153
Nanostructured Selenium for Biomedical Applications: from Theory to Practice By Phong A. Tran Brown University, 2010 A Dissertation Submitted in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy in the Department of Physics at Brown University Providence, Rhode Island May 2010

Nanostructured Selenium for Biomedical Applications: from

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Nanostructured Selenium for Biomedical Applications: from

Nanostructured Selenium for Biomedical Applications:

from Theory to Practice

By Phong A. Tran

Brown University, 2010

A Dissertation Submitted in Partial Fulfillment of the

Requirements for the Degree of Doctor of Philosophy

in the Department of Physics at Brown University

Providence, Rhode Island

May 2010

Page 2: Nanostructured Selenium for Biomedical Applications: from

© Copyright 2010 by Phong A. Tran

Page 3: Nanostructured Selenium for Biomedical Applications: from

iii

This dissertation by Phong A. Tran is accepted in its present form

by the Department of Physics as satisfying the

dissertation requirement for the degree of Doctor of Philosophy.

Date __________________ ___________________________

Thomas J. Webster, Advisor

Recommended to the Graduate Council

Date __________________ ___________________________

Thomas J. Webster, Reader

Date __________________ ___________________________

James M. Valles Jr., Reader

Date __________________ ___________________________

Derek Stein, Reader

Approved by the Graduate Council

Date _____________ __________________________________

Sheila Bonde, Dean of the Graduate School

Page 4: Nanostructured Selenium for Biomedical Applications: from

iv

Curriculum Vitae

Phong A. Tran

Address: 38 East Street, Floor 3, Providence, RI 02906 Phone: (401) 855-4925 Email: [email protected] Date of birth: 01/04/1981 Place of birth: Bacgiang, Vietnam

EDUCATION Brown University, Providence, RI, USA 2005 - 2010 Ph.D. in Physics Area of focus: Nanostructured selenium for anti-cancer, anti-bacterial applications Advisor: Professor Thomas J. Webster, Division of Engineering and Department of Orthopedics, Brown University Co-Advisor: Professor Derek Stein, Department of Physics, Brown University Brown University M.Sc. in Electrical Sciences and Computer Engineering 2005 - 2009 Hanoi University of Technology, Hanoi, Vietnam 2000 - 2004 B.Sc. in Engineering Physics Honor Training Program Area of focus: Magnetic materials

HONORS AND AWARDS STAR (Student Travel Achievement Recognition) Honorable Mention, Society for Biomaterials (2010) Dissertation Fellowship, Brown University (2009-2010) Research image highlighted on front cover of Journal of Materials Chemistry, 19 (18) (2009) Invited instant insight: Bone repair breakthrough. Chemical Technology (http://www.rsc.org/Publishing/ChemTech/Volume/2009/04/bone_repair_breakthrough.asp) (2009) Paper (“Enhanced osteoblast adhesion on nanostructured selenium compacts for anti-cancer orthopedic applications,” International Journal of Nanomedicince, 3, 391-396) in the top 5% most downloaded for all articles published from that time in the journal (2008) Vietnam Education Foundation Fellowship. 1 of 51 selected from approximately 2000 applicants (2005 - 2010) Conference Travel Awards, awarded by the Vietnam Education Foundation (2007, 2008) Brown University Graduate School Fellowships (2005, 2006) Award for best senior research project, Hanoi University of Technology (2004) Awarded Highest Honors for Honor’s Thesis, Hanoi University of Technology (2004)

Page 5: Nanostructured Selenium for Biomedical Applications: from

v

‘Rencontres du Vietnam’ scholarships (awarded to less than 1% of physics student body) (2001, 2002) Bronze Medal in National Mechanical Olympiad, Vietnam (2001) Hanoi University of Technology undergraduate scholarships (2000 - 2004)

RESEARCH EXPERIENCE Lab of Nanomedicine, Brown University Research Assistant 2006 - 2010 Conceptualize idea of combining chemopreventive properties of selenium and

mechanical properties of conventional orthopedic materials to create composites for anti-cancer orthopedic applications

Fabricate anti-cancer, anti-bacterial biomaterials by three methods: (i) pressing selenium particles into compacts, (ii) coating selenium nanoclusters on materials and (iii) synthesizing selenium nanoparticles

Characterize material structure and surface chemistry by scanning electron microscopy (SEM), transmission electron microscopy (TEM), X-ray photoelectron spectroscopy (XPS), atomic force microscopy (AFM), energy dispersive X-ray spectroscopy (EDS), contact angle measurement, dynamic light scattering (DLS), Zeta potential, and X-ray diffraction (XRD)

Investigate functions of primary osteoblast cells on materials in vitro using cell assays (ELISA, CyQUANT, cell viability (MTT), Live/Dead, ALP activities, calcium deposition and protein synthesis)

Investigate functions of osteosarcoma cells on the materials using assays (Annexin V cell apoptosis, cell cycle arrest), fluorescence microscopy and flow cytometry

Investigate bacterial functions and formation of bacterial biofilms on the materials using Live/Dead assays, confocal microscopes, fluorescence microscopes, microplate readers and SEM

Develop fluorescence microscopy method and image processing programs to track the proliferation of primary osteoblast cells and osteosarcoma cells in co-culture

Create programs in ImageJ and Matlab to eliminate manual cell counting, reduce image processing, data processing time and increase the accuracy in determining cell count and cell spreading. The programs are being used across several labs

Design a multi-pressing mold that can be used to fabricate simultaneously multiple samples from powder or particles. The mold helped reduce sample preparation time approximately 12 times

Train 7 lab members on the mammalian co-culturing system and image/data processing programs

Modify the Cassie-Baxter model of contact angle on surface to quantify contribution of nanometer roughness to air-pocket formation during contact angle measurement

Develop program in Matlab to simulate adsorption of fibronectin on nanometer rough surfaces. The simulation results fit well with experimental results

Lab of Amorphous and Nanocrystalline Materials, Hanoi University of Technology Senior Research Project 2003, 2004 Designed and setup a system to measure giant magneto resistance (GMR) effects Measured GMR effects and hysteresis loops in rapidly quenched Cu-Co alloys

Page 6: Nanostructured Selenium for Biomedical Applications: from

vi

Used Langevin theory of paramagnetism to estimate particle sizes from the measured hysteresis loops

Trained 3 lab members on using the GMR measuring system Supervised a group of two junior students to design a GMR-based magnetic sensor

PAPERS/BOOK CHAPTERS 1. P. A. Tran, L. Sarin, R. H. Hurt and T. J. Webster. Differential effects of nanoselenium doping on healthy and cancerous osteoblasts in co-culture on titanium. International Journal of Nanomedicine, (in press). 2. P. A. Tran, L. Sarin, R. H. Hurt and T. J. Webster. Titanium surfaces with adherent selenium nanoclusters as a novel anti-cancer orthopedic material. Journal of Biomedical Materials Research part A, online ahead of print, DOI: 10.1002/jbm.a.32631 (2009). 3. P. A. Tran, L. Zhang and T. J. Webster. Carbon nanofibers and carbon nanotubes in regenerative medicine. Advanced Drug Delivery Reviews, 61, 1097-111 (2009). 4. P. A. Tran, L. Sarin, R. H. Hurt and T. J. Webster. Opportunities for nanotechnology-enabled bioactive bone implants. Journal of Materials Chemistry, 19, 2653–2659 (2009). 5. P. A. Tran and T. J. Webster. Nanotechnologies for cancer diagnostics and treatment. In Charles J. Dixon and Ollin W. Curtines, editors, Nanotechnology: nanofabrication, patterning, and self assembly. Nova Science Publisher (2009). 6. P. A. Tran and T. J. Webster. Enhanced osteoblast adhesion on nanostructured selenium compacts for anti-cancer orthopedic applications. International Journal of Nanomedicince, 3, 391-396 (2008). 7. N. N. Hoang, P. A. Tran, et al. The influence of heat treatment on magnetoresistance effect in granular Cu-Co alloys prepared by rapid quenching. Adv. In Tech. of Mat. and. Mat. Proc.J (ATM).Vol.6 (1), 83-86 (2005).

CONFERENCE PROCEEDINGS 1. P. A. Tran, L. Sarin, R. H. Hurt and T. J. Webster. Selenium nanocluster coatings: transforming current orthopedic materials into inhibiting bone cancer. Materials Science Forum, 638-642: 718-723, Trans Tech Publications, Switzerland (2010). 2. P. A. Tran, E. Taylor, L. Sarin, R. H. Hurt and T. J. Webster. Selenium nanocluster coatings for anti-cancer, anti-bacterial orthopedic applications. Proceedings of 2009 AIChE Annual Meeting, Nashville, TN (in press). 3. P. A. Tran, L. Sarin, R. H. Hurt and T. J. Webster. Transforming orthopedic biomaterials into bone cancer inhibiting implants: The role of selenium nanoclusters. in Materials in Tissue Engineering, edited by Webster, T.J. (Mater. Res. Soc. Symp. Proc. Volume 1136E, Warrendale, PA (2009).

CONFERENCE PRESENTATIONS 1. P. A. Tran, E. Taylor, L. Sarin, R. H. Hurt and T. J. Webster. Novel anti-cancer, anti-bacterial coatings for biomaterial applications: selenium nanoclusters. 2009 Materials Research Society Meeting, Boston, MA (2009) 2. P. A. Tran, L. Sarin, R. H. Hurt and T. J. Webster. Imparting anti-cancer properties to orthopedic materials: The role of selenium nanoclusters. Society for Biomaterials 2009 Annual Meeting, San Antonio, TX (2009).

Page 7: Nanostructured Selenium for Biomedical Applications: from

vii

3. P. A. Tran, L. Sarin, R. H. Hurt and T. J. Webster. Selenium nanocluster coatings for anti-cancer orthopedic applications. 35th Northeast Bioengineering Conference, Boston, MA (2009). 4. P. A. Tran, L. Sarin, R. H. Hurt and T. J. Webster. Promising orthopedic materials for bone cancer patients: titanium coated with selenium nanoclusters. 34th Northeast Bioengineering Conference, Providence, RI (2008). 5. P. A. Tran, L. Sarin, R. H. Hurt and T. J. Webster. Nanoselenium cluster coatings for anti-cancer orthopedic applications. Society for Biological Engineering’s 4th International Conference on Bioengineering and Nanotechnology Dublin, Ireland (2008). 6. P. A. Tran and T. J. Webster. Promising anti-cancer orthopedic material: Nanostructured selenium. Biomedical Engineering Society Fall Meeting, Los Angeles, CA (2007). 7. P. A. Tran and T. J. Webster. Novel anti-cancer orthopedic materials: Nanostructured selenium. 33rd Northeast Bioengineering Conference, Stony Brook, NY (2007).

COMPUTER SKILLS ImageJ, Image Pro Plus, Matlab, Assembly, Pascal, Microsoft Excel, Microsoft Word, Microsoft PowerPoint

Page 8: Nanostructured Selenium for Biomedical Applications: from

viii

ACKNOWLEDGEMENTS

I am greatly grateful to my advisor, Professor Thomas J. Webster who has guided

me and provided financial support to me through my academic career at Brown

University. I am also grateful to my co-advisor, Professor Derek Stein who has given me

advice on my research and also on the career track in general. I deeply appreciate

Professor James M. Valles, Jr.’s interest in my research and serving in my preliminary

exam and thesis defense committees, helping me to finish my graduate career. I would

like to thank all members in the Laboratory of Nanomedicince Research for their support

and help during my time in the laboratory.

Most of all, I thank my parents who gave birth to me, raise me, love me and guide

me to the career path today. I also thank other members of my family, my sister, my

brother, who are always a constant source of inspiration for me to overcome challenges in

my studies.

I thank the Vietnam Education Foundation for their partial financial support for

my studies at Brown University.

Page 9: Nanostructured Selenium for Biomedical Applications: from

ix

TABLE OF CONTENTS PAGE

LIST OF TABLES ............................................................................................................ xi LIST OF ILLUSTRATIONS............................................................................................ xii CHAPTER 1. INTRODUCTION ....................................................................................... 1

1.1. Bone cancers ............................................................................................................ 1 1.2. Implant bacterial infection ....................................................................................... 1 1.3. Selenium as an anti-cancer, anti-bacterial material ................................................. 4 1.4. Focus of thesis.......................................................................................................... 6

CHAPTER 2. SELENIUM NANOCLUSTER COATINGS ON METALS FOR ANTI-CANCER ORTHOPEDIC MATERIALS .......................................................................... 7

2.1. Introduction.............................................................................................................. 7 2.1.1. Problems with currently used orthopedic implants........................................... 7 2.1.2. The role of nano-structured surfaces on orthopedic implants........................... 8 2.1.3. Nano-structured selenium coatings: a novel approach of using selenium to create anti-cancer, anti-bacterial biomaterials ............................................................ 8

2.2. Materials and methods ............................................................................................. 9 2.2.1. Materials ........................................................................................................... 9 2.2.2. Material characterization ................................................................................ 11 2.2.3. In vitro biological assays for uncoated and selenium coated metallic substrates................................................................................................................................... 14 2.2.4. Fibronectin adsorption on uncoated and selenium coated titanium substrates19 2.2.5. Statistical analysis........................................................................................... 20

2.3. Results.................................................................................................................... 20 2.3.1. Surface characterization.................................................................................. 21 2.3.2. Cell adhesion and proliferation....................................................................... 28 2.3.3. Healthy osteoblast differentiation ................................................................... 35 2.3.4. Selenium release ............................................................................................. 37 2.3.5. Effects of selenium released into culture media on healthy and cancerous cells................................................................................................................................... 38 2.3.6. Intracellular thiol assays ................................................................................. 40 2.3.7. Fibronectin adsorption on uncoated and selenium-coated titanium substrates41

2.4. Results on stainless steel samples .......................................................................... 45 2.4.1. Material characterization ................................................................................ 45 2.4.2. Cell experiments on stainless steel substrates................................................. 48

2.5. Conclusions and summary ..................................................................................... 52 CHAPTER 3. COARSE - GRAINED MONTE CARLO COMPUTER SIMULATION OF FIBRONECTIN ADSORPTION ON NANOMETER ROUGH SURFACES .......... 54

3.1. Introduction............................................................................................................ 54 3.2. Protein adsorption onto biomaterials ..................................................................... 55 3.3. Roles of rough surfaces.......................................................................................... 58 3.4. Simple method to generate rough surfaces with pre-defined RMS roughness...... 59 3.5. RSA model of fibronectin adsorption on uncoated and selenium coated titanium surfaces ......................................................................................................................... 64 3.6. Summary and conclusions ..................................................................................... 76

Page 10: Nanostructured Selenium for Biomedical Applications: from

x

CHAPTER 4. A MODIFIED CONTACT ANGLE MODEL TO FIT EXPERIMENTAL RESULTS ......................................................................................................................... 78

4.1. Introduction: role of contact angles in biomaterials .............................................. 78 4.2. Contact angle and Young’s equation ..................................................................... 81 4.3. Wenzel model versus Cassie-Baxter model on contact angles.............................. 82 4.4. The needs to modify the Wenzel model and Cassie-Baxter model to fit experimental results ...................................................................................................... 88 4.5. Modification of the Cassie-Baxter model .............................................................. 92 4.6. Summary and conclusions ..................................................................................... 96

CHAPTER 5. SELENIUM COATINGS ON POLYMERS FOR ANTI-INFECTION BIOMATERIAL APPLICATIONS.................................................................................. 97

5.1. Introduction............................................................................................................ 97 5.2. Materials and methods ........................................................................................... 99

5.2.1. Materials ......................................................................................................... 99 5.2.2. Material characterization .............................................................................. 100 5.2.3. In vitro assays for the polymeric substrate-bacterial experiments................ 101 5.2.4. Statistical analysis......................................................................................... 103

5.3. Results and discussion ......................................................................................... 103 5.3.1. SEM and EDS............................................................................................... 103 5.3.2. S. aureus response on uncoated and selenium-coated polymeric substrates 107 5.3.3. Thiol assays................................................................................................... 110

5.4. Discussion ............................................................................................................ 111 5.5. Conclusions.......................................................................................................... 112

CHAPTER 6. SELENIUM NANOPARTICLES FOR ANTI-BACTERIAL APPLICATIONS ............................................................................................................ 114

6.1. Introduction.......................................................................................................... 114 6.2. Materials and methods ......................................................................................... 115

6.2.1. Material synthesis ......................................................................................... 115 6.2.2. Material characterization .............................................................................. 115 6.2.3. Bacteria assays .............................................................................................. 116 6.2.4. Superoxide anion assays ............................................................................... 118 6.2.5. Statistical analysis......................................................................................... 119

6.3. Results.................................................................................................................. 119 6.3.1. TEM, DLS and XPS results on size distribution and oxidation state of selenium nanoparticles............................................................................................ 119 6.3.2. Effects of selenium nanoparticles on S. aureus in solution .......................... 122 6.3.3. Intracellular thiol assays ............................................................................... 122 6.3.4. Superoxide assays ......................................................................................... 124

6.4. Summary and conclusions ................................................................................... 125 CHAPTER 7. CONCLUSIONS ..................................................................................... 127

Page 11: Nanostructured Selenium for Biomedical Applications: from

xi

LIST OF TABLES PAGE Table 2.1. Reagent volumes used in the colloidal synthesis of Se nanoclusters in the

presence of various base substrates............................................................................. 10 Table 2.2. Fitting parameter and the goodness of fitting (R2) for growth curve and

calculated doubling time of healthy osteoblasts cultured on uncoated and selenium-coated titanium substrates ........................................................................................... 33

Table 2.3.Fitting parameter and goodness of the fitting (R2) for growth curve and calculated doubling time of cancerous osteoblasts cultured on uncoated and selenium-coated titanium substrates. .......................................................................................... 35

Table 3.1. Convergence property of the proposed method to generate surface High-nSe-Ti with a defined RMS roughness. Dimensionless simulated roughness, Ro, is the square root of variance of the heights of all points on the simulated surface. Simulated roughness is then determined as R=RoDo (in this case, Do=9nm)............................... 63

Table 3.2. Parameters for simulation (a) calculated from experimental RMS roughness as described in section 3.5 and the RMS roughness of the simulated surfaces. The simulated surfaces resemble the experimental surfaces quite well in terms of RMS roughness. ................................................................................................................... 69

Table 4.1. The selenium surface fraction (ρ, determined from SEM images), relative surface roughness ratio (r, normalized by that of the uncoated substrates), the relative root mean square roughness (R, normalized to that of uncoated substrates), and cos(θ), of the substrates of interest to the studies. ...................................................... 90

Table 6.1. Superoxide anion assay reagent mixtures...................................................... 118

Page 12: Nanostructured Selenium for Biomedical Applications: from

xii

LIST OF ILLUSTRATIONS PAGE Figure 1.1. A scanning electron microscopy image of a Staphylococcus biofilm on the

inner surface of a catheter showing bacteria and polymeric matrix (reprinted with permission from [11]). .................................................................................................. 3

Figure 2.1. Image of DSA100 (Updated Version of DSA-10, Kruss, Germany: Image Adopted from www.kruss.com). Capability of equipment: The range of contact angle measurement is between 0 and 180° with a resolution of ±0.1 and a range of surface tension between 10-2 and 100mJ/m2, with a resolution of 0.01mJ/m2. ....................... 13

Figure 2.2. Representative SEM images of: (A) uTi; (B) Low-nSe-Ti; (C) Medium-nSe-Ti and (D) High-nSe-Ti. The surface area coverage of selenium on the selenium coated substrates were determined (using an image processing program, ImageJ) to be 2.7%, 5.1% and 7.5% for Low-nSe-Ti, Medium-nSe-Ti and High-nSe-Ti, respectively. ................................................................................................................ 21

Figure 2.3. SEM image, taken at a 45° tilt, of High-nSe-Ti showing the hemispherical shape of the selenium nanoclusters on the titanium surface. ...................................... 22

Figure 2.4. Morphology of a selenium-coated titanium surface appeared the same before (A) and after sonication (B) and UV radiation (C). Bars=500 nm. ............................ 23

Figure 2.5. All coated substrates showed peaks of selenium that were not detected in uncoated substrate....................................................................................................... 24

Figure 2.6. XPS profiles of High-nSe-Ti showing the characteristic binding energy peak for elemental selenium at 55.2 eV. ............................................................................. 24

Figure 2.7. Representative AFM images and line analysis of uTi (A), Low-nSe-Ti (B), Medium-nSe-Ti (C) and High-nSe-Ti (D). Nano-scale roughness was created by the selenium coatings and increased with increasing selenium coating density............... 26

Figure 2.8. RMS of the substrates increased with increasing selenium coating density. Data = mean ± standard error of the mean; N=3; * p<0.05 compared to all other substrates; ** p<0.01 compared to Medium-nSe-Ti and High-nSe-Ti....................... 27

Figure 2.9. Water contact angles on the uncoated and coated Ti substrates. Contact angles increased on the substrates coated with selenium nanoclusters. Data = mean ± standard error of the mean; N=3; * p<0.05 compared to all the coated substrates. There was no significant difference among the contact angles on the coated substrates. ........................................................................................................ 28

Figure 2.10. Increased healthy osteoblast densities after 4 hrs, 1 day and 3 days. Data = mean ± standard error of the mean; N=3, * p<0.05 compared to Medium-nSe-Ti (compared at same time period), ** p<0.05 compared to uTi (compared at same time period), *** p< 0.05 compared to Low-nSe-Ti (compared at same time period), # p<0.05 compared to uTi (compared at same time period), & p =0.06 compared to Low-nSe-Ti. (compared at same time period). ........................................................... 29

Figure 2.11. Cancerous osteoblast densities after 4 hrs, 1 day, and 3 days. Decreased cancerous osteoblast densities on selenium coated titanium substrates after 4 hrs, 1 day and 3 days. Data = mean ± standard error of the mean; N=3, * p<0.05, ** p< 0.01 compared to High-nSe-Ti. (Compared at same time period); # p<0.1 compared to Medium-nSe-Ti. (Compared at same time period). There are no significant differences in cell densities among substrates after 1 day. ......................................... 30

Page 13: Nanostructured Selenium for Biomedical Applications: from

xiii

Figure 2.12. Representative fluorescence microscopy images of healthy osteoblasts on: (A) uTi; (B) Low-nSe-Ti; (C) Medium-nSe-Ti and (D) High-nSe-Ti after 1 day (Magnification = 10X). Scale bars = 200 μm. ............................................................ 31

Figure 2.13. Representative fluorescence microscopy images of cancerous osteoblasts on: (A) uTi; (B) Low-nSe-Ti; (C) Medium-nSe-Ti and (D) High-nSe-Ti after 3 days (Magnification = 20X). Scale bars = 100 μm. ............................................................ 32

Figure 2.14. Fitting exponential growth curves to experiment data for densities of healthy osteoblasts cultured on uncoated titanium and on selenium-coated titanium substrates. Growth curves of healthy osteoblasts cultured on higher selenium coating densities had large slopes........................................................................................................... 33

Figure 2.15. Fitting exponential growth curves to experiment data for densities of cancerous osteoblasts cultured on uncoated titanium and selenium-coated titanium substrates. Curves of cancerous osteoblasts cultured on lower selenium coating densities had large slopes with the uncoated titanium substrates having the largest slope. ........................................................................................................................... 34

Figure 2.16. Increased alkaline phosphatase (ALP) activity of osteoblasts on High-nSe-Ti compared to all other substrates. Data = mean ± standard error of the mean; N=3, * p<0.01 compared to uTi, Low-nSe-Ti and Medium-nSe-Ti.................................... 36

Figure 2.17. Extracellular calcium deposition by osteoblasts on uTi, Low-nSe-Ti, Medium-nSe-Ti and High-nSe-Ti after 14 days. Data = mean ± standard error of the mean; N=3, * p<0.05 compared to uTi. ...................................................................... 37

Figure 2.18. Decreased thiol content of healthy osteoblasts cultured in supernatant from High-nSe-Ti compared to uTi. Data = mean ± standard error of the mean; N=3; * p<0.1 ........................................................................................................................... 40

Figure 2.19. Decreased thiol content of cancerous osteoblasts cultured in supernatant from High-nSe-Ti compared to uTi. Data = mean ± standard error of the mean; N=3; * p<0.05. ..................................................................................................................... 41

Figure 2.20. Increased fibronectin adsorption from DMEM onto Medium-nSe-Ti and High-nSe-Ti substrates. DMEM (supplemented with 10% FBS and 10% P/S) was estimated to have fibronectin concentration of 3µg/mL. Each sample was immersed in 0.5 mL of DMEM solution for 24 hrs under standard condition (37oC, 5% CO2, 95% humidified air) and amount of absorbed fibronectin was determined via optical density of ABTS solution using ELISA method. A standard curve relating optical density to amount of fibronectin was used to determine fibronectin concentration from optical densities of ABTS solution. Finally, amount of fibronectin was normalized to geometrical surface areas of the substrates. Data = mean ± standard error of the mean; N=3, * p<0.05 compared to uTi, ** p<0.05 compared to uTi and Low-nSe-Ti. ................................................................................................................ 43

Figure 2.21. Increased fibronectin adsorption from fibronectin solution (in PBS) on Medium-nSe-Ti and High-nSe-Ti substrates. Each sample was immersed in 0.5 mL of 5 µg/mL fibronectin solution for 24 hrs under standard condition (37oC, 5% CO2, 95% humidified air) and amount of absorbed fibronectin was determined via optical density of ABTS solution using ELISA method. A standard curve relating optical density to amount of fibronectin was used to determine fibronectin concentration from optical densities of ABTS solution. Finally, amount of fibronectin was normalized to geometrical surface areas of the substrates. Data = mean ± standard

Page 14: Nanostructured Selenium for Biomedical Applications: from

xiv

error of the mean; N=3, * p<0.05 compared to uTi and Low-nSe-Ti. ** p=0.055 compared to Medium-nSe-Ti...................................................................................... 44

