by two independent groups.

(Evans and Kaufman, 1981; Martin, 1981) More than 15 years of research on mouse ES

cells led to the breakthrough of human ES cells. In 1998, Thomson et. al. successfully

isolated hES cells from human blastocysts and grew the cells in the laboratory.

ES cells have two important characteristics that distinguish them from other types

of stem cells. The first is their developmental potential called pluripotency, which means

that they are able to differentiate into any of the three primary germ layers and thus into

different types of mammalian tissues. Second, ES cells can proliferate for one year or

more in cell culture without differentiating while maintaining the karyotype, which

allows the production of unlimited numbers of undifferentiated cells. Because of the

capability of self-renewal and vast differentiation, ES cells are identified as a promising

cell source that could be used therapeutically to treat tissue injury as well as genetic

disorders.

In general, ES cells are maintained in the undifferentiated state on feed layers of

embryonic fibroblasts or on feeder-free substrate (gelatin for mouse ES cells and Matrigel

for human ES cells) in the presence of basic factors (leukemia inhibitory factor for mouse

ES cells, and fibroblast growth factor for human ES cells). (Carpenter et al., 2003; Xu et

Page 47: Fei Zhengzheng

-26-

al., 2001) When placed in suspension cultures, ES cells spontaneously form three-

dimensional cell spheroid-like aggregates called embryoid bodies (EBs), and differentiate

into various cell types, e.g. neural cells muscle cells, and blood cells. (Drukker and

Benvenisty, 2003) Cultured ES cells can also be induced to differentiate in vitro to

specific cell types by adding growth factors to the media. This provides the possibility for

cell-based therapies to repair damaged or destroyed cells or tissues.

2.3.2 Gene delivery to embryonic stem cells

Over the past decade, the use of genetically modified ES cells has become an

attractive tool for fundamental studies as well as clinical applications. (Ben-Nuna and

Benvenisty, 2006; O’Connor and Crystal1, 2006; Strulovici1 et al., 2007) Recent studies

in cell culture systems indicate that gene delivery to ES cells has the potential to treat

tissue injury and cure genetic disorders. (Fuchs and Segre, 2000) For example,

introducing the gene Nurr1 into ES cells has made it possible to regulate the formation of

dopamine-producing nerve cells for the treatment of Parkinson’s disease. (Kim et al.,

2002; Lindval and Kokaia, 2006)

Figure 2.5 summarizes three pathways to possibly reprogram multi-potent stem

cells for treatment of human disorders. (Fuchs and Segre, 2000) Two of the pathways

involve controlling in vitro differentiation of primary ES cells, and transplanting

differentiated cells. Control of ES cells differentiation to specific cell lineages in vitro is

Page 48: Fei Zhengzheng

-27-

very challenging due to the complexity of signals involved in determining the fate of

stem cells. (Odorico et al., 2001) Genetically modifying ES cells with certain

transcription factors, growth factors and/or other signal molecules can guide their

differentiation process to specific cell types. (Zeng et al., 2003) Efficient gene delivery

techniques are the key to achieving the full potential of ES cells.

Figure 2.5: Three pathways to possibly reprogram multi-potent stem cells for treatment

of human disorders. (Fuchs and Segre, 2000)

Page 49: Fei Zhengzheng

-28-

2.3.3 Electroporation of embryonic stem cells

Electroporation is one of the most popular non-viral gene transfer methods for ES

cell transfection. (Tompers and Labosky, 2004) However, high cell concentration (at

least five million per milliliter) and a large number of cells (at least half million for a

cuvette) are often required to obtain acceptable gene transfection and cell viability,

because the entire cell membrane is affected by the applied high electric field and the

electric field distribution is non-uniform due to randomly suspended cells and genes

during the electroporation process. Amaxa nucleofection (now Lonza), the best

commercial electroporation-based technique, demonstrated good transfection efficiency:

more than 85% transfection efficiency in mouse ES cells (Lorenz and Harnack, 2004),

and over 66% in human ES cells (Hohenstein et al., 2008; Siemen et al., 2005). However,

nucleofection relies on an expensive electroporation buffer that varies from cell to cell.

Neither the recipes of the cell-dependent electroporation buffer nor the electric

parameters are disclosed by the manufacturer. As designed for individual cell suspension

electroporation, it is hard to obtain good transfection of cell colonies or EBs with limited

programs suggested by the manufacturer. Also, there are concerns that the unknown

addictives in the electroporation buffer might affect the behavior of transfection cells,

limiting its use for research only.

Page 50: Fei Zhengzheng

-29-

2.4 Generation of transgene-free induced pluripotent stem cells by electroporation

2.4.1 Induced pluripotent stem cells

Induced pluripotent stem (iPS) cells, just as the name implies, are artificial

pluripotent stem cells derived from non-pluripotent cells transfected by specific stem

cell-associated genes. Takahashi and Yamanaka of Kyoto University in Japan first

announced successful iPS cell generation in June 2006. Through direct reprogramming of

mouse skin cells, they obtained iPS cells by introducing four transcription factors,

Oct3/4, Sox2, c-Myc, and Klf4. (Takahashi and Yamanaka, 2006) In 2007, Yamanaka’s

group and Thomson’s group successfully generated iPS cells from human somatic cells.

Their work was published in Cell and Science respectively almost at the same time.

(Takahashi et al., 2007; Yu et al., 2007)

As natural ES cells, iPS cells possess high nucleus-to-cytoplasma ratio and typical

compact colony morphology. Chromosome analysis and flow cytometry expression

analysis shows that iPS cells have normal karyotypes, and expresse ES cell-specific cell

surface marker and genes. More importantly, iPS cells can be differentiatioted into

derivatives of all three germ layers. (Takahashi et al., 2006 and 2007; Yu et al., 2007 and

2009)

iPS cell lines provide an alternative source of autologous tissue for transplantation

in regenerative therapy, especially patient-specific cell-based therapy, as shown in Figure

2.6 (Passier et al., 2008; Maherali and Hochedlinger, 2008; Lensch, 2009) Not only can

Page 51: Fei Zhengzheng

-30-

they be used for studing disease models and testing new drugs against those diseases, iPS

cells may also contribute to the fast screening of inducing pluripontent cells to

differentiate into desired cell types. (Figure 2.6) Besides its promise in scientific field,

iPS cells also solved ethical issues associated with the use of fertilized embryos to obtain

ES cells and oocytes for somatic cell nuclear transfer.

Figure 2.6: Schematic of iPS-cell-based treatment. (Passier et al., 2008)

Page 52: Fei Zhengzheng

-31-

2.4.2 Strategies to generate induced pluripotent stem cells

There are three strategies to generate iPS cells as shown in Figure 2.7. (Lowry

and Plath, 2008) Prior to 2008, the generation of iPS cells was based on the integration

of the reprogramming genes into the host-cell genome by retroviral transduction, which

may also turn on cancer-causing genes. (Takahashi et al., 2006 and 2007; Yu et al.,

2007) To avoid the danger of insertional mutagenesis or potential oncogenesis using

retroviruses, scientists started to pursue the vector-free and transgene-free reprogramming

methods. Adenoviral transduction without genome integration (Stadtfeld et al., 2008)

and repeated plasmid transfection (Okita et al., 2008) were reported to achieve

reprogramming of mouse iPS cells. However, the reprogramming efficiencies of vector-

free and transgene-free approaches were less than 0.01%, much lower than that of

retroviral transduction. (Lowry and Plath, 2008) Using non-integrating episomal vectors,

Yu et al. (2009) has reported the generation of human iPS cells completely free of vector

and transgene sequences. Although the programming efficiency was improved to ~ 0.1%

using episomal vectors, it is still only one tenth of the highest efficiency achieved by

retroviral transduction.

Page 53: Fei Zhengzheng

-32-

Figure 2.7: Three strategies to generate induced pluripotent stem cells: (a) retroviral or

lentiviral transduction, (b) adenoviral transduction, and (c) plasmid transfection. (Lowry

and Plath, 2008)

2.4.3 Electroporation for transgene-free induced pluripotent stem cells

To address the challenge of low efficiency of vector-free and transgene-free

reprogramming, scientists developed an alternative strategy involving delivery of a single

vector with all the required reprogramming genes by electroporation, and using piggyBac

Page 54: Fei Zhengzheng

-33-

transposon technology to integrate the vector into the host-genome. (Kaji et al., 2009;

Nagy et al. 2009; Stadtfeld and Hochedlinger, 2009; Yusa et al., 2009) Kaji et al. (2009)

constructed all four genes, Oct4, Sox2, c-Myc, and Klf4, into a single plasmid vector

with a 2A-linkage. By using the 2A-peptide sequence, the vector is able to undergo self-

removal from a peptide undergoing translation, and thus realized complete removal of

vectors and transgenes after reprogramming (Pera, 2009; Stadtfeld and Hochedlinger,

2009) Using piggyBac technology, the single vector can easily integrate into the host-

genome. (Nagy et al. 2009; Stadtfeld and Hochedlinger, 2009) Through transient

expression of the transposase enzyme, the integrated genes can also be removed from the

host genome in a in a highly efficient and seamless fashion. (Nagy et al. 2009; Yusa et

al., 2009)

Page 55: Fei Zhengzheng

-34-

CHAPTER 3: MEMBRANE SANDWICH ELECTROPORATION

3.1 Introduction

Among the physical and mechanical methods, electroporation-enhanced delivery

of plasmid vectors is gaining acceptance for both in vitro and in vivo applications. The

electroporation process induces transient openings in the plasma membrane by executing

electric pulses on cells and driving genes or drugs into the cytoplasm. It is applicable to a

wide variety of animal cells and tissues, simple to perform, and easy to use. However,

conventional bulk electroporation requires the use of a high electric voltage, leading to

low cell viability and limited transfection efficiency.

In this chapter, we present a much less invasive and more efficient gene delivery

method, called membrane sandwich electroporation (MSE). We trapped cells on a track-

etched polyethylene terephthalate (PET) membrane. Cell immobilization on a porous

surface leads to localized cell electroporation, allowing the use of a low applied voltage

to achieve temporarily dielectric breakdown of the cell membrane. When we placed

another track-etched PET membrane on the top of the immobilized cells and sandwiched

the cells between the two membranes, we observed a significant improvement of gene

transfection with minimal cell damage.

Page 56: Fei Zhengzheng

-35-

3.2 Materials and methods

3.2.1 DNA preparation

Reporter Plasmids, gWizTM green flurescence protein vector (GFP, 5757 bp) and

secreted alkaline phosphatase vector (SEAP, 6569 bp), were purchased from Aldevron

(Fargo, ND), and purified with an EndoFree Plasmid Maxi Kit from Qiagen (Valencia,

CA, USA) according to the manufacturer’s instructions. Figure 3.1 shows the maps of

gWiz GFP and SEAP.

(Continued)

Figure 3.1: Plasmid maps of gWizTM green flurescence protein vector (GFP, 5757 bp)

and secreted alkaline phosphatase vector (SEAP, 6569 bp).

Page 57: Fei Zhengzheng

-36-

Figure 3.1 continued

3.2.2 NIH 3T3 fibroblast culture and preparation

NIH 3T3 fibroblasts (mouse embryonic fibroblast cell line, CRL-1658) were

purchased from American Type Culture Collection (ATCC, Manassas, VA), and were

grown in the culture medium. The culture medium consists of Dulbecco’s modified

Eagle’s medium: Nutrient Mix F-12 (D-MEM/F-12, Catalog No.10565), 2 mM L-

glutamine (Catalog No. 25030), 1mM MEM sodium pyruvate (Catalog No. 11360), and

10% (v/v) newborn calf serum (NCS, heat-inactivated, Catalog No. 26010). NIH 3T3

Page 58: Fei Zhengzheng

-37-

cells were maintained in 25 cm2 T-flasks at 37oC with 5% CO2 and passaged every 2~3

days using 0.05% trypsin with 1mM EDTA·4Na (Catalog No. 25300).

Before experimentation, NIH 3T3 cells were plated on a Ф35mm plastic petri

dish and allowed to grow till 70 ~ 90% confluence. Cells were harvested by

trypsinization, and pelleted via centrifuge. The pellet was washed with Dulbecco's

Phosphate Buffered Saline (D-PBS, Catalog No. 14190) without calcium or magnesium

at least twice, and resuspended in GIBCO Opti-MEM I reduced-serum medium (Catalog

No. 51985). Cells were counted with a hemocytometer, and cell suspensions were

adjusted to desired cell concentration for electroporation.

All the media and solutions for cell culture and transfection were purchased from

invitrogen (Carlsbad, CA) unless otherwise specified.

3.2.3 Experimental set-up

The experimental set-up of membrane sandwich electroporation (MSE) is shown

in Figure 3.2(a). The MSE device and platform was fabricated using a high precision

computer numerically controlled (CNC) machine (AeroTech Inc, Pittsburgh, PA). The

MSE platform is connected with a square wave pulse generator designed and built in our

lab. The technical specifications of the square wave pulse generator are given in Table

3.1.

Page 59: Fei Zhengzheng

-38-

Figure 3.2: Experimental set-up (a) and fluidic device (b) of membrane sandwich

electroporation (MSE). (Designed and fabricated by Mr. Shi-Chiung Yu, Dr. Weixiong

Wang, and Dr. Chuhe Zhang, 2006)

Page 60: Fei Zhengzheng

-39-

Table 3.1: Technical specifications of the square wave pulse generator.

Voltage Pulse length Pulse interval No. Pulses

(maximum)

Current

(maximum)

10 ~ 240 V 1 ~ 999 ms 1 ~ 999 ms 6 180 mA

3.2.4 Fabrication and assembly of microfluidic device

The fluidic device consists of a pair of cross channels connected by a center hole

as shown in Figure 3.2(b). One channel is present on the top of the device while the

other is on the bottom. Both channels are 500 µm in width and depth. The channel on the

top of the device intersects with a 1 cm diameter reservoir located at the center of the

device where membranes can be fixed to the device.

The MSE device was fabricated in an Acrylite® acrylic plastic sheet (thickness:

1/16", US Plastic Corporation, Lima, OH) using the CNC machine (AeroTech Inc,

Pittsburgh, PA). A 50-µm thick polymethylmethacrylate (PMMA) film (Fisher Scientific

Inc., USA) was welded on the backside of the device using a thermal film laminator

(Catena 35, GBC, Addison, IL), enclosing the bottom channel but allowing top access via

through reservoirs at the ends.

In order to remove any contaminants on the surface, the MSE device was cleaned

in acetone ultra-sonication bath for 10 min and successively rinsed with isopropyl alcohol

Page 61: Fei Zhengzheng

-40-

(IPA) and deionized water (DI water) before thermal bonding. Each MSE device was

reused more than 20 times.

3.2.5 Electroporation procedure

3.2.5.1 Bulk electroporation

Bulk electroporation of NIH 3T3 fibroblasts was carried out using Bio-Rad Gene

Pulser XcellTM unit (Catalog No. 165) with CE module and ShockPod. 100 µL of

suspended cells (1 × 105 cells) and 5µg DNA sample was loaded into the 2-mm gap

electroporation cuvet. The electroporation conditions were chosen according to Tekle et

al. (1991)’s work, and are given in Table 3.2. After electroporation, every 10 µL of the

cell suspension was transfer to a well of the 24-well plate with 240 µL culture media. The

transfection efficiency and cell viability were measured at 24 or 48 hours after

electroporation.

Page 62: Fei Zhengzheng

-41-

Table 3.2: Comparison of conventional bulk electroporation (BE), localized cell

electroporation (LCE), and membrane sandwich electroporation (MSE).

LCE and MSE

Method BE

(Tekle et al., 1991) Electroporation DNA Attraction

Field Strength (V/cm) 1,600 35 3.5

Pulse Frequency (Hz) 40 1 100

Pulse duration (ms) 0.4 500 5

No. of pulses 1 5 300

Electroporator Bio-Rad Gene Pulser Home-made

Note: The pulse type is a biopolar square wave. Electric field strength is defined as the

voltage amplitude divided by the distance between two electrodes. Voltage amplitude is

the absolute value of peak value minus zero.

3.2.5.2 Localized cell electroporation and MSE

For localized cell electroporation and MSE, a 3-mm diameter track-etched

polyethylene terephthalate (PET) membrane (BD Biosciences, San Jose, CA) was used as

the support membrane with an average pore size of 400 nm, and fixed at the center

reservoir of the fluidic device by sealing tape (shown in Figure 3.1(a)). First, a 10-µL

drop of suspended cells ( 4101× cells) was loaded onto the support membrane, and a

Page 63: Fei Zhengzheng

-42-

vacuum of 334 ± kPa was used to trap the cells on the support membrane. Next, another

3-mm diameter track-etched PET membrane with an average pore size of 3 µm was

added on top of the immobilized cells with a spacer of ~ 10 µm between the two

membranes. Opti-MEM I reduced-serum medium was then loaded into the channels and

the center reservoir. Two thin silver wire electrodes were placed in the inlet and outlet

reservoirs, and 0.5 µg DNA sample was loaded into the reservoir with the cathode.

Finally, a DNA attraction step was performed, followed by electroporation (The

conditions are given in Table 3.2). The DNA molecules were migrated from the cathode

side to the anode side as shown in Figure 3.3(b). After 15 to 20 minutes, the support

membrane with the cells was transferred to a 24-well plate with 250 µL culture media in

each well. The transfection efficiency and cell viability were measured at 24 or 48 hours

after electroporation.

Page 64: Fei Zhengzheng

-43-

Figure 3.3: Schematic drawing of (a) cell-binding substrate in MSE disk and (b) DNA

migration path during electroporation.

3.2.6 Detection of green fluorescence protein (GFP) expression

The transfection efficiency of plasmid GFP (pGFP) was qualified by the

percentage of the cells with green fluorescence among the cells observed by phase

contrast with the same visual area. An inverted digital microscope (Eclipse TS100,

Nikon, USA) equipped with X-Cite 120 fluorescence illumination system (EXFO Life

Sciences Division, Canada) was used to detect GFP expression and cell viability 24 hours

after electroporation. For each visual field, cells were first observed by phase contrast,

Page 65: Fei Zhengzheng

-44-

then by fluorescence. B-2E/C fluorescence filter set (Excitation filter wavelengths: 465 –

495 nm, Dichouromatic mirror cut-on wavelength: 505 nm, Barrier filter wavelengths:

515 – 555 nm; Nikon, USA) was used for green fluorescence detection. Both phase

contrast and fluorescence images were taken with a digital camera (SPOT Insight 2MP

Firewire Color Mosaic, Diagnostic Instruments, Inc., Sterling Heights, MI) set gain at 4,

and controlled by the SPOT Advanced software.

3.2.7 Assay for secreted alkaline phosphatase (SEAP) Activity

The transfection efficiency of plasmid SEAP (pSEAP) was expressed as the total

SEAP activity per ten thousand initial input cells. Samples of culture media were

collected 48 hours after electroporation and determined by a colorimetric assay based on

the hydrolysis of p-Nitrophenyl phosphate (pNPP). pNPP substrate solution was fresh

prepared using SIGMAFAST™ pNPP tablets (Sigma-Aldrich, Catalog No. N1891, St.

Louis, MO). 100 µL of culture media and 25 µL of pNPP substrate solution were added

into each well of a 96-well plate. The plate was incubated in the dark for approximately

15 minutes at room temperature, and read at the wavelength of 405 nm on a multi-well

plate reader (GENios Pro, Tecan, Durham, NC, USA). A standard curve of absorbance

value at 405nm versus total SEAP activity (mU) was generated (Appendix A), and then

the experimental readings at 405nm were normalized to total SEAP activity per ten

thousand initial input cells.

Page 66: Fei Zhengzheng

-45-

3.2.8 Cell viability

NIH 3T3 cells were collected in a single-cell suspension after electroporation, and

mixed with equal volume of trypan blue stain (Invitrogen, Catalog No. 15250). 10 µL of

the mixture was loaded onto a hemocytometer (a counting chamber covered with a cover

slide), and then counted under a microscope. The trypan blue stainable cells were the

dead cells, and thus the viability of the cells was calculated in accordance with the

percentage of the cells excluded from staining.

3.2.9 DNA distribution study by spin-disk confocal microscopy

To facilitate visualization, large λ-DNA molecules stained with the dye YOYO-1

were used in the confocal microscopic experiments instead of plasmids GFP and SEAP.

Lambda DNA (λ-DNA, N6-methyladenine-free, 48502 bp) was purchased New England

Biolabs (Ipswich, MA) and used as received without further purification. According to

the staining procedure given by Perkins et al. (1997), the λ-DNA solution was diluted to

the concentration of 0.03 µg/ml in the imaging buffer, and incubated with YOYO -1

(EM491/EX509, Molecular Probes, Eugene, OR) at a dye-base pair ratio of 1:4 for 1 ~ 2

hours. The imaging buffer consisted of YOYO-1 iodide and 20% 2-mercaptoethanol

(Sigma-Aldrich, Catalog No. M7154) in sterile Tris-EDTA (TE) buffer (Fluka, Catalog

No. 93302). YOYO-1 iodide is an intercalating dye that stains the DNA backbone and

makes it possible to visualize the DNA. 2-Mercaptoethanol is a strong reducing agent that

Page 67: Fei Zhengzheng

-46-

retards photo bleaching of the YOYO-1 fluorescence dye by scavenging oxygen from the

solution. (Xiao et al., 2007)

A spin-disk confocal microscope (VisiTech International, Alexandria, VA) with

Z-stacking was used to trace the location of DNA molecules in the gap of the MSE setup.

The Yokogawa spin-disk scanning unit (CSU-22) synchouronized with Hamamatsu EM

CCD camera was connected to an inverted microscope (Olympus IX-81, Tokyo, Japan).

A solid-state laser line was employed with the supplying 50 mW at 491 nm wavelength.

A Jena piezo-controller was mounted underneath the 60X oil objective to carry out a Z-

direction scans with submicron accuracy (the minimum distance is 100 nm). The entire

system was controlled using VoxCell Scan software from VisiTech.

Table 3.3: Optical set-up of spin-disk confocal system. (Hemminger et al., 2007)

Parameter Value

Pinhole diameter 50 µm

Magnification (M) 60

Numerical Aperture (NA) 1.42

Refractive index (n) 1.52

Excitation wavelength 491 nm

Emission wavelength 515 nm

Page 68: Fei Zhengzheng

-47-

3.3 Results and discussions

3.3.1 Comparison of MSE with bulk electroporation

Using plasmid GFP and NIH 3T3 fibroblasts as reporter gene and model cells,

bulk electroporation and MSE were tested. A significant improvement was observed of

green fluorescence protein (GFP) expression by using the MSE method over the

conventional bulk electroporation method (Figure 3.4(a, b)). In the conventional bulk

electroporation, a layer of foam was observed due to cell lysis in the high intensity

electric field (1,600 V/cm). In comparison, a much low electric field with the amplitude

of 35 V/cm was applied in our MSE method, and more than 90% of cell viability was

achieved. Cell viability was quantified right after electroporaion with the trypan blue

method.

3.3.2 Comparison of MSE with localized electroporation

Using plasmid GFP and NIH 3T3 fibroblasts as reporter gene and model cells,

two different cases were tested for localized cell electroporation. Cells and genes were

placed on either opposite sides (Figure 3.4(c)) or the same side (Figure 3.4(d)) of the

support membrane. In both cases, only a slight improvement was observed of green

fluorescence protein (GFP) expression over the conventional bulk electroporation method

(Figure 3.4(a)). When the MSE method was used, most cells survived after the

treatment, and GFP expression (Figure 3.4(b)) was much higher than in localized cell

Page 69: Fei Zhengzheng

-48-

electroporation (Figure 3.4(c, d)). Using another plasmid SEAP, the levels of transgene

expression mediated by localized cell electroporation and MSE were quantified. The

amount of secreted alkaline phosphatase (SEAP) expression mediated by MSE was

improved about 40% over localized cell electroporation (Figure 3.5).

(Continued)

Figure 3.4: Comparison of membrane sandwich electroporation (MSE) with

conventional bulk electroporation and localized cell electroporation (LCE) using plasmid

GFP. The green fluorescence indicated green fluorescence protein (GFP) expression 24

hours after bulk electroporation (a), MSE (b), and LCE with genes and cells on (c)

opposite sides and (d) the same side of the support membrane.

Page 70: Fei Zhengzheng

-49-

Figure 3.4 continued

0

5

10

15

20

25

30

35

SE

AP

Activity (

mU

)

Cell & gene on opposite sides

Cell & gene on the same side

MSE

Figure 3.5: Comparison of membrane sandwich electroporation (MSE) with localized

cell electroporation (LCE) using plasmids SEAP. The bars indicated the activity levels of

secreted alkaline phosphatase (SEAP) expressed by NIH 3T3 cells 48 hours after

electroporation. Data were plotted with the standard deviation from the mean (n=3).

Page 71: Fei Zhengzheng

-50-

3.3.3 Mechanism analysis by a spin-disk confocal microscope

To explain why the MSE method promotes transgene delivery, Z-stacking was

carried out using a spin-disk confocal microscope with a 60X oil objective to trace the

local concentration of DNA molecules during the electroporation process. This system is

similar to the Confocal Laser Scanning Microscopy (CLSM) micro-PIV system (Park et

al., 2004), but equipped with a Yokogawa CSU-22 spin-disk unit and Hamamatsu EM

CCD camera. It is capable of scanning more than 120 full-frame images (1024×1024

pixels) per second, sufficient to directly measure the DNA distribution inside the

sandwich gap in the MSE setup.

To facilitate visualization, YOYO-1 conjugated large λ-DNA molecules were

used instead of plasmids GFP and SEAP. The DNA solution was loaded to the cathode

side, while the anode side was loaded with the buffer solution only. Scanning was carried

out every 0.4 µm across the 10-µm gap between two membranes. The time for DNA to

diffuse across the two planes was calculated according to Fick’s second laws

D

lt

2

2

≈ (3.1)

where

t: time for DNA to diffuse across the two planes,

D : diffusion coefficient of λ-DNA molecules,

4.0=l µm: diffusion distance.

Page 72: Fei Zhengzheng

-51-

The diffusion of λ-DNA molecules is predicted according to the Zimm model in

good solvent conditions. (Smith et al., 1996; Hur and Shaqfeh, 2001)

36.0~ ≈−v

DNALD µm2/s (3.2)

where

6.0=ν : scaling exponent,

17=DNAL µm contour length for λ-DNA.

Thus the time for DNA to diffuse across the two planes is 2.0~ second, while the

scanning time interval between two adjacent planes is 04.0~ second, about one fifth of

the time for DNA to diffuse across the two planes. Therefore, DNA molecules observed

in adjacent planes must be different individual molecules.

Three sets of consecutive images were analyzed at each z slice and experimental

results were then plotted in Figure 3.6. A large number of DNA molecules were found in

the gap after electroporation, with a decreasing number of DNA molecules near the

membrane surface. Without the top membrane, DNA molecules were hardly seen this

time within the same distance of 10 µm from the bottom membrane. During

electroporation, the extent of cell permeabilization is dependent on the amplitude of

electric pulses, while the transport of the polyanionic DNA molecules into the cells is

driven by the electrophoretic force (Mir et al., 2005) and is dependent on the duration and

number of electric pulses (Gabriel and Teissie, 1997). The nano-scale pores in the

support membranes allow a focused electric field on the cell membrane, and thus enhance

Page 73: Fei Zhengzheng

-52-

cell permeabilization at a low electric voltage. However, negatively charged DNA

molecules quickly migrate away from the negatively charged cell surface after the pulse

duration because of electric repulsion. This can seriously limit gene transfer into the cells.

When a negatively charged track-etched PET membrane is placed on top of the cells,

DNA molecules are prevented from diffusing away, as demonstrated in Figure 3.6. Thus,

the sandwich membrane configuration is able to provide better gene confinement near the

cell surface to facilitate genes transport into the cells.

3.4 Conclusion

A new membrane sandwich electroporation (MSE) approach was demonstrated

using plasmids GFP and SEAP as model materials. NIH 3T3 fibroblasts were tested and a

significant improvement in transgene expression was observed compared to current

electroporation techniques. In the MSE method, the focused electric field enhances cell

permeabilization at a low electric voltage, leading to high cell viability; more important,

the sandwich membrane configuration is able to provide better gene confinement near the

cell surface, facilitating gene delivery into the cells.

Page 74: Fei Zhengzheng

-53-

Figure 3.6: (a) DNA distribution in the gap between two membranes in the observed

domain. The zero position is set at the surface of the top membrane. 3 sets of consecutive

images were analyzed at each z slice. (b, c) Confocal images of the slices near the top

membrane (x = 0) and in the middle of the two membranes (x = 5.2µm).

Page 75: Fei Zhengzheng

-54-

Successful examples of in vitro electroporation trials have been done on animal

and human patients. Since typically cells or tissues from the patients are very limited and

therapeutic materials such as plasmids and oligonucleotides are very expensive, our MSE

method with the ability to deal with small number of cells with high transfection

efficiency and cell viability, offers a great impossibility for ex vivo gene therapy. The

applicability of the MSE method to primary cells and hard-to-transfect cells (such as

mouse embryonic stem cells and leukemia cells) is currently under investigation in our

laboratory.

Page 76: Fei Zhengzheng

-55-

CHAPTER 4: MICRO-NOZZLE ARRAY ENHANCED MEMBRANE

SANDWICH ELECTROPORATION

4.1 Introduction

In Chapter 3, we demonstrated a membrane sandwich electroporation (MSE)

technique. However, the design could not provide a uniform electric field distribution to

each cell because of randomly distributed pores on the track-etched membrane.

Consequently, the gene transfection efficiency was limited.

To address this limitation, we formed well-defined micro-pore array on nano-

porous polyethylene terephthalate (PET) track-etched membranes using femtosecond

pulsed laser ablation. By adjusting the laser output powers and laser beam focus points,

we were able to produce both converging micro-nozzle and straight micro-channel arrays

on the membrane. This new design was tested by plasmid gWiz SEAP transfection of

mouse embryonic stem (ES) cells. The micro-nozzle array enhanced MSE method was

optimized. Effect of membrane porosity and pore shape was explored. The observed

transfection results are further explained by numerical calculations of the transmembrane

potential distribution on the cell surface.

Page 77: Fei Zhengzheng

-56-

4.2 Fabrication of micro-pore arrays on gelatin-treated polyethylene

terephthalate (PET) track-etched membrane by femtosecond laser ablation

4.2.1 Micro-patterning of pores by femtosecond pulsed laser ablation

In the past several decades, the pulsed laser beam systems have been used in

micro-manufacturing a wide range of materials. Due to its flexibility and non-cleanroom

operation, laser micro-machining offers rapid and cost-effective evaluation of design

concepts during prototyping phases. Recent studies demonstrated that there is nearly no

thermal damage on the surrounding material if the laser-material interaction time is

picoseconds or shorter. (Aguliar et al., 2005; Varel et al., 1997) The major benefits of a

femtosecond (or ultrashort) laser pulse include its ability to produce very high peak intensity

( 1610≥ W/cm2) and rapid deposition of energy into the material. The difference between a

long laser pulse and ultrafast laser pulses on material removal and heat transfer is shown

in Figure 4.1. Therefore, femtosecond laser pulsing appears to be a very promising tool

for biomedical application (Serafetinides, 1997), such as surface modification of poly(ε-

caprolactone) (PCL) membrane for tissue engineering application (Tiaw et al., 2005), and

patterning of collagen for directed cell attachment (Liu et al., 2005).

Page 78: Fei Zhengzheng

-57-

Figure 4.1: Physical phenomena that are present when machining with a long laser pulse

(a) and ultrafast laser pulses (b). (http://www.cmxr.com/Industrial/Handbook.htm)

Page 79: Fei Zhengzheng

-58-

4.2.2 Femtosecond laser system used in this study

Figure 4.2 shows the femtosecond laser system used this study. An ultrashort

CPA laser system (Model 2161, Clark-MXR, Dexter, MI) was used as a femtosecond

laser energy source. It has a central wavelength of 775 nm, pulse width of 150

femtoseconds, and pulse repetition frequency of 3kHz. The system has a single-mode

erbium (Er) fiber oscillator to provide a seed pulse for chirped pulse amplification in a

Ti:Al2O3 –based regenerative amplifier.

Figure 4.2: Femtosecond laser CPA system (Model 2161, Clark-MXR) with micro-

station. The arrow indicates the laser pathway.

Page 80: Fei Zhengzheng

-59-

The seed pulse from the frequency doubled SErF laser (λ=775nm) is stretched to

protect optics during pulse amplification. The stretched seed pulse is then transferred to a

regenerative amplifier to amplify the laser intensity. The optimum injection or cavity

dumping time is controlled by electronic devices (DT-505, Clark-MXR) and is critical for

the power and pulse duration. As shown in Figure 4.3, the injected seed laser is amplified

through an Ti:Al2O3 cavity which is pumped by a frequency doubled Nd:YAG laser

(λ=532nm) through a dichouroic mirror. After multiple passes of amplification, typically

5 times, the polarization of a Pockels cell is changed to eject the amplified beam to the

compressor.

Figure 4.3: Schematic drawing of regenerative amplifier, including High Reflective (HR)

mirror, Faraday Rotator (FR), Pockels cell (PC), Dichouroic mirror (DM), and

Radiofrequency (RF) unit. (Clark-MXR, CPA 2110 User manual. 2nd Edition, 2004)

Page 81: Fei Zhengzheng

-60-

Figure 4.4 presents the block diagram of the optical beam delivery system. The

maximum output power of the system is 2.5 W. Since the typical range of power for

micromachining is less than 10 mW, it was attenuated to the mW scale by thin-film

polarizing beam splitters (PBS) and a λ/2 wave plate. The PBS regulates the transmission

and reflection rate based on the polarization direction, which can be adjusted by a λ/2

wave plate and incidence angle of laser beam.

Figure 4.4: Block diagram of the beam delivery system set-up.

(Farson et al., 2006 and 2008)

Page 82: Fei Zhengzheng

-61-

The diameter of raw beam was measured to be 5 mm and the beam quality, M2,

was measured by using auto correlator to be 1.3 where a perfect Gaussian beam is M2=1.

The laser beam can be turned on or off using an external high speed mechanical shutter

with 4 ms shutter control time (LS055, NMlaser Inc), and was finally delivered to the

target material on a high precision X-Y stage (Parker-Hannifin) with 0.5 µm resolution

and 50 mm travel distance. The focusing optics was mounted on the Z axis with 0.5 µm

resolution and 25 mm travel distance. A 50x infinity corrected microscope objective lens

with numerical aperture (NA) of 0.42 (50x M Plan Apo NIR, Mitutotyo) was used for

fine focusing. Attenuated laser power was measured by a power meter (PM100, Thorlab)

placed right under the laser focusing lens. (Farson et al., 2006 and 2008)

For the consistency and easy focusing purpose, a coaxial vision system was

installed, allowing the material to be visually located at the focus within ±1 µm range

where focal depth was 1.6 µm for this selected optics.