Figure 2.22. SEM images of stainless steel substrates: (A) uncoated; (B), (C), and (D) coated with Se concentration in the final solution which equaled 0.42 mM, 0.83 mM ,and 1.25 mM, respectively. Some Se nanoclusters are indicated by arrows. These clusters are approximately 80nm in diameters. Bars = 1µm....................................... 46

Figure 2.23. Water contact angles on stainless steel substrates. Selenium nanocluster coating makes the stainless steel substrates become more hydrophobic. ................... 47

Figure 2.24. Maintained healthy osteoblast densities on selenium-coated stainless steel substrates compared to the uncoated substrate after 3 days of culturing. Data = mean ± standard error of the mean; N=3. ....................................................... 48

Figure 2.25. Representative fluorescence images of healthy osteoblast densities on uSS (A), Low-nSe-SS (B), Medium-nSe-SS (C) and High-nSe-SS (D). Magnification = 10X.............................................................................................................................. 49

Figure 2.26. Decreased cancerous osteoblast densities after 3 days of culturing on Se coated stainless steel. Cell densities on Medium-nSe-SS and High-nSe-SS were significantly less than those on uSS and Low-nSe-SS. * p<0.05 compared to uSS and Low-nSe-SS; ** p<0.01 compared to Medium-nSe-SS. Data = mean ± standard error of the mean; N = 3. ..................................................................................................... 50

Figure 2.27. Representative fluorescence images of cancerous osteoblast densities on uSS (A), Low-nSe-SS (B), Medium-nSe-SS (C) and High-nSe-SS (D). Magnification = 20X.............................................................................................................................. 51

Figure 3.1. Schematic structure of fibronectin showing two monomers linked by disulfide bonds (Reprinted with permission from [72]). ........................................................... 57

Figure 3.2. Representative regions on the four rough surfaces generated in this study with the RMS roughness values equal to experimental values (unit lengh Do=9nm, N=500). Colors are to aid in visualization only.......................................................... 64

Figure 3.3. Gallery of transmission electron microscopy images of individual fibronectin molecules freeze-dried in vacuum and shadowed by tungsten-tantalum. (Freeze-drying technique was used to preserve the native structure of fibronectin and tungsten-tantalum shadowing was used to give contrast for imaging). Most fibronectin molecules had elongated shapes with the length of ~ 15 nm and the width of ~9 nm. Magnification is 300,000. (Adapted with permission from [98]) . ........... 66

Figure 3.4. A 2D picture of fibronectin adsorbing onto a rough surface. Fibronectin was modeled as a dimmer consisting of 2 monomers and the surface was modeled as consisting of columns. Each point on the surface was defied by two coordinates (i,j). Fibronectin can adsorb on all the "unoccupied" sites (in this picture, the right and the left sides of point (i,j) are unoccupied, the top and bottom sides of point (i,j) are occupied)..................................................................................................................... 67

Figure 3.5. Examples of the simulation of protein adsorption onto a rough surface. For the case of the right side of the point specified by (i,j) (the bold side), there are two possible directions the protein can adsorb: horizontal direction (a) or vertical direction (b) (Note, horizontal adsorption to the left of the bold side is prohibited due to steric constraints). In this case, configuration (a) will have 4 contacts while (b) will have only 2 contacts upon adsorption, therefore, configuration (a) is more favorable for protein adsorption. For the case of the point on the far right (bold side), both the (c)

Page 15: Nanostructured Selenium for Biomedical Applications: from

xv

and the (d) configuration will have the same number of contacts (3 contacts) upon adsorption, therefore, the protein will adsorb following either (c) or (d) with equal probability. .................................................................................................................. 68

Figure 3.6. Overhang is not allowed. Overhang would be immediately shifted downward as indicated by the dashed arrows............................................................................... 69

Figure 3.7. A snapshot of simulation process in 2D with N=500 and M=25000. Blue columns represent proteins. Grey columns represent rough surfaces. The adsorption did not reach saturation yet as some areas on the surface are still unoccupied. ......... 71

Figure 3.8. Simulation results of fibronectin saturated adsorption on all four substrates of interest. The protein amount was normalized to a monolayer of protein on a perfectly flat surface with the same geometrical surface area N x N. Substrates were generated to have experimental RMS roughness values. Protein adsorption simulation was implemented with the size of the substrates N x N=500x500 and the number of simulation M=6,250,000. Data = mean ± standard error of the mean (n=3 runs). ..... 72

Figure 3.9. Comparison of the increase in the amount of fibronectin adsorption (expressed as times increase compared to uTi) between simulation results and experimental results on uncoated and selenium coated titanium substrates. .............. 73

Figure 3.10. Simulation and experiment results on the amount of adsorbed fibronectin normalized by geometrical surface area. The simulation values of normalized fibronectin adsorption are greater than experimental values. ..................................... 74

Figure 3.11. Protein adsorption on a flat surface and a rough surface. Assuming the same "binding site" densities, the protein density (per unit surface area) d1 and d2 should be the same. ..................................................................................................................... 75

Figure 3.12. Computer simulation results on fibronectin density expressed as microgram of fibronectin per cm2 of the surface area. Surfaces with higher RMS roughness had lower fibronectin density. ........................................................................................... 75

Figure 3.13. An example illustrating that increased surface area does not necessarily lead to equal increase in protein adsorption. In this example, the rough surface area increases ~67% but the amount of fibronectin increases only 25%. .......................... 76

Figure 4.1. Atomic and ribbon chains of an extended human fibronectin molecule. Inset shows that hydrophilic parts are blue colored regions while hydrophobic parts are red colored in modules (Reprinted with permission from [100]). .................................... 80

Figure 4.2. Contact angle, θ, of a liquid on a solid surface depends on the interaction of

the liquid with the solid, liquid with vapor and solid with vapor. SL , SV , and LV are interfacial energies at the solid-liquid interface, solid-vapor interface and liquid-vapor interface, respectively. ................................................................................................ 81

Figure 4.3.Schematic demonstration of the derivation of Young's equation. When the solid-liquid interface is displaced by dx, the increase in liquid-vapor interface is dx.cos(θ)...................................................................................................................... 82

Figure 4.4. Schematic demonstration of the derivation of the Wenzel equation. When the solid-liquid interface is displaced by dx, the increase in liquid-vapor interface is dx.cos(θ), while the actual liquid-solid interface increase associated with this displacement is r.dx. ................................................................................................... 84

Figure 4.5. Homogenous wetting assumption in the Wenzel model. The liquid-solid interface is assumed to follow a solid surface. ........................................................... 85

Page 16: Nanostructured Selenium for Biomedical Applications: from

xvi

Figure 4.6. A schematic demonstration of the derivation of the Cassie-Baxter model. When the solid-liquid interface increases by dx, the liquid-vapor interface increases by dx.cos(θ)................................................................................................................. 86

Figure 4.7. The Cassie-Baxter model of heterogeneous wetting. In this case, the liquid is considered to contact a surface consisting of the solid material and air. The width of spikes as well as the spacing between them is very small compared to the size of the liquid drop. The scale used in this picture is to aid in visualization. .......................... 87

Figure 4.8.Water contact angles on uncoated titanium and titanium coated with increasing selenium nanocluster densities. Data = mean ± standard error of the mean, n=5. ..... 89

Figure 4.9.The Cassie-Baxter model curve (dashed line) and the experimental curve (solid line) of the cosine of water contact angles on uncoated titanium substrates and titanium substrates coated with various amount of selenium nanoclusters. ............... 91

Figure 4.10. Possible mechanisms for the plateau in contact angles on the substrates with different selenium coating densities. Increasing coating density of selenium nanoclusters increased roughness that resulted in both increases in the roughness factor (which is the actual surface area to geometric area ratio) and the increase in the formation of air-pockets. The former made contact angles decrease while the latter made contact angles increase. The opposite contribution of the roughness factor and air-pockets is likely the mechanism for similar contact angles for the three selenium coating densities on titanium in here. ......................................................................... 92

Figure 4.11. Schematic representation of the modified Cassie-Baxter model on the wetting process. A drop of liquid makes contact to all three materials: titanium, selenium nanoclusters, and air when placed on a selenium-titanium composite surface. ........................................................................................................................ 93

Figure 5.1. SEM images of uncoated (left panel) and selenium (Se)-coated substrates (right panel). Arrows indicate selenium nanoclusters. The selenium nanoclusters had sizes ranging from approximately 80nm to 200nm and were uniformly coated on the substrates................................................................................................................... 104

Figure 5.2. EDS spectra (right panel) of an uncoated region (top) and a coated region (bottom) on a selenium-coated silicone substrate. Selenium peaks were detected in the coated region demonstrating that the clusters were selenium. Peaks of gold were from the sputter-coated layer. ............................................................................................ 105

Figure 5.3. EDS spectra (right panel) of an uncoated region (top) and a coated region (bottom) on a selenium-coated PU substrate. Selenium peaks were detected in the coated region demonstrating that the clusters were selenium. Peaks of gold were from the sputter-coated layer. ............................................................................................ 106

Figure 5.4. EDS spectra (right panel) of an uncoated region on selenium-coated PVC (top) and a coated region on selenium-coated PVC (bottom). Selenium peaks were detected in the selenium-coated PVC demonstrating that the clusters were selenium. Peaks of gold were from the sputter-coated layer..................................................... 107

Figure 5.5. Representative SEM images showing reduced S. aureus colonization on selenium-coated PU compared to uncoated PU after 8 hrs of inoculation. Arrows indicate cells.............................................................................................................. 108

Figure 5.6. Representative SEM images showing reduced S. aureus colonization on selenium-coated PVC compared to uncoated PVC. Arrows show bacteria. ............ 108

Page 17: Nanostructured Selenium for Biomedical Applications: from

xvii

Figure 5.7. Representative SEM images showing reduced S. aureus colonization on selenium-coated compared to uncoated silicone. Arrows indicate cells................... 108

Figure 5.8. Decreased S. aureus densities on polymeric substrates coated with selenium nanoclusters. Data = mean ± standard error of the mean, N=3; *p<0.05 compared to the uncoated substrate (compared in the same material group), ** p<0.05 compared to silver-coated PVC. .................................................................................................... 109

Figure 5.9. Representative SEM images of S. aureus colonization on silver-coated and Se-coated PVC. Bacterial colonization on Se-coated PVC was reduced compared to silver-coated PVC after 8 hrs of incubation.............................................................. 110

Figure 5.10. Decreased intracellular thiol level in S. aureus cultured in the supernatant from selenium-coated PVC compared to control. Data = mean ± standard error of the mean; N=3; * p<0.05 compared to the control. ........................................................ 111

Figure 6.1. TEM image of selenium nanoparticles stabilized in BSA and dispersed in water. Selenium nanoparticles had average sizes of approximately 100 nm............ 120

Figure 6.2. Size-distribution profile of selenium nanoparticles in solution as measured by the dynamic light scattering technique. Particle sizes centered around 100 nm....... 121

Figure 6.3. XPS profile of selenium nanoparticles. Peak at 55.2 eV confirmed a zero oxidation state of selenium. ...................................................................................... 121

Figure 6.4. Inhibited growth of S. aureus in the presence of selenium nanoparticles at all three selenium nanoparticles concentrations: 12.5 µg/mL, 6.2 µg/mL and 3.1 µg/mL at 4 hrs, 12 hrs and 24 hrs. Data = mean ± standard error of the mean, N=3. *p<0.05 compared to bacteria treated with 6.2 µg Se/mL, 3.1 µg Se/mL and control (0 µg Se/mL) (compared at same time period); ** p<0.05 compared to bacteria treated with 3.1 µg Se/mL and control (0 µg Se/mL) (compared at same time period); ***p<0.05 compared to control (compared at same time period); #p<0.05 compared to control (compared at same time period). Dose-dependent inhibition was observed at 4hrs and 12hrs.......................................................................................................................... 122

Figure 6.5. Decreased intracellular thiol levels in S. aureus cultured with selenium nanoparticles compared to the control (i.e., S. aureus cultured in TSB without selenium nanoparticles). Bacteria were cultured at a density of 50,000 cells/mL for 4 hrs in either TSB or TSB added with selenium nanoparticle (at a concentration of 3.1µg/mL). Data = mean ± standard error of the mean; N=3; * p< 0.05........... 123

Figure 6.6. Superoxide anion assays for selenium particles (nSe). No significant difference was found among the test tubes. Selenium nanoparticles were mixed with GSH and luminol was used as the probes for superoxide anions. Superoxide anions oxidize luminol in a reaction that produces photons. An enhancer was used to amplify the chemiluminescence. Data = mean ± standard error of the mean, N=3. .............. 125

Page 18: Nanostructured Selenium for Biomedical Applications: from

1

CHAPTER 1. INTRODUCTION

1.1. Bone cancers

It is estimated that 2,380 individuals (1,270 men and 1,110 women) will be

diagnosed with bone and joint cancer and 1,470 individuals will die from primary bone

and joint cancer in 2008 in the U.S. [1]. Primary bone cancer is rare as usually bone

cancer is a result of the spread of cancer from other organs (such as the lungs, breasts and

the prostate [2]). Because many deaths are officially attributed to the original cancer

source, the true numbers of bone-cancer deaths are underreported. A common technique

to treat bone cancer is the surgical removal of the cancerous tissue followed by insertion

of an orthopedic implant to restore patient function. Sometimes cancerous cells are not

completely removed, therefore, the remaining cancerous cells will proliferate and cancer

can reoccur at the implant site. In these cases, it would clearly be beneficial to have

implants specifically designed to not only promote healthy bone tissue growth but also to

prevent the reoccurrence of bone cancer.

1.2. Implant bacterial infection

Bacterial infection is a problem for all implants. For example, infection is one of

the most common causes for failure of a hip implant, responsible for 14% of the total

number of revision surgeries [3]. Another example of implant infection relates to

catheters which is the most common and complicated problem associated with catheter

Page 19: Nanostructured Selenium for Biomedical Applications: from

2

usage. Infections are the most serious complication of tunneled dialysis catheters,

resulting in serious systemic infections, including endocarditis, osteomyelitis, epidural

abscess, septic arthritis, and even death [4]. Infections lead to implant failure, extended

hospital stay, and additional treatment/surgeries. Bacteria infect up to 54% of all catheters

[5] and cause many serious complications including patient death. For example, catheter

infection is associated with a mortality rate of 12% to 25% among critically ill patients

[6]. Catheter-associated urinary tract infection (CAUTI) is the most common type

(accounting for 40%) of hospital-acquired infection (“nosocomial infection”) resulting in

serious complications such as bloodstream infection, and even death [7]. Each year, in

U.S. acute-care hospitals and extended-care facilities, CAUTI affects approximately 1

million patients who then will have increased institutional death rates [8]. Chronic

indwelling urinary catheters also increase the risk of infection, accounting for 80% of all

nosocomial urinary tract infections [9]. Significantly, 14 % of deaths in people

undergoing dialysis in 1996 were due to infection [5]. The cost incurred by infections in

the U.S. is nearly $11 billion annually [10].

Implant associated infections are difficult to treat because of biofilm formation.

After implantation, bacteria (from the patient’s own skin, hospital personnel or

equipment) quickly attach and adhere irreversibly to the implant surface, secreting a

polymeric-like substance (composed of mostly polysaccharides) and form a biofilm

which consists of bacteria and a polymeric-like matrix (Figure 1.1).

Page 20: Nanostructured Selenium for Biomedical Applications: from

3

Figure 1.1. A scanning electron microscopy image of a Staphylococcus biofilm on the inner surface of a catheter showing bacteria and polymeric matrix (reprinted with permission from [11]).

Bacteria in a biofilm can escape from the film and enter the blood, lungs, etc.,

causing serious problems. Biofilms tenaciously bind to surfaces. More importantly,

bacteria in biofilms are extremely resistant to antibiotic treatment due to the slow

transport of antibiotic molecules through the polymeric-like biofilm substance, altered

micro-environment within the biofilm and higher number of “persister” cells (cells that

are resistant to many types of stress) within the biofilm [compared to planktonic (free-

floating) cells] [12].

Among the most common pathogens found on infected implants are S.

epidermidis (comprising 24% of all bacteria for catheter infections and 15% for

orthopedic implant infections), S. aureus (20% for catheter infections, 35% for

orthopedic implant infections), and P. aeruginosa (25% for catheter infection) [11, 13,

14].

Page 21: Nanostructured Selenium for Biomedical Applications: from

4

1.3. Selenium as an anti-cancer, anti-bacterial material

Selenium belongs to the group of metalloids, which are neither fully metals nor

non-metals but share characteristics of both [15]. In the periodic table of elements,

selenium is in the same column with oxygen and sulfur, and in the same row with

germanium, arsenic and bromine. The outermost electron configuration of selenium is

4s23d104p4. Selenium is naturally found in humans and animals and selenoproteins are

important proteins in antioxidant defense systems, thyroid hormone metabolism and

redox control of cell reactions [16]. Animal studies have shown that selenium intake in

excess of the nutritional requirement can inhibit and/or retard carcinogenesis [17].

Moreover, studies have shown that people in areas of low soil selenium (lower than 0.05

ppm) and people with decreased plasma selenium levels (below 128 ng/ml) have higher

cancer incidence and/or cancer mortality [18, 19]. High levels of selenium in the blood

(~154 µg/ml) have been correlated with reduced numbers of cancers including

pancreatic, gastric, lung, nasopharyngeal, breast, uterine, respiratory, digestive,

hematological and gynecological [20]. The strongest evidence for the effect of selenium

in reducing cancers was shown for lung cancer (46% lower incidence) [21], esophageal

and gastric-cardiac cancers [22] and especially prostate cancer (63% lower incidence)

[23, 24].

In vitro research has also shown the inhibitory effects of selenium on the growth

of many cancerous cell lines [25-29]. Apostolou and co-workers [30] added selenium (in

the form of either sodium selenite (Na2SeO3), selenomethionine (SeMet), or

selenocysteine (Sec)) to culture media and found a concentration-dependent inhibition of

malignant mesothelioma (MM) cells but normal mesothelial cells were not affected. They

Page 22: Nanostructured Selenium for Biomedical Applications: from

5

also showed that selenium induced apoptosis in MM cells by a mechanism related to

SEP15 which is a gene encodinga 15-kDa selenium containing protein. Menter and

colleagues studied the effect of sodium selenite or SeMet on the growth of normal

primary prostate cells and prostate cancer cells [31] by adding either sodium selenite or

SeMet to culture media. They observed a dose-dependent growth inhibition and apoptosis

in the cancer cells treated with either sodium selenite or SeMet. Meanwhile, growth of

the normal cells was not suppressed by the selenium treatment. Narayan and co-workers

demonstrated that selenium (in the form of sodium selenite) caused growth inhibition and

apoptosis in human brain tumor cells. Interestingly, in spite of a great number of studies

on the effect of selenium on cancer, there are very few experiments focusing on the effect

of elemental selenium on cancer growth, especially the use of selenium in anti-cancer

applications. The mechanisms of selenium-based chemoprevention are also complex and

incompletely understood [32]. Selenium was shown to inhibit angiogenesis which is the

growth of blood vessels around the tumor without which the tumor will not be able to

grow [33]. Selenium was also demonstrated to induce apoptosis, which is a natural-cell-

death process [31, 34]. Evidence also exists to show that selenium produces superoxide

anions that kill cancer cells [35].

In the meantime, studies have also provided evidence of the anti-bacterial

properties of many selenium compounds. For example, selenium-enriched probiotics

have been shown to strongly inhibit the growth of pathogenic E. coli in vivo and in vitro

[36]. Specially, selenium-enriched probiotics (the selenium concentration was

approximately 0.5 µg/mg) were grown and co-cultured with E. coli solutions and were

shown to inhibit the growth of E. coli approximately 10 times compared to the control

Page 23: Nanostructured Selenium for Biomedical Applications: from

6

(no probiotic treatment). In addition, mice fed with selenium-enriched probiotics had the

lowest mortality rate upon exposure to pathogenic E. coli [36]. A series of

organoselenium compounds (such as 2,4,6-tri-para-methoxyphenylselenopyrylium

chloride, 9-para-chlorophenyloctahydroselenoxanthene, perhydroselenoxanthene , and 9-

para-fluorophenyloctahydroselenoxanthene) have been synthesized and have shown anti-

bacterial activities in vitro, especially against S. aureus [37-39]. For example, all

organoselenium compounds synthesized and tested in [38] were shown to be as effective

as penicillin, a common antibiotic, in inhibiting S. aureus growth in solution in vitro. The

anti-bacterial properties of selenium are believed to be attributed to the ability of

selenium to generate superoxide radicals [35] and/or to catalyze oxidation of intracellular

thiol causing thiol depletion that leads to cell death [40].

1.4. Focus of thesis

This thesis is focused on exploring the potential of using nanometer scale

structured selenium for anti-cancer and anti-bacterial applications. The thesis is organized

as follows. Chapter 2 will describe selenium coatings on metallic orthopedic materials

(titanium and stainless steel) with a focus especially on titanium which is the most

common orthopedic material. Chapter 3 will propose a computer simulation method to

reproduce the experimental results on protein adsorption presented in Chapter 2. Chapter

4 will propose a modified contact angle model to fit the experimental results presented in

Chapter 2. Chapter 5 will describe the potential of using selenium coatings for anti-

bacterial applications. Chapter 6 will describe the use of selenium nanoparticles for anti-

bacterial applications. Finally, a conclusion chapter will summarize the content of the

thesis in the biomaterials field at large.

Page 24: Nanostructured Selenium for Biomedical Applications: from

7

CHAPTER 2. SELENIUM NANOCLUSTER COATINGS ON METALS FOR ANTI-CANCER ORTHOPEDIC MATERIALS

2.1. Introduction

2.1.1. Problems with currently used orthopedic implants

The most commonly used materials for bone implants are metals (such as

titanium, stainless steel) and polymers (such as ultra high molecular weight polyethylene)

[41]. Several problems exist with current generation implants including: (i) insufficient

bonding between the implanted material and juxtaposed bone [42-45], (ii) different

mechanical properties between bone and the implant leading to stress shielding [43-45],

and (iii) wear debris generated at articulating surfaces of orthopedic implants that may

lead to cell and bone death [44, 45]. Insufficient new bone growth on implants is the main

cause of insufficient bonding between the implant and surrounding bone. Therefore, one

of the most important characteristics that the new generation of orthopedic implant

materials should have is better biocompatibility to support faster new bone growth. The

above-mentioned problems with implants usually lead to implant failure which results in

surgery to remove the failed implant and insertion of a new one (i.e., revision surgery).

Revision surgery is not only painful for patients but also costly. The number of revision

surgeries due to implant failure has increased steadily over the last decade and is

expected to continue increasing into the next decade.

Page 25: Nanostructured Selenium for Biomedical Applications: from

8

2.1.2. The role of nano-structured surfaces on orthopedic implants

It is well known that tissues in the body possess nano-scale features. For example,

hydroxyapatite crystals, the major inorganic component of bone, are about 2 -5 nm in

width and 50 nm in length. Type I collagen, the major organic component of bone, has

fibrils 300 nm in length, 0.5 nm in width and has a periodicity of 67 nm [46]. Numerous

reports in the literature indicate that osteoblast function is greater on nano-structured

compared to current implant surfaces (which are micron-scale rough and nano-scale

smooth). Studies have shown optimal initial protein adsorption from serum onto nano-

structured ceramics [47, 48] and nano-phase metals [49] leading to greater osteoblast

functions. It has also been demonstrated that increased osteoblasts (i.e., bone-forming

cells) functions on nano-phase compared to micron, conventional ceramics is

independent of surface chemistry and material crystalline phase [50].

2.1.3. Nano-structured selenium coatings: a novel approach of using selenium to create

anti-cancer, anti-bacterial biomaterials

As discussed in Chapter 1, selenium has been shown to possess anti-cancer

properties. It has also been discussed in the previous section that nano-structured

materials can promote bone cell functions. Therefore, there has been efforts to create

nano-rough selenium by chemical etching from selenium compacts [51]. This mode of

selenium addition, however, can limit the mechanical properties of the implant as

selenium, being a metalloid, does not have sufficient mechanical strength needed for

orthopedic applications (such as hip implants). Moreover, this methodology does not

provide control over the release or the dose of selenium. Considering the toxicity of

Page 26: Nanostructured Selenium for Biomedical Applications: from

9

selenium at high doses [52], stability and control over its release would be a very

desirable attribute. The colloidal synthesis of selenium nano-particles is known [53-56]

and yields stable dispersions of free nano-particles in the presence of surface stabilizing

agents [53-56]. It was hypothesized that the addition of nano-particles to implant surfaces

could provide a desirable nano-rough morphology, but would require strong adhesion to

fabricate a surface stable to vibration, sonication, or other common processing steps. A

second target would be to control the time release and the surface density of the selenium,

which determines the selenium dose to the adjacent tissue. To address these needs, a

novel fabrication method was developed here in which colloidal selenium synthesis was

carried out in the presence of substrates to create a coating of uniformly adherent

selenium nanoclusters instead of (or in addition to) free selenium nano-particles in

solution.

2.2. Materials and methods

A detailed fabrication and characterization method for coating titanium and

stainless steel with selenium nanoclusters is introduced here. Cell and bacterial assays to

investigate the functions of healthy osteoblasts and cancerous osteoblasts on the uncoated

and selenium coated materials is also described.

2.2.1. Materials

Metallic substrates (titanium, stainless steel substrates [Alfa Aesar]) were

individually degreased and sonicated in acetone and ethanol for 10 min according to

established procedures [57] with a slight modification comprising of an additional step in

which the substrates were hand-wiped with acetone soaked tissue before degreasing.

Page 27: Nanostructured Selenium for Biomedical Applications: from

10

Degreased titanium substrates were then sterilized by autoclaving at 121oC for 30 min.

Cleaned and sterilized metallic substrates (titanium or Ti, and stainless steel or SS) were

used as a base substrate for colloidal decoration with selenium nanoclusters. The

substrates were exposed to 4:1 molar mixtures of glutathione (GSH, reduced form, TCI

America) and sodium selenite (Na2SeO3, Alfa Aesar) in the concentration ranges shown

in Table 2.1.