4.2.3 Thermal effect of femtosecond laser fabrication on gelatin-coated

polyethylene terephthalate surface

Femtosecond laser is an excellent tool for direct micro-patterning of biomaterials.

Because of the ultrashort contact between the laser beam and materials, there is very low

heat transfer to surrounding materials. (Farson et al., 2006 and 2008) Gelatin has been

widely used as a feeder-free substrate for maintaining mouse embryonic stem (ES) cells

Page 83: Fei Zhengzheng

-62-

in an undifferentiated status. Different output laser beam powers up to 6 mW were tested

to determine the desirable power range for a minimal thermal effect on the surrounding

gelatin-coated polyethylene terephthalate (PET) surface.

A PET surface was coated with 0.1 % (w/v) gelatin (from porcine skin, Catalog

No. G2500, Sigma-Aldrich, St. Louis, MO) at room temperature for 30 minutes, yielding

a thin layer with thickness of 2 to 3 µm. Multi micro-wells were produced by scanning

the focused femtosecond pulsed laser beams under various powers over the gelatin-

coated PET surface. The G-code was programmed to fabricate 100 µm diameter micro-

wells with a center-to-center distance of 200 µm in a 22 × array. The G-code was

imported into a computer controlled motion system (MX80L, Parker Hannifin, Rohnert

Park, CA) with a 0.5-µm resolution in the X, Y, and Z axes. After Ultraviolet (UV)

sterilization, mouse ES cells were seeded at desired density to determine the heat effect

on the bioactivity of the surrounding gelatin-coated surface. Figure 4.5 shows the

morphology and distribution of mouse ES cells after 2-day culture. Mouse ES cells grew

very well around the wells at both 2 and 4 mW, while nearly no mouse ES cells were

observed on the surrounding surface of the wells milled at 6 mW. This implies that an

average output laser beam power up to 4 mW can provide proper irradiance incident to

fabricate the gelatin-coated PET surface without denaturing the gelatin coating.

Page 84: Fei Zhengzheng

-63-

Figure 4.5: Heat effect of various laser beam power on the surrounding gelatin-coated

polyethylene terephthalate (PET) surface.

4.2.4 Femtosecond laser drilling of gelatin-treated polyethylene terephthalate

track-etched membrane with micro-pore arrays

Polyethylene terephthalate (PET) track-etched membranes (thickness: 10 µm,

pore size: 400 nm) were coated with 0.1 % (w/v) gelatin at room temperature for 30

minutes, yielding a thin layer with thickness of 2 to 3 µm. After gelatin coating, most of

the original submicron pores were blocked (Figure 4.6). A range of laser beam power

from 1 to 4 mW was then used to drill a periodic array of micro-pores in a gelatin-coated

PET membrane. A scanning electron microscope (SEM, Hitachi S-4300 Field Emission)

was used to characterize the dimension and surface roughness of the pores on the gelatin-

Page 85: Fei Zhengzheng

-64-

coated PET membrane after femtosecond drilling. To perform SEM, a thin gold layer (50

nm) was sputter-coated on the Samples by Emitech k550x sputter coater.

Figure 4.6: SEM image of PET track-etched membrane with average pore size of 400

nm after coating with gelatin. White arrows point out the pores blocked with gelatin.

By varying the laser beam power pattern, the shape and the size of micro-pores

could be controlled. The relationship between pore dimension and laser beam power is

plotted in Figure 4.7(a). Smooth and well-defined micro-pores with minimal gelatin

Page 86: Fei Zhengzheng

-65-

denaturing was obtained by femtosecond laser drilling at relatively low pulse energies in

our work, but a pulse energy below 2 mW was not high enough to penetrate the gelatin-

coated PET membrane.

(Continued)

Figure 4.7: (a) shape and size of the micro-pores produced under various laser beam

power up to 4 mW; (b) SEM images of micro-pores on the gelatin coating side produced

at the average laser beam power of 2.5 (upper) and 3.5 mW (lower).

Page 87: Fei Zhengzheng

-66-

Figure 4.7 continued

Two different micro-pore shapes were generated: a converging nozzle at the laser

output power between 2 and 3.5 mW, and a straight channel at the output laser power

higher than 3.5 mW. Figure 4.7(b) shows SEM images of micro-pores on the gelatin

coating side produced under an average output power of 2.5 and 3.5 mW respectively on

gelatin-coated PET membrane. For 2.5 mW, the average pore size was 3.5 µm on the

PET membrane side and 1.5 µm on the gelatin coating side, while the average pore size

on both sides was 3.5 µm under 3.5 mW.

4.3 Micro-nozzle enhanced sandwich electroporation

4.3.1 Experimental

4.3.1.1 Reporter plasmids

Reporter plasmids pmaxGFP (3486 bp), encoding the new green fluorescet

protein from Pontellina sp., were purchased from Amaxa (now Lonza, Switzerland), and

Page 88: Fei Zhengzheng

-67-

gWiz SEAP (6569 kbp) from Aldevron (Fargo, ND). Figure 4.8 shows the map of

pmaxGFP, and the map of gWiz SEAP was shown in Figure 3.1(b).

Figure 4.8: Plasmid map of pmaxGFP, encoding the new green fluorescent protein from

Pontellina sp. (http://www.lonzabio.com/uploads/tx_mwaxmarketingmaterial/

amaxa_Newsletter_amaxa-news-03.pdf)

Page 89: Fei Zhengzheng

-68-

4.3.1.2 Culture of mouse embryonic stem cells

Mouse embryonic stem (ES) cells (CCE strain) were purchased from StemCell

Technologies (Vancouver, BC, Canada). All mES CCE cell culture reagents were

purchased from Invitrogen (Carlsbad, CA) unless otherwise specified.

Mouse ES cells were maintained in an undifferentiated state on 0.1% (w/v)

gelatin coated dishes in high glucose Dulbecco’s Modified Eagle’s Medium (DMEM

with 4500 mg D-glucose/L, StemCell Technologies, Catalog No. 36250) supplemented

with 15% ES-Cult fetal bovine serum (FBS, StemCell Technologies, Catalog No. 06952),

2 mM L-glutamine (Catalog No. 25030), 1 mM MEM sodium pyruvate (Catalog No.

11360), 1000 U/mL recombinant mouse leukemia inhibitory factor (rm LIF, Millipore,

Catalog No. LIF2010), 100 U/ml penicillin G + 10 µg/ml streptomycin (Catalog No.

15140), 0.1 mM MEM non-essential amino acids (NEAA, Catalog No. 11140), and 150

µM monothioglycerol (MTG, Sigma-Aldrich, Catalog No. M6145).

Mouse ES cells were cultured at 37°C with 5% CO2, and passaged every 3 days

by trypsinization. Mouse ES cells were typically used when reaching 50~70% confluent.

For single cell suspension, 0.25% trypsin with EDTA·4Na (Catalog No. 25200) was used

for harvesting mouse ES cells; for cell colony suspension, 0.05% trypsin with EDTA ·

4Na (Catalog No. 25300) was used.

Page 90: Fei Zhengzheng

-69-

4.3.1.3 Experimental set-up

The second generation of MSE system, including a multi-functional pulse

generator and a platform, is shown in Figure 4.9. The multi-functional pulse generator is

able to generate either square wave pulse or exponentially decay pulse. The technical

specifications of the multi-functional pulse generator are given in Table 4.1. The new

platform is able to handle three fluidic devices (Figure 3.2) in parallel.

Figure 4.9: The second generation of MSE system, including a multi-functional pulse

generator, and a platform, which is able to handle three fluidic devices in parallel.

(Designed and fabricated by Mr. Mr. Shi-Chiung Yu and Dr. Shengnian Wang, 2008)

Page 91: Fei Zhengzheng

-70-

Table 4.1: Technical specifications of the multi-functional pulse generator.

Pulse Type Voltage Capacitance

No. Pulses

(maximum)

Current

(maximum)

Exponentially

decay

10 ~ 300 V 100 ~ 1000 µF 10 20mA

Voltage Pulse length Pulse interval

No. Pulses

(maximum)

Current

(maximum)

Square wave 10 ~ 300 V 1 ~ 999 ms 1 ~ 999 ms 10 20mA

4.3.1.4 Electroporation procedure

Nucleofection by Amaxa Biosystem:

Mouse ES cells were pelleted via centrifugation and washed twice with

Dulbecco's phosphate-buffered saline (D-PBS, pH 7.4, Invitrogen, Catalog No. 14190).

One million ( 6101× ) cells were then resuspended in transfection solution from Mouse ES

Cell Nucleofector® Kit (Amaxa, Catalog No. VPH-1001) with 200 ng/µL plasmids and

transferred to the 2-mm gap nucleofection cuvette. Mouse ES cells were nucleoporated at

Program A-13, A-23, A-24, and A-30 according to manufactory’s suggestion, and A-23

was selected in this study (Appendix A).

Page 92: Fei Zhengzheng

-71-

Bulk electroporation by Bio-Rad Gene Pulser X-Cell system:

Mouse ES cells were pelleted via centrifugation and washed twice with D-PBS.

The pellet was resuspended in high glucose DMEM, and then transferred to the 1-mm or

2-mm electroporation cuvette (Bio-Rad). Two different cell concentrations were tested as

shown in Table 4.2. The electroporation parameters were optimized (Appendix B) and

the best electroporation condition is shown in Table 4.2.

Table 4.2: Comparison of nucleofection, conventional bulk electroporation by Bio-Rad

Gene Pulser XCell system, and membrane sandwich electroporation (MSE).

Method Nucleofection Bulk electroporation MSE

Pulse type Unknown Exponentially decay

Pulse number Unknown 1 1

Field Strength (V/cm) Unknown 500 150

Capacitance (µF) Unknown 500 500

DNA concentration (µg/mL) 200 200 50 5

Initial cell number 6101× 6101× 5101× 4101×

Cuvette 2-mm 2-mm 1-mm N/A

Total buffer volume (µL) 100 100 100 200

Page 93: Fei Zhengzheng

-72-

Micro-pore array enhanced MSE:

A gelatin-coated PET membrane with micro-pore arrays drilled by femtosecond

pulsed laser was used as the cell binding membrane. Briefly, the cell binding membrane

was mounted at the center reservoir of the fluidic device (Figure 3.2). A drop of

suspended cells was loaded onto the support membrane, and a vacuum of 334 ± kPa was

used to trap the cells on the support membrane. Another PET track-etched membrane

with an average pore size of 1 µm was placed over the immobilized cells. The cells were

sandwiched between two membranes sealed together (Figure 4.10(a)). The bottom

channel and inlet reservoirs were loaded with high glucose DMEM with 5 ng/µL DNA,

and the top channel and outlet reservoirs were then loaded with high glucose DMEM. A

low DC voltage of 10 V was applied to the system for 5 seconds; DNA molecules were

migrated from the cathode to the MSE disk (Figure 4.10(b)) and concentrated in the

micro-pores of the cell binding membrane. Finally, the cells were pulsed with a single

exponentially decay pulse at 150 V/cm and 500 µF (Table 4.2), and the MSE disk was

transferred to a 48-well plate with 250 µL culture media in each well.

Page 94: Fei Zhengzheng

-73-

Figure 4.10: (a) Schematic of MSE disk set-up; (b) Schematic of DNA migration path in

the MSE device. DNA molecules migrate from cathode to anode.

4.3.1.5 Assay for transfection efficiency and cell proliferation

The transfection efficiency of pmaxGFP in mouse ES cells was qualified by the

percentage of the cells with green fluorescence, and the transfection efficiency of pSEAP

was quantified by the activity level of secreted alkaline phosphatase by the transfected

cells, determined by a colorimetric assay based on the hydrolysis of p-nitrophenyl

Page 95: Fei Zhengzheng

-74-

phosphate (pNPP). The experimental procedures were the same as described in Sections

3.2.6 and 3.2.7.

CellTiter 96® non-radioactive cell proliferation assay (MTS) was used to measure

the cell viability at 24-hour post-electroporation. MTS reagent was diluted in culture

media at 1:10, and the work solution was then added to the culture wells. The absorbance

at 570 nm (OD570) was recorded after 2 hours of incubation at 37°C with 5% CO2.

Background absorbance can be corrected by including negative control wells on each

plate to measure the absorbance from culture medium in the absence of cells. A set of

positive control wells containing untreated cells performed in sister cultures was settled

as 100% cell viability. The cell viability of each sample was calculated as

%100570570

570570×

−=

backgroundcell

backgroundsmaple

ODOD

ODODViabilityCell (4.1)

4.3.1.6 Statistical analysis

Data analyses were performed using Student’s t-test and are expressed as

arithmetic mean + s.d.; t-test values of *P < 0.05, **P < 0.01, ***P < 0.005 were

considered statistically significant.

Page 96: Fei Zhengzheng

-75-

4.3.2 System optimization

4.3.4.1 Converging micro-nozzle vs straight micro-channel

A gelatin-coated PET membrane with micro-pore arrays drilled by femtosecond

pulsed laser was used as the cell binding membrane. As mentioned in Section 4.2.4, two

different micro-pore shapes were generated: a converging micro-nozzle and a straight

channel. To determine the effect of micro-pore shape on the MSE performance, mouse

ES cells were transfected by the MSE method with the converging micro-nozzle arrays

and the straight micro-channel arrays. The converging micro-nozzles were obtained at the

output laser power of 2.5 mW (Average pore size: 3.5 µm on the PET membrane side and

1.5 µm on the gelatin coating side), and the straight micro-channels was fabricated under

3.5 mW (Average pore size: 3.5 µm for both sides). Plasmids gWiz SEAP were used as

reporter genes, and transfection efficiency and cell viability of pSEAP transfection were

evaluated 24 hours after electroporation. As shown in Figure 4.11, SEAP expression

using the micro-nozzles almost doubled over that with micro-channels, mainly because

the electric field is more concentrated at the small-end of the micro-nozzle, resulting in

better localized electroporation. Additionally, pSEAP are relative large plasmids, and

they may experienced strong stretching along the axis direction in the converging

direction of the micro-nozzle, (Hu et al., 2009; Wang et al., 2008), leading to easier

delivery through the cell membrane.

Page 97: Fei Zhengzheng

-76-

0

15

30

45

60

75

To

tal S

EA

P A

ctivity m

U/

10,0

00 I

nitia

l se

ed

ing

ce

lls

Micro-channel Micro-nozzle

Figure 4.11: Effect of different pore shapes, micro-channel () and micro-nozzle (), on

mouse ES cell transfection by MSE. The bars indicate total activity of SEAP expression

24 hours after MSE under the optimized electrical field (Appendix B).

4.3.4.2 Effect of porosity and micro-pore shape of top membrane

With initial ten thousands ( 4101× ) mouse ES cells trapped on a 100100 × micro-

nozzle array, five different commercially available membranes were investigated as the

top membrane in MSE to determine the effect of porosity on transfection efficiency and

cell viability. Table 4.3 shows the properties of these membranes (Cases 1 to 5).

Page 98: Fei Zhengzheng

-77-

Table 4.3: Top membranes with different pore size, pore density, and pore shape

Case No.

Pore size

(µm)

Pore density

(pores /cm2)

Porosity

(%)

Pores per cell Pore shape

1 a 0.4 4 x 106 0.5 4 ~ 5

2 a 3 2 x 106 14.1 2 ~ 3

3 a 0.4 1.6 x 106 0.2 2 ~ 3

4 b 1 1.6 x 106 1.3 2 ~ 3

5 b 3 8 x 105 5.7 1 ~ 2

Straight

micro-channel

6 Small end: 1;

Large end: 3

1 x 106 1.9 1 ~ 2 Converging

micro-nozzle

a: Polyester (Corning, Lowell, MA);

b: PET (BD Biosciences, San Jose, CA).

pSEAP transfection experiments were carried out and the transfection results are

shown in Figure 4.12.Since the micro-pore shape is a straight channel in the five

commercially available membranes (Cases 1 to 5), there are two main determinants of the

porosity: pore size and pore density. In Cases 2 to 4, the membranes have the similar pore

density but different average pore size. When average pore size increased from 400 nm in

Case 3 to 1 µm in Case 4, the SEAP expression largely increased while the cell viability

slightly decreased. Although the average pore size in Case 2 is 3 µm, three times larger

Page 99: Fei Zhengzheng

-78-

than 1 µm in Case 4, there is no obviously difference observed in the transfection results

in both cases. For Cases with the same pore size but different pore density (e.g. Cases 1

and 3, Cases 2 and 5), either the SEAP expression or the cell viability is similar. Based on

the value of pore size and pore density, the porosity was calculated and the value is given

in Table 4.3. The transfection results can be divided into two groups: Group A includes

Cases 2, 4, 5 with porosity larger than 1% and Group B includes Cases 1 and 3 with less

than 0.5% porosity. The total SEAP activity in Group A is about one third higher than

that in Group B, while losing 10 ~ 15 % cell viability.

Transfection experiments of mouse ES cells using top membrane with either the

micro-nozzle array or the straight micro-channel array were also carried out to determine

the effect of pore shape of top membrane on the MSE design. Due to the limitation of

femtosecond laser micro-fabrication, the minimum distance between two converging

micro-nozzles (center to center) was 10 µm. It means the maximum pore density of the

micro-nozzle array is about 6101× pores/cm2, and thus there are 1 ~ 2 pores per cell. To

minimize the effect of porosity, Case 5 in Table 4.3, was used for the corresponding

straight micro-channel comparison. No significant difference of the transfection results

between two cases was observed (Figure 4.12). The results showed evidence that for the

top membrane the porosity effect is more important than micro-pore shape effect on cell

transfection.

Page 100: Fei Zhengzheng

-79-

Figure 4.12: Comparison of top membrane with different micro-pore size and micro-pore

density in MSE: (a) transfection efficiency and (b) cell viability of mouse ES cells 24

hours after MSE.

Page 101: Fei Zhengzheng

-80-

4.3.4.3 Effect of top membrane location

Mouse ES cell transfection by pSEAP with different top membrane positions

(Table 4.4) was also investigated. As shown in Figure 4.13, SEAP expression of Cases I

to III decreased with increasing the distance between two membranes. The cell viability

of the three cases with a top membrane is better than the case without the top membrane

(Case IV). It means the existence of the top membrane helps to protect the cells during

the electroporation process.

It is noticed that besides slightly deformation due to the gravity effect of cell

itself, there’s a compression effect of the top membrane on cell when the distance

between two membranes is less than 5 µm. As observed from an inverted microscope, the

average diameter of the cells is 15 µm in the X-Y cross-section area in Case I, and 12 µm

at the remaining three cases. It has been reported that cell viability decreased when cells

are compressed to ~ 50 % of their original diameter. (Takamatsu and Rubinsky, 1999)

This explained why the cell viability of Case II was better than that of Case I.

Page 102: Fei Zhengzheng

-81-

Table 4.4: Top membranes with different distance to cell binding membrane.

Cell dimension

Case No. With top

membrane

With Spacer Spacer thickness

(µm) Major axis

(X-Y)

Minor axis

(X-Z)

I Yes No N/A 15 4.5

II Yes Yes 10~12 12 7

III Yes Yes 16~20 12 7

IV No No N/A 12 7

Note: A PET membrane (BD Biosciences, San Jose, CA) was used as the top membrane

with an average pore size of 3 µm. The spacers were spin-coated PCL membranes with a

3 mm diameter hole in the middle.

Page 103: Fei Zhengzheng

-82-

Figure 4.13: Effect of top membrane location in MSE on cell transfection: (a)

Transfection efficiency and (b) cell viability of mouse ES cells 24 hours after

electroporation.

Page 104: Fei Zhengzheng

-83-

4.3.3 Comparison of MSE with bulk electroporation and nucleofection

pSEAP transfection of mouse ES cells by micro-nozzle array enhanced MSE,

bulk electroporation using Bio-Rad Gene Pulser, and nucleofection was compared

(Figure 4.14). Although the transfection result of MSE was still not as good as that of

nucleofection using unknown cell-specific reagents, the amount of SEAP expression

mediated by the MSE method was higher than that in bulk electroporation by Bio-Rad

Gene Pulser XCell system using the same electroporation buffer. The main reason is that

the MSE method is able to pre-concentrate the genes near the cell surface and to focus the

electric field strength around the micro-pores for better gene transport during

electroporation. Also, the electric field strength applied in MSE was 150 V/cm, much less

than 500 V/cm used in bulk electroporation. Therefore, mouse ES cells experienced an

average survival rate of ~ 75 % in MSE with an initial cell number of 4101× , similar to

bulk electroporation with 6101× initial cells and much higher than bulk electroporation

with 5101× initial cells.

MSE provided the high transgene efficiency and cell viability, because the area of

the cell membrane with the highest transmembrane potential is exactly the same location

that negatively charged DNA molecules could permeate into the cells (Golzio et al.,

2002b). As comparison, a much larger area of cell surface experiences high electric field

strength in the bulk electroporation, and DNA molecules can’t transport across most of

the affected cell surface. Since only a small area of cell membrane was affected during

Page 105: Fei Zhengzheng

-84-

MSE and each cell experienced a similar electric field with the micro-array, a very small

number of cells could be uniformly transfected.

(Continued)

Figure 4.14: Comparison of mouse ES cell transfection by micro-nozzle array enhanced

MSE, bulk electroporation by Bio-Rad Gene Pulser, and nucleofection. (a) Transfection

efficiency and (b) cell viability 24 hours after electroporation. From left to right, bulk

electroporation with initial input cell number of 6101× ( ) and 5101× ( ); micro-nozzle

enhanced MSE with initial input cell number of 4101× ( ); and nucleofection with initial

input cell number of 6101× ( ).

Page 106: Fei Zhengzheng

-85-

Figure 4.14 continued

A hundred of mouse ES cells trapped on a 1010 × micro-nozzle array were

transfected by pmaxGFP. Mouse ES cells were analyzed 24 hours after electroporation

by phase contrast and fluorescence microscopy using a GFP filter. The best result is

given in Figure 4.15. Almost all the cells were transfected, and remained alive. This

indicated the potential of micro-nozzle array enhanced MSE for hard-to-harvest cells.

However, it is very difficult to get such a perfect result, because it is hard to handle such

a small amount of cells.

Page 107: Fei Zhengzheng

-86-

Figure 4.15: GFP transfection of mouse ES cell by micro-nozzle enhanced MSE. A

hundred of cells were trapped on a 1010 × micro-nozzle array, and (a) phase contrast and

(b) fluorescent images were taken 24 hours after electroporation.

4.4 Simulation of transmembrane potential distribution

Although there are still arguments on the creation and evolution of nano-pores on

cell membrane during the electroporation, it is well established that the size of nano-pore

Page 108: Fei Zhengzheng

-87-

is directly related to the value of transmembrane potential. (Weaver and Chizmadzhev,

1996) When the transmembrane potential is lower than the threshold value, there is no

formation of nano-pores in cell membrane. In a successful gene electrotransfection

process, nano-pores should be large enough for gene to diffuse through. A higher

transmembrane potential leads to larger nano-pores formed on the cell membrane, thus

the more DNA molecules can be delivered to achieve the higher transgene efficiency.

(Krassowska and Filev, 2007) Therefore, we can propose the possible reason for the

results in Sections 4.3.2 and 4.3.3 through the calculated distribution of transmembrane

potential for bulk electroporation, localized electroporation, and different set-ups of

MSE.

4.4.1 Three-layer model

Since the distribution of transmembrane potential is extremely important to the

cell transfection, the consideration of cell membrane should be indispensable in the

numerical simulation of cell electroporation. In this study, we simplify the cell with a

three-layer (or single-shell) model and solved the equations in the cytoplasm, the external

medium, and the cell membrane, respectively.

The Laplace equation of the three-layer model (Stewart et al., 2005; Zudans et al.,

2007) was used to calculate the distribution of electric field and transmembrane potential

(the potential difference across the cell membrane) of a single cell in bulk electroporation

and MSE:

Page 109: Fei Zhengzheng

-88-

0)( =∇⋅∇ ii φσ (4.2)

0)( =∇⋅∇ ee φσ (4.3)

0)( =∇⋅∇ mm φσ (4.4)

where:

φ : electric potential,

σ : electric conductivity,

Subscript i: cytoplasm,

Subscript e: external medium,

Subscript m: cell membrane.

Correspondingly, there are two boundary conditions (B.C.s) for two interfaces,

i.e., the external interface between the external medium and the cell membrane, and the

inner interface between the cytoplasm and the cell membrane.

nn

mm

ee

∂=

∂ φσ

φσ

(4.5)

me φφ =

B.C.s on external interface

(4.6)

nn

mm

ii

∂=

∂ φσ

φσ

(4.7)

mi φφ =

B.C.s on internal interface

(4.8)

In our experiment, low-conductive PET membrane was used, and thus the electric

conductivity of the membrane can be neglected. Therefore, all the solid walls were

Page 110: Fei Zhengzheng

-89-

treated as insulators. Since no electric field can penetrate into the walls, different

potentials are imposed at inlet and outlet in order to create a voltage drop. Once the

electric potentials were calculated, the electric field E could be known at different layers:

φ−∇=E (4.9)

The transmembrane potential was obtained as

)()( inlmexlmm SSV φφ −=∆ (4.10)

where

S: surface of cell membrane;

Subscript exi: external;

Subscript inl: internal.

Compared with the two-layer model, the three-layer model is more complicated

and very dense mesh needs to be generated near the cell membrane, but it can be easily

applied to calculate the transmembrane potential around electroporated cells with nano-

pores formed on the cell membrane.

4.4.2 Two-dimensional (2-D) simulation process

A two-dimensional (2-D) simulation with the three-layer model was carried out

using commercial FEM software, COMSOL (Mathworks, Natick, MA), and parameters

used in the simulation are given in Table 4.5.

Page 111: Fei Zhengzheng

-90-

Table 4.5: Parameters of the three-layer model (Kotnik et al., 1997)

Symbol Value Definition

d nm 5 Cell membrane thickness

mS /2.0 Electric conductivity of external medium

mS /2.0 Electric conductivity of cytoplasm

mσ mS /105 7−× Electric conductivity of cell membrane

For a single cell in bulk electroporation, Zudans et al. (2007) have shown that 2-D

simulation results agree well with the experiments, and thus the accuracy of the 2-D

numerical simulation has been verified. The analytical solution of transmembrane

potential for a 2-D spherical cell was obtained (Appendix B):

)cos(θfERVm =∆ (4.11)

where

E: external electric field strength, V/cm;

R: radius of the cell, µm;

θ : angle between Eext and the point on the cell membrane;

f: shape factor, f = 2.

Mouse ES cells with an average diameter of 10 µm were used in the experiments.

In bulk electroporation, cells are suspended in buffer solution, and thus are simplified as a

Page 112: Fei Zhengzheng

-91-

sphere shape with the diameter of 10 µm. In localized electroporation and MSE, the cell

is sit down on a flat surface, and slightly deformed as a result of the gravity. As observed

from inverted microscope, the average diameter of the cells is 12 µm in the X-Y cross-

section area, and thus we simplify its shape as an ellipse with the major axis 12 µm and

the minor axis 7 µm in localized electroporation. It is noticed that besides slightly

deformation due to the gravity effect of cell itself, there’s compression effect of the top

membrane on cell when the membrane distance is less than 5 µm. From the top view, the

average diameter of the cells is 15 µm, and thus we simplify its shape as an ellipse with

the major axis 15 µm and the minor axis 4.5 µm in MSE. 2-D simulation of

transmembrane potential for an oval cell was preformed using the finite element methods,

as used in Agawal et al. (2007). The mesh of finite elements was generated, and the

electric potential inside and outside the cell was then computed by solving Eq. (4.10).

4.4.3 Simulation results

4.4.3.1 Effect of cell shape

The transmembrane potential distribution of four cases with different top

membrane location (Table 4.4) was calculated and the results are presented in Figure

4.16. The cell shape is simplify as an ellipse with the major axis 15 µm and the minor

axis 4.5 µm in Case I, while in Cases II to IV the major axis is 12 µm and the minor axis

7 is µm (Figure 4.16(a)). The perk value of the transmembrane potential at Case I is

higher than the other three case as a result of the shape effect, which explained the

Page 113: Fei Zhengzheng

-92-

transfection results in Figure 4.13 that better transfection efficiency in Case I than the

other three cases.

a

(Continued)

Figure 4.16: Simulation comparison of top membranes with different pore size, pore

density, and pore shape. (a) Schematic diagram of Cases I to IV, from left to right, with

electric field lines across/around a single cell; (b) calculated transmembrane potential

distribution. θ is the angle around cell surface.

Page 114: Fei Zhengzheng

-93-

Figure 4.16 continued

b

4.4.3.2 Effect of porosity and pore shape of top membrane

Figure 4.17 shows the calculated transmembrane potential distributions of

different top membranes as described in Table 4.3. According to the calculation in

Figure 4.17, the membrane with lager pores at a certain pore density has higher

transmembrane potential; for a fixed pore size, the membrane with a higher pore density

has a higher transmembrane potential in the model. The transmembrane potential at θ = π

Page 115: Fei Zhengzheng

-94-

increased with the increase of the porosity in the simulation results (Figure 4.17). The

transfection results in Figure 4.12 are divided into two groups, indicating that there is a

threshold transmembrane potential to permeate cell membrane. Since Cases 2, 4 and 5

have better transfection than Cases 1 and 3, the critical transmembrane potential is around

0.4 V according to the calculation in Figure 4.17.

Figure 4.17: Simulation comparison of transmembrane potential distribution of top

membranes with different pore size, pore density, and pore shape.

Page 116: Fei Zhengzheng

-95-

4.4.3.3 Converging micro-nozzle vs straight micro-channel

For simplicity, we assume micro-channels on the top membrane are uniformly

distributed in the simulation. The thickness of top membrane is 10 µm, and the size of a

micro-channel is ~1 µm. According to the pore density of 6106.1 × pores/cm2 given in the

manufactory’s instruction, the center-to-center distance between two micro-channels is

7.7 µm. For the cell binding membrane, the thickness is 12 µm and the center-to-center

distance between two micro-nozzles or micro-channels is 20 µm. The simulation results

are given in Figure 4.18.

Figure 4.18(a) shows that the electric field is concentrated around the micro-hole,

and more electric field lines are forced to penetrate through the cell at the small-end of

the micro-nozzle. Correspondingly, the transmembrane potential at the small-end of the

micro-nozzle is much higher than that of the straight channel as shown in Figure 4.18(b).

This explains why our micro-nozzle MSE can provide better delivery efficiency than the

micro-channel MSE, as shown in Figure 4.11.

Page 117: Fei Zhengzheng

-96-

Figure 4.18: Simulation comparison between support membranes with micro-nozzles

and micro-channels (a) electric potential distribution and electric field lines across/around

a single cell near a micro-nozzle (left) and micro-channel (right); (b) calculated

transmembrane potential distribution. θ is the angle around cell surface.

Page 118: Fei Zhengzheng

-97-

4.4.3.4 Comparison of micro-nozzle enhanced sandwich electroporation with

bulk electroporation

Micro-nozzle enhanced MSE was also compared with bulk electroporation. The

transgene efficiency doesn’t depend on the total area of cell membrane with openings, but

on the effective area facing the cathode. (Golzio et al., 2002b) Although a much larger

area of cell surface experiences a high electric field strength in the bulk electroporation

(Figure 4.18(b)), the area of the cell membrane with the highest transmembrane potential

in MSE is at exactly the same location that negatively charged DNA molecules could

permeate into the cells. Furthermore, since only a small area of cell membrane was

affected in MSE and each cell experienced a similar electric field with the micro-array, a

very small number of cells could be uniformly transfected. In addition, the size of nano-

pores on the cell membrane surface is smaller, and thus DNA molecules can’t transport

across most of the affected cell surface.

4.5 Conclusion

In this study, we demonstrated the use of a femtosecond laser fabricated micro-

nozzle arrays on a gelatin-coated PET membrane for membrane sandwich electroporation

(MSE). Using micro-nozzle array enhanced MSE, we observed high and uniform gene

transfection, and good cell viability of mouse ES cells compare to the bulk

electroporation. The ability to treat a small number of cells (i.e. a hundred) offers great

potential to work with hard-to-harvest patient cells for pharmaceutical kinetic studies.

Page 119: Fei Zhengzheng

-98-

Numerical calculation of transmembrane potential qualitatively explains the

observed differences among different cases of MSE and bulk electroporation. Since

there’s a reasonably well correlation between transfection results and transmembrane

potential calculations, the simulation process with the threshold experiments can be used

to predict the transfection results, and thus largely reduced the trial-and-error window

size.

Page 120: Fei Zhengzheng

-99-

CHAPTER 5: NANOFIBER BASED MEMBRANE SANDWICH

ELECTROPORATION

5.1 Introduction

Over the last decade, the clinical applications of genetically modified embryonic

stem (ES) cells have increased considerably. For clinical trails, it is essential to have high

transfection efficiency as well as high cell viability. It has been showed that ES cells,

especially human ES cells, grow better when forming highly compact colonies. (Amit et

al., 2000) However, commercially available electroporation systems, especially the

leading Nucleofection, are based on single cell suspension. As a result, it is hard to treat

cell colonies using conventional bulk electroporation, especially if repeated transfection

is required. In this chapter, the membrane sandwich electroporation (MSE) method is

integrated into cell culture such that ES cells can form desirable colonies, be transfected,

and further cultured on a same substrate before and after MSE.

Electrospinning is a simple and versatile technique that can produce a porous

scaffold comprised randomly submicron-diameter polymeric fibers (or nanofibers) (Chew

et al., 2006). Electrospun nanofiber scaffolds characterized by a high surface area for cell

attachment and three-dimensional (3-D) microenvironment for cell-cell interaction

Page 121: Fei Zhengzheng

-100-

provide stronger topographic cues by mimicking the filamentary extracellular matrics,

and enhance cell adhesion and proliferation compared to micro-porous polymer

membrane. (Wnek et al., 2003; Yashimoto et al., 2003) Poly(ε-caprolactone) (PCL), a

Food and Drug Administration (FDA) approved biocompatible polymer, has been widely

used as a tissue engineering scaffold to support cellular in-growth and proliferation for

the generation of biological tissues. However, it was reported that cell adhesion,

migration, proliferation, and differentiation were reduced due to its poor hydrophilicity.