Table 2.1. Reagent volumes used in the colloidal synthesis of Se nanoclusters in the presence of various base substrates.

Preparation Method

Reagent *[Se]=0.42mM “Low Dose”

[Se]=5mM “Medium Dose”

[Se]=11.7mM “High Dose”

Deionized water 14.5 ml 9 ml 1 ml 100 mM GSH 0.25 ml 3 ml 7 ml

25 mM Na2SeO3 0.25 ml 3 ml 7 ml Final volume 15 ml 15 ml 15 ml

*[Se] = Final concentration of selenium in the colloidal synthesis solution.

Specifically, the cleaned and sterilized substrates were first placed in a 50 ml

beaker with the side to be decorated facing upward. The reduced glutathione solution was

added to the beaker followed by the sodium selenite solution. Three different solution

concentrations (as shown in Table 2.1) were used to achieve different doses denominated

as low dose (LD), medium dose (MD) and high dose (HD). After a gentle mixing of the

solutions in the reaction beaker, 1mL of 1M NaOH was introduced to bring the pH into

the alkaline regime. The reaction mixture was once again gently mixed and left

undisturbed for 10 min. The substrates were removed from the beaker and rinsed in

deionized water. The uncoated and coated metallic substrates (Ti and SS) were exposed

Page 28: Nanostructured Selenium for Biomedical Applications: from

11

to ultra-violet light for 24 hrs on each side to sterilize them before use in cell

experiments.

2.2.2. Material characterization

2.2.2.1. Topography and chemical characterization

Surfaces of uncoated and selenium coated metallic substrates (Ti and SS) were

visualized (without a conductive coating) using a scanning electron microscope (SEM,

LEO 1530VP FE-4800) with an accelerating voltage from 3 kV to 10 kV. SEM produces

images by scanning an electron beam across the surface of interest in a rectangular

pattern recording the reflected electrons. SEM can achieve much higher magnification

than conventional light microscopy because the wavelengths of electrons used in SEM

are much shorter than that of photons. For chemical characterization, X-ray photoelectron

spectroscopy (XPS, Perkin-Elmer PHI. 5500 Multi-Technique System) was used to

confirm the presence of selenium and relatively compare selenium concentrations on the

different substrates. In XPS, a sample surface is hit by a beam of X-rays that provide

energy to electrons in a thin top layer (about 1 to 10 nm) on the surface. Some electrons

receive enough energy to escape from the sample. XPS measures the kinetic energy and

the number of the escaped electrons that contain information about the elemental

composition, chemical state and quantity of the elements that exist in the material.

2.2.2.2. Strength of attachment of selenium on substrates

To test the strength of attachment, the selenium-coated substrates were placed in a

beaker filled with water. The beaker was subjected to sonication for 10 min at 90 W

Page 29: Nanostructured Selenium for Biomedical Applications: from

12

(Ultrasonic Cleaner 75D, VWR). SEM was used to visualize the surfaces of the

substrates before and after sonication.

2.2.2.3. Surface roughness characterization

Surface roughness can affect protein adsorption and cellular adhesion. To

determine the surface roughness of the substrates (uncoated and selenium coated

substrates), atomic force microscopy (AFM, Autoprobe CP, Park Scientific Instrument)

was used with commercially available AFM tips (radius of tip curvature was less than 10

nm, NSC15/ALBS, Micro-Masch, OR) in tapping mode. The relatively high resolution of

~10nm in the lateral scale and 1 Ao in the vertical scale allows for accurate measurement

of nanometer dimensions (~40-90 nm) on the samples. AFM works by using a tip

attached to a cantilever to scan surface of sample and measure the deflection of the

cantilever caused by forces between atoms at the tip and those on the scanned area. The

scanning area was 5 µm x 5 µm and each substrate was scanned at 5 random regions. The

conventional roughness is defined as the deviation from median height (root mean

square):

2

1

( )N

i avei

Z ZRMS

N

(Equation 2.1)

Here, aveZ is the average Z value within a given area and N is the number of data points

within a given area.

Page 30: Nanostructured Selenium for Biomedical Applications: from

13

2.2.2.4. Contact angle measurements

Contact angles contain information about hydrophilicity or hydrophobicity of a

surface which plays an important role in the interaction of the surface with biological

objects (such as proteins and cells). A drop shape analysis system (DSA-10, Kruss,

Germany) (Figure 2.1) with analysis software (DSA1 v 1.80) was used to determine the

surface contact angles for the samples.

Figure 2.1. Image of DSA100 (Updated Version of DSA-10, Kruss, Germany: Image Adopted from www.kruss.com). Capability of equipment: The range of contact angle measurement is between 0 and 180° with a resolution of ±0.1 and a range of surface tension between 10-2 and 100mJ/m2, with a resolution of 0.01mJ/m2.

To measure water contact angles, a 2µL drop of water was placed onto the

substrates and contact angles were recorded within 10 seconds at room temperature.

Pipettes with controlled suction speed (R-10, Rainin Instrument, MA) were used with

Page 31: Nanostructured Selenium for Biomedical Applications: from

14

appropriately sized pipette tips (Optimum #7560-100) to ensure the accurate liquid

volumes. Measurements were repeated 9 times for each sample.

2.2.3. In vitro biological assays for uncoated and selenium coated metallic substrates

2.2.3.1. Bone cell adhesion and proliferation assays on uncoated and selenium coated Ti

substrates

To investigate osteoblast (bone-forming cell) functions on titanium substrates

coated with selenium, human osteoblasts (will be referred to as healthy osteoblasts, CRL-

11372; American Type of Culture Collection-ATCC, population numbers 6-10) in

Dulbecco's Modified Eagle Media (DMEM) supplemented with 10% fetal bovine serum

(FBS, Hyclone) and 1% Penicillin / Streptomycin (P/S, Hyclone) were seeded at a

density of 3500 cells/cm2 and placed in an incubator under standard cell culture

conditions (37oC, 5% CO2, 95% humidified air) for 4 hrs and 1 day. This human

osteoblast cell line was chosen because they have been shown to be able to differentiate

into mature osteoblasts expressing a normal osteoblast phenotype [58].

To study cancerous bone cell functions on the samples, mouse osteosarcoma

osteoblasts (will be referred to as cancerous osteoblasts; CRL-2837; ATCC, population

numbers 14-17) in DMEM supplemented with 10% FBS (Hyclone) and 1% P/S

(Hyclone) were seeded at a density of 3500 cells/cm2 and placed in an incubator under

standard cell culture conditions (37oC, 5% CO2, 95% humidified air) for 4 hrs as well as

1 and 3 days. Media was exchanged every other day. This mouse osteosarcoma cell line

was chosen because this cell line has been shown to be able to develop primary tumors

and pulmonary metastases with histology consistent with osteosarcoma in human patients

Page 32: Nanostructured Selenium for Biomedical Applications: from

15

[59]. In addition, for this cell line, “expression of bone sialoprotein, biglyan, decorrin,

and osteopontin was suggestive of the bone lineage cells” [59].

After the desired time periods, cells were fixed using formaldehyde 4% (Sigma),

stained with 4',6-diamidino-2-phenylindole (DAPI) (ATCC) and counted under

fluorescence microscopy (Zeiss Axiovert 200M Light Microscope) in five random fields

and averaged for each substrate.

2.2.3.2. Healthy osteoblast differentiation assays

Healthy osteoblasts (100,000 cells/cm2) were seeded onto the substrates of

interest to the present study and cultured in DMEM (supplemented with 10% FBS, 1%

P/S, 50 μg/ml L-ascorbate (Sigma) and 10 mM β-glycerophosphate (Sigma)) under

standard cell culture conditions for 14 days. Medium was replaced every other day. At

the end of the prescribed time period, the substrates were rinsed with tris buffered saline

(pH 7.2, Sigma Aldrich) for three times. The remaining osteoblasts on the substrates were

lysed using 1 mL of deionized water and three freeze-thaw cycles. The supernatant

lysates were transferred into microtubes for determining intracellular protein synthesis

and alkaline phosphatase (ALP) activity, while the remaining substrates were used for

determining extracellular calcium deposition.

2.2.3.2.1. Total intracellular protein content

Total protein is proportional to total number of cells. Total protein content in the

cell lysates was determined spectrophotometrically using a commercially available kit

(Pierce Chemical Co.) and following manufacturer's instructions. For this purpose,

aliquots of each protein-containing, distilled-water supernatant were incubated with a

Page 33: Nanostructured Selenium for Biomedical Applications: from

16

solution of copper sulfate and bicinchoninic acid at 37°C for 2 hrs. Light absorbance of

these samples was measured at 562 nm on a MR600 Spectrophotomectric Microplate

Reader (Dynatech). Total intracellular protein (expressed as mg) synthesized by

osteoblasts cultured on the substrates of interest to the present study was determined from

a standard curve of absorbance versus known concentrations of albumin run in parallel

with experimental samples. Total intracellular protein synthesized by osteoblasts cultured

on conventional, uncoated titanium served as controls.

2.2.3.2.2. Intracellular alkaline phosphatase (ALP) activity

Intracellular ALP activity is an indicator of how active the osteoblast cells are in

producing new bone. Intracellular human osteoblast ALP activity was analyzed by a

commercial ALP activity detection kit (Upstate Cell Signal Solutions) following

manufacturer's instructions. The ALP activity in cell lysates was spectrophotometrically

detected by adding a Malachite Green solution compared to a standard curve of

absorbance versus known concentrations of converted phosphate, which reported ALP

activity as the amount of converted phosphate. ALP activity was further normalized by

total intracellular protein content and substrate surface area.

2.2.3.2.3. Calcium deposition in the extracellular matrix

Calcium deposition is an important indicator of the bone-forming process by

osteoblasts. The remaining extracellular matrix on the substrates after 14 days of

culturing healthy osteoblasts was treated with 0.6 M HCL in an incubator (37oC) for 24 h.

After the prescribed time period, calcium present in the acidic supernatant was quantified

using a commercially available kit (Calcium Reagent Set, Pointe Scientific) following

Page 34: Nanostructured Selenium for Biomedical Applications: from

17

manufacturer's instructions. Light absorbance of the acidic supernatant solution with the

addition of a calcium reagent was read at 570 nm. Calcium content was determined from

a standard curve of absorbance versus known concentrations of calcium run in parallel

with the acidic supernatant solution, and the results were normalized over the substrate

surface area.

2.2.3.3. Release of selenium in cell culture media

To study the release of selenium from the substrate surface into the cell culture

media, substrates (uncoated and low, medium and high selenium dose coated titanium)

were cultured without cells in DMEM supplemented with 1% P/S (Hyclone) for 3 days.

To quantify the release of selenium to the fluid phase, the media was exchanged after 2

days (to be consistent with the proliferation experiments where media was exchanged

every other day) and the spent media collected for graphite furnace atomic absorption

spectroscopy (AAS) using Zeeman background correction (41002L Perkin-Elmer GF-

AAS). The instrument was calibrated and tested using the samples with known selenium

concentrations. All the samples were acidified to 2% nitric acid prior to the GF-AAS run.

The wavelength was 196 nm with a bandwidth of 2 nm. A five-step graphite furnace

program was used: (1) 110oC, ramp for 1 s, hold for 20 s; (2) 130oC, ramp for 5 s, hold

for 30 s; (3) 650oC, ramp for 10 s, hold for 20 s; (4) 1750oC, hold for 6 s; and (5) 2400oC,

ramp for 1 s, hold for 2 s. To reduce any matrix effects, 4 µl of a matrix modifier ([1 µg

Pd + 0.6 µg Mg(NO3)2]/µl in 0.5 mol/l HNO3) was used per 20 µl of sample size.

Page 35: Nanostructured Selenium for Biomedical Applications: from

18

2.2.3.4. Effects of selenium released into culture media on healthy and cancerous cells

Either uncoated titanium or titanium coated with the highest selenium density

were cultured (without cells) in DMEM supplemented with 10%FBS and 1% P/S for 2

days. After 2 days, the media (supernatant) was collected and used to culture either

osteoblast or osteosarcoma cells (described in 2.2.3.1) in the wells of 6 well-tissue culture

plates. Fresh DMEM supplemented with 10%FBS and 1% P/S was used as the control.

After 3 days of culture, cell densities were determined by the fluorescence microscopy

techniques as just described in 2.2.3.1.

2.2.3.5. Intracellular thiol assays

Intracellular thiol groups are crucial for the overall health of cells. Intracellular

thiols protect the cells from damages caused by free radicals, toxins, carcinogens, etc. To

investigate the influence of selenium released into the culture media on the level of

intracellular thiols, uTi and High-nSe-Ti were immersed in DMEM (supplemented with

10% FBS and 1% P/S) without cells for 2 days. After the indicated time, supernatants

collected and used to culture either osteoblast or osteosarcoma cells (described in 2.2.3.1)

for 3 days. After the indicated time, cells were collected and analyzed for thiol content

using a glutathione assay kit (CS1020, Sigma Aldrich) following manufacturer’s

instruction. The assay uses a thiol probe (monochlorobimane) which passes through cell

membranes to detect the level of reduced glutathione, the major free thiol in most living

cells. Glutathione is involved in many biological processes such as removal of

hydroperoxides and detoxification of xenobiotics. Intracellular reduced glutathione level

is a sensitive indicator of the overall health of a cell. If the probe binds to reduced

Page 36: Nanostructured Selenium for Biomedical Applications: from

19

glutathione in cells, it forms a fluorescent product whose fluorescence intensity can be

measured. The unbound probe does not show fluorescence.

2.2.4. Fibronectin adsorption on uncoated and selenium coated titanium substrates

Fibronectin is one of important cell-adhesive proteins that regulate interaction

between osteoblasts and substrates. Adsorption of fibronectin on implant is, therefore,

important for cell adhesion, migration and spreading. Several methods exist to measure

protein adsorption on implant surfaces (such as XPS, Raman spectroscopy, fluorescence

spectroscopy, infrared spectroscopy, enzyme-linked immunosorbent assay (ELISA) or

surface plasmon resonance). Among all the methods, ELISA is simple, requiring less

equipment but still provide accurate results. In ELISA, fibronectin is directly linked to

antibodies which are then conjugated with fluorescent dyes and detected by a fluorescent

signal detector. Experiments were conducted to investigate adsorption of fibronectin from

either DMEM (supplemented with 10%FBS and 1% P/S) or 5µg/mL fibronectin in

phosphate-buffered solution (PBS, Sigma) onto the substrates of interest. Specifically,

ELISA experiments were conducted following these steps below:

1. Uncoated and selenium-coated titanium samples were immersed in 0.5 mL of

either DMEM (supplemented with 10% FBS and 1% P/S) or 5µg/mL fibronectin solution

(in PBS) in a 24 well culture dish (Corning) in an incubator (37oC, 5%CO2) for 24 hrs.

Samples were immersed in either DMEM (with no FBS or P/S supplement) or PBS (with

no fibronectin) were used as controls to subtract a nominal fluorescence signal later.

2. Samples were rinsed in a PBS three times. Bovine serum albumin (BSA,

Sigma) (2% by wt % in PBS) was used to block all substrate areas that did not react with

proteins.

Page 37: Nanostructured Selenium for Biomedical Applications: from

20

3. A rabbit anti-bovine fibronectin antibody (AB2047, Chemicon) was added to

each well at concentration of 6 µg/mL (in 1% BSA) and was incubated for 1 hr to link to

fibronectin.

4. Samples were rinsed in 0.05% Tween 20 (Sigma) three times. A second

antibody, goat anti-rabbit conjugated with horseradish peroxidase (HRP) was added to

each well at a concentration of 10 µg/mL (in 1% BSA) and incubated for 1 hr.

5. Samples were transferred to new 24 well plates and rinsed twice in 0.05%

Tween 20. An 2,2'-azino-bis (3-ethylbenzthiazoline-6-sulfonic acid) (ABTS) substrate kit

(SK-4500, Vector Laboratories) was used. ABTS reacts only with HRP to develop a

water soluble, green colored product. Different intensities of fluorescence (for each

sample) were detected by a spectrophotometer (SpectraMax 190, Molecular Devices) at

405nm. Fluorescence intensities were subtracted from those of the corresponding control.

2.2.5. Statistical analysis

Experiments were conducted in triplicate and repeated three times. Data were

collected and the significant differences were assessed with the probability associated

with a one- tailed Student’s t test.

2.3. Results

In this section, results from material characterization studies and cell tests on

titanium substrates will be presented first. Results on stainless steel substrates will follow

to demonstrate that the same synthesis method can be used on other orthopedic metals to

achieve similar results.

Page 38: Nanostructured Selenium for Biomedical Applications: from

21

2.3.1. Surface characterization

Selenium nanoclusters were clearly present on titanium substrates following the in

situ reaction under the low, medium and high dose conditions (which corresponded to

increasing the overall reagent concentrations as detailed in Table 2.1) (Figure 2.2).

Figure 2.2. Representative SEM images of: (A) uTi; (B) Low-nSe-Ti; (C) Medium-nSe-Ti and (D) High-nSe-Ti. The surface area coverage of selenium on the selenium coated substrates were determined (using an image processing program, ImageJ) to be 2.7%, 5.1% and 7.5% for Low-nSe-Ti, Medium-nSe-Ti and High-nSe-Ti, respectively.

The clusters were approximately 80nm in diameter and had hemispherical shapes

(Figure 2.3). Higher surface densities of selenium nanoclusters were observed on the

titanium substrates for the higher dose cases (prepared at higher selenite concentrations in

the final synthesis solution). The substrates will be referred to as uTi, Low-nSe-Ti,

Medium-nSe-Ti and High-nSe-Ti for uncoated titanium, and low, medium and high

Page 39: Nanostructured Selenium for Biomedical Applications: from

22

selenium nanocluster density coated titanium substrates, respectively. The surface

coverage of selenium determined from SEM images using an image processing program,

ImageJ (NIH), was approximately 2.7% for Low-nSe-Ti, 5.1% for Medium-nSe-Ti and

7.5% for High-nSe-Ti. The surface coverage of selenium is not linearly dependent on the

concentration of selenium in the synthesis solution (see Table 2.1)

Figure 2.3. SEM image, taken at a 45° tilt, of High-nSe-Ti showing the hemispherical shape of the selenium nanoclusters on the titanium surface.

The morphology in the high-magnification tilted SEM image (Figure 2.3) strongly

suggested the selenium was bonded to the substrate. Together with the fact that cleaning

of the titanium surface before coating was found to be crucial for forming of selenium

nanoclusters on the surfaces, it is speculated that selenium nanoclusters were formed

Page 40: Nanostructured Selenium for Biomedical Applications: from

23

through nucleation processes. The nucleation sites on the titanium surface (e.g. defects,

impurities) served as high energy regions for the nucleation process to happen.

To test the strength of attachment, the composite substrate was subjected to

sonication for 10 min at 90 W (Ultrasonic cleaner 75D, VWR). The morphology of the

substrates and the densities of coatings appeared similar by SEM before and after

sonication. The morphology was also found to be unchanged after UV treatment for 24

hrs (Figure 2.4).

Figure 2.4. Morphology of a selenium-coated titanium surface appeared the same before (A) and after sonication (B) and UV radiation (C). Bars=500 nm.

The presence of elemental selenium on the titanium surfaces was further

confirmed by XPS analysis. All XPS profiles of the selenium coated substrates showed

peaks of selenium not observed for the conventional uncoated titanium substrates (Figure

2.5).

Page 41: Nanostructured Selenium for Biomedical Applications: from

24

Figure 2.5. All coated substrates showed peaks of selenium that were not detected in uncoated substrate.

The position of the peak at a binding energy of 55.2 eV (Figure 2.6) confirmed

the elemental state of selenium [60].

Figure 2.6. XPS profiles of High-nSe-Ti showing the characteristic binding energy peak for elemental selenium at 55.2 eV.

The elemental form is also produced by the colloidal synthesis of free particles

[61], which indicated that the substrate did not alter the basic reaction chemistry except to

Page 42: Nanostructured Selenium for Biomedical Applications: from

25

introduce nucleation sites. Selenium peaks were more pronounced both in height and in

number on the substrates coated with higher doses of selenium.

The AFM images and analysis (Figure 2.7) confirmed again the present and

hemispherical morphology of selenium nanoclusters on the coated Ti substrates (Figure

2.7).

Page 43: Nanostructured Selenium for Biomedical Applications: from

26

Figure 2.7. Representative AFM images and line analysis of uTi (A), Low-nSe-Ti

(B), Medium-nSe-Ti (C) and High-nSe-Ti (D). Nano-scale roughness was created by the

selenium coatings and increased with increasing selenium coating density.

More importantly, all coated substrates possess nano-scale roughness (Figure 2.8)

contributed by the nanostructure of selenium coatings. The normalized RMS roughness

Page 44: Nanostructured Selenium for Biomedical Applications: from

27

values were 8.07±0.71 nm for uTi, 9.72±0.41 nm for Low-nSe-Ti, 12.56±0.79 nm for

Medium-nSe-Ti and 14.11±1.08 nm for High-nSe-Ti.

Figure 2.8. RMS of the substrates increased with increasing selenium coating density. Data = mean ± standard error of the mean; N=3; * p<0.05 compared to all other substrates; ** p<0.01 compared to Medium-nSe-Ti and High-nSe-Ti.

Increased coating levels (provided by increasing selenium concentration in the

synthesis solution and as indicated by the densities of selenium clusters in SEM images

and in AFM images) created surfaces with increased nanometer scale roughness (Figures

2.7 and 2.8).

Uncoated Ti substrates were hydrophilic with an average water contact angle of

approximately 64 degrees. Importantly, coating Ti with selenium nanoclusters made the

Ti surfaces more hydrophobic with an average water contact angles of approximately 74

degrees (Figure 2.9). No significant difference was found among contact angles on all

selenium coated substrates.

Page 45: Nanostructured Selenium for Biomedical Applications: from

28

Figure 2.9. Water contact angles on the uncoated and coated Ti substrates. Contact angles increased on the substrates coated with selenium nanoclusters. Data = mean ± standard error of the mean; N=3; * p<0.05 compared to all the coated substrates. There was no significant difference among the contact angles on the coated substrates.

The plateau of water contact angle values for different selenium coating densities

was hypothesized to be attributed to the formation of air-pockets during the contact angle

measurement. More detailed discussion of this will be presented in Chapter 4.

2.3.2. Cell adhesion and proliferation

Most importantly, after 4 hrs of culture, healthy osteoblast densities significantly

increased on the High-nSe-Ti compared to uTi and Low-nSe-Ti (Figure 2.10). After 4

hrs, healthy osteoblast density increased 12% on High-nSe-Ti compared to uTi. After 1

day, healthy osteoblast density increased 68% on High-nSe-Ti, 37% on Medium-nSe-Ti

Page 46: Nanostructured Selenium for Biomedical Applications: from

29

and 13% on Low-nSe-Ti compared to uTi, respectively. After 3 days, healthy osteoblast

density increased 60% on High-nSe-Ti, 37% on Medium-nSe-Ti and 22% on Low-nSe-Ti

compared to uTi, respectively.

Figure 2.10. Increased healthy osteoblast densities after 4 hrs, 1 day and 3 days. Data = mean ± standard error of the mean; N=3, * p<0.05 compared to Medium-nSe-Ti (compared at same time period), ** p<0.05 compared to uTi (compared at same time period), *** p< 0.05 compared to Low-nSe-Ti (compared at same time period), # p<0.05 compared to uTi (compared at same time period), & p =0.06 compared to Low-nSe-Ti. (compared at same time period).

In contrast, for cancerous cells, after 4 hrs, cell densities were significantly higher

on the uTi and Low-nSe-Ti than on the High-nSe-Ti (Figure 2.11).

Page 47: Nanostructured Selenium for Biomedical Applications: from

30

Figure 2.11. Cancerous osteoblast densities after 4 hrs, 1 day, and 3 days. Decreased cancerous osteoblast densities on selenium coated titanium substrates after 4 hrs, 1 day and 3 days. Data = mean ± standard error of the mean; N=3, * p<0.05, ** p< 0.01 compared to High-nSe-Ti. (Compared at same time period); # p<0.1 compared to Medium-nSe-Ti. (Compared at same time period). There are no significant differences in cell densities among substrates after 1 day.

Cancerous osteoblast densities after 4 hrs were 5365 ±145 cells/cm2 on uTi, 4779

± 443 cells/cm2 on Low-nSe-Ti, 4526 ± 308 cells/cm2 on Medium-nSe-Ti and 3547 ± 47

cells/cm2 on High-nSe-Ti. Cancerous osteoblast density decreased approximately 1.5

times on High-nSe-Ti compared to uTi.

After 1 day of cell culture, healthy osteoblast densities increased on Medium-nSe-

Ti (8165± 1509 cells/cm2) and High-nSe-Ti (9983± 1075 cells/cm2) compared to that on

uTi (5955± 561 cells/cm2) (Figure 2.12). Healthy cell density increased approximately

1.7 times on High-nSe-Ti and 1.4 times on Medium-nSe-Ti compared to that on uTi. No

significant difference was found for the cancerous osteoblast densities on the four types

of titanium substrates (i.e., uTi, Low-nSe-Ti, Medium-nSe-Ti, High-nSe-Ti) after one day

of culture.

Page 48: Nanostructured Selenium for Biomedical Applications: from

31

Figure 2.12. Representative fluorescence microscopy images of healthy osteoblasts on: (A) uTi; (B) Low-nSe-Ti; (C) Medium-nSe-Ti and (D) High-nSe-Ti after 1 day (Magnification = 10X). Scale bars = 200 μm.

However, after 3 days of culture, cancerous osteoblast density on the High-nSe-Ti

was significantly reduced in comparison to all other substrates (Figure 2.13). Cancerous

osteoblast densities on Medium-nSe-Ti were also significantly lower than that on uTi.

Cancerous osteoblast densities on the substrates after 3 days were 141882 ±36572

cells/cm2 for uTi, 111581±17101 cells/cm2 for Low-nSe-Ti, 72187±13702 cells/cm2 for

Medium-nSe-Ti and 31885±6933 cells/cm2 for High-nSe-Ti. Cancerous cell density

decreased approximately 4.5 times on High-nSe-Ti, 1.9 times on Medium-nSe-Ti

compared to that on uTi after 3 days.

Page 49: Nanostructured Selenium for Biomedical Applications: from

32

Figure 2.13. Representative fluorescence microscopy images of cancerous osteoblasts on: (A) uTi; (B) Low-nSe-Ti; (C) Medium-nSe-Ti and (D) High-nSe-Ti after 3 days (Magnification = 20X). Scale bars = 100 μm.

To quantitatively assess the effects of selenium nanocluster coatings on growth of

healthy osteoblasts, exponential growth curves were fit to the experiment data (Figure

2.14)

Page 50: Nanostructured Selenium for Biomedical Applications: from

33

Figure 2.14. Fitting exponential growth curves to experiment data for densities of healthy osteoblasts cultured on uncoated titanium and on selenium-coated titanium substrates. Growth curves of healthy osteoblasts cultured on higher selenium coating densities had large slopes.

The fitting parameter and doubling time of healthy osteoblast determined from

the exponential fitting curves were summarized in Table 2.2.

Table 2.2. Fitting parameter and the goodness of fitting (R2) for growth curve and calculated doubling time of healthy osteoblasts cultured on uncoated and selenium-coated titanium substrates

Fitting parameters for y~ebx b Doubling time (hours)

τ = ln(2)/b R2

uTi 0.0167 41.505 0.9125 Low-nSe-Ti 0.0193 35.914 0.9067

Medium-nSe-Ti 0.0198 35.007 0.8559 High-nSe-Ti 0.0208 33.324 0.8131

Doubling time of healthy osteoblasts decreased from ~ 41 hrs when cultured on

uncoated titanium to ~ 33 hrs when cultured on High-nSe-Ti, i.e., the healthy osteoblasts

divided ~ 1.2 times faster on High-nSe-Ti compared to uTi.

Page 51: Nanostructured Selenium for Biomedical Applications: from

34

Exponential growth curves fitted from experiment data for densities of cancerous

osteoblast cultured on the uncoated and selenium coated substrates showed a reverse

trend (Figure 2.15); growth curves of cells cultured on selenium-coated surfaces had

smaller slopes.

Figure 2.15. Fitting exponential growth curves to experiment data for densities of cancerous osteoblasts cultured on uncoated titanium and selenium-coated titanium substrates. Curves of cancerous osteoblasts cultured on lower selenium coating densities had large slopes with the uncoated titanium substrates having the largest slope.

The fitting parameter and doubling time of cancerous osteoblast determined from

the exponential fitting curves were summarized in Table 2.3.

Page 52: Nanostructured Selenium for Biomedical Applications: from

35

Table 2.3.Fitting parameter and goodness of the fitting (R2) for growth curve and calculated doubling time of cancerous osteoblasts cultured on uncoated and selenium-coated titanium substrates.

Fitting parameters for y~ebx b Doubling time (hours)

τ = ln(2)/b R2

uTi 0.0499 13.89 0.9794 Low-nSe-Ti 0.0478 14.5 0.9844

Medium-nSe-Ti 0.0471 14.716 0.99 High-nSe-Ti 0.0322 21.526 0.999

Doubling time of cancerous osteoblasts increased from ~ 14 hrs when cultured on

uncoated titanium to ~ 14.5 hrs on Low-nSe-Ti, 14.7 hrs on Medium-nSe-Ti and 21.5 hrs

when cultured on High-nSe-Ti.

2.3.3. Healthy osteoblast differentiation

To investigate the long-term functions of healthy osteoblasts on substrates coated

with Se, a variety of osteoblast differentiation makers (including intracellular protein

synthesis, ALP activity and extracellular calcium deposition) were examined in the

present study. Uncoated Ti substrates were used as the control. The value of ALP activity

(normalized to total protein content) was 136.78 ±3.7 pmol/µg.cm2 for uTi, 120.16 ±5.5

pmol/µg.cm2 for Low-nSe-Ti, 151.93 ±5.0 pmol/µg.cm2 for Medium-nSe-Ti and 226.75

±28.4 pmol/µg.cm2 for High-nSe-Ti. The results showed that after 14 days of culturing,

ALP activity of healthy osteoblasts on High-nSe-Ti was higher than on uTi, Low-nSe-Ti

and Medium-nSe-Ti (Figure 2.16). Compared to the uncoated substrate (uTi), the

substrate coated with highest amount of selenium (High-nSe-Ti) has ALP activity

increased approximately 1.7 times.

Page 53: Nanostructured Selenium for Biomedical Applications: from

36

Figure 2.16. Increased alkaline phosphatase (ALP) activity of osteoblasts on High-nSe-Ti compared to all other substrates. Data = mean ± standard error of the mean; N=3, * p<0.01 compared to uTi, Low-nSe-Ti and Medium-nSe-Ti.

There was no significant difference between ALP activity on uTi, Low-nSe-Ti

and Medium-nSe-Ti. Higher ALP activity which was normalized by intracellular protein

content per unit area on High-nSe-Ti indicated that bone formation on High-nSe-Ti was

more active than on uTi, Low-nSe-Ti and Medium-nSe-Ti.

Calcium deposition in the extracellular matrix is one of the important markers of

bone formation and osteoblast differentiation. Calcium deposition (normalized to surface

area) was 0.97 ±0.03 (µg/cm2) for uTi, 1.003 ±0.029 (µg/cm2) for Low-nSe-Ti, 1.004

±0.038 (µg/cm2) for Medium-nSe-Ti and 1.093 ±0.04 (µg/cm2) for High-nSe-Ti.

Calcium deposition by osteoblasts on High-nSe-Ti was significantly greater than on uTi

(Figure 2.17). There were no significant difference between calcium deposition by

osteoblasts on uTi, Low-nSe-Ti and Medium-nSe-Ti.

Page 54: Nanostructured Selenium for Biomedical Applications: from

37

Figure 2.17. Extracellular calcium deposition by osteoblasts on uTi, Low-nSe-Ti, Medium-nSe-Ti and High-nSe-Ti after 14 days. Data = mean ± standard error of the mean; N=3, * p<0.05 compared to uTi.

These results are consistent with the results of proliferation of healthy osteoblasts

after 1 day and 3 days (Figures 2.10 and 2.12) and suggest promoted active osteoblast

functions on High-nSe-Ti compared to uTi.

2.3.4. Selenium release

To test the release of selenium coatings into culture media, titanium substrates

coated with selenium were immersed in culture media (DMEM supplemented with 10%

FBS and 1% P/S) without cells for 3 days. Culture media were then used to analyze

selenium content. A dose dependent release of selenium from the substrates to cell

culture medium was observed by GF-AAS analysis (Figure 2.18).

Page 55: Nanostructured Selenium for Biomedical Applications: from

38

Figure 2.18. Total selenium released into cell culture media after 2 and 3 days. Data = mean ± standard error of the mean; N=2, * p<0.0001 compared to High-nSe-Ti. (Compared at same time period); † p<0.0001 compared to Low-nSe-Ti. (Compared at same time period).

After 2 days, approximately 3.6% of the total selenium (estimated using SEM

images) was released for the High-nSe-Ti. Release then slowed greatly, with a negligible

selenium release occurring on the third day. The release from High-nSe-Ti was

significantly greater than that from Medium-nSe-Ti and that from Low-nSe-Ti (compared

at the same time period). The release from Medium-nSe-Ti was also significantly greater

than from Low-nSe-Ti (compare at the same time periods). The amount of selenium

released into culture media was also not linearly dependent on the coating density.

2.3.5. Effects of selenium released into culture media on healthy and cancerous cells

To further explore why proliferation of cancerous osteoblasts was inhibited on

selenium coated Ti, the supernatant experiments were conducted which consisted of

Page 56: Nanostructured Selenium for Biomedical Applications: from

39

immersing the substrates in culture media with no cells, collecting the media (i.e.,

supernatant) after 2 days and culturing cells (either healthy osteoblasts or cancerous

osteoblasts) in the supernatant. The results (Figure 2.19) showed significantly lower

(approximately 1.5 times smaller) cancerous cell densities cultured with the supernatant

from High-nSe-Ti than cancerous cell cultured with fresh media (p<0.01) after 3 days of

culture. The density of cancerous cells cultured with the supernatant from High-nSe-Ti

was also smaller than that of cells cultured with the supernatant from uTi (p=0.12). On

the other hand, after 3 days, no significant difference was found among healthy osteoblast

densities cultured in fresh media, the supernatant from uTi or supernatant from High-nSe-

Ti (Figure 2.19). The release of selenium into culture media showed no toxic effects on

healthy osteoblasts in this study.

Figure 2.19. Decreased cancerous osteoblast and maintained healthy osteoblast densities in media collected from supernatant experiment with High-nSe-Ti. Data = mean ± standard error of the mean; N=3, ** p<0.01 compared to High-nSe-Ti. Density of cancerous cells cultured in media collected from the supernatant experiment with uncoated Ti was greater (but not significant, * p=0.12) than that of cells cultured in media collected from the supernatant experiment with High-nSe-Ti. There was no significant difference among the densities of healthy osteoblasts.

Page 57: Nanostructured Selenium for Biomedical Applications: from

40

2.3.6. Intracellular thiol assays

To have a mechanistic understanding of the results presented in Figure 2.19,

intracellular thiol assays experiments were conducted for two types of substrates, uTi

(control) and High-nSe-Ti (which is the Se coated substrate that had the highest normal

healthy osteoblast density and the lowest cancerous osteoblast density from individual

cell experiments). Depletion of intracellular thiol content has been shown to cause cell

death (either apoptosis or necrosis or both) [62-65]. It has been shown that elemental

selenium in the form of nanoparticles caused thiol depletion in leukemia cells leading to

cell death [40]. Therefore, this part of the thesis tested the hypothesis that selenium

released into culture media caused depletion of intracellular thiols. For this, cells (either

healthy osteoblasts or cancerous osteoblasts) were cultured in supernatants (i.e., culture

media collected after 3 days of immersing in DMEM with no cells) from uTi or High-

nSe-Ti. After 1 day, cells were analyzed for content of intracellular thiols.

Thiol content of healthy osteoblasts cultured in supernatant from High-nSe-Ti was

lower than that of healthy osteoblasts cultured in supernatant from uTi (Figure 2.20).

Figure 2.18. Decreased thiol content of healthy osteoblasts cultured in supernatant from High-nSe-Ti compared to uTi. Data = mean ± standard error of the mean; N=3; * p<0.1

Page 58: Nanostructured Selenium for Biomedical Applications: from

41

Similarly, content of intracellular thiols in cancerous osteoblasts cultured in

supernatant from High-nSe-Ti was lower than uTi (Figure 2.21).

Figure 2.19. Decreased thiol content of cancerous osteoblasts cultured in supernatant from High-nSe-Ti compared to uTi. Data = mean ± standard error of the mean; N=3; * p<0.05.

As can be seen from the results (Figures 2.20 and 2.21), thiol content was

depleted in both healthy osteoblasts and cancerous osteblasts when cultured in

supernatant from High-nSe-Ti. Since cancerous cells are more susceptible to oxidative

stress than their normal counterparts [66-68], it is likely that, in this study, the cancerous

cells were more sensitive to oxidative stress caused by thiol depletion compared to the

healthy cells. It is also possible that many cancer cells are more sensitive to thiol

depletion compared to normal counterparts because the level of thiols in those cancer

cells are close to the level required for cell survivals [69].

2.3.7. Fibronectin adsorption on uncoated and selenium-coated titanium substrates

Fibronectin is one of the most important cell adhesive proteins that regulate the

interactions between osteoblasts and surface of bone implants. To understand the

Page 59: Nanostructured Selenium for Biomedical Applications: from

42

mechanism of improved healthy osteoblast functions (adhesion, proliferation and

differentiation) on the selenium-coated titanium substrates, fibronectin adsorption

experiments were conducted. Amount of fibronectin was determined using ELISA

method. In this method, fibronectin was attached to antibodies which were attached to a

“color” molecule. The “color” molecules were then solubilized by ABTS solution to give

a soluble green color product whose optical density can be measured. A standard curve

was constructed from a series of known fibronectin concentrations to give relationship

between optical density and amount of fibronectin. Results from experiments of

fibronectin adsorption from DMEM (supplemeted with 10% FBS and 10% P/S) onto

substrates of interests showed that fibronection adsorption was greater on Medium-nSe-

Ti and High-nSe-Ti compared to uTi while fibronectin adsorption on Low-nSe-Ti was

not significantly different from that on uTi (Figure 2.22).

Page 60: Nanostructured Selenium for Biomedical Applications: from

43

Figure 2.20. Increased fibronectin adsorption from DMEM onto Medium-nSe-Ti and High-nSe-Ti substrates. DMEM (supplemented with 10% FBS and 10% P/S) was estimated to have fibronectin concentration of 3µg/mL. Each sample was immersed in 0.5 mL of DMEM solution for 24 hrs under standard condition (37oC, 5% CO2, 95% humidified air) and amount of absorbed fibronectin was determined via optical density of ABTS solution using ELISA method. A standard curve relating optical density to amount of fibronectin was used to determine fibronectin concentration from optical densities of ABTS solution. Finally, amount of fibronectin was normalized to geometrical surface areas of the substrates. Data = mean ± standard error of the mean; N=3, * p<0.05 compared to uTi, ** p<0.05 compared to uTi and Low-nSe-Ti.

Results on fibronectin adsorption from fibronectin solution (in PBS) onto the

uncoated and coated substrates also showed similar trend (Figure 2.21): Samples with

higher coating densities had higher amount of protein adsorbed.

Page 61: Nanostructured Selenium for Biomedical Applications: from

44

Figure 2.21. Increased fibronectin adsorption from fibronectin solution (in PBS) on Medium-nSe-Ti and High-nSe-Ti substrates. Each sample was immersed in 0.5 mL of 5 µg/mL fibronectin solution for 24 hrs under standard condition (37oC, 5% CO2, 95% humidified air) and amount of absorbed fibronectin was determined via optical density of ABTS solution using ELISA method. A standard curve relating optical density to amount of fibronectin was used to determine fibronectin concentration from optical densities of ABTS solution. Finally, amount of fibronectin was normalized to geometrical surface areas of the substrates. Data = mean ± standard error of the mean; N=3, * p<0.05 compared to uTi and Low-nSe-Ti. ** p=0.055 compared to Medium-nSe-Ti.

It is noted that the concentration of fibronectin in DMEM is approximately the

same as the concentration of fibronectin in PBS used in adsorption experiments.

However, the amount of fibronectin adsorbed from fibronectin solution in PBS (no other

protein present) was approximately 10 times higher than that from DMEM. This is likely

due to the competitive adsorption of other proteins (such as albumin, laminin, vitronectin,

etc.) in the DMEM onto the substrate surface.

The increase in fibronectin adsorption onto the selenium nanocluster coated

surfaces can be attributed to factors such as surface area increase, surface roughness/

morphology, surface charges or surface hydrophilicity/hydrophobicity properties.

However, in this study, the surface area increase was not significant among the uncoated

Page 62: Nanostructured Selenium for Biomedical Applications: from

45

and coated surfaces. In addition, hydrophobicity/hydrophilicity properties were not

significantly different among the coated surfaces. Therefore, it is hypothesized that the

increase in fibronectin adsorption on the higher density selenium-coated substrates was

due to the surface roughness/morphology. Detail discussion of this subject is in Chapter

3.

2.4. Results on stainless steel samples

Stainless steel is another common material used in orthopedics. As a

demonstration of the applicability of the fabrication method to create nanocluster

selenium coatings for anti-cancer orthopedic application, results on scanning electron

microscopy investigation of the stainless steel surfaces and results on culturing healthy

osteoblasts, cancerous osteoblasts on the substrates for 3 days were presented here.

2.4.1. Material characterization

Selenium nanoclusters were clearly present on stainless steel following the in situ

reaction under low, medium and high dose conditions, (which corresponded to increasing

overall reagent concentrations as detailed in Table 2.1) (Figure 2.22)

Page 63: Nanostructured Selenium for Biomedical Applications: from

46

Figure 2.22. SEM images of stainless steel substrates: (A) uncoated; (B), (C), and (D) coated with Se concentration in the final solution which equaled 0.42 mM, 0.83 mM ,and 1.25 mM, respectively. Some Se nanoclusters are indicated by arrows. These clusters are approximately 80nm in diameters. Bars = 1µm.

Higher surface densities of Se nanoclusters were observed on the stainless steel

substrates with higher doses (prepared at higher selenite concentrations). The surface area

coverage of selenium nanoclusters (determined from SEM images using imageJ, an

image processing program) was approximately 0.6% in (B), 1.5% in (C) and 4.5% in (D).

The substrates will be referred to as uSS, Low-nSe-SS, Medium-nSe-SS and High-nSe-

SS for uncoated stainless steel, and low, medium and high Se nanocluster density coated

stainless steel substrates, respectively.

One can notice that, compared to the selenium coated titanium substrates, the

selenium coated stainless steel substrates prepared using the same concentration of

selenium in the synthesis solutions had lower coating densities of selenium nanoclusters.

This might be due to the difference in densities of nucleation sites present on the surfaces

Page 64: Nanostructured Selenium for Biomedical Applications: from

47

of titanium and stainless steel substrates. The sizes of the selenium nanoclusters on

stainless steel substrates were similar to those on titanium substrates.

Water contact angles were measured on the stainless steel substrates. Uncoated

stainless steel substrate was hydrophilic and had a contact angle of 70.2±2 degrees.

Coating stainless steel substrates with selenium nanoclusters makes the substrates

become more hydrophobic (Figure 2.23).

Figure 2.23. Water contact angles on stainless steel substrates. Selenium nanocluster coating makes the stainless steel substrates become more hydrophobic.

Contact angles on the stainless steel substrates coated with three different

selenium nanocluster densities did not show significant difference. This trend was also

observed in titanium substrates coated with selenium nanoclusters. Again, this trend is

hypothesized to be attributed to the formation of air-pockets when a drop of liquid was

placed on the surface to measure the contact angle. Detailed discussion of the contact

angle results is presented in Chapter 4.

Page 65: Nanostructured Selenium for Biomedical Applications: from

48

2.4.2. Cell experiments on stainless steel substrates

After 3 days of culture, there was no significant difference in non-cancerous

osteoblast densities among the substrates (i.e., selenium-coated stainless steel and

uncoated stainless steel) (Figures 2.26 and 2.27)

Figure 2.24. Maintained healthy osteoblast densities on selenium-coated stainless steel substrates compared to the uncoated substrate after 3 days of culturing. Data = mean ± standard error of the mean; N=3.

Page 66: Nanostructured Selenium for Biomedical Applications: from

49

Figure 2.25. Representative fluorescence images of healthy osteoblast densities on uSS (A), Low-nSe-SS (B), Medium-nSe-SS (C) and High-nSe-SS (D). Magnification = 10X.

Compared to results on titanium substrates, this result showed that the stainless

steel substrates coated with selenium nanoclusters maintained densities of healthy

osteoblasts. This means that the coated stainless steel substrates had similar

biocompatibilities toward healthy bone cells compared to the uncoated substrate. It was

Page 67: Nanostructured Selenium for Biomedical Applications: from

50

hypothesized that the selenium nanocluster coating densities were not high enough to

produce significantly promoted healthy osteoblast functions on stainless steel substrates.

In contrast, for cancerous osteoblasts, after 3 days of culture, cell densities were

significantly lower on High-nSe-SS and Medium-nSe-SS than on Low-nSe-SS, uSS

(Figures 2.28 and 2.29). Cell density was decreased approximately 1.5 times on High-

nSe-SS and 1.2 times on Medium-nSe-SS compared to that on uSS. In addition,

comparing between High-nSe-SS and Medium-nSe-SS substrates, High-nSe-SS had a

significantly lower number of cancerous osteoblasts than Medium-nSe-SS substrates.

Figure 2.26. Decreased cancerous osteoblast densities after 3 days of culturing on Se coated stainless steel. Cell densities on Medium-nSe-SS and High-nSe-SS were significantly less than those on uSS and Low-nSe-SS. * p<0.05 compared to uSS and Low-nSe-SS; ** p<0.01 compared to Medium-nSe-SS. Data = mean ± standard error of the mean; N = 3.

Page 68: Nanostructured Selenium for Biomedical Applications: from

51

Figure 2.27. Representative fluorescence images of cancerous osteoblast densities on uSS (A), Low-nSe-SS (B), Medium-nSe-SS (C) and High-nSe-SS (D). Magnification = 20X.

This trend (i.e., decreased cancerous cell densities on substrates with increased

selenium coating densities) was similar to what observed in titanium samples. However,

the level of decrease in cancerous cell densities on selenium-coated stainless steel

substrates (compared to uncoated stainless steel substrate) was much less than the level of

decrease in cancerous cell densities on selenium-coated titanium substrates (compared to

Page 69: Nanostructured Selenium for Biomedical Applications: from

52

uncoated titanium substrate); High-nSe-SS had cell density decreased 1.5 times compared

to uSS while High-nSe-Ti had cell density decreased 4.5 times compared to uTi. Again,

this difference in inhibition of cancerous cell growth between selenium-coated stainless

steel substrates and titanium substrate is likely due to the smaller coating densities on

stainless steel substrates.

2.5. Conclusions and summary

To investigate the potential of using selenium nanoclusters as a coating material

for orthopedic applications, selenium nanoclusters were coated on conventional

orthopedic materials including titanium and stainless steel. The coatings on titanium

samples were investigated in more details because titanium was chosen as the model base

material due to its popularity in orthopedics. Titanium samples with higher coating

densities of selenium nanoclusters showed greater osteoblast adhesion, proliferation and

differentiation compared to lower coating densities of selenium nanoclusters as well as

uncoated titanium. Proliferation of cancerous osteoblasts was reduced on the samples

coated with selenium nanoclusters. The inhibition of cancerous osteoblasts is likely due

to the release of selenium into the culture media from the coatings. The promoted

functions of healthy osteoblasts on titanium surfaces coated with selenium nanoclusters

can be correlated to the increased adsorption of fibronectin which is an important protein

mediating osteoblast adhesion on implant surface.

Stainless steel samples were also investigated to demonstrate that the coating

process can be applied to other orthopedic metals. Stainless steel samples were coated

with selenium nanoclusters with various coating densities. The selenium coated samples

showed maintained proliferation of healthy osteoblasts compared to uncoated samples.

Page 70: Nanostructured Selenium for Biomedical Applications: from

53

Importantly, similar to the results obtained in selenium coated titanium samples,

proliferation of cancerous osteoblasts was inhibited on selenium coated stainless steel

samples.

Together, the results of this chapter suggest that selenium nanoclusters is a

promising coating material for promoting healthy bone growth while inhibiting cancer

growth for orthopedic applications. To fully understand the observed healthy osteoblast

functions, more intensive studies of protein adsorption on selenium coated surfaces from

other adhesion proteins such as vitronectin and laminin may be required because they are

also important adhesion proteins for osteoblast functions. To fully understand the

observed inhibition of cancerous osteoblasts functions on selenium-coated substrates,

more in-depth analysis of effects of selenium released into culture media on cancerous

osteoblasts should be conducted.

Page 71: Nanostructured Selenium for Biomedical Applications: from

54

CHAPTER 3. COARSE - GRAINED MONTE CARLO COMPUTER SIMULATION OF FIBRONECTIN ADSORPTION ON NANOMETER

ROUGH SURFACES

3.1. Introduction

It was shown in Chapter 2 that increasing selenium nanocluster coating density

led to the increase in fibronectin adsorption and that the increased fibronectin adsorption

promoted healthy osteoblast functions. The objective of this part of the thesis is to

understand the factor that gave rise to increased fibronectin adsorption. The increase in

fibronectin adsorption might be attributed to factors such as substrate surface

hydrophobicity/hydrophilicity, morphology, roughness or a combination of these factors.

For example, elongated proteins (such as fibronectin) can adsorb in different directions.

They may adsorb onto a surface with the main axis oriented perpendicularly to or along

the surface. A rough surface with the rough features in the same length scale with the

proteins may have more proteins adsorbed perpendicularly to the surface because of the

increase in steric hindrance caused by the increased surface roughness. A protein

adsorbed perpendicularly to the surface will occupy less area on the surface. Therefore,

the surface with increased roughness may have increased protein adsorption compared to

a flat surface.

To further understand the influence of surface roughness on fibronectin

adsorption, in this chapter, a coarse-grained Monte Carlo computer simulation was used

to reproduce experimental results. First, a general

Page 72: Nanostructured Selenium for Biomedical Applications: from

55

introduction of protein adsorption onto a biomaterial surface is presented, followed by a

quick summary of some of the established physical models of protein adsorption. Then a

simple computer method was used to simulate rough surfaces. Finally, surfaces with pre-

determined root mean square (RMS) roughness were generated and used for simulating

fibronectin adsorption. Simulation results and experimental results were then compared.

3.2. Protein adsorption onto biomaterials

Protein adsorption onto solid surfaces is a complex yet interesting problem for

many fields of studies (such as biotechnology, biomedicine, biology, biochemical

engineering etc.). Especially, in biomaterial applications, where a material is designed to

interact with biological objects, protein adsorption on a surface plays an important role as

it is the first event that occurs when a material is placed in contact with biological fluids.

This event will affect subsequent reactions of biological entities (such as cells and

tissues) with a material. For example, when a material is inserted into humans or animals,

proteins adsorb onto the surface of the material almost instantaneously (within seconds)

after implantation. Cells arrive to the surface after much longer times (on the order of

minutes). Therefore, cells will interact directly with the adsorbed protein layer instead of

the material. Thus, early cellular events (such as cellular adhesion and morphology)

depend largely on the adsorbed protein layers (specifically, what proteins are present,

their amount and bioactivity).