(Kohl et al., 2005; Ingber, 2005) As a natural biopolymer derived from collagen by

controlled hydrolysis, gelatin is blended with PCL to obtain a scaffold with good

biocompatibility and improved mechanical, physical and chemical properties. (Chong et

al., 2007; Zhang et al., 2005) Additionally, electrospun PCL/gelatin nanofiber scaffolds

may serve as safe substitutes to Matrigel, a gelatinous protein mixture secreted by mouse

tumor cells. In a recent study, Gauthaman et al. (2008) evaluated the influence of

electrospun nanofibrous (PCL/collagen and PCL/gelatin) scaffolds for human ES cell

proliferation. Increased colony-formation, self-renewal properties, undifferentiation and

retention of stemness were observed in their study.

A combination of MSE and electrospun PCL/gelatin nanofiber scaffolds allows

gene transfection at different time points during cell colony formation without repeating

cell trypsinization, and thus may enhance the cell viability. Since the size of ES cell

colonies can influence pluripotent cell differentiation trajectories (Bauwens et al., 2008;

Page 122: Fei Zhengzheng

-101-

Tsuruma et al., 2008), cell colony formation may be further controlled with micro-

patterned spacer bonded on nanofiber surface.

5.2 Materials and methods

5.2.1 Cell culture

NIH 3T3 fibroblasts and mouse embryonic stem (mES) CCE cells were cultured

and prepared according to the procedure described in Sections 3.2.2 and 4.3.1.2.

5.2.2 Fabrication and characterization of nanofiber scaffolds with micro-well

spacers

5.2.2.1 Preparation of electrospun poly (ε-caprolactone) (PCL) /gelatin nanofiber

scaffolds

A 6.7 % (w/v) solution of poly (ε-caprolactone) (PCL; Sigma-Aldrich, Mw =

65,000) in 1,1,1,3,3,3-Hexafluoro-2-propanol (HFIP, Sigma-Aldrich) and a 6.7 % (w/v)

solution of porcine gelatin (Sigma-Aldrich, Catalog. No. G2500) in HFIP was prepared

by stirring overnight at room temperature, separately. The two solutions were mixed

together at a 1:1 weight ratio, and stirred for 2 minutes. The mixture was then placed in a

60 mL syringe with a 20 gauge blunt tip needle, and electrospun using a high voltage DC

power supply (Glassman High Voltage, High Bridge, NJ) set to -26 kV voltage and 15

Page 123: Fei Zhengzheng

-102-

mL/h flow rate. (Johnson et al., 2008, Lim et al., 2009) The support membrane was

placed on an aluminum foil with a 20 cm distance between the tip and the collector. The

electrospun sheet with nanofibers was deposited onto the support membrane, and the

thickness of the sheet was controlled by the electrospinning time. The electrospun sheet

was then placed in a vacuum oven overnight to ensure removal of residual HFIP. Figure

5.1 shows the electrospining process.

Figure 5.1: Schematic diagram of fiber formation by electrospining process where a

droplet of a polymer solution is elongated by a high electrical field.

(http://nano.mtu.edu/Electrospinning_start.html)

Page 124: Fei Zhengzheng

-103-

5.2.2.2 Fabrication of PCL /gelatin nanofiber scaffolds with polystyrene (PS)

micro-well arrays

An 8 ~ 10 µm thick polystyrene (PS) (Sigma-Aldrich, melt flow index 4.0) sheet

with arrays of 100 or 300 µm diameter micro-wells was fabricated via soft lithography

(Figure 5.2). Standard photolithography using negative tone photoresists, NANOTM SU-8

(Microchem Corp., Newton, MA), was used to achieve an 8 ~ 10 µm thick master mold

with micro-well arrays. The fabrication parameters, including spin speed, exposure dose,

post-exposure bake times, and develop time, were set up according to the manufacturer’s

suggestion. The master mold was then used to create poly(dimethyl siloxane) (PDMS)

stamp with micro-pillar arrays. The PDMS stamp was spin-coated with 15 ~ 20 %

polystyrene (PS) solution in anisole under spin-speed of 3,000 rpm for 1 minute. The

PDMS molds were then placed on a hotplate at 100°C for 5 minutes to drive off the

residual solvent and anneal the PS sheet. After cooling, the PS sheet left in-between the

pillars was manually peeled off. The PS sheet with micro-well arrays was thermally

bonded to PCL/gelatin nanofibers with support membrane at 80 ~ 90°C.

Page 125: Fei Zhengzheng

-104-

Figure 5.2: Schematic of fabricating electrospun nanofiber scaffold with polystyrene

(PS) micro-well arrays. (1) PDMS stamp with micro-pillar arrays; (2) Drop-cast PS

solution; (3) PS solution is spin-coated and it de-wets on the surface of the PDMS stamp;

(4) PS in-between the features is removed; (5) PS micro-well arrays were bonded to

electrospun nanofiber scaffold by thermal bonding. (Gallego et al., In preparation)

5.2.2.3 Structure characterization by scanning electron microscopy (SEM)

SEM was used to observe the structure of PCL/gelatin nanofiber scaffolds. The

samples were coated with 15 nm of osmium (model OPC-80T, SPI Supplies) prior to

viewing in a scanning electron microscope (Sirion FEI). The use of osmium plasma

deposition instead of gold or gold-palladium sputtering eliminated concerns regarding

PCL melting and allowed for higher resolution imaging of the fiber surface. SEM images

Page 126: Fei Zhengzheng

-105-

of the surface of randomly distributed PCL/gelatin nanofibers and nanofiber scaffolds

with 300 µm PS micro-wells are shown in Figure 5.3.

Figure 5.3: SEM image of PCL/gelatin nanofiber scaffolds with 300 µm PS micro-wells.

5.2.3 Experimental set-up for nanofiber based MSE

The experimental set-up for electrospun nanofiber based MSE is shown in Figure

5.4. The cell binding substrate consists of a support membrane and a thin sheet of

electrospun nanofiber scaffold. Membranes with different properties were tested as the

Page 127: Fei Zhengzheng

-106-

support membrane for nanofiber scaffolds. The electrospun fibers are randomly

distributed poly-ε-caprolactone (PCL) / gelatin blends (weight ratio: 50:50, mean fiber

diameter: 300-400 nm, thickness: 5-20 µm), directly bonded to 8 ~ 10 µm thick

polystyrene (PS) spacer with 100-500 µm micro-well arrays.

Figure 5.4: Schematic drawing of electrospun nanofiber based MSE.

5.2.4 Electric Resistance Measurements

The electric resistance of a sample was determined by the ratio of voltage to

current at room temperature, in accordance with Ohm's law:

I

VR = (5.1)

where:

Page 128: Fei Zhengzheng

-107-

R: electric resistance, KΩ

V: voltage, V

I: current, mA.

The sample was placed in the central reservoir, and the inlet and outlet channels

and reservoirs were filled with high glucose DMEM for mouse ES cells. Two wire

electrodes were placed in the inlet and outlet reservoirs, and connected to PowerPac

Basic Power Supply (Bio-Rad, Catalog No. 164-5050). A DC voltage that increased from

a 30 to 60 V with increment of 10V was applied to the system, and the value of the

current at each specific voltage was recorded. The graph was plotted with voltage as Y-

axis and current as X-axis, and a linear curve was fitted to the data points. The calculated

slope of the line is numerically equal to the electric resistance of the sample.

5.2.5 Electroporation procedure

5.2.5.1 Single cell electroporation

For NIH 3T3 fibroblasts, electroporation followed the same procedure described

in Section 3.2.4.

For mES cells, single cell suspension was obtained using 0.25% trypsin with

EDTA·4Na. Bulk electroporation and nucleofection followed the same procedure given

in Section 4.3.1.4, and nanofiber based MSE followed the procedure described in Section

3.2.5.2.

Page 129: Fei Zhengzheng

-108-

5.2.5.2 Cell colony electroporation

Mouse ES cell colony suspension was obtained using 0.05% trypsin with

EDTA·4Na (Catalog No. 25300).

Bulk electroporation and nucleofection followed the procedure given in Section

4.3.1.4.

For nanofiber based MSE, cells were seeded in the micro-wells on the cell

binding substrate and grow 1 to 2 days to form colonies. A PET track-etched membrane

with an average pore size of 3 µm was placed over the colonies, and the cell colonies

were sandwiched between nanofiber scaffold and PET membrane, which are sealed

together. The whole MSE disk was then sealed at the center reservoir of the MSE device

by sealing tape. The bottom channel and reservoirs (cathode side) were loaded with high

glucose DMEM, and the top channel and reservoirs were then loaded with high glucose

DMEM with 100 ng/µL DNA molecules. A low DC electric field of 3 V/cm was applied

to the system for 5 seconds to concentrate DNA molecules in the micro-pores, and the

cells were then electroporated by a single exponentially decay pulse at 150 V/cm and 500

µF.

Page 130: Fei Zhengzheng

-109-

5.2.6 Assays for transfection efficiency and cell proliferation

As described in Sections 3.2.6 and 3.2.7, the transfection efficiency of pGFP in

NIH 3T3 fibroblasts and mouse ES cells was qualified by the percentage of cells with

green fluorescence, while the transfection efficiency of pSEAP was quantified by the

activity level of secreted alkaline phosphatase by the transfected cells, determined by a

colorimetric assay based on the hydrolysis of p-nitrophenyl phosphate (pNPP).

AlamarBlue® cell proliferation assay (Molecular Probes, Catalog No. DAL-1100)

was used to quantify the cell viability at 24 and 48 hours post-electroporation. The

alamarBlue® dye was diluted in the culture media at 1:10, and the work solution was

then added to the culture wells. The plate was shaken gently and incubated for 3 hours at

37°C with 5% CO2. The fluorescence detection was performed using a fluorescence

excitation wavelength of 570 nm, and the reading at emission wavelength of 595 nm was

recorded. (Nociari et al., 1998) Fluorescence background can be corrected by including

negative control wells on each plate to measure the fluorescence from culture medium in

the absence of cells. A set of positive control wells containing untreated cells performed

in sister cultures was settled as 100% cell viability. The cell viability of each sample was

calculated as

%100595595

595595×

−=

backgroundcell

backgroundsmaple

ODOD

ODODViabilityCell (5.2)

Page 131: Fei Zhengzheng

-110-

5.2.7 Cell morphology characterization by confocal microscopy

Cell morphology on PCL/gelatin nanofibers was characterized by laser scanning

confocal microscopy (Zeiss LSM 510 Meta, Zeiss, Germany). Cells were fixed directly

with 70 % ethanol solution for 1 hour at room temperature, and labeled with the red-

fluorescent nucleic acid stain, propidium iodide (PI; Molecular Probe, Catalog No.

P3566). After incubated with 2 µg/ml PI for 15 minutes at 37°C and washed twice with

D-PBS, cells were then visualized by employing the excitation filter, 543 nm.

5.3 Optimization of nanofiber based membrane sandwich electroporation

5.3.1 Effect of support membrane

Preliminary experiments were carried out using electrospun nanofiber scaffolds

without any support membrane as a cell-binding substrate, and the results showed that it

is difficult to transfer the nanofiber scaffolds to membrane sandwich electroporation

(MSE) platform after overnight cell culture. To make the cell-binding substrate stronger,

a micro- / nano-porous membrane was used as a support membrane.

A track-etched polyethylene terephthalate (PET) membrane (average pore size: 3

µm) with a layer of PCL/ gelatin nanofibers was tested. Although the cell viability

increased, the transfection efficiency decreased compared to the transfection results using

PET membrane only (Figure 5.5(a,b)). Since the electric resistance of the cell-binding

substrate may increase by adding a layer of low-conductance nanofibers, the resistance

Page 132: Fei Zhengzheng

-111-

measurement using PET membrane with and without nanofibers was carried out and the

data of voltage against current are plotted as shown in Figure 5.5(c). As the resistance is

determined by the gradient of the straight fitting line, a steeper curve corresponds to

increased resistance. By adding a layer of nanofiber scafflolds, the resistance value

almost doubled over that using the PET membrane only.

(Continued)

Figure 5.5: Comparison of different cell-binding substrates used for membrane

sandwich electroporation. The transfection efficiency (a) and cell viability (b) of mouse

embryonic stem cells, and the resistance (R) of MSE disk (c) were presented using PET

membrane only ( ), aluminum oxide membrane only ( ), PET membrane with

nanofibers ( ), and aluminum oxide membrane with nanofibers ( ).

Page 133: Fei Zhengzheng

-112-

Figure 5.5 continued

To reduce the resistance of the MSE disk, AnodiscTM aluminum oxide membrane,

was tested as a support membrane. The properties of the membrane from the

manufactory’s website are given in Table 5.1. Without nanofiber scaffolds, the resistance

Page 134: Fei Zhengzheng

-113-

of an aluminum oxide (Al2O3) membrane is similar to that of a PET membrane (Figure

5.5(c)); however, cell viability using the Al2O3 membrane significantly decreased

(Figure 5.5(b)). With a nanofiber scaffold, the resistance of the MSE disk using the

Al2O3 membrane as the support membrane is much lower than that of the PET

membrane; and good transfection results (Figure 5.5(a, b)) were achieved by using the

nanofiber scaffold supported with the Al2O3 membrane. Besides the material difference,

the Al2O3 membrane has more than 4 times higher porosity than the PET membrane.

Table 5.1: Properties of three different types of membranes used as support membrane.

Material

Pore size

(µm)

Porosity

(%)

Thickness

(µm)

Note

PET 3 6 10 BD Biosciences,

San Jose, CA, USA

Aluminum oxide 0.2 25 60 Anodisc™, Waltman/GE,

Piscataway, NJ, USA

Polycarbonate 0.2 20 20 Isopore™,

Millipore, USA

Page 135: Fei Zhengzheng

-114-

To understand why Al2O3 membrane works better, another polymeric membrane,

polycarbonate (PC) membrane (Table 5.1) with a similar porosity as the Al2O3

membrane were tested. There was not any obvious difference observed in the electric

resistance measurements and transfection resultus between PC and PET membrane with

nanofiber scaffolds. Therefore, the difference in electric properties is more important than

the porosity difference. The most possible reason is that Al2O3 membrane became

electrically conductive under the external electric field, and consequencely reduced the

electric resistance of nanofiber scaffold covered support membrane.

5.3.2 Effect of nanofiber thickness

As shown in Figure 5.5, the resistance of the cell-binding substrate was increased

by adding PCL/gelatin nanofiber scaffold. As the PCL/gelatin nanofiber scaffold is a

non-conductive polymeric/protein matrix, the resistance of the cell-binding substrate

increased with increasing the thickness of the nanofiber scaffold layer, which was

controlled by the electrospinning time (Table 5.2). If the nanofiber layer was too thick

( 15≥ m), the transfection efficiency (Figure 5.6(a)) decreased greatly as the result of

significant resistance increase. However, if the nanofiber layer is too thin (≤ 5 µm), a

large number of cells would contact the Al2O3 membrane and thus died during the

electroporation process (Figure 5.6(b)). The optimal nanofiber scaffold layer thickness is

10 µm in the MSE set-up.

Page 136: Fei Zhengzheng

-115-

Table 5.2: Thickness and corresponding resistance of nanofiber layer controlled by

electrospinning time.

Electrospining time (min) Thickness (µm) Resistance (KΩ)

0 0 5.46

1 4~5 6.19

2 8~10 6.46

4 15~20 8.11

(Continued)

Figure 5.6: Effect of electrospun nanofiber thickness on the transfection efficiency (a)

and cell viability (b) of mouse embryonic stem cells. The thickness of nanofiber layer

corresponds to electrospinning time as shown in Table 5.2.

Page 137: Fei Zhengzheng

-116-

Figure 5.6 continued

5.4 Nanofiber based MSE of mouse embryonic stem (ES) cell colony

5.4.1 Mouse ES cell colony formation with controlled size

Figure 5.7 shows confocal images of mouse ES cells after 24 and 48 hours

culture on PCL/gelatin nanofiber scaffolds with and without 100 µm PS micro-wells.

With cell seeding density at 5,000 / mm2, mouse ES cells formed colonies with a uniform

size after 24 hours culture on PCL/gelatin nanofiber scaffolds with micro-wells. After 48

hours culture, mouse ES cells were expanded outside the 10 µm thick micro-wells.

Page 138: Fei Zhengzheng

-117-

Figure 5.7: Confocal images of mouse ES cell colonies after cultured 24 (a, c) and 48 (b,

d) hours on randomly distributed PCL/gelatin nanofiber scaffolds without (a, b) and with

(c, d) 100 µm PS micro-wells. The cell seeding density was 5,000 / mm2. Cells were

fixed with 70% ethanol and stained with PI dye. The length of the standard bars is 100

µm.

Page 139: Fei Zhengzheng

-118-

5.4.2 Nanofiber based MSE with vs without micro-well spacer

Mouse ES cells were seeded on the cell binding substrates, PCL/gelatin nanofiber

scaffolds with and without a micro-well spacer, respectively. The cell seeding density

was 5,000 / mm2. As shown in Figure 5.7, the PCL/gelatin nanofiber surface with 100

µm micro-wells were fully covered with mouse ES cells after 24-hour culture. After 48-

hour culture, mouse ES cells were grown beyond confluence on the flat nanofiber surface

and expanded outside the 10 µm thick micro-wells. To perform nanofiber based MSE at

confluent mouse ES cells with no more than double layers, SEAP transfection was

carried out after 24-hour pre-culture. Since the solid spacer blocked the passage of the

electric field, the amount of SEAP expression from mouse ES cell colonies cultured on

PCL/gelatin nanofiber scaffolds with the micro-well spacer is much higher than that

without micro-well spacer (Figure 5.8). More importantly, the size and thickness of cell

colonies was more uniform with the micro-well spacer (Figure 5.9).