The interaction between proteins and a solid surface depends on both properties of

the protein (such as hydrophobicity/hydrophilicity, charge, conformation, etc.) and the

properties of the surface (such as the hydrophobicity/hydrophilicity, elemental

composition, charge, roughness, morphology, etc.). Protein adsorption is a very complex

Page 73: Nanostructured Selenium for Biomedical Applications: from

56

process driven by many surface-protein interactions (such as hydrophobic/hydrophilic

interactions, electrostatic interactions, van der Waals interactions and formation of ionic,

hydrogen or covalent bonds [70]). For example, hydrophobic regions on a protein adsorb

to hydrophobic surface regions and similarly, hydrophilic regions of a protein adsorb to

hydrophilic surface regions. Since proteins usually have regions with different charges

(even if the whole protein is neutral), electrostatic interactions play an important role if

the solid surface acquires some surface charge in solution. The electrostatic interaction

usually becomes complex if there are ions (such as calcium ions) in solution which will

also interact with the protein and the solid surface.

Fibronectin is one of the most important adhesive proteins that mediate

anchorage-dependent cell (such as osteoblasts, fibroblasts, endothelial cells, etc.)

adhesion. Fibronectin is a high molecular weight extra-cellular matrix protein consisting

of two nearly identical ~ 250kDa monomers connected by disulfide bonds [71] (Figure

3.1).

Page 74: Nanostructured Selenium for Biomedical Applications: from

57

Figure 3.1. Schematic structure of fibronectin showing two monomers linked by disulfide bonds (Reprinted with permission from [72]).

Within seconds after implantation, proteins (including fibronectin) arrive and

adsorb onto the surface of the implant. Anchorage-dependent cells arrive at the implant

surface later and, therefore, will interact directly with the adsorbed protein layer. Among

many proteins that can adsorb on the surface of an orthopedic implant material (such as

albumin, laminin, vitronectin, etc.), adhesive proteins (such as fibronectin and

vitronectin) play an important role for cell adhesion which is a crucial prerequisite for

subsequent cell functions (such as proliferation, and for osteoblasts, synthesis of

extracellular matrix proteins and formation of calcium-containing mineral deposits).

These adhesive proteins have certain peptide sequences that specific cells can recognize

and adhere to. For example, osteoblasts recognize and adhere to the arginine-glycine-

aspartic acid (RGD) peptide sequence in fibronectin. Studies have shown that interrupting

Page 75: Nanostructured Selenium for Biomedical Applications: from

58

the interactions between osteoblasts and fibronectin in cell culture media results in a

sharp decrease in osteoblast functions. For example, osteblasts were grown in culture

media added with anti-fibronectin (an antibody to fibronectin at a concentration of

100µg/mL) and showed a 10% reduction of mineralization by osteoblasts compared to

cells grown in control conditions (i.e., culture media without anti-fibronectin) [73].

Interfering with fibronectin-osteoblast interactions was also reported to result in reduced

osteoblast attachment [74] and even complete nonadherence of osteoblasts to implant

materials [75]. Coating a material surface with fibronectin (at a concentration of

~10µg/mL) has been a method to increase the attachment of osteoblasts to an implant

surface [74, 76-79]. The concentration of fibronectin in plasma is from 300-400 µg/mL

[80]. Osteoblasts have also been shown to adhere more strongly to higher density

fibronectin-coated surfaces [81, 82]. Specifically, using the micropipette aspiration

technique, osteoblasts were determined to adhere 80% more strongly on a titanium

surface coated with fibronectin compared to an uncoated substrate [82].

3.3. Roles of rough surfaces

Rough surfaces are encountered more often than flat surfaces. More strictly

speaking, there is hardly any surface which is perfectly flat. For many problems where

interaction of molecules (such as proteins) with a surface is investigated using computer

simulation, there is a great interest in generating rough surfaces with pre-determined

properties such as the dimension of surface structures (for example, pores, groves, etc.),

hydrophobicity/hydrophilicity properties, the roughness of the surfaces, etc. Among

many surface properties that require simulating, surface roughness is a very important

factor. Surface roughness can affect electrical properties in semiconductor materials

Page 76: Nanostructured Selenium for Biomedical Applications: from

59

(such as electrical capacity, electronic conductivity, peak electric field and sheet

resistance [91]). In biomaterials, surface roughness can influence the adsorption,

conformation, and denaturation of proteins when the material is placed in contact with

biological fluids (as this thesis has demonstrated). In addition, there is a growing interest

in computer simulation of the interactions of biological entities (such as proteins and

cells) with biomaterial surfaces because it is important to have a mechanistic

understanding of these interactions in order to engineer desired materials. However, most

of the methods employed in the literature (for example, the Fast Fourier Transform (FFT)

method and artificial neural networks method) are complex and involve complicated

mathematic manipulations. In the next section, a simple method to generate rough

surfaces with pre-determined RMS values is proposed.

3.4. Simple method to generate rough surfaces with pre-defined RMS roughness

To generate a rough surface having a random morphology, but defined RMS

roughness, the idea of variance of uniformly-distributed random numbers was employed

here. Consider a set of random variables, X, which have a uniform continuous distribution

in the interval [a,b]: {X}=U[a,b]:

bxaforab

bxoraxforxf1

0)(

(Equation 3.1)

where f(x) is the probability function, a and b are the two values defining the interval

(a<b).

Page 77: Nanostructured Selenium for Biomedical Applications: from

60

The variance of X, defined as Var(X)=E[X-E(X) 2], where E(X) is the mean of X,

is:

12

)(2

)(

abXVar uniform

(Equation 3.2)

If instead of a continuous uniform distribution, X has a discrete uniform

distribution in the interval [a,b]; i.e., {X}=Udiscrete[a,b]:

bkaforn

bkorakforkp1

0)((Equation 3.3)

where n = b-a+1 and p(k) is the probability function, the variance of X is given as:

12

1)(

2

)(

nXVar discrete

(Equation 3.4)

Now, the explicit expression of the variance of discrete, uniformly distributed

variables X is:

N

xxXVar

N

imeani

1

2)()(

(Equation 3.5)

Now, consider the definition of RMS:

Page 78: Nanostructured Selenium for Biomedical Applications: from

61

N

xxRMS

N

imeani

1

2)((Equation 3.6)

It can be seen that:

)(XVarRMS (Equation 3.7)

Therefore, a random, rough surface with a pre-defined RMS roughness can be

generated by using the concept of uniformly distributed discrete variables. Specifically,

the problem of generating a 3D rough surface with RMS roughness values of R using

Matlab consists of the following steps:

Step 1: Determine the interval [0, a] of random numbers to be generated. The

value of a is determined from the requirement that the RMS is equal to R:

12)(

aXVarRMSR

(Equation 3.8 )

Step 2: Define a dimensionless roughness parameter, r, which is the normalization

(i.e., dividing) of the actual RMS value R by the dimension Do of an appropriately chosen

parameter. For example, if adsorption of a specific protein onto the surface is the problem

of interest, Do can be the smallest dimension of the protein. Do is the unit length in the

simulation (i.e., the smallest length scale in the problem).

0D

ar

(Equation 3.9)

Page 79: Nanostructured Selenium for Biomedical Applications: from

62

Step 3: Generate a matrix NxN of random numbers, X, that have a uniform,

continuous distribution in [0, r].

Step 4: Convert X into X* whose elements are the nearest integers greater than or

equal to elements in X.

If N is large enough, the X* “surface” will have a random morphology with the

RMS value r.

The method described above was used here to create the High-nSe-Ti surface in

this study with an RMS value, R, equal to 14.113 nm. From the value of R, the value of a

was calculated to be equal to 48.878nm from the equation R2=a2/12. If the problem of

interest in this study is the adsorption of a globular protein whose diameter Do is 9nm, the

dimensionless parameter, r, is calculated by dividing a by Do, or r = 5.431. Now, a N x N

matrix of random variables X whose distribution is uniform in the interval [0, 5.431] can

be generated by using the rand() function in Matlab: X=5.431 rand(N,N). However, in

order to use the surface in computer simulation, the height of the points on the surface

needs to be discretized, this can be done by using the ceil() function in Matlab which

rounds up the value of the height to the nearest integer greater than or equal to the height:

X*=ceil(X). As can be from Table 3.1, the method converges as N increases.

Page 80: Nanostructured Selenium for Biomedical Applications: from

63

Table 3.1. Convergence property of the proposed method to generate surface High-nSe-Ti with a defined RMS roughness. Dimensionless simulated roughness, Ro, is the square root of variance of the heights of all points on the simulated surface. Simulated roughness is then determined as R=RoDo (in this case, Do=9nm).

N Dimensionless simulated

roughness, Ro Simulated

roughness, R(nm) Actual

roughness (nm)50 1.594 14.346

100 1.576 14.184 250 1.575 14.175 500 1.570 14.130

14.113

This method was used to generate surfaces having RMS roughness values the

same as the experimental substrates (i.e., uTi, Low-nSe-Ti, Medium-nSe-Ti and High-

nSe-Ti) with unit length Do=9nm and N=500. The generated surfaces are referred to as

S-uTi, S-Low-nSe-Ti, S-Medium-nSe-Ti and S-High-nSe-Ti. Figure 3.2 shows

representative regions on the generated surfaces.

Page 81: Nanostructured Selenium for Biomedical Applications: from

64

Figure 3.2. Representative regions on the four rough surfaces generated in this study with the RMS roughness values equal to experimental values (unit lengh Do=9nm, N=500). Colors are to aid in visualization only.

3.5. RSA model of fibronectin adsorption on uncoated and selenium coated titanium

surfaces

The ELISA method used in this study to detect and quantify the amount of the

adsorbed fibronectin on substrate surfaces relies on the recognition and attachment of

antibodies to proteins. As the result, the ELISA method detects only the top layer of

adsorbed proteins (if there is multi-layer adsorption). Therefore, the simulation in this

study will assume a monolayer adsorption of fibronectin on the substrates.

Page 82: Nanostructured Selenium for Biomedical Applications: from

65

The concentration of fibronectin used in this study (i.e., adsorption onto surfaces

of fibronectin from the solution of fibronectin in PBS) was 5µg/mL. This value of

fibronectin concentration is smaller than physiological concentrations of fibronectin

(which is ~ 300 µg/mL in blood plasma) but is close to the concentration of fibronectin in

the cell culture media (i.e., DMEM supplemented with 10% FBS, 1% P/S, which is the

optimal medium for culturing osteoblasts in vitro has approximately 2.5 µg/mL of

fibronectin). This concentration of fibronectin (i.e., 5µg/mL) is in the range of

concentrations that the RSA model fits best [89, 92]. In addition, many reports in the

literature use the RSA model to describe the adsorption of proteins onto solid surfaces

with good agreement between experimental results and predictions of the RSA model

[93-95]. The size of fibronectin depends on factors such as its folding and the

surrounding environment ionic strength, therefore, dimensions of fibronectin reported in

the literature vary significantly. Unfolded fibronectin is 140nm in length and 2nm in

diameter [96]. Fibronectin is in more compact form in near - physiological conditions

[97]. Fibronectin were reported to be in a compact, elongated structure with the length of

approximately 15nm and the width of approximately 9nm (Figure 3.3)[98].

Page 83: Nanostructured Selenium for Biomedical Applications: from

66

Figure 3.3. Gallery of transmission electron microscopy images of individual fibronectin molecules freeze-dried in vacuum and shadowed by tungsten-tantalum. (Freeze-drying technique was used to preserve the native structure of fibronectin and tungsten-tantalum shadowing was used to give contrast for imaging). Most fibronectin molecules had elongated shapes with the length of ~ 15 nm and the width of ~9 nm. Magnification is 300,000. (Adapted with permission from [98]) .

An elongated shape of fibronectin molecule in physiological pH (7.4) with the

length of ~ 24 nm and the width of ~ 16 nm was also observed using transmission

electron microscopy [99]. Therefore, in this thesis, fibronectin molecule was modeled as

a hard rod with dimensions of 18nm x 9nm x 9nm. The smallest dimension value, 9nm,

was chosen to be the width of fibronectin reported in [98] and the length of the rod,

18 nm, was chosen to be near the average of the two values of the length of fibronectin

molecule reported in [98] (15 nm) and [99] (24 nm). Desorption and diffusion of

Page 84: Nanostructured Selenium for Biomedical Applications: from

67

adsorbed protein on the surface was assumed to be negligible. Fibronectin is assumed to

adsorb randomly on the surface with equal probability, i.e., any point on the surface can

be the point for protein adsorption. For simulation, a fibronectin molecule was modeled

as a dimmer consisting of 2 monomers (see Figure 3.4).

Figure 3.4. A 2D picture of fibronectin adsorbing onto a rough surface. Fibronectin was modeled as a dimmer consisting of 2 monomers and the surface was modeled as consisting of columns. Each point on the surface was defied by two coordinates (i,j). Fibronectin can adsorb on all the "unoccupied" sites (in this picture, the right and the left sides of point (i,j) are unoccupied, the top and bottom sides of point (i,j) are occupied).

In a 3D simulation, each point on the surface was specified by 3 coordinates

(x,y,z) where z is the height. Fibronectin was assumed to adsorb sequentially onto the

surface. Each adsorption trial started by randomly choosing a point (specified by x, y and

z) on the rough surface for one end of the protein to “attach” to (this end will then have a

“contact”). The direction of the protein was then varied. If the adsorption of one of the

monomers was impossible due to steric constraints, the trial was rejected. If steric

constraints were satisfied, proteins would adsorb in the direction that had more contacts.

Figure 3.5 demonstrates examples to clarify this description.

Page 85: Nanostructured Selenium for Biomedical Applications: from

68

Figure 3.5. Examples of the simulation of protein adsorption onto a rough surface. For the case of the right side of the point specified by (i,j) (the bold side), there are two possible directions the protein can adsorb: horizontal direction (a) or vertical direction (b) (Note, horizontal adsorption to the left of the bold side is prohibited due to steric constraints). In this case, configuration (a) will have 4 contacts while (b) will have only 2 contacts upon adsorption, therefore, configuration (a) is more favorable for protein adsorption. For the case of the point on the far right (bold side), both the (c) and the (d) configuration will have the same number of contacts (3 contacts) upon adsorption, therefore, the protein will adsorb following either (c) or (d) with equal probability.

“Overhang” is not allowed, i.e., it is prohibited to have protein adsorption at a site

in which the site immediately underneath is not occupied. Overhang will be immediately

shifted downward to reach the nearest occupied site. Figure 3.6 demonstrates examples of

overhangs which would be immediately shifted downward (dashed arrows).

Page 86: Nanostructured Selenium for Biomedical Applications: from

69

Figure 3.6. Overhang is not allowed. Overhang would be immediately shifted downward as indicated by the dashed arrows.

To generate surfaces of uncoated and selenium-coated titanium substrates, the

method described in section 3.5 was employed here. Specifically, with the RMS

roughness determined experimentally, the parameters for the surface were calculated and

summarized in (Table 3.2).

Table 3.2. Parameters for simulation (a) calculated from experimental RMS roughness as described in section 3.5 and the RMS roughness of the simulated surfaces. The simulated surfaces resemble the experimental surfaces quite well in terms of RMS roughness.

Sample Experimental RMS roughness (nm)

a Simulated RMS roughness (dimensionless) L=500x500

Simulated RMS roughness (nm) L=500x500

uTi 8.07 27.97 0.88 7.93 Low-nSe-Ti 9.72 33.67 1.08 9.73 Medium-nSe-Ti 12.56 43.53 1.34 12.51 High-nSe-Ti 14.11 48.88 1.58 14.22

Page 87: Nanostructured Selenium for Biomedical Applications: from

70

As mentioned, the generated surfaces will be referred to as S-uTi (simulated-uTi),

S-Low-nSe-Ti, S-Medium-nSe-Ti and S-High-nSe-Ti. Simulation was implemented with

size of the substrate NxN=500x500 and the total number of adsorption trials,

M=6,250,000. With this choice of N and M, each point on the surface of all the substrates

had at least 5 protein adsorption trials. This choice of M was shown (by simulation) to be

efficient enough to achieve saturation adsorption on all the surfaces (as indicated by less

than 2.6% of the surface area unoccupied after M adsorption trials).

The amount of adsorbed protein was normalized by a monolayer of proteins

adsorbed on a perfectly flat surface which had the same geometric surface area.

Simulation was implemented first in 2D to verify the validity of the algorithm. A

snapshot of the surfaces during the simulation is shown in Figure 3.7.

Page 88: Nanostructured Selenium for Biomedical Applications: from

71

Figure 3.7. A snapshot of simulation process in 2D with N=500 and M=25000. Blue columns represent proteins. Grey columns represent rough surfaces. The adsorption did not reach saturation yet as some areas on the surface are still unoccupied.

Page 89: Nanostructured Selenium for Biomedical Applications: from

72

With the choice of N x N and M indicated, simulations in 3D were implemented

with 3 runs for each substrate; results reported as the average of 3 runs are shown in

Figure 3.8.

Figure 3.8. Simulation results of fibronectin saturated adsorption on all four substrates of interest. The protein amount was normalized to a monolayer of protein on a perfectly flat surface with the same geometrical surface area N x N. Substrates were generated to have experimental RMS roughness values. Protein adsorption simulation was implemented with the size of the substrates N x N=500x500 and the number of simulation M=6,250,000. Data = mean ± standard error of the mean (n=3 runs).

Simulation results showed a small increase in fibronectin adsorption on the Low-

nSe-Ti substrate compared to uTi. However, fibronectin adsorption on Medium-nSe-Ti

and High-nSe-Ti significantly increased compared to both uTi and Low-nSe-Ti. This

trend is similar to the trend in the experimental results (Figure 2.23 in Chapter 2).

To better compare the simulation results with experimental results, the results

were expressed as the increase compared to uTi (the number of times the amount of

fibronectin increased compared to the uncoated substrate). As shown in Figure 3.9,

computer simulation results fit quite well to the experimental results.

Page 90: Nanostructured Selenium for Biomedical Applications: from

73

Figure 3.9. Comparison of the increase in the amount of fibronectin adsorption (expressed as times increase compared to uTi) between simulation results and experimental results on uncoated and selenium coated titanium substrates.

The good agreement between computer simulation and experimental results

indicated that the simple RSA model employed in this study is appropriate to describe the

adsorption process of fibronectin on nanometer rough surfaces. It is appropriate here to

evaluate some other physical quantities from the simulation (such as the amount of

adsorbed fibronectin normalized to geometrical surface area of the substrate). Recall that

the unit length in the simulation corresponded to 9nm in the experiment. The molecular

weight of fibronectin is approximately 440kDa. From the total protein adsorbed on the

substrates, the normalized protein amount (expressed as µg/cm2) was calculated and

presented together with experimental results for comparison in Figure 3.10.

Page 91: Nanostructured Selenium for Biomedical Applications: from

74

Figure 3.10. Simulation and experiment results on the amount of adsorbed fibronectin normalized by geometrical surface area. The simulation values of normalized fibronectin adsorption are greater than experimental values.

As can be seen from Figure 3.10, simulation results are approximately 80% higher

than corresponding experimental results. The difference in the values for the simulation

and experiment results is an indication of the simplicity of the RSA model employed in

this study. However, the trend is similar for both simulation and experiment results (as

can be seen from Figure 3.8 and previously analyzed and presented in Figure 3.9). It is

likely that the higher values of the simulation results on fibronectin adsorption compared

to experimental results are due to the fact that desorption of adsorbed fibronectin was

ignored in the simulation model.

Another interesting physical quantity that can be extracted from the computer

simulation is the amount of adsorbed protein per unit area of the rough surface. As the

roughness of a surface is increased, the surface area of the surface also increases. Assume

that the “binding site” densities on a rough surface and a perfectly flat surface (of the

same material and the same geometrical area) are the same. Intuitively, it is expected that

Page 92: Nanostructured Selenium for Biomedical Applications: from

75

the densities of a protein adsorbed on the flat surface and the rough surface (per unit

surface area) are the same (Figure 3.11).

Figure 3.11. Protein adsorption on a flat surface and a rough surface. Assuming the same "binding site" densities, the protein density (per unit surface area) d1 and d2 should be the same.

However, as seen in Figure 3.12, computer simulation predicts that surfaces with

increased roughness will have decreased adsorbed protein density per unit surface area.

Figure 3.12. Computer simulation results on fibronectin density expressed as microgram of fibronectin per cm2 of the surface area. Surfaces with higher RMS roughness had lower fibronectin density.

This prediction by computer simulation can be understood as follows. The model

employed in this study assumes that proteins will try to maximize contact with the

Page 93: Nanostructured Selenium for Biomedical Applications: from

76

substrate during adsorption. In addition, in this model, proteins can only adsorb in one of

two directions: horizontal direction and vertical direction. Therefore, an increase in

surface area will not necessarily lead to an equal increase in protein adsorption. An

example showing that increases in surface area do not necessarily lead to equal increases

in protein adsorption is shown in Figure 3.13.

Figure 3.13. An example illustrating that increased surface area does not necessarily lead to equal increase in protein adsorption. In this example, the rough surface area increases ~67% but the amount of fibronectin increases only 25%.

Future experiments that can accurately determine the density of adsorbed

fibronectin per unit surface area (rather than per unit geometrical surface area) need to be

implemented to verify this prediction.

3.6. Summary and conclusions

In this chapter, a simple method for generating rough surfaces with pre-

determined RMS roughness was proposed using the idea of uniformly-distributed random

variables. The method was used to generate four substrates with RMS roughness values

equal to the experimental values of the uncoated and selenium coated titanium substrates.

A simple random sequential adsorption model was used to simulate the adsorption of

Page 94: Nanostructured Selenium for Biomedical Applications: from

77

fibronectin on the generated surfaces. Simulation results of the increase in fibronectin

adsorption on coated surfaces compared to uncoated surfaces were found to be in a good

agreement with experimental results. Further development is needed to generate rough

surfaces, for example, to simulate not only random but also ordered-rough surfaces. The

RSA model also needs to be improved to include more complex process in the adsorption

(such as desorption, diffusion, and denaturation of the adsorbed proteins). For example,

to account for denaturation process, the adsorbed proteins can be modeled to have a size

that is larger than that of the non-adsorbed ones.

Page 95: Nanostructured Selenium for Biomedical Applications: from

78

CHAPTER 4. A MODIFIED CONTACT ANGLE MODEL TO FIT EXPERIMENTAL RESULTS

In this chapter, a brief introduction to the roles of contact angles in biomaterials

will be presented and followed by an introduction to two theoretical models of contact

angles (i.e., the Wenzel model and the Cassie-Baxter model). After that, the needs to

modify the Wenzel model and Cassie-Baxter model to fit the experimental results on

contact angles presented in Chapter 2 will be discussed. Finally, a modified contact angle

model will be proposed to fit the experimental results.

4.1. Introduction: role of contact angles in biomaterials

Contact angles easily and quickly identify the hydrophobicity/hydrophilicity of a

proposed material for medical applications; this is of paramount importance for the

interaction of a material with biological entities in the body. When an implant is inserted

into humans or animals, proteins absorb onto the surface of the implant within seconds

after implantation. Cells arrive at the implant surface much later (within minutes),

therefore, directly interacting with such absorbed proteins. As the result, early responses

of cells to a biomaterial (such as adhesion, migration, spreading, etc.) depend largely on

the adsorption of proteins (i.e., which type, amount and consequent bioactivity) onto the

material surface which, in turn, depends on the specific properties of the material surface

(such as hydrophilicity/hydrophobicity, surface charge, surface morphology, surface

roughness, etc.). Among the protein-surface interactions, hydrophobic/hydrophilic

Page 96: Nanostructured Selenium for Biomedical Applications: from

79

interactions play an important role. Hydrophilicity and hydrophobicity are clearly the

water-attracting and water-repelling properties of a material. If the material is water-

repelling, it is hydrophobic. If it is water-attracting, it is hydrophilic. A more rigorous

definition of hydrophilic and hydrophobic materials will be given in the next section.

Nanostructured materials, especially materials having surface with nanofeatures have

increased surface area, different surface morphology, surface roughness, etc., and,

importantly, will have different hydrophilicity properties compared to flat surfaces (of the

same chemistry).

Proteins usually have both hydrophobic regions and hydrophilic regions (for

example, see Figure 4.1 for the hydrophobic and hydrophilic regions in a fibronectin

molecule). Hydrophobic regions on a protein attach to hydrophobic regions on a material

surface and, vice versa, hydrophilic regions attach to hydrophilic regions on material

surface.

Page 97: Nanostructured Selenium for Biomedical Applications: from

80

Figure 4.1. Atomic and ribbon chains of an extended human fibronectin molecule. Inset shows that hydrophilic parts are blue colored regions while hydrophobic parts are red colored in modules (Reprinted with permission from [100]).

In this part of the study, an introduction to contact angle fundamentals and the

Young’s equation will be discussed first followed by a brief introduction into two

physical models of contact angles on rough and composite surfaces. Finally, a modified

contact angle model will be proposed and discussed to fit the experimental results from

this study.

Page 98: Nanostructured Selenium for Biomedical Applications: from

81

4.2. Contact angle and Young’s equation

The hydrophilicity/ hydrophobicity of a material is usually determined by

measuring contact angles of drops of liquid on the material surface (as illustrated in

Figure 4.2).

Figure 4.2. Contact angle, θ, of a liquid on a solid surface depends on the interaction of

the liquid with the solid, liquid with vapor and solid with vapor. SL , SV , and LV are interfacial energies at the solid-liquid interface, solid-vapor interface and liquid-vapor interface, respectively.

If the material is water-repelling (hydrophobic), a water drop on the surface of a

material will have a contact angle, θ, greater than 90 degrees. Otherwise, if the material is

water-attracting (hydrophilic), the contact angle will be less than 90 degrees. Consider a

drop of liquid on a solid substrate. The contact angle at equilibrium is θ. Imagine if the

liquid-solid interface is displaced by dx, the increase in the liquid-vapor interface is

dx.cos(θ) (Figure 4.3).

Page 99: Nanostructured Selenium for Biomedical Applications: from

82

Figure 4.3.Schematic demonstration of the derivation of Young's equation. When the solid-liquid interface is displaced by dx, the increase in liquid-vapor interface is dx.cos(θ).

The change in total energy associated with this displacement is

)cos()( dxdxdE LVSVSL (Equation 4.1)

A stable contact angle corresponds to dE/dx=0:

0)cos()( LVSVSL(Equation 4.2)

or

LV

SLSV

)cos((Equation 4.3)

again where SL , SV , and LV are interfacial energies at the solid-liquid interface, solid-

vapor interface and liquid-vapor interface, respectively. This is Young’s equation of

contact angle.

Young’s equation provides a quantitative relationship between the contact angles

and physical properties of a solid, liquid and vapor. In Young’s equation, the solid

surface is assumed to be perfectly flat.

4.3. Wenzel model versus Cassie-Baxter model on contact angles

To describe the effects of different surface roughness on the wetting behavior of a

solid surface, Wenzel considered a drop of water on a solid surface. The surface energy

Page 100: Nanostructured Selenium for Biomedical Applications: from

83

of the wetted area (i.e., under the drop), σwetted, will be, in general, different from that of

the non-wetted area (i.e., not covered by the drop), σnon-wetted. Let’s assume for a moment

σwetted < σnon-wetted . In this case, the wetting process is involved in the decrease of the

surface energy or wetting is energetically favorable, in another words, the surface is

hydrophilic. Now, if the surface under the drop is rough instead of flat, the decrease in

surface energy of the wetting process will be greater. Therefore, rough surfaces will be

more easily wetted, or more hydrophilic.

The same reasoning applies to the case of a surface that has σwetted > σnon-wetted , or

a water-repelling (hydrophobic) surface. There will be a net increase in surface energy

during the wetting progresses. If the surface under the drop is rough instead of flat, the

actual wetted surface area will be greater than the geometric area. As the result, a net

increase in surface energy during wetting will be greater. In this case, roughness will

make the surface more hydrophobic.

Wenzel introduced the concept of a “roughness factor”, r, that describes the

increase in surface area of a rough surface compared to the geometric area:

areasurfacegeometric

areasurfaceactualr

(Equation 4.4)

where r is always greater than 1 except in the case of a perfectly flat surface where the

actual surface and geometric surface are identical.

Now, consider a drop of a liquid on a rough surface which has roughness factor of

r. At equilibrium, the contact angle is θ. Imagine if the liquid-solid interface is displaced

Page 101: Nanostructured Selenium for Biomedical Applications: from

84

by dx, the increase in the liquid-vapor interface is dx.cos(θ), while the actual liquid-solid

interface increase associated with this displacement is r.dx (Figure 4.4).

Figure 4.4. Schematic demonstration of the derivation of the Wenzel equation. When the solid-liquid interface is displaced by dx, the increase in liquid-vapor interface is dx.cos(θ), while the actual liquid-solid interface increase associated with this displacement is r.dx.

The total energy change is then:

)cos()( dxrdxdE LVSVSL (Equation 4.5)

where E is the total energy of the system, SL , SV , and LV are the surface tensions at

the solid-liquid interface, solid-vapor interface and liquid-vapor interface, respectively.

A stable contact angle θ* corresponds to dE/dx=0; or

0)cos()( * LVSVSL r (Equation 4.6)

or:

Page 102: Nanostructured Selenium for Biomedical Applications: from

85

)cos()(

)cos( *

rr

LV

SLSV

(Equation 4.7)

where θ is the contact angle of the same solid, but with a smooth surface.

The above equation indicates that if the surface which is hydrophilic (contact

angle θ < 90 degrees, cos(θ) >0) is roughened, cos(θ*) will be greater than cos(θ), or θ*

will be smaller than θ or the surface becomes more hydrophilic after being roughened. In

the other words, hydrophilic surfaces become more hydrophilic after being roughened.

Similarly, if θ is greater than 90 degrees, or cos(θ) < 0, cos(θ*) will be more negative

(compared to cos(θ)), or θ* will become greater than θ (and greater than 90 degrees). In

the other words, hydrophobic surfaces become more hydrophobic after being roughened.

The underlying assumption in the Wenzel model is that the liquid-solid interface

follows the solid surface, i.e., wetting is homogeneous (Figure 4.5).

Figure 4.5. Homogenous wetting assumption in the Wenzel model. The liquid-solid interface is assumed to follow a solid surface.

Page 103: Nanostructured Selenium for Biomedical Applications: from

86

Cassie and Baxter proposed a model for the wetting on a heterogeneous surface

consisting of two different materials. The derivation of Cassie-Baxter equation is

summarized here. Consider a composite surface of two species with a surface fraction of

f1 and f2 (f1+f2=1). Suppose that the liquid drop is at equilibrium with a contact angle θ.

Now imagine the interfacial boundary is displaced by a small amount, dx (Figure 4.6).

Figure 4.6. A schematic demonstration of the derivation of the Cassie-Baxter model. When the solid-liquid interface increases by dx, the liquid-vapor interface increases by dx.cos(θ).

The energy change associated with this increase will be:

)cos()()( ,2,22,1,11 dxdxfdxfdE LVSVSLSVSL (Equation 4.8)

where SL , SV , LV are surface tensions at the solid-liquid interface, solid-vapor

interface and liquid-vapor interface, respectively. Indices 1 and 2 denote species 1 and 2,

respectively.

A stable contact angle θ corresponds to dE/dx=0, or

0)cos()()( ,2,22,1,11 LVSVSLSVSL ff (Equation 4.9)

Page 104: Nanostructured Selenium for Biomedical Applications: from

87

The above equation can be re-written to obtain the Cassie-Baxter equation as

follows:

0)cos()()( ,2,2

2,1,1

1

LV

SVSL

LV

SVSL ff(Equation 4.10)

or

)cos()cos()cos( 2211 ff (Equation 4.11)

The Cassie-Baxter model of wetting can be applied to the case of a rough surface

where the roughness can create air-pockets when wetting happens (Figure 4.7).

Figure 4.7. The Cassie-Baxter model of heterogeneous wetting. In this case, the liquid is considered to contact a surface consisting of the solid material and air. The width of spikes as well as the spacing between them is very small compared to the size of the liquid drop. The scale used in this picture is to aid in visualization.

Page 105: Nanostructured Selenium for Biomedical Applications: from

88

In this case, because the contact angle of a liquid drop on air is considered be 180

degrees, the above equation can be re-written as:

)180cos().1()cos(.)cos( 0111 ff (Equation 4.12)

or

1))cos(1()1()cos(.)cos( 11111 fff (Equation 4.13)

where f1 is the area fraction of the solid-liquid interface.

It was shown that the Wenzel model has been typically applied to surfaces with

low or moderate roughness (r less than 1.2) [101, 102]. However, the range of roughness

ratio, r, for the Cassie-Baxter “regime” is less consistent among different reports. Some

have reported the Cassie-Baxter regime at r values greater than 1.2 [101], others have

reported the regime at r values greater than 2.3 [101, 102]. In practical term, the latter is

usually used for surfaces that have spike-structures where air can be trapped between the

spikes.

4.4. The needs to modify the Wenzel model and Cassie-Baxter model to fit experimental

results

As presented in chapter 2, contact angle measurements on uncoated titanium and

selenium-coated titanium substrates possessed some interesting findings (Figure 4.8).

Page 106: Nanostructured Selenium for Biomedical Applications: from

89

Figure 4.8.Water contact angles on uncoated titanium and titanium coated with increasing selenium nanocluster densities. Data = mean ± standard error of the mean, n=5.

Specifically, uncoated titanium substrates were hydrophilic (i.e., had contact

angle less than 90 degrees). Coating titanium substrates with selenium increased the

substrate roughness (r greater than 1). Therefore, if differences in chemistry between

selenium and titanium are ignored, the coated substrates should become more hydrophilic

(i.e., the contact angle should decrease) according to the Wenzel theory. However, the

experimental data showed an opposite trend. Therefore, the differences in chemistry of

selenium and titanium should be taken into account. Thus, the Cassie-Baxter model,

which describes the contact angle of a composite surface as a function of contact angles

of the components, was used here to explain the wetting behavior of the selenium coated

surfaces:

)cos()cos()cos( 2211 ff (Equation 4.14)

where is the water contact angle on the selenium-coated titanium substrates, 1 is the

water contact angle on the uncoated titanium substrate, 2 is the water contact angle on

the selenium substrate, f1 is the surface fraction of titanium, and f2 is the surface area

Page 107: Nanostructured Selenium for Biomedical Applications: from

90

coverage fraction of selenium, ρ, (f1 +ρ =1); ρ was determined from SEM images and is

given in Table 4.1.

Table 4.1. The selenium surface fraction (ρ, determined from SEM images), relative surface roughness ratio (r, normalized by that of the uncoated substrates), the relative root mean square roughness (R, normalized to that of uncoated substrates), and cos(θ), of the substrates of interest to the studies.

Substrate Selenium surface fraction, ρ, (%)

Relative surface roughness ratio, r

Relative RMS, R

cos(θ)

uTi 0 1 1 0.433 Low-nSe-Ti 2.4 1.0373 1.204 0.274

Medium-nSe-Ti 5.0 1.050 1.556 0.262 High-nSe-Ti 6.2 1.054 1.748 0.220

The relative surface roughness ratio was calculated by dividing the surface

roughness ratio of the sample by the surface roughness ratio of the uncoated substrates.

Similarly, the relative RMS was calculated by dividing the RMS of the sample by the

RMS of the uncoated substrate. 2 was measured on a flat elemental selenium surface

using the same method as described in section 2.2.2.4. 2 was determined to be 83.5±1.4

degrees.

Using Equation 4.12, a cos( )-versus-f2 curve was constructed based on the

Cassie-Baxter model on the selenium-titanium composite surfaces. However, the

experiment data did not fit well to this curve (Figure 4.9). The experimental values of

cos( ) were lower than predicted by the Cassie-Baxter model.

Page 108: Nanostructured Selenium for Biomedical Applications: from

91

Figure 4.9.The Cassie-Baxter model curve (dashed line) and the experimental curve (solid line) of the cosine of water contact angles on uncoated titanium substrates and titanium substrates coated with various amount of selenium nanoclusters.

From the comparison of the cos(θ)-versus-selenium surface percentage curves

between the Cassie-Baxter model and the experiment data, it can be speculated that

during contact angle measurements on the selenium coated surfaces, there was formation

of air-pockets on the selenium coated samples that made the contact angles larger

(consequentially, smaller values of cos(θ)). It can be imagined that the higher selenium

coating density had higher RMS roughness values that, according to the Wenzel theory,

made the contact angles decrease. However, the surfaces with higher roughness could

also have more air-pockets that made the contact angles increase. Therefore, with the

contribution from both roughness and air-pockets, the contact angles of the three different

coating densities were not significantly different (Figure 4.10).

Page 109: Nanostructured Selenium for Biomedical Applications: from

92

Figure 4.10. Possible mechanisms for the plateau in contact angles on the substrates with different selenium coating densities. Increasing coating density of selenium nanoclusters increased roughness that resulted in both increases in the roughness factor (which is the actual surface area to geometric area ratio) and the increase in the formation of air-pockets. The former made contact angles decrease while the latter made contact angles increase. The opposite contribution of the roughness factor and air-pockets is likely the mechanism for similar contact angles for the three selenium coating densities on titanium in here.

To quantify the contribution of each factor to contact angles, in this part of the

chapter, a modified Cassie-Baxter model will be introduced that will take into account the

contribution of surface roughness, selenium chemistry and the contribution of air-

pockets.

4.5. Modification of the Cassie-Baxter model

It can be seen from the discussion in the Cassie-Baxter model that air-pockets

formed during contact angle measurements will increase the contact angle compared to

the case of no air-pockets formed. Therefore, from the comparison of the cos(θ)-versus-

selenium surface percentage curves between the Cassie-Baxter model and experimental

data, it can be speculated that there were air-pockets which increased contact angles

(consequently, decreased values of cos(θ)). Also, because the roughness ratio, r, of the

selenium coated substrates were smaller than 1.1, it is reasonable to adapt the Wenzel

model to the Cassie-Baxter model as follows:

Page 110: Nanostructured Selenium for Biomedical Applications: from

93

)cos()]cos()cos([)cos( 332211 fffr (Equation 4.15)

where r is the surface roughness ratio as defined by Equation 4.2; f1, f2 and f3 are the

surface area fractions of titanium, selenium and air, respectively; and θ1 ,θ2 , and θ3 are

contact angles on titanium, selenium and air, respectively (f1 + f2 + f3 =1 and θ3 =180o).

In this modified model the liquid drop contacts with all three materials: titanium,

selenium and air (as seen in Figure 4.10).

Figure 4.11. Schematic representation of the modified Cassie-Baxter model on the wetting process. A drop of liquid makes contact to all three materials: titanium, selenium nanoclusters, and air when placed on a selenium-titanium composite surface.

Clearly, f2 now is not necessarily the same as the surface fraction, ρ, of the

selenium nanoclusters as determined from SEM pictures. In the most general sense, all

three parameter (f1, f2 and f3) are dependent on ρ, r and R. To simplify this problem, it is

assumed that f2 and ρ are equal and f3 is a function of ρ, R and r. To further simplify the

Page 111: Nanostructured Selenium for Biomedical Applications: from

94

model, since the values of r do not vary significantly among the substrates (see Table

4.1), it was assumed that f2 had no dependence on r and had the following form:

RCRBARff )1(),(3(Equation 4.16)

where A, B and C are fitting parameters. Because ρ, r and R are dimensionless by their

definitions (see Table 4.1), A, B and C are also dimensionless. Physically, A and B

represents the contribution to the air-pocket formation from (i) the surface coverage of

selenium on titanium and (ii) the increased RMS roughness of the composite surface,

respectively; C represents the inter-play of both (i) and (ii) to air-pocket formation

because they are not independent (specifically, increasing ρ increases R). The factor (R-1)

comes from the requirement that f3 is 0 when ρ=0 and R=1.

Now f1 can be calculated by:

))1((11 321 CRRBAfff (Equation 4.17)

Equation 4.13 can now be re-written as:

)cos(1

)cos())]cos()(cos()[cos()1(

1

121

r

rCRRBA

(Equation 4.18)

Solving a set of 3 linear equations with 3 unknowns one can determine that

A = 0.089, B = -0.184 and C = -0.030.

Recall that:

])cos()cos([)cos( 32211 fffr (Equation 4.19)

Page 112: Nanostructured Selenium for Biomedical Applications: from

95

and

RCRBARff )1(),(3(Equation 4.20)

The positive value of A and negative value of B indicate the different contribution

to the air-pocket formation and, eventually, to the value of cos(θ). This different (in fact,

counteractive) contribution explains the plateau of the values of cos(θ) for the three

coating densities in this study. However, the negative value of B implies that increasing

surface roughness led to a decrease in air-pocket formation which is non-intuitive

because one expects there would be more air-pockets trapped between the gaps created

by the rough features on the selenium coated surface. This analysis can be understood as

follows. The underlying assumption of the Wenzel model is that the liquid-solid interface

follows the solid surface. In the meantime, the underlying assumption in the Cassie-

Baxter model for the case of air-pocket formation is that the liquid-solid interface does

not follow the solid surface. Therefore, the formation of air-pockets can be considered as

the indication of the degree of “non-Wenzel properties” in the contact angle

measurement. In other words, f3 also contains information about the “Wenzel properties”.

The fact that B is negative is in an agreement with this statement; increasing surface

roughness (i.e., increasing R) led to decreased f3, therefore, increased value of cos(θ) or

decreased contact angle.

Similarly, negative value of C indicates another contribution of Wenzel model in

the modified model proposed in this thesis.

Page 113: Nanostructured Selenium for Biomedical Applications: from

96

4.6. Summary and conclusions

In this chapter, a modified Cassie-Baxter model for contact angles was proposed

to fit the experimental data. Air-pockets were hypothesized to form when measuring

contact angles of water on the selenium-coated substrates. The modified model was used

to quantify the contribution of surface chemistry, surface area and surface roughness to

the formation of air-pockets. It was shown that the formation of air-pockets was

attributed mostly to the surface roughness brought about by the selenium coatings.

However, it should be noted that air-pockets can also occur on the uncoated titanium

substrate when measuring water contact angles. Experimental methods should be

developed to visualize and quantify air-pockets formed during the contact angle

measurements.

Page 114: Nanostructured Selenium for Biomedical Applications: from

97

CHAPTER 5. SELENIUM COATINGS ON POLYMERS FOR ANTI-INFECTION BIOMATERIAL APPLICATIONS

In previous chapters, it has already been shown that selenium can be readily

coated on a substrate surface to create various densities of nanoclusters. In addition to the

anti-cancer properties of selenium reported in Chapter 2 and in the literature, selenium

has also been shown to have anti-bacterial properties. Therefore, in this chapter, the use

of selenium nanocluster coatings for anti-bacterial applications was investigated. This

chapter is focused on bacterial infection on polymeric catheters as catheter-related

infection is the leading cause for extended hospital stays and increased cost of treatment.

5.1. Introduction

As discussed in Chapter 1, implant infection is a serious and complicated

problem. A common method to treat implant infection is though systemic antibiotic

therapy. However, this method has not shown satisfactory results to date. For example,

systemic antibiotic (vancomycin and gentamicin) administration alone without catheter

removal is only 22 to 37% effective in treating blood infection associated with using

long-term (more than 2 weeks) central venous catheters [4]. In addition, this mode of

antibiotic administration has the disadvantage of causing side effects and creating a pool

of antibiotic-resistant bacteria [10]. Another treatment method is the local delivery of

antibiotics. In this method, antibiotic molecules are incorporated (for example, by

adsorption or impregnation) onto the surface or into coating of the implants so that they

Page 115: Nanostructured Selenium for Biomedical Applications: from

98

will slowly release locally over an extended period of time after implantation. Because of

the use of antibiotics, this method also has significant draw backs including

ineffectiveness against antibiotic-resistant bacteria (such as methicillin-resistant

staphylococcus aureus or MRSA, important bacterium that infects all implants). An

emerging approach to prevent implant infection is engineering raw implant surfaces to

resist bacterial attachment and colonization.

Since the implant surface itself is an important source of infection and bacterial

adhesion to the implant surface, is important in the pathogenesis of infection, the most

promising and straight forward strategy towards decreasing infection is to fabricate

implant materials that resist bacteria attachment [103]. However, currently used materials

used as implant have been shown to be ineffective to resist bacterial infection [104-107].

For example, silver and silver compounds (such as chlorohexidine-silver sulfadiazine),

one of the most widely studied materials for anti-bacterial coatings on catheters or for

impregnating catheters, have yielded mixed results towards preventing infection. Some

studies as early as the 1950s and later in the 1970s and 1990s, showed that silver coatings

could reduce the incidence of catheter-associated infection [108-111]. In contrast, a large

number of recent studies have demonstrated that catheters coated or impregnated with

silver or silver-compounds either did not reduce the incidence of bacterial colonization

and catheter-associated infection [104-107] or lacked efficacy [7, 112]. Lastly, coating

catheters with silver is expensive; endotracheal tubes coated with silver can cost as much

as $100 per tube.

As discussed in Chapter 1, selenium has been shown to have anti-bacterial

properties. Therefore, the specific objective of this part of the study was to test the anti-

Page 116: Nanostructured Selenium for Biomedical Applications: from

99

bacterial properties of selenium nanocluster coatings on polymers that were created using

the novel method described in Chapter 2. Due to the popularity of catheter usage and

common catheter-related infections, this part of study focused on using selenium for anti-

bacterial catheter applications. Selenium nanoclusters were coated on three materials

commonly used for catheters: polyvinyl chloride (PVC), polyurethane (PU) and silicone,

using the same fabrication technique described in Chapter 2 and bacteria functions on the

selenium coated materials were determined.

The rest of this chapter is organized into sections on materials and methods,

results and conclusions. The materials and methods section describes the methods used

for creating, characterizing and testing the polymers coated with selenium. In the results

section, the highly anti-bacterial property of selenium nanocluster coatings was

demonstrated as direct evidence of the potential use of the coatings for anti-infection

catheters.

5.2. Materials and methods

A detailed fabrication and characterization method for coating PVC, PU and

silicone with selenium nanoclusters is introduced here. Bacterial culture assays to

investigate functions of bacteria on the polymer selenium coated materials is also

described.

5.2.1. Materials

Polymeric substrates (PVC, PU and silicone, in the form of discs 6mm in

diameter and approximately 2mm in height) were cut from PU (U.S. Plastic, catalog

number 48222), commercial PVC endotracheal tubes (Hudson RCI, catalog number 5-

Page 117: Nanostructured Selenium for Biomedical Applications: from

100

10412), and commercial silicone tracheostomy tubes (Bivona, catalog number 60PFS30).

The polymeric substrates were then immersed in ethanol for 10 min, sonicated in ethanol

for 10 min for sterilization and air-dried in a sterile hood overnight. Cleaned and

sterilized polymeric substrates (PVC, PU, and silicone) were used as base substrates for

the colloidal decoration with selenium nanoclusters. For this, 10ml of deionized water

and 10ml of 100mM glutathione (GSH, reduced form, TCI America) were added to a

beaker followed by 10ml of 25mM sodium selenite (Na2SeO3, Alfa Aesar). After gentle

mixing, 2ml of 1N NaOH was added to the beaker to bring the pH of the mixture into the

alkaline regime. The beaker was left untouched for 10 minutes. The substrates were

removed and rinsed for 24 hrs in deionized water. The uncoated and coated polymeric

substrates (PVC, PU, and silicone) were immersed in ethanol for 30 min for sterilization

and air-dried in a sterile hood overnight. Bard’s (Covington, GA) commercially available

Agento I.C. Silver-Coated (Bard Medical, catalog number 500375) endotracheal tubes

were also used as a comparison.

5.2.2. Material characterization

Polymeric substrates (PVC, PU, and silicone) were sputter-coated with a thin

layer (approximately 90 Ao) of palladium and observed under SEM at 5 kV. For chemical

characterization, energy dispersed X-ray spectroscopy (EDS) with an acceleration voltage

of 20kV was employed. EDS was used instead of XPS because it is more convenient and

can be done at the same time as SEM surface characterization. For EDS, a beam of

electrons is focused onto the sample’s surface. The energy of the electron beam excites

electrons in the inner shells of atoms and creates holes in the inner shells. Electrons from

the outer shells (higher energy levels) follow a transition to fill the holes and the energy is

Page 118: Nanostructured Selenium for Biomedical Applications: from

101

released in the form of X-rays. The energy of the X-rays is characteristic of the elemental

composition being analyzed.

5.2.3. In vitro assays for the polymeric substrate-bacterial experiments

5.2.3.1. SEM visualization of bacteria on the uncoated and coated substrates

S. aureus (ATCC, 25923), an important bacteria infecting all catheters was

prepared from a frozen vial received from the vendor. A sterile loop was used to

withdraw some of the bacteria from the frozen vial, streaked onto tryptic soy agar plates

(30g Tryptone [MP Biomedicals] and 15g agar [Sigma] per litter of distilled water) and

incubated for 12 hrs. Another sterile loop was used to select one colony from the agar

plates to place into 3ml trypic soy broth (TSB, 30g/L) and was incubated for 16 hrs.

Uncoated and selenium-coated substrates were placed into the wells of 24 well plates.

The bacteria were then seeded in TSB at approximately 20,000,000 bacteria per well and

inoculated for 8 hrs in a standard bacteria culture incubator (37ºC, 95% humidified air,

5% CO2 environment, non-shaking). At the end of the indicated time, the substrates were

removed from culture and treated according to standard protocols for SEM visualization.

Briefly, substrates were rinsed in 0.1M sodium cacodylate buffer (SCB) solution, fixed in

2.5% glutaraldehyde in 0.1M SCB for 30 min at room temperature (RT), rinsed in SCB,

dehydrated in a series of ethanol solutions (30%, 50%, 70%, 90%, 95%, 100% ethanol,

10 min each), left in 100% ethanol for 15 min, critically-point dried, sputter-coated with

gold and finally visualized under SEM (LEO 1530VP) at a voltage from 3 to 10 kV.

Page 119: Nanostructured Selenium for Biomedical Applications: from

102

5.2.3.2. Crystal violet assays for bacteria quantification

To quantitatively compare the adherent bacteria on the different substrates, crystal

violet assays were used. After incubation of bacteria for 8 hrs, substrates were rinsed in

1% NaCl solution to remove all unattached bacteria. The adherent bacteria on the

surfaces were then stained with crystal violet (CV) which penetrated cell walls and cell

membranes. For this, substrates were immersed in 300µL of 0.01% CV (in water) for 15

minutes. After the indicated time, unbound CV was washed using 1% NaCl. The

substrates were then left air-dried overnight to make sure all the liquid was evaporated.

The bound CV (to cells) was solubilized in 100% ethanol. For this, 300µL of 100%

ethanol was added to each substrate and incubated at room temperature for 30 minutes.

Optical densities (ODs) of the soluble CV solutions (which correlate to total cell number)

were measured using a microplate reader (SpectraMax, Molecular Devices) and

compared among the substrates.

5.2.3.3. Intracellular thiol assays

Uncoated and selenium-coated PVC substrates were immersed in 1ml TSB for 8

hrs in an incubator (37o, 5% CO2). After the indicated time, the supernatant was collected

and used for culturing S. aureus in 15ml centrifuge tubes. For this, S. aureus (prepared as

described in 5.2.3.1) were cultured in the supernatant at the concentration of 10,000,000

bateria/mL for 8 hrs. After the indicated time, bacteria were collected and analyzed for

intracellular thiol content using a glutathione assay kit (CS 1020, Sigma) following

manufacturer’s instructions. The assay uses a thiol probe (monochlorobimane) which

passes through cell membranes to detect the level of glutathione, the major free thiol in

Page 120: Nanostructured Selenium for Biomedical Applications: from

103

most living cells. If the probe binds to reduced glutathione in cells, it forms a fluorescent

product. The unbound probe does not show fluorescence.