Page 140: Fei Zhengzheng

-119-

Figure 5.8: SEAP transfection of mouse embryonic stem cells by nanofiber based MSE

without ( ) and with ( ) micro-well spacer, bulk electroporation by Bio-Rad Gene Pulser

X-Cell system ( ), and nucleofection ( ): (a) transfection efficiency (b) cell viability 24

hours after electroporation.

Page 141: Fei Zhengzheng

-120-

Figure 5.9: Confocal images of mouse ES cells 6 (a, c) and 30 (b, d) hours after

nanofiber based MSE without (a, b) and with (c, d) 100 µm micro-well spacer. Mouse ES

cells were fixed with 70% ethanol and stained with PI dye. The length of the standard

bars is 100 µm.

Page 142: Fei Zhengzheng

-121-

5.4.3 Nanofiber based MSE vs Bulk electroporation of cell colony

SEAP transfection of mouse ES cell colonies by nanofiber based MSE was further

compared with bulk electroporation and nucleofection. Nanofiber based MSE with 100

µm PS micro-well spacers was performed as described in Section 5.4.2. For bulk

electroporation and nucleofection, half million ( 5105× ) mouse ES cells were seeded in

each well of the gelatin coated 6-well plate. After 24 hours pre-culture, mouse ES cell

colonies were harvested, and the total cell number from each well was around one million

( 6101× ). Colonies of mouse ES cells collected from each well were used for one bulk

electroporation or nucleofection sample. With only one percent of initial input cells, the

transfection efficiency of nanofiber based MSE with the micro-well spacer was

comparable to bulk electroporation (Figure 5.8(a)). Also, the electric field strength

applied in nanofiber based MSE was 150 V/cm, less than one third of 500 V/cm used in

bulk electroporation. Consequently, mouse ES cells experienced an average survival rate

of ~ 66 % in nanofiber based MSE, much higher than ~ 42 % in bulk electroporation

(Figure 5.8(b)). Although the transfection efficiency of nanofiber based MSE was not as

good as nucleofection, the survival rate of mouse ES cell colonies using nanofiber based

MSE was much better, ~ 35 % higher than that of nucleofection.

Page 143: Fei Zhengzheng

-122-

5.5 Nanofiber based MSE of NIH 3T3 fibroblasts

Previous studies (Rodolfa and Eggan, K. 2006; Lowry and Plath, 2008; Pera,

2009) have shown that transient expression of reprogramming factors into embryonic

fibroblasts is sufficient to obtain iPS cells. However, the very low success rate (less than

10 colonies per one million input cells) and a substantial degree of clone-to-clone

variation limited its clinical applications of iPS cells. In addition, the reprogramming

factors are diluted during cell division in the case of plasmid transfection. (Okita et al.,

2008) To achieve better reprogramming efficiency, longer exposure to programming

factors and thus repeated transfection is needed.

5.5.1 NIH 3T3 fibroblasts with micro-well spacer

NIH 3T3 fibroblasts were seeded on the PET membrane surface with and without

micro-well spacer. Figure 5.10 shows the phase contrast images of NIH 3T3 fibroblasts

after 48 hours culture. With the micro-well spacer, NIH 3T3 fibroblasts did not spread

out to the fiber shape as observed on flat cell culture surfaces, instead formed high

compact cell clusters, similar to the morphology of mouse ES cell colonies. This suggests

the potential of using MSE for obtaining induced pluripotent stem (iPS) by plasmid

transfection.

Page 144: Fei Zhengzheng

-123-

Figure 5.10: Morphology of NIH 3T3 fibroblasts with (a) and without (b) 300 µm micro-

well spacer after 48 hours.

5.5.2 Repeated SEAP transfection of NIH 3T3 fibroblasts

The repeated SEAP transfection results of NIH 3T3 fibroblasts by MSE (Figure

5.11) demonstrate that cell viability significantly decreased, resulting in a decrease of the

Page 145: Fei Zhengzheng

-124-

transfection efficiency after the 2nd treatment. Since fibroblast proliferation was found to

be much better on nanofiber scaffolds (Park et al., 2007), the cell-binding substrate was

changed to nanofiber scaffolds. Nanofiber based MSE was conducted at day 0 and 1, and

the transfection results are shown in Figure 5.11. With repeated nanofiber based MSE,

NIH 3T3 fibroblasts expressed SEAP stably. Using electrospun nanofiber scaffolds as the

cell-binding substrate, cell viability was also higher than that using micro-porous

membrane, because the nanofibrous surface provides better cell adhesion and faster cell

recover rate after electroporation. The combination of MSE and electrospun nanofiber

scaffolds allows repeated plasmid transfection at different time points during cell culture.

Furthermore, the micro-well spacer provides well-defined microenvironment to form cell

colony with uniform size, and thus can reduce the transfection difference from colony to

colony.

Page 146: Fei Zhengzheng

-125-

Figure 5.11: Repeated SEAP transfection of NIH 3T3 fibroblasts using PET micro-

porous membrane () and electrospun PCL/gelatin nanofiber scaffolds (): (a)

transfection efficiency (b) cell viability at day 1 and 2 post-electroporation. Both cell-

binding substrates had the micro-well spacer.

Page 147: Fei Zhengzheng

-126-

5.6 Conclusions

The successful combination of MSE and electrospun nanofiber scaffold allowed

better plasmid transfection during colony formation of mouse ES cells. The micro-well

spacer provides well-defined microenvironment to form cell colony with uniform size,

and thus can reduce the transfection difference from colony to colony. Furthermore, the

electrospun nanofiber scaffold as cell-binding substrate improved the cell survival and

recover rate of embryonic fibroblasts during repeated plasmid SEAP transfection,

demonstrating the potential for better iPS cell generation.

Page 148: Fei Zhengzheng

-127-

CHAPTER 6: CONCLUSIONS AND RECOMMENDATIONS

6.1 Conclusions

A new membrane sandwich electroporation (MSE) approach was developed and

demonstrated using plasmids GFP and SEAP as reporter genes. Several cell lines, such as

NIH 3T3 fibroblasts and mouse embryonic stem (ES) cells were tested and a significant

improvement in transgene expression and cell viability was observed comparing to

current electroporation techniques. In the MSE method, the focused electric field

enhances cell permeabilization at a lower electric voltage, leading to high cell viability.

Furthermore, the sandwich membrane configuration is able to provide better gene

confinement near the cell surface, facilitating gene delivery into the cells.

We also demonstrated the use of femtosecond laser fabricated micro-nozzle arrays

on a gelatin-coated PET membrane for MSE. Using micro-nozzle array enhanced MSE,

we observed higher and more uniform gene transfection with excellent cell viability of

mouse ES cells comparing to that achieved with the bulk electroporation methods.

The strategy of ex vivo human-stem-cell-based therapy for tissue regeneration

includes transfecting human stem cells in vitro to express a certain transgene or to

differentiate to certain cells, and then implanting them in vivo under pharmaceutical or

Page 149: Fei Zhengzheng

-128-

blood bank standards of Good Manufacturing Practice (GMP). Successful pre-clinical

trials of genetically modificated human stem cells using in vitro electroporation has been

demonstrated for bone repair (Aslan et al., 2006) and dentin formation (Nakashima et al.,

2004). Since cells or tissues from the patients are often very limited and therapeutic

materials such as plasmids and oligonucleotides are very expensive, the ability to treat a

small number of cells (i.e. hundreds) instead of millions offers great potential for hard-to-

harvest patient cells in patient-specific ex vivo gene therapy and in vitro pharmaceutical

kinetic studies.

We also successfully integrated the electrospun nanofiber scaffold as a cell-

binding substrate in MSE, i.e. nanofiber based MSE. With a micro-well spacer, uniformly

sized colonies of mouse ES cells were obtained, and plasmid transfection by

electroporation was performed during colony formation. Also, repeated plasmid SEAP

transfection of NIH 3T3 fibroblasts was tested, and better cell survival and recover rates

were observed comparing to that using a micro-porous membrane. Due to its ability to

repeated transfection with reprogramming factors, the nanofiber based MSE method has

potential for efficient iPS cell generation by repeated plasmid transfection.

Page 150: Fei Zhengzheng

-129-

6.2 Recommendations

6.2.1 Individual cell array trapping

In this study, we demonstrated the ability to treat a small number of cells using

membrane sandwich electroporation (MSE), and observed high gene transfection and cell

viability comparing to current bulk electroporation techniques. Using a 10 x 10 micro-

nozzle array, we once achieved very exciting transfection results that almost all 100 cells

were transfected with pmaxGFP and remained alive. However, it is very difficult to

repeat such a perfect experiment because of difficulty to align every individual cell on a

micro-nozzle using the vacuum-assisted seeding process. More robust and reliable

techniques should be developed allowing precisely control of the position of individual

cells on a large cell array without causing cell damage. For example, optical trapping

using an optical tweezer array as shown in Figure 6.1 can be a potential solution. In this

method, an array of individual cells is optically trapped by an optical tweezer array

created by focused laser beam through a micro-lens array. By designing a suitable micro-

lens array with the distance of two adjacent optical traps matching that of the adjacent

micro-nozzles, the trapped cells can be aligned right on top of micro-nozzles.

Page 151: Fei Zhengzheng

-130-

Figure 6.1: Schematic illustration of individual cell array trapping by an optical tweezer

array created by focused laser beam through a micro-lens array.

6.2.2 Cell membrane permeability experiments

Our numerical simulation of transmembrane potential distribution qualitatively

explained the effects of cell shape, porosity and pore shape of top and bottom membranes

on MSE. However, there remained large discrepancy in some cases. This is because the

electroporation mediated gene transfection process (Figure 6.2) includes two parts: (1)

cell membrane break-down and reseal, and (2) genes bounding to the cell membrane

during the electroporation and entering cell plasma by endocytosis. The simulation in

Chapter 4 provides the distribution of initial transmembrane potential under an imposed

external electric field. If the transmembrane potential is larger than the critical

Page 152: Fei Zhengzheng

-131-

transmembrane potential, cell membrane becomes permeable. Therefore, the membrane

permeability experiments should be carried out to obtain the critical transmembrane

potential for each cell type in order to link the simulation with transfection prediction in

both MSE and bulk electroporation. A method using the spin-disk confocal microscopy

with appropriate fluidic system and cell impermanent dye, propidium iodide (PI), is being

developed in our lab. Time series studies to track how the dye enters the cell should be

further explored such that the critical transmembrane potential and the cell permeable

location can be measured. Furthermore, the electroporation experiments with different

size of fluorescein-labeled dextrans under different electric fields should be used to

quantify the relationship between the transmembrane potential and size of nanopores on

cell membrane.

Figure 6.2: Electroporation of a cell. The electroporation mediated gene transfection

process includes two parts: (1) cell membrane break-down and reseal, (2) genes bounding

to the cell membrane during the electroporation, and entering cell plasma by endocytosis.

(http://www.inovo.com)

Page 153: Fei Zhengzheng

-132-

In Chapter 5, we have demonstrated the possibility of nanofiber based MSE for

iPS cell generation by repeated plasmid transfection. During the repeated electroporation,

the MSE disks have to be transferred from culture wells to the current fluidic device

(Figure 3.2), and then placed back to culture wells every time. Since the cell-binding

substrates could not be perfectly fixed on the current fluidic devices after culture in the

media, the electric field might leak through the edge of the MSE disk. To address this

problem, a new design that allows cell culture with the cell-binding substrate was

developed, and Figure 6.3 shows this latest MSE system including a MSE stage with two

MSE disks and an electroporation box. Each MSE disk consists of a bottom and a top

piece. The cell-binding substrate and nanofiber scaffolds with micro-well spacer can be

fixed on the bottom piece, and two bottom pieces can be cultured together in one well on

a 6-well plate. Advantages of this new design should be further investigated, particularly

for transfection of delicate cells such as human stem cells and iPS cells.

Page 154: Fei Zhengzheng

-133-

Figure 6.3: The third generation of MSE system, including (a) a MSE stage with two

MSE disks and (b) an electroporation box, which is able to connect with one AC pulse

generator and one DC power supply. Each MSE disk consists of a bottom and a top piece.

(Designed and fabricated by Mr. Shi-Chiung Yu and Dr. Weixiong Wang, 2009)

Page 155: Fei Zhengzheng

-134-

One advantage of the MSE method is able to provide better gene confinement

near the cell surface, facilitating gene delivery into the cells. Although a low DC electric

field was applied to pre-concentrate genes in this study, the negatively charged DNA

molecules tended to migrate away from the negatively charged cell membrane when the

applied electric field was switched from the low DC electric field to a high voltage AC

electric pulse during. This time lapse was around 3-5 seconds. To avoid this problem, a

low DC electric field should be imposed on DNA molecules before, during and after

electroporation. A new electroporation box (Figure 6.3(b)) is constructed, which is able

to connect with one AC pulse generator with one DC power supply. This new device and

box is our third generation MSE system, and currently is under investigation.

Comparing the single cell electroporation and cell colony electroporation results,

we found that single cell nucleofection provided much better transfection efficiency,

while our nanofiber based MSE provided the highest cell survival and recover rates.

Therefore, an ideal procedure for iPS cell generation may include single cell transfection

by nucleofection, cell seeding on nanofiber scaffolds with a micro-well spacer, and

repeated cell colony transfection by nanofiber based MSE. Since one of the most

important factors of successful nucleofection is the cell-specific electroporation buffer, a

combination of nucleofection solution with our MSE method could provide a synergistic

effect. This idea should also be explored in the future study.

Page 156: Fei Zhengzheng

-135-

6.3 Possible ways of in vivo applications

One possible way of in vivo applications is the combination of MSE with the

cannula for thymic cancer or breast cancer treatment. A good example is thymomas.

Although rare, thymomasis the most common type of thymic cancer and is hard to detect

at the early stage. Thymoma arises from thymic epithelial cells, which cover the thymus.

The thymus is a small organ that locats in the upper chest just below the neck. Some

preliminary results of transfecting embryoid bodies using the MSE method (Appendix E)

demonstrated that only few outer layers (less than 3 layers) were transfected. Since it is

nearly impossible to remove thymoma by surgery, MSE may be very suitable for

treatment in such case, because our MSE method is able to treat thymoma derived from

the outer thymic epithelial cells without disturbing the inside thymus.

Another possible in vivo application of MSE is to use a double-balloon

enteroscope to treat the solid tumors in the gastrointestinal system. The double-balloon

enteroscope (or push-and-pull enteroscope) was developed by Yamamoto et al. in 2001

and has been used for diagnosis and control of bleeding with electrocoagulation

(Nishimura et al., 2004). For the malignant tumors that are hard to be surgically resected,

MSE with the double-balloon enteroscope is a promising technique for performing

electrochemotherapy or electrogenetherapy in targeted tissue with minimal the damage to

surrounding tissues. The outer-balloon should be designed using micro-porous

membrane, and the inner-balloon should carry out a bundle of micro-electrodes and drugs

and genes. When the outer-balloon wraps up the targeted tumor, the drugs or genes are

Page 157: Fei Zhengzheng

-136-

able to be injected into the tumor by electroporation. Before in vivo treatment, in vitro

test of a tumor sample from the patient by MSE should be first investigated, followed by

pre-clinical testing in small and large animal models.

Page 158: Fei Zhengzheng

-137-

REFERENCES

Aguliar, C. A.; Lu, Y.; Mao, S.; Chen, S. (2005) Direct micro-patterning of biodegradable polymers using ultraviolet and femtosecond lasers. Biomaterials 26(36): 7642-7649.

Amit, M.; Carpenter, M. K.; Inokuma, M. S.; Chiu, C-P.; Harris, C. P.; Waknitz, M. A.; Itskovitz-Eldor, J.; Thomson, J. A. (2000) Clonally derived human embryonic stem cell lines maintain pluripotency and proliferative potential for prolonged periods of culture. Developmental Biology 227: 271–278.

Anderson, M. L. M.; Spandidos, D. A.; Coggins, J. R. (1991) Electroporation of lymphoid cells: factors affecting the efficiency of transfection. Journal of Biochemical

and Biophysical Methods 22: 207-222.

Aslan, H.; Zilberman, Y.; Arbeli, V.; Sheyn, D.; Matan, Y.; Liebergall, M.; Li, J-Z.; Helm, G. A.; Gazit, D.; Gazit, Z. (2006) Nucleofection-based ex vivo nonviral gene delivery to human stem cells as a platform for tissue regeneration. Tissue Engineering 12(4): 877-889.

Ben-Nuna, I. F.; Benvenisty, N. (2006) Human embryonic stem cells as a cellular model for human disorders. Molecular and Cellular Endocrinology 252 (1-2): 154-159.

Blaese, R. M.; Culver, K. W.; Miller, A. D.; et al. (1995) T lymphocyte-directed gene therapy for ADA- SCID: initial trial results after 4 years. Science 270: 475-480.

Bier, M.; Chen, W.; Gowrishankar, T. R.; Astumian, R. D.; Lee, R. C. (2002) Resealing dynamics of a cell after electroporation. Physics Review E 66(062905): 1-4.

Carpenter, M. K.; Xu, C.; Daigh, C. A.; et al. (2003) Protocols for the Isolation and Maintenance of Human Embryonic Stem Cells. Human Embyronic Stem Cells edited by Chiu, M., Rao, M.S.

Chang, D. C.; Reese, T. S. (1990) Changes in membrane structure induced by electroporation as revealed by rapid-freezing electron microscopy. Biophysical Journal 58(1): 1–12.

Page 159: Fei Zhengzheng

-138-

Chang, D. C.; Chassy, B. M. and Saunder, J. A. (1992) Guide to electroporation and electrofusion, Academic, San Diego, CA.

Chen, W.; Lee, R. C. (1994) An improved double vaseline voltage clamp to study electroporated skeletal muscle fibers. Biophysical Journal 66: 700–709.

Chernomordik, L. V.; Sukharev, S. I.; Abidor, I. G.; Chizmadzher, Y. A. (1983) Breakdown of lipid bilayer membranes in an electric field. Biochimica et Biophysica Acta

736: 203-213.

Chernomordik, L. V.; Sukharev, S. I.; Popov, S. V.; Pastushenko, V. F.; Sokirko, A. V.; Abidor, I. G.; Chizmadzher, Y. A. (1987) The electrical breakdown of cell and lipid membranes: the similarity of phenomenologies. Biochimica et Biophysica Acta 902(3): 360–373.

Chong, E. J.; Phan, T. T.; Lim, I. J; Zhang, Y. Z; Bay, B. H.; Ramakrishna, S.; Lim, C. T. (2007) Evaluation of electrospun PCL/gelatin nanofibrous scaffold for wound healing and layered dermal reconstitution. Acta Biomaterialia 3(3): 321-330.

Chu, G.; Hayakawa, H.; Berg, P. (1987) Electroporation for the efficient transfection of mammalian cells with DNA. Nucleic Acids Research 15(3): 1311-1326.

Glogauer, M.; Lee, W.; McCulloch, C. A. (1993) Induced endocytosis in human fibroblasts by electrical fields. Experimental Cell Research 208(1): 232-240.

Coulberson, A. L.; Hud, N. V.; LeDoux, J. M.; et al. (2003) Gene packaging with lipids, peptides and viruses inhibits transfection by electroporation in vitro. Journal of

Controlled Release 86: 361–370.

Delteil, C.; Teissie, J.; Rols, M. P. (2000) Effect of serum on in vitro electrically mediated gene delivery and expression in mammalian cells. Biochime Biophysics Acta 1467: 362-368.

Drukker, M.; Benvenisty, N. (2003) Genetic Manipulation of Human Embryonic Stem Cells. Human Embryonic Stem Cells edited by Chiu, A.Y., Rao, M.S. 265 – 284.

Evans, M. J.; Kaufman, M. H. (1981) Establishment in culture of pluripotential cells from mouse embryos. Nature 292: 154–156.

Farsona, D. F.; Choi, H. W.; Lu, C.; Lee, J. L. (2006) Femtosecond laser bulk micromachining of microfluid channels in poly-methylmethacrylate. Journal of Laser

Applications 18(3): 210-215.

Page 160: Fei Zhengzheng

-139-

Farson, D. F.; Choi, H. W.; Zimmerman, L. B.; Steach, J. K.; Chalmers, J. J.; Olesik, S. V.; Lee, L. J. (2008) Femtosecond laser micromachining of dielectric materials for biomedical applications. Journal of Micromechanics and Microengineering 18(3): 035020/1-9.

Fei, Z.; Wang, S.; Xie, Y.; Henslee, B. E.; Koh, C. G.; Lee, L. J. (2007) Gene Transfection of Mammalian Cells Using Membrane Sandwich Electroporation. Analytical Chemistry 79 (15): 5719 -5721.

Fuchs, E.; Segre, J. A. (2000) Stem cells: A new lease of life. Cell 100(1): 143-155.

Gabriel, B.; Teissie, J. (1997) Direct observation in the millisecond time range of fluorescent molecule asymmetrical interaction with the electropermeabilized cell membrane. Biophysical Journal 73: 2630-2637.

Gabriel, B.; Teissie, J. (1999) Time courses of mammalian cell electropermeabilization observed milisecond imaging of membrane property changes during the pulse. Biophysical Journal 76: 2158–2165.

Gallego, D.; Higuita, N.; Sharma, S.; Reen, R.; Gooch, K.; Palmer, A.; Lee, L. J.; Lannutti, J.; Hansford, D. J. A micro-nano device for high throughput assembly of 3D clusters. Submitted to Nature Methods

Gehl, J. (2003) Electroporation: theory and methods, perspectives for drug delivery, gene therapy and research. Acta Physiol Scand 177: 437-447.

Glaser, R. W.; Leikin, S. L.; Chernomordik, L. V.; Pastushenko, V. L.; Sokirko, A. I. (1988) Reversible electrical breakdown of lipid bilayers: formation and evolution of pores. Biochimica et Biophysica Acta 940: 275-287.

Golzio, M.; Mora, M-P.; Raynaud, C.; Delteil, C.; Teissié, J.; Rols, M-P. (1998) Control by osmotic pressure of voltage-induced permeabilization and gene transfer in mammalian cells. Biophysics Journal 74(6): 3015-3022.

Golzio, M.; Teissié, J.; Rols, M-P. (2002) Cell synchronization effect on mammalian cell permeabilization and gene delivery by electric field. Biochimica et Biophysica Acta 1563: 23– 28.

Golzio, M.; Teissié, J.; Rols, M-P. (2002) Direct visualization at the single-cell level of electrically mediated gene delivery. Proceedings of the National Academy of Sciences,

USA 99(3): 1292-1297.

Page 161: Fei Zhengzheng

-140-

Hemminger, O.; Yu, Z.; Zhang, C.; Lee, L. J.; Fan, L .S. (2007) Microfluidic Velocity Measurements using Three-Dimensional Confocal Micro Particle Tracking Velocimetry (CM-PTV). The 12th International Conference on Fluidization - New Horizons in

Fluidization Engineering RP4: 489.

Hohenstein, K. A.; Pyle, A. D.; Chern, J. Y.; Lock, L. F.; Donovan, P. J. (2008) Nucleofection Mediates High-efficiency Stable Gene Knockdown and Transgene Expression in Human Embryonic Stem Cells. Stem Cells 26: 1436-1443.

Hollon, T. (2000)Researchers and regulators reflect on first gene therapy death. Nature

Medicine 6: 6.

Hu, X.; Wang, S.; Lee, L. J. (2009) Single-molecule DNA dynamics in tapered contraction-expansion microchannels under electrophoresis. Physical Review E 79(4): 041911.

Huang, Y.; Rubinsky, B. (1999) Micro-electroporation: improving the efficiency and understanding of electrical permeabilization of cells. Biomedical Microdevices 2(2): 145-150.

Huang, Y.; Rubinsky, B. (2003) Flow-through micro-electroporation chip for high efficiency single-cell genetic manipulation. Sensor Actuators A 104: 205-212.

Hur, J. S.; Shaqfeh, E. S. G. (2001) Dynamics of dilute and semidilute DNA solutions in the start-up of shear flow. Journal of Rheology 45(2): 421-450.

Ingber, D. E. (2005) Mechanical control of tissue growth: function follows form. Proceedings of the National Academy of Sciences, USA 102(33): 11,471 – 11, 572.

Izumikawa, M.; Minoda1, R.; Kawamoto, K.; et al. (2005) Auditory hair cell replacement and hearing improvement by Atoh1 gene therapy in deaf mammals. Nature Medicine 11: 271-276.

Jacobsen, F.; Mertens-Rill, J.; Beller, J.; Hirsch, T.; Daigeler, A.; Langer, S.; Lehnhardt, M.; Steinau, H.; Steinstraesser, L. (2006) Nucleofection: A new method for coetaneous gene transfer. Journal of Biomedicine and Biotechnology 2006: 26060.

Jaroszeski, M. J.; Gilbert, R.; Nicolau, C.; Heller, R. (1999) In vivo gene delivery by electroporation. Advanced Drug Delivery Reviews 35(1): 131-137.

Kaji, K.; Norrby, K.; Paca, A.; Mileikovsky, M.; Mohseni, P.; Woltjen, K. (2009) Virus-free induction of pluripotency and subsequent excision of reprogramming factors. Nature 458: 771-775.

Page 162: Fei Zhengzheng

-141-

Kanduser, M.; Fosnari,c M.; Sentjurc, M.; Kralj-Iglic, V.; Hagerstrand, H.; Iglic, A.; Miklavcic, D. (2003) Effect of surfactant polyoxyethylene glycol (C12E8) on electroporation of cell line DC3F. Colloids and Surfaces A: Physicochemical and

Engineering Aspects 214(1): 205-217.

Klenchin, V. A.; Sukharev, S. I.; Serov, S. M.; Chernomordik, L. V.; Chizmadzhev Y.A. (1991) Electrically induced DNA uptake by cells is a fast process involving DNA electrophoresis. Biophysical Journal 60(4): 804-811.

Khine, M.; Lau, A.; Ionescu-Zanetti, C.; Seo. J.; Lee, L-P. (2005) A single cell electroporation chip. Lab on a Chip 5(1): 38-43.

Kim, J. H.; Auerbach, J. M.; Rodriguez-Gomez, J. A.; et al. (2002) Dopamine neurons derived from embryonic stem cells function in an animal model of Parkinson's disease. Nature 418(6893): 50-56.

Kong, H. J.; Liu, J.; Riddle, K.; Matsumoto, T.; Leach, K.; Mooney, D. J. (2005) Non-viral gene delivery regulated by stiffness of cell adhesion substrates. Nature Materials 4: 460-464.

Krassowska, W.; Filev, P. D. (2007) Modeling Electroporation in a Single Cell. Biophysical Journal 92: 404-417.

Kotnik, T.; Bobanović,F. and Miklavčič, D. (1997) Sensitivity of transmembrane voltage induced by applied electric fields-a theoretical analysis. Bioelectrochemistry and

Bioenergetics 43(2): 285-291.

Kotnik, T.; Pucihar, G.; Rebersek, M.; Mir, L. M.; Miklavcic, D. (2003) Role of pulse shape in cell membrane electropermeabilization. Biochimica et Biophysica Acta 1614:193–200.

Le Meur, G.; Stieger, K.; Smith, A. J.; et al. (2007) Restoration of vision in RPE65-deficient Briard dogs using an AAV serotype 4 vector that specifically targets the retinal pigmented epithelium. Gene Therapy 14(4): 292-303.

Lee, S-W.; Tai, Y-C. (1999) A micro cell lysis device. Sensor Actuators A 73(1–2): 74 - 79.

Lensch, M. W. (2009) Cellular reprogramming and pluripotency induction. British

Medical Bulletin 90:19-35.

Lim, Y. C.; Johnson, J.; Fei, Z.; Farson, D. F.; Lannutti, J.; Lee, L. J. 3-D micropocket fabrication on electrospun PCL/gelatin nanofiber tissue scaffolds by direct write

Page 163: Fei Zhengzheng

-142-

femtosecond laser ablation for mouse embryonic stem cell growth. Submitted to Tissue

Engineering C.

Lin, Y-C.; Jen, C-M.; Huang, M-Y.; Wu, C-Y.; Lin, X-Z. (2001) Electroporation microchips for continuous gene transfection. Sensor Actuators B 79: 137–143.

Lindvall, O.; Kokaia, Z. (2006) Stem cells for the treatment of neurological disorders. Nature 441 (7097): 1094 – 1096.

Liu, Y. M.; Sun, S.; Singha, S.; Cho, M. R.; Gordon, R. J. (2005) 3D femtosecond laser patterning of collagen for directed cell attachment. Biomaterials 81(3): 298-305.

Lorenz, U. P.; Harnack, R. M. (2004) Efficient gene transfer into murine embryonic stem cells by nucleofection. Biotechnology Letter 26: 1589-1592.

Lowry, W. E.; Plath, K. (2008) The many ways to make an iPS cell. Nature

Biotechnology 26: 1246-1248.

Lu, H.; Schmidt, M.A.; Jensen, K. F. (2005) A microfluidic electroporation device for cell lysis. Lab on a Chip 5: 23–29.

Maherali, N.; Hochedlinger, K. (2008) Guidelines and techniques for the generation of induced pluripotent stem cells. Cell Stem Cell 3: 595-605.

Martin, G.R. (1981) Isolation of a pluripotent cell line from early mouse embryos cultured in medium conditioned by teratocarcinoma stem cells. Proceedings of the

National Academy of Sciences, USA 78: 7634–7638.

Mehier-Humbert, S.; Guy, R. H. (2005) Physical Methods for Gene Transfer: Improving the Kinetics of Gene Delivery into Cells. Advance Drug Delivery Review 57: 733-753.

Miklavcic1 D.; L. Towhidi. (2008) Time course of electrical and diffusional parameters during and after electroporation. 4th European Conference of the International

Federation for Medical and Biological Engineering: ECIFMBE 2008, IFMBE

Proceedings 22: 2659–2663.

Mir, L. M.; Bureau, M. F.; Gehl, J.; Daniel, et al. (1999) High-efficiency gene transfer into skeletal muscle mediated by electric pulses. Proceedings of the National Academy of

Sciences, USA 96(8): 4262-4267.

Mir, L.M.; Moller, P.H.; Andre´, F.; Gehl, J. (2005)Electric pulse-mediated gene delivery to various animal tissues. Advances in Genetics 54: 83-114.

Page 164: Fei Zhengzheng

-143-

Mohour, J. C.; dePablo, J. J.; Palecek, S. P. (2006) Electroporation of human embryonic stem cells: small and macromolecule loading and DNA transfection. Biotechnology

Progress 22: 825-834.

Mounir T. (2005) Membrane Electroporation: A Molecular Dynamics Simulation. Biophysical Journal 88: 4045–4053.

Mulligan, R. C. (1993) The basic science of gene therapy. Science 260: 926–932.

Myers, A.K.; Tisa, L.S. (2003) Effect of electroporation conditions on cell viability of Frankia EuI1c. Plant and Soil 254 (1): 83-88.

Nakashima, M.; Iohara, K.; Ishikawa, M.; Ito, M.; Tomokiyo, A.; Tanaka, T.; Akamine, A. (2004) Stimulation of Reparative Dentin formation by ex vivo gene therapy using dental pulp stem cells electrotransfected with growth/differentiation factor 11 (Gdf11). Human Gene Therapy 15(11): 1045-1053.

Neumann, E.; Kakorin, S.; Toensing, K. (1999) Fundamentals of electroporative delivery of drugs and genes. Bioelectrochemistry and Bioenergetics 48: 3–16.

Niidome, T.; Huang, L. (2002) Gene therapy progress and prospects: nonviral vectors. Gene Therapy 9: 1647-1652.

Nishimura, M.; Yamamoto, H.; Kita, H.; Yano, T.; Sunada, K.; Miyata, T.; Sugimoto, T.; Iino, S.; Sekine, Y.; Iwamoto, M.; Ohnishi, N.; Kuno, A.; Ohnishi, H.; Sakurai, S.; Ido, K.; Sugano. K. (2004) Gastrointestinal stromal tumor in the jejunum: diagnosis and control of bleeding with electrocoagulation by using double-balloon enteroscopy. Journal

of Gastroenterology 39(10):1001–1004.

Nociari, M. M.; Shalev, A.; Benias, P.; Russo, C. (1998) A novel one-step, highly sensitive fluorometric assay to evaluate cell-mediated cytotoxicity. Journal of

Immunological Method 213:157–167.

O’Connor, T. P.; Crystal1, R. G. (2006) Genetic medicines: treatment strategies for hereditary disorders. Nature Reviews Genetics 7: 261-276.

Odorico, J. S.; Kaufman, D. S.; Thomson, J. A. (2001) Multilineage differentiation from human embryonic stem cell lines. Stem cells 19: 193-204.

Okita, K.; Nakagawa, M.; Hyenjong, H.; Ichisaka, T,; Yamanaka, S. (2008) Generation of mouse induced pluripotent stem cells without viral vectors. Science 322: 949-953.

Page 165: Fei Zhengzheng

-144-

Park, J. S.; Choi, C. K.; Kihm, K. D. (2004) Optically sliced micro-PIV using confocal laser scanning microscopy (CLSM). Experiments in Fluid 37: 105-119.

Park, K.; Ju, Y. M.; Son, J. S.; Ahn, K.; Han, D. K. (2007) Surface modification of biodegradable electrospun nanofiber scaffolds and their interaction with fibroblasts. Journal of biomaterials science 18 (4): 369-382.

Passier, R.; van Laake, L. W.; Mummery, C. L. (2008) Stem-cell-based therapy and lessons from the heart. Nature 453: 322-329.

Pera, M. F. (2009) Stem cells: Low-risk reprogramming. Nature 458: 715-716.

Perkins, T. T.; Smith, D. E.; Chu, S. (1997) Single polymer dynamics in an elongational flow. Science 276: 2016 -2021.

Potter, H. (1993) Application of electroporation in recombinant DNA technology. Mehoths of Enzymology 217: 461-478.

Rodolfa, K. T.; Eggan, K. (2006) A transcrptional logic for nuclear reprogramming. Cell 126: 652 – 655.

Rols, M. P.; Teissie, J. (1990) Electropermeabilization of mammalian cells: Quantitative analysis of the phenomenon. Biophysical Journal 58: 1089-1098.

Saulis, G. (1997) Pore disappearance in a cell after electroporation: theoretical simulation and comparison with experiments. Biophysical Journal 73(3): 1299-1309.

Schindhelm, K.; Nordon, R. Ex vivo cell therapy, Academic, London (1999).

Serafetinides, A. A. (1997) Short pulse laser beam interactions with polymers biocompatible materials and tissue. Proceeding of SPIE (International Society for Optical

Engineering) 3052: 111-122.

Siemen, H.; Nix, M.; Endl, E.; et al. (2005) Nucleofection of human embryonic stem cells. Stem Cell Development 14: 378-383.