5.2.4. Statistical analysis

Experiments were conducted in triplicate and repeated three times. Data were

collected and the significant differences were assessed with the probability associated

with one-tailed Student’s t tests.

5.3. Results and discussion

In this section, the results from SEM visualization of the selenium coated

polymers as well as EDS analysis are presented. Bacteria (S. aureus) response on the

substrates is also discussed.

5.3.1. SEM and EDS

Importantly, this study elucidated a process to coat all three polymeric substrates

with selenium nanoclusters. At the same concentration of selenium in the synthesis

solution, coating densities on 3 different types of polymeric substrates (silicone, PU and

PVC) were different as shown by SEM images (Figure 5.1). This difference is likely due

to the difference in nucleation sites (e.g. defects) present on the surfaces of the materials

to be coated. Compared to the selenium coatings on titanium and stainless steel in

Chapter 2, the morphology of selenium nanoclusters on the polymeric substrates

appeared similar. The size of the selenium nanoclusters on polymeric substrates was

approximately 150 nm in diameter which was larger than on titanium and stainless steel

substrates (which was ~80nm in diameter). The surface coverage of selenium

(determined from SEM images using an image processing program, imageJ (National

Page 121: Nanostructured Selenium for Biomedical Applications: from

104

Institute of Health)) was approximately 10% on PVC substrates, 4% on PU substrates and

7% on silicone substrates.

Figure 5.1. SEM images of uncoated (left panel) and selenium (Se)-coated substrates (right panel). Arrows indicate selenium nanoclusters. The selenium nanoclusters had sizes ranging from approximately 80nm to 200nm and were uniformly coated on the substrates.

EDS spectra were used to confirm the chemistry of selenium nanoclusters.

Selenium peaks observed in the area coated with selenium nanoclusters confirmed that

the clusters were selenium (Figures 5.2, 5.3, and 5.4).

Page 122: Nanostructured Selenium for Biomedical Applications: from

105

Figure 5.2. EDS spectra (right panel) of an uncoated region (top) and a coated region (bottom) on a selenium-coated silicone substrate. Selenium peaks were detected in the coated region demonstrating that the clusters were selenium. Peaks of gold were from the sputter-coated layer.

Page 123: Nanostructured Selenium for Biomedical Applications: from

106

Figure 5.3. EDS spectra (right panel) of an uncoated region (top) and a coated region (bottom) on a selenium-coated PU substrate. Selenium peaks were detected in the coated region demonstrating that the clusters were selenium. Peaks of gold were from the sputter-coated layer.

Page 124: Nanostructured Selenium for Biomedical Applications: from

107

Figure 5.4. EDS spectra (right panel) of an uncoated region on selenium-coated PVC (top) and a coated region on selenium-coated PVC (bottom). Selenium peaks were detected in the selenium-coated PVC demonstrating that the clusters were selenium. Peaks of gold were from the sputter-coated layer.

5.3.2. S. aureus response on uncoated and selenium-coated polymeric substrates

The uncoated substrates were colonized with S. aureus after 8 hrs of inoculation

while selenium-coated substrates had very little S. aureus colonization (Figures 5.5, 5.6

and 5.7).

Page 125: Nanostructured Selenium for Biomedical Applications: from

108

Figure 5.5. Representative SEM images showing reduced S. aureus colonization on selenium-coated PU compared to uncoated PU after 8 hrs of inoculation. Arrows indicate cells.

Figure 5.6. Representative SEM images showing reduced S. aureus colonization on selenium-coated PVC compared to uncoated PVC. Arrows show bacteria.

Figure 5.7. Representative SEM images showing reduced S. aureus colonization

on selenium-coated compared to uncoated silicone. Arrows indicate cells.

Page 126: Nanostructured Selenium for Biomedical Applications: from

109

Bacteria densities determined by crystal violet assays showed that the bacterial

colonization on the selenium coated substrates significantly decreased compared to

controls (to 80%, 20% and 40% of the control, i.e., uncoated substrates, for PU, PVC and

silicone, respectively) (Figure 5.8).

Figure 5.8. Decreased S. aureus densities on polymeric substrates coated with selenium nanoclusters. Data = mean ± standard error of the mean, N=3; *p<0.05 compared to the uncoated substrate (compared in the same material group), ** p<0.05 compared to silver-coated PVC.

Importantly, compared to the commercial Bard’s silver-coated PVC endotracheal

tube, selenium nanocluster-coated PVC tubes were shown to be approximately twice

more effective in preventing bacteria attachment after 8 hrs (Figures 5.8 and 5.9).

Page 127: Nanostructured Selenium for Biomedical Applications: from

110

Figure 5.9. Representative SEM images of S. aureus colonization on silver-coated and Se-coated PVC. Bacterial colonization on Se-coated PVC was reduced compared to silver-coated PVC after 8 hrs of incubation.

For the first time, selenium nanocluster coatings were demonstrated to inhibit S.

aureus colonization. Especially, selenium nanocluster coatings were approximately twice

more effective than silver coatings in preventing S. aureus colonization after 8 hrs.

Research has shown that elemental selenium could catalyze the oxidation of intracellular

thiols resulting in cellular thiol depletion that causes cell death [40]. It has also been

demonstrated (in Chapter 2) that selenium was released from selenium coatings on

titanium and that the released selenium induced cellular thiol depletion in normal

osteoblasts and cancerous osteoblasts. Therefore, this part of the study also tested if the

selenium nanocluster coatings on polymers inhibited S. aureus growth via the thiol-

depletion mechanism. The results will be presented in the following section.

5.3.3. Thiol assays

To test the hypothesis that selenium nanoclusters caused intracellular thiol

depletion in S. aureus, the bacteria were cultured in the supernatant from either uncoated

PVC or selenium coated PVC substrates. After 8 hrs of incubation, intracellular thiol

levels in bacteria were determined using the Glutathione assay kit (Sigma). S. aureus

Page 128: Nanostructured Selenium for Biomedical Applications: from

111

cultured with fresh bacterial culture media (i.e., tryptic soy broth) were used as controls.

As can be seen from Figure 5.10, the thiol level in the bacteria cultured in the supernatant

from the selenium-coated PVC was significantly lower than that of bacteria cultured in

the supernatant from uncoated PVC.

Figure 5.10. Decreased intracellular thiol level in S. aureus cultured in the supernatant from selenium-coated PVC compared to control. Data = mean ± standard error of the mean; N=3; * p<0.05 compared to the control.

Therefore, it is likely that selenium in the selenium nanocluster coatings released

into the bacteria culture media and this released selenium induced depletion of

intracellular thiol which caused bacteria death.

5.4. Discussion

Attaching selenium to a material to transform an otherwise bacterial colonizing

surface into a material that is anti-bacterial has also been explored by other research

groups. For example, a group at Texas Tech University (Lubbock, TX) developed a

technology that covalent bonds organo-selenium molecules (whose name was unrevealed

for proprietary reasons) to silicone hydrogel contact lenses to transform the lenses into

Page 129: Nanostructured Selenium for Biomedical Applications: from

112

anti-bacterial materials [113]. Specifically, the contact lenses were coated with the

organo-selenium molecules and tested against colonization of Pseudomonas aeruginosa

and S. aureus. The results showed little or no bacterial biofilm formation on the coated

lenses after 4 days of incubating the lenses in bacteria solutions. In contrast, uncoated

lenses showed extensive bacterial biofilm formation. The safety of the organo-selenium

coated lenses was also tested by placing them into the eyes of New Zealand albino rabbits

for up to 2 months. There was no sign of toxicity for the eyes wearing the coated lenses

as well as the eyes wearing uncoated lenses. However, because the method used by this

research group relies on the covalent bonding between selenium compounds and the

material, their method can only be used to impart selenium onto organic materials (such

as polymers). The method used in this thesis, relying on the nucleation of selenium onto

surfaces, creates selenium coatings not only on polymers but also on metals (as shown in

Chapter 2). Therefore, the process employed in this thesis has the ability to create

selenium coatings on a broad spectrum of biomaterials.

5.5. Conclusions

Selenium nanoclusters were coated on polymeric substrates using the same

synthesis methods utilized to coat selenium nanoclusters on metallic substrates in Chapter

2. Similar morphology and chemistry of the coatings (compared to the coating on

titanium and stainless steel substrates) was observed. Compared to the uncoated

substrates, substrates coated with selenium nanoclusters had significantly lower

attachment of S. aureus after 8 hrs of incubation. Importantly, compared to commercially

available silver coated endotracheal tubes, the selenium coated tubes were approximately

twice as more effective at reducing S. aureus attachment. The decreased bacterial

Page 130: Nanostructured Selenium for Biomedical Applications: from

113

attachment on selenium coated substrates was likely due to the released selenium that

depleted thiol levels in the cells causing cell death. This result, for the first time,

demonstrates the potential of using selenium nanoclusters as a coating material for

numerous anti-bacterial applications.

Page 131: Nanostructured Selenium for Biomedical Applications: from

114

CHAPTER 6. SELENIUM NANOPARTICLES FOR ANTI-BACTERIAL APPLICATIONS

6.1. Introduction

In the previous chapters, it has been demonstrated that selenium nanocluster

coatings on traditional implant materials promoted healthy bone growth while inhibiting

bone cancer growth as well as inhibiting bacterial growth in vitro. However, how

selenium nanoparticles act as opposed to adherent nanoclusters on implants on bacteria

(such as S. aureus, S. epidermidis) remains largely unknown.

Nanotechnology has enable researchers to synthesize nano-size particles (that is,

particles that have sizes less than 100nm in at least 1 dimension) for a wide range of

applications. Nanoparticles have increased surface areas, therefore, have potentially

increased interactions with biological targets (such as bacteria) compared to

conventional, micro-size particles. As a result, nano-antibacterial particles will likely

exert a stronger effect on bacteria than their microparticle-counterparts.

Therefore, the objective of this part of the study was to examine the functions of

S. aureus in the presence of selenium nanoparticles. In doing so, this study hopes to

reveal a new type of anti-bacterial nanoparticles for decreasing S. aureus adhesion and

colonization.

This chapter is organized into a materials and methods section where the methods

to synthesize and characterize selenium nanoparticles are described, a results section and

a summary and conclusion section.

Page 132: Nanostructured Selenium for Biomedical Applications: from

115

6.2. Materials and methods

6.2.1. Material synthesis

Selenium nanoparticles were synthesized by reduction of sodium selenite (Alfa

Aesar) by glutathione (reduced form, GSH, TCI America) stabilized by bovine serum

albumin (BSA, Sigma Aldrich). Specifically, 3ml of 25mM Na2SeO3, 3ml of 100mM

GSH and 0.15g BSA were added to 9ml double distilled water in a sterile cabinet. BSA

was used as a surfactant to keep the selenium nanoparticles well dispersed. All solutions

were made in a sterile environment by using a sterile cabinet and double distilled water.

After the reactant solution was mixed, 1M NaOH was added to bring the pH of the

solution to the alkaline regime. Selenium nanparticles were formed immediately

following the addition of NaOH as visualized by a color change of the reactant solution

from clear white to clear red. Selenium nanoparticles were then collected by centrifuging

at 13000 rounds per minute and were re-suspended in deionized water five times before

using in bacteria experiments.

6.2.2. Material characterization

The size and morphology of selenium nanoparticles were investigated using a

transmission electron microscope (TEM). For this, the dispersions of nanoparticles in

deionized water were allowed to slowly dry on formvar-coated copper grids. All imaging

was carried out using a Philips 410 TEM (New York, NY, USA) at 80kV.

The size distribution of selenium nanoparticles was investigated by dynamic light

scattering (DLS) technique using a Zetasizer-Nano-S90 (Malvern Instrument).

Chemistry of selenium nanoparticles was characterized using X-ray photoelectron

Page 133: Nanostructured Selenium for Biomedical Applications: from

116

spectroscopy (XPS, Perkin-Elmer PHI. 5500 Multi-Technique System). For this, a drop

of selenium nanoparticle solution was placed on a glass coverslip, air-dried and analyzed

under XPS.

6.2.3. Bacteria assays

6.2.3.1. Bacteria preparation

Bacteria solutions were prepared as described in section 5.2.3.1 in Chapter 5.

Briefly, a bacterial cell line of biofilm-producing S. aureus was obtained in freeze-dried

form from the American Type Culture Collection (catalog number 25923). The cells were

propagated in 30mg/mL tryptic soy broth (TSB, MP Biomedicals). A sterile loop was

used to withdraw some of the bacteria from the frozen vial, streaked onto tryptic soy agar

plates (30g Tryptone [MP Biomedicals] and 15g agar [Sigma] per litter of distilled water)

and incubated for 12 hrs. Another sterile loop was used to select one colony from the agar

plates to place into 3ml trypic soy broth (TSB, 30g/L) and was incubated for 16 hrs.

Bacteria concentration was assessed by measuring optical density of bacterial solution at

562nm and using a standard curve correlating optical densities and bacterial

concentrations. The bacterial solution was prepared at a concentration of 50,000 cells/mL

for all experiments.

6.2.3.2. Effects of selenium nanoparticles on planktonic S. aureus

Three concentrations of selenium (Se) nanoparticles were tested against S. aureus

growth: 3.1µg Se/mL, 6.2 µg Se/mL and 12.5 µg Se/mL with TSB (0 µg Se/mL) as the

control. Selenium nanoparticles were mixed with bacterial solutions and cultured for 4,

12 and 24 hrs in an incubator (37°C, humidified, 5% CO2) shaking at 250 rounds per

Page 134: Nanostructured Selenium for Biomedical Applications: from

117

minute. Blank solutions were prepared by adding selenium nanoparticles into TSB

without bacteria at the above concentrations of particles. Blank solutions of TSB without

bacteria and without selenium nanoparticles were used as the blank for the controls.

At the end of the prescribed time periods, optical densities (or the degree of

“cloudiness”) of the bacteria solutions (which is proportional to bacteria densities) were

measured. For this, 200µL of the bacteria solution, control, or blank were added to wells

of a 96 well plate and optical densities were measured at 562nm. Measured optical

densities of bacterial solutions were subtracted by that of the corresponding blanks to

remove the optical density resulting from the nanoparticles alone.

6.2.3.3. Intracellular thiol assays

S. aureus (at a density of ~50,000 cells/mL) were cultured with selenium

nanoparticles (at a concentration of 3.1 µg/mL in tryptic soy broth) for 4 hrs under

standard conditions (37oC, 5%CO2, 95% humidified air) shaking at 250 rounds per

minute. After the indicated time, cells were collected and analyzed for thiol content using

a glutathione assay kit (CS1020, Sigma Aldrich) following manufacturer’s instruction. S.

aureus cultured in normal cell culture media without selenium nanoparticles were used as

a control. The glutathione assay kit uses a thiol probe (monochlorobimane) which passes

through cell membranes to detect the level of glutathione, the major free thiol in most

living cells. Glutathione is involved in many biological processes such as removal of

hydroperoxides and detoxification of xenobiotics. Intracellular reduced glutathione level

is a sensitive indicator of the overall health of a cell. If the probe binds to reduced

glutathione in cells, it forms a fluorescent product whose fluorescence intensity can be

measured. The unbound probe does not show fluorescence.

Page 135: Nanostructured Selenium for Biomedical Applications: from

118

6.2.4. Superoxide anion assays

Some reports in the literature show that selenium (in the presence of thiols) can

generate superoxide anion which is a short-lived (half-life time ~ 0.05 seconds) free

radical that kills bacteria [35, 114, 115]. Therefore, in this chapter, superoxide anion

assays were conducted to test if selenium nanoparticles in this thesis generated

superoxide anion. The assays were conducted using a superoxide anion kit (LumiMax

Superoxide Anion Detection kit, Strategene, catalog number 204525). Reagents for

superoxide anion assays were prepared following manufacturer’s instruction with some

modifications. Briefly, three types of reagents were prepared: test samples which had

luminol, enhancer, glutathione (GSH) and selenium nanoparticles (referred to as nSe),

control 1 which did not have nSe and control 2 which did not have the enhancer solution

(Table 6.1).

Table 6.1. Superoxide anion assay reagent mixtures

Assay

medium (µL)

Luminol solution

(µL)

Enhancer (µL)

12.5 µg/mL nSe solution

(µL)

1µg/mL GSH (µL)

Test sample 89 5 5 5 5 Control 1

(without selenium) 94 5 5 0 5

Control 2 (without enhancer)

94 5 0 5 5

The mixture was mixed and immediately measured for photon emission using an

illuminometer (Lumat LB9510, Berthold Technologies). Due to the short life-time of

superoxide anions (~0.05 seconds), photon emission was read every 10 seconds for 30

Page 136: Nanostructured Selenium for Biomedical Applications: from

119

seconds immediately after mixing. Photon emission was then reported as the total relative

light unit emitted during these 30 seconds.

6.2.5. Statistical analysis

All experiments were conducted in triplicate and repeated three times. Data were

collected and the significant differences were assessed with the probability associated

with one- tailed Student’s t tests.

6.3. Results

6.3.1. TEM, DLS and XPS results on size distribution and oxidation state of selenium

nanoparticles

TEM images of selenium nanoparticles show that the particles were spherical and

about 40-100 nm in diameter (Figure 6.1). Further investigation of size distribution of the

selenium nanoparticles by DLS revealed that most of the particles had diameters of

100nm (Figure 6.2). The sizes observed by DLS were larger than those determined by

TEM images because water molecules bound to the surface of the nanoparticles creating

a “hairy layer” that made the particles appear larger [116, 117].

Page 137: Nanostructured Selenium for Biomedical Applications: from

120

Figure 6.1. TEM image of selenium nanoparticles stabilized in BSA and dispersed in water. Selenium nanoparticles had average sizes of approximately 100 nm.

Page 138: Nanostructured Selenium for Biomedical Applications: from

121

Figure 6.2. Size-distribution profile of selenium nanoparticles in solution as measured by the dynamic light scattering technique. Particle sizes centered around 100 nm.

The size-distribution profile demonstrated that the synthesis method yielded

selenium nanoparticles of a narrow size range and stable in water. Nanometer scale sizes

of the synthesized nanoparticles rendered the nanoparticles large surface areas that are

important towards increasing the particles’ interaction with bacteria.

The peak at 55.2 eV (Figure 6.3) demonstrated that the selenium nanoparticles

had a zero oxidation state (i.e., they were elemental selenium).

Figure 6.3. XPS profile of selenium nanoparticles. Peak at 55.2 eV confirmed a zero oxidation state of selenium.

Page 139: Nanostructured Selenium for Biomedical Applications: from

122

6.3.2. Effects of selenium nanoparticles on S. aureus in solution

When the selenium nanoparticles were mixed with the bacterial solution, the

growth of bacteria was significantly inhibited after 4 hrs (compared to the controls -

0µg/mL). The inhibitory effects continued after 24 hrs (Figure 6.4). A dose-dependent

inhibition of bacterial growth was observed at 4 hrs and 12 hrs.

Figure 6.4. Inhibited growth of S. aureus in the presence of selenium nanoparticles at all three selenium nanoparticles concentrations: 12.5 µg/mL, 6.2 µg/mL and 3.1 µg/mL at 4 hrs, 12 hrs and 24 hrs. Data = mean ± standard error of the mean, N=3. *p<0.05 compared to bacteria treated with 6.2 µg Se/mL, 3.1 µg Se/mL and control (0 µg Se/mL) (compared at same time period); ** p<0.05 compared to bacteria treated with 3.1 µg Se/mL and control (0 µg Se/mL) (compared at same time period); ***p<0.05 compared to control (compared at same time period); #p<0.05 compared to control (compared at same time period). Dose-dependent inhibition was observed at 4hrs and 12hrs.

6.3.3. Intracellular thiol assays

As it was demonstrated in Chapter 5 that selenium can be released from selenium

coatings on polymers into the bacteria culture media causing depletion of intracellular

Page 140: Nanostructured Selenium for Biomedical Applications: from

123

thiol in S. aureus. To test if selenium nanoparticles inhibited S. aureus in this study via a

similar mechanism, thiol assays were conducted to determine thiol content in S. aureus

cultured with selenium nanoparticles. S. aureus cultured in normal bacteria culture media

(tryptic soy broth (TBS)) were used as the control. The results (Figure 6.5) showed that

the intracellular thiol levels in S. aureus cultured with selenium nanoparticles was

significantly decreased (to approximately 90% of the control) compared to that of the

control. This means that S. aureus treated with selenium nanoparticles had depleted thiol

levels which is known to cause cell death.

Figure 6.5. Decreased intracellular thiol levels in S. aureus cultured with selenium nanoparticles compared to the control (i.e., S. aureus cultured in TSB without selenium nanoparticles). Bacteria were cultured at a density of 50,000 cells/mL for 4 hrs in either TSB or TSB added with selenium nanoparticle (at a concentration of 3.1µg/mL). Data = mean ± standard error of the mean; N=3; * p< 0.05.

Page 141: Nanostructured Selenium for Biomedical Applications: from

124

Similar effects (i.e., selenium induced intracellular thiol depletion) were observed

in selenium nanocluster coatings on metals (Chapter 2) and polymers (Chapter 5). These

results indicated that intracellular thiol depletion is likely the mechanisms of anti-cancer

and anti-bacterial properties of nanometer structured selenium in this thesis.

6.3.4. Superoxide assays

Some reports in the literature demonstrate that selenium, in the presence of thiols

(such as glutathione), can generate superoxide anion ( 2O ) that kills bacteria [114, 115].

To test if the selenium nanoparticles in this study generated superoxide anions which

killed S. aureus, superoxide anion assays were conducted. The assay used luminol which

reacts with superoxide anions to generate photons of light which can be measured using

an illuminometer. The assays also used an enhancer solution which amplifies the

luminescence signal to increase the sensitivity of the measurement. For this, selenium

nanoparticles (referred to as nSe) were mixed with reduced glutathione (GSH, Sigma),

luminol and enhancer solutions and the photon emission from this mixture was measured.

Two controls (luminol + enhancer and luminol + GSH+ nSe) were used for comparisons.

Results showed that there was no significant difference among the tested sample (i.e.,

nSe) and controls (i.e., luminol + enhancer or luminol + GSH + nSe) (Figure 6.6).

Therefore, it was concluded that elemental selenium nanoparticles in this study did not

generate superoxide anions.

Page 142: Nanostructured Selenium for Biomedical Applications: from

125

Figure 6.6. Superoxide anion assays for selenium particles (nSe). No significant difference was found among the test tubes. Selenium nanoparticles were mixed with GSH and luminol was used as the probes for superoxide anions. Superoxide anions oxidize luminol in a reaction that produces photons. An enhancer was used to amplify the chemiluminescence. Data = mean ± standard error of the mean, N=3.

6.4. Summary and conclusions

In this chapter, it has been demonstrated that selenium nanoparticles were

synthesized with a narrow size distribution centered around the value of 100nm. More

importantly, selenium nanoparticles were shown to significantly inhibit the growth of S.

aureus in solution. It was further shown that selenium nanoparticles induced thiol

depletion that caused death of S. aureus. Future studies should evaluate the anti-bacterial

properties of selenium nanoparticles against other bacteria (such as Pseudomonas

aeruginosa and S. epidermidis) to fully explore the potential of using selenium

nanoparticles for anti-infection applications. Toxicity studies should also be conducted to

determine safety concentrations of selenium nanoparticles to use in humans. Specifically,

in vitro studies should be implemented to evaluate response of mammalian cells (for

example fibroblasts) to selenium nanoparticles at different concentrations. Further studies

Page 143: Nanostructured Selenium for Biomedical Applications: from

126

should then be completed to evaluate the anti-bacterial properties as well as safety

concentration of selenium nanoparticles in animals and humans.

Page 144: Nanostructured Selenium for Biomedical Applications: from

127

CHAPTER 7. CONCLUSIONS

This thesis presented a comprehensive study of the potential of nanometer

structured selenium for various biomedical applications. Specifically, in this study,

selenium nanoclusters were coated on orthopedic implant metals (such as titanium and

stainless steel). The coatings were fully characterized using techniques such as electron

scanning microscopy, X-ray photoelectron spectroscopy, atomic force microscopy, and

contact angle measurements. Most importantly, compared to uncoated metals, the

coatings on orthopedic metals promoted healthy bone cell functions. This result by itself

suggests much promise for selenium coated metals for any orthopedic application.

Moreover, cancerous bone cell functions were inhibited on selenium nanocluster coatings

compared to uncoated metals. This result suggests that selenium coated implants should

be tested for numerous cancer applications, in particular, bone cancer applications. This

study also examined the possible mechanisms of the promoted healthy bone-cell

properties of selenium nanocluster coatings on metals. It was shown that the increased

nanometer roughness of the selenium coated surfaces promoted the adsorption of

fibronectin (which is one of the most important cell adhesive proteins that regulate

interactions between osteoblasts and an implant surface). The experimental results of

protein adsorption were complimented by results from computer simulation. The anti-

cancer properties of selenium nanocluster coatings were attributed in this thesis to

Page 145: Nanostructured Selenium for Biomedical Applications: from

128

selenium released from the coatings into cell culture media to induce cellular thiol

depletion which caused cell death in cancerous osteosarcoma cells.

To further demonstrate the versatility of nanometer scale structured selenium in

biomedical applications, the selenium nanoclusters were also coated on common catheter

polymers (such as polyvinyl chloride, polyurethane and silicone). The selenium coatings

were shown to be highly effective against the colonization of S. aureus, an important

bacterium that infects all implants, especially catheters. This result opens the possibility

of using nanostructured selenium for any anti-bacterial application. This thesis also

demonstrated that the selenium released from the coatings induced cellular thiol depletion

which caused cell death in S. aureus.

The study also investigated the potential of using selenium nanoparticles in

various anti-bacterial medical applications. Selenium nanoparticles were synthesized and

characterized by techniques such as transmission electron microscopy, X-ray

photoelectron microscopy and dynamic light scattering. Selenium nanoparticles were

demonstrated to be highly effective at inhibiting the growth of S. aureus. Intracellular

thiol levels of the bacteria treated with selenium nanoparticles also decreased compared

to those of the bacteria treated with normal cell culture media. This result opens the

possibility of using nanoparticulate selenium in numerous infection applications (such as

implantation, catheterization, wound repair, etc.)

In conclusion, this thesis demonstrated a number of promising directions for using

nanometer scale structured selenium in diverse biomedical applications (including bone-

promoting, cancer-inhibiting and bacteria-inhibiting applications). The anti-cancer

properties of nanometer structured selenium should be further tested on other cancerous

Page 146: Nanostructured Selenium for Biomedical Applications: from

129

cell types (such as breast cancer, lung cancer, etc.) to fully evaluate its applications in

cancer treatment. The anti-bacterial properties of nanometer structured selenium should

also be tested on other types of bacteria (such as P. aeruginosa, S. epidermidis, etc.) to

fully investigate its anti-infection properties and its anti-infection applications. In vivo

studies should be conducted to further investigate the potential of this biomaterial.

Page 147: Nanostructured Selenium for Biomedical Applications: from

130

References:

1. ACS. Cancer Statistics. 2008 [cited 2009 2/12]; Available from: http://www.cancer.org/docroot/STT/STT_0.asp

2. Miller D WT. Anticancer orthopedic implants. In: Nalwa H WTJ, editor. Cancer nanotechnology. California: American Scientific, 2007. p. 307-316.

3. Orthopedics today; 29:6 2009 [cited 2009 2/7/2010]; Available from: http://www.orthosupersite.com/view.asp?rID=44153

4. Krishnasami Z, Carlton D, Bimbo L, Taylor ME, Balkovetz DF, Barker J, et al. Management of hemodialysis catheter-related bacteremia with an adjunctive antibiotic lock solution. Kidney Int 2002 Mar;61(3):1136-1142.

5. Trerotola SO. Hemodialysis catheter placement and management. Radiology 2000 Jun;215(3):651-658.

6. Sanders J, Pithie A, Ganly P, Surgenor L, Wilson R, Merriman E, et al. A prospective double-blind randomized trial comparing intraluminal ethanol with heparinized saline for the prevention of catheter-associated bloodstream infection in immunosuppressed haematology patients. J Antimicrob Chemother 2008 Oct;62(4):809-815.

7. Johnson JR, Kuskowski MA, Wilt TJ. Systematic review: antimicrobial urinary catheters to prevent catheter-associated urinary tract infection in hospitalized patients. Ann Intern Med 2006 Jan 17;144(2):116-126.

8. Maki DG, Tambyah PA. Engineering out the risk for infection with urinary catheters. Emerg Infect Dis 2001 Mar-Apr;7(2):342-347.

9. Roe D, Karandikar B, Bonn-Savage N, Gibbins B, Roullet JB. Antimicrobial surface functionalization of plastic catheters by silver nanoparticles. J Antimicrob Chemother 2008 Apr;61(4):869-876.

10. Schierholz JM, Beuth J. Implant infections: a haven for opportunistic bacteria. Journal of Hospital Infection 2001;49(2):87-93.

11. Donlan RM. Biofilms and device-associated infections. Emerg Infect Dis 2001 Mar-Apr;7(2):277-281.

12. Davies D. Understanding biofilm resistance to antibacterial agents. Nat Rev Drug Discov 2003;2(2):114-122.

13. Chastre J, Fagon J-Y. Ventilator-associated Pneumonia. Am J Respir Crit Care Med 2002 April 1, 2002;165(7):867-903.

14. Sanderson PJ. Infection in orthopaedic implants. Journal of Hospital Infection 1991;18(Supplement 1):367-375.

15. Styblo M, Thomas DJ. Selenium modifies the metabolism and toxicity of arsenic in primary rat hepatocytes. Toxicol Appl Pharmacol 2001 Apr 1;172(1):52-61.

16. McDowell L. Minerals in human nutrition. Amsterdam:: Elsevier, 2003. 17. Combs GFJ, Combs S. The role of selenium in nutrition. Orlando: Academic press, 1986. 18. Clark LC, Cantor KP, Allaway WH. Selenium in forage crops and cancer mortality in

U.S. counties. Arch Environ Health 1991 Jan-Feb;46(1):37-42. 19. Combs GF, Jr., Gray WP. Chemopreventive agents: selenium. Pharmacol Ther 1998

Sep;79(3):179-192. 20. Navarro-Alarcon M, Lopez-Martinez MC. Essentiality of selenium in the human body:

relationship with different diseases. Sci Total Environ 2000 Apr 17;249(1-3):347-371.

Page 148: Nanostructured Selenium for Biomedical Applications: from

131

21. Clark LC, Combs GF, Jr., Turnbull BW, Slate EH, Chalker DK, Chow J, et al. Effects of selenium supplementation for cancer prevention in patients with carcinoma of the skin. A randomized controlled trial. Nutritional Prevention of Cancer Study Group. JAMA 1996 December 25, 1996;276(24):1957-1963.

22. Wei W-Q, Abnet CC, Qiao Y-L, Dawsey SM, Dong Z-W, Sun X-D, et al. Prospective study of serum selenium concentrations and esophageal and gastric cardia cancer, heart disease, stroke, and total death. Am J Clin Nutr 2004 January 1, 2004;79(1):80-85.

23. Clark LC, Dalkin B, Krongrad A, Combs GF, Jr., Turnbull BW, Slate EH, et al. Decreased incidence of prostate cancer with selenium supplementation: results of a double-blind cancer prevention trial. Br J Urol 1998 May;81(5):730-734.

24. Rayman MP. Selenium in cancer prevention: a review of the evidence and mechanism of action. Proceedings of the Nutrition Society 2005;64(04):527-542.

25. Lu J, Jiang C, Kaeck M, Ganther H, Ip C, Thompson H. Cellular and metabolic effects of triphenylselenonium chloride in a mammary cell culture model. Carcinogenesis 1995 Mar;16(3):513-517.

26. Ronai Z, Tillotson JK, Traganos F, Darzynkiewicz Z, Conaway CC, Upadhyaya P, et al. Effects of organic and inorganic selenium compounds on rat mammary tumor cells. Int J Cancer 1995 Nov 3;63(3):428-434.

27. Redman C, Scott JA, Baines AT, Basye JL, Clark LC, Calley C, et al. Inhibitory effect of selenomethionine on the growth of three selected human tumor cell lines. Cancer Lett 1998 Mar 13;125(1-2):103-110.

28. Kaeck M, Lu J, Strange R, Ip C, Ganther HE, Thompson HJ. Differential induction of growth arrest inducible genes by selenium compounds. Biochem Pharmacol 1997 Apr 4;53(7):921-926.

29. Lu J, Pei H, Ip C, Lisk DJ, Ganther H, Thompson HJ. Effect on an aqueous extract of selenium-enriched garlic on in vitro markers and in vivo efficacy in cancer prevention. Carcinogenesis 1996 Sep;17(9):1903-1907.

30. Apostolou S, Klein JO, Mitsuuchi Y, Shetler JN, Poulikakos PI, Jhanwar SC, et al. Growth inhibition and induction of apoptosis in mesothelioma cells by selenium and dependence on selenoprotein SEP15 genotype. Oncogene 2004 Jun 24;23(29):5032-5040.

31. Menter DG, Sabichi AL, Lippman SM. Selenium effects on prostate cell growth. Cancer Epidemiol Biomarkers Prev 2000 Nov;9(11):1171-1182.

32. Combs GF, Jr. Impact of selenium and cancer-prevention findings on the nutrition-health paradigm. Nutr Cancer 2001;40(1):6-11.

33. Cheng J, Weiqin J, Clement I, Howard G, Junxuan L. Selenium-induced inhibition of angiogenesis in mammary cancer at chemopreventive levels of intake. Molecular Carcinogenesis 1999;26(4):213-225.

34. Jiang C, Wang Z, Ganther H, Lu J. Caspases as Key Executors of Methyl Selenium-induced Apoptosis (Anoikis) of DU-145 Prostate Cancer Cells. Cancer Res 2001 April 1, 2001;61(7):3062-3070.

35. Spallholz JE. On the nature of selenium toxicity and carcinostatic activity. Free Radic Biol Med 1994 Jul;17(1):45-64.

36. Yang J, Huang K, Qin S, Wu X, Zhao Z, Chen F. Antibacterial action of selenium-enriched probiotics against pathogenic Escherichia coli. Dig Dis Sci 2009 Feb;54(2):246-254.

Page 149: Nanostructured Selenium for Biomedical Applications: from

132

37. Ratushnaya EV, Kirova YI, Suchkov MA, Drevko BI, Borodulin VB. Synthesis and Antibacterial Activity of Organoselenium Compounds. Pharmaceutical Chemistry Journal 2002;36(12):652-653.

38. Pietka-Ottlik M, Wojtowicz-Mlochowska H, Kolodziejczyk K, Piasecki E, Mlochowski J. New organoselenium compounds active against pathogenic bacteria, fungi and viruses. Chem Pharm Bull (Tokyo) 2008 Oct;56(10):1423-1427.

39. Kucukbay H, Durmaz R, Orhan E, Gunal S. Synthesis, antibacterial and antifungal activities of electron-rich olefins derived benzimidazole compounds. Farmaco 2003 Jun;58(6):431-437.

40. Sieber F, Günther WHH, Daziano J-p, Krieg-kowald M, Bousbaa J, Bula RJ, inventors. Method of making, and the use of cytotoxic agents containing elemental selenium, 2007.

41. Crompton K. Joint replacements become more routine here. J Business-Spokane 2004;19:10.

42. Buser D, Nydegger T, Oxland T, Cochran DL, Schenk RK, Hirt HP, et al. Interface shear strength of titanium implants with a sandblasted and acid-etched surface: a biomechanical study in the maxilla of miniature pigs. J Biomed Mater Res 1999 May;45(2):75-83.

43. Webster T. Nanophase ceramics: the future of orthopedic and dental implant material. In: Ying J, editor. Nanostructured Materials. New York: Academy Press, 2001. p. 125-166.

44. Kaplan FS HW, Keaveny TM., Boskey A, Einhorn TA, Iannotti JP. Biomaterials. In: SP S, editor. Orthopedic Basic Science. Columbus: American Academy of Orthopedic Surgeons, 1994. p. 460-478.

45. Kaplan FS LW, Keaveny TM, Boskey A, Einhorn TA, Iannotti JP. . Forms and functions of bone. In: SP S, editor. Orthopedic basic science. Columbus: American Academy of Orthopedic Surgeons, 1994. p. 127-185.

46. Park JB, RS L. Biomaterials-An introduction. New York: Plenum Press, 1992. 47. Webster TJ, Ergun C, Doremus RH, Siegel RW, Bizios R. Enhanced functions of

osteoblasts on nanophase ceramics. Biomaterials 2000 Sep;21(17):1803-1810. 48. Webster TJ, Siegel RW, Bizios R. Osteoblast adhesion on nanophase ceramics.

Biomaterials 1999 Jul;20(13):1221-1227. 49. Webster TJ, Ejiofor JU. Increased osteoblast adhesion on nanophase metals: Ti,

Ti6Al4V, and CoCrMo. Biomaterials 2004 Aug;25(19):4731-4739. 50. Webster TJ, Ergun C, Doremus RH, Siegel RW, Bizios R. Specific proteins mediate

enhanced osteoblast adhesion on nanophase ceramics. J Biomed Mater Res 2000 Sep 5;51(3):475-483.

51. Perla V, Webster TJ. Better osteoblast adhesion on nanoparticulate selenium- A promising orthopedic implant material. J Biomed Mater Res A 2005 Nov 1;75(2):356-364.

52. Whanger P, Vendeland S, Park YC, Xia Y. Metabolism of subtoxic levels of selenium in animals and humans. Ann Clin Lab Sci 1996 Mar-Apr;26(2):99-113.

53. Zhang S-Y, Zhang J, Wang H-Y, Chen H-Y. Synthesis of selenium nanoparticles in the presence of polysaccharides. Materials Letters 2004;58(21):2590-2594.

54. Kopeikin V, Valueva S, Kipper A, Borovikova L, Filippov A. Synthesis of selenium nanoparticles in aqueous solutions of poly(vinylpyrrolidone) and morphological characteristics of the related nanocomposites. Polymer science Series A 2003(45(4)):374-379.

Page 150: Nanostructured Selenium for Biomedical Applications: from

133

55. Lin Z-H, Chris Wang CR. Evidence on the size-dependent absorption spectral evolution of selenium nanoparticles. Materials Chemistry and Physics 2005;92(2-3):591-594.

56. Wang H, Wei W, Zhang SY, Shen YX, Yue L, Wang NP, et al. Melatonin-selenium nanoparticles inhibit oxidative stress and protect against hepatic injury induced by Bacillus Calmette-Guerin/lipopolysaccharide in mice. J Pineal Res 2005 Sep;39(2):156-163.

57. Thapa A, Miller DC, Webster TJ, Haberstroh KM. Nano-structured polymers enhance bladder smooth muscle cell function. Biomaterials 2003 Aug;24(17):2915-2926.

58. ATCC. 2010 [cited 2010 2/27]; Available from: http://www.atcc.org/ATCCAdvancedCatalogSearch/ProductDetails/tabid/452/Default.aspx?ATCCNum=CRL-11372&Template=cellBiology

59. Khanna C, Prehn J, Yeung C, Caylor J, Tsokos M, Helman L. An orthotopic model of murine osteosarcoma with clonally related variants differing in pulmonary metastatic potential. Clinical and Experimental Metastasis 2000;18(3):261-271.

60. Moulder JF, Stickle WF, Sobol PE, Bomben KD. Handbook of X-ray photoelectron spectroscopy: A reference book of standard spectra for identification and interpretation of xps data. Minnesota: Perkin-Elmer, 1992.

61. Zhang JS, Gao XY, Zhang LD, Bao YP. Biological effects of a nano red elemental selenium. Biofactors 2001;15(1):27-38.

62. Yang CF, Shen HM, Ong CN. Ebselen induces apoptosis in HepG2 cells through rapid depletion of intracellular thiols. Archives of Biochemistry and Biophysics 2000;374(2):142-152.

63. Iwata S, Hori T, Sato N, Hirota K, Sasada T, Mitsui A, et al. Adult T cell leukemia (ATL)-derived factor/human thioredoxin prevents apoptosis of lymphoid cells induced by L-cystine and glutathione depletion: possible involvement of thiol-mediated redox regulation in apoptosis caused by pro-oxidant state. The Journal of Immunology 1997;158(7):3108.

64. Borutaite V, Brown GC. Nitric oxide induces apoptosis via hydrogen peroxide, but necrosis via energy and thiol depletion. Free Radical Biology and Medicine 2003;35(11):1457-1468.

65. Aoshiba K, Yasui S, Nishimura K, Nagai A. Thiol depletion induces apoptosis in cultured lung fibroblasts. American journal of respiratory cell and molecular biology 1999;21(1):54.

66. Oberley TD, Oberley LW. Antioxidant enzyme levels in cancer. Histology and histopathology 1997;12(2):525.

67. Hileman EO, Liu J, Albitar M, Keating MJ, Huang P. Intrinsic oxidative stress in cancer cells: a biochemical basis for therapeutic selectivity. Cancer chemotherapy and pharmacology 2004;53(3):209-219.

68. Coursin DB, Cihla HP, Sempf J, Oberley TD, Oberley LW. An immunohistochemical analysis of antioxidant and glutathione S-transferase enzyme levels in normal and neoplastic human lung. Histology and histopathology 1996;11(4):851.

69. Anderson M, Meister A, inventors. Method for depletion of glutathione, 1995. 70. Zhdanov VP, Kasemo B. Van der Waals interaction during protein adsorption on a solid

covered by a thin film. Langmuir 2001;17(18):5407-5409.

Page 151: Nanostructured Selenium for Biomedical Applications: from

134

71. Besenbacher F, Chevallier J, Hovgaard MB, Rechendorff K, Foss M. Fibronectin Adsorption on Tantalum: The Influence of Nanoroughness. Journal of Physical Chemistry B 2008;112(28):8241-8249.

72. Magnusson MK, Mosher DF. Fibronectin: structure, assembly, and cardiovascular implications. Arteriosclerosis, thrombosis, and vascular biology 1998;18(9):1363.

73. Moursi AM, Damsky CH, Lull J, Zimmerman D, Doty SB, Aota S, et al. Fibronectin regulates calvarial osteoblast differentiation. Journal of Cell Science 1996;109(6):1369.

74. Puleo DA, Bizios R. Mechanisms of fibronectin-mediated attachment of osteoblasts to substrates in vitro. Bone and mineral 1992;18(3):215-226.

75. Lamponi S, Forbicioni M, Barbucci R. The role of fibronectin in cell adhesion to spiral patterned TiO2 nanoparticles. J Appl Biomater Biomech 2009 May-Aug;7(2):104-110.

76. Verrier S, Pallu S, Bareille R, Jonczyk A, Meyer J, Dard M, et al. Function of linear and cyclic RGD-containing peptides in osteoprogenitor cells adhesion process. Biomaterials 2002;23(2):585-596.

77. Ponik SM, Pavalko FM. Formation of focal adhesions on fibronectin promotes fluid shear stress induction of COX-2 and PGE2 release in MC3T3-E1 osteoblasts. Journal of Applied Physiology 2004;97(1):135.

78. Schneider G, Burridge K. Formation of focal adhesions by osteoblasts adhering to different substrata. Experimental cell research 1994;214(1):264-269.

79. Klee D, Ademovic Z, Bosserhoff A, Hoecker H, Maziolis G, Erli HJ. Surface modification of poly (vinylidenefluoride) to improve the osteoblast adhesion. Biomaterials 2003;24(21):3663-3670.

80. Mosher DF. Plasma fibronectin concentration: a risk factor for arterial thrombosis? Arteriosclerosis, thrombosis, and vascular biology 2006;26(6):1193.

81. Altankov G, Groth T. Reorganization of substratum-bound fibronectin on hydrophilic and hydrophobic materials is related to biocompatibility. Journal of Materials Science: Materials in Medicine 1994;5(9):732-737.

82. Nugiel DJ, Wood DJ, Sung KLP. Quantification of adhesiveness of osteoblasts to titanium surfaces in vitro by the micropipette aspiration technique. Tissue Engineering 1996;2(2):127-140.

83. Feder J. Random sequential adsorption. Journal of Theoretical Biology 1980;87(2):237-254.

84. Pomeau Y. Some asymptotic estimates in the random parking problem. Journal of Physics A: Mathematical and General 1980;13:L193-L196.

85. Schaaf P, Voegel J-C, Senger B. From Random Sequential Adsorption to Ballistic Deposition: A General View of Irreversible Deposition Processes. The Journal of Physical Chemistry B 2000;104(10):2204-2214.

86. Talbot J, Tarjus G, Van Tassel PR, Viot P. From car parking to protein adsorption: an overview of sequential adsorption processes. Colloids and Surfaces A: Physicochemical and Engineering Aspects 2000;165(1-3):287-324.

87. Elaissari A, Haouam A, Huguenard C, Pefferkorn E. Kinetic factors in polymer adsorption at solid/liquid interfaces. Methods of study of the adsorption mechanism. Journal of Colloid and Interface Science 1992;149(1):68-83.

88. Ramsden JJ. Concentration scaling of protein deposition kinetics. Physical Review Letters 1993;71(2):295.

Page 152: Nanostructured Selenium for Biomedical Applications: from

135

89. Van Tassel PR, Guemouri L, Ramsden JJ, Tarjus G, Viot P, Talbot J. A particle-level model of irreversible protein adsorption with a postadsorption transition. Journal of Colloid and Interface Science 1998;207(2):317-323.

90. Elaissari A, Pefferkorn E. Polyelectrolyte adsorption at solid/liquid interfaces: A simple model for the structural relaxation and excluded area effects. Journal of Colloid and Interface Science 1991;143(1):85-91.

91. Patrikar RM. Modeling and simulation of surface roughness. Applied Surface Science 2004;228(1-4):213-220.

92. Hedsten K, Fonollosa J, Enoksson P, Modh P, Bengtsson J, Sutherland DS, et al. Optical Label-Free Nanoplasmonic Biosensing Using a Vertical-Cavity Surface-Emitting Laser and Charge-Coupled Device. Analytical Chemistry:442-651.

93. Dimilla PA, Albelda SM, Quinn JA. Adsorption and elution of extracellular matrix proteins on non-tissue culture polystyrene petri dishes. Journal of Colloid and Interface Science 1992;153(1):212-225.

94. Guemouri L, Ogier J, Zekhnini Z, Ramsden JJ. The architecture of fibronectin at surfaces. The Journal of Chemical Physics 2000;113:8183.

95. Hemmersam AG, Rechendorff K, Foss M, Sutherland DS, Besenbacher F. Fibronectin adsorption on gold, Ti-, and Ta-oxide investigated by QCM-D and RSA modelling. Journal of Colloid and Interface Science 2008;320(1):110-116.

96. Erickson HP, Carrell N, McDonagh J. Fibronectin molecule visualized in electron microscopy: a long, thin, flexible strand. J Cell Biol 1981 December 1, 1981;91(3):673-678.

97. Rocco M, Infusini E, Daga MG, Gogioso L, Cuniberti C. Models of fibronectin. The EMBO journal 1987;6(8):2343.

98. Victor EK, Marina AG, Marina VB, Vladimir NS, Valdimir VF, Olga MZ, et al. A Study of the Structure of Fibronectin. European Journal of Biochemistry 1981;119(3):619-624.

99. Tooney NM, Mosesson MW, Amrani DL, Hainfeld JF, Wall JS. Solution and surface effects on plasma fibronectin structure. J Cell Biol 1983 December 1, 1983;97(6):1686-1692.

100. Khang D. adhesion protein adsorption and bone cell growth on carbon nanotube composite materials: Perdue University; 2003.

101. Yoshimitsu Z, Nakajima A, Watanabe T, Hashimoto K. Effects of Surface Structure on the Hydrophobicity and Sliding Behavior of Water Droplets. Langmuir 2002;18(15):5818-5822.

102. Ishino C, et al. Wetting transitions on rough surfaces. EPL (Europhysics Letters) 2004;68(3):419.

103. Maki DG, Stolz SM, Wheeler S, Mermel LA. Prevention of Central Venous Catheter-Related Bloodstream Infection by Use of an Antiseptic-Impregnated Catheter: A Randomized, Controlled Trial. Ann Intern Med 1997 August 15, 1997;127(4):257-266.

104. Yahav D, Rozen-Zvi B, Gafter-Gvili A, Leibovici L, Gafter U, Paul M. Antimicrobial lock solutions for the prevention of infections associated with intravascular catheters in patients undergoing hemodialysis: systematic review and meta-analysis of randomized, controlled trials. Clin Infect Dis 2008 Jul 1;47(1):83-93.

105. Trerotola SO, Johnson MS, Shah H, Kraus MA, McKusky MA, Ambrosius WT, et al. Tunneled hemodialysis catheters: use of a silver-coated catheter for prevention of infection--a randomized study. Radiology 1998 May 1998;207(2):491-496.

Page 153: Nanostructured Selenium for Biomedical Applications: from

136

106. Logghe C, Van Ossel C, D'Hoore W, Ezzedine H, Wauters G, Haxhe JJ. Evaluation of chlorhexidine and silver-sulfadiazine impregnated central venous catheters for the prevention of bloodstream infection in leukaemic patients: a randomized controlled trial. J Hosp Infect 1997 Oct;37(2):145-156.

107. Dunser MW, Mayr AJ, Hinterberger G, Florl CL, Ulmer H, Schmid S, et al. Central venous catheter colonization in critically ill patients: a prospective, randomized, controlled study comparing standard with two antiseptic-impregnated catheters. Anesth Analg 2005 Dec;101(6):1778-1784.

108. Winson L. Catheterization: a need for improved patient management. Br J Nurs 1997 Nov 27-Dec 10;6(21):1229-1232, 1234, 1251-1222.

109. Akiyama H, Okamoto S. Prophylaxis of indwelling urethral catheter infection: clinical experience with a modified Foley catheter and drainage system. J Urol 1979 Jan;121(1):40-42.

110. Liedberg H, Lundeberg T. Silver alloy coated catheters reduce catheter-associated bacteriuria. Br J Urol 1990 Apr;65(4):379-381.

111. Kumon H, Hashimoto H, Nishimura M, Monden K, Ono N. Catheter-associated urinary tract infections: impact of catheter materials on their management. International journal of antimicrobial agents 2001;17(4):311-316.

112. Bach A, Eberhardt H, Frick A, Schmidt H, Bottiger BW, Martin E. Efficacy of silver-coating central venous catheters in reducing bacterial colonization. Crit Care Med 1999 Mar;27(3):515-521.

113. Mathews SM, Spalholz JE, Dubielzig RR, Grimson MJ, Reid TW. The Efficacy of an Organo-Selenium Coating on Silicone Hydrogel Contact Lenses. Investigative Ophtalmology and Visual Science 2002;43(12):3109.

114. Spallholz JE, Shriver BJ, Reid TW. Dimethyldiselenide and methylseleninic acid generate superoxide in an in vitro chemiluminescence assay in the presence of glutathione: implications for the anticarcinogenic activity of L-selenomethionine and L-Se-methylselenocysteine. Nutrition and cancer 2001;40(1):34-41.

115. Tran PL, Hammond AA, Mosley T, Cortez J, Gray T, Colmer-Hamood JA, et al. An Organo-Selenium Coating on Cellulose Inhibits Pseudomonas aeruginosa and Staphylococcus aureus Biofilm Formation. Appl Environ Microbiol 2009 April 3, 2009:AEM.02683-02608.

116. Brouwer WM, Zsom RLJ. Polystyrene latex particle surface characteristics. Colloids and Surfaces 1987;24(2-3):195-208.

117. Seebergh JE, Berg JC. Evidence of a hairy layer at the surface of polystyrene latex particles. Colloids and Surfaces A: Physicochemical and Engineering Aspects 1995;100:139-153.