Smith, D. E.; Perkins, T. T.; Chu, S. (1996) Dynamical Scaling of DNA Diffusion Coefficients. Macromolecules 29: 1372.

Stadtfeld, M.; Nagaya, M.; Utikal, J.; Weir, G.; Hochedlinger, K. (2008) Induced pluripotent stem cells generated without viral integration. Science 322: 945-949.

Stadtfeld, M.; Hochedlinger, K. (2009) Without a trace? PiggyBac-ing toward pluripotency. Nature Methods 6: 329-330.

Page 166: Fei Zhengzheng

-145-

Stewart, D. A.; Gowrishankar, T. R.; Smith, K. C.; Weaver, J. C. (2005) Cylindrical cell membranes in uniform applied electric fields: validation of a transport lattice method. IEEE Transactions on Biomedical Engineering 52(10): 1643-1653.

Strulovici1, Y., Leopold, P. L., O’Connor, T. P., et al. (2007) Human Embryonic Stem Cells and Gene Therapy. Molecular Therapy 15(5): 850-866.

Sukharev, S. I.; Klenchin, V. A.; Serov, S. M.; Chernomordik, L. V.; Chizmadzhev, Y. A. (1992) Electroporation and electrophoretic DNA transfer into cells. The effect of DNA interaction with electropores. Biophysical Journal 63(5): 1320–1327.

Takahashi, K.; Tanabe, K.; Ohnuki, M.; Narita, M.; Ichisaka, T.; Tomoda, K.; Yamanaka, S. (2007) Induction of pluripotent stem cells from adult human fibroblasts by defined factors. Cell 131: 861-872.

Takahashi, K.; Yamanaka, S. (2006) Induction of pluripotent stem cells from mouse embryonic and adult fibroblast cultures by defined factors. Cell 126: 663-676.

Takamatsu, H.; Rubinsky, B. (1999) Viability of deformed cells. Cryobiology 39, 243 -251.

Teissie, J.; Rols, M. P. (1993) An experimental evaluation of the critical potential difference inducing cell membrane electropermeabilization. Biophysical Journal 65(1):

409-413.

Thomas, C. E.; Ehourhardt, A.; Kay, M. A. (2003) Progress and problems with the use of viral vectors for gene therapy. Nature Reviews Genetics 4: 346-358.

Thomson, J. A.; Itskovitz-Eldor, J.; Shapiro, S. S.; et al. (1998) Embryonic Stem Cell Lines Derived from Human Blastocysts. Science 282: 1145-1147.

Tiaw, K.-S.; Goh, S.-W.; Hong, M.; Wang, Z.; Lan, B.; Teoh, S.-H. (2005) Laser surface modification of poly(ε-caprolactone) (PCL) membrane for tissue engineering applications. Biomaterials 26(7): 763-769.

Tompers, D.M.; Labosky, P.A. (2004) Electroporation of murine embryonic stem cells: step-by-step guide. Stem Cells 22: 243-249.

Tsong 1991 Electroporation of cell membrane. Biophysical Journal

Tsuruma, A.; Tanaka, M.; Yamamoto, S.; Shimonura, M. (2008) Control of neural stem cell differentiation on honeycomb films. Colloids and Surfaces A-Physicochemical and

Engineering Aspects 313: 536-540.

Page 167: Fei Zhengzheng

-146-

Varel, H.; Ashkenasi, D.; Rosenfeld, A.; et al. (1997) Micromachining of quartz with ultrashort pulses. Applied Physiscs A 65(4-5): 367-373.

Wang, S.; Hu, X.; Lee, L. J. (2008) Electrokinetics induced asymmetric transport in polymeric nanonozzles. Lab on a Chip 8: 573-581.

Weaver, J. C. Electroporation: A general Phenomenon for manipulating cells and tissues

Weaver, J. C.; Chizmadzhev, Y. A. (1996) Theory of electroporation: A review. Biochemistry and Bioenergetics 41: 135-160.

Wells, D. J. (2004) Gene Therapy Progress and Prospects: Electroporation and other physical methods. Gene Therapy 11: 1363–1369.

Wnek, G. E.; Carr, M. E.; Simpson, D. G.; Bowlin, G. L. (2003) Electrospinning of nanofiber fibrinogen structures. Nano Letters 3(2):213-216.

Woltjen, K.; Michael, I. P.; Mohseni, P.; Desai, R.; Mileikovsky, M.; Hamalainen, R.; Cowling, R.; Wang, W.; Liu, P.; Gertsenstein, M.; Kaji, K.; Sung, H. K.; Nagy, A. (2009) piggyBac transposition reprograms fibroblasts to induced pluripotent stem cells. Nature 458: 766-770.

Xiao, M.; Phong, A.; Ha, Chan, T-F.; Cai, D.; Leung, L.; Wan, E.; Kistler, A. L.; DeRisi, J. L.; Selvin, P. R.; Kwok, P-Y. (2007) Rapid DNA mapping by fluorescent single molecule detection. Nucleic Acids Research 35(3): e16.

Xu, C.; Inokuma, M. S.; Denham, J.; Golds, K.; Kundu, P.; Gold, J. D.; Carpenter, M. K. (2001) Feeder-free growth of undifferentiated human embryonic stem cells. Nature

Biotechnology 19: 971 – 974.

Yamamoto, H.; Sekine, Y.; Sato, Y.; Higashizawa, T.; Miyata, T.; Iino, S.; Ido, K.; Sugano, K. (2001) Total enteroscopy with a nonsurgical steerable double-balloon method. Gastrointestinal Endoscopy 53(2): 216-220.

Yashimoto, H.; Shin, Y. M.; Terai, H.; Vacanti, J. P. (2003) A biodegradable nanofiber scaffold by electrospinning and its potential for bone tissue engineering. Biomaterials 24(12): 2077-2082.

Yu, J.; Vodyanik, M. A.; Smuga-Otto, K.; Antosiewicz-Bourget, J.; Frane, J. L.; Tian, S.; Nie, J.; Jonsdottir, G. A.; Ruotti, V.; Stewart, R.; Slukvin, II.; Thomson, J. A. (2007) Induced pluripotent stem cell lines derived from human somatic cells. Science 318:1917-1920.

Page 168: Fei Zhengzheng

-147-

Yu, J.; Hu, K.; Smuga-Otto, K.; Tian, S.; Stewart, R.; Slukvin, II.; Thomson, J. A. (2009) Human induced pluripotent stem cells free of vector and transgene sequences. Science 324: 797-801.

Yusa, K.; Rad, R.; Takeda, J.; Bradley, A. (2009) Generation of transgene-free induced pluripotent mouse stem cells by the piggyBac transposon. Nature Methods 6: 363-369.

Zhang, Y.; Ouyang, H.; Lim, C. T.; Ramakrishna, S.; Huang, Z. (2005) Electrospinning of gelatin fibers and gelatin/PCL composite fibrous scaffolds. Journal of Biomed

Research Part B: Applied Biomaterial 72B(1): 156-165.

Zeng, X.; Chen, J.; Sanchez, J. F.; et al. (2003) Stable expression of hourGFP by mouse embryonic stem cells: promoter activity in the undifferentiated state and during dopaminergic neural differentiation. Stem Cells 21: 647-653.

Zimmermann, U.; Vienken, J. (1982) Electric field-induced cell-to-cell fusion. Journal of

Membrane Biology 67: 165-182.

Zudans, I.; Agarwal, A.; Orwar, O.; Weber, S. G. (2007) Numerical calculation of single-cell electroporation with an electrolyte-filled capillary. Biophysical Journal 92(10): 3696-3705.

Page 169: Fei Zhengzheng

-148-

APPENDIX A: STANDARD CURVE

y = 72.68x

R2 = 0.9908

0

10

20

30

40

50

60

70

0.0 0.2 0.4 0.6 0.8 1.0

OD@405nm

SE

AP

Acti

vit

y (

mU

)

Figure A.1: Standard curve of SEAP activity (mU) vs absorbance reading at 405 nm

(OD@405nm). 100 µL of standard alkaline phosphatase solution of 0, 5, 10, 20, 40, and

60 mU and 25 µL of pNPP substrate solution were added into each well of a 96-well

plate. The plate was incubated in the dark for approximately 15 minutes at room

temperature, and read at the wavelength of 405 nm on a multi-well plate reader.

Page 170: Fei Zhengzheng

-149-

APPENDIX B: OPTIMIZATION OF BULK ELECTROPORATION AND

NUCLEOFECTION OF MOUSE EMBRYONIC STEM (ES) CELLS

Figure B.1: Effect of initial cell number on bulk electroporation of mouse ES cells

using Bio-Rad Gene Pulser X-Cell system: (a) transfection efficiency and (b) cell

viability 24 hours after BE.

0

20

40

60

80

1.00E+06 5.00E+05 1.00E+05

Initial Cell Number

To

tal

SE

AP

Acti

vit

y m

U/

Init

ial

10,0

00 s

eed

ing

cell

s

0

20

40

60

80

100

1.00E+06 5.00E+05 1.00E+05

Initial Cell Number

Cell

Via

bil

ity (

%)

a b

Page 171: Fei Zhengzheng

-150-

Figure B.2: Optimization of bulk electroporation of mouse embryonic stem cells using

Bio-Rad Gene Pulser X-Cell system: (a) transfection efficiency and (b) cell viability 24

hours after BE.

0

25

50

75

375 500 625 750

Electrical field strength (V/cm)

Tota

l S

EA

P A

ctivity m

U/

10,0

00 I

nitia

l seedin

g c

ells

0

25

50

75

100

375 500 625 750

Electrical field strength (V/cm)

Cell

Via

bili

ty (

%)

a

b

Page 172: Fei Zhengzheng

-151-

Figure B.3: Optimization of nucleofection of mouse ES cells: (a) transfection

efficiency and (b) cell viability 24 hours after nucleofection.

0

25

50

75

100

125

A-13 A-23 A-24 A-30

Program

Tota

l S

EA

P A

ctivity m

U/

10,0

00 I

nitia

l seedin

g c

ells

0

25

50

75

100

A-13 A-23 A-24 A-30

Electrical field strength (V/mm)

Ce

ll V

iab

ility

(%

)

a

b

Page 173: Fei Zhengzheng

-152-

0

20

40

60

80

100 150 200

Electrical Field Strength (V/cm)

Rela

tive S

EA

P A

cti

vit

y

(mU

/10,0

00 c

ell

s)

0

20

40

60

80

100

100 150 200

Electrical Field Strength (V/cm)

Cell

Via

bil

ity (

%)

Figure B.4: SEAP transfection of mouse ES cell by micro-nozzle enhanced MSE at

different electrical field strength. Initially, ten thousands of cells were trapped on a 100 x

100 micro-nozzle array, and (a) transfection efficiency and (b) cell viability were

quantified 24 hours after electroporation.

Page 174: Fei Zhengzheng

-153-

Figure B.5: Optimization of bulk electroporation (BE) of mouse ES cell colonies using

Bio-Rad Gene Pulser X-Cell system: (a) transfection efficiency and (b) cell viability 24

hours after BE.

0

2

4

6

8

10

375 500 625 750

Electric field strength V/cm

Tota

l S

EA

P A

cti

vity

mU

/

Initia

l 35

,00

0 s

ee

din

g c

ells

0

20

40

60

80

375 500 625 750

Electric field strength V/cm

Cell

Via

bili

ty (

%)

a

b

Page 175: Fei Zhengzheng

-154-

Figure B.6: Optimization of nucleofection of mouse ES cell colonies: (a) transfection

efficiency and (b) cell viability 24 hours after nucleofection.

0

5

10

15

20

A13 A23 A24 A30

Program#

To

tal S

EA

P A

cti

vit

y m

U/

Init

ial 3

5,0

00

se

ed

ing

ce

lls

0

15

30

45

60

A13 A23 A24 A30

Program #

Ce

ll V

iab

ilit

y (

%)

a

b

Page 176: Fei Zhengzheng

-155-

APPENDIX C: ANALYTICAL SOLUTION OF TRANSMEMBRANE

POTENTIAL FOR A TWO-DIMENTIONAL (2-D) CELL IN BULK

For a 2-D cylindrical cell (three-layer model) in bulk electric field shown in

Figure C.1, the governing equation 0)( =∇⋅∇ φσ can be simplified as

01

)(1

2

2

2

2 =∂

∂+

∂=∇

θ

φφφ

rrr

rr (C.1)

Figure C.1: Schematics of a 2D cylindrical cell in bulk electric field.

Page 177: Fei Zhengzheng

-156-

Equation (B-1) can be solved by the method of separation of variable )()( θφ grf=

and usually we take θθ cos)( =g , then 02 =−′+′′ ffrfr . And we have

rBArrf /)( += , or

θθφ cos)/(),( rBArr += (C.2)

So the electric field φ−∇=E is given by

θθ θ sin)/(cos)/( 22rBArBAr +++−= eeE (C.3)

where the relation between two coordinates systems is given as: θθ θ sincos eee −= rx

and θθ θ cossin eee += ry . Generally, we need to solve the electric potential layer by

layer.

(1) Layer of external medium

Since xe E eE 0−= for ∞→r , where 0E is the electric field far away from the cell,

we get 0EAe = , thus θθφ cos)/(),( 0 rBrEr ee += ,

θθ θ sin)/(cos)/( 2

0

2

0 rBErBE eere +++−= eeE .

(2) Layer of inner cytoplasm

iφ is finite at 0=r , then 0=iB , thus θθφ cos),( rAr ii = ,

θθ θ sincos)( iiri AA eeE +−= .

(3) Layer of cell membrane

θθφ cos)/(),( rBrAr mmm += ,

Page 178: Fei Zhengzheng

-157-

θθ θ sin)/(cos)/( 22rBArBA mmmmrm +++−= eeE .

Boundary conditions are given as:

nn

m

m

e

e∂

∂=

∂ φσ

φσ (or rmmree eEeE ⋅=⋅ σσ ),

me φφ = on external interface Rr = (C.4)

nn

m

m

i

i∂

∂=

∂ φσ

φσ (or rmmrii eEeE ⋅=⋅ σσ ),

mi φφ = on inner interface dRr −= (C.5)

Thus we have totally 4 unknowns and 4 equations as follows:

=

−−−

−−−−−

0

0

0)/(

//0

0)/(1)()(

/1/10

0

0

2

22E

RE

B

B

A

A

dR

RR

dRdRdR

RRR

e

e

m

m

i

mmi

emm σ

σσσ

σσσ (C.6)

The solution is:

))(())(()(

)(222

0

2

memimemi

emi

mRdR

ERA

σσσσσσσσ

σσσ

++−−−−

+−= (C.7)

))(())(()(

)()(222

0

22

memimemi

emi

mRdR

EdRRB

σσσσσσσσ

σσσ

++−−−−

−−= (C.8)

))(())(()(

422

0

2

memimemi

em

iRdR

ERA

σσσσσσσσ

σσ

++−−−−

−= (C.9)

))(())(()(

))(())(()(22

22

0

2

memimemi

memimemi

eRdR

RdRERB

σσσσσσσσ

σσσσσσσσ

++−−−−

−+−+−−= (C.10)

The transmembrane potential is given as )()( dRRV mmm −−= φφ and we have

Page 179: Fei Zhengzheng

-158-

))(())(()(

)2(cos2

22

22

0

memimemi

emei

mRdR

dRddREV

σσσσσσσσ

σσσσθ

++−−−−

−−= (C.11)

Because im σσ << , and em σσ << , you can treat 0=mσ , thus the transmembrane

potential for a 2D cylindrical cell in the bulk electric field is simplified as

θσσ

σσθ cos2

))((

)2(cos2 022

2

0 RERdR

RddREV

ei

ei

m =−−

−= (C.12)

Thus we complete the deduction of the distribution of transmembrane potential for a

2D cylindrical cell in the bulk electric field.

Page 180: Fei Zhengzheng

-159-

APPENDIX D: G-CODE GENERATION FOR FABRICATING MICRO-PORE

ARRAYS BY FEMTOSECOND LASER

Figure D.1: Interface of G-code generation software for femtosecond laser fabrication of

micro-pore arrays. (Programmed by Hae Woon Choi, December 2007)

Page 181: Fei Zhengzheng

-160-

APPENDIX E: ELECTROPORATION OF MOUSE EMBRYOID BODIES

Figure E.1: Formation of mouse embryoid bodies (EBs) by handing drop culture. The

initial cell number per drop is ten thousand in (a-c), and twenty thousand in (d-f). Phase

contrast images were taken at day 3 (a, d), day 5 (b, e) and day 8 (c, f).

Page 182: Fei Zhengzheng

-161-

Figure E.2: Electroporation of EBs by plasmid RFP at day 8. The initial size of EBs is ~

400 µm in (a, b), and ~ 500 µm in (c, d). RFP stands for red fluorescence protein. (a, c)

Shape of EBs; (b,d) RFP transfection results.

Page 183: Fei Zhengzheng

-162-

APPENDIX F: MEMBRANE PERMEABILITY EXPERIMENT

0.1

0.15

0.2

0.25

0.3

0.35

4 6 8 10 12

Diameter of cells (mm)

Tra

ns

me

mb

ran

e p

ote

nti

al

(V) 500V/cm

375V/cm

250V/cm

Figure F.1: Transmembrane potential vs cell size at different external electric field. The

pulse type is exponentially decay, and pulse capacitance is 500 µF. Cell impermanent

dye, propidium iodide (PI), was used in the experiment. Each data point presents that at

least 50% of transfected cells at such size were observed